0% found this document useful (0 votes)
529 views417 pages

NSBA Routine Steel Bridge Design Guide

Uploaded by

kejspm
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
529 views417 pages

NSBA Routine Steel Bridge Design Guide

Uploaded by

kejspm
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 417

Navigating Routine

Navigating Routine
Steel Bridge Design
Steel
AASHTO Bridge
LRFD Bridge Design
Design
Specifications, 9th Edition
AASHTO LRFD Bridge Design
Specifications, 9th Edition
© AISC 2021

by

American Institute of Steel Construction

All rights reserved. This book or any part thereof must not be reproduced
in any form without the written permission of the publisher.
The AISC logo is a registered trademark of AISC.

The information presented in this publication has been prepared following recognized principles of design
and construction. While it is believed to be accurate, this information should not be used or relied upon
for any specific application without competent professional examination and verification of its accuracy,
suitability and applicability by a licensed engineer or architect. The publication of this information is not a
representation or warranty on the part of the American Institute of Steel Construction, its officers, agents,
employees or committee members, or of any other person named herein, that this information is suitable
for any general or particular use, or of freedom from infringement of any patent or patents. All represen-
tations or warranties, express or implied, other than as stated above, are specifically disclaimed. Anyone
making use of the information presented in this publication assumes all liability arising from such use.

Caution must be exercised when relying upon standards and guidelines developed by other bodies and
incorporated by reference herein since such material may be modified or amended from time to time sub-
sequent to the printing of this edition. The American Institute of Steel Construction bears no responsibility
for such material other than to refer to it and incorporate it by reference at the time of the initial publication
of this edition.

NSBA Guide to Navigating Routine Steel Bridge Design / i


FOREWORD
The design of bridges requires the efforts of engineers educated and experienced in structural
design, specifically the unique aspects of bridge design. However, bridge design need not be
complicated or challenging, particularly for the more routine bridges which form a large part of
the inventory of transportation structures in the United States. In particular, the design of “routine
steel I-girder bridges,” a workhorse structure type, can be relatively simple if the engineer knows
where to focus their efforts and is provided with guidance on how to streamline the more
predictable, repetitive aspects of the design effort.
To this end, the National Steel Bridge Alliance (NSBA), a division of the American Institute of
Steel Construction (AISC) has developed this Guide to Navigating Routine Steel Bridge Design.
The goal of this Guide is to help designers navigate the comprehensive design provisions of the
AASHTO LRFD Bridge Design Specifications (AASHTO LRFD BDS), identifying just the
provisions that are applicable to the design of routine steel I-girder bridges, explaining how to
apply those provisions, recommending practices proven to lead to economical designs, and
suggesting ways to streamline the design effort.
This Guide is meant to be used as an interactive reference, rather than as a textbook read from
cover to cover. The general flow of design tasks is outlined, and at any point in the design process
the reader can quickly jump to a detailed Discussion of any particular AASHTO LRFD BDS
provision. Through this process, a bridge engineer will not only find answers to specific questions,
but can also increase their familiarity with, and understanding of, the AASHTO LRFD BDS.
We would like to gratefully acknowledge the support of Christopher Garrell (NSBA), who steered
the development of this Guide. Credit is also due to Kaylene Callicoatt (HDR), who helped
assemble and edit this Guide. Finally, we would like to thank the numerous professionals who
provided invaluable peer review comments during the writing of this Guide, including: Travis Butz
(Burgess and Niple), Matt Farrar (Idaho DOT), Jamie Farris (Texas DOT), Karl Frank
(Consultant), Christina Freeman (Florida DOT), Dennis Golabek (WSP), John Holt (Modjeski
and Masters), Ted Kniazewycz (Tennessee DOT), Shane Kuhlman (New Mexico DOT), Ronnie
Medlock (High Steel Structures), Adam Price (Tennessee DOT), Curtis Rockiki (Texas DOT),
Kevin Sear (AECOM), Tony Shkurti (HNTB), Jason Stith (Michael Baker International), Greg
Turco (Texas DOT), Jeffrey Vetter (Idaho DOT), Dayi Wang (Federal Highway Administration),
Wagdy Wassef (WSP), and Jaclyn Whelan (AECOM).

Michael Grubb (M.A. Grubb & Associates, LLC)


Domenic Coletti, Aleksander Nelson, Anthony Ream (HDR)
November 2020

NSBA Guide to Navigating Routine Steel Bridge Design / ii


Page intentionally left blank

NSBA Guide to Navigating Routine Steel Bridge Design / iii


TABLE OF CONTENTS
SCOPE OF THIS GUIDE ............................................................................................................... 1
DETERMINATION DEFINITIONS.............................................................................................. 3
TERMINOLOGY ........................................................................................................................... 4
DEFINITION OF A “ROUTINE STEEL I-GIRDER BRIDGE” .................................................. 5
DEFINITION CHECKLIST FOR A “ROUTINE STEEL I-GIRDER BRIDGE” ......................... 9
USEFUL REFERENCES ............................................................................................................. 11
GENERAL FLOW OF DESIGN TASKS .................................................................................... 15
GRAPHICAL INDEX OF DESIGN TASKS ............................................................................... 16
DESIGN TASK QUICK LINKS .................................................................................................. 20
General Considerations ............................................................................................................. 21
Deck Design .............................................................................................................................. 22
Resistance Factors and Load Modifiers .................................................................................... 23
Load Combinations and Load Factors ...................................................................................... 24
Live Load Force Effects - Introduction ..................................................................................... 25
Live Load Force Effects - Flexure ............................................................................................ 26
Live Load Force Effects - Shear ............................................................................................... 27
Other Load Effects and Factors Affecting Load Effect Calculations ....................................... 28
Girder Flexure Design – General .............................................................................................. 29
Girder Flexure Design – Constructibility .................................................................................. 30
Girder Flexure Design – Service Limit State ............................................................................ 31
Girder Flexure Design – Fatigue and Fracture Limit State ....................................................... 32
Girder Flexure Design – Strength Limit State .......................................................................... 33
Girder Shear Design .................................................................................................................. 34
Stiffener Design ......................................................................................................................... 35
Shear Connector Design ............................................................................................................ 36
Splice Design............................................................................................................................. 37
Cross-Frame/Diaphragm Design ............................................................................................... 38
Bolted Connection Design ........................................................................................................ 39
Welded Connection Design ....................................................................................................... 40
Connection Design – Miscellaneous Checks ............................................................................ 41
SECTION 1: INTRODUCTION .................................................................................................. 42
SECTION 2: GENERAL DESIGN AND LOCATION FEATURES ......................................... 48
SECTION 3: LOADS AND LOAD FACTORS .......................................................................... 53
SECTION 4: STRUCTURAL ANALYSIS AND EVALUATION ............................................ 82
SECTION 6: STEEL STRUCTURES ....................................................................................... 118
CONCLUSION ........................................................................................................................... 412

NSBA Guide to Navigating Routine Steel Bridge Design / iv


SCOPE OF THIS GUIDE
This NSBA Guide to Navigating Routine Steel Bridge Design (this Guide) primarily address the
design of steel superstructures for “routine steel I-girder bridges.” The intent is to illustrate which
provisions of the AASHTO LRFD Bridge Design Specifications (AASHTO LRFD BDS) are
applicable to the design of these types of structures, and perhaps more importantly, which
provisions are not applicable or are perhaps only partially or conditionally applicable or beyond
the scope of superstructure design. In doing so, it is hoped that this Guide will help designers to
streamline the design process, avoid unnecessary or misguided effort, and simplify their approach
to the necessary tasks associated with the design of a more routine steel I-girder bridge. The
intended audience includes, but is not limited to, those designers who may be less experienced
with steel bridge design. This Guide is intended to be used in combination with the AASHTO
LRFD BDS and should not be used as a substitute for the AASHTO LRFD BDS itself.
The definition of a “routine steel I-girder bridge” is provided below and is intended to encompass
a large family of straight steel I-girder bridges with little or no skew, “routine” span lengths, and
commonly used framing layouts and details. To keep the Guide focused, the following items are
specifically excluded from the scope of this Guide:
• Barrier rail design: Standard details, often mandated by the local Owner-agency, are
generally used, so design guidance per se is not needed.
• Deck design: The design of concrete decks for steel I-girder bridges is often governed by
Owner-agency policy manuals (e.g., standard designs, pre-calculated design tables, etc.),
and so their design is not addressed herein.
• Substructure and foundation design: The wide variety of types, configurations, materials,
and conditions associated with substructure and foundation design make it difficult to
provide a succinct set of guidelines that would be broadly applicable over all parts of the
U.S.
• Bearing design: The variety of types of bearings and the associated variety of local Owner-
agency or regional bearing design preferences make it difficult to provide a succinct set of
guidelines that would be broadly applicable over all parts of the U.S. For the purposes of
this Guide, the use of “routine” bearings (as defined under “Definition of a ‘Routine Steel
I-Girder Bridge’ below) is assumed.
Given these scope limitations, consideration of the Extreme Event limit state, outside of the
identification of the seismic zone for design, is also omitted from this Guide.
If a given bridge somehow falls partially outside the limits of the definition of a “routine steel I-
girder bridge” or outside the exclusions of this scope, this Guide may still provide value to
designers; in such cases, senior bridge engineers with extensive experience in steel bridge design
should be consulted when determining if and how to apply any of the recommendations provided
herein.

NSBA Guide to Navigating Routine Steel Bridge Design / 1


Conversely, designers should be cognizant of local Owner-agency policies which may supersede
the recommendations and information presented in this Guide. In such cases senior bridge
engineers with extensive experience in steel bridge design should be consulted when determining
how to apply this Guide in conjunction with Owner-agency policy.
Finally, while the scope of this Guide is limited to helping bridge designers understand and
correctly apply specific provisions of the AASHTO LRFD BDS, it is worthwhile to briefly address
the broader scope of bridge design.
• Correctness of structural design – It is of the utmost importance to correctly design a
bridge with sufficient strength to carry the intended loads without suffering undue distress
or structural failure. The safety of the general public who travel on and under bridges is the
overriding concern of a bridge designer.
• Correctness of geometric information – It is also important that pertinent geometric
information be correctly, accurately, and precisely calculated and presented on the plans.
Errors in the calculation and presentation of geometric information can result in costly
delays, rework, and claims.
• Correctness of estimated quantities – It is also important that required estimated
quantities, particularly quantities which form the basis of payment for a bridge, be
correctly, accurately, and precisely calculated and presented on the plans. Errors in the
calculation and presentation of estimated quantities can result in costly delays, rework, and
claims.

NSBA Guide to Navigating Routine Steel Bridge Design / 2


DETERMINATION DEFINITIONS
Each Article of the AASHTO LRFD BDS is assigned a Determination of Applicability to the
design of routine steel I-girder bridges. These Determinations of Applicability do not relieve
designers of their responsibility to read, understand, and correctly apply the provisions of the
AASHTO LRFD BDS, but instead are intended to aid designers in navigating and understanding
those provisions.
The various Determinations are defined as follows:
1. Applicable: The Article, in its entirety, is fully applicable to the design of routine steel I-
girder bridges
2. Partially Applicable: Parts of the Article are applicable to the design of routine steel I-
girder bridges, other parts are not applicable; see the Discussion for explanation
3. Conditionally Applicable: Some or all of the Article may be applicable to the design of
routine steel I-girder bridges depending on the circumstances; see the Discussion for
explanation
4. Not Applicable: None of the Article is applicable to the design of routine steel I-girder
bridges
5. Beyond Scope of Superstructure Design: Some or all of the Article may be applicable
to some aspect of the design of routine steel I-girder bridges, but is not applicable to
superstructure design; see the Discussion for explanation

NSBA Guide to Navigating Routine Steel Bridge Design / 3


TERMINOLOGY
For the purposes of this Guide, the following terminology is defined to avoid confusion:
AASHTO LRFD BDS: The term “AASHTO LRFD BDS” refers to the AASHTO LRFD Bridge
Design Specifications, 9th Edition, 2020.
Article: The term “Article” (capitalized) refers to a specific numbered Article in the AASHTO
LRFD BDS.
Commentary: The term “Commentary” (capitalized) refers to a specific numbered commentary
section related to an Article in the AASHTO LRFD BDS.
Determination: The term “Determination” (capitalized) refers to the determination of
applicability of a given Article in the AASHTO LRFD BDS to the design of routine steel I-girder
bridges.
Discussion: The term “Discussion” (capitalized) refers to the discussion explaining the rationale
behind a specific Determination in this Guide and/or providing guidance on how to streamline the
design tasks or actions associated with the Article referenced.
Guide: The term “Guide” (capitalized) refers to this NSBA Guide to Navigating Routine Steel
Bridge Design.

NSBA Guide to Navigating Routine Steel Bridge Design / 4


DEFINITION OF A “ROUTINE STEEL I-GIRDER BRIDGE”
For the purposes of implementing the recommendations of this Guide, a “routine steel I-girder
bridge” is defined as the superstructure of a bridge meeting the following characteristics:
• Straight (non-curved) girders.
• Straight (non-curved) deck.
• Framing such that flange lateral bending can be neglected, specifically:
o Skew not more than 20 degrees, where skew is measured as the angular deviation
of the orientation of the supports from perpendicular to the centerline of the bridge.
o Parallel supports or supports which are within 10 degrees of being parallel.
o Contiguous cross-frames or diaphragms.
o Parallel girders.
• Constant deck width.
• Skew Index less than or equal to 0.30, where the Skew Index is as defined in Eq. 4.6.3.3.2-
2.
• Constant depth girders (i.e., girders with parallel flanges – no haunched or tapered girders).
• Superstructure designed to meet the live load deflection limits outlined in Article 2.5.2.6.2.
• Span lengths not exceeding 200 feet.
• No lateral bracing.
• Non-hybrid girders (i.e., girder flanges and webs in all spans shall be fabricated using steel
of the same grade/yield strength).
• Steel grades/yield strengths of 36 ksi or 50 ksi.
• Structural steel using any of the following corrosion protection systems: painted steel,
galvanized steel, metallized steel, or uncoated weathering steel.
• Stringer-type cross-section (no girder-substringer systems), with four or more girders in
the cross-section.
• No longitudinal web stiffeners.
• Typical round, headed, stud-type shear connectors.

NSBA Guide to Navigating Routine Steel Bridge Design / 5


• Cast-in-place concrete composite decks, formed using one of the following methods:
o Conventional (removable) wood or steel forms.
o Stay-in-place corrugated metal forms.
o Partial-depth precast concrete deck panels.
• Routine barrier rail heights (up to 3’-6” tall.)
• No sound walls, noise walls, or other solid barriers atop or adjacent to the barrier rails.
• Bolted field splices.
• Welded steel bearing stiffeners, intermediate stiffeners, and cross-frame/diaphragm
connection plates.
• Solid web steel diaphragms (e.g., bent plate, rolled channel shape, rolled I-section, or built-
up plate girder diaphragms) bolted to connection plates, or truss-type steel cross-frames
bolted or welded to gusset plates or connection plates.
• Routine bearings, i.e., those types of bearings which do not provide restraint of girder major
axis bending behavior (bearings which do not restrain girder major axis bending end
rotations, do not affect flange stresses, etc.) and thus do not affect the behavior of the
superstructure in a manner which would invalidate the results of a typical line girder
analysis. Examples of routine bearings include steel-reinforced elastomeric bearing pads,
high-load multi-rotational bearings such as disc or pot bearings, or roller and rocker
bearings.
• No cover plates.
• Generally Seismic Zone 1 only. The guidance provided in this Design Guide may be useful
in the design of bridges in higher seismic zones, with input from a senior bridge engineer
with extensive experience in the design of steel girder bridges for high-seismic zones,
particularly when it is appropriate to use a spline model or other simplified analysis method
for the seismic analysis, such that the need for a seismic analysis model does not imply the
need for a refined analysis model of the steel superstructure.
• Highway bridges only. Railroad and transit bridges, and bridge intended for use solely by
pedestrians, are excluded from the definition of a “routine steel I-girder bridge” for the
purposes of this Guide. However, highway bridges with sidewalk loading are included.
• Bridges for which the superstructure is not subject to stream flow loading, ice loading, or
vessel collision loading. This implies, at a minimum, that there is sufficient freeboard
between the low chord of the superstructure and the high-water elevation during design
flood events.
• No redistribution of negative moments in continuous beam or girder bridges.

NSBA Guide to Navigating Routine Steel Bridge Design / 6


• Only single-phase construction or simple multi-phase construction. The definition of
“simple multi-phase construction” is as follows: Each phase of construction must meet the
various definitions of a “routine steel I-girder bridge,” including the requirements for
overall deck and framing plan geometry and minimum number of girders in the cross-
sections of each phase of construction. In addition, a closure bay and closure pour must be
provided, whereby erection and deck placement of each independent phase of
superstructure construction is completed prior to installation of cross-frames or diaphragms
in the closure bay between the adjacent phases of construction, and installation of the
closure bay cross-frames or diaphragms is completed prior to placement of the deck closure
pour between the adjacent phases of construction. Note that in addition to evaluating each
phase of construction, the fully-completed bridge must also be evaluated. For further
guidance, see Section 6.3.2.5.4 of the Reference Manual for NHI Course 130081, Load and
Resistance Factor Design (LRFD) for Highway Bridge Superstructures. For cases which
differ from these conditions, consultation with a senior bridge engineer with extensive
experience in steel girder bridge design can potentially determine ways to apply some or
all of the guidance in this Guide.
• Only full bridge replacement, new bridge construction, or simple bridge widening
construction. The definition of “simple bridge widening” is as follows: Both the existing
structure (including consideration of any partial demolition or partial removal of existing
superstructure) and the widened portion of the superstructure must meet all other
definitions of a “routine steel I-girder bridge,” including the requirements for overall deck
and framing plan geometry and minimum number of girders in the cross-sections of both
the existing superstructure and the widened portion of the superstructure. In addition, a
closure bay and closure pour must be provided, whereby erection and deck placement of
each independent portion of superstructure construction is completed prior to installation
of cross-frames or diaphragms in the closure bay between the adjacent portion of
construction, and installation of the closure bay cross-frames or diaphragms is completed
prior to placement of the deck closure pour between the adjacent portions of construction.
Note that in addition to evaluating each stage of construction, the fully-completed bridge
must also be evaluated. For further guidance, see Section 6.3.2.5.4 of the Reference Manual
for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For cases which differ from these conditions, consultation with a senior
bridge engineer with extensive experience in steel girder bridge design can potentially
determine ways to apply some or all of the guidance in this Guide.
These assumptions are intended to limit the scope of the guidance presented in this Guide. Also,
by inference many of these assumptions allow for the use of line girder methods of analysis.
Refined methods of analysis (such as 2D grid analysis, 2D plate-and-eccentric beam analysis, 3D
finite element analysis, etc.) are not required, nor recommended, for the design of routine steel I-
girder bridges as defined herein. The use of the empirical live load distribution factors associated
with line girder analysis methods usually result in a slightly, but not excessively, more
conservative live load distribution than would be determined using a refined method of analysis;

NSBA Guide to Navigating Routine Steel Bridge Design / 7


nevertheless, refined analysis methods are not required, nor recommended, for the design of
routine steel I-girder bridges as defined herein.
Using refined analysis methods for the design of a routine steel I-girder bridge requires additional
effort that would be better spent on other tasks such as exploring framing plan and girder design
refinement options, improving plan clarity, or implementing more robust checking and QC
procedures. Refined analysis methods are also inherently more complex than line girder analysis
methods, introducing more opportunities for errors. Finally, and importantly, using a refined
method of analysis to decrease conservatism in live load distribution could lead a designer to
reduce girder flange or web sizes during the initial design of a bridge. This could result in
difficulties later when the Owner-agency performs periodic routine load rating analyses of the
bridge. For the sake of practicality, most Owner-agencies default to using line girder analysis
methods for these load rating analyses – they have hundreds or thousands of bridges to load rate
each year and cannot afford to perform labor-intensive refined analyses when line girder analysis
methods would suffice. It is problematic when a bridge exhibits an insufficient load rating due
solely to the minor conservatism of line girder analysis methods, forcing the Owner-agency to
invest limited resources in performing a refined analysis to demonstrate that a bridge has sufficient
load-carrying capacity.

NSBA Guide to Navigating Routine Steel Bridge Design / 8


DEFINITION CHECKLIST FOR A “ROUTINE STEEL I-GIRDER BRIDGE”
Answer all questions with “Yes” or “No”. If any questions are answered “No”, the bridge does not
satisfy the definition of a routine steel I-girder bridge for the purposes of this Guide. For further
detail on any of these criteria, please see the preceding section of the Guide: DEFINITION OF A
“ROUTINE STEEL I-GIRDER BRIDGE”
If a given bridge somehow falls partially outside the limits of the definition of a “routine steel I-
girder bridge”, or outside the exclusions of this scope, this Guide may still provide value to
designers; in such cases, senior bridge engineers with extensive experience in steel bridge design
should be consulted when determining if and how to apply any of the recommendations provided
herein.
 Are the girders straight (non-curved)?
 Is the deck straight?
 Is the skew not more than 20 degrees?
 Are all supports parallel (or within 10 degrees of being parallel)?
 Are the cross-frames contiguous?
 Are the girders parallel?
 Is the deck constant width?
 Is the Skew Index (Eq. 4.6.3.3.2-2) less than or equal to 0.30?
 Do the girders have a constant web depth?
 Is the superstructure designed to meet the live-load deflection limits outlined in Article
2.5.2.6.2?
 Are all spans less than 200 feet?
 Does the design not include the use of lateral bracing?
 Are all girder flanges and webs fabricated using steel of the same grade/yield strength?
 Is the steel grade/yield strength of the girders 36 ksi or 50 ksi?
 Will one of the following corrosion protection systems – painted steel, galvanized steel,
metallized steel, or uncoated weathering steel – be used?
 Does the bridge feature a stringer-type configuration, with four or more girders in the cross-
section?
 Are the girders to be designed without the use of longitudinal web stiffeners?
 Are typical round, headed, stud-type shear connectors to be used?

NSBA Guide to Navigating Routine Steel Bridge Design / 9


 Will a cast-in-place composite concrete deck, formed using one of the following methods
– conventional (removable) wood or steel forms, stay-in-place corrugated metal forms,
partial-depth precast concrete deck panels – be used?
 Will the bridge be provided with routine barrier rail heights (no more than 3’-6” tall)?
 Will the bridge be free of superstructure-mounted sound walls, noise walls, or other solid
barriers atop or adjacent to the barrier rails?
 Will any needed field splices be constructed using bolted connections?
 Will any bearing stiffeners, intermediate stiffeners, and/or cross-frame or diaphragm
connection plates be fabricated from welded steel plates or bar stock?
 Will either solid web steel diaphragms (e.g., bent plate, rolled channel shape, rolled I-
section, or built-up plate girder diaphragms) bolted to connection plates, or truss-type steel
cross-frames bolted or welded to gusset plates or connection plates, be used to brace the
girders?
 Will only routine bearing types be used?
 Will the bridge girders be free of cover plates?
 Is the bridge located in Seismic Zone 1?
 Is the bridge carrying only highway vehicular traffic?
 Is the superstructure not subject to stream flow loading, ice loading, and vessel collision
loading?
 Will the use of design methods involving the redistribution of negative moments in
continuous beam or girder bridges be avoided?
 Will the bridge be constructed using only single-phase construction or simple multi-phase
construction?
 Will the construction of the bridge fall into one of these three categories – full bridge
replacement, new bridge construction, or simple bridge widening construction?

NSBA Guide to Navigating Routine Steel Bridge Design / 10


USEFUL REFERENCES
A list of useful references is provided below. These references are frequently cited in the Guide.
Some of the references may be slightly dated but represent the best guidance available at the time
of writing of this Guide. In general, the reader can typically compensate for the dated nature of
any particular reference by recognizing where references to specific provisions of the AASHTO
LRFD BDS may be out of date; in such cases the current provisions of the AAHTO LRFD BDS
should be followed and the guidance provided in the reference should be interpreted accordingly.
• AASHTO’s AASHTO Guide Specifications for Wind Loads on Bridges During Construction:
At the time of the writing of this Guide, the 1st Edition of this guide specification had been
published by AASHTO in 2017.

o https://store.transportation.org/Item/PublicationDetail?ID=3728

• AASHTO-NSBA Steel Bridge Collaboration’s Guideline G1.4-2006 Guidelines for Design


Details: At the time of the writing of this Guide, this guideline had been published by the
AASHTO/NSBA Steel Bridge Collaboration in 2006.

o https://www.aisc.org/globalassets/nsba/aashto-nsba-collab-docs/g-1.4-2006-
guidelines-for-design-details.pdf

• AASHTO-NSBA Steel Bridge Collaboration’s Guideline G12.1-2020 Guidelines to Design


for Constructability and Fabrication: At the time of the writing of this Guide, this guideline
had been published by the AASHTO/NSBA Steel Bridge Collaboration in 2020.

o https://www.aisc.org/globalassets/nsba/aashto-nsba-collab-docs/nsbagdc-3.pdf

• AASHTO-NSBA Steel Bridge Collaboration’s Guideline G13.1-2019 Guidelines for Steel


Girder Bridge Analysis: At the time of the writing of this Guide, this guideline had been
published by the AASHTO/NSBA Steel Bridge Collaboration in 2019.

o https://www.aisc.org/globalassets/nsba/aashto-nsba-collab-docs/g-13.1-2019-
guidelines-for-steel-girder-bridge-analysis.pdf

• AASHTO-NSBA Steel Bridge Collaboration’s Guide Specification S2.1-2018 Steel Bridge


Fabrication Guide Specification: At the time of the writing of this Guide, this guide
specification had been published by the AASHTO/NSBA Steel Bridge Collaboration in 2018.

o https://www.aisc.org/globalassets/nsba/aashto-nsba-collab-docs/s2.1-2018-steel-
bridge-fabrication-guide-specification.pdf

• AISC’s AISC Design Guide 17 High Strength Bolts - A Primer for Engineers: At the time of
the writing of this Guide, the 1st Edition of AISC Design Guide 17 had been published in 2002.

o https://www.aisc.org/products/publication/design-guide/design-guide-17-high-
strength-bolts--a-primer-for-structural-engineers/

NSBA Guide to Navigating Routine Steel Bridge Design / 11


• AISC’s Database of Rolled Steel Shape Section Properties: At the time of the writing of this
Guide, Version 15.0 of this database had been released in November 2017.

o https://www.aisc.org/globalassets/aisc/manual/v15.0-shapes-database/aisc-shapes-
database-v15.0.xlsx

• AISC’s Specifications for Structural Steel Buildings and Commentary: At the time of the
writing of this Guide, the current edition of this specification was dated July 7, 2016.

o https://www.aisc.org/globalassets/aisc/publications/standards/a360-16-spec-and-
commentary.pdf

• ASCE’s ASCE 7-10, Minimum Design Loads for Buildings and Other Structures: At the time
of the writing of this Guide, this guideline had been published by the American Society of
Civil Engineers in 2010. A later version had been published, but the AASHTO wind load
provisions are based on the 2010 version.

o https://sp360.asce.org/PersonifyEbusiness/Merchandise/Product-
Details/productId/232961952

• Coletti, D.A., and M.A. Grubb, “Practical Implementation of Stability Bracing Strength and
Stiffness Guidelines for Steel I-Girder Bridges,” Proceedings of the 2016 World Steel Bridge
Symposium, April 14, 2016, Orlando, FL.

o https://www.aisc.org/globalassets/nsba/conference-proceedings/2016/coletti_grubb---
2016-wsbs-final.pdf

• FHWA’s Steel Bridge Design Handbook: At the time of the writing of this Guide, the FHWA
had published this 19 volume guide, plus 6 full design examples, as Report No. FHWA-HIF-
16-002, dated December, 2015, at which time it was based on the 7th Edition, 2014, of the
AASHTO LRFD BDS. However, subsequently the FHWA had transferred the responsibility
for maintaining and updating the Steel Bridge Design Handbook to the NSBA and an effort
was underway to update the document for conformance with the 9th Edition of the AASHTO
LRFD BDS.

o https://www.fhwa.dot.gov/bridge/steel/pubs/hif16002/

• FHWA’s FHWA Bridge Welding Reference Manual: At the time of the writing of this Guide,
this manual had been published by the FHWA as Report No. FHWA-HIF-19-088, dated
September 2019.

o https://www.fhwa.dot.gov/bridge/steel/pubs/hif19088.pdf

NSBA Guide to Navigating Routine Steel Bridge Design / 12


• NHI and FHWA’s Reference Manual for NHI Course 130081, Load and Resistance Factor
Design (LRFD) for Highway Bridge Superstructures: At the time of the writing of this Guide,
this Reference Manual had been issued by the FHWA as Report No. FHWA-NHI-15-047,
dated July 2015. It was based on the 7th Edition, 2014, of the AASHTO LRFD BDS.

o https://www.fhwa.dot.gov/bridge/pubs/nhi15047.pdf

• NHI and FHWA’s Reference Manual for NHI Course 130102, Engineering for Structural
Stability in Bridge Construction: At the time of the writing of this Guide, this Reference
Manual had been issued by the FHWA as Report No. FHWA-NHI-15-044, dated April 2015.
It was based on the 6th Edition, 2012, of the AASHTO LRFD BDS.

o https://www.fhwa.dot.gov/bridge/pubs/nhi15044.pdf

• NHI and FHWA’s Reference Manual for NHI Course 130122, Design and Evaluation of Steel
Bridges for Fatigue and Fracture: At the time of the writing of this Guide, this Reference
Manual had been issued by the FHWA as Report No. FHWA-NHI-16-016, dated December
2016. It was based on the 7th Edition, 2014, of the AASHTO LRFD BDS, with Interim
Revisions through 2015.

o https://www.fhwa.dot.gov/bridge/steel/pubs/nhi16016.pdf

• NSBA’s Bolted Field Splices for Steel Bridge Flexural Members – Overview and Design
Examples: At the time of the writing of this Guide, the current version of this document was
Version 2.03, dated April 2020.

o https://www.aisc.org/globalassets/nsba/design-resources/bolted-field-splices-for-
steel-bridge-flexural-members.pdf

• NSBA’s NSBA Splice Microsoft Excel-based bolted field splice design spreadsheet: At the
time of the writing of this Guide, NSBA’s Splice Microsoft Excel-based bolted field splice
design program functioned in a manner consistent with the 9th Edition, 2020, of the AASHTO
LRFD Bridge Design Specifications (AASHTO LRFD BDS).

o https://www.aisc.org/nsba/design-resources/nsba-splice/

• NSBA’s LRFD Simon line-girder analysis and design program: At the time of the writing of
this Guide, NSBA’s LRFD Simon program functioned in a manner consistent with the 8th
Edition, 2017, of the AASHTO LRFD Bridge Design Specifications (AASHTO LRFD BDS).

o https://www.aisc.org/nsba/design-resources/simon/

• NSBA’s Continuous Span Standards: At the time of the writing of this Guide, NSBA’s
Continuous Span Standards were prepared using NSBA’s LRFD Simon program, version
10.12, which functioned in a manner consistent with the 8th Edition, 2017, of the AASHTO
LRFD BDS.

o https://www.aisc.org/nsba/design-resources/continuous-span-standards/

NSBA Guide to Navigating Routine Steel Bridge Design / 13


• NSBA’s Span-to-Weight Curves: At the time of the writing of this Guide, NSBA’s Span-to-
Weight curves were based on over based upon over 800 preliminary designs the NSBA has
done through the years up to 2016.

o https://www.aisc.org/nsba/design-resources/span-to-weight-curves/

• The NSBA brief guide to Skewed and Curved Steel I-Girder Bridge Fit (Executive Summary):
At the time of the writing of this Guide, NSBA had published a short executive summary on
the top of steel I-girder bridge fit.

o https://www.aisc.org/globalassets/nsba/technical-documents/skewed-curved-steel-
bridges-august-2016-summary-final.pdf

• The NSBA in-depth guide to Skewed and Curved Steel I-Girder Bridge Fit (Full White Paper):
At the time of the writing of this Guide, NSBA had published an longer, in-depth white-paper
on this topic.

o https://www.aisc.org/globalassets/nsba/technical-documents/skewed-curved-steel-
bridges-august-2016-final.pdf

• Short Span Steel Bridge Alliance’s Technical Design Resources for Short Span Steel Bridges:
At the time of the writing of this Guide, the Short Span Steel Bridge Alliance was posting
several design resources at this web page, including access to their eSPAN140 interactive web-
based preliminary design aid.

o https://www.shortspansteelbridges.org/resources/design/

• Research Council on Structural Connections (RCSC)’s Specification for Structural Joints


Using High-Strength Bolts: At the time of the writing of this Guide, the Research Council on
Structural Connections had published this specification on August 1, 2014, revised to account
for April, 2015 errata.

o https://www.boltcouncil.org/files/2014RCSCSpecification-withErrata.pdf

NSBA Guide to Navigating Routine Steel Bridge Design / 14


GENERAL FLOW OF DESIGN TASKS
Listed below are the general Design Tasks associated with the typical flow of design of a routine
steel I-girder bridge superstructure. The list of Design Tasks is presented in roughly the typical
order that they occur in the superstructure design process. However, as noted below, some topics
apply to several Design Tasks. And, of course, the process of designing a bridge typically involves
some degree of iteration; the initial results of later Design Tasks may suggest that revising part of
the design which occurred earlier in the process might be beneficial. When iterating through a
design in this manner, the designer is reminded that all steps of the design process should be
checked to see if the revision of one part of the design might affect other parts. Each task/topic
below is hyperlinked to its associated Design Task Quick Links page.
General Flow of Design Tasks:
1. General Considerations

2. Deck Design

3. Resistance Factors and Load Modifiers

4. Load Combinations and Load Factors

5. Live Load Force Effects - Introduction

6. Live Load Force Effects - Flexure

7. Live Load Force Effects - Shear

8. Other Load Effects and Factors Affecting Load Effect Calculations

9. Girder Flexure Design – General

10. Girder Flexure Design – Constructibility

11. Girder Flexure Design – Service Limit State

12. Girder Flexure Design – Fatigue and Fracture Limit State

13. Girder Flexure Design – Strength Limit State

14. Girder Shear Design

15. Stiffener Design

16. Shear Connector Design

17. Splice Design

18. Cross-Frame/Diaphragm Design

Topics Which May Apply to Several Design Tasks:


• Bolted Connection Design

• Welded Connection Design

• Connection Design – Miscellaneous Checks

NSBA Guide to Navigating Routine Steel Bridge Design / 15


GRAPHICAL INDEX OF DESIGN TASKS
The general Design Tasks associated with the typical steps in the design of a routine steel I-girder
bridge superstructure are grouped graphically below. Each task/topic is hyperlinked to its
associated Design Task Quick Links page.

General Considerations
General considerations prior to beginning detailed superstructure design include understanding
the LRFD design philosophy and the concept of limit states design, selecting basic design
parameters such as target girder depth and spacing, identifying superstructure materials, and
deciding whether or not to make the deck composite with the girders. The following Design
Tasks apply – each task is hyperlinked to its associated Design Task Quick Links page.
• General Considerations

Deck Design
The design of decks for routine steel I-girder bridges is beyond the scope of this Guide; see
the Owner-agencies design policy manual for standard deck designs or guidance on acceptable
deck design methods, or design the deck per the provisions of Chapter 9 of the AASHTO
LRFD BDS. The following Design Tasks apply – each task is hyperlinked to its associated
Design Task Quick Links page.
• Deck Design

NSBA Guide to Navigating Routine Steel Bridge Design / 16


Loads
The identification and calculation of various load effects is typically accomplished early in the
design process. It is difficult to design the superstructure of a routine steel I-girder bridge if
the applicable loads are not known. The process of determining those loads begins with
identification of the applicable limit states, definition of their associated load combinations,
and quantification of the various load modifiers and load factors. Resistance factors are
typically identified and quantified as well. Then the specific loading effects are calculated.
The following Design Tasks apply – each task is hyperlinked to its associated Design Task
Quick Links page.
• Resistance Factors and Load Modifiers
• Load Combinations and Load Factors
• Live Load Force Effects - Introduction
• Live Load Force Effects - Flexure
• Live Load Force Effects - Shear
• Other Load Effects and Factors Affecting Load Effect Calculations

NSBA Guide to Navigating Routine Steel Bridge Design / 17


Girder Design
Once the loads are defined, girder design follows. Various limit states must be addressed in
girder flexure design, and the design must reflect the noncomposite or composite nature of the
superstructure at the time each load is applied.
Generally, it is most efficient to perform the flexural design first, and then design for shear
afterwards. The following Design Tasks apply to girder flexure design – each task is
hyperlinked to its associated Design Task Quick Links page.
• Girder Flexure Design – General
• Girder Flexure Design – Constructibility
• Girder Flexure Design – Service Limit State
• Girder Flexure Design – Fatigue and Fracture Limit State
• Girder Flexure Design – Strength Limit State
Once the initial flexure design is completed, shear design of the web follows. It may be
appropriate or necessary to iterate back through more than one cycle of flexure design and
shear design. The following Design Task applies to girder shear design – the task is
hyperlinked to its associated Design Task Quick Links page.
• Girder Shear Design

NSBA Guide to Navigating Routine Steel Bridge Design / 18


Design of Details and Bracing
Once the basic girder design is established, design of details and bracing can begin. The design
of several details is associated directly with the girders, including stiffener design, shear
connector design, and bolted field splice design. The following Design Tasks apply to the
design of girder-related details – each task is hyperlinked to its associated Design Task Quick
Links page.
• Stiffener Design
• Shear Connector Design
• Splice Design
Next the bracing members (cross-frames or diaphragms) can be designed. The following
Design Task applies to bracing design – the task is hyperlinked to its associated Design Task
Quick Links page.
• Cross-Frame/Diaphragm Design

Connection Design Topics


Several design topics related to connection design are applicable to one or more Design Tasks.
These topics are grouped here for convenience. The following Design Tasks apply to these
connection design topics – each task is hyperlinked to its associated Design Task Quick Links
page.
• Bolted Connection Design
• Welded Connection Design
• Connection Design – Miscellaneous Checks

NSBA Guide to Navigating Routine Steel Bridge Design / 19


DESIGN TASK QUICK LINKS
The design of a routine steel I-girder bridge can be broken down into several tasks, each one quite
manageable. Each of these tasks can be made even easier when the designer has access to three
things:
• Quick Links to applicable AASHTO LRFD BDS provisions, with Discussion
o Clicking on the AASHTO LRFD BDS Article number (in parenthesis) will take the
reader directly to the Discussion of that Article
• Quick Links to helpful industry design guidelines, references, and examples
o Clicking on the hyperlink for the given reference will take the reader directly to a
free copy of that reference on the Internet
• Quick Links to useful tools
o Clicking on the hyperlink to the NSBA Simon program or the NSBA Splice
spreadsheet will take the reader directly to the page on the NSBA website where
these tools can be downloaded for free.
The Quick Links to applicable AASHTO LRFD BDS provisions include hyperlinks to many, but
not all, of the Articles for which there are Discussions in this Guide. When looking for the
Discussion of a particular Article not cited in the Quick Links, the reader can use the “bookmarks”
in the PDF version of the Guide; each Article’s Discussion is bookmarked.

NSBA Guide to Navigating Routine Steel Bridge Design / 20


GENERAL CONSIDERATIONS
Quick links to applicable AASHTO LRFD BDS provisions, with Discussion
General considerations prior to beginning detailed superstructure design include:
• Understanding the LRFD Design Philosophy (1.3.1)
• Understanding LRFD Limit States (1.3.2.1, 1.3.2.2, 1.3.2.3, 1.3.2.4, 1.3.2.5)
• Selecting Composite (6.10.1.1) or Noncomposite (6.10.1.2) Design– Routine steel I-girder bridges are
composite.
• Selecting Hybrid or Nonhybrid (6.10.1.3) Design – Routine steel I-girder bridges are nonhybrid
• Selecting Constant or Variable Web Depth (6.10.1.4) – Routine steel I-girder bridges use constant web
depth.
• Optional criteria for span-to-depth ratios (2.5.2.6.3)
• Cross-section Proportion Limits (6.10.2.1.1, 6.10.2.2)

Quick links to helpful industry design guidelines, references, and examples


For further background and explanation of general considerations, see:
• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures
o Sections 1.2 (LRFD Design Philosophy), 1.3 (Limit States), 6.3.3.2 (Girder Depth), 6.3.4 (I-Girder
Design and Sizing), 6.3.4.4.5 (Sizing Flanges for Efficient Fabrication), 6.4.2.3.2 (Sections in
Positive Flexure), 6.4.2.3.3 (Sections in Negative Flexure), 6.4.2.4.1 (Steel Girder), 6.4.2.4.2
(Concrete Deck), and pages 6.194 and 6.195 (Effects of Creep and Shrinkage)
• FHWA’s Steel Bridge Design Handbook
o Volume 10 - Limit States
o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge
• The AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to Design for
Constructability and Fabrication
o Including: Section 1.5 and Tables 1.4.1.A, 1.4.2.A, and 1.5.2.A

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website and is also a valuable tool for the design of routine steel I-girder bridges. It calculates loads and resistances in
accordance with the provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. It contains a useful web-depth optimization option that automatically generates a series of trial-design input
files from an acceptable starting design input file. Other commercial software packages with the ability to analyze and
design routine steel I-girder bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 21


DECK DESIGN

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design Conventionally Reinforced Concrete Deck – Deck design is beyond the scope of this guide; see Owner-
Agency policy manuals for standard deck design if available, or design the deck per chapter 9 of the AASHTO LRFD
BDS.

Quick links to helpful industry design guidelines, references, and examples


For further background and explanation of deck design, see:
• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures
o Sections 7 (Decks and Deck Systems), and specifically Section 7.3 (Concrete Deck Slabs), 7.3.1
(General), 7.3.2 (Traditional Design Method), 7.3.3 (Empirical Design Method), and 7.3.4 (Deck
Overhang Design).

• FHWA’s Steel Bridge Design Handbook

o Volume 17 – Bridge Deck Design

Quick links to useful tools


N/A

NSBA Guide to Navigating Routine Steel Bridge Design / 22


RESISTANCE FACTORS AND LOAD MODIFIERS

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Select resistance factors for:

• Strength limit state (6.5.4.2)

Select load modifiers for:

• Ductility (1.3.3)

• Redundancy (1.3.4)

• Operational Importance (1.3.5)

Quick links to helpful industry design guidelines, references, and examples


For further background and explanation of resistance factors and load modifiers, see:
• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures
o Sections 1.2 (LRFD Design Philosophy) and 1.3 (Limit States)
• FHWA’s Steel Bridge Design Handbook
o Volume 7 - Loads and Load Combinations
o Volume 10 - Limit States

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website and is also a valuable tool for the design of routine steel I-girder bridges. It calculates loads and resistances in
accordance with the provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-girder bridges are
also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use, including verifying and/or modifying the resistance factors and load modifiers as appropriate.

NSBA Guide to Navigating Routine Steel Bridge Design / 23


LOAD COMBINATIONS AND LOAD FACTORS

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


• Select load combinations and load factors (3.4.1) for the following limit states:

o Strength limit state (6.5.4.1, 6.10.6.1)

o Service limit state (6.5.2, 6.10.4.2.1)

o Fatigue and fracture limit state (6.5.3)

o Constructibility (3.4.2.1, 3.4.2.2, , 6.10.3.1)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of the determination of load combinations and load factors, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Section 1.3 (Limit States) and Chapter 3 (Loads and Load Factors)

• FHWA’s Steel Bridge Design Handbook

o Volume 7 - Loads and Load Combinations


o Volume 10 - Limit States

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website and is also a valuable tool for the design of routine steel I-girder bridges. It calculates loads and resistances in
accordance with the provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-girder bridges are
also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use, including verifying and/or modifying the load combinations and load factors as appropriate.

NSBA Guide to Navigating Routine Steel Bridge Design / 24


LIVE LOAD FORCE EFFECTS - INTRODUCTION

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Calculate live load force effects, which includes consideration of:

• Live Loads (3.6.1.2.1, 3.6.1.2.2, 3.6.1.2.3, 3.6.1.2.4, 3.6.1.3.1, 3.6.1.3.2, 3.6.1.3.3, 3.6.1.4.1, 3.6.1.4.2,
3.6.1.6)

• Number of Lanes (3.6.1.1.1)

• Multiple Presence (3.6.1.1.2)

• Dynamic Load Allowance (3.6.2)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of the determination of live load force effects, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures (2015)

o Sections 3.4 (Live Loads) and 4.4 (Live Load)

• FHWA’s Steel Bridge Design Handbook

o Volume 7 - Loads and Load Combinations

o Volume 8 - Structural Analysis

o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge


o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website and is also a valuable tool for the design of routine steel I-girder bridges. It calculates loads and resistances in
accordance with the provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Simon automatically calculates force effects on girders due to standard AASHTO LRFD vehicular live load.
Other commercial software packages with the ability to analyze and design routine steel I-girder bridges are also
available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 25


LIVE LOAD FORCE EFFECTS - FLEXURE

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


• Determine distribution factors for moment, considering:

o Interior beams with concrete decks (4.6.2.2.2b)

o Exterior beams (4.6.2.2.2d)

o Skewed bridges (4.6.2.2.2e)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of the determination of live load force effects with regards to flexure, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 4.4.1 (General), 4.4.2 (Live Load Distribution Factors), 4.4.3 (Influence Lines and
Influence Surfaces)

• FHWA’s Steel Bridge Design Handbook

o Volume 7 - Loads and Load Combinations

o Volume 8 - Structural Analysis

o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge


o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It can automatically calculate the live load
distribution factors necessary for the analysis, greatly reducing the time and effort required of the designer. Other
commercial software packages with the ability to analyze and design routine steel I-girder bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 26


LIVE LOAD FORCE EFFECTS - SHEAR

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


• Determine distribution factors for shear, including consideration of:

o Interior Beams (4.6.2.2.3a)

o Exterior Beams (4.6.2.2.3b)

o Skewed Bridges (4.6.2.2.3c)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of the determination of live load force effects with regards to shear, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 4.4.1 (General), 4.4.2 (Live Load Distribution Factors), 4.4.3 (Influence Lines and
Influence Surfaces)

• FHWA’s Steel Bridge Design Handbook

o Volume 7 - Loads and Load Combinations

o Volume 8 - Structural Analysis

o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge


o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It can automatically calculate the live load
distribution factors necessary for the analysis, greatly reducing the time and effort required of the designer. Other
commercial software packages with the ability to analyze and design routine steel I-girder bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 27


OTHER LOAD EFFECTS AND FACTORS AFFECTING LOAD EFFECT
CALCULATIONS

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Consider other loads (beyond gravity loads such as dead load and live load) which may affect the design of routine
steel I-girder bridges. Also consider the effects of the composite concrete deck on the distribution of moment and
shear in multispan continuous bridges.

• Consider the effects of the composite concrete deck on the stiffness of multispan continuous bridges
(6.10.1.5)

• Consider the effects of wind loading, including flange lateral bending effects (4.6.2.7.1)

• Consider seismic loads (6.16.1, 6.16.3, 3.10.9.2, 4.7.4.1, 4.7.4.2, 4.7.4.3, 4.7.4.4)

• Calculate force effects from other loads such as construction loads (6.10.3.1)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of the determination of other load effects and factors affecting load effect
calculation, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 3.3 (Construction Loads), 3.5 (Wind Loads), 3.6 (Seismic Loads), 6.5.3 (LRFD
Constructibility Design), 6.5.6.5.1 (Wind Loads on I-Sections)

• FHWA’s Steel Bridge Design Handbook

o Volume 7 - Loads and Load Combinations

o Volume 8 - Structural Analysis

o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge


o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It calculates the section properties for the
stiffness analysis in accordance with the provisions of the AASHTO LRFD BDS, greatly reducing the time and effort
required of the designer. Other commercial software packages with the ability to analyze and design routine steel I-girder
bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 28


GIRDER FLEXURE DESIGN – GENERAL

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design girders for flexure, considering the following general topics:

• Composite Section Stresses (6.10.1.1.1a, 6.10.1.1.1b, 6.10.1.1.1c, 6.10.1.1.1d, 6.10.1.1.1e)

• Flange Stresses and Member Bending Moments (6.10.1.6)

• Fundamental Section Properties (D6.1, D6.2.1, D6.2.2, D6.2.3, D6.3.1, D6.3.2)

• Materials (6.4)

• Material Thickness (6.7.3)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of flexure design, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 6.4.5.2 (Plastic Moment), 6.4.5.3 (Yield Moment), 6.4.5.4.1 (Depth of Web in
Compression in the Elastic Range), 6.4.5.4.2 (Depth of Web in Compression at the Plastic
Moment), and 6.5.2 (LRFD Flexural Design Resistance Equations)

• FHWA’s Steel Bridge Design Handbook

o Volume 1 – Bridge Steels and Their Mechanical Properties


o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

• The Reference Manual for NHI Course 130102, Engineering for Structural Stability in Bridge Construction

In addition, sanity check initial design results by comparing them to NSBA’s Span-to-Weight Curves

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It calculates the stresses in the section in
accordance with the provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. NOTE that the Simone software currently does not include the capability to design the girders using the
provisions of Appendix A6 to account for the ability of certain compact and noncompact web I-sections to develop
flexural resistances significantly greater than the yield moment, My. Other commercial software packages with the ability
to analyze and design routine steel I-girder bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 29


GIRDER FLEXURE DESIGN – CONSTRUCTIBILITY

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design girders for flexure with regards to constructibility, considering the following:

• Constructibility (6.10.3.1, 6.5.4.1), Flowchart (C6.4.1)

• Flexure (6.10.3.2, 6.10.1.8, 6.10.1.9, 6.10.1.10.1, 6.10.8.2, A6.3.3—optional)

• Shear (6.10.3.3)

• Deck placement (6.10.3.4)

• Dead load deflections (6.10.3.5)

• Tension flanges with holes (6.10.1.8)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of flexure design with regards to constructibility, see:
• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures
o Sections 1.3 (Limit States), 6.4.5.5 (Web Bend Buckling Resistance), 6.5.3 (LRFD Constructibility
Design), and 6.5.6 (LRFD Strength Limit State for Flexure)
• FHWA’s Steel Bridge Design Handbook
o Volume 10 - Limit States
o Volume 11 - Design for Constructability
o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge
• The AASHTO-NSBA Steel Bridge Collaboration Guidelines
o G12.1-2020 Guidelines to Design for Constructability and Fabrication
In addition, sanity check initial design results by comparing them to NSBA’s Span-to-Weight Curves

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It calculates the design loads and resulting
stresses, and the corresponding resistances in accordance with the provisions of the AASHTO LRFD BDS, including
the constructibility checks of Article 6.10.3, greatly reducing the time and effort required of the designer. Other
commercial software packages with the ability to analyze and design routine steel I-girder bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 30


GIRDER FLEXURE DESIGN – SERVICE LIMIT STATE

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design girders to meet the requirements of the service limit state for flexure, considering the following:

• Service limit state (6.5.2, 6.10.4), Flowchart (C6.4.2)

• Optional live-load deflection limits (2.5.2.6.2) – coordinate with Owner-agency policy

• Elastic deformations (6.10.4.1)

• Permanent deformations (6.10.4.2, 6.10.4.2.1, 6.10.4.2.2)

• Flexure (6.10.1.9.1, 6.10.1.10.1, 6.10.1.10.2)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of flexure design with regards to service limit state, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 1.3 (Limit States), 6.4.5.5 (Web Bend Buckling Resistance), and 6.5.4 (LRFD Service
Limit State Design)

• FHWA’s Steel Bridge Design Handbook

o Volume 10 - Limit States

o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge


o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge

o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

In addition, sanity check initial design results by comparing them to NSBA’s Span-to-Weight Curves

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It calculates the design loads and resulting
stresses, and the corresponding resistances, in accordance with the provisions of the AASHTO LRFD BDS, including
the service limit state checks of Article 6.10.4, greatly reducing the time and effort required of the designer. Other
commercial software packages with the ability to analyze and design routine steel I-girder bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 31


GIRDER FLEXURE DESIGN – FATIGUE AND FRACTURE LIMIT STATE

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design girders to meet the requirements of the fatigue and fracture limit state for flexure, considering the following:

• Fatigue and Fracture Limit State (6.5.3, 6.10.5.1, 6.10.5.2), Flowchart (C6.4.3)

• Fatigue (6.6.1.1, 6.6.1.2, 6.6.1.3, 6.10.5.1)

• Fracture (6.6.2.1, 6.10.5.2)

• Special Fatigue Requirement for Webs (6.10.5.3)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of flexure design with regards to fatigue and fracture limit state, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 1.3 (Limit States), 6.5.5 (LRFD Fatigue and Fracture Limit State Design), and 6.6.2 (Shear
Connectors)

• FHWA’s Steel Bridge Design Handbook

o Volume 10 - Limit States

o Volume 12 - Design for Fatigue


o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge
o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge

o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

• The AASHTO-NSBA Steel Bridge Collaboration Guidelines

o G1.4-2006 Guidelines for Design Details

o G12.1-2020 Guidelines for Design for Constructabilty

In addition, sanity check initial design results by comparing them to NSBA’s Span-to-Weight Curves

• links to useful tools


Quick
NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It calculates the design loads and resulting
stresses, and the corresponding resistances, in accordance with the provisions of the AASHTO LRFD BDS, greatly
reducing the time and effort required of the designer. It will calculate the fatigue stress range for either the Fatigue I or
Fatigue II limit-state load combination at multiple points along the length of the girder and compare it to the nominal
fatigue resistance. Other commercial software packages with the ability to analyze and design routine steel I-girder
bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 32


GIRDER FLEXURE DESIGN – STRENGTH LIMIT STATE

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design girders for flexure to meet the requirements of the strength limit state, considering the following:

• Strength limit state (6.5.4.1, 6.5.4.2, 6.10.6.1, 6.10.6.2.1), Flowchart (C6.4.4)

• Composite sections in positive flexure (6.10.6.2.2, 6.10.7.1.1, 6.10.7.1.2, 6.10.7.2.2, 6.10.7.3), Flowchart
(C6.4.5)

• Composite sections in negative flexure and noncomposite sections (6.10.6.2.3, 6.10.8.1.1, 6.10.8.1.2,
6.10.8.1.3, 6.10.8.2.1, 6.10.8.2.2, 6.10.8.2.3, 6.10.8.3), Flowchart (C6.4.6) (APPENDIX A6—optional),
Flowchart (C6.4.7−optional) (D6.4—optional)

• Tension flanges with holes (6.10.1.8)

• Flange-strength Reduction Factors (6.10.1.10.1, 6.10.1.10.2)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of flexure design with regards to strength limit state, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 1.3 (Limit States) and 6.5.6 (LRFD Strength Limit State Design for Flexure)

• FHWA’s Steel Bridge Design Handbook

o Volume 10 - Limit States

o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge


o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge

o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

In addition, sanity check initial design results by comparing them to NSBA’s Span-to-Weight Curves

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It calculates the design loads and resulting
stresses, and the corresponding resistances, in accordance with the provisions of the AASHTO LRFD BDS, including
the strength limit state checks of Article 6.10.6, greatly reducing the time and effort required of the designer. Other
commercial software packages with the ability to analyze and design routine steel I-girder bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 33


GIRDER SHEAR DESIGN

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design girders for shear to meet the requirements of the strength limit state, considering the following:

• General provisions (6.10.9.1), Flowchart (Figure C6.10.9.1-1)

• Nominal resistance of unstiffened webs (6.10.9.2)

• Nominal resistance of stiffened webs

o General provisions (6.10.9.3.1)

o Nominal resistance of interior panels (6.10.9.3.2)

o Nominal resistance of end panels (6.10.9.3.3)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of girder shear design, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Section 6.5.7 (LRFD Strength Limit State Design for Shear)

• FHWA’s Steel Bridge Design Handbook

o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge


o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge

o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

• The AASHTO-NSBA Steel Bridge Collaboration Guidelines

o G12.1-2020 Guidelines for Design for Constructabilty

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It calculates the design shear loads, and the
corresponding shear resistances, in accordance with the provisions of the AASHTO LRFD BDS, greatly reducing the
time and effort required of the designer. Other commercial software packages with the ability to analyze and design
routine steel I-girder bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 34


STIFFENER DESIGN

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design stiffeners, considering the following:

• Web transverse stiffener design provisions (6.10.11.1.1, 6.10.11.1.2, 6.10.11.1.3)

• Bearing stiffener design provisions (6.10.11.2.1, 6.10.11.2.2, 6.10.11.2.3, 6.10.11.2.4a, 6.10.11.2.4b)

• Provisions for concentrated loads applied to webs without bearing stiffeners (D6.5.1, D6.5.2, D6.5.3)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of stiffener design, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 6.6.6.2 (Transverse Web Stiffeners), 6.6.6.3 (Bearing Stiffeners)

• FHWA’s Steel Bridge Design Handbook

o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge


o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge

• The AASHTO-NSBA Steel Bridge Collaboration Guidelines

o G1.4-2006 Guidelines for Design Details

o G12.1-2020 Guidelines to Design for Constructability and Fabrication

• The FHWA Bridge Welding Reference Manual

o Sections 9.2.5 and 9.3.2

Quick links to useful tools


Various aspects of stiffener design are sometimes addressed in commercial line-girder analysis and design software, but
more often the calculations are performed by hand or are automated by designers in spreadsheets. However, even if the
detailed design of the stiffeners is performed by hand or spreadsheet, certain design variables such as shear demand and
resistance values or bearing reactions are still obtained from the line-girder analysis and design software. NSBA's LRFD
Simon line-girder analysis and design software is available for free download from the NSBA website is also a valuable
tool for the design of routine steel I-girder bridges. It calculates bearing reactions and shear deman and the corresponding
resistances in accordance with the provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required
of the designer. Other commercial software packages with the ability to analyze and design routine steel I-girder bridges
are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 35


SHEAR CONNECTOR DESIGN

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design shear connectors for the fatigue and strength limit states, considering the following:
• General provisions (6.10.10.1, 6.10.10.1.1, 6.10.10.1.2, 6.10.10.1.3, 6.10.10.1.4)
• Fatigue resistance (6.10.10.2)
• Special requirements for points of permanent load contraflexure (6.10.10.3)

• Strength limit state (6.10.10.4.1, 6.10.10.4.2, 6.10.10.4.3)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of the determination of the design of shear connectors at the fatigue and strength
limit states, see:

• The Reference Manual for NHI Course 130122, Design and Evaluation of Steel Bridges for Fatigue and
Fracture

o Section 6.3.6.3 (Shear Studs)

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Section 6.6.2 (Shear Connectors)

• FHWA’s Steel Bridge Design Handbook

o Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge

o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge

o Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge

Quick links to useful tools


NSBA's LRFD Simon line-girder analysis and design software. Simon is available for free download from the NSBA
website is also a valuable tool for the design of routine steel I-girder bridges. It performs design calculations addressing
the demand on, and resistance of, shear connectors at the fatigue and strength limit states in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the designer. Other commercial
software packages with the ability to analyze and design routine steel I-girder bridges are also available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 36


SPLICE DESIGN

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design field splices (if present), considering the following:

• Bolted field splices of flexural members

o General considerations (6.13.6.1.3a)

o Flange splices (6.13.6.1.3b)

o Web splices (6.13.6.1.3c)

• Welded splices (6.13.6.2)

• Minimum thickness requirements (6.7.3)

Determine flange sizes and locations of welded shop splices, considering the following:

• Welded splices (6.13.6.2)

• Minimum thickness requirements (6.7.3)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of field splice design, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 6.6.5 (Splices), especially 6.6.5.2 (Flexural Members) (NOTE: The explanations in these
references are written in the context of the bolted field splice provisions prior to publication of the
8th Edition of the AASHTO LRFD BDS and are thus out of date).

• The AASHTO-NSBA Steel Bridge Collaboration Guidelines G12.1-2020 Guidelines to Design for
Constructability and Fabrication

o Section 1.5.3 (Flange Plate Width) and Table 1.5.2.A, Section 2.2.1 (Field Connections)

• NSBA’s Bolted Field Splices for Steel Bridge Flexural Members – Overview and Design Examples

Quick links to useful tools


The NSBA Splice Microsoft Excel-based bolted field splice design spreadsheet is available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It performs the design of a bolted
field splice for a steel I-girder in accordance with the provisions of Article 6.13.6.1.3, greatly reducing the time and effort
required of the designer. Other commercial software packages with the ability to design bolted field splices are also
available.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 37


CROSS-FRAME/DIAPHRAGM DESIGN

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design cross-frames or diaphragms, considering the following:

• General considerations (6.7.4.1, 6.7.4.2), minimum thicknesses (6.7.3)

• Design for tension (6.8.1, 6.8.2.1, 6.8.2.2, 6.8.3, 6.8.4)

• Design for compression (6.9.1, 6.9.2.1, 6.9.2.2.1, 6.9.3, 6.9.4.1.1, 6.9.4.1.2, 6.9.4.1.3, 6.9.4.2.1, 6.9.4.2.2a,
6.9.4.2.2b, 6.9.4.4)

• Design considerations for miscellaneous flexural members (6.12.1.1, 6.12.1.2.1, 6.12.1.2.2, 6.12.1.2.3a,
6.12.2.1, 6.12.2.2.4a, 6.12.2.2.4b, 6.12.2.2.4c, 6.12.2.2.4d, 6.12.2.2.4e, 6.12.2.2.5)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of the determination of the design of shear connectors at the
fatigue and strength limit states, see:
• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures
o Section 6.6.3 (Bracing Member Design)
• FHWA’s Steel Bridge Design Handbook
o Volume 13 – Bracing System Design
o Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge
• The AASHTO-NSBA Steel Bridge Collaboration Guidelines
o G1.4-2006 Guidelines for Design Details
o G12.1-2020 Guidelines to Design for Constructability and Fabrication
• Coletti, D.A., and M.A. Grubb, “Practical Implementation of Stability Bracing Strength and Stiffness
Guidelines for Steel I-Girder Bridges,”
• AISC’s Specifications for Structural Steel Buildings and Commentary
o Article D3, Table D3.1, and the Commentary for Article D3
• AISC’s Database of Rolled Steel Shape Section Properties

Quick links to useful tools


The calculations associated with cross-frame and diaphragm design for routine steel I-girder bridges are typically
performed by hand or in spreadsheets. Some commercial bridge design software packages offer some capabilities
associated with cross-frame or diaphragm design, but those capabilities are typically limited to refined analysis
software packages, not line girder analysis and design packages.

Users should verify the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

NSBA Guide to Navigating Routine Steel Bridge Design / 38


BOLTED CONNECTION DESIGN

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design bolted connections for cross-frames or diaphragms and for bolted field splices, considering the following (also
see the Quick Links for field splice design):

• General provisions (6.13.1, 6.13.2.1)

• Bolt, nut, washer, and bolt hole provisions (6.13.2.3.1, 6.13.2.3.2, 6.13.2.4.1a, 6.13.2.4.1b, 6.13.2.4.1c,
6.13.2.4.1d, 6.13.2.4.2, 6.13.2.5)

• Bolt spacing, edge and end distances (6.13.2.6.1, 6.13.2.6.2, 6.13.2.6.3, 6.13.2.6.4, 6.13.2.6.5, 6.13.2.6.6)

• Net area (6.8.3)

• Factored resistance of bolted connections (6.13.2.2)

• Slip critical bolt resistance (6.13.2.1.1) (6.13.2.8)

• Bearing connections (6.13.2.1.2), bolt shear resistance (6.13.2.7), bearing resistance at bolt holes (6.13.2.9)

• Bolt tensile resistance (6.13.2.10.1, 6.13.2.10.2, 6.13.2.10.3, 6.13.2.10.4, 6.13.2.11)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of bolted connection design, see:
• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures
o Sections 6.6.3.2 (Stability Bracing Requirements), 6.6.3.3.2.3 (Net Area), 6.6.4.2 (Bolted
Connections) and associated subsections

• The AISC Design Guide 17 High Strength Bolts - A Primer for Engineers

• The RCSC Specifications for Structural Joints Using High-Strength Bolts

• The AASHTO-NSBA Steel Bridge Collaboration Guide Specification

o S2.1-2018 Steel Bridge Fabrication Guide Specification

o G12.1-2020 Guidelines to Design for Constructability and Fabrication

Quick links to useful tools


The calculations associated with most bolted connection designs (including cross-frame bolted connections and
general bolted connections) are typically performed by hand or in spreadsheets.

See also the Design Task Quick Links for Field Splice Design for reference to the NSBA Splice spreadsheet.

NSBA Guide to Navigating Routine Steel Bridge Design / 39


WELDED CONNECTION DESIGN

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Design welded connections considering the following:

• General provisions (6.13.3.16.13.3)

• Factored resistance (6.13.3.2.1)

• Complete joint penetration welded connections (6.13.3.2.2a, 6.13.3.2.2b)

• Partial penetration groove-welded connections (6.13.3.2.3a, 6.13.3.2.3b)

• Fillet-welded connections (6.13.3.2.4, 6.13.3.4, 6.13.3.5, 6.13.3.6, 6.13.3.7)

• Effective area (6.13.3.3)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of weld connection design, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 6.6.4.3 (Welded Connections)

• The AASHTO-NSBA Steel Bridge Collaboration Guidelines

o G1.4-2006 Guidelines for Design Details

o G12.1-2020 Guidelines to Design for Constructability and Fabrication

• The FHWA Bridge Welding Reference Manual

Quick links to useful tools


The calculations associated with welded connection design are typically performed by hand or in spreadsheets.

NSBA Guide to Navigating Routine Steel Bridge Design / 40


CONNECTION DESIGN – MISCELLANEOUS CHECKS

Quick links to applicable AASHTO LRFD BDS provisions, with Discussion


Evaluate all connections (including elements in bolted and welded connections such as splice plates, gusset plates,
brackets, etc.) for the following considerations:

• Block shear rupture resistance (6.13.4)

• Connection elements - tension (6.13.5.2)

• Connection elements - shear (6.13.5.3)

Quick links to helpful industry design guidelines, references, and examples


For more explanation and examples of miscellaneous checks of connection design, see:

• The Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway
Bridge Superstructures

o Sections 6.6.3.3.2.5 (Block Shear Rupture Resistance), 6.6.4.2.5.6.1 (Tensile Resistance of a


Connected Element), and 6.6.4.2.5.6.2 (Shear Resistance of a Connected Element)

• NSBA’s Bolted Field Splices for Steel Bridge Flexural Members – Overview and Design Examples

Quick links to useful tools


The calculations associated with miscellaneous connection design are typically performed by hand or in spreadsheets.

See also the Design Task Quick Links for Field Splice Design for reference to the NSBA Splice spreadsheet.

NSBA Guide to Navigating Routine Steel Bridge Design / 41


SECTION 1: INTRODUCTION

TABLE OF CONTENTS

1.1 SCOPE OF THE SPECIFICATIONS ............................................................................. 43


1.2 DEFINITIONS ................................................................................................................ 43
1.3 DESIGN PHILOSOPHY ................................................................................................ 43
1.3.1 General .................................................................................................................. 43
1.3.2 Limit States ........................................................................................................... 43
1.3.2.1 General ............................................................................................................ 43
1.3.2.2 Service Limit State.......................................................................................... 44
1.3.2.3 Fatigue and Fracture Limit State..................................................................... 45
1.3.2.4 Strength Limit State ........................................................................................ 45
1.3.2.5 Extreme Event Limit State .............................................................................. 45
1.3.3 Ductility ................................................................................................................ 46
1.3.4 Redundancy........................................................................................................... 46
1.3.5 Operational Importance ........................................................................................ 46
1.4 REFERENCES ................................................................................................................ 47

NSBA Guide to Navigating Routine Steel Bridge Design / 42


1.1 SCOPE OF THE SPECIFICATIONS

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
This Article discusses the overall scope of the AASHTO LRFD BDS, and from that perspective is
fairly self-explanatory. Key guidance includes the statement, “These Specifications are not
intended to supplant proper training or the exercise of judgment by the Designer, and state only
the minimum requirements necessary to provide for public safety.” In a similar vein, this “Guide
for streamlined design of routine steel I-girder bridges” is also only another tool. Ultimately, the
design of any bridge should be performed, or at least directly supervised, by engineers who are
experienced and qualified to do the work. No amount of specifications, commentary, guidelines,
or pre-packaged design software can take the place of proper training, experience, and oversight.
The Article continues by discussing basic concepts upon which the AASHTO LRFD BDS are
based, and providing references to associated specifications which cover topics not directly
addressed in the AASHTO LRFD BDS.
The Commentary for this Article provides important definitions of the terms “notional,” “shall,”
“should,” and “may,” which have specific connotations throughout the specifications.

1.2 DEFINITIONS

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
The terms listed are either directly applicable to, or help define provisions which are not applicable
to, the design of routine steel I-girder bridges.

1.3 DESIGN PHILOSOPHY

1.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The design philosophy of the AASHTO LRFD BDS encompasses the design of routine steel I-
girder bridges, among other structures.

1.3.2 Limit States

1.3.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 43


Discussion:
The basic load and resistance factor design (LRFD) equation and the associated equations in this
Article limiting the combined values of the load modifiers apply to all bridges, including the
routine steel I-girder bridges covered by this Guide. The basic LRFD equation is intended to check
that the force effects caused by factored loads do not exceed the factored resistance of the
component under consideration. The specified load and resistance factors are statistically based to
provide a targeted level of reliability, or probability of exceedance of a given limit state, over the
75-year design life of the bridge. A resistance factor of 1.0 is typically applied to the nominal
resistance at all non-strength limit states unless otherwise specified.
The load modifier in the basic equation is a factor that is used in a subjective fashion to account
for the ductility, redundancy, and operational classification of the bridge (see the Discussion of
Articles 1.3.3, 1.3.4, and 1.3.5 in this Guide). For a routine steel I-girder bridge, the load modifiers
for ductility and redundancy are only applicable at the strength limit state; their value is taken as
1.0 at other limit states. Also, Eq. 1.3.2.1-3 is only applicable for the calculation of the load
modifier when dead- and live-load force effects are of opposite sign and the minimum load factor
specified in Table 3.4.1-2 is applied to the dead-load force effects (e.g., when investigating for
uplift at a support or when designing bolted field splices located near points of permanent load
contraflexure); otherwise, Eq. 1.3.2.1-2 is to be used.
A limit state is defined as a condition beyond which the bridge or component ceases to satisfy the
provisions for which it was designed. The various limit states defined in the specifications, which
are considered of equal importance, were established to allow categorization of the evaluation of
different combinations of loads (including both unfactored and factored loads) and corresponding
resistance values representing different aspects of structural performance. In general, structures
are required to provide different levels of performance for routine, frequent loading conditions
versus infrequent or extreme loading conditions.
For further background and explanation of the LRFD design philosophy and the basic LRFD
design equation, consult Section 1.2 of the Reference Manual for NHI Course 130081, Load and
Resistance Factor Design (LRFD) for Highway Bridge Superstructures. For further background
and explanation of limit states, consult the FHWA’s Steel Bridge Design Handbook, Volume 10 -
Limit States, and Section 1.3 of the Reference Manual for NHI Course 130081, Load and
Resistance Factor Design (LRFD) for Highway Bridge Superstructures.

1.3.2.2 Service Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
All steel girder bridges, including the routine steel I-girder bridges covered by this Guide, are
subject to limits on stresses and deformations under service limit state loading conditions, which
generally refers to normal operational use of the bridge under service conditions. The load factors
for the service limit state are generally (but not always) set at 1.0.

NSBA Guide to Navigating Routine Steel Bridge Design / 44


Steel bridges are not subject to specific limits on crack widths, which in the context of this Article
refer to the width of cracks in concrete structures, not fatigue cracks in steel structures.
For further background and explanation of the service limit state, consult the FHWA’s Steel Bridge
Design Handbook, Volume 10 - Limit States, and Section 1.3 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.

1.3.2.3 Fatigue and Fracture Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
Fatigue is defined in Section 6 of the AASHTO LRFD BDS as, “The initiation and/or propagation
of cracks due to a repeated variation of normal stress with a tensile component.” Fracture is the
partial or total severing of an element under the action of force, particularly a tensile force. Fatigue
and fracture are phenomena which can easily occur in an improperly designed or detailed steel
structure. The fatigue and fracture limit state is directly applicable to the design of all steel bridges,
including the routine steel I-girder bridges covered by this Guide.
The discussion of cracks in the commentary refers to fatigue cracks in steel structures, not tension
or shear cracks in concrete structures.
For further background and explanation of the fatigue and fracture limit state, consult the FHWA’s
Steel Bridge Design Handbook, Volume 10 - Limit States, and Section 1.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures.

1.3.2.4 Strength Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The strength limit state addresses the investigation of stability and/or yielding of structural
elements. The strength limit state is applicable to the design of all structures, including the routine
steel I-girder bridges covered by this Guide.
For further background and explanation of the strength limit state, consult the FHWA’s Steel
Bridge Design Handbook, Volume 10 - Limit States, and Section 1.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures.

1.3.2.5 Extreme Event Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.

NSBA Guide to Navigating Routine Steel Bridge Design / 45


Discussion:
The extreme event limit state applies primarily to substructure design, which is excluded from the
scope of this Guide. The rare cases of applicability of the extreme event limit state to the design
of steel I-girder bridge superstructures are limited to situations which have been excluded from the
definition given herein of a routine steel I-girder bridge.

1.3.3 Ductility
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The structural members and details used in the design of routine steel I-girder bridge
superstructures are specifically designed and configured to inherently exhibit ductile behavior, so
the specified value of 1.0 for the load modifier for ductility for conventional designs and details
should always be used for the design of routine steel I-girder bridge superstructures.
The use of a load modifier for ductility with a value other than 1.0 may be appropriate for design
of other elements in a given bridge, such as the substructure. In such cases, designers are advised
to carefully identify and differentiate elements which are subject to the application of such a
ductility load modifier in their design.

1.3.4 Redundancy
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The configuration, associated structural member designs, and details of a routine steel I-girder
bridge, as defined for the purposes of this Guide, inherently provide multiple, redundant load paths,
in both simple span and multiple-span continuous bridges. Therefore, the specified value of 1.0
for the load modifier for redundancy for conventional levels of redundancy should always be used
for the design of routine steel I-girder bridge superstructures.

1.3.5 Operational Importance


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The value of the load modifier for operational importance should be chosen based on a careful
evaluation of a given bridge in the larger context of the transportation network in which it
functions. Generally, this evaluation is performed by the Owner-agency or their designated
representative. Alternately, the Owner-agency or their designated representative may provide
explicit guidance on how to perform such an evaluation for bridges within their transportation
network. In either case, the value of the load modifier for operational importance should be chosen
with input from the Owner-agency. In the absence of such input, the load modifier for operational
importance should be taken as 1.0.

NSBA Guide to Navigating Routine Steel Bridge Design / 46


1.4 REFERENCES

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
Helpful reference material is provided for more in-depth discussion of the various provisions
within Section 1 of the AASHTO LRFD BDS.

NSBA Guide to Navigating Routine Steel Bridge Design / 47


SECTION 2: GENERAL DESIGN AND LOCATION FEATURES

TABLE OF CONTENTS
2.1 SCOPE .................................................................... not addressed in this Guide
2.2 DEFINITIONS ........................................................ not addressed in this Guide
2.3 LOCATION FEATURES ....................................... not addressed in this Guide
2.3.1 Route Location ........................................................ not addressed in this Guide
2.3.1.1 General .................................................................... not addressed in this Guide
2.3.1.2 Waterway and Floodplain Crossings ...................... not addressed in this Guide
2.3.2 Bridge Site Arrangement ........................................ not addressed in this Guide
2.3.2.1 General .................................................................... not addressed in this Guide
2.3.2.2 Traffic Safety .......................................................... not addressed in this Guide
2.3.2.2.1 Protection of Structures........................................... not addressed in this Guide
2.3.2.2.2 Protection of Users .................................................. not addressed in this Guide
2.3.2.2.3 Geometric Standards ............................................... not addressed in this Guide
2.3.2.2.4 Road Surfaces ......................................................... not addressed in this Guide
2.3.2.2.5 Vessel Collisions..................................................... not addressed in this Guide
2.3.3 Clearances ............................................................... not addressed in this Guide
2.3.3.1 Navigational ............................................................ not addressed in this Guide
2.3.3.2 Highway Vertical .................................................... not addressed in this Guide
2.3.3.3 Highway Horizontal ................................................ not addressed in this Guide
2.3.3.4 Railroad Overpass ................................................... not addressed in this Guide
2.3.4 Environment ............................................................ not addressed in this Guide
2.4 FOUNDATION INVESTIGATION ...................... not addressed in this Guide
2.4.1 General .................................................................... not addressed in this Guide
2.4.2 Topographic Studies ............................................... not addressed in this Guide
2.5 DESIGN OBJECTIVES ......................................... not addressed in this Guide
2.5.1 Safety ...................................................................... not addressed in this Guide
2.5.1.1 Structural Survival .................................................. not addressed in this Guide
2.5.1.2 Limited Serviceability ............................................. not addressed in this Guide
2.5.1.3 Immediate Use ........................................................ not addressed in this Guide
2.5.2 Serviceability .......................................................... not addressed in this Guide
2.5.2.1 Durability ................................................................ not addressed in this Guide
2.5.2.1.1 Materials ................................................................. not addressed in this Guide
2.5.2.1.2 Self-Protecting Measures ........................................ not addressed in this Guide
2.5.2.2 Inspectability ........................................................... not addressed in this Guide
2.5.2.3 Maintainability ........................................................ not addressed in this Guide
2.5.2.4 Rideability ............................................................... not addressed in this Guide
2.5.2.5 Utilities.................................................................... not addressed in this Guide
2.5.2.6 Deformations................................................................................................... 50
2.5.2.6.1 General ....................................................................................................... 50
2.5.2.6.2 Criteria for Deflection ................................................................................ 50
2.5.2.6.3 Optional Criteria for Span-to-Depth Ratios ............................................... 51
2.5.2.7 Consideration of Future Widening...........................not addressed in this Guide
2.5.2.7.1 Exterior Beams on Girder System Bridges .........not addressed in this Guide

NSBA Guide to Navigating Routine Steel Bridge Design / 48


2.5.2.7.2 Substructure .........................................................not addressed in this Guide
2.5.3 Constructibility .................................................................not addressed in this Guide
2.5.4 Economy ...........................................................................not addressed in this Guide
2.5.4.1 General .....................................................................not addressed in this Guide
2.5.4.2 Alternative Plans ......................................................not addressed in this Guide
2.5.5 Bridge Aesthetics ..............................................................not addressed in this Guide
2.6 HYDROLOGY AND HYDRAULICS ....................................not addressed in this Guide
2.6.1 General ..............................................................................not addressed in this Guide
2.6.2 Site Data ............................................................................not addressed in this Guide
2.6.3 Hydrologic Analysis .........................................................not addressed in this Guide
2.6.4 Hydraulic Analysis............................................................not addressed in this Guide
2.6.4.1 General .....................................................................not addressed in this Guide
2.6.4.2 Stream Stability........................................................not addressed in this Guide
2.6.4.3 Bridge Waterway .....................................................not addressed in this Guide
2.6.4.4 Bridge Foundations ..................................................not addressed in this Guide
2.6.4.4.1 General ................................................................not addressed in this Guide
2.6.4.4.2 Bridge Scour ........................................................not addressed in this Guide
2.6.4.5 Roadway Approaches to Bridge ..............................not addressed in this Guide
2.6.5 Culvert Location, Length, and Waterway Area ................not addressed in this Guide
2.6.6 Roadway Drainage ............................................................not addressed in this Guide
2.6.6.1 General .....................................................................not addressed in this Guide
2.6.6.2 Design Storm ...........................................................not addressed in this Guide
2.6.6.3 Type, Size, and Number of Drains...........................not addressed in this Guide
2.6.6.4 Discharge from Deck Drains ...................................not addressed in this Guide
2.6.6.5 Drainage of Structures .............................................not addressed in this Guide
2.7 BRIDGE SECURITY...............................................................not addressed in this Guide
2.7.1 General ..............................................................................not addressed in this Guide
2.7.2 Design Demand .................................................................not addressed in this Guide
2.8 REFERENCES .........................................................................not addressed in this Guide

NSBA Guide to Navigating Routine Steel Bridge Design / 49


2.5.2.6 Deformations

2.5.2.6.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
Only the first paragraph of the provision is applicable to the routine steel I-girder bridges covered
by this Guide. The remainder of the provisions in this article related to dynamic analysis and to
straight skewed and horizontally curved steel girder bridges are not applicable to these bridges.

2.5.2.6.2 Criteria for Deflection


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
These criteria are optional at the direction of the Owner-agency; review Owner-agency guidelines
in conjunction with the provisions of this Article. In the absence of Owner-agency guidelines,
design routine steel I-girder bridges to meet the applicable live load deflection limits presented in
this article for “steel, aluminum, and/or concrete vehicular bridges.” Only the provisions directly
related to straight steel I-girder bridges apply to the design of the routine steel I-girder bridges
covered by this Guide.
Designs in which the live load deflections of the superstructure exceed the limits suggested in this
Article may be at risk of experiencing adverse dynamic response (vibrations) under routine service
loading. The use of dynamic analysis to investigate and/or substantiate the dynamic performance
of a design in which the live load deflections exceed these limits is not recommended for routine
steel I-girder bridges.
Furthermore, designs which do not meet these simple live load deflection limits are generally
uneconomical, may be difficult to construct, and/or may have difficulty meeting other design
criteria.
The special live load specified in Article 3.6.1.3.2 should be used to evaluate live load deflection.
The specified load is intended to produce live load deflections similar to those produced by HS20
loading, which was the basic design live load specified in the AASHTO Standard Specifications
for Highway Bridges (see the Discussion of Article 3.6.1.3.2 in this Guide). The load factor on the
special live load is taken equal to 1.0. For the routine steel I-girder bridges covered by this Guide,
all design lanes should be loaded and the beams and girders should be assumed to deflect equally;
for multi-girder bridges, this is equivalent to saying that the distribution factor for calculating live
load deflection should be taken equal to the appropriate multiple presence factor, m, given in Table
3.6.1.1.2-1 (see the Discussion of Article 3.6.1.1.2 in this Guide) times the corresponding number
of design lanes loaded divided by the number of girders in the cross-section. The live load
deflection is typically limited to L/800 for bridges carrying vehicular loading only and L/1000 for
bridges carrying vehicular and pedestrian loading, where L is the span length in feet.
Concrete barriers and sidewalks, and even railings, often contribute to the stiffness of composite
superstructures at service load levels. Therefore, this Article permits the entire width of the

NSBA Guide to Navigating Routine Steel Bridge Design / 50


roadway and the structurally continuous portions of railings, sidewalks and barriers (i.e.,
continuous cast-in-place barriers) to be included in determining the composite stiffness for
deflection calculations. However, inclusion of concrete items other than the deck complicates the
calculation of the composite stiffness of the superstructure and is virtually never considered for
routine steel I-girder bridges. Barriers are generally located at the edges of the deck, where they
tend to stiffen and draw load to the exterior girders. Thus any beneficial stiffening of the system
tends to be counterbalanced by unequal distribution of the loading among the girders and the
associated reduction in computed deflections resulting from consideration of the barriers tends to
be negligible.
For further discussion on span-to-depth ratios and live load deflection, consult Sections 6.3.3.1
and 6.5.4.2.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating span-to-depth ratio
and live load deflection calculations, consult FHWA’s Steel Bridge Design Handbook, Design
Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge, Design Example
2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B,
Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.

2.5.2.6.3 Optional Criteria for Span-to-Depth Ratios


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
These criteria are optional at the direction of the Owner-agency; review Owner-agency guidelines
in conjunction with the provisions of this Article. In the absence of Owner-agency guidelines,
following the guidance on minimum superstructure depths presented in Table 2.5.2.6.3-1 for steel
I-beam bridges is recommended, and typically results in more economical, better-performing
designs. Only the provisions directly related to straight steel I-girder bridges apply to routine steel
I-girder bridges.
To help control elastic deformations at the service limit state, the optional span-to-depth ratios
given for a constant-depth superstructure in Table 2.5.2.6.3-1 should be met for all routine steel I-
girder bridges to establish a reasonable minimum web depth for the design in the absence of
specific depth restrictions. Shallower girders may be used when clearance limits girder depth if
permitted by the Owner-agency. For continuous spans, the suggested limits include a built-in factor
of 0.8 to reflect an effective span length based on an approximate distance within the span between
points of permanent load contraflexure. Typically, the longest span length is used to establish the
limit. For end spans, a depth-to-span ratio of 0.9 of the simple-span ratio might be considered to
better account for only one end of the span being restrained by continuity. Although the suggested
minimum depths are taken to apply to the overall depth of the steel girder, it is suggested that they
be applied to the web depth for simplicity. The greatest depth determined from the applicable
equation for each span in a continuous girder should be used. Girder depths at, or most often
exceeding, these suggested minimum depths typically provide the most economical girders. In
many cases, the optimum web depth (see the Discussion of Article 6.10.2.1.1 in this Guide) will
be significantly greater than the minimum depth based on the traditional span-to-depth ratios. The

NSBA Guide to Navigating Routine Steel Bridge Design / 51


provisions in this Article regarding the suggested minimum span-to-depth ratios for curved steel
girder systems do not apply to the routine steel I-girder bridges covered by this Guide.
As noted in the AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to
Design for Constructability and Fabrication (see commentary section C1.5.3), “Deeper girders are
generally more economical, but only up to a point. To assess overall economy, it may be valuable
to perform a web depth study where the web depth is incrementally increased, the girder is
redesigned (targeting a partially stiffened web design), and the resulting girder weight versus depth
is recorded. These data points (girder weight versus web depth) can then be plotted to determine
the optimum (minimum girder weight) web depth. Some steel girder design software packages
(e.g., NSBA's LRFD Simon line-girder analysis and design program) offer automated web depth
study features; otherwise the study can be performed by simply iterating the design with different
web depths.” Users should verify the capabilities, assumptions, and general correctness of any
program’s calculations prior to initial use. Other industry guideline documents present similar
recommendations. See the Discussions of Articles 6.10.4.1 and 6.10.2.1.1 in this Guide for further
information on the computation of span-to-depth ratios and on web-depth optimization,
respectively, for routine steel I-girder bridges.
For further information on span-to-depth ratios, consult Section 6.3.3.1 of the Reference Manual
for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples illustrating span-to-depth ratio calculations, consult the
FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous Straight
Composite Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous Straight Composite
Steel I-Girder Bridge, and Design Example 2B, Two-Span Continuous Straight Composite Steel
Wide-Flange Beam Bridge.

NSBA Guide to Navigating Routine Steel Bridge Design / 52


SECTION 3: LOADS AND LOAD FACTORS

TABLE OF CONTENTS
3.1 SCOPE..................................................................................... not addressed in this Guide
3.2 DEFINITIONS ........................................................................ not addressed in this Guide
3.3 NOTATION ............................................................................ not addressed in this Guide
3.3.1 General .......................................................................... not addressed in this Guide
3.3.2 Load and Load Designation .......................................... not addressed in this Guide
3.4 LOAD FACTORS AND COMBINATIONS.................................................................. 58
3.4.1 Load Factors and Load Combinations .................................................................. 58
3.4.2 Load Factors for Construction Loads ................................................................... 59
3.4.2.1 Evaluation at the Strength Limit State ............................................................ 59
3.4.2.2 Evaluation of Deflection at the Service Limit State ....................................... 60
3.4.3 Load Factors for Jacking and Post-Tensioning Forces ......................................... 61
3.4.3.1 Jacking Forces ................................................................................................. 61
3.4.3.2 Force for Post-Tensioning Anchorage Zones ................................................. 61
3.4.4 Load Factors for Orthotropic Decks ..................................................................... 61
3.5 PERMANENT LOADS .................................................................................................. 62
3.5.1 Dead Loads: DC, DW, and EV .............................................................................. 62
3.5.2 Earth Loads: EH, ES, and DD............................................................................... 62
3.6 LIVE LOADS.................................................................................................................. 62
3.6.1 Gravity Loads: LL and PL..................................................................................... 62
3.6.1.1 Vehicular Live Load ....................................................................................... 62
3.6.1.1.1 Number of Design Lanes ............................................................................ 62
3.6.1.1.2 Multiple Presence of Live Load ................................................................. 63
3.6.1.2 Design Vehicular Live Load ........................................................................... 64
3.6.1.2.1 General ....................................................................................................... 64
3.6.1.2.2 Design Truck .............................................................................................. 64
3.6.1.2.3 Design Tandem........................................................................................... 65
3.6.1.2.4 Design Lane Load ...................................................................................... 65
3.6.1.2.5 Tire Contact Area ....................................................................................... 66
3.6.1.2.6 Distribution of Wheel Load through Earth Fills ........................................ 66
3.6.1.3 Application of Design Vehicular Live Loads ................................................. 67
3.6.1.3.1 General ....................................................................................................... 67
3.6.1.3.2 Loading for Optional Live Load Deflection Evaluation ............................ 68
3.6.1.3.3 Design Loads for Decks, Deck Systems, and the Top Slabs of Box Culverts
.................................................................................................................... 68
3.6.1.3.4 Deck Overhang Load.................................................................................. 68
3.6.1.4 Fatigue Load ................................................................................................... 68
3.6.1.4.1 Magnitude and Configuration .................................................................... 68
3.6.1.4.2 Frequency ................................................................................................... 69
3.6.1.4.3 Load Distribution for Fatigue ..................................................................... 70
3.6.1.5 Rail Transit Load ............................................................................................ 70
3.6.1.6 Pedestrian Loads ............................................................................................. 70
3.6.1.7 Loads on Railings ........................................................................................... 71

NSBA Guide to Navigating Routine Steel Bridge Design / 53


3.6.2 Dynamic Load Allowance: IM.............................................................................. 71
3.6.2.1 General ............................................................................................................ 71
3.6.2.2 Buried Components ........................................................................................ 71
3.6.2.3 Wood Components.......................................................................................... 71
3.6.3 Centrifugal Forces: CE ......................................................................................... 72
3.6.4 Braking Force: BR................................................................................................. 72
3.6.5 Vehicular Collision Force: CT .............................................................................. 73
3.6.5.1 Protection of Structures................................................................................... 73
3.6.5.2 Vehicle Collision with Barriers ...................................................................... 73
3.7 WATER LOADS: WA..................................................................................................... 73
3.7.1 Static Pressure ....................................................................................................... 73
3.7.2 Buoyancy .............................................................................................................. 73
3.7.3 Stream Pressure ..................................................................................................... 74
3.7.3.1 Longitudinal .................................................................................................... 74
3.7.3.2 Lateral ............................................................................................................. 74
3.7.4 Wave Load ............................................................................................................ 74
3.7.5 Change in Foundations Due to Limit State for Scour ........................................... 74
3.8 WIND LOAD: WL AND WS .......................................................................................... 75
3.8.1 Horizontal Wind Load .......................................................................................... 75
3.8.1.1 Exposure Conditions ....................................................................................... 75
3.8.1.1.1 General ....................................................................................................... 75
3.8.1.1.2 Wind Speed ................................................................................................ 75
3.8.1.1.3 Wind Direction for Determining Wind Exposure Category ...................... 76
3.8.1.1.4 Ground Surface Roughness Categories ...................................................... 76
3.8.1.1.5 Wind Exposure Categories ......................................................................... 76
3.8.1.2 Wind Load on Structures: WS ........................................................................ 76
3.8.1.2.1 General ....................................................................................................... 76
3.8.1.2.2 Loads on the Superstructure ....................................................................... 77
3.8.1.2.3 Loads on the Substructure .......................................................................... 77
3.8.1.2.4 Wind Loads on Sound Barriers .................................................................. 78
3.8.1.3 Wind Load on Live Load: WL........................................................................ 78
3.8.2 Vertical Wind Load............................................................................................... 78
3.8.3 Wind-Induced Bridge Motions ............................................................................. 79
3.8.3.1 General ............................................................................................................ 79
3.8.3.2 Wind-Induced Motions ................................................................................... 79
3.8.3.3 Control of Dynamic Responses ...................................................................... 79
3.8.4 Site-Specific and Structure-Specific Studies ........................................................ 79
3.9 ICE LOADS: IC ...................................................................... not addressed in this Guide
3.9.1 General ...........................................................................not addressed in this Guide
3.9.2 Dynamic Ice Forces on Piers .........................................not addressed in this Guide
3.9.2.1 Effective Ice Strength ..............................................not addressed in this Guide
3.9.2.2 Crushing and Flexing ...............................................not addressed in this Guide
3.9.2.3 Small Streams ..........................................................not addressed in this Guide
3.9.2.4 Combination of Longitudinal and Transverse Forces ..........................................
..................................................................................not addressed in this Guide
3.9.2.4.1 Piers Parallel to Flow ..........................................not addressed in this Guide

NSBA Guide to Navigating Routine Steel Bridge Design / 54


3.9.2.4.2 Piers Skewed to Flow ..........................................not addressed in this Guide
3.9.2.5 Slender and Flexible Piers .......................................not addressed in this Guide
3.9.3 Static Ice Loads on Piers ................................................not addressed in this Guide
3.9.4 Hanging Dams and Ice Jams ..........................................not addressed in this Guide
3.9.5 Vertical Forces Due to Ice Adhesion .............................not addressed in this Guide
3.9.6 Ice Accretion and Snow Loads on Superstructures .......not addressed in this Guide
3.10 Earthquake Effects: EQ ................................................................................................... 79
3.10.1 General ...........................................................................not addressed in this Guide
3.10.2 Seismic Hazard ..................................................................................................... 79
3.10.2.1 General Procedure ........................................................................................... 79
3.10.2.2 Site-Specific Procedure ............................................not addressed in this Guide
3.10.3 Site Effects ............................................................................................................ 80
3.10.3.1 Site Class Definitions...............................................not addressed in this Guide
3.10.3.2 Site Factors...................................................................................................... 80
3.10.4 Seismic Hazard Characterization .......................................................................... 80
3.10.4.1 Design Response Spectrum......................................not addressed in this Guide
3.10.4.2 Elastic Seismic Response Coefficient ............................................................. 80
3.10.5 Operational Classification ..............................................not addressed in this Guide
3.10.6 Seismic Performance Zones ...........................................not addressed in this Guide
3.10.7 Response Modification Factors......................................not addressed in this Guide
3.10.7.1 General .....................................................................not addressed in this Guide
3.10.7.2 Application ...............................................................not addressed in this Guide
3.10.8 Combination of Seismic Force Effects ..........................not addressed in this Guide
3.10.9 Calculation of Design Forces ................................................................................ 80
3.10.9.1 General .....................................................................not addressed in this Guide
3.10.9.2 Seismic Zone 1 ................................................................................................ 80
3.10.9.3 Seismic Zone 2 .........................................................not addressed in this Guide
3.10.9.4 Seismic Zones 3 and 4 .............................................not addressed in this Guide
3.10.9.4.1 General ................................................................not addressed in this Guide
3.10.9.4.2 Modified Design Forces. .....................................not addressed in this Guide
3.10.9.4.3 Inelastic Hinging Forces ......................................not addressed in this Guide
3.10.9.4.3a General ............................................................not addressed in this Guide
3.10.9.4.3b Single Columns and Piers ...............................not addressed in this Guide
3.10.9.4.3c Piers with Two or More Columns...................not addressed in this Guide
3.10.9.4.3d Column and Pile Bent Design Forces .............not addressed in this Guide
3.10.9.4.3e Pier Design Forces ..........................................not addressed in this Guide
3.10.9.4.3f Foundation Design Forces ..............................not addressed in this Guide
3.10.9.5 Longitudinal Restrainers ..........................................not addressed in this Guide
3.10.9.6 Hold-Down Devices.................................................not addressed in this Guide
3.10.10 Requirements for Temporary Bridges and Stage Construction ................................
........................................................................................not addressed in this Guide
3.11 EARTH PRESSURE: EH, ES, LS, AND DD .......................not addressed in this Guide
3.11.1 General ...........................................................................not addressed in this Guide
3.11.2 Compaction ....................................................................not addressed in this Guide
3.11.3 Presence of Water ..........................................................not addressed in this Guide

NSBA Guide to Navigating Routine Steel Bridge Design / 55


3.11.4 Effect of Earthquake ......................................................not addressed in this Guide
3.11.5 Earth Pressure: EH .........................................................not addressed in this Guide
3.11.5.1 Lateral Earth Pressure ..............................................not addressed in this Guide
3.11.5.2 At-Rest Lateral Earth Pressure Coefficient, ko........not addressed in this Guide
3.11.5.3 Active Lateral Earth Pressure Coefficient, ka..........not addressed in this Guide
3.11.5.4 Passive Lateral Earth Pressure Coefficient, kp ........not addressed in this Guide
3.11.5.5 Equivalent-fluid Method of Estimating Rankine Lateral Earth Pressures ...........
..................................................................................not addressed in this Guide
3.11.5.6 Lateral Earth Pressures for Nongravity Cantilevered Walls ................................
..................................................................................not addressed in this Guide
3.11.5.7 Apparent Earth Pressure (AEP) for Anchored Walls ..........................................
..................................................................................not addressed in this Guide
3.11.5.7.1 Cohesionless Soils ...............................................not addressed in this Guide
3.11.5.7.2 Cohesive Soils .....................................................not addressed in this Guide
3.11.5.7.2a Stiff to Hard ....................................................not addressed in this Guide
3.11.5.7.2b Soft to Medium Stiff .......................................not addressed in this Guide
3.11.5.8 Lateral Earth Pressures for Mechanically Stabilized Earth Walls .......................
..................................................................................not addressed in this Guide
3.11.5.8.1 General ................................................................not addressed in this Guide
3.11.5.8.2 Internal Stability ..................................................not addressed in this Guide
3.11.5.9 Lateral Earth Pressures for Prefabricated Modular Walls ...................................
..................................................................................not addressed in this Guide
3.11.5.10 Lateral Earth Pressures for Sound Barriers Supported on Discrete and
Continuous Vertical Embedded Elements ...............not addressed in this Guide
3.11.6 Surcharge Loads: ES and LS .........................................not addressed in this Guide
3.11.6.1 Uniform Surcharge Loads (ES) ...............................not addressed in this Guide
3.11.6.2 Point, Line, and Strip Loads (ES): Walls Restrained from Movement ...............
..................................................................................not addressed in this Guide
3.11.6.3 Strip Loads (ES): Flexible Walls .............................not addressed in this Guide
3.11.6.4 Live Load Surcharge (LS) .......................................not addressed in this Guide
3.11.6.5 Reduction of Surcharge............................................not addressed in this Guide
3.11.7 Reduction Due to Earth Pressure ...................................not addressed in this Guide
3.11.8 Downdrag .......................................................................not addressed in this Guide
3.12 FORCE EFFECTS DUE TO SUPERIMPOSED DEFORMATIONS: TU, TG, SH, CR,
SE, PS ....................................................................................not addressed in this Guide
3.12.1 General ...........................................................................not addressed in this Guide
3.12.2 Uniform Temperature ....................................................not addressed in this Guide
3.12.2.1 Temperature Range for Procedure A .......................not addressed in this Guide
3.12.2.2 Temperature Range for Procedure B .......................not addressed in this Guide
3.12.2.3 Design Thermal Movements ....................................not addressed in this Guide
3.12.3 Temperature Gradient ....................................................not addressed in this Guide
3.12.4 Differential Shrinkage ....................................................not addressed in this Guide
3.12.5 Creep ..............................................................................not addressed in this Guide
3.12.6 Settlement ......................................................................not addressed in this Guide
3.12.7 Secondary Forces from Post-Tensioning, PS.................not addressed in this Guide

NSBA Guide to Navigating Routine Steel Bridge Design / 56


3.13 FRICTION FORCES: FR .....................................................not addressed in this Guide
3.14 VESSEL COLLISION: CV...................................................not addressed in this Guide
3.14.1 General ...........................................................................not addressed in this Guide
3.14.2 Owner’s Responsibility ..................................................not addressed in this Guide
3.14.3 Operational Classification ..............................................not addressed in this Guide
3.14.4 Design Vessel.................................................................not addressed in this Guide
3.14.5 Annual Frequency of Collapse.......................................not addressed in this Guide
3.14.5.1 Vessel Frequency Distribution.................................not addressed in this Guide
3.14.5.2 Probability of Aberrancy..........................................not addressed in this Guide
3.14.5.2.1 General ................................................................not addressed in this Guide
3.14.5.2.2 Statistical Method ................................................not addressed in this Guide
3.14.5.2.3 Approximate Method ..........................................not addressed in this Guide
3.14.5.3 Geometric Probability ..............................................not addressed in this Guide
3.14.5.4 Probability of Collapse ............................................not addressed in this Guide
3.14.5.5 Protection Factor ......................................................not addressed in this Guide
3.14.6 Design Collision Velocity ..............................................not addressed in this Guide
3.14.7 Vessel Collision Energy .................................................not addressed in this Guide
3.14.8 Ship Collision Force on Pier ..........................................not addressed in this Guide
3.14.9 Ship Bow Damage Length .............................................not addressed in this Guide
3.14.10 Ship Collision Force on Superstructure .........................not addressed in this Guide
3.14.10.1 Collision with Bow ..................................................not addressed in this Guide
3.14.10.2 Collision with Deck House ......................................not addressed in this Guide
3.14.10.3 Collision with Mast ..................................................not addressed in this Guide
3.14.11 Barge Collision Force on Pier ........................................not addressed in this Guide
3.14.12 Barge Bow Damage Length ...........................................not addressed in this Guide
3.14.13 Damage at the Extreme Limit State ...............................not addressed in this Guide
3.14.14 Application of Impact Force ..........................................not addressed in this Guide
3.14.14.1 Substructure Design .................................................not addressed in this Guide
3.14.14.2 Superstructure Design ..............................................not addressed in this Guide
3.14.15 Protection of Substructures ............................................not addressed in this Guide
3.14.16 Security Considerations .................................................not addressed in this Guide
3.15 BLAST LOADING: BL ........................................................not addressed in this Guide
3.15.1 Introduction ....................................................................not addressed in this Guide
3.16 REFERENCES ......................................................................not addressed in this Guide
APPENDIX A3 SEISMIC DESIGN FLOWCHARTS ................not addressed in this Guide
APPENDIX B3 OVERSTRENGTH RESISTANCE...................not addressed in this Guide

NSBA Guide to Navigating Routine Steel Bridge Design / 57


3.4 LOAD FACTORS AND COMBINATIONS

3.4.1 Load Factors and Load Combinations


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article defines the factored load combinations necessary for bridge design. The limit-state
load combinations and associated load factors are specified in Table 3.4.1-1. A number of limit-
state load combinations apply to the design of routine steel I-girder bridges. Strength I and II apply.
Strength III and V also apply, but only to the analysis of interim construction conditions and to
cross-frame and substructure design. Service I and II apply. Fatigue I applies when infinite fatigue
life design is required per Article 6.6.1.2.3; otherwise, Fatigue II applies for finite fatigue life. See
the Discussion of Articles 3.6.1.4.2 and 6.6.1.2.3 in this Guide for more guidance on the Fatigue I
and II limit-state load combinations.
Conversely, a number of limit-state load combinations do not apply to, and should not be
considered in, the design of the routine steel I-girder bridges covered by this Guide. Strength IV
is intended to address structures with a high dead load to live load ratio exceeding 7.0, but the
routine steel I-girder bridges covered by this Guide do not exhibit such high dead load to live load
ratios. Extreme Event I may apply in areas subject to significant seismic effects (Seismic Zones 2,
3, and 4), but design for those types of conditions is considered beyond the definition given herein
of a routine steel I-girder bridge. Extreme Event I may apply in the design of bridges subject to
less severe seismic effects (Seismic Zone 1), but only to bearing and substructure design, which
are beyond the scope of this Guide. Extreme Event II only applies on a case-by-case basis and only
to substructure design; typically, this limit-state load combination would not be applicable for the
design of routine steel I-girder bridges. Service III and IV are specifically applicable only to the
design of segmental concrete girders and concrete substructures, respectively, and thus do not
apply to the design of the routine steel I-girder bridges covered by this Guide.
Only the permanent loads DC and DW and the transient loads IM, LL, WL, and WS, as defined in
Article 3.3.2, are applicable to the design of the routine steel I-girder bridges covered by this Guide.
The permanent loads CR, DD, EH, EL, ES, EV and PS and the transient loads BL, BR, CE, CT,
CV, EQ, FR, IC, LS, SE, TG, TU, and WA do not affect the design of superstructures and should
not be considered in the design of routine steel I-girder bridge superstructures.
In Table 3.4.1-1, the load factors for the permanent loads DC and DW in the strength limit state
load combinations are not provided with singular numeric values, but instead are designated with
a variable, γp, to allow the designation of minimum and maximum values. This addresses the need
to consider potential variability in these loads, where a minimum value might create a certain
critical loading condition, while a maximum value might create a different critical loading
condition. The values of γp for the permanent loads DC and DW are specified in Table 3.4.1-2;
other load factors specified in Table 3.4.1-2 do not affect the design of superstructures. Also, only
the load factors for DC specified on the first line of the table are applicable for routine steel I-
girder bridge superstructures (i.e., the load factors specified for Strength IV do not apply). The
specified minimum load factors for DC and DW in Table 3.4.1-2 are only to be applied when the

NSBA Guide to Navigating Routine Steel Bridge Design / 58


dead- and live-load force effects are of opposite sign; e.g., when investigating for uplift at end
supports or when designing bolted field splices located near points of permanent load
contraflexure. The load factors specified in Tables 3.4.1-3 through 3.4.1-5 are not applicable for
the design of routine steel I-girder bridge superstructures.
For further discussion on loads and limit-state load combinations, consult Chapter 3 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures.

3.4.2 Load Factors for Construction Loads

3.4.2.1 Evaluation at the Strength Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies the minimum required load factors for constructibility checks. For routine
steel I-girder design, the constructibility checks apply to the timeframe prior to when the concrete
deck is cured (i.e., the final structural condition).
The standard of care in many jurisdictions is that the designer need only perform a non-structural
review of the conceptual erection sequence for the structural steel framing, primarily to
demonstrate that a viable erection scheme exists (i.e., an erection sequence that is feasible given
the known site conditions and constraints, specified maintenance-of-traffic sequence and
requirements, etc.), including consideration of the location of shoring towers, lifting and holding
cranes, etc. Owner-agencies in these jurisdictions expect detailed erection engineering to be
performed by the Contractor’s engineer, not by the bridge’s designer. However, some Owner-
agencies do require that the designer perform some level of detailed erection engineering. Review
local Owner-agency design policies and construction specifications and the local standard of care
to determine the requirements in any given specific jurisdiction. Note that the performance of
detailed erection engineering is beyond the scope of this Guide.
Once the structural steel framing system is fully erected, the designer is responsible for checking
that the structural steel has sufficient strength and stiffness to resist construction loads. The
noncomposite steel superstructure should be evaluated for constructibility using the load factors
prescribed in this Article. The constructibility checks typically involve more than just
consideration of the weight of the wet concrete deck on the noncomposite girders, which is usually
applied sequentially, but also construction equipment and worker loads, wind loads, and deck
overhang falsework and formwork loads. The load factors and load combinations used for these
constructibility checks are specified in this article, and include load combinations for the Strength
I and Strength III limit states, as well as a special load combination discussed in the last paragraph
of the article, in which a load factor of 1.4 is applied to the DC and construction loads. This special
load combination typically controls the constructibility checks, but all of the specified load
combinations must be investigated. The specified load factors for Strength III involve design
checks for wind loads acting on the fully erected steelwork (see the Discussion of Article 3.8 in
this Guide).

NSBA Guide to Navigating Routine Steel Bridge Design / 59


For the design constructibility checks, the positive moment regions of the exterior (fascia) girders
in their noncomposite condition are typically evaluated for the combined effects of the self-weight
of the girders, the sequential deck-casting sequence, and the deck overhang and wind loads (as
appropriate) for a “construction active” case, and self-weight plus wind for a “construction
inactive” case. Consult Owner-agency policy with regard to specific loads and load combinations
to be investigated during construction. The constructibility checks generally control the design of
the top flange of the girder and the cross-frame or diaphragm spacing in the positive moment
region. Most commercial steel bridge design software will perform a sequential deck-casting
analysis, but may not necessarily evaluate the additional effects of deck overhang and wind loads
acting on the exterior girders; therefore, the combined effects may need to be evaluated through
separate computations.
See the Discussion of Article 6.10.3 in this Guide for more detailed information on the required
design constructibility checks utilizing the load factors specified in this Article. The Reference
Manual for NHI Course 130102, Engineering for Structural Stability in Bridge Construction
provides further discussion of constructibility checks and the associated loads. In addition, the
AASHTO Guide Specifications for Wind Loads on Bridges During Construction, presents wind
load provisions addressing wind loads that may occur during the length of time between erection
of girders and placement of the deck, and can be consulted in the absence of Owner-agency
guidance.

3.4.2.2 Evaluation of Deflection at the Service Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article specifies load factors for evaluating deflections during construction. In terms of
vertical deflections, this Article does not typically apply to routine steel I-girder bridges unless site
constraints or contract requirements limit superstructure deflections under construction loading. In
terms of horizontal deflections, several Owner-agencies require evaluation of lateral deflections of
the steel framing under wind loads; check local Owner-agency policy to see if this requirement
applies. The wind pressure used for checking deflections during construction is usually of lower
magnitude than the wind loading applied to the final structure; this reflects the shorter duration of
construction vs. the anticipated service life of the bridge and thus the lower probability of an
extreme storm event.
These checks are generally performed using fairly simple calculations, and are generally
performed by hand or programmed into a simple spreadsheet. Typically, the routine steel I-girder
bridges covered by this Guide will pass these checks due to their relatively limited span lengths.
Excessive lateral deflection of routine steel I-girder bridges under wind loading is likely a sign that
the flanges are too narrow; consider using wider flanges as the initial step to address the situation.
The introduction of lateral bracing to control horizontal deflections prior to construction of the
composite concrete decks adds significant complexity to the design and construction and is beyond
the scope of this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 60


Section 7.4.4 of the Reference Manual for NHI Course 130102, Engineering for Structural
Stability in Bridge Construction provides a good discussion of the calculation of horizontal wind
loading on the non-composite structural steel framing. Section 6.5.3.6 of the Reference Manual
for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures provides a good discussion of the evaluation of horizontal wind loading on the
non-composite structural steel framing, including a simplified example calculation. FHWA’s Steel
Bridge Design Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-
Girder Bridge also provides a good discussion of the evaluation of horizontal wind loading on the
non-composite structural steel framing, including a simplified example calculation. In addition,
the AASHTO Guide Specifications for Wind Loads on Bridges During Construction, presents
wind load provisions addressing wind loads that may occur during the length of time between
erection of girders and placement of the deck, and can be consulted in the absence of Owner-
agency guidance.

3.4.3 Load Factors for Jacking and Post-Tensioning Forces

3.4.3.1 Jacking Forces


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article specifies the minimum load factors for dead load and live load associated with the
design of elements affected by future jacking of the bridge (for example, to replace bearings) while
the bridge remains in service. This Article is only applicable when the Owner-agency requires that
the bridge be designed to accommodate future jacking. In these situations, the designer first
determines appropriate locations for the placement of the jacks, and then must design or check the
structural elements that will be affected by the loads applied by the jacks when lifting the bridge.
Depending on the method and location of jacking, the girders and/or cross-frames may be affected,
as well as the substructures. Design details to address jacking loads, including items such as
jacking diaphragms, jacking stiffeners on the girder or diaphragms at supports, etc., may need to
be added.

3.4.3.2 Force for Post-Tensioning Anchorage Zones


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the load factor for post-tensioning anchor zones. The routine steel I-girder
bridges covered by this Guide do not contain post-tensioning; therefore, this Article is not
applicable.

3.4.4 Load Factors for Orthotropic Decks


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 61


Discussion:
This Article specifies additional fatigue load factors for the design orthotropic decks. The routine
steel I-girder bridges covered by this Guide do not use orthotropic decks; therefore, this Article is
not applicable.

3.5 PERMANENT LOADS

3.5.1 Dead Loads: DC, DW, and EV


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article defines the permanent dead loads applicable to the design of bridges and the unit
weights that may be applied to those loads. DC and DW are applicable to superstructure design.
Vertical earth loads, EV, do not apply to the superstructure design of the routine steel I-girder
bridges covered by this Guide.

3.5.2 Earth Loads: EH, ES, and DD


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article defines earth loads, which are applicable to the design of substructures, culverts,
retaining walls, tunnels, and similar structures, but not to routine steel I-girder bridge
superstructures.

3.6 LIVE LOADS

3.6.1 Gravity Loads: LL and PL

3.6.1.1 Vehicular Live Load

3.6.1.1.1 Number of Design Lanes


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies the method for determining the number of design lanes for applying live load
to the structure, which is applicable to all routine steel I-girder bridges. Owners occasionally may
specify more conservative guidelines for the determination of the number of design lanes which
may supersede the provisions of this Article, however the effects of such guidance are generally
more significant for substructure design than superstructure design. For routine steel I-girder
design, which is based on line girder analysis (such as NSBA's LRFD Simon line-girder analysis
and design program) using the approximate live load distribution factors presented in Article
4.6.2.2, the effect of variations in the number of design lanes generally has no impact on the

NSBA Guide to Navigating Routine Steel Bridge Design / 62


superstructure design. Users should verify the capabilities, assumptions, and general correctness
of any program’s calculations prior to initial use.

3.6.1.1.2 Multiple Presence of Live Load


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies the multiple presence factors applied to live load effects, which account for
the probability of concurrent vehicles in adjacent lanes. These provisions are applicable to the
design of bridges carrying vehicular traffic, including the routine steel I-girder bridges covered by
this Guide.
Care should be taken in the application of this factor when using the approximate live load
distribution factors of Articles 4.6.2.2, as is the case in the line girder analysis methods used to
design routine steel I-girder bridges (such as NSBA's LRFD Simon line-girder analysis and design
program). Users should verify the assumptions and general correctness of any program’s
calculations prior to initial use. As noted in the first paragraph of the Commentary, the multiple
presence factors are already included in the approximate equations for the live load distribution
factors given in the tables in Articles 4.6.2.2 and 4.6.2.3 and should not be applied when using
these equations. Also, as noted in the first paragraph of this Article, the multiple presence factors
are not to be applied at the fatigue limit state. Therefore, when investigating fatigue using the
fatigue design load placed in a single lane using the single-lane distribution factor equations given
in Articles 4.6.2.2 and 4.6.2.3, as applicable, the computed distribution factor from the equation
must be divided by the specified multiple presence factor for one-lane loaded of 1.2 (Table
3.6.1.1.2-1). When utilizing the lever rule or the special rigid cross-section requirement for
evaluating the single-lane live-load distribution at the fatigue limit state to the exterior girder in
steel I-girder bridges (see the Discussion of Articles 4.6.2.2.2d and 4.6.2.2.2e in the Guide), the
multiple presence factor of 1.2 should not be applied. However, when utilizing the lever rule or
the special rigid cross-section requirement for evaluating the live-load distribution to the exterior
girder in steel I-girder bridges at the strength and service limit states, the appropriate multiple
presence factor specified in Table 3.6.1.1.2-1 must be applied. For further description of the lever
rule, see the Commentary for Article 4.6.2.2.1.
Designers should carefully read and fully understand the provisions of this Article and its
associated Commentary when determining when and how to apply the multiple presence factors.
This is especially the case for routine steel I-girder bridges with sidewalk loading
For examples illustrating the proper application of the multiple presence factors in the computation
of the live load distribution factors, consult Section 4.4.2 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures (2015)
and FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous Straight
Composite Steel I-Girder Bridge.

NSBA Guide to Navigating Routine Steel Bridge Design / 63


3.6.1.2 Design Vehicular Live Load

3.6.1.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article defines the components (truck, tandem and lane) comprising the HL-93 design
vehicle, which is applicable to the design of all bridges, including routine steel I-girder bridges.
Note that most commercial line girder analysis steel bridge design programs, such as NSBA's
LRFD Simon line-girder analysis and design program, are pre-programmed with the AASHTO
LRFD BDS standard live loads as either the default live load or a selectable live load option.
Designers should verify their understanding of the program’s live load model and how it is used
in the analysis prior to initial use of the program but should rarely have to separately program the
standard AASHTO LRFD BDS live loads.
Note that some Owners may prescribe that designs use a modified version of these live load
components or may specify that designs also consider additional Owner-specific live loads; review
Owner-agency guidelines in conjunction with the provisions of this Article.
For further information on live loads, consult section 3.4 and 4.4 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures
(2015).

3.6.1.2.2 Design Truck


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article defines loading and axle spacing of the HL-93 design truck, which is applicable to all
routine steel I-girder bridges, and is applied in conjunction with the design lane load (see the
Discussion of Articles 3.6.1.2.4 and 3.6.1.3 in this Guide). The dynamic load allowance, IM, of
33 percent specified in Article 3.6.2.1 (see the Discussion of Article 3.6.2.1 in this Guide) is
applied to the design truck only.
Note that most commercial line girder analysis steel bridge design programs, such as NSBA's
LRFD Simon line-girder analysis and design program, are preprogrammed with the AASHTO
LRFD BDS standard live loads as either the default live load or a selectable live load option.
Designers should verify their understanding of the program’s live load model and how it is used
in the analysis prior to initial use of the program but should rarely have to separately program the
standard AASHTO LRFD BDS live loads.
Note that some Owners may prescribe that designs use a modified version of these live load
components or may specify that designs also consider additional Owner-specific live loads; review
Owner-agency guidelines in conjunction with the provisions of this Article.

NSBA Guide to Navigating Routine Steel Bridge Design / 64


For further information on live loads, consult section 3.4 and 4.4 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures
(2015).

3.6.1.2.3 Design Tandem


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article defines loading and axle spacing of the HL-93 design tandem, which is applicable to
all routine steel I-girder bridges, and is applied in conjunction with the design lane load (see the
Discussion of Articles 3.6.1.2.4 and 3.6.1.3 in this Guide). The dynamic load allowance, IM, of 33
percent specified in Article 3.6.2.1 (see the Discussion of Article 3.6.2.1 in this Guide) is applied
to the design tandem only. This combination of the design tandem and design lane load will
typically only control the live-load force effects for short-span bridges.
Note that most commercial line girder analysis steel bridge design programs, such as NSBA's
LRFD Simon line-girder analysis and design program, are preprogrammed with the AASHTO
LRFD BDS standard live loads as either the default live load or a selectable live load option.
Designers should verify their understanding of the program’s live load model and how it is used
in the analysis prior to initial use of the program but should rarely have to separately program the
standard AASHTO LRFD BDS live loads.
Note that some Owners may prescribe that designs use a modified version of these live load
components or may specify that designs also consider additional Owner-specific live loads; review
Owner-agency guidelines in conjunction with the provisions of this Article.
For further information on live loads, consult section 3.4 and 4.4 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures
(2015).

3.6.1.2.4 Design Lane Load


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article defines loading and lane with of the HL-93 lane load, which is applicable to all routine
steel I-girder bridges, and is applied in conjunction with the design truck or design tandem (see
the Discussion of Articles 3.6.1.2.2, 3.6.1.2.3, and 3.6.1.3 in this Guide). The dynamic load
allowance, IM, of 33 percent specified in Article 3.6.2.1 (see the Discussion of Article 3.6.2.1 in
this Guide) is not applied to the design lane load.
Note that most commercial line girder analysis steel bridge design programs, such as NSBA's
LRFD Simon line-girder analysis and design program, are preprogrammed with the AASHTO
LRFD BDS standard live loads as either the default live load or a selectable live load option.
Designers should verify their understanding of the program’s live load model and how it is used
in the analysis prior to initial use of the program but should rarely have to separately program the
standard AASHTO LRFD BDS live loads.

NSBA Guide to Navigating Routine Steel Bridge Design / 65


Note that some Owners may prescribe that designs use a modified version of these live load
components or may specify that designs also consider additional Owner-specific live loads; review
Owner-agency guidelines in conjunction with the provisions of this Article.
For further information on live loads, consult section 3.4 and 4.4 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures
(2015).

3.6.1.2.5 Tire Contact Area


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article defines the tire contact area for the design truck and tandem. The tire contact area is
generally considered only in the local design bridge decks. For the superstructure (girder) design
of routine bridges, idealizing the wheel or axle loads as point loads is the standard accepted
method.

3.6.1.2.6 Distribution of Wheel Load through Earth Fills

3.6.1.2.6a General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the method of distributing live load forces through earth fill for buried
structures, which does not apply to the design of the routine steel I-girder bridges covered by this
Guide.

3.6.1.2.6b Traffic Parallel to the Culvert Span


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies live load distribution factors for culverts subjected to traffic parallel to the
structure, which do not apply to the design of the routine steel I-girder bridges covered by this
Guide.

3.6.1.2.6c Traffic Perpendicular to the Culvert Span


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies live load distribution factors for culverts subjected to traffic perpendicular
to the structure, which do not apply to the design of the routine steel I-girder bridges covered by
this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 66


3.6.1.3 Application of Design Vehicular Live Loads

3.6.1.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article describes the method for applying the design vehicular live loads to the structure.
Most, but not all, of the provisions are applicable to the design of the routine steel I-girder bridges
covered by this Guide as described further below. Note that axle loads or lengths of the design
lane that do not contribute to the extreme force effect under consideration are to be neglected.
The particular item in this Article related to the design of deck overhangs is not applicable to the
design of routine steel I-girder bridge superstructures in the context of this Guide, since the design
of the deck (and thus the deck overhang) is beyond the scope of this Guide. When calculating the
effect of live loads on the steel superstructure, the other provision which states that the center of
any wheel load not be closer than 2.0 feet from the edge of the design lane is applicable,
particularly when utilizing the lever rule or the special rigid cross-section requirement for
evaluating the live-load distribution to the exterior girder in steel I-girder bridges (see the
Discussion of Articles 4.6.2.2.2d and 4.6.2.2.2e in this Guide).
For simple span bridges, the particular item concerning the effect of loading by 90% of two design
trucks in a single lane with a minimum headway of 50 feet (and a rear-axle spacing fixed at 14
feet) combined with 90% of the design lane load for negative moments between points of
permanent load contraflexure is not applicable to the design of simple span bridge superstructures,
which are subject only to positive moment loading. This item may be applicable to the calculation
of reactions at interior piers in bridges with multiple simple spans, but substructure design is
beyond the scope of this Guide.
For multi-span continuous bridges, the particular item concerning the effect of loading by 90% of
two design trucks in a single lane with a minimum headway of 50 feet (and a rear-axle spacing
fixed at 14 feet) combined with 90% of the design lane load is applicable when calculating negative
moments between points of permanent load contraflexure and reactions at interior supports in
routine multi-span continuous rolled beam and plate girder bridges. The negative moments
between points of permanent load contraflexure and interior-support reactions due to this loading
are compared to the negative moments in these regions and interior-support reactions due to the
HL-93 loading and the governing moment and reaction applies. Due to the minimum 50-foot
headway between trucks and the 0.90 reduction factor, this loading will generally not control for
continuous structures with short spans (i.e., less than 50 feet). The dynamic load allowance, IM,
of 33 percent specified in Article 3.6.2.1 (see the Discussion of Article 3.6.2.1 in this Guide) is
applied to the two design trucks only.
For further information on live loads, consult section 3.4 and 4.4 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures
(2015).

NSBA Guide to Navigating Routine Steel Bridge Design / 67


3.6.1.3.2 Loading for Optional Live Load Deflection Evaluation
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article defines the special live loading for the optional live load deflection evaluation
specified in Article 2.5.2.6.2 (see the Discussion of Article 2.5.2.6.2 in this Guide). This special
loading should be used to evaluate the live load deflection in the routine steel I-girder bridges
covered by this Guide, unless specified otherwise by the Owner-agency. The dynamic load
allowance, IM, of 33 percent specified in Article 3.6.2.1 (see the Discussion of Article 3.6.2.1 in
this Guide) is applied to the design truck portion of the loading only.
For further information on live loads, consult section 3.4 and 4.4 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures
(2015).

3.6.1.3.3 Design Loads for Decks, Deck Systems, and the Top Slabs of Box Culverts
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article describes the live loading of decks and top slabs of culverts. Only the provisions
related to deck design are applicable to the routine steel I-girder bridges covered by this Guide;
however, deck design is outside the scope of this Guide.

3.6.1.3.4 Deck Overhang Load


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article describes the live loading for the design of deck overhangs, which is outside the scope
of this Guide.

3.6.1.4 Fatigue Load

3.6.1.4.1 Magnitude and Configuration


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides the axle loads and spacing for the fatigue live load vehicle, which is
equivalent to the design truck specified in Article 3.6.1.2.2 but with a constant rear-axle spacing
of 30 feet. The fatigue live load vehicle is placed in a single lane and the design lane load is not
applied. The dynamic load allowance, IM, of 15 percent specified in Article 3.6.2.1 (see the
Discussion of Article 3.6.2.1 in this Guide) is applied to the fatigue live load vehicle. Fatigue
loading is applicable to all steel girder bridges carrying vehicular live load.

NSBA Guide to Navigating Routine Steel Bridge Design / 68


The third paragraph and associated figure concerning orthotropic decks is not applicable to the
routine steel I-girder bridges covered by this Guide, which do not use orthotropic steel decks.

3.6.1.4.2 Frequency
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article defines the average number of trucks per day in a single lane averaged over the design
life of the structure, ADTTSL. This is used to determine if a particular fatigue detail should be
designed for finite or infinite fatigue life in order to calculate the appropriate nominal fatigue
resistance for the detail in question (see the Discussion of Article 6.6.1.2.3 in this Guide regarding
the use of the ADTTSL in determining whether a fatigue detail is to be designed for finite or infinite
life).
The ADTTSL is calculated from Eq. 3.6.1.4.2-1 using the average daily truck traffic in one direction
averaged over the design life, or ADTT. The ADTT should be determined in consultation with the
traffic engineers to obtain a best estimate of traffic over the life of the structure. In most cases,
traffic count data is only available for the “current year” and a “design year,” which is often 20
years into the future. The traffic count data for the future “design year” is generally what should
be used as the basis for fatigue analysis calculations; the current year should not be used, and there
is no need to try to extrapolate for the 75-year or 100-year service life of the structure. Should a
bidirectional ADTT be provided, the Commentary for this Article recommends designing for 55
percent of the bidirectional ADTT to determine the ADTT in one direction. The fraction of truck
traffic in a single lane, p, determined from Table 3.6.1.4.2-1 should be based on the number of
design lanes (see the Discussion of Article 3.6.1.1.1 in this Guide).
If a reasonable estimate of the ADTTSL cannot be made due to lack of traffic data or some other
reason, consideration may be given to conservatively designing fatigue details for infinite fatigue
life, which does not require the ADTTSL. However, in the routine multi-span continuous I-girder
bridges covered by this Guide, the Category C' fatigue check of the connection plate-to-bottom
flange weld in regions near the points of permanent load contraflexure may control the size of the
bottom flange. Therefore, a reasonable estimate of the ADTTSL should be made, if possible, to
perform a more accurate assessment of the nominal fatigue resistance for either finite or infinite
life, as applicable.
The welded connections typically used to attach angle- or tee-section (WT) cross-frame members
to gusset plates in the truss-type cross-frames used in many steel I-girder bridges are identified as
Category E' details, and as such have very low fatigue resistance. As a result, typically the only
way to achieve a reasonable cross-frame design is to evaluate the cross-frames for finite life if the
ADTTSL is such that finite-life design is permitted for a Category E' detail. However, as explained
in the Discussion of Article 6.6.1.2.1 in this Guide, designers need not be concerned about
performing a fatigue analysis of cross-frame or diaphragm members in routine steel I-girder
bridges; due to the nature of the geometry of the framing plan and overall layout of routine steel
I-girder bridges, the live load force effects (and the resulting live load stress ranges) in the cross-
frames or diaphragms are not significant. Also, cross-frame force effects cannot be directly

NSBA Guide to Navigating Routine Steel Bridge Design / 69


calculated by a line-girder analysis. Thus, it is reasonable, permitted, and recommended that
designers do not try to evaluate the cross-frames or diaphragms in routine steel I-girder bridges for
fatigue.

3.6.1.4.3 Load Distribution for Fatigue

3.6.1.4.3a Refined Methods


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the placement of live load for refined methods of analysis, such as 2D grid
or grillage analysis, 2D plate-and-eccentric-beam analysis, or 3D analysis. For routine steel I-
girder bridges, only line girder analysis (such as the analysis performed by NSBA's LRFD Simon
line-girder analysis and design program) is needed and therefore this Article is not applicable.
Users should verify the capabilities, assumptions, and general correctness of any program’s
calculations prior to initial use.

3.6.1.4.3b Approximate Methods


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies the distribution factor to be used for a single lane of fatigue live load when
using approximate methods of analysis such as line girder analysis; therefore this Article is
applicable to the design of the routine steel I-girder bridges covered by this Guide.

3.6.1.5 Rail Transit Load


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article discusses rail transit live loads, which are considered not applicable in the design of
routine steel I-girder bridges for the purposes of this Guide.

3.6.1.6 Pedestrian Loads


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article defines pedestrian loads, which are only applicable to highway bridges providing
pedestrian access (i.e., sidewalks). The Article also specifies the method for combining pedestrian
and vehicular loads to determine the total live load on a girder. Where vehicular traffic can occupy
the sidewalk in the as-designed condition or a potential future configuration, the design live load
should be considered to act over the width of the sidewalk. Depending on sidewalk width, simple
hand calculations may demonstrate that this condition (vehicular traffic over full width of the

NSBA Guide to Navigating Routine Steel Bridge Design / 70


structure) controls over the pedestrian plus vehicular traffic condition, thereby eliminating the need
to analyze the additional configurations.

3.6.1.7 Loads on Railings


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article refers to Section 13 for the applicable live load on railings. Railing, or barrier, design
is beyond the scope of this Guide.

3.6.2 Dynamic Load Allowance: IM

3.6.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides the dynamic load allowance, IM, factors to apply to the static force effects
of the design truck or design tandem to account for dynamic effects as the vehicle crosses the
structure. The dynamic load allowance is not to be applied to the force effects due to the design
lane load or due to pedestrian loads. The provisions directly related to the design of bridges are
applicable to the design of the routine steel I-girder bridges covered by this Guide.
The provisions mentioned in this Article for the application of the dynamic load allowance for
buried components and for deck joints are not applicable to the routine steel I-girder bridge
superstructures covered by this Guide.
In addition, the provision which allows for the reduction of the dynamic load allowance “if justified
by sufficient evidence” generally should not be used in the design of routine steel I-girder bridges
without truly sufficient evidence (such as site-specific experimental testing) and the input and
approval of the Owner-agency.

3.6.2.2 Buried Components


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides the dynamic load allowance for buried components, which is not applicable
to the superstructure design of the routine steel I-girder bridges covered by this Guide.

3.6.2.3 Wood Components


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 71


Discussion:
This Article discusses the dynamic load allowance for wood components, which is not applicable
to the design of the routine steel I-girder bridges covered by this Guide.

3.6.3 Centrifugal Forces: CE


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the magnitude of lateral loading, or centrifugal force, acting on vehicles as
they traverse a bridge on a horizontally curved alignment. Routine steel I-girder bridges, as defined
for the purposes of this Guide, do not have curved decks and thus presumably would not have
curved lanes; in such cases there would not be any centrifugal force effects to consider.
However, even in the rare instance of a bridge with a straight deck but curved lanes, consideration
of centrifugal force effects in the design of the superstructure is unlikely to be warranted. The
design of routine steel I-girder bridges is, by definition for the purposes of this Guide, based on
line girder analysis (such as NSBA's LRFD Simon line-girder analysis and design program). The
effects of centrifugal force on the distribution of loads to the left and right wheels of a vehicle
cannot be considered in a line girder analysis, in which the distribution of live loads is based on
approximate live load distribution factors, not on refined distribution of individual wheel loads.
These approximate live load distribution factors are conservative, and it is unlikely that there
would be a case of a bridge with a straight deck (part of the definition of a “routine steel I-girder
bridge” for the purposes of this Guide) with lanes curved significantly enough for centrifugal force
effects to be of concern.
Centrifugal force effects should always be considered in the design of substructures for bridges
with curved decks and/or curved lanes, but substructure design is beyond the scope of this Guide.

3.6.4 Braking Force: BR


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article defines a longitudinal load acting 6 feet above the deck level to account for the force
effects of vehicular braking. The principal effects of this loading are on the bearings, substructures,
and foundations. The applied braking loads do not have a measurable influence on the design
forces of girders (moment and shear) or cross-frames. It can be shown that any redistribution in
vehicular weight to the axle loads due to braking will have a negligible effect on girder design.
Therefore, this Article is not applicable to the superstructure design of routine steel I-girder bridges
and consideration of the effects of braking forces are beyond the scope of this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 72


3.6.5 Vehicular Collision Force: CT

3.6.5.1 Protection of Structures


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article defines which substructure items are subject to vehicular collision forces, the
associated loading conditions, and means of mitigating or avoiding the subjecting of substructure
units to crash loads. Vehicular collision loads are typically not a consideration in new structures
where site constraints allow either a geometric layout that provides a proper clear zone or an
appropriate barrier to protect the substructure units. Regardless, vehicular collision is primarily a
substructure design consideration in most bridges, and the definition of a routine steel I-girder
bridge for the purposes of this Guide specifically excludes bridges where the superstructure may
be subject to vehicular collision forces.

3.6.5.2 Vehicle Collision with Barriers


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article covers the design of rigid barriers used to protect substructure elements, which is
beyond the scope of this Guide.

3.7 WATER LOADS: WA

3.7.1 Static Pressure


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article discusses the static pressure of water exerted on water-retaining structures, which is
not applicable to the superstructure design of the routine steel I-girder bridges covered by this
Guide.

3.7.2 Buoyancy
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies buoyancy uplift forces for structures inundated by water. This Article only
applies for stream or river crossings when the design flood level is above the bottom flange for the
entire length of span, and only when ventilation is not available. The definition of a routine steel
I-girder bridge for the purposes of this Guide specifically excludes bridges where there is

NSBA Guide to Navigating Routine Steel Bridge Design / 73


insufficient freeboard between the superstructure low chord and the high-water elevation
associated with design flood events.

3.7.3 Stream Pressure

3.7.3.1 Longitudinal
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article defines the longitudinal stream pressure force acting on substructure units subjected
to flowing water. This Article is only applicable to substructure design and is beyond the scope of
this Guide. Recall that the definition of a routine steel I-girder bridge for the purposes of this Guide
specifically excludes bridges where there is insufficient freeboard between the superstructure low
chord and the high-water elevation associated with design flood events.

3.7.3.2 Lateral
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article defines the lateral stream pressure force acting on substructure units subjected to
flowing water. This Article is only applicable to substructure design and is beyond the scope of
this Guide. Recall that the definition of a routine steel I-girder bridge for the purposes of this Guide
specifically excludes bridges where there is insufficient freeboard between the superstructure low
chord and the high-water elevation associated with design flood events.

3.7.4 Wave Load


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies that wave action, where applicable, shall be considered. This Article only
applies to bridges where wave action is a consideration and would be very rare. Recall that the
definition of a routine steel I-girder bridge for the purposes of this Guide specifically excludes
bridges where there is insufficient freeboard between the superstructure low chord and the high-
water elevation associated with design flood events.

3.7.5 Change in Foundations Due to Limit State for Scour


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.

NSBA Guide to Navigating Routine Steel Bridge Design / 74


Discussion:
This Article discusses scour and only applies to the substructures of stream or river crossing
bridges. New routine steel I-girder bridges are generally designed to avoid scour activity affecting
the superstructure; therefore, this Article does not apply.

3.8 WIND LOAD: WL AND WS

3.8.1 Horizontal Wind Load

3.8.1.1 Exposure Conditions

3.8.1.1.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article defines the superstructure area exposed to wind and the associated attack angles. For
the routine steel I-girder bridges covered by this Guide, wind loading is to be considered when
calculating force effects and deflections in the girders prior to deck placement, and possibly during
deck placement (see the Discussion of Article 3.8.1.2.1 in this Guide), before the top flange is
continuously braced by the concrete deck. After the deck is placed, wind loading is to be
considered when determining flange lateral bending moments and stresses in the exterior girder
bottom flange, as well as forces in the cross-frame members, due to loading on the exterior girder
web.
Note that the Commentary for Article 4.6.2.7.1 provides approximate methods for determining
these forces due to wind loading.

3.8.1.1.2 Wind Speed


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies the design wind speeds to use at the applicable strength and service limit
states.
For routine steel I-girder bridges, the values shown in Figure 3.8.1.1.2-1 are generally sufficient
as the basis for calculating design wind pressures; a site-specific wind study is rarely, if ever,
warranted.
For the determination of wind speeds during construction and their application, see the AASHTO
Guide Specifications for Wind Loads on Bridges During Construction, and also consider local
Owner-agency policies which may address this topic. In addition, the Reference Manual for NHI
Course 130102, Engineering for Structural Stability in Bridge Construction provides guidance on
selection of design wind speeds during construction.

NSBA Guide to Navigating Routine Steel Bridge Design / 75


3.8.1.1.3 Wind Direction for Determining Wind Exposure Category
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies the wind directions for determining the exposure category. The Article states
that for “typical bridges, the wind exposure category as specified in Article 3.8.1.1.5 shall be
perpendicular to the bridge.” Routine steel I-girder bridges generally fall under this designation;
consideration of multiple directions is generally unnecessary for these types of structures.

3.8.1.1.4 Ground Surface Roughness Categories


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article defines the ground surface roughness categories used for determining applicable static
wind load pressures. The categories match those used in ASCE 7-10, Minimum Design Loads for
Buildings and Other Structures. Photographs of typical Category B and C conditions are shown in
ASCE 7-10 at the end of Chapter C26, and are also provided in Section 7 of the Reference Manual
for NHI Course 130102, Engineering for Structural Stability in Bridge Construction. For a
particular structure, if the category isn’t clear (e.g., Category B or C), choosing the higher category
(i.e., C), will result in higher loads. For the routine steel I-girder bridges covered by this Guide,
the wind load often does not dictate the girder dimensions.

3.8.1.1.5 Wind Exposure Categories


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article defines the wind exposure category based on ground surface roughness as determined
in Article 3.8.1.1.4 and its prevalence upwind from the structure. Section C26.7 of ASCE 7-10,
Minimum Design Loads for Buildings and Other Structures, provides figures that help clarify the
definition of conditions upwind from the structure.

3.8.1.2 Wind Load on Structures: WS

3.8.1.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies the equation for determining the static design wind pressure acting on the
structure and the associated variables. The Design 3-second Gust Wind Speed, V, is taken from
Article 3.8.1.1.2 for the various final condition limit states. The equations to determine the pressure
exposure and elevation coefficient, KZ, for the Strength III and Service IV limit states are presented
in this Article; however, the commentary for this Article provides predetermined values for the
various exposure categories at regular intervals of structure height. For routine steel I-girder bridge

NSBA Guide to Navigating Routine Steel Bridge Design / 76


superstructure design, any refinement beyond using the maximum superstructure height about the
ground or water is generally not warranted. Additionally, for routine steel I-girder bridges without
sound barriers, the Gust Effect Factor, G, and Drag Coefficient, CD, are typically 1.00 and 1.3,
respectively.
Designers should also consider including the effects of wind loading on the structure during
construction for the constructibility checks of the girders (see Discussion of Article 6.10.3.1 and
related Articles), including consulting with Owner-agency policy regarding loads and load
combinations to be evaluated during construction. For the determination of applicable wind
pressure and associated variables during construction, see the AASHTO Guide Specifications for
Wind Loads on Bridges During Construction, and also consider any local Owner-agency policies
on this subject. In addition, the Reference Manual for NHI Course 130102, Engineering for
Structural Stability in Bridge Construction provides guidance on the determination of wind
pressures during construction. Generally, two levels of wind should be considered, an inactive and
active wind. The higher inactive wind load is applied to the fully erected steel structure to check
loads and stability prior to deck placement. The lower active wind load is the force to apply during
the placement of the noncomposite, or wet, concrete weight, which corresponds to the maximum
expected wind speed that would occur during deck placement operations. Often, the magnitude of
this active wind load pressure may be found to be very small and may be neglected in the analysis
of deck placement loads.

3.8.1.2.2 Loads on the Superstructure


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article discusses the wind load and directions of application for the design of superstructures.
For the routine steel I-girder bridges covered by this Guide, specifically for the design of the
superstructure, this discussion should be viewed in two contexts: wind loads during construction;
and wind loads on the completed structure. During construction, the steel superstructure elements
typically investigated for the effects of wind load include the noncomposite girder top and bottom
flanges (evaluated as part of the girder constructibility checks), and the cross-frames or
diaphragms. In the completed structure, steel superstructure elements typically investigated for the
effects of wind load include the girder bottom flanges and the cross-frames or diaphragms. The
wind effects in these elements are controlled by wind acting perpendicular to the bridge; other
wind skew angles do not need to be investigated.

3.8.1.2.3 Loads on the Substructure

3.8.1.2.3a Loads from the Superstructure


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.

NSBA Guide to Navigating Routine Steel Bridge Design / 77


Discussion:
This Article discusses the application of wind effects from the superstructure as they are applied
to the substructure including the various angles of attack to be considered. In general, this Article
is only applicable to substructure design and is beyond the scope of this Guide.

3.8.1.2.3b Loads Applied Directly to the Substructure


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article provides requirements for applying wind loads directly to the substructure elements,
which is beyond the scope of this Guide.

3.8.1.2.4 Wind Loads on Sound Barriers


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article only applies to bridges supporting sound barriers. For the purposes of this Guide, the
definition of a routine steel I-girder bridge excludes bridges which have sound barriers.

3.8.1.3 Wind Load on Live Load: WL


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article defines the transverse wind acting on live load and acting 6 feet above the deck level.
The applied wind on live load does not have a measurable influence on the design forces of girders
(moment and shear) or intermediate cross-frames. Wind on live load is primarily a design
consideration for bearing and substructure design. However, the transmission of the load from the
superstructure (resisted by diaphragm action of the concrete deck) to the bearings though the cross-
frames or diaphragms at the supports must be considered in the design of those elements. Similar
to wind load acting on the superstructure, wind on live load acting perpendicular to the bridge is
generally the controlling direction for the design of cross-frames or diaphragms at the supports.
Wind on live load acting at other angles of attack as discussed in this Article is applicable to the
design of the bearings and substructure, which is beyond the scope of this Guide.

3.8.2 Vertical Wind Load


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article specifies an upward, or overturning, wind load acting on the superstructure in
conjunction with the transverse superstructure wind for the Strength III and Service IV limit states.
For the routine steel I-girder bridges covered by this Guide, the effect on the superstructure design

NSBA Guide to Navigating Routine Steel Bridge Design / 78


(girders and cross-frames) is negligible and can be ignored. The uplift wind must be considered,
however, in the design of the bearings and substructure, which is beyond the scope of this Guide.

3.8.3 Wind-Induced Bridge Motions

3.8.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article and other sub-Articles of Article 3.8.3 specify requirements for bridges subjected to
wind-induced vibrations. The routine steel I-girder bridges covered by this Guide do not have a
span-to-depth or length-to-width ratio exceeding 30, cable supports, or, in general, fundamental
vertical or translational periods greater than 1 second; therefore, this Article is not applicable.

3.8.3.2 Wind-Induced Motions


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The routine steel I-girder bridges covered by this Guide are not subject to wind-induced vibrations;
therefore this Article is not applicable.

3.8.3.3 Control of Dynamic Responses


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The routine steel I-girder bridges covered by this Guide are not subject to wind-induced vibrations;
therefore this Article is not applicable.

3.8.4 Site-Specific and Structure-Specific Studies


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The routine steel I-girder bridges covered by this Guide are not subject to wind-induced vibrations;
therefore this Article is not applicable.

3.10 EARTHQUAKE EFFECTS: EQ

3.10.2 Seismic Hazard

3.10.2.1 General Procedure


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 79


Discussion:
This Article presents horizontal peak ground acceleration coefficients which are used to determine
the peak ground acceleration coefficient, PGA. The PGA is used to calculate the design forces
discussed in Article 3.10.9.2 (see the Discussion of Article 3.10.9.2 in this Guide). Other values
such as the short- and long-period spectral acceleration coefficients are not pertinent or applicable
to the design of the routine steel I-girder bridge superstructures covered by this Guide.

3.10.3 Site Effects

3.10.3.2 Site Factors


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article presents various site factors. The only site factor which is applicable to the design of
routine steel I-girder bridge superstructures is the Site Factor Fpga. The Site Factor Fpga is found in
Table 3.10.3.2-1. For Bridges in Seismic Zone 1, always use Site Class B, and identify the value
of the Site Factor Fpga from the table as a function of the peak ground acceleration coefficient,
PGA, determined in Article 3.10.2.1 (see the Discussion of Article 3.10.2.1 in this Guide).
The Site Factor Fpga is used to calculate the design forces discussed in Article 3.10.9.2. Other
values Site Factors Fa and Fv are not pertinent or applicable to the design of the routine steel I-
girder bridges covered by this Guide.

3.10.4 Seismic Hazard Characterization

3.10.4.2 Elastic Seismic Response Coefficient


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article discusses elastic seismic response coefficients. Only bridges in Seismic Zone 1 meet
the definition of a routine steel I-girder bridge for the purposes of this Guide, and thus only Article
3.10.9.2 is applicable, and Article 3.10.9.2 only refers to this Article (3.10.4.2) for the purposes of
determining the applicable acceleration coefficient, As, as specified in Eq. 3.10.4.2-2. The two
variables in Eq. 3.10.4.2-2 are the peak ground acceleration coefficient on rock (Site Class B),
PGA, and the Site Factor Fpga which are specified in Articles 3.10.2.1 and 3.10.3.2, respectively.

3.10.9 Calculation of Design Forces

3.10.9.2 Seismic Zone 1


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies horizontal design connection forces between the superstructure and
substructure for bridges in Seismic Zone 1. Bridges in Seismic Zone 1 meet the definition of a

NSBA Guide to Navigating Routine Steel Bridge Design / 80


routine steel I-girder bridge for the purposes of this Guide, so the provisions of this Article are
applicable to their design. The provisions for the calculations of the forces are a function of the
acceleration coefficient, which is specified in Article 3.10.4.2 (see the Discussion of Article
3.10.4.2 in this Guide).
Note that the design forces discussed this Article are only for the “connection” of the superstructure
to the substructure, and the definition of “connections” (as discussed in the Commentary for Article
3.10.7.1) includes only fixed bearings, expansion bearings with either restrainers, STUs, or
dampers, and shear keys. There is no need to apply these forces to the design of pier or end cross-
frames or other elements of the superstructure. See the related Discussion of Article 6.16.3 in this
Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 81


SECTION 4: STRUCTURAL ANALYSIS AND EVALUATION

TABLE OF CONTENTS
4.1 SCOPE..................................................................................... not addressed in this Guide
4.2 DEFINITIONS ........................................................................ not addressed in this Guide
4.3 NOTATION ............................................................................ not addressed in this Guide
4.4 ACCEPTABLE METHODS OF STRUCURAL ANALYSIS ....................................... 86
4.5 MATHEMATICAL MODELING .................................................................................. 86
4.5.1 General .................................................................................................................. 86
4.5.2 Structural Material Behavior................................................................................. 87
4.5.2.1 Elastic Versus Inelastic Behavior ................................................................... 87
4.5.2.2 Elastic Behavior .............................................................................................. 88
4.5.2.3 Inelastic Behavior ........................................................................................... 88
4.5.3 Geometry............................................................................................................... 89
4.5.3.1 Small Deflection Theory ................................................................................. 89
4.5.3.2 Large Deflection Theory ................................................................................. 89
4.5.3.2.1 General ....................................................................................................... 89
4.5.3.2.2 Approximate Methods ................................................................................ 90
4.5.3.2.3 Refined Methods ........................................................................................ 90
4.5.4 Modeling Boundary Conditions ............................................................................ 91
4.5.5 Equivalent Members ............................................................................................. 91
4.6 STATIC ANALYSIS ...................................................................................................... 92
4.6.1 Influence of Plan Geometry .................................................................................. 92
4.6.1.1 Plan Aspect Ratio ............................................................................................ 92
4.6.1.2 Structures Curved in Plan ............................................................................... 92
4.6.1.2.1 General ................................................................not addressed in this Guide
4.6.1.2.2 Single-Girder Torsionally Stiff Superstructures ..not addressed in this Guide
4.6.1.2.3 Concrete Box Girder Bridges ..............................not addressed in this Guide
4.6.1.2.4 Steel Multiple-Beam Superstructures ..................not addressed in this Guide
4.6.1.2.4a General ............................................................not addressed in this Guide
4.6.1.2.4b I-Girders ..........................................................not addressed in this Guide
4.6.1.2.4c Closed Box and Tub Girders ...........................not addressed in this Guide
4.6.2 Approximate Methods of Analysis ....................................................................... 92
4.6.2.1 Decks............................................................................................................... 92
4.6.2.1.1 General ....................................................................................................... 92
4.6.2.1.2 Applicability ........................................................not addressed in this Guide
4.6.2.1.3 Width of Equivalent Interior Strips .....................not addressed in this Guide
4.6.2.1.4 Width of Equivalent Strips at Edges of Slabs .....not addressed in this Guide
4.6.2.1.4a General ............................................................not addressed in this Guide
4.6.2.1.4b Longitudinal Edges .........................................not addressed in this Guide
4.6.2.1.4c Transverse Edges ............................................not addressed in this Guide
4.6.2.1.5 Distribution of Wheel Loads ...............................not addressed in this Guide
4.6.2.1.6 Calculation of Force Effects ................................not addressed in this Guide
4.6.2.1.7 Cross-Sectional Frame Action.............................not addressed in this Guide

NSBA Guide to Navigating Routine Steel Bridge Design / 82


4.6.2.1.8 Live Load Force Effects for Fully and Partially Filled Grids and for
Unfilled Grid Decks Composite with Reinforced Concrete
Slabs ....................................................................not addressed in this Guide
4.6.2.1.9 Inelastic Analysis ................................................not addressed in this Guide
4.6.2.2 Beam-Slab Bridges ......................................................................................... 92
4.6.2.2.1 Application ................................................................................................. 92
4.6.2.2.2 Distribution Factor Method for Moment and Shear ................................... 93
4.6.2.2.3 Distribution Factor Method for Shear ........................................................ 96
4.6.2.2.4 Curved Steel Bridges .................................................................................. 99
4.6.2.2.5 Special Loads with Other Traffic ............................................................... 99
4.6.2.3 Equivalent Strip Widths for Slab-Type Bridges ........................................... 100
4.6.2.4 Truss and Arch Bridges ................................................................................ 100
4.6.2.5 Effective Length Factor, K ........................................................................... 100
4.6.2.6 Effective Flange Width ................................................................................. 101
4.6.2.6.1 General ..................................................................................................... 101
4.6.2.6.2 Segmental Concrete Box Beams and Single-Cell, Cast-in-Place Box Beams
.................................................................................................................. 101
4.6.2.6.3 Cast-in-Place Multicell Superstructures ................................................... 101
4.6.2.6.4 Orthotropic Steel Decks ........................................................................... 102
4.6.2.6.5 Transverse Floorbeams and Integral Bent Caps ....................................... 102
4.6.2.7 Lateral Wind Load Distribution in Girder System Bridges .......................... 102
4.6.2.7.1 I-Sections .................................................................................................. 102
4.6.2.7.2 Box Sections ............................................................................................. 102
4.6.2.7.3 Construction ............................................................................................. 103
4.6.2.8 Seismic Lateral Load Distribution ................................................................ 103
4.6.2.8.1 Applicability ............................................................................................. 103
4.6.2.8.2 Design Criteria ......................................................................................... 103
4.6.2.8.3 Load Distribution ..................................................................................... 104
4.6.2.9 Analysis of Segmental Concrete Bridges ..................................................... 104
4.6.2.9.1 General ..................................................................................................... 104
4.6.2.9.2 Strut-and-Tie Models ............................................................................... 104
4.6.2.9.3 Effective Flange Width ............................................................................ 104
4.6.2.9.4 Transverse Analysis ................................................................................. 104
4.6.2.9.5 Longitudinal Analysis .............................................................................. 105
4.6.2.10 Equivalent Strip Widths for Box Culverts .................................................... 105
4.6.2.10.1 General ..................................................................................................... 105
4.6.2.10.2 Case 1: Traffic Travels Parallel to Span ................................................... 106
4.6.2.10.3 Case 2: Traffic Travels Perpendicular to Span ......................................... 106
4.6.2.10.4 Precast Box Culverts ................................................................................ 106
4.6.3 Refined Methods of Analysis.............................................................................. 106
4.6.3.1 General .......................................................................................................... 106
4.6.3.2 Decks............................................................................................................. 108
4.6.3.2.1 General ..................................................................................................... 108
4.6.3.2.2 Isotropic Plate Model ............................................................................... 108
4.6.3.2.3 Orthotropic Plate Model ........................................................................... 108
4.6.3.2.4 Refined Orthotropic Deck Model ............................................................. 109

NSBA Guide to Navigating Routine Steel Bridge Design / 83


4.6.3.3 Beam-Slab Bridges ....................................................................................... 109
4.6.3.3.1 General ..................................................................................................... 109
4.6.3.3.2 Grid and Plate and Eccentric Beam Analyses of Curved and/or Skewed
Steel I-Girder Bridges .............................................................................. 109
4.6.3.3.3 Curved Steel Bridges ................................................................................ 110
4.6.3.3.4 Cross-Frames and Diaphragms ................................................................ 110
4.6.3.4 Cellular and Box Bridges .............................................................................. 111
4.6.3.5 Truss Bridges ................................................................................................ 111
4.6.3.6 Arch Bridges ................................................................................................. 111
4.6.3.7 Cable-Stayed Bridges.................................................................................... 111
4.6.3.8 Suspension Bridges ....................................................................................... 111
4.6.4 Redistribution of Negative Moments in Continuous Beam Bridges .................. 111
4.6.4.1 General .......................................................................................................... 111
4.6.4.2 Refined Method ............................................................................................ 113
4.6.4.3 Approximate Procedure ................................................................................ 113
4.6.5 Stability ............................................................................................................... 113
4.6.6 Analysis for Temperature Gradient .................................................................... 114
4.7 DYNAMIC ANALYSIS ............................................................................................... 115
4.7.1 Basic Requirements of Structural Dynamics ....................not addressed in this Guide
4.7.1.1 General .....................................................................not addressed in this Guide
4.7.1.2 Distribution of Masses .............................................not addressed in this Guide
4.7.1.3 Stiffness....................................................................not addressed in this Guide
4.7.1.4 Damping ...................................................................not addressed in this Guide
4.7.1.5 Natural Frequencies .................................................not addressed in this Guide
4.7.2 Elastic Dynamic Responses ..............................................not addressed in this Guide
4.7.2.1 Vehicle-Induced Vibration.......................................not addressed in this Guide
4.7.2.2 Wind-Induced Vibration ..........................................not addressed in this Guide
4.7.2.2.1 Wind Velocities ...................................................not addressed in this Guide
4.7.2.2.2 Dynamic Effects ..................................................not addressed in this Guide
4.7.2.2.3 Design Considerations .........................................not addressed in this Guide
4.7.3 Inelastic Dynamic Responses ...........................................not addressed in this Guide
4.7.3.1 General .....................................................................not addressed in this Guide
4.7.3.2 Plastic Hinges and Yield Lines ................................not addressed in this Guide
4.7.4 Analysis for Earthquake Loads ........................................................................... 115
4.7.4.1 General .......................................................................................................... 115
4.7.4.2 Single-Span Bridges...................................................................................... 116
4.7.4.3 Multispan Bridges ......................................................................................... 116
4.7.4.3.1 Selection of Method ............................................not addressed in this Guide
4.7.4.3.2 Single-Mode Methods of Analysis ......................not addressed in this Guide
4.7.4.3.2a General ............................................................not addressed in this Guide
4.7.4.3.2b Single-Mode Spectral Method ........................not addressed in this Guide
4.7.4.3.2c Uniform Load Method ....................................not addressed in this Guide
4.7.4.3.2 Multimode Spectral Method ................................not addressed in this Guide
4.7.4.3.3 Time-History Method ..........................................not addressed in this Guide
4.7.4.3.4a General ............................................................not addressed in this Guide

NSBA Guide to Navigating Routine Steel Bridge Design / 84


4.7.4.3.4b Acceleration Time Histories ...........................not addressed in this Guide
4.7.4.4 Minimum Support Length Requirements ..................................................... 116
4.7.4.5 P- Requirements .................................................................................117
4.7.5 Analysis for Collision Loads ............................................not addressed in this Guide
4.7.6 Analysis of Blast Effects ...................................................not addressed in this Guide
4.8 ANALYSIS BY PHYSICAL MODELS ................................not addressed in this Guide
4.8.1 Scale Model Testing .........................................................not addressed in this Guide
4.8.2 Bridge Testing ...................................................................not addressed in this Guide
4.9 REFERENCES .......................................................................not addressed in this Guide
APPENDIX A4 DECK SLAB DESIGN TABLE .........................not addressed in this Guide

NSBA Guide to Navigating Routine Steel Bridge Design / 85


4.4 ACCEPTABLE METHODS OF STRUCURAL ANALYSIS

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
This Article discusses acceptable methods of structural analysis. While many of these methods
could be used to analyze routine steel I-girder bridges, only “classical force and displacement
methods,” specifically line girder analysis, is necessary. While more refined, and more complex,
analysis methods might yield nominally more efficient designs, the benefits rarely justify the
increased level of effort for these types of structures. Additionally, most Owners prefer to use line
girder analysis methods for load rating routine steel I-girder bridges; they rarely have the time or
budget to use refined analysis for load rating the simpler structures in their inventory. “Optimizing”
a design through the use of refined analysis methods may result in a situation where the bridge
cannot be successfully rated in the future using the Owner’s typical line girder analysis rating
methods.

4.5 MATHEMATICAL MODELING

4.5.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article and its associated Commentary discuss a variety of modeling topics. Most of these
topics are not applicable to the design of routine steel I-girder bridges and should not be explored
for those types of designs.
This Article discusses the consideration of continuous composite barriers as part of the
superstructure stiffness in analysis. For the design of routine steel I-girder bridges, relying on
inclusion of barrier stiffness to meet deflection, service, and fatigue requirements is neither
necessary nor advisable. Inability to meet these requirements using the stiffness of the girders and
deck alone typically indicates that the depth, size, and/or spacing of the girders are inadequate.
The Article also mentions that the analysis should recognize the vertical freedom of the girder at
bearings where lift-off is indicated. This would necessitate consideration of geometric nonlinearity
in the analysis. Line girder analysis programs (such as NSBA's LRFD Simon line-girder analysis
and design program) typically cannot address nonlinear behavior, and the use of refined analysis
methods which could address nonlinear behavior is not warranted for routine steel I-girder bridges.
Instead, routine steel I-girder bridges should be designed such that they are not subject to lift-off
at any bearings. If the analysis of a routine steel I-girder bridge suggests there might be uplift/lift-
off at any bearing, steps should be taken to eliminate or prevent uplift/lift-off. Typically, if a design
is subject to uplift, it will be indicated in a line girder analysis if the net bearing reactions reported
by the analysis program are “negative” (i.e., if the reactions are acting to hold down the
superstructure at a given bearing). Note that “live load uplift” (i.e., “negative” bearing reactions
under the specific load case of live load alone) may not be a problem, as long as the uplift under

NSBA Guide to Navigating Routine Steel Bridge Design / 86


live load is less than the downward bearing reaction due to dead load (with appropriate load factors
applied to all load cases to determine the overall net reaction for the load combination).
Uplift at bearings (under either service level or factored loading) should not be permitted in a
routine steel I-girder bridge, and in fact most Owner-agencies have policies (either explicit, written
policies or implicit, unwritten policies) prohibiting uplift at bearings. If the line girder analysis
results are showing “negative” net reactions, steps should be taken to eliminate the uplift. The best
solution for addressing uplift in a routine steel I-girder bridge is to revise the span arrangement /
span lengths in such a way that net uplift does not occur. If this is not possible due to site
constraints, and uplift is unavoidable, Article 14.6.1 requires that bearings subject to net uplift at
any limit state be secured by tie-downs or anchorages, which are generally very complex, costly,
and maintenance-prone devices, and which complicate the behavior of the structure. Alternately,
counterweights could be provided to eliminate the uplift, but in most cases only relatively small
counterweights are actually practical.
The Commentary for this Article discusses the use of elastic and inelastic analysis assumptions.
Routine steel I-girder bridges should be analyzed as fully elastic for all limit states, per the
definition of these structures for the purposes of this Guide. Inelastic redistribution of negative
bending moments introduces levels of complexity and effort which are not warranted for routine
steel I-girder bridges.
The other items discussed in this Article are not applicable to the superstructure design of routine
steel I-girder bridges.

4.5.2 Structural Material Behavior

4.5.2.1 Elastic Versus Inelastic Behavior


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article defines the basic assumption of elastic-inelastic behavior of structural materials for
analysis. Inelastic behavior should not be considered for routine steel I-girder bridges.
For straight continuous-span steel I-girder bridges, the optional provisions of Appendix B6 of the
AASHTO LRFD BDS provide rational approaches for calculating moment redistribution from
interior-pier sections due to the effects of yielding (see the Discussion of Article 4.6.4.1 in this
Guide). This can potentially produce more economical designs, but at the cost of additional
analysis and design effort. These approaches utilize elastic moment envelopes, and do not require
the direct use of any inelastic analysis methods, but the associated analysis and design
considerations are unfamiliar to most designers and commercial software packages do not
currently include the capability to automate the associated calculations. As a result, designs which
rely on moment redistribution in order to satisfy AASHTO design criteria will also be more
difficult for the Owner-agency to load rate in the future. Therefore, most Owner-agencies currently
discourage or prohibit the use of moment redistribution methods for continuous-span steel I-girder
bridges.

NSBA Guide to Navigating Routine Steel Bridge Design / 87


Additionally, routine steel I-girder bridges, as defined for the purposes of this Guide, do not
generally need to recognize inelastic behavior in determining the resistance to extreme event
loadings (e.g., seismic loadings).

4.5.2.2 Elastic Behavior


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article discusses elastic behavior as well as stiffness properties of concrete.
The Article suggests that changes in material properties due to concrete maturity should be
included in analysis models, where appropriate. This statement warrants discussion and
clarification. In situations where the deck is placed in stages, specified by a deck-placement
sequence, the stiffness of previously-placed and cured sections of the deck should be considered
in the analysis, particularly with regard to dead load deflections used to develop the camber
diagrams and web-cutting ordinates for the girders, and especially in cases of irregular span
balance. Most Owner-agency policies and construction specifications require that previously-
placed sections of the deck achieve a specified minimum age, minimum strength, or other
performance criteria before placing the next section when a staged deck-placement sequence is
used; it is generally reasonable in that circumstance to assume the previously-placed sections of
the deck are close enough to “fully-hardened” to assume they have achieved full stiffness.
However, some Owner-agencies’ design policies include consideration of a reduced modulus for
previously-placed portions of the deck when a staged deck-placement sequence is used. See the
Discussion of Article 6.10.3.4.1 in this Guide for additional guidance on this topic.
Conversely, it is neither necessary nor recommended to address changes to material properties
during the curing cycle of the deck, as these changes have negligible effects on the behavior or
performance of the routine steel I-girder bridges covered in this Guide.
Also, when considering the appropriate level of refinement for the evaluation of dead load
deflections, designers are encouraged to keep in mind the typical construction tolerances
associated with steel girder fabrication, and the fact that the haunch between the deck and the
girder top flange serves to provide geometric adjustability in the field. Trying to quantify
deflections to an accuracy of a small fraction of an inch does not provide significant value.
The second paragraph of the Article states that the “stiffness characteristics of beam-slab-type
bridges may be based on full participation of concrete decks.” Most line girder analysis programs
(such as NSBA's LRFD Simon line-girder analysis and design program) assume the full, uncracked
concrete deck for computing stiffness along the girder, including in negative moment regions of
continuous span structures. This assumption is reasonable, is standard practice, and is supported
by the discussion in the Commentary for Article 6.10.1.5 (refer also to the Discussion of Article
6.10.1.5 in this Guide). Users should verify the capabilities, assumptions, and general correctness
of any program’s calculations prior to initial use. Article 6.10.1.5 in this Guide.

4.5.2.3 Inelastic Behavior


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 88


Discussion:
This Article discusses inelastic behavior and analysis methods. Inelastic behavior should not be
considered for routine steel I-girder bridges.
For straight continuous-span steel I-girder bridges, the optional provisions of Appendix B6 of the
AASHTO LRFD BDS provide rational approaches for calculating moment redistribution from
interior-pier sections due to the effects of yielding (see the Discussion of Articles 4.6.4.1 and
Appendix B6 in this Guide). This can potentially produce more economical designs, but at the cost
of additional analysis and design effort. These approaches utilize elastic moment envelopes, and
do not require the direct use of any inelastic analysis methods, but the associated analysis and
design considerations are unfamiliar to most designers and commercial software packages do not
currently include the capability to automate the associated calculations. As a result, designs which
rely on moment redistribution to satisfy AASHTO design criteria will also be more difficult for
the Owner-agency to load rate in the future. Therefore, most Owner-agencies currently discourage
or prohibit the use of moment redistribution methods for continuous-span steel I-girder bridges.

4.5.3 Geometry

4.5.3.1 Small Deflection Theory


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article discusses the limits for the consideration of small deflection theory (i.e., no
consideration of second-order effects). Line girder analysis is based on small deflection theory;
the moments and shears determined from the elastic analysis are not increased for any second-
order effects. Cross-frame axial wind loads calculated by hand are also not increased for second-
order effects; however, if the cross-frame members are connected eccentrically, second-order
effects must be considered for the resulting moments as detailed in the Discussion of Article
4.5.3.2.1 in this Guide.

4.5.3.2 Large Deflection Theory

4.5.3.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article discusses large deflection theory, which is generally not applicable to the
superstructure design of routine steel I-girder bridge. One exception is the design of eccentrically
connected cross-frame members in compression. Eccentrically connected members are subjected
to bending moments when axially loaded. For single-angle members, the resistance equations of
Article 6.9.4.4 directly account for these secondary forces and no further action is required.
However, for other members, such as WT shapes in compression connected only through their
flanges, the designer must account for the increase in moment due to secondary effects when

NSBA Guide to Navigating Routine Steel Bridge Design / 89


checking the moment-axial force interaction per Article 6.9.2.2. These secondary moments may
be approximated based on the provisions in Articles 4.5.3.2.2a and 4.5.3.2.2b.
Substructure units of routine steel I-girder bridges may be subjected to second-order effects as
discussed in this Article (for example, P-delta moment magnification effects in tall piers);
however, substructure design is beyond the scope of this Guide.

4.5.3.2.2 Approximate Methods

4.5.3.2.2a General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article simply identifies that the use of approximate methods for evaluating the effects of
deflections on beam-columns and arches is acceptable. The Commentary for this Article mentions
an alternate method, which is generally considered inappropriate and unnecessary for use in
routine steel I-girder bridges, and provides comments about limitations on actual movements
which apply to substructure design, not superstructure design.

4.5.3.2.2b Moment Magnification – Beam Columns


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article presents equations associated with an approximate method for moment magnification
of beam columns. This is particularly useful in the design of tall, slender substructure elements
(columns or piles) but that application is beyond the scope of this Guide. For the design of routine
steel I-girder bridges, these provisions are typically most applicable for the evaluation of tee
sections (WT members) with eccentric connections, such as those which may be used in truss-type
cross-frames.

4.5.3.2.2c Moment Magnification – Arches


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article specifically apply to the evaluation of arches, and are not applicable
to the design of routine steel I-girder bridges.

4.5.3.2.3 Refined Methods


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides general requirements for refined methods of analysis which may be used to
evaluate structures subject to second-order large deflection theory. The use of these types of

NSBA Guide to Navigating Routine Steel Bridge Design / 90


refined methods of analysis for the design of routine steel I-girder bridges is neither necessary nor
recommended.
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges.

4.5.4 Modeling Boundary Conditions


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article requires that boundary conditions represent the actual characteristics of support and
continuity. For routine steel I-girder bridges, analyzed using line girder analysis methods, simple
boundary conditions are assumed. Line girder analysis programs are one-dimensional analysis
methods; the boundary conditions are, by definition, located at the neutral axis of the beam. Line
girder analysis also only considers vertical (gravity) loading effects (dead load and live load).
Consequently, the boundary conditions for any line girder analysis are limited to the following:
• the vertical translation degree of freedom (DOF) is fixed at all supports;
• the longitudinal translation DOF is fixed at one support (which support does not matter),
in order to provide stability for the model;
• all other DOFs (transverse translation and all rotations) are free.
An actual routine steel I-girder bridge, obviously, is a three-dimensional construct, and the
bearings are typically located under the bottom flanges of the girders. In the actual structure, the
choice of which bearings are fixed against horizontal translation affects the performance of the
structure and influences the distribution of horizontal loads among the various bearings and
substructures.
Designers should not overthink the boundary conditions in a line girder analysis. But they should
think very carefully about how to establish the bearing articulation in the plans.
The AASHTO-NSBA Steel Bridge Collaboration Guideline G13.1-2019 Guidelines for Steel
Girder Bridge Analysis provides a discussion of boundary conditions in Section 3.14.

4.5.5 Equivalent Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article discusses the methods for modeling nonprismatic (varying) members. For constant-
depth girders, such as those used in routine steel I-girder bridges as defined for the purposes of
this Guide, most commercial line girder analysis methods (such as NSBA's LRFD Simon line-
girder analysis and design program) use appropriate means for discretizing the girder into

NSBA Guide to Navigating Routine Steel Bridge Design / 91


segments, accounting for changes in flange sizes, etc., such that the requirements of this provision
are met; users should verify the capabilities, assumptions, and general correctness of any
program’s calculations prior to initial use. Performing sensitivity studies of the degree of
discretization for specific bridge designs is not necessary, warranted, or recommended.

4.6 STATIC ANALYSIS

4.6.1 Influence of Plan Geometry

4.6.1.1 Plan Aspect Ratio


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article discusses the approximate modeling methods for torsionally-stiff closed cross-
sections, such as steel or concrete box-girders. I-girders are open cross-sections, and straight I-
girders like those used in routine steel I-girder bridges are quite flexible torsionally.

4.6.1.2 Structures Curved in Plan


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article and its associated sub-Articles discuss the approximate modeling methods for
structures curved in plan. For the purposes of this Guide, the definition of routine steel I-girder
bridges has been limited to structures which are straight in plan.

4.6.2 Approximate Methods of Analysis

4.6.2.1 Decks

4.6.2.1.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article and its associated sub-Articles address the design of bridge decks. The design of
concrete decks for steel I-girder bridges is typically governed by Owner-agency policy manuals
(e.g., standard designs, pre-calculated design tables, etc.), and so is not addressed herein.

4.6.2.2 Beam-Slab Bridges

4.6.2.2.1 Application
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 92


Discussion:
This Article provides the general bounding criteria for use of the AASHTO LRFD BDS
approximate live load distribution factors for moment and shear for various types of bridges.
Routine steel I-girder bridges as defined for the purposes of this Guide meet these general
bounding criteria.
The longitudinal stiffness parameter, Kg, (Eq. 4.6.2.2.1-1) and the appropriate value of the span
length, L, (Table 4.6.2.2.1-2) will need to be determined for use in Article 4.6.2.2.2. Routine steel
I-girder bridges fall under Table 4.6.2.2.1-1 Case (a).
When line-girder analyses are performed, proper distribution of the loads to the individual girders
becomes important to establish the total demand moments in the interior and exterior girders. Since
the routine steel I-girder bridges as defined for the purposes of this Guide meet the general
bounding criteria given in this Article, the 11th paragraph of this Article permits the DC loads
applied to the noncomposite section (referred to herein as DC1 loads) to be distributed equally to
all of the girders of the cross-section for the line-girder analysis. The DC1 loads, consist primarily
of the self-weight of the steel, the weight of any stay-in-place forms, and the weight of the wet
concrete in the deck. These loads should be assigned equally between all girders in the cross-
section if the girders are of approximately equal stiffness at the cross-frame connection points,
which is the case for the routine steel I-girder bridges defined for the purposes of this Guide. The
intermediate cross-frames or cross-frames act to equalize the girder deflections within a cross-
section and nearly equalize the load in equal-stiffness noncomposite girders regardless of the
amount of load applied to the individual girders. Using this assumption in these cases, in lieu of
the more traditional tributary area assumption applied to the weight of the wet deck concrete and
forms, is particularly important in helping to determine more accurate noncomposite deflections,
which are used in establishing girder cambers.
To better simulate the actual distribution of the barrier loads, or DC loads applied to the composite
section (referred to herein as DC2 loads) when line-girder analyses are performed, consider
assigning a larger percentage of these loads to the exterior girders and the adjacent interior girders,
which is a better assumption than a uniform distribution of these loads to all the girders based on
an examination of refined analysis results for several cases. Some Owner-agencies have a specific
design policy on the distribution of barrier loads; for example, one Owner-agency recommends,
for bridges over 44ˊ in width, distributing the barrier load to the first three exterior girders with
44% of the load applied to the exterior girder, 33% applied to the first interior girder, and 23%
applied to the second interior girder.
For the wearing surface load, DW, an equal distribution of the load to all the girders is a reasonable
assumption and has been the customary practice.

4.6.2.2.2 Distribution Factor Method for Moment and Shear

4.6.2.2.2a Interior Beams with Wood Decks


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 93


Discussion:
Wood decks are not included in the definition of routine steel I-girder bridges for the purposes of
this Guide.

4.6.2.2.2b Interior Beams with Concrete Decks


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article addressed the live load distribution factor for moment on interior beams. The
equations in the third row of Table 4.6.2.2.2b-1 (“Concrete Deck or Filled Grid, Partially Filled
Grid, or Unfilled Grid Deck Composite with Reinforced Concrete Slab on Steel or Concrete
Beams; Concrete T-beams, T- and Double T-sections”) are the only equations applicable for
calculation of live load distribution factors for moment on interior beams of routine steel I-girder
bridges. The routine steel I-girder bridges covered by this Guide satisfy the limitations specified
in the table for the use of these distribution factors.
Note that the live load distribution factor equations of Table 4.6.2.2.2b-1 inherently include
consideration of multiple presence (Article 3.6.1.1.2) as discussed in this Article, in Article
3.6.1.1.2, and associated Commentary for both articles (see also the Discussion of Article 3.6.1.1.2
in this Guide). When evaluating the live load distribution for interior girders at the strength and
service limit states, the live load distribution factors calculated from the formulas given in the table
should not be modified to account for multiple presence. However, as discussed in the
Commentary for Article 3.6.1.1.2 and in the Discussion of Article 3.6.1.1.2 in this Guide, the
multiple presence factor of 1.20 should be removed from the one-lane-loaded live load distribution
factor for interior girders calculated from the formula given in the table for evaluation of the fatigue
limit state.
Section 4.4.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures provides an extensive and helpful discussion of the
AASHTO LRFD BDS approximate live load distribution factors for moment in interior girders,
including example calculations.
Most commercial line girder analysis programs (such as NSBA's LRFD Simon line-girder analysis
and design program) automatically calculate the live load distribution factors necessary for the
analysis. Users should verify the capabilities, assumptions, and general correctness of any
program’s calculations of the live load distribution factors prior to initial use.

4.6.2.2.2c Interior Beams with Corrugated Steel Decks


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Corrugated steel decks are not included in the definition of routine steel I-girder bridges for the
purposes of this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 94


4.6.2.2.2d Exterior Beams
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article addressed the live load distribution factor for moment on exterior beams. The
equations the third row of Table 4.6.2.2.2d-1 (“Concrete Deck or Filled Grid, Partially Filled Grid,
or Unfilled Grid Deck Composite with Reinforced Concrete Slab on Steel or Concrete Beams;
Concrete T-beams, T- and Double T-sections”) are the only equations applicable for calculation
of live load distribution factors for moment on exterior beams of routine steel I-girder bridges. The
routine steel I-girder bridges covered by this Guide satisfy the limitations specified in the table for
the use of these distribution factors.
Note that the live load distribution factor equations of Table 4.6.2.2.2d-1 inherently include
consideration of multiple presence (Article 3.6.1.1.2) as discussed in this Article, in Article
3.6.1.1.2, and associated Commentary for both articles (see also the Discussion of Article 3.6.1.1.2
in this Guide). When evaluating live load distribution for exterior girders at the strength and service
limit states, the live load distribution factor calculated from the formula given in the table for the
case of two or more lanes loaded should not be modified to account for multiple presence.
For situations where only one design lane is loaded, the lever rule is used to calculate the
distribution factor for moment in an exterior girder. For further description of the lever rule, see
the Commentary for Article 4.6.2.2.1. The provisions of Article 3.6.1.1.1 regarding the placement
of the design lanes and the placement of the wheel loads within those lanes should be followed
when utilizing the lever rule. When evaluating the live load distribution for exterior girders for
one-lane loaded at the fatigue limit state utilizing the lever rule, the multiple presence factor of 1.2
should not be applied. When evaluating the live load distribution for exterior girders utilizing the
lever rule for situations where only one design lane is loaded at the strength and service limit states,
the appropriate multiple presence factor specified in Table 3.6.1.1.2-1 must be applied. The
presence or absence of cross-frames or diaphragms is not considered when calculating distribution
factors using the lever rule, which only considers the deck acting as a lever supported by the
exterior and first interior girder.
In addition, this Article specifies that for steel bridge cross-sections with cross-frames or
diaphragms, the live load distribution factor for the exterior girder is not to be taken less than that
which would be obtained by assuming the cross-section deflects and rotates as a rigid cross-
section. This special analysis is specified because the empirical distribution factors for moment
given in the specification table were determined without consideration of cross-frames or
diaphragms; hence, while they are conservative for interior girders, they are generally
unconservative for exterior girders in steel multi-girder bridges. Therefore, the distribution factor
for moment in the exterior girders determined from this special analysis will usually control and
should always be employed for routine steel I-girder bridges since the exterior girder is typically
the critical girder for moment. It is recommended that Eq. C4.6.2.2.2d-1 be used to satisfy this
assumption; the equation should be evaluated for one lane loaded and also for two or more lanes
loaded (up to the total number of design lanes the design roadway width can accommodate). The
provisions of Article 3.6.1.1.1 regarding the placement of the design lanes and the placement of

NSBA Guide to Navigating Routine Steel Bridge Design / 95


the wheel loads within those lanes should also be followed. When evaluating the live load
distribution for exterior girders for one-lane loaded at the fatigue limit state utilizing the special
analysis, the multiple presence factor of 1.2 should not be applied. When evaluating the live load
distribution for exterior girders for any number of design lanes loaded at the strength and service
limit states utilizing the special analysis, the appropriate multiple presence factor specified in Table
3.6.1.1.2-1 must be applied.
Section 4.4.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures provides an extensive and helpful discussion of the
AASHTO LRFD BDS approximate live load distribution factors for moment in exterior girders,
including example calculations utilizing the specification formulas, the lever rule, and the special
rigid cross-section analysis.
Most commercial line girder analysis programs (such as NSBA's LRFD Simon line-girder analysis
and design program) automatically calculate the live load distribution factors necessary for the
analysis. Users should verify the capabilities, assumptions, and general correctness of any
program’s calculations of the live load distribution factors prior to initial use. Note that the LRFD
Simon program does not currently perform the special rigid cross-section analysis.

4.6.2.2.2e Skewed Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides factors which may be used to reduce the live load distribution factor for
moment in the design of skewed bridges. The use of these particular factors is typically at the
discretion of the Owner-agency. However, the range of applicability for these provisions is limited
to bridges with support skew angles between 30 and 60 degrees. Since the routine steel I-girder
bridges as defined for the purposes of this Guide are limited to support skew angles of 20 degrees
or less, the reduction factors discussed in this Article cannot be used regardless of the Owner-
agency policy.

4.6.2.2.2f Flexural Moments and Shear in Transverse Floorbeams


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Routine steel I-girder bridges as defined for this Guide do not have transverse floorbeams.

4.6.2.2.3 Distribution Factor Method for Shear

4.6.2.2.3a Interior Beams


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 96


Discussion:
This Article addressed the live load distribution factor for shear on interior beams. The equations
in the third row of Table 4.6.2.2.3a-1 (“Concrete Deck or Filled Grid, Partially Filled Grid, or
Unfilled Grid Deck Composite with Reinforced Concrete Slab on Steel or Concrete Beams;
Concrete T-beams, T- and Double T-sections”) are the only equations applicable for calculation
of live load distribution factors for shear on interior beams of routine steel I-girder bridges. The
routine steel I-girder bridges covered by this Guide satisfy the limitations specified in the table for
the use of these distribution factors.
Note that the live load distribution factor equations of Table 4.6.2.2.3a-1 inherently include
consideration of multiple presence (Article 3.6.1.1.2) as discussed in this Article, in Article
3.6.1.1.2, and associated Commentary for both articles (see also the Discussion of Article 3.6.1.1.2
in this Guide). When evaluating the live load distribution for interior girders at the strength and
service limit states, the live load distribution factors calculated from the formulas given in the table
should not be modified to account for multiple presence. However, as discussed in the
Commentary for Article 3.6.1.1.2 and in the Discussion of Article 3.6.1.1.2 in this Guide, the
multiple presence factor of 1.20 should be removed from the one-lane-loaded live load distribution
factor for interior girders calculated from the formula given in the table for evaluation of the special
fatigue requirement for webs specified in Article 6.10.5.3 (see the Discussion of Article 6.10.5.3
in this Guide). Similarly, the multiple presence factor of 1.20 should also be removed for the
determination of the longitudinal shear range for the fatigue design of shear connectors (see the
Discussion of Article 6.10.10.1.2 in this Guide). Note that the interior girder is typically the critical
girder for shear in a routine steel I-girder bridge.
Section 4.4.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures provides an extensive and helpful discussion of the
AASHTO LRFD BDS approximate live load distribution factors for shear in interior girders,
including example calculations.
Most commercial line girder analysis programs (such as NSBA's LRFD Simon line-girder analysis
and design program) automatically calculate the live load distribution factors necessary for the
analysis. Users should verify the capabilities, assumptions, and general correctness of any
program’s calculations of the live load distribution factors prior to initial use.

4.6.2.2.3b Exterior Beams


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article addressed the live load distribution factor for shear on exterior beams. The equations
in the third row of Table 4.6.2.2.3b-1 (“Concrete Deck or Filled Grid, Partially Filled Grid, or
Unfilled Grid Deck Composite with Reinforced Concrete Slab on Steel or Concrete Beams;
Concrete T-beams, T- and Double T-sections”) are the only equations applicable for calculation
of live load distribution factors for shear on exterior beams of routine steel I-girder bridges. The
routine steel I-girder bridges covered by this Guide satisfy the limitations specified in the table for
the use of these distribution factors.

NSBA Guide to Navigating Routine Steel Bridge Design / 97


Note that the live load distribution factor equations of Table 4.6.2.2.3b-1 inherently include
consideration of multiple presence (Article 3.6.1.1.2) as discussed in this Article, in Article
3.6.1.1.2, and associated Commentary for both articles (see also the Discussion of Article 3.6.1.1.2
in this Guide). When evaluating the live load distribution for exterior girders at the strength and
service limit states, the live load distribution factor calculated from the formula for the case of two
or more lanes loaded should not be modified to account for multiple presence.
For situations where only one design lane is loaded, the lever rule is used to calculate the
distribution factor. For further description of the lever rule, see the Commentary for Article
4.6.2.2.1. The provisions of Article 3.6.1.1.1 regarding the placement of the design lanes and the
placement of the wheel loads within those lanes should be followed when utilizing the lever rule.
When utilizing the lever rule to determine the live load distribution for exterior girders for one-
lane loaded for evaluation of the special fatigue requirement for webs specified in Article 6.10.5.3
(see the Discussion of Article 6.10.5.3 in this Guide) and for the determination of the longitudinal
shear range for the fatigue design of shear connectors (see the Discussion of Article 6.10.10.1.2 in
this Guide), the multiple presence factor of 1.2 should not be applied. When utilizing the lever rule
to determine the live load distribution for exterior girders for situations where only one design lane
is loaded at the strength and service limit states, the appropriate multiple presence factor specified
in Table 3.6.1.1.2-1 must be applied. The presence or absence of cross-frames or diaphragms is
not considered when calculating distribution factors using the lever rule, which only considers the
deck acting as a lever supported by the exterior and first interior girder.
In addition, this Article specifies that for steel bridge cross-sections with cross-frames or
diaphragms, the live load distribution factor for the exterior girder is not to be taken less than that
which would be obtained by assuming the cross-section deflects and rotates as a rigid cross-
section. It is recommended that Eq. C4.6.2.2.2d-1 be used to satisfy this assumption; the equation
should be evaluated for one lane loaded and also for two or more lanes loaded (up to the total
number of design lanes the design roadway width can accommodate). The provisions of Article
3.6.1.1.1 regarding the placement of the design lanes and the placement of the wheel loads within
those lanes should also be followed. When utilizing the special analysis to determine the live load
distribution for exterior girders for one-lane loaded for evaluation of the special fatigue
requirement for webs specified in Article 6.10.5.3 (see the Discussion of Article 6.10.5.3 in this
Guide) and for the determination of the longitudinal shear range for the fatigue design of shear
connectors (see the Discussion of Article 6.10.10.1.2 in this Guide), the multiple presence factor
of 1.2 should not be applied. When utilizing the special analysis to evaluate the live load
distribution for exterior girders for any number of design lanes loaded at the strength and service
limit states, the appropriate multiple presence factor specified in Table 3.6.1.1.2-1 must be applied.
Note that the special rigid-cross section analysis typically does not control for shear.
Section 4.4.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures provides an extensive and helpful discussion of the
AASHTO LRFD BDS approximate live load distribution factors for shear in exterior girders,
including example calculations utilizing the specification formulas, the lever rule, and the special
rigid cross-section analysis.

NSBA Guide to Navigating Routine Steel Bridge Design / 98


Most commercial line girder analysis programs (such as NSBA's LRFD Simon line-girder analysis
and design program) automatically calculate the live load distribution factors necessary for the
analysis. Users should verify the capabilities, assumptions, and general correctness of any
program’s calculations of the live load distribution factors prior to initial use. Note that the LRFD
Simon program does not currently perform the special rigid cross-section analysis.

4.6.2.2.3c Skewed Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article contains the equations for the correction factor for live load distribution factors for
shear in girders at and adjacent to the obtuse corners of skewed supports; the factor is to be applied
at all skewed supports (i.e., at both end and interior skewed supports). The equations the first row
of Table 4.6.2.2.3c-1 (“Concrete Deck or Filled Grid, Partially Filled Grid, or Unfilled Grid Deck
Composite with Reinforced Concrete Slab on Steel or Concrete Beams; Concrete T-beams, T- and
Double T-sections”) are the only equations applicable for calculation of correction factors for
routine steel I-girder bridges. These correction factors apply only to the shear distribution factors
for the exterior and first interior girder at and adjacent to the obtuse corner of the end and/or interior
support and decrease linearly to 1.0 at midspan. The routine steel I-girder bridges covered by this
Guide satisfy the limitations specified in the table for the use of the correction factor, and the
correction factor is applicable to routine steel I-girder bridges with skewed supports. Be aware that
a correction factor is not provided for the dead load shears at skewed supports.
Section 4.4.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures provides an extensive and helpful discussion of the
AASHTO LRFD BDS approximate live load distribution factors, including example calculations.
Most commercial line girder analysis programs (such as NSBA's LRFD Simon line-girder analysis
and design program) automatically calculate the live load distribution factors necessary for the
analysis. Users should verify the capabilities, assumptions, and general correctness of any
program’s calculations of the live load distribution factors prior to initial use.

4.6.2.2.4 Curved Steel Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
By the definition used in this Guide, routine steel I-girder bridges are not curved.

4.6.2.2.5 Special Loads with Other Traffic


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article only applies when the Owner-agency does not have an overriding policy on how to
address situations where one lane is loaded with overweight or permit vehicles mixed with routine

NSBA Guide to Navigating Routine Steel Bridge Design / 99


traffic in the other lanes. The live load distribution equation in this Article is to be used only for
cases involving two or more design lanes and is not to be used when use of the lever rule or the
rigid cross-section assumption is required by the related Articles for the calculation of live load
distribution factors since both of those methods could potentially be used to compute the
distribution factor directly.
Section 4.4.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures provides an extensive and helpful discussion of the
AASHTO LRFD BDS approximate live load distribution factors, including example calculations.
Commercial line girder analysis programs may not have the built-in capability to address the case
of special loads combined with routine traffic. Some programs may allow the user to overwrite the
program’s calculated live load distribution factors and substitute factors calculated by the user
outside of the program. Users should verify the capabilities, assumptions, and general correctness
of any program’s calculations of the live load distribution factors prior to initial use.

4.6.2.3 Equivalent Strip Widths for Slab-Type Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article only applies to slab-type bridges, which are bridges where the main spanning element
is a concrete or wood slab, without supporting girders, beams or stringers. Routine steel I-girder
bridges are categorized as “beam-slab bridges” in the AASHTO LRFD BDS.

4.6.2.4 Truss and Arch Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article only applies to truss and arch bridges. Routine steel I-girder bridges are categorized
as “beam-slab bridges” in the AASHTO LRFD BDS.

4.6.2.5 Effective Length Factor, K


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides guidance on the effective length factor, K, based on various end conditions
for the design of compression members. For the design of superstructures for routine steel I-girder
bridges, the provisions in this Article apply primarily to the design of truss-type cross-frame
members; the appropriate approximate values given in the bulleted items in this Article are
typically used. The provisions of this article do not apply to the design of the girders.
The provisions of this Article also apply to a variety of other structural elements which may be
present in a bridge, such as columns or piles, but the design of substructure and foundation
elements is beyond the scope of this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 100


4.6.2.6 Effective Flange Width

4.6.2.6.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
The provisions in this Article are used to determine the “effective flange width” of the concrete
deck. The effective flange width is used for computing the composite section properties for the
composite girder cross-section for determining the composite cross-section stiffness for the
analysis and for determining the flexural resistance of the composite section (see the Discussion
of Article 6.10.1.1 and the associated sub-Articles in this Guide for further information on the
computation of composite section properties).
In composite steel girders subject to major-axis bending, longitudinal stresses are distributed to
the various components of the cross-section, including the concrete deck, by in-plane shear stresses
resulting in shear deformations. As a result of the corresponding shear deformations in the deck –
which is wider and less efficient than the steel girder in distributing the stresses -- plane sections
do not remain plane and the longitudinal stresses across the deck are non-uniform; a phenomenon
referred to as shear lag. The effective flange width is the width of deck over which the assumed
uniformly distributed longitudinal stresses result in approximately the same deck force and
member moments calculated from elementary beam theory (i.e. assuming plane sections remain
plane) as would be produced by the actual non-uniform stress distribution. As described in the first
paragraph of this Article, for the routine steel I-girder bridges covered by this Guide, the effective
flange width of the concrete deck should be taken as the corresponding tributary width of the deck
perpendicular to the axis of the member. Provisions related to other types of systems mentioned in
the remainder of this Article are not applicable. The provisions which allow for extending the
deck overhang width used for the analysis to account for the presence of a continuous concrete
barrier rail should not be used for routine steel I-girder bridge design. If design requirements cannot
be met using the section properties and associated strength of the girder and deck alone, this
typically indicates that the depth, size, and/or spacing of the girders are inadequate.

4.6.2.6.2 Segmental Concrete Box Beams and Single-Cell, Cast-in-Place Box Beams
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
These provisions do not apply to the design of steel I-girder bridges.

4.6.2.6.3 Cast-in-Place Multicell Superstructures


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
These provisions do not apply to the design of steel I-girder bridges.

NSBA Guide to Navigating Routine Steel Bridge Design / 101


4.6.2.6.4 Orthotropic Steel Decks
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
These provisions do not apply to the design of the routine steel I-girder bridges covered by this
Guide, which are assumed not to have orthotropic steel decks.

4.6.2.6.5 Transverse Floorbeams and Integral Bent Caps


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
These provisions do not apply to the design of the routine steel I-girder bridges covered by this
Guide, which are assumed not to have transverse floorbeams or integral bent caps.

4.6.2.7 Lateral Wind Load Distribution in Girder System Bridges

4.6.2.7.1 I-Sections
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article addressed the loading of wind perpendicular to the span of the superstructure. Routine
steel I-girder bridges, in their final fully-constructed condition, have a concrete deck that acts as a
lateral diaphragm to transmit lateral wind loads from the top half of the girder directly through the
deck. The lateral wind loads on the bottom half of the girder are resisted by lateral bending in the
bottom flange and transmitted up to the deck by the cross frames or diaphragms. At the support
locations, the transverse wind loads are transmitted from the deck to the support through the cross-
frames or diaphragms at those locations.
The Commentary for this Article provides a simplified procedure for calculating the various
horizontal loading effects associated with transmitting wind loads through the load paths discussed
above. The equations in the procedure are derived from classical equations for moments and
reactions in beams, occasionally with modified factors to account for some degree of continuity in
the girder flange acting as a beam supported at multiple points by the cross-frames.
Sections 3.5 and 6.5.6.5.1 of the Reference Manual for NHI Course 130081, Load and Resistance
Factor Design (LRFD) for Highway Bridge Superstructures provides an extensive and helpful
discussion of the evaluation of the effects of wind loading on the superstructure of steel I-girder
bridges, including explanation and background of the provisions of this Article and example
calculations.

4.6.2.7.2 Box Sections


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 102


Discussion:
These provisions do not apply to the design of steel I-girder bridges.

4.6.2.7.3 Construction
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
During construction, the absence of the hardened, composite concrete deck means that routine
steel I-girder bridges behave as structures lacking a deck to provide the horizontal diaphragm
action discussed in Article 4.6.2.7.1. Consequently, the design of routine steel I-girder bridges
should include constructibility checks of the steel superstructure in the non-composite condition
to resist lateral wind loads, in conjunction with other construction loads, as applicable. These
constructibility checks and their associated loads are mentioned in other Discussions in this Guide,
such as the Discussion of Article 3.4.2.1, the Discussion of Article 3.8.1.2.2, and the Discussion
of Article 6.10.3.1. The Reference Manual for NHI Course 130081, Load and Resistance Factor
Design (LRFD) for Highway Bridge Superstructures and the Reference Manual for NHI Course
130102, Engineering for Structural Stability in Bridge Construction provide discussion of
constructibility checks and the associated loads.
Sections 3.5, 6.3.2.10.2.1, 6.5.3.1, 6.5.3.6, and 6.5.6.5.1 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures provide
extensive and helpful discussions of how to approach the evaluation of wind loads during
construction. These discussions focus on investigations into the possible need to provide lateral
bracing to help resist wind loads and limit lateral displacements of the girders prior to the
placement of the concrete deck and an example calculation is included.

4.6.2.8 Seismic Lateral Load Distribution

4.6.2.8.1 Applicability
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article are only applicable for bridges in Seismic Zones, 2, 3, or 4. For the
purposes of this Guide, the definition of routine steel I-girder bridges only includes bridges in
Seismic Zone 1. Consequently, detailed discussion of this Article is beyond the scope of this
Guide.

4.6.2.8.2 Design Criteria


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.

NSBA Guide to Navigating Routine Steel Bridge Design / 103


Discussion:
The provisions of this Article are only applicable for bridges in Seismic Zones, 2, 3, or 4. For the
purposes of this Guide, the definition of routine steel I-girder bridges only includes bridges in
Seismic Zone 1. Consequently, detailed discussion of this Article is beyond the scope of this
Guide.

4.6.2.8.3 Load Distribution


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article are only applicable for bridges in Seismic Zones, 2, 3, or 4. For the
purposes of this Guide, the definition of routine steel I-girder bridges only includes bridges in
Seismic Zone 1. Consequently, detailed discussion of this Article is beyond the scope of this
Guide.

4.6.2.9 Analysis of Segmental Concrete Bridges

4.6.2.9.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of segmental concrete bridges, and do not
apply to the design of steel I-girder bridges.

4.6.2.9.2 Strut-and-Tie Models


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of segmental concrete bridges, and do not
apply to the design of steel I-girder bridges.

4.6.2.9.3 Effective Flange Width


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of segmental concrete bridges, and do not
apply to the design of steel I-girder bridges.

4.6.2.9.4 Transverse Analysis


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 104


Discussion:
The provisions of this Article only apply to the analysis of segmental concrete bridges, and do not
apply to the design of steel I-girder bridges.

4.6.2.9.5 Longitudinal Analysis


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of segmental concrete bridges, and do not
apply to the design of steel I-girder bridges.

4.6.2.9.5a General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of segmental concrete bridges, and do not
apply to the design of steel I-girder bridges.

4.6.2.9.5b Erection Analysis


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of segmental concrete bridges, and do not
apply to the design of steel I-girder bridges.

4.6.2.9.5c Analysis of the Final Structural System


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of segmental concrete bridges, and do not
apply to the design of steel I-girder bridges.

4.6.2.10 Equivalent Strip Widths for Box Culverts

4.6.2.10.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of concrete box culverts, and do not apply
to the design of steel I-girder bridges. The provisions of this Article do not even apply to the design
of decks for routine steel I-girder bridges, or the decks of any other girder bridge type.

NSBA Guide to Navigating Routine Steel Bridge Design / 105


4.6.2.10.2 Case 1: Traffic Travels Parallel to Span
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of concrete box culverts, and do not apply
to the design of steel I-girder bridges. The provisions of this Article do not even apply to the design
of decks for routine steel I-girder bridges, or the decks of any other girder bridge type.

4.6.2.10.3 Case 2: Traffic Travels Perpendicular to Span


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of concrete box culverts, and do not apply
to the design of steel I-girder bridges. The provisions of this Article do not even apply to the design
of decks for routine steel I-girder bridges, or the decks of any other girder bridge type.

4.6.2.10.4 Precast Box Culverts


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article only apply to the analysis of concrete box culverts, and do not apply
to the design of steel I-girder bridges. The provisions of this Article do not even apply to the design
of decks for routine steel I-girder bridges, or the decks of any other girder bridge type.

4.6.3 Refined Methods of Analysis

4.6.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges.
Note that there are several commercial line girder analysis programs available to help automate
and streamline the analysis and design of routine steel I-girder bridges, including NSBA's LRFD
Simon line-girder analysis and design program. Users should verify the capabilities, assumptions,
and general correctness of any program’s calculations prior to initial use.
Refined methods of analysis are formally defined in the AASHTO LRFD BDS as "Methods of
structural analysis that consider the entire superstructure as an integral unit and provide the
required deflections and actions.” More to the point, refined methods of analysis, in the context of

NSBA Guide to Navigating Routine Steel Bridge Design / 106


girder bridges (or “beam-slab bridges”), directly model the girders, the cross-frames/diaphragms,
the deck, and their interaction. In a refined analysis model, the distribution of vertical loads (gravity
loads) is determined by consideration of the relative stiffness of the girders, cross-
frames/diaphragms, and deck. Examples of refined methods of analysis for steel girder bridges
include 2D grid or grillage analysis, 2D plate-and-eccentric-beam analysis, and 3D finite element
analysis.
Approximate methods of analysis, on the other hand, typically consider only an individual girder,
isolated from the rest of the structural system. The distributions of dead loads such as the weight
of the wet concrete deck, future wearing surface, and barrier rails are typically determined using
approximations such as uniform distribution of the deck weight to the girders, or semi-arbitrary
percentage distribution factors. The distribution of live loads is typically determined using
approximate live load distribution factors such as those found in Article 4.6.2.2 and its associated
sub-Articles. The most common example of an approximate method of analysis for steel girder
bridges is line girder analysis.
A routine steel I-girder bridge could be analyzed using a refined method of analysis; doing so
would result in a more refined, theoretically less conservative distribution of loads and a more
refined, theoretically less conservative determination of design force effects in individual
structural elements such as the girders and cross-frames. However, the use of refined methods of
analysis takes more time and effort on the part of the designer, and involves the use of more
complicated analysis models, which are more prone to the introduction of inadvertent modeling
errors and which are more difficult to check. Approximate methods of analysis, specifically line
girder analysis, are simpler to use, involving much less effort and time on the part of the designer.
Line girder analysis is also less complicated, less error-prone, and easier to check. While the use
of a line girder analysis may introduce some measure of conservatism, particularly in the
distribution of live loads, this additional conservatism is not excessive and is considered well
within the standards of acceptability in the bridge design industry. The extra refinement associated
with the use of a refined method of analysis versus an approximate method of analysis is simply
not worth the additional analysis time, cost, and complexity for the design of routine steel I-girder
bridges; the extra time and effort that would be expended performing a refined analysis for a
routine steel I-girder bridge would be far better spent on other activities, such as optimizing the
framing plan or the girder sizing, preparing a clear and well-laid out set of plans, and/or performing
robust checking and quality control reviews.
Finally, and importantly, using a refined method of analysis to primarily decrease conservatism in
live load distribution could lead a designer to reduce girder flange or web sizes during the initial
design of a bridge. This could result in difficulties later when the Owner-agency performs periodic
routine load rating analyses of the bridge. For the sake of practicality, most Owner-agencies default
to using line girder analysis methods for these load rating analyses – they have hundreds or
thousands of bridges to load rate each year and cannot afford to perform labor-intensive refined
analyses when line girder analysis methods would suffice. It is problematic when a bridge exhibits
an insufficient load rating due solely to the minor conservatism of line girder analysis methods,
forcing the Owner-agency to invest limited resources in performing a refined analysis to
demonstrate that a bridge has sufficient load-carrying capacity.

NSBA Guide to Navigating Routine Steel Bridge Design / 107


4.6.3.2 Decks

4.6.3.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges. See the Discussion of Article 4.6.3.1 for more information about
refined versus approximate methods of analysis.
Note that there are several commercial line girder analysis programs available to help automate
and streamline the analysis and design of routine steel I-girder bridges, including NSBA's LRFD
Simon line-girder analysis and design program. Users should verify the capabilities, assumptions,
and general correctness of any program’s calculations prior to initial use.

4.6.3.2.2 Isotropic Plate Model


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges. See the Discussion of Article 4.6.3.1 for more information about
refined versus approximate methods of analysis.
Note that there are several commercial line girder analysis programs available to help automate
and streamline the analysis and design of routine steel I-girder bridges, including NSBA's LRFD
Simon line-girder analysis and design program. Users should verify the capabilities, assumptions,
and general correctness of any program’s calculations prior to initial use.

4.6.3.2.3 Orthotropic Plate Model


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges. See the Discussion of Article 4.6.3.1 for more information about
refined versus approximate methods of analysis.

NSBA Guide to Navigating Routine Steel Bridge Design / 108


Note that there are several commercial line girder analysis programs available to help automate
and streamline the analysis and design of routine steel I-girder bridges, including NSBA's LRFD
Simon line-girder analysis and design program. Users should verify the capabilities, assumptions,
and general correctness of any program’s calculations prior to initial use.

4.6.3.2.4 Refined Orthotropic Deck Model


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges. See the Discussion of Article 4.6.3.1 for more information about
refined versus approximate methods of analysis.
Note that there are several commercial line girder analysis programs available to help automate
and streamline the analysis and design of routine steel I-girder bridges, including NSBA's LRFD
Simon line-girder analysis and design program. Users should verify the capabilities, assumptions,
and general correctness of any program’s calculations prior to initial use.

4.6.3.3 Beam-Slab Bridges

4.6.3.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges. See the Discussion of Article 4.6.3.1 for more information about
refined versus approximate methods of analysis.
Note that there are several commercial line girder analysis programs available to help automate
and streamline the analysis and design of routine steel I-girder bridges, including NSBA's LRFD
Simon line-girder analysis and design program. Users should verify the capabilities, assumptions,
and general correctness of any program’s calculations prior to initial use.

4.6.3.3.2 Grid and Plate and Eccentric Beam Analyses of Curved and/or Skewed Steel I-Girder
Bridges
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 109


Discussion:
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges. See the Discussion of Article 4.6.3.1 for more information about
refined versus approximate methods of analysis.
Note that there are several commercial line girder analysis programs available to help automate
and streamline the analysis and design of routine steel I-girder bridges, including NSBA's LRFD
Simon line-girder analysis and design program. Users should verify the capabilities, assumptions,
and general correctness of any program’s calculations prior to initial use.

4.6.3.3.3 Curved Steel Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges. See the Discussion of Article 4.6.3.1 for more information about
refined versus approximate methods of analysis.
Note that there are several commercial line girder analysis programs available to help automate
and streamline the analysis and design of routine steel I-girder bridges, including NSBA's LRFD
Simon line-girder analysis and design program. Users should verify the capabilities, assumptions,
and general correctness of any program’s calculations prior to initial use.

4.6.3.3.4 Cross-Frames and Diaphragms


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A routine steel I-girder bridge, as defined for the purposes of this Guide, is configured in such a
manner that refined methods of analysis are not required for its design; instead, routine steel I-
girder bridges can, and should, be designed using approximate methods of analysis, specifically
line girder analysis methods. As a result, the provisions of this Article do not apply to the design
of routine steel I-girder bridges. See the Discussion of Article 4.6.3.1 for more information about
refined versus approximate methods of analysis.
Note that there are several commercial line girder analysis programs available to help automate
and streamline the analysis and design of routine steel I-girder bridges, including NSBA's LRFD
Simon line-girder analysis and design program. Users should verify the capabilities, assumptions,
and general correctness of any program’s calculations prior to initial use.

NSBA Guide to Navigating Routine Steel Bridge Design / 110


4.6.3.4 Cellular and Box Bridges
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article does not apply to the design of routine steel I-girder bridges, or to any beam-slab,
girder-type bridge.

4.6.3.5 Truss Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article does not apply to the design of routine steel I-girder bridges, or to any beam-slab,
girder-type bridge.

4.6.3.6 Arch Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article does not apply to the design of routine steel I-girder bridges, or to any beam-slab,
girder-type bridge.

4.6.3.7 Cable-Stayed Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article does not apply to the design of routine steel I-girder bridges, or to any beam-slab,
girder-type bridge.

4.6.3.8 Suspension Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article does not apply to the design of routine steel I-girder bridges, or to any beam-slab,
girder-type bridge.

4.6.4 Redistribution of Negative Moments in Continuous Beam Bridges

4.6.4.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 111


Discussion:
Allowing redistribution of negative moments in multi-span continuous steel I-girder bridges can
potentially produce more economical designs, but the associated analysis and design
considerations are unfamiliar to most designers and most Owner-agencies currently do not permit
or encourage the use of moment redistribution methods. As a result, the use of moment
redistribution methods has been specifically excluded from the scope of this Guide for the design
of routine steel I-girder bridges.
A basic explanation of moment distribution methods and their associated advantages and
disadvantages is provided below for information only.
In conventional elastic analysis and design, moment and shear envelopes are typically determined
by elastic analysis with no consideration of redistribution due to the effects of yielding. Even if
localized yielding at some section is permitted for a specific girder design check, redistribution of
moments and shears to account for such local yielding is not addressed by the basic provisions of
Section 6 of the AASHTO LRFD BDS. Under these provisions, when performing conventional
elastic analysis and design, the girder sections must be proportioned to provide resistance equal to
or greater than that required by the moment and shear envelopes determined by the elastic analysis.
The requirement to meet these moment and shear demands can lead to designs which are less than
optimal from a performance or economical standpoint. For instance, the designer may choose to
use oversized sections or reduce beam spacing and add more beams to the cross section. In welded
beams, multiple flange transitions might be added, resulting in increased fabrication costs.
On the other hand, accounting for the redistribution of moments (where appropriate) can make it
possible to eliminate such details by using prismatic sections along the entire length of the bridge
or between field splices, thus providing fabrication economies and improving the overall fatigue
resistance. This is made possible by removing restrictions on the flexural resistance in the regions
adjacent to interior piers from which moments are redistributed at both the service and strength
limit states by accounting for the strength and ductility of the pier sections directly within the
procedures used to calculate the redistribution moments.
For straight continuous-span steel I-girder bridges, the optional Appendix B6 of the AASHTO
LRFD BDS provides both an approximate procedure and a refined method to calculate the
redistribution moments (see the Discussion of Appendix B6 in this Guide). The provisions of
Appendix B6 may be applied only to straight continuous-span I-section members whose support
lines are not skewed more than 10 degrees from radial and along which there are no staggered
cross-frames. Cross-sections throughout the unbraced lengths immediately adjacent to interior-
pier sections from which moments are redistributed must also satisfy certain specified restrictions
to provide adequate robustness to redistribute the moments.
Although members in routine steel I-girder bridges covered by this Guide may satisfy these
restrictions, the use of moment redistribution should only be undertaken with the full knowledge
and consent of the Owner, and only with a full understanding of the implications. While use of this
method can potentially result in a more economical design in terms of smaller/lighter girder/beam
sections, the necessary analysis is somewhat more time-consuming than that for a design in which

NSBA Guide to Navigating Routine Steel Bridge Design / 112


the section is designed to remain elastic since commercial software packages do not currently
include the capability for automating the associated calculations.
For further information on the provisions of Appendix B6, consult Section 6.5.6.6 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures.

4.6.4.2 Refined Method


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The use of moment redistribution methods has been specifically excluded from the scope of this
Guide for the design of routine steel I-girder bridges. See the Discussion of Article 4.6.4.1 for
more information.

4.6.4.3 Approximate Procedure


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The use of moment redistribution methods has been specifically excluded from the scope of this
Guide for the design of routine steel I-girder bridges. See the Discussion of Article 4.6.4.1 for
more information.

4.6.5 Stability
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The investigation of stability utilizing large deflection theory is not applicable to routine steel I-
girder bridge superstructures but may be applicable for the design of substructures (specifically
tall piers).
The design of most steel girder bridges, including routine steel I-girder bridges, is typically based
on small-deflection theory. Small-deflection theory is the basis for methods of analysis where the
effects of deformation upon force effects in the structure are neglected. In small-deflection theory
analyses, second-order geometric nonlinear behavior is not considered. Instead it is assumed that
the deformations of the structure are small enough that they do not lead to second-order
amplification of member loads. This is a perfectly rational and reasonable assumption for the
design of routine steel I-girder bridges.
There are provisions in the specifications, specifically in Article 6.10.1.6, where second-order
compression-flange lateral bending stresses are approximated using a simple formula to amplify
the first-order values. However, these provisions are specifically intended only for amplification
of compression-flange lateral bending stresses due to torsion, such as those that occur due to the

NSBA Guide to Navigating Routine Steel Bridge Design / 113


effect of deck overhang loads acting on exterior (fascia) girders (see the Discussions of Article
6.10.1.6 and the associated sub-Articles of Article 6.10.3 in this Guide).
Cases where it may be appropriate to base the analysis of a steel girder bridge on large deflection
theory include structure types which have been specifically excluded from the definition of routine
steel I-girder bridges. Examples include, but are not limited to, narrow, slender steel I-girder
superstructures with three or fewer girders during the deck placement (e.g., in a phased
construction situation), which may experience global amplification of lateral-torsional
deformations and potentially be subject to global lateral-torsional buckling. Again, the definition
of a routine steel I-girder bridge for the purposes of this Guide includes limitations which preclude
concern about this type of behavior (in this case, the specific requirement that the cross-section
include four or more girders with no consideration of phased construction).

4.6.6 Analysis for Temperature Gradient


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Consideration of temperature gradient effects is unnecessary for the design of routine steel I-girder
bridges.
Typically, the main reason for performing a temperature gradient analysis in a steel girder bridge
is to evaluate the potential for uplift at the bearings. This is generally more of a concern in narrow
curved and/or skewed steel girder bridges. It is highly unlikely that a temperature gradient loading
would produce uplift in a bridge meeting the definition of a routine steel I-girder bridge.
It is also highly unlikely that temperature gradient loading would contribute to a controlling load
case in terms of stresses or forces in a routine steel I-girder bridge. Consider the recommended
load factors for temperature gradient presented in the AASHTO LRFD BDS: 0.0 for the strength
and extreme event limit states, 1.0 at the service limit state when live load is not present, and 0.50
at the service limit state when live load is considered. It is difficult to imagine a situation where
the effects of temperature gradient on member forces or stresses would contribute to a controlling
load combination when such load factors are used.
Most guideline documents either explicitly or implicitly recommend neglecting analysis of the
temperature gradient effect for the design of bridges which meet the definition of a routine steel I-
girder bridge. For example, the FHWA Steel Bridge Design Handbook, Design Example 1: Three-
Span Continuous Straight Composite Steel I-Girder Bridge, explicitly neglects consideration of
the effects of temperature gradient loading. The Reference Manual for NHI Course 130081, Load
and Resistance Factor Design (LRFD) for Highway Bridge Superstructures discusses the
temperature gradient (TG) load case in general terms in Chapter 3, Loads and Load Factors,
mentions it several times in Chapter 5, Concrete Girder Superstructures, but does not mention this
load case in Chapter 6, Steel Girder Superstructures. The Reference Manual for NHI Course
130095, Analysis and Design of Skewed and Curved Steel Bridges with LRFD, mentions
temperature gradient specifically as a loading which can potentially contribute to uplift in narrow
curved and/or skewed steel girder bridges and which should be considered in determining bearing
reactions.

NSBA Guide to Navigating Routine Steel Bridge Design / 114


Finally, it should also be noted that the effects of temperature gradient cannot be captured in a
typical line girder analysis, and that line girder analysis is widely acknowledged as being perfectly
appropriate for the design of routine steel I-girder bridges.

4.7 DYNAMIC ANALYSIS

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
The provisions of Article 4.7 and its associated sub-Articles are intended to address a wide variety
of dynamic analyses which may be applicable to the design of bridges. For the purposes of this
Guide, it is helpful to group these potential applications for discussion.
Dynamic analysis may be appropriate for evaluation of vehicle- and/or wind-induced vibrations of
certain types of bridge superstructures. However, this type of analysis is typically only necessary
for extremely flexible bridges. Bridges which meet the description of a routine steel I-girder
bridge, as defined for the purposes of this Guide, generally exhibit sufficient stiffness to avoid
harmful dynamic response to vehicle or wind loading. Consequently, dynamic analysis of routine
steel I-girder bridges is not necessary to investigate vehicle- or wind-induced vibrations.
Dynamic analysis is also appropriate for the evaluation of the response of bridges to various lateral
loading cases, such as wind loading, vessel collision, blast forces, or seismic loading. Vessel
collision and blast loading are considered beyond the scope of this Guide; designers faced with the
need to evaluate these types of loading conditions are encouraged to consult with experienced
senior bridge engineers to define the scope and approach for such analyses. Wind loading typically
does not induce harmful dynamic response in superstructures proportioned and configured in a
manner that would meet the description of a routine steel I-girder bridge as defined by this Guide;
typically, wind loading is more of a concern for particularly tall or long span bridges. Seismic
loading can be a concern for virtually any bridge, depending on the nature and magnitude of that
loading. However, the routine steel I-girder bridges covered by this Guide are assumed, by
definition, to be in Seismic Zone 1. Dynamic analysis for seismic loads is only applicable for
bridges in Seismic Zones, 2, 3, or 4; therefore, detailed discussion of dynamic analysis for seismic
loading is beyond the scope of this Guide. Overall, dynamic analysis of routine steel I-girder
bridges is not necessary to evaluate the effects of lateral loading.

4.7.4 Analysis for Earthquake Loads

4.7.4.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article specifies minimum analysis requirement for seismic effects, based on the seismic zone
in which the bridge resides. Only bridges in Seismic Zone 1 meet the definition of a routine steel
I-girder bridge for the purposes of this Guide, so only the provisions related to bridges in Seismic

NSBA Guide to Navigating Routine Steel Bridge Design / 115


Zone 1 apply. Note that this Article states that bridges in Seismic Zone 1 are subject to the
provisions of Articles 4.7.4.4 and 3.10.9.

4.7.4.2 Single-Span Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article states that seismic analysis is not required for single-span bridges, regardless of
seismic zone. The Article also references Article 3.10.9 for minimum forces requirements for the
design of connections between the bridge superstructure and abutments. While these requirements
apply to the routine steel I-girder bridges covered by this Guide, the design of bearings and
substructures is considered beyond the scope of superstructure design.

4.7.4.3 Multispan Bridges


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides the minimum seismic analysis requirements for multispan bridges. Only
bridges in Seismic Zone 1 meet the definition of a routine steel I-girder bridge for the purposes of
this Guide, and the provisions of this Article indicate that no seismic analysis is required for
bridges in Seismic Zone 1. As a result the associated sub-Articles under Article 4.7.3 are not
discussed in this Guide since none of them are applicable.
Note that Article 4.7.4.1 states that bridges in Seismic Zone 1 are subject to the provisions of
Articles 4.7.4.4 and 3.10.9 (see the Discussion of Article 4.7.4.1 in this Guide).

4.7.4.4 Minimum Support Length Requirements


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article specifies the minimum support length requirements for bridges. Only bridges in
Seismic Zone 1 meet the definition of a routine steel I-girder bridge for the purposes of this Guide.
For bridges in Seismic Zone 1, this Article states that the requirements of Article 4.7.4.3 do not
apply, and only the provisions of this Article (4.7.4.4) need to be considered.
This Article provides a simple formula for the nominal empirical minimum support length, N,
measured normal to the centerline of bearing at expansion bearings without restrainers, STUs, or
dampers. The Article also includes Table 4.7.4.4-1, which specifies the percentage of the minimum
support length, N, which must actually be provided, as a function of the Acceleration Coefficient,
As (see the Discussion of Article 3.10.4.2 in this Guide); otherwise, longitudinal restrainers must
be provided in accordance with Article 3.10.9.5.

NSBA Guide to Navigating Routine Steel Bridge Design / 116


4.7.4.5 P- Requirements
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article discusses the design of columns and piers to limit seismic displacements so that P-Δ
effects will not significantly affect the response of the bridge during an earthquake, which is not
applicable to bridges in Seismic Zone 1 and is also beyond the scope of superstructure design.

NSBA Guide to Navigating Routine Steel Bridge Design / 117


SECTION 6: STEEL STRUCTURES

TABLE OF CONTENTS
6.1 SCOPE........................................................................................................................... 130
6.2 DEFINITIONS .............................................................................................................. 130
6.3 NOTATION .................................................................................................................. 130
6.4 MATERIALS ................................................................................................................ 131
6.4.1 Structural Steels .................................................................................................. 131
6.4.2 Pins, Rollers, and Rockers .................................................................................. 131
6.4.3 Bolts, Nuts, and Washers .................................................................................... 132
6.4.3.1 High-Strength Structural Fasteners ............................................................... 132
6.4.3.1.1 High Strength Bolts .................................................................................. 132
6.4.3.1.2 Nuts Used with ASTM F3125 Bolts ........................................................ 132
6.4.3.1.3 Washers Used with ASTM F3125 Bolts .................................................. 133
6.4.3.1.4 Direct Tension Indicators ......................................................................... 133
6.4.3.2 Low-Strength Steel Bolts .............................................................................. 133
6.4.3.3 Fasteners for Structural Anchorage .............................................................. 134
6.4.3.3.1 Anchor Rods ............................................................................................. 134
6.4.3.3.2 Nuts Used with Anchor Rods ................................................................... 134
6.4.4 Stud Shear Connectors ........................................................................................ 134
6.4.5 Weld Metal.......................................................................................................... 134
6.4.6 Cast Metal ........................................................................................................... 135
6.4.6.1 Cast Steel and Ductile Iron ........................................................................... 135
6.4.6.2 Malleable Castings ........................................................................................ 135
6.4.6.3 Cast Iron ........................................................................................................ 135
6.4.7 Stainless Steel ..................................................................................................... 135
6.4.8 Cables .................................................................................................................. 136
6.4.8.1 Bright Wire ................................................................................................... 136
6.4.8.2 Galvanized Wire ........................................................................................... 136
6.4.8.3 Epoxy-Coated Wire ...................................................................................... 136
6.4.8.4 Bridge Strand ................................................................................................ 136
6.4.9 Dissimilar Metals ................................................................................................ 136
6.5 LIMIT STATES ............................................................................................................ 137
6.5.1 General ................................................................................................................ 137
6.5.2 Service Limit State .............................................................................................. 137
6.5.3 Fatigue and Fracture Limit State......................................................................... 138
6.5.4 Strength Limit State ............................................................................................ 138
6.5.4.1 General .......................................................................................................... 138
6.5.4.2 Resistance Factors ......................................................................................... 139
6.5.5 Extreme Event Limit State .................................................................................. 139
6.6 FATIGUE AND FRACTURE CONSIDERATIONS ................................................... 139
6.6.1 Fatigue................................................................................................................. 139
6.6.1.1 General .......................................................................................................... 139
6.6.1.2 Load-Induced Fatigue ................................................................................... 140
6.6.1.2.1 Application ............................................................................................... 140

NSBA Guide to Navigating Routine Steel Bridge Design / 118


6.6.1.2.2 Design Criteria ......................................................................................... 141
6.6.1.2.3 Detail Categories ...................................................................................... 142
6.6.1.2.4 Detailing to Reduce Constraint ................................................................ 144
6.6.1.2.5 Fatigue Resistance .................................................................................... 144
6.6.1.3 Distortion-Induced Fatigue ........................................................................... 146
6.6.1.3.1 Transverse Connection Plates .................................................................. 146
6.6.1.3.2 Lateral Connection Plates ......................................................................... 147
6.6.1.3.3 Orthotropic Decks .................................................................................... 147
6.6.2 Fracture ............................................................................................................... 147
6.6.2.1 Member or Component Designations and Charpy V-Notch Testing
Requirements .................................................................................................................... 147
6.6.2.2 Fracture-Critical Members ............................................................................ 148
6.7 GENERAL DIMENSION AND DETAIL REQUIREMENTS .................................... 148
6.7.1 Effective Length of Span .................................................................................... 148
6.7.2 Dead Load Camber and Detailing of Structural Components ............................ 148
6.7.3 Minimum Thickness of Steel .............................................................................. 150
6.7.4 Diaphragms and Cross-Frames ........................................................................... 150
6.7.4.1 General .......................................................................................................... 150
6.7.4.2 I-Section Members ........................................................................................ 152
6.7.4.3 Composite Box-Section Members ................................................................ 154
6.7.4.4 Noncomposite Box-Section Members .......................................................... 154
6.7.4.4.1 General ..................................................................................................... 154
6.7.4.4.2 Square and Rectangular HSS Members ................................................... 154
6.7.4.4.3 Welded and Nonwelded Built-Up Noncomposite Box-Section Members154
6.7.4.5 Trusses and Arches ....................................................................................... 155
6.7.5 Lateral Bracing.................................................................................................... 155
6.7.5.1 General .......................................................................................................... 155
6.7.5.2 I-Section Members ........................................................................................ 155
6.7.5.3 Tub Section Members ................................................................................... 155
6.7.5.4 Trusses .......................................................................................................... 156
6.7.6 Pins...................................................................................................................... 156
6.7.6.1 Location ........................................................................................................ 156
6.7.6.2 Resistance ..................................................................................................... 156
6.7.6.2.1 Combined Flexure and Shear ................................................................... 156
6.7.6.2.2 Bearing ..................................................................................................... 157
6.7.6.3 Minimize Size Pin for Eyebars ..................................................................... 157
6.7.6.4 Pins and Pin Nuts .......................................................................................... 157
6.7.7 Heat-Curved Rolled Beams and Welded Plate Girders ...................................... 158
6.7.7.1 Scope ............................................................................................................. 158
6.7.7.2 Geometric Limitations .................................................................................. 158
6.7.8 Bent Plates .......................................................................................................... 158
6.8 TENSION MEMBERS ................................................................................................. 159
6.8.1 General ................................................................................................................ 159
6.8.2 Tensile Resistance ............................................................................................... 160
6.8.2.1 General .......................................................................................................... 160
6.8.2.2 Reduction Factor, U ...................................................................................... 161

NSBA Guide to Navigating Routine Steel Bridge Design / 119


6.8.2.3 Combined Axial Tension, Flexure, and Flexural and/or Torsional Shear .... 163
6.8.2.3.1 General ..................................................................................................... 163
6.8.2.3.2 Interaction with Torsional and/or Flexural Shear ..................................... 164
6.8.2.3.3 Tension Rupture Under Axial Tension or Compression Combined with
Flexure ...................................................................................................... 164
6.8.3 Net Area .............................................................................................................. 165
6.8.4 Limiting Slenderness Ratio for Tension Members ............................................. 166
6.8.5 Built-Up Members .............................................................................................. 167
6.8.5.1 General .......................................................................................................... 167
6.8.5.2 Perforated Plates ........................................................................................... 168
6.8.6 Eyebars ................................................................................................................ 168
6.8.6.1 Factored Resistance ...................................................................................... 168
6.8.6.2 Proportions .................................................................................................... 168
6.8.6.3 Packing .......................................................................................................... 168
6.8.7 Pin-Connected Plates .......................................................................................... 169
6.8.7.1 General .......................................................................................................... 169
6.8.7.2 Pin Plates....................................................................................................... 169
6.8.7.3 Proportions .................................................................................................... 169
6.8.7.4 Packing .......................................................................................................... 169
6.9 COMPRESSION MEMBERS ...................................................................................... 170
6.9.1 General ................................................................................................................ 170
6.9.2 Compressive Resistance...................................................................................... 170
6.9.2.1 Axial Compression........................................................................................ 170
6.9.2.2 Combined Axial Compression, Flexure, and Flexural and/or Torsional Shear ..
....................................................................................................................... 171
6.9.2.2.1 General ..................................................................................................... 171
6.9.2.2.2 Interaction with Torsional and/or Flexural Shear ..................................... 173
6.9.3 Limiting Slenderness Ratio for Compression Members ..................................... 174
6.9.4 Noncomposite Members ..................................................................................... 175
6.9.4.1 Nominal Compressive Resistance................................................................. 175
6.9.4.1.1 General ..................................................................................................... 175
6.9.4.1.2 Elastic Flexural Buckling Resistance ....................................................... 177
6.9.4.1.3 Elastic Torsional Buckling and Flexural-Torsional Buckling Resistance 178
6.9.4.2 Effects of Local Buckling on the Nominal Compressive Resistance ........... 179
6.9.4.2.1 Classification of Cross-Section Elements ................................................ 179
6.9.4.2.2 Slender Longitudinally Unstiffened Cross-Section Elements .................. 180
6.9.4.3 Built-Up Members ........................................................................................ 182
6.9.4.3.1 General ..................................................................................................... 182
6.9.4.3.2 Perforated Plates ....................................................................................... 182
6.9.4.4 Single-Angle Members ................................................................................. 182
6.9.4.5 Plate Buckling Under Service and Construction Loads ................................ 184
6.9.5 Composite Members ........................................................................................... 184
6.9.5.1 Nominal Compressive Resistance................................................................. 184
6.9.5.2 Limitations .................................................................................................... 185
6.9.5.2.1 General ..................................................................................................... 185
6.9.5.2.2 Concrete-Filled Tubes .............................................................................. 185

NSBA Guide to Navigating Routine Steel Bridge Design / 120


6.9.5.2.3 Concrete-Encased Shapes......................................................................... 185
6.9.6 Composite Concrete-Filled Steel Tubes (CFSTs)............................................... 186
6.9.6.1 General .......................................................................................................... 186
6.9.6.2 Limitations .................................................................................................... 186
6.9.6.3 Combined Axial Compression and Flexure .................................................. 186
6.9.6.3.1 General ..................................................................................................... 186
6.9.6.3.2 Axial Compressive Resistance ................................................................. 187
6.9.6.3.3 Nominal Flexural Composite Resistance ................................................. 187
6.9.6.3.4 Nominal Stability-Based Interaction Curve ............................................. 187
6.10 I-SECTION FLEXURAL MEMBERS ......................................................................... 188
6.10.1 General ................................................................................................................ 188
6.10.1.1 Composite Sections....................................................................................... 189
6.10.1.1.1 Stresses ..................................................................................................... 189
6.10.1.2 Noncomposite Sections ................................................................................. 195
6.10.1.3 Hybrid Sections ............................................................................................. 196
6.10.1.4 Variable Web Depth Members ..................................................................... 196
6.10.1.5 Stiffness......................................................................................................... 197
6.10.1.6 Flange Stresses and Member Bending Moments .......................................... 198
6.10.1.7 Minimum Negative Flexure Concrete Deck Reinforcement ........................ 200
6.10.1.8 Tension Flanges with Holes .......................................................................... 202
6.10.1.9 Web Bend-Buckling Resistance ................................................................... 203
6.10.1.9.1 Webs without Longitudinal Stiffeners ..................................................... 203
6.10.1.9.2 Webs with Longitudinal Stiffeners........................................................... 205
6.10.1.10 Flange-Strength Reduction Factors............................................................... 205
6.10.1.10.1 Hybrid Factor, Rh ..................................................................................... 205
6.10.1.10.2 Web Load-Shedding Factor, Rb................................................................ 205
6.10.2 Cross-Section Proportion Limits......................................................................... 207
6.10.2.1 Web Proportions ........................................................................................... 207
6.10.2.1.1 Webs without Longitudinal Stiffeners ..................................................... 207
6.10.2.1.2 Webs with Longitudinal Stiffeners........................................................... 208
6.10.2.2 Flange Proportions ........................................................................................ 208
6.10.3 Constructibility ................................................................................................... 210
6.10.3.1 General .......................................................................................................... 210
6.10.3.2 Flexure .......................................................................................................... 213
6.10.3.2.1 Discretely Braced Flanges in Compression.............................................. 213
6.10.3.2.2 Discretely Braced Flanges in Tension ...................................................... 214
6.10.3.2.3 Continuously Braced Flanges in Tension or Compression ...................... 215
6.10.3.2.4 Concrete Deck .......................................................................................... 215
6.10.3.3 Shear ............................................................................................................. 216
6.10.3.4 Deck Placement ............................................................................................ 217
6.10.3.4.1 General ..................................................................................................... 217
6.10.3.4.2 Global Displacement Amplification in Narrow I-Girder Bridge Units .... 220
6.10.3.5 Dead Load Deflections ................................................................................. 221
6.10.4 Service Limit State .............................................................................................. 221
6.10.4.1 Elastic Deformations ..................................................................................... 221
6.10.4.2 Permanent Deformations .............................................................................. 221

NSBA Guide to Navigating Routine Steel Bridge Design / 121


6.10.4.2.1 General ..................................................................................................... 221
6.10.4.2.2 Flexure ...................................................................................................... 222
6.10.5 Fatigue and Fracture Limit State......................................................................... 224
6.10.5.1 Fatigue........................................................................................................... 224
6.10.5.2 Fracture ......................................................................................................... 225
6.10.5.3 Special Fatigue Requirement for Webs ........................................................ 225
6.10.6 Strength Limit State ............................................................................................ 226
6.10.6.1 General .......................................................................................................... 226
6.10.6.2 Flexure .......................................................................................................... 226
6.10.6.2.1 General ..................................................................................................... 226
6.10.6.2.2 Composite Sections in Positive Flexure ................................................... 227
6.10.6.2.3 Composite Sections in Negative Flexure and Noncomposite Sections .... 228
6.10.6.3 Shear ............................................................................................................. 231
6.10.6.4 Shear Connectors .......................................................................................... 231
6.10.7 Flexural Resistance—Composite Sections in Positive Flexure .......................... 232
6.10.7.1 Compact Sections.......................................................................................... 232
6.10.7.1.1 General ..................................................................................................... 232
6.10.7.1.2 Nominal Flexural Resistance.................................................................... 233
6.10.7.2 Noncompact Sections.................................................................................... 234
6.10.7.2.1 General ..................................................................................................... 234
6.10.7.2.2 Nominal Flexural Resistance.................................................................... 235
6.10.7.3 Ductility Requirement ................................................................................... 236
6.10.8 Flexural Resistance—Composite Sections in Negative Flexure and Noncomposite
Sections ............................................................................................................................. 237
6.10.8.1 General .......................................................................................................... 237
6.10.8.1.1 Discretely Braced Flanges in Compression.............................................. 237
6.10.8.1.2 Discretely Braced Flanges in Tension ...................................................... 239
6.10.8.1.3 Continuously Braced Flanges in Tension or Compression ...................... 240
6.10.8.2 Compression-Flange Flexural Resistance ..................................................... 242
6.10.8.2.1 General ..................................................................................................... 242
6.10.8.2.2 Local Buckling Resistance ....................................................................... 243
6.10.8.2.3 Lateral Torsional Buckling Resistance..................................................... 245
6.10.8.3 Flexural Resistance Based on Tension Flange Yielding............................... 247
6.10.9 Shear Resistance ................................................................................................. 248
6.10.9.1 General .......................................................................................................... 248
6.10.9.2 Nominal Resistance of Unstiffened Webs .................................................... 250
6.10.9.3 Nominal Resistance of Stiffened Webs ........................................................ 251
6.10.9.3.1 General ..................................................................................................... 251
6.10.9.3.2 Interior Panels........................................................................................... 253
6.10.9.3.3 End Panels ................................................................................................ 254
6.10.10 Shear Connectors ................................................................................................ 255
6.10.10.1 General .......................................................................................................... 255
6.10.10.1.1 Types ........................................................................................................ 255
6.10.10.1.2 Pitch .......................................................................................................... 256
6.10.10.1.3 Transverse Spacing................................................................................... 257
6.10.10.1.4 Cover and Penetration .............................................................................. 257

NSBA Guide to Navigating Routine Steel Bridge Design / 122


6.10.10.2 Fatigue Resistance ........................................................................................ 258
6.10.10.3 Special Requirements for Points of Permanent Load Contraflexure ............ 259
6.10.10.4 Strength Limit State ...................................................................................... 259
6.10.10.4.1 General ..................................................................................................... 259
6.10.10.4.2 Nominal Shear Force ................................................................................ 260
6.10.10.4.3 Nominal Shear Resistance ........................................................................ 261
6.10.11 Web Stiffeners .................................................................................................... 262
6.10.11.1 Web Transverse Stiffeners ............................................................................ 262
6.10.11.1.1 General ..................................................................................................... 262
6.10.11.1.2 Projecting Width....................................................................................... 263
6.10.11.1.3 Moment of Inertia ..................................................................................... 263
6.10.11.2 Bearing Stiffeners ......................................................................................... 264
6.10.11.2.1 General ..................................................................................................... 264
6.10.11.2.2 Minimum Thickness ................................................................................. 265
6.10.11.2.3 Bearing Resistance ................................................................................... 266
6.10.11.2.4 Axial Resistance of Bearing Stiffeners .................................................... 267
6.10.11.3 Web Longitudinal Stiffeners ......................................................................... 269
6.10.11.3.1 General ..................................................................................................... 269
6.10.11.3.2 Projecting Width....................................................................................... 269
6.10.11.3.3 Moment of Inertia and Radius of Gyration .............................................. 269
6.10.12 Cover Plates ........................................................................................................ 269
6.10.12.1 General .......................................................................................................... 269
6.10.12.2 End Requirements ......................................................................................... 269
6.10.12.2.1 General ..................................................................................................... 269
6.10.12.2.2 Welded Ends............................................................................................. 270
6.10.12.2.3 Bolted Ends .............................................................................................. 270
6.11 COMPOSITE BOX-SECTION FLEXURAL MEMBERS .......................................... 270
6.11.1 General ................................................................................................................ 270
6.11.1.1 Stress Determinations ................................................................................... 271
6.11.1.2 Bearings ........................................................................................................ 271
6.11.1.3 Flange-to-Web Connections ......................................................................... 271
6.11.1.4 Access and Drainage ..................................................................................... 272
6.11.2 Cross-Section Proportion Limits......................................................................... 272
6.11.2.1 Web Proportions ........................................................................................... 272
6.11.2.1.1 General ..................................................................................................... 272
6.11.2.1.2 Webs without Longitudinal Stiffeners ..................................................... 272
6.11.2.1.3 Webs with Longitudinal Stiffeners........................................................... 273
6.11.2.2 Flange Proportions ........................................................................................ 273
6.11.2.3 Special Restrictions on Use of Live Load Distribution Factor for Multiple Box
Sections ....................................................................................................................... 273
6.11.3 Constructibility ................................................................................................... 273
6.11.3.1 General .......................................................................................................... 273
6.11.3.2 Flexure .......................................................................................................... 274
6.11.3.3 Shear ............................................................................................................. 274
6.11.4 Service Limit State .............................................................................................. 275
6.11.5 Fatigue and Fracture Limit State......................................................................... 275

NSBA Guide to Navigating Routine Steel Bridge Design / 123


6.11.6 Strength Limit State ............................................................................................ 275
6.11.6.1 General .......................................................................................................... 275
6.11.6.2 Flexure .......................................................................................................... 276
6.11.6.2.1 General ..................................................................................................... 276
6.11.6.2.2 Sections in Positive Flexure ..................................................................... 276
6.11.6.2.3 Sections in Negative Flexure .................................................................... 276
6.11.6.3 Shear ............................................................................................................. 276
6.11.6.4 Shear Connectors .......................................................................................... 277
6.11.7 Flexural Resistance—Sections in Positive Flexure ............................................ 277
6.11.7.1 Compact Sections.......................................................................................... 277
6.11.7.1.1 General ..................................................................................................... 277
6.11.7.1.2 Nominal Flexural Resistance.................................................................... 278
6.11.7.2 Noncompact Sections.................................................................................... 278
6.11.7.2.1 General ..................................................................................................... 278
6.11.7.2.2 Nominal Flexural Resistance.................................................................... 279
6.11.8 Flexural Resistance—Sections in Negative Flexure ........................................... 279
6.11.8.1 General .......................................................................................................... 279
6.11.8.1.1 Box Flanges in Compression .................................................................... 279
6.11.8.1.2 Continuously Braced Flanges in Tension ................................................. 280
6.11.8.2 Flexural Resistance of Box Flanges in Compression ................................... 280
6.11.8.2.1 General ..................................................................................................... 280
6.11.8.2.2 Unstiffened Flanges .................................................................................. 280
6.11.8.2.3 Longitudinally Stiffened Flanges ............................................................. 281
6.11.8.3 Flexural Resistance Based on Tension Flange Yielding............................... 281
6.11.9 Shear Resistance ................................................................................................. 282
6.11.10 Shear Connectors ................................................................................................ 282
6.11.11 Stiffeners ............................................................................................................. 283
6.11.11.1 Web Stiffeners .............................................................................................. 283
6.11.11.2 Longitudinal Compression-Flange Stiffeners ............................................... 283
6.12 MISCELLANEOUS FLEXURAL MEMBERS ........................................................... 284
6.12.1 General ................................................................................................................ 284
6.12.1.1 Scope ............................................................................................................. 284
6.12.1.2 Strength Limit State ...................................................................................... 285
6.12.1.2.1 Flexure ...................................................................................................... 285
6.12.1.2.2 Combined Flexure, Axial Load, and Flexural and/or Torsional Shear .... 285
6.12.1.2.3 Flexural Shear and/or Torsion .................................................................. 287
6.12.1.2.4 Special Provisions for HSS Members ...................................................... 288
6.12.2 Nominal Flexural Resistance .............................................................................. 288
6.12.2.1 General .......................................................................................................... 288
6.12.2.2 Noncomposite Members ............................................................................... 288
6.12.2.2.1 I- and H-Shaped Members ....................................................................... 288
6.12.2.2.2 Rectangular Box-Section Members.......................................................... 289
6.12.2.2.3 Circular Tubes and Round HSS ............................................................... 292
6.12.2.2.4 Tees and Double Angles........................................................................... 292
6.12.2.2.5 Channels ................................................................................................... 295
6.12.2.2.6 Single Angles ........................................................................................... 295

NSBA Guide to Navigating Routine Steel Bridge Design / 124


6.12.2.2.7 Rectangular Bars and Solid Rounds ......................................................... 296
6.12.2.3 Composite Members ..................................................................................... 296
6.12.2.3.1 Concrete-Encased Shapes......................................................................... 296
6.12.2.3.2 Concrete-Filled Tubes .............................................................................. 296
6.12.2.3.3 Composite Concrete-Filled Steel Tubes (CFSTs) .................................... 297
6.12.3 Nominal Shear Resistance of Composite Members............................................ 297
6.12.3.1 Concrete-Encased Shapes ............................................................................. 297
6.12.3.2 Concrete-Filled Tubes ................................................................................... 298
6.12.3.2.1 Rectangular Tubes .................................................................................... 298
6.12.3.2.2 Composite Concrete Filled Tubes ............................................................ 298
6.13 CONNECTIONS AND SPLICES................................................................................. 298
6.13.1 General ................................................................................................................ 298
6.13.2 Bolted Connections ............................................................................................. 302
6.13.2.1 General .......................................................................................................... 302
6.13.2.1.1 Slip-Critical Connections ......................................................................... 303
6.13.2.1.2 Bearing-Type Connections ....................................................................... 304
6.13.2.2 Factored Resistance ...................................................................................... 305
6.13.2.3 Bolts, Nuts, and Washers .............................................................................. 306
6.13.2.3.1 Bolts and Nuts .......................................................................................... 306
6.13.2.3.2 Washers .................................................................................................... 306
6.13.2.4 Holes ............................................................................................................. 307
6.13.2.4.1 Type .......................................................................................................... 307
6.13.2.4.2 Size ........................................................................................................... 308
6.13.2.5 Size of Bolts .................................................................................................. 309
6.13.2.6 Spacing of Bolts ............................................................................................ 309
6.13.2.6.1 Minimum Spacing and Clear Distance ..................................................... 309
6.13.2.6.2 Maximum Spacing for Sealing Bolts ....................................................... 309
6.13.2.6.3 Maximum Pitch for Stitch Bolts ............................................................... 310
6.13.2.6.4 Maximum Pitch for Stitch Bolts at the End of Compression Members ... 310
6.13.2.6.5 End Distance............................................................................................. 311
6.13.2.6.6 Edge Distances ......................................................................................... 311
6.13.2.7 Shear Resistance ........................................................................................... 312
6.13.2.8 Slip Resistance .............................................................................................. 314
6.13.2.9 Bearing Resistance at Bolt Holes .................................................................. 315
6.13.2.10 Tensile Resistance ......................................................................................... 317
6.13.2.10.1 General ..................................................................................................... 317
6.13.2.10.2 Nominal Tensile Resistance ..................................................................... 317
6.13.2.10.3 Fatigue Resistance .................................................................................... 318
6.13.2.10.4 Prying Action ........................................................................................... 318
6.13.2.11 Combined Tension and Shear ....................................................................... 319
6.13.2.12 Shear Resistance of Anchor Rods ................................................................. 320
6.13.3 Welded Connections ........................................................................................... 320
6.13.3.1 General .......................................................................................................... 320
6.13.3.2 Factored Resistance ...................................................................................... 322
6.13.3.2.1 General ..................................................................................................... 322
6.13.3.2.2 Complete Penetration Groove-Welded Connections ............................... 323

NSBA Guide to Navigating Routine Steel Bridge Design / 125


6.13.3.2.3 Partial Penetration Groove-Welded Connections..................................... 325
6.13.3.2.4 Fillet-Welded Connections ....................................................................... 327
6.13.3.3 Effective Area ............................................................................................... 329
6.13.3.4 Size of Fillet Welds ....................................................................................... 329
6.13.3.5 Minimum Effective Length of Fillet Welds.................................................. 331
6.13.3.6 Fillet Weld End Returns................................................................................ 331
6.13.3.7 Fillet Welds for Sealing ................................................................................ 332
6.13.4 Block Shear Rupture Resistance ......................................................................... 333
6.13.5 Connection Elements .......................................................................................... 334
6.13.5.1 General .......................................................................................................... 334
6.13.5.2 Tension .......................................................................................................... 334
6.13.5.3 Shear ............................................................................................................. 336
6.13.6 Splices ................................................................................................................. 336
6.13.6.1 Bolted Splices ............................................................................................... 336
6.13.6.1.1 Tension Members ..................................................................................... 336
6.13.6.1.2 Compression Members ............................................................................. 337
6.13.6.1.3 Flexural Members..................................................................................... 338
6.13.6.1.4 Fillers ........................................................................................................ 345
6.13.6.2 Welded Splices.............................................................................................. 347
6.13.7 Rigid Frame Connections ................................................................................... 348
6.13.7.1 General .......................................................................................................... 348
6.13.7.2 Webs ............................................................................................................. 348
6.14 PROVISIONS FOR STRUCTURE TYPES ................................................................. 349
6.14.1 Through-Girder Spans ........................................................................................ 349
6.14.2 Trusses ................................................................................................................ 349
6.14.2.1 General .......................................................................................................... 349
6.14.2.2 Truss Members.............................................................................................. 349
6.14.2.3 Secondary Stresses ........................................................................................ 350
6.14.2.4 Diaphragms ................................................................................................... 350
6.14.2.5 Camber .......................................................................................................... 350
6.14.2.6 Working Lines and Gravity Axes ................................................................. 350
6.14.2.7 Portal and Sway Bracing ............................................................................... 351
6.14.2.7.1 General ..................................................................................................... 351
6.14.2.7.2 Through-Truss Spans ............................................................................... 351
6.14.2.7.3 Deck Truss Spans ..................................................................................... 351
6.14.2.8 Gusset Plates ................................................................................................. 352
6.14.2.8.1 General ..................................................................................................... 352
6.14.2.8.2 Multilayered Gusset and Splice Plates ..................................................... 352
6.14.2.8.3 Shear Resistance ....................................................................................... 352
6.14.2.8.4 Compressive Resistance ........................................................................... 353
6.14.2.8.5 Tensile Resistance .................................................................................... 353
6.14.2.8.6 Chord Splices ........................................................................................... 354
6.14.2.8.7 Edge Slenderness ...................................................................................... 354
6.14.2.9 Half Through-Trusses ................................................................................... 354
6.14.2.10 Factored Resistance ...................................................................................... 355
6.14.3 Orthotropic Deck Superstructures....................................................................... 355

NSBA Guide to Navigating Routine Steel Bridge Design / 126


6.14.3.1 General .......................................................................................................... 355
6.14.3.2 Decks in Global Compression ...................................................................... 355
6.14.3.2.1 General ..................................................................................................... 355
6.14.3.2.2 Local Buckling ......................................................................................... 355
6.14.3.2.3 Panel Buckling ......................................................................................... 355
6.14.3.3 Effective Width of Deck ............................................................................... 356
6.14.3.4 Superposition of Global and Local Effects ................................................... 356
6.14.4 Solid Web Arches ............................................................................................... 356
6.14.4.1 General .......................................................................................................... 356
6.14.4.2 Web Slenderness ........................................................................................... 356
6.14.4.3 Moment Amplification.................................................................................. 357
6.14.4.4 Nominal Compressive Resistance................................................................. 357
6.14.4.5 Nominal Flexural Resistance ........................................................................ 357
6.14.4.6 Combined Axial Compression or Tension with Flexural and Torsion ......... 357
6.15 PILES ............................................................................................................................ 357
6.15.1 General ................................................................................................................ 357
6.15.2 Structural Resistance ........................................................................................... 358
6.15.3 Compressive Resistance...................................................................................... 358
6.15.3.1 Axial Compression........................................................................................ 359
6.15.3.2 Combined Axial Compression and Flexure .................................................. 359
6.15.3.3 Buckling ........................................................................................................ 359
6.15.4 Maximum Permissible Driving Stresses ............................................................. 360
6.16 PROVISIONS FOR SEISMIC DESIGN ...................................................................... 360
6.16.1 General ................................................................................................................ 360
6.16.2 Materials ............................................................................................................. 361
6.16.3 Design Requirements for Seismic Zone 1 .......................................................... 361
6.16.4 Design Requirements for Seismic Zones 2, 3, or 4............................................. 362
6.16.4.1 General .......................................................................................................... 362
6.16.4.2 Deck .............................................................................................................. 362
6.16.4.3 Shear Connectors .......................................................................................... 362
6.16.4.4 Elastic Superstructures .................................................................................. 363
6.17 REFERENCES .............................................................................................................. 363
APPENDIX A6 FLEXURAL RESISTANCE OF COMPOSITE I-SECTIONS IN
NEGATIVE FLEXURE AND NONCOMPOSITE I-SECTIONS WITH COMPACT OR
NONCOMPACT WEBS IN STRAIGHT BRIDGES ................................................... 363
A6.1 GENERAL .................................................................................................................... 363
A6.1.1 Sections with Discretely Braced Compression Flanges...................................... 366
A6.1.2 Sections with Discretely Braced Tension Flanges .............................................. 367
A6.1.3 Sections with Continuously Braced Compression Flanges ................................ 369
A6.1.4 Sections with Continuously Braced Tension Flanges ......................................... 369
A6.2 WEB PLASTIFICATION FACTORS .......................................................................... 370
A6.2.1 Compact Web Sections ....................................................................................... 370
A6.2.2 Noncompact Web Sections ................................................................................. 371
A6.3 FLEXURAL RESISTANCE BASED ON THE COMPRESSION FLANGE ............. 372
A6.3.1 General ................................................................................................................ 372
A6.3.2 Local Buckling Resistance .................................................................................. 374

NSBA Guide to Navigating Routine Steel Bridge Design / 127


A6.3.3 Lateral Torsional Buckling Resistance ............................................................... 375
A6.4 FLEXURAL RESISTANCE BASED ON TENSION FLANGE YIELDING ............. 378
APPENDIX B6 MOMENT REDISTRIBUTION FROM INTERIOR-PIER I-SECTIONS
IN STRAIGHT CONTINUOUS-SPAN BRIDGES ..................................................... 379
B6.1 GENERAL .................................................................................................................... 379
B6.2 SCOPE........................................................................................................................... 380
B6.2.1 Web Proportions ................................................................................................. 380
B6.2.2 Compression Flange Proportions ........................................................................ 380
B6.2.3 Section Transitions.............................................................................................. 381
B6.2.4 Compression Flange Bracing .............................................................................. 381
B6.2.5 Shear ................................................................................................................... 381
B6.2.6 Bearing Stiffeners ............................................................................................... 382
B6.3 SERVICE LIMIT STATE ............................................................................................. 382
B6.3.1 General ................................................................................................................ 382
B6.3.2 Flexure ................................................................................................................ 382
B6.3.2.1 Adjacent to Interior-Pier Sections ................................................................. 382
B6.3.2.2 At All Other Locations.................................................................................. 383
B6.3.3 Redistribution Moments...................................................................................... 383
B6.3.3.1 At Interior-Pier Sections ............................................................................... 383
B6.3.3.2 At All Other Locations.................................................................................. 383
B6.4 STRENGTH LIMIT STATE......................................................................................... 384
B6.4.1 Flexural Resistance ............................................................................................. 384
B6.4.1.1 Adjacent to Interior-Pier Sections ................................................................. 384
B6.4.1.2 At All Other Locations.................................................................................. 384
B6.4.2 Redistribution Moments...................................................................................... 385
B6.4.2.1 At Interior-Pier Sections ............................................................................... 385
B6.4.2.2 At All Other Sections .................................................................................... 385
B6.5 EFFECTIVE PLASTIC MOMENT .............................................................................. 385
B6.5.1 Interior-Pier Sections with Enhanced Moment-Rotation Characteristics ........... 385
B6.5.2 All Other Interior-Pier Sections .......................................................................... 386
B6.6 REFINED METHOD .................................................................................................... 386
B6.6.1 General ................................................................................................................ 386
B6.6.2 Nominal Moment-Rotation Curves..................................................................... 387
APPENDIX C6 BASIC STEPS FOR STEEL BRIDGE SUPERSTRUCTURES ............. 387
C6.1 GENERAL .................................................................................................................... 387
C6.2 GENERAL CONSIDERATIONS ................................................................................. 388
C6.3 SUPERSTRUCTURE DESIGN.................................................................................... 388
C6.4 FLOWCHARTS FOR FLEXURAL DESIGN OF I-SECTION MEMBERS .............. 388
C6.4.1 Flowchart for LRFD Article 6.10.3 .................................................................... 388
C6.4.2 Flowchart for LRFD Article 6.10.4 .................................................................... 388
C6.4.3 Flowchart for LRFD Article 6.10.5 .................................................................... 389
C6.4.4 Flowchart for LRFD Article 6.10.6 .................................................................... 389
C6.4.5 Flowchart for LRFD Article 6.10.7 .................................................................... 389
C6.4.6 Flowchart for LRFD Article 6.10.8 .................................................................... 389
C6.4.7 Flowchart for Appendix A6 ................................................................................ 390
C6.4.8 Flowchart for Article D6.4.1 ............................................................................... 391

NSBA Guide to Navigating Routine Steel Bridge Design / 128


C6.4.9 Flowchart for Article D6.4.2 ............................................................................... 392
C6.4.10 Moment Gradient Modifier, Cb (Sample Cases) ................................................ 393
C6.5 FLOWCHARTS FOR LRFD ARTICLES 6.9.4 AND 6.12.2.2.2 ................................ 394
C6.5.1 Flowchart for LRFD Article 6.9.4 ...................................................................... 394
C6.5.2 Flowchart for LRFD Article 6.12.2.2.2 .............................................................. 394
APPENDIX D6 FUNDAMENTAL CALCULATIONS FOR FLEXURAL MEMBERS . 395
D6.1 PLASTIC MOMENT .................................................................................................... 395
D6.2 YIELD MOMENT ........................................................................................................ 397
D6.2.1 Noncomposite Sections ....................................................................................... 397
D6.2.2 Composite Sections in Positive Flexure ............................................................. 398
D6.2.3 Composite Sections in Negative Flexure ............................................................ 399
D6.2.4 Sections with Cover Plates.................................................................................. 400
D6.3 DEPTH OF THE WEB IN COMPRESSION ............................................................... 401
D6.3.1 In the Elastic Range (Dc) .................................................................................... 401
D6.3.2 At Plastic Moment (Dcp) ..................................................................................... 403
D6.4 LATERAL TORSIONAL BUCKLING EQUATIONS FOR CB > 1.0, WITH
EMPHASIS ON UNBRACED LENGTH REQUIREMENTS FOR DEVELOPMENT
OF THE MAXIMUM FLEXURAL RESISTANCE .................................................... 404
D6.4.1 By the Provisions of Article 6.10.8.2.3 ............................................................... 404
D6.4.2 By the Provisions of Article A6.3.3 .................................................................... 405
D6.5 CONCENTRATED LOADS APPLIED TO WEBS WITHOUT BEARING
STIFFENERS ................................................................................................................ 406
D6.5.1 General ................................................................................................................ 406
D6.5.2 Web Local Yielding ............................................................................................ 407
D6.5.3 Web Crippling ..................................................................................................... 409
APPENDIX E6 NOMINAL COMPRESSIVE RESISTANCE OF NONCOMPOSITE
MEMBERS CONTAINING LONGITUDINALLY STIFFENED PLATES ............... 410
E6.1 NOMINAL COMPRESSIVE RESISTANCE .............................................................. 410
E6.1.1 General ................................................................................................................ 410
E6.1.2 Classification of Longitudinally Stiffened Plate Panels ..................................... 410
E6.1.3 Nominal Compressive Resistance and Effective Area of Plates with Equally-
spaced Equal-size Longitudinal Stiffeners .......................................................................... 410
E6.1.4 Longitudinal Stiffeners ....................................................................................... 411
E6.1.5 Transverse Stiffeners .......................................................................................... 411
E6.1.5.1 General .......................................................................................................... 411
E6.1.5.2 Moment of Inertia ......................................................................................... 411

NSBA Guide to Navigating Routine Steel Bridge Design / 129


6.1 SCOPE

Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.


Discussion:
This Article provides the overall scope of what types of steel components and systems are
addressed by the design provisions given in Section 6. The Article also discusses the limitations
of the design provisions as they relate specifically to horizontally curved-girder systems, which
are not applicable to the routine steel-girder bridges covered by this Guide. The Article also points
to Appendix C6, which provides a helpful outline of the basic steps for the design of steel-girder
bridges; note that only the steps shown in the outline for I-girder bridges are applicable to the
routine steel-girder bridges covered by this Guide (see the Discussions of Articles C6.1 through
C6.3 in this Guide).
The Commentary for this Article provides a discussion of the overall organization of Section 6.
Articles 6.10 and 6.13 are the Articles primarily applicable to routine steel I-girder bridges;
portions of Article 6.8, 6.9, and 6.12 may also be partially or conditionally applicable, or beyond
the scope of superstructure design, as discussed further below. The application of advanced
analysis methods, as discussed in the Commentary, is typically neither necessary nor
recommended for the routine steel I-girder bridges covered by this Guide.

6.2 DEFINITIONS

Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.


Discussion:
This Article provides an alphabetic listing of the specific definitions for commonly used terms
throughout Section 6. Designers should refer to these definitions whenever a term is encountered
in Section 6 to clearly understand the specific meaning that the code writers intended to apply to
that term. Note that many, but not all, of the definitions will apply to routine steel I-girder bridges.

6.3 NOTATION

Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.


Discussion:
This Article provides an alphabetical listing of the definitions for the variables used in the text and
in the equations throughout Section 6. Some of the variables are re-used and have multiple
definitions. The Article number where a specific definition for a given variable first appears in
Section 6 is provided in parentheses at the end of the definition(s). Note that many, but not all, of
the variables and their definitions will apply to the design of routine steel I-girder bridges.

NSBA Guide to Navigating Routine Steel Bridge Design / 130


6.4 MATERIALS

6.4.1 Structural Steels


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides the material specification names and minimum material properties for the
applicable plate and rolled steel used for steel bridges. The material specification for structural
steel for bridges is ASTM A709/A709M, with the “M” referring to the metric version of the
specification, which is not applicable. The AASHTO M270M/M270 material specification is
equivalent to the ASTM A709/A709M specification but is typically several versions behind the
ASTM specification. The specification that is applicable depends on the specific Owner-agency
standards.
Only Grades 36, 50, 50S, and 50W are considered applicable for the routine steel I-girder bridges
covered by this Guide, with Grades 50 and 50W being the standard grades of structural plate steel
specified for the girders in routine plate-girder bridges (with the “W” referring to weathering steel),
and Grades 50S and 50W being the only grades of structural steel that can be specified for the
rolled wide-flange beams in routine rolled-beam bridges (with the “S” referring to shape). There
is virtually no price advantage associated with the use of Grade 36 steels, so it is rarely if ever used
for the design of plate girders or rolled beams in modern designs. “HPS” designates “high
performance steel” grades which have enhanced weldability and toughness; however, the use of
HPS 50W is typically not warranted since there have been few weldability problems reported in
the non-HPS Grade 50 and 50W steels, and the enhanced toughness is generally of more value for
certain fracture-critical members with low redundancy such as tension ties in tied-arch bridges.
Other structural steels listed in this Article are not applicable to routine steel-girder bridges.
Note that rolled wide-flange beams used as diaphragms and structural tees used as cross-frame
members in routine steel-girder bridges should be specified as Grade 50S or 50W; structural tees
are typically split by the fabricator from rolled wide-flange beams. Angles and channels used as
cross-frame or diaphragm members are available as either Grade 36, 50, or 50W and should be
specified accordingly. Detail steel, e.g., steel used to fabricate connection plates and stiffeners, is
typically specified as Grade 50 or 50W but Grade 36 is sometimes used; in some cases, Owner-
agency policy may specify that Grade 36 is permitted to be substituted in place of Grade 50 in
detail steel applications.
Table 6.4.1-1 provides the specified minimum yield strength, Fy, and the specified minimum
tensile strength, Fu, for Grades 36, 50, 50S, and 50W, and the ranges of plate thicknesses available
in Grades 36, 50, and 50W.
For additional information, consult the FHWA’s Steel Bridge Design Handbook, Volume 1 –
Bridge Steels and Their Mechanical Properties.

6.4.2 Pins, Rollers, and Rockers


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 131


Discussion:
Pins, rollers, and rockers are found on older existing bridges, and are not typically used on modern
steel bridges, including the routine steel I-girder bridges covered by this Guide. Therefore, the
provisions in this Article are considered not applicable to such bridges.

6.4.3 Bolts, Nuts, and Washers

6.4.3.1 High-Strength Structural Fasteners

6.4.3.1.1 High Strength Bolts


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article defines the material specification for high-strength bolts, which are used in bolted
field splices and in cross-frame or diaphragm connections in routine steel I-girder bridges. The
specification is the ASTM F3125/3125M specification, with the “M” referring to the metric
version of the specification which is not applicable. Grade A325 bolts are the most common high-
strength bolts under the ASTM F3125 specification but Grade A490 bolts may sometimes be used
to reduce the number of bolts required in the connection if permitted by Owner-agency policy. The
twist-off equivalents of Grade A325 and A490 bolts, i.e., Grade F1852 and F2280 respectively,
may alternatively be used depending on the preferences of the Owner-agency. The ASTM F3125
specification also defines the types of coatings that may be applied to the bolts in non-weathering
steel applications. Type 3 bolts are used with weathering steel. Galvanizing of Grade A490 and
F2280 bolts is not permitted. The rotational capacity testing of the fastener assemblies mentioned
in the Commentary is performed by the bolt manufacturer.
Table 6.4.3.1.1-1 provides the specified minimum tensile strength of the bolts, Fub, which is used
to compute the factored shear resistance of the bolts according to the provisions of Article 6.13.2.7
(see the Discussion of Article 6.13.2.7 in this Guide).
For further information on high-strength bolts, including installation provisions and verification
procedures, consult the AISC Design Guide 17 High Strength Bolts - A Primer for Engineers, the
RCSC Specifications for Structural Joints Using High-Strength Bolts available from the Research
Council on Structural Connections (RCSC), the AASHTO-NSBA Steel Bridge Collaboration
Guide Specification S2.1-2018 Steel Bridge Fabrication Guide Specification, the AASHTO LRFD
Bridge Construction Specifications, and the applicable Owner-agency standards.

6.4.3.1.2 Nuts Used with ASTM F3125 Bolts


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article simply indicates that the nuts recommended or suitable for use with the various grades
of high-strength bolts described above are listed in the ASTM F3125 specification.

NSBA Guide to Navigating Routine Steel Bridge Design / 132


6.4.3.1.3 Washers Used with ASTM F3125 Bolts

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
This Article simply indicates that the hardened washers recommended or suitable for use with the
various grades of high-strength bolts described above are listed in the ASTM F3125 specification.
The need for washers and the selection of them is typically done by the fabricator.

6.4.3.1.4 Direct Tension Indicators

Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.


Discussion:
This Article provides the requirements for direct tension indicators, or so-called “DTIs” which
may be used, depending on the preferences and policies of the Owner-agency, in conjunction with
high-strength bolts, nuts, and washers in order to verify the required bolt installation tension. DTIs
are washers which include mechanical features (typically small arch-shaped protrusions) which
compress in response to the pretension developed in the bolt. When correctly calibrated, the
amount of pretension can be determined by measuring the gap remaining between the washer and
the connected element.
The material specification for DTIs is ASTM A959/A959M, with the “M” referring to the metric
version of the specification which is not applicable. Two alternative DTIs known as captive
DTI/nuts and self-indicating DTIs may be used if permitted by Owner-agency policy. When used,
refer to the reference documents cited in the Discussion of Article 6.4.3.1.1 for further information
on DTIs and their specific installation provisions and verification procedures.

6.4.3.2 Low-Strength Steel Bolts

Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.


Discussion:
This Article provides the requirements for low-strength steel bolts, which may be used to connect
non-structural items in routine steel-girder bridges, if permitted by Owner-agency policy. The
material specification for these bolts is the ASTM A307 specification, which covers two grades of
bolts (Grades A and B). The specified minimum tensile strength of the bolts, Fub, is given as 60
ksi, which is used to compute the factored shear resistance of the bolts according to the provisions
of Article 6.13.2.7 using the equation for threads included in the shear plane (see the Discussion
of Article 6.13.2.7 in this Guide).

NSBA Guide to Navigating Routine Steel Bridge Design / 133


6.4.3.3 Fasteners for Structural Anchorage

6.4.3.3.1 Anchor Rods

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
This Article defines the material specification for anchor rods, i.e., ASTM F1554, which are
typically only used at bearings in most routine steel-girder bridges. Anchor rods are also commonly
referred to as “anchor bolts”; the two terms are considered synonymous.

6.4.3.3.2 Nuts Used with Anchor Rods

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
This Article simply indicates the nuts that must be used with ASTM F1554 anchor rods (also
sometimes called “anchor bolts”), which are typically only used at bearings in most routine steel-
girder bridges.

6.4.4 Stud Shear Connectors


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides the specified minimum yield strength, Fy, and specified minimum tensile
strength, Fu, of stud shear connectors, which are used in routine steel-girder bridges to develop
composite action by preventing slip between the concrete deck and the steel beam or girder. Fu is
used in the computation of the nominal shear resistance, Qn, of a stud shear connector at the
strength limit state in Eq. 6.10.10.4.3-1 (see the Discussion of Article 6.10.10.4.3 in this Guide).
Consult the AASHTO/AWS D1.5M/D1.5 Bridge Welding Code for further information on the
specifications for material, manufacturing, physical properties, certification, and welding of stud
shear connectors.

6.4.5 Weld Metal


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article refers to the AASHTO/AWS D1.5M/D1.5 Bridge Welding Code (BWC) for the
specific requirements for weld metal, including the specification of the weld metal and flux by the
appropriate AWS designation. Virtually all routine steel I-girder bridges utilize structural steel
welds. Using AWS classifications, the BWC prescribes which consumables may be used with
various base metals and the rules for their use. Fabricators choose welding processes and

NSBA Guide to Navigating Routine Steel Bridge Design / 134


associated consumables to be used on the bridge and develop and follow welding procedure
specifications (WPSs) that conform to BWC requirements.
Consult the FHWA Bridge Welding Reference Manual for additional information on weld metal,
welding processes, and the appropriate designations of welding consumables.

6.4.6 Cast Metal

6.4.6.1 Cast Steel and Ductile Iron

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
This Article provides the appropriate material specifications for cast steel and ductile iron castings.
Routine steel I-girder bridges do not typically contain castings and so this Article is designated as
not applicable.

6.4.6.2 Malleable Castings

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
This Article provides the appropriate material specification and the minimum specified yield
strength for malleable castings. Routine steel I-girder bridges do not typically contain castings and
so this Article is designated as not applicable.

6.4.6.3 Cast Iron

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
This Article provides the appropriate material specification for cast iron castings. Routine steel I-
girder bridges do not typically contain castings and so this Article is designated as not applicable.

6.4.7 Stainless Steel


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article contains requirements related to the material specifications for stainless steel. Routine
steel I-girder bridges do not typically utilize stainless steel for structural members such as girders,
cross-frames, or diaphragms, but the provisions may apply for bearings with stainless steel sliding
surfaces.

NSBA Guide to Navigating Routine Steel Bridge Design / 135


6.4.8 Cables

6.4.8.1 Bright Wire

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
This Article provides the appropriate material specification for bright wire used in bridge cables.
Routine steel I-girder bridges do not make use of cables and so this Article is designated as not
applicable.

6.4.8.2 Galvanized Wire

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
This Article provides the appropriate material specification for galvanized wire used in bridge
cables. Routine steel I-girder bridges do not make use of cables and so this Article is designated
as not applicable.

6.4.8.3 Epoxy-Coated Wire

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
This Article provides the appropriate material specification for epoxy-coated wire used in bridge
cables. Routine steel I-girder bridges do not make use of cables and so this Article is designated
as not applicable.

6.4.8.4 Bridge Strand

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
This Article provides the appropriate material specification for strand used in bridge cables.
Routine steel I-girder bridges do not make use of cables and so this Article is designated as not
applicable.

6.4.9 Dissimilar Metals


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article provides requirements to help prevent galvanic corrosion when steel components are
coupled with aluminum components in the presence of an electrolyte. Aluminum is not used in the

NSBA Guide to Navigating Routine Steel Bridge Design / 136


fabrication of structural members such as girders or cross-frames/diaphragms for routine steel I-
girder bridges. This Article would only be applicable to routine steel I-girder bridges in situations
where aluminum components (such as sign support brackets, for example) might be fastened to
the structural steel.

6.5 LIMIT STATES

6.5.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article presents the four limit states for which structural steel-bridge components must be
proportioned to satisfy the applicable design requirements specified at each of these limit states.
Three of these limit states are applicable or partially applicable to routine steel I-girder bridges, as
explained further in the Discussions for Articles 6.5.2, 6.5.3, 6.5.4, and 6.5.5 in this Guide.
For further information on each limit state, consult the FHWA’s Steel Bridge Design Handbook,
Volume 10 - Limit States, and Section 1.3 of the Reference Manual for NHI Course 130081, Load
and Resistance Factor Design (LRFD) for Highway Bridge Superstructures).

6.5.2 Service Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The service limit state is taken as restrictions on stress, deformation, and deck crack width under
regular service conditions. This Article requires that flexural members in routine steel I-girder
bridges satisfy the service limit state checks specified in Article 6.10.4 (see the Discussion of
Article 6.10.4 in this Guide) to prevent objectionable permanent deformations under expected
severe traffic loadings that may impair rideability (i.e., under the Service II load combination
specified in Table 3.4.1-1, which applies a load factor of 1.30 to the live load force effects).
In addition, Article 6.10.1.7 (see the Discussion of Article 6.10.1.7 in this Guide) presents
provisions intended to control the width of cracks in the concrete decks in the negative moment
regions of multi-span continuous steel-girder bridges under regular service conditions (i.e., under
the Service II load combination specified in Table 3.4.1-1). This Article also points to optional
span-to-depth ratios and live-load deflection requirements specified in Article 2.5.2.6. Most
Owners choose to enforce a live-load deflection requirement at the Service I limit state (which
applies a load factor of 1.0 to the live load force effects); consult the applicable Owner-agency
policy. See the Discussion of Article 2.5.2.6.2 in this Guide for further information on the
computation of the live-load deflection.
For further background and explanation of the service limit state, consult the FHWA’s Steel Bridge
Design Handbook, Volume 10 - Limit States, and Section 1.3 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.

NSBA Guide to Navigating Routine Steel Bridge Design / 137


Note that the resistance factors, ϕ, used in service limit state calculations are implicitly taken equal
to 1.0 for all members and components.

6.5.3 Fatigue and Fracture Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
The fatigue limit state is taken as restrictions on stress ranges resulting from the passage of a single
design truck specified in Article 3.6.1.4 occurring at the number of expected stress range cycles.
The fracture limit state is taken as a set of material toughness requirements. This Article requires
that components and details in routine steel I-girder bridges satisfy the fatigue limit state checks
and fracture toughness considerations specified in Article 6.6 (see the Discussion of Article 6.6 in
this Guide). In addition, flexural members in routine steel I-girder bridges must satisfy the
additional fatigue limit state checks specified in Article 6.10.
The provisions of Article 6.13.2.10.3 for bolts subject to tensile fatigue (see the Discussion of
Article 6.13.2.10.3 in this Guide) are not applicable for the routine steel I-girder bridges covered
by this Guide because the typical bolted connections in such bridges are not subject to axial
tension.
For further background and explanation of the fatigue and fracture limit state, consult the FHWA’s
Steel Bridge Design Handbook, Volume 10 - Limit States, and Section 1.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures.
Note that the resistance factors, ϕ, used in fatigue limit state calculations are implicitly taken equal
to 1.0 for all members and components.

6.5.4 Strength Limit State

6.5.4.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The strength limit state is investigated to check that strength and both local and global stability are
provided to resist the statistically significant load combinations that a bridge is expected to
experience over its design life. This Article requires that steel I-girder bridges satisfy the applicable
strength and stability checks in Section 6 in the final condition for the factored force effects at the
strength limit state calculated using the appropriate strength load combinations specified in Table
3.4.1-1, and also for the force effects acting on the fully erected steelwork during the deck
placement calculated using the special load combination specified in Article 3.4.2.1.
As can be seen by reviewing these load combinations, many of the load factors are greater than
1.0; the load factors in Table 3.4.1-1 were calculated by means of statistical analyses to envelope
possible overload conditions which are considered “statistically significant” (i.e., overload
conditions which have a certain probability of occurring over the anticipated life of the structure).

NSBA Guide to Navigating Routine Steel Bridge Design / 138


The load factor of 1.4 in the special load combination specified in Article 3.4.2.1 for evaluating
the constructibility of primary steel superstructure components for the force effects applied to the
fully erected steelwork was not statistically calibrated; instead, it was selected to provide a level
of strength and stability during critical construction stages (where unintended events could
potentially lead to significantly larger force effects than those predicted during the design) which
at least approaches that attained in the past using previous design methodologies. A similar load
combination is specified in the AISC LRFD Specification for Structural Steel Buildings.
For further background and explanation of the strength limit state, consult the FHWA’s Steel
Bridge Design Handbook, Volume 10 - Limit States, and Section 1.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures.

6.5.4.2 Resistance Factors


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article lists the resistance factors, ϕ, applied to the nominal resistance, Rn, of members and
components at the strength limit state to compute the factored resistance, Rr. Some, but not all, the
resistance factors are applicable to routine steel I-girder bridges.
The resistance factors are implicitly taken equal to 1.0 for members and components at the service
and fatigue and fracture limit states.
Designers should thoroughly review the entire list of resistance factors to identify the correct
resistance factor for each specific design calculation.

6.5.5 Extreme Event Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The extreme event limit state is investigated to check that the bridge can survive a specified major
earthquake, flood, vessel collision, vehicle collision, or ice collision event, possibly under scoured
conditions. Routine steel I-girder bridges as defined for the purposes of this Guide are assumed to
be located in Seismic Zone 1 and not subject to stream flow loading, ice loading, or vessel collision
loading of the superstructure. Also, the provisions of this Article are not applicable to the
superstructure design for routine steel I-girder bridges located in Seismic Zone 1. Therefore, this
Article is not applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.6 FATIGUE AND FRACTURE CONSIDERATIONS

6.6.1 Fatigue

6.6.1.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 139


Discussion:
This Article categorizes fatigue into load-induced fatigue (Article 6.6.1.2) and distortion-induced
fatigue (Article 6.6.1.3).
For further information on fatigue design, consult the FHWA’s Steel Bridge Design Handbook,
Volume 12 - Design for Fatigue, the Reference Manual for NHI Course 130122, Design and
Evaluation of Steel Bridges for Fatigue and Fracture, and Section 6.5.5 of the Reference Manual
for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples illustrating load-induced fatigue design computations,
consult the FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous
Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous Straight
Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span Continuous Straight
Composite Steel Wide-Flange Beam Bridge.

6.6.1.2 Load-Induced Fatigue

6.6.1.2.1 Application
Determination of applicability, Simple Span Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
Load-induced fatigue is defined as fatigue effects due to the in-plane stresses for which
components and details are explicitly designed. This Article indicates that the force effect to be
considered in the load-induced fatigue design of components and details in routine steel I-girder
bridges is the live load stress range, or the algebraic difference between the maximum and
minimum fatigue live-load stresses in the component or at the detail under consideration due to
the fatigue live load placed in a single lane.
In routine steel multi-span continuous I-girder bridges with shear connectors provided throughout
their length and longitudinal deck reinforcement satisfying the provisions of Article 6.10.1.7 (see
the Discussion of Article 6.10.1.7 in this Guide), the concrete deck may be considered effective in
tension for computing the stress due to the negative (minimum) fatigue live-load moment (using
the short-term modular ratio, n, to transform the concrete deck) when calculating the stress range.
This is strongly recommended when computing the stress range at details on beams or girders in
routine steel multi-span continuous I-girder bridges satisfying the preceding criteria; recognition
of this behavior will significantly reduce the fatigue stress ranges at details on or adjacent to the
top flanges in regions of negative flexure or stress reversal. The concrete deck may also be
considered effective in tension in such cases when calculating the stress due to the unfactored
permanent loads applied to the composite section, i.e., DC2 and DW loads, at the fatigue limit state
(using the long-term modular ratio, 3n, in this case to transform the concrete deck).
This Article also provides the criterion to determine if a component or detail is subject to a net
tensile stress and therefore must be checked for fatigue. This criterion is applicable to components

NSBA Guide to Navigating Routine Steel Bridge Design / 140


and details in routine steel multi-span continuous I-girder bridges located in regions where the
unfactored permanent loads produce compression. In such cases, fatigue is to be checked only if
the factored tensile stress due to the negative (minimum) fatigue live-load moment (factored for
the Fatigue I load combination specified in Table 3.4.1-1; i.e., with a load factor of 1.75 applied to
the fatigue live-load moment) exceeds the unfactored permanent load compressive stress in the
component or at the detail. Of course, this discussion does not apply for simple span routine steel
I-girder bridges as they are only subject to positive moments.
The discussion in the Commentary for this Article related to the checking of fatigue in cross-frames
or diaphragms using the force effects computed from a refined analysis does not apply to the
routine steel I-girder bridges covered by this Guide. Live-load force effects in cross-frame or
diaphragm members are not available from the line girder analysis methods, which are normally
used for the design of routine steel I-girder bridges. Designers need not be concerned about
performing a fatigue analysis of cross-frame or diaphragm members in routine steel I-girder
bridges; that is, due to the nature of the geometry of the framing plan and overall layout of routine
steel I-girder bridges, the live load force effects (and the resulting live load stress ranges) in the
cross-frames or diaphragms of these bridges are typically not significant.
For further information on fatigue design, consult the FHWA’s Steel Bridge Design Handbook,
Volume 12 - Design for Fatigue, the Reference Manual for NHI Course 130122, Design and
Evaluation of Steel Bridges for Fatigue and Fracture, and Section 6.5.5 of the Reference Manual
for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples illustrating load-induced fatigue design computations,
consult the FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous
Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous Straight
Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span Continuous Straight
Composite Steel Wide-Flange Beam Bridge.
Note that many commercial steel bridge design programs, such as NSBA's LRFD Simon line-
girder analysis and design program, can perform the fatigue evaluation of the details on the girders
of a routine steel I-girder bridge subject to a net tensile stress as specified in this Article. These
programs will calculate the appropriate fatigue stress range in the girder flanges at these details for
either the Fatigue I or Fatigue II limit-state load combination, as applicable, and compare it to the
nominal fatigue resistance of the detail, determined for either an infinite life or a finite life
evaluation as appropriate, based on parameters such as the detail category and the Average Daily
Truck Traffic in a single lane in one direction, (ADTT)SL, input by the user. Users should verify
the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.6.1.2.2 Design Criteria


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides the basic equation for checking load-induced fatigue for components and
details in steel I-girder bridges subject to a net tensile stress as specified in Article 6.6.1.2.1 (see

NSBA Guide to Navigating Routine Steel Bridge Design / 141


the Discussion of Article 6.6.1.2.1 in this Guide); i.e., the live-load stress range due to the passage
of the fatigue live load in a single lane (see the Discussion of Article 3.6.1.4.1 in this Guide),
factored for either the Fatigue I or Fatigue II load combination specified in Table 3.4.1-1, as
applicable, must not exceed the nominal fatigue resistance of the component or detail (see the
Discussion of Article 6.6.1.2.5 in this Guide).
For further information on fatigue design, consult the FHWA’s Steel Bridge Design Handbook,
Volume 12 - Design for Fatigue, the Reference Manual for NHI Course 130122, Design and
Evaluation of Steel Bridges for Fatigue and Fracture, and Section 6.5.5 of the Reference Manual
for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures.

6.6.1.2.3 Detail Categories


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides an important table that defines the specific Detail Category for the load-
induced fatigue design of typical components and details that may be encountered on steel bridges
(Table 6.6.1.2.3-1). For the routine steel I-girder bridges covered by this Guide, the most common
fatigue details that must be examined as a minimum when subject to a net tensile stress as specified
in Article 6.6.1.2.1 (see the Discussion of Article 6.6.1.2.1 in this Guide) are the base metal away
from welds and connections (Condition 1.1 in the table – Category A for all steels except for
uncoated weathering steel; Condition 1.2 – Category B for uncoated weathering steel); base metal
adjacent to continuous flange-to-web fillet welds (Condition 3.1 – Category B); the base metal at
the toe of transverse stiffener-to-flange fillet welds and transverse stiffener-to-web fillet welds,
including bearing stiffener and cross-frame or diaphragm connection plate welds (Condition 4.1 –
Category C'); the base metal adjacent to complete joint penetration groove-welded flange splices
(Condition 5.1 – Category B) and; the base metal at stud-type shear connectors attached to the top
flange by welds (Condition 9.1 – Category C). Other conditions in the table typically do not apply
to routine steel I-girder bridges, including Conditions 8.1 through 8.9 covering orthotropic deck
details.
In routine multi-span continuous I-girder bridges, the Category C' fatigue check of the connection
plate-to-bottom flange weld (Condition 4.1) may control the size of the bottom flange in regions
of positive flexure. If the nominal fatigue resistance is exceeded, consideration should be given to
either moving the connection plate to a slightly different location or increasing the size of the
bottom flange; bolting the connection plate to the flange to raise the fatigue category to Category
B (Condition 2.1) is not recommended as this detail is typically more expensive to fabricate and
there is a fatigue Category C' detail at the termination of the connection plate-to-web weld only a
short distance above the flange.
Fatigue does not typically control for bolted field splices The nominal fatigue resistance of base
metal at the gross section adjacent to slip-critical bolted connections with the bolts installed in
holes drilled full size or subpunched and reamed to size is Category B (Condition 2.1); the
combined areas of the flange and web splice plates must equal or exceed the areas of the smaller

NSBA Guide to Navigating Routine Steel Bridge Design / 142


flanges and web to which they are attached, and the flanges and web are usually checked separately
for either equivalent or more critical fatigue category details.
The welded connections typically used to attach angle or tee section (WT) cross-frame members
to gusset plates in the truss-type cross-frames used in many steel I-girder bridges are designated
as Category E' details (Condition 7.2), and as such have very low fatigue resistance. Bolted
connections of cross-frame members to gusset plates, which are less commonly used and are not
recommended, are typically punched full size by the fabricator and are designated as Category D
details (Conditions 2.3 and 2.5). Where permitted for use, this Article states that unless specific
information is available to the contrary, bolt holes in bracing members and their connection plates
are to be assumed for design to be punched full size. However, as explained in the Discussion of
Article 6.6.1.2.1 in this Guide, designers need not be concerned about performing a fatigue analysis
of cross-frame or diaphragm members in routine steel I-girder bridges; that is, due to the nature of
the geometry of the framing plan and overall layout of routine steel I-girder bridges, the live load
force effects (and the resulting live load stress ranges) in the cross-frames or diaphragms of these
bridges are typically not significant.
This Article also provides another important table giving the 75-year Average Daily Truck Traffic
in a single lane in one direction, (ADTT)SL, Equivalent to Infinite Life for each fatigue-detail
category (Table 6.6.1.3.2-2). The 75-year (ADTT)SL may be calculated as the fraction of traffic in
a single lane, p, given in Table 3.6.1.4.2-1 times the ADTT in one direction averaged over the
design life. Article C3.6.1.4.2 provides helpful recommendations on how to estimate the 75-year
ADTT in one direction (see the Discussion of Article 3.6.1.4.2 in this Guide). For the component
or detail under consideration, if the calculated 75-year (ADTT)SL in a single lane in one direction
exceeds the value of the 75-year (ADTT)SL Equivalent to Infinite Life specified in Table 6.6.1.2.3-
2 for the corresponding fatigue-detail category, that component or detail is to be designed for
infinite life using the Fatigue I load combination in Table 3.4.1-1. Otherwise, the component or
detail is to be designed for finite life using the Fatigue II load combination (see the Discussion of
Article 6.6.1.2.5 in this Guide). In such cases, details on the routine steel I-girder bridges covered
by this Guide should not be designed for infinite life, unless required to do so by Owner-agency
policy. The values given in Table 6.6.1.2.3-2 are calculated from Eq. C6.6.1.2.3-1. For a number
of stress range cycles per truck passage, n, other than 1.0, the values in Table 6.6.1.2.3-2 should be
modified by dividing the values by the appropriate value of n taken from Table 6.6.1.2.5-2. For values
of the fatigue design life other than 75 years, the values in Table 6.6.1.2.3-2 should be modified by
multiplying the values by the ratio of 75 divided by the fatigue life sought in years.
The remaining provisions in this Article dealing with Fracture-Critical Members and orthotropic deck
components and details are not applicable to the routine steel I-girder bridges covered by this Guide.
For further information on fatigue design, consult the FHWA’s Steel Bridge Design Handbook,
Volume 12 - Design for Fatigue, the Reference Manual for NHI Course 130122, Design and
Evaluation of Steel Bridges for Fatigue and Fracture, and Section 6.5.5 of the Reference Manual
for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures.
Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006 Guidelines for
Design Details and G12.1-2020 Guidelines to Design for Constructability and Fabrication for

NSBA Guide to Navigating Routine Steel Bridge Design / 143


recommendations, commentary, and sample design details that allow for more economical
fabrication and erection.
Note that many commercial steel bridge design programs, such as NSBA's LRFD Simon line-
girder analysis and design program, can perform the fatigue evaluation of the details on the girders
of a routine steel I-girder bridge subject to a net tensile stress as specified in Article 6.6.1.2.1.
These programs will calculate the appropriate fatigue stress range in the girder flanges for either
the Fatigue I or Fatigue II limit-state load combination, as applicable, and compare it to the nominal
fatigue resistance of the detail, determined for either an infinite life or a finite life evaluation as
appropriate, based on parameters such as the detail category and the Average Daily Truck Traffic
in a single lane in one direction, (ADTT)SL, input by the user. Users should verify the capabilities,
assumptions, and general correctness of any program’s calculations prior to initial use.

6.6.1.2.4 Detailing to Reduce Constraint


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides recommended guidelines for specific details involving intersecting welded
elements to avoid highly constrained joints and crack-like geometric discontinuities that could
potentially be susceptible to constraint-induced fracture (CIF). Routine steel I-girder bridges
covered by this Guide do not typically contain such details (e.g., longitudinal web stiffeners
intersecting transverse web stiffeners or lateral connection plates intersecting transverse web
stiffeners), and so the provisions of this Article are not applicable to the routine steel I-girder
bridges covered by this Guide.

6.6.1.2.5 Fatigue Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides the provisions necessary to compute the nominal fatigue resistance, (ΔF)n,
of components or details subject to a net tensile stress as specified in Article 6.6.1.2.1 (see the
Discussion of Article 6.6.1.2.1 in this Guide) in steel I-girder bridges.
For components and details to be designed for infinite life under the Fatigue I load combination,
which is representative of the maximum stress range of the truck population, the nominal fatigue
resistance is simply taken equal to the constant-amplitude fatigue threshold, (ΔF)TH, which is given
for each fatigue-detail category in Table 6.6.1.2.5-3. This occurs whenever the calculated 75-year
(ADTT)SL exceeds the value of the 75-year (ADTT)SL Equivalent to Infinite Life specified in Table
6.6.1.2.3-2 for the applicable fatigue-detail category (see the Discussion of Article 6.6.1.2.3 in this
Guide).
For components and details to be designed for finite life under the Fatigue II load combination,
which is representative of the effective stress range of the truck population, the nominal fatigue
resistance is taken equal to (A/N)1/3, where A is the y-intercept of the sloping portion of the S-N
curve given for each fatigue-detail category as specified in Table 6.6.1.2.5-1 and N is the number

NSBA Guide to Navigating Routine Steel Bridge Design / 144


of stress range cycles given by Eq. 6.6.1.2.5-3. This occurs whenever the calculated 75-year
(ADTT)SL is less than or equal to the value of the 75-year (ADTT)SL Equivalent to Infinite Life
specified in Table 6.6.1.2.3-2 for the applicable fatigue-detail category (see the Discussion of
Article 6.6.1.2.3 in this Guide). In such cases, details on the routine steel I-girder bridges covered
by this Guide should not be designed for infinite life, unless required to do so by Owner-agency
policy. If a fatigue design life other than 75 years is sought, a number other than 75 may be inserted
in the equation for N. The necessary adjustments to the tabulated values given in Table 6.6.1.2.3-
2 for a fatigue design life other than 75 years are discussed in Article C6.6.1.2.3 (see the Discussion
of Article 6.6.1.2.3 in this Guide). Values of A and (ΔF)n specified for bolts subject to axial tension
in Tables 6.6.1.2.5-1 and 6.6.1.2.5-3, respectively, are not applicable to routine steel I-girder
bridges.
Eq. 6.6.1.2.5-4 for calculating the nominal fatigue resistance of base metal at details where loaded
discontinuous plate elements are connected with a pair of fillet welds or partial joint penetration
groove welds on opposite sides of the plate normal to the direction of primary stress, or where
partial joint penetration groove welds are transversely loaded in tension, is not applicable to the
routine steel I-girder bridges covered by this Guide. These types of details, including details
utilizing partial joint penetration groove welds, are not typically used in these bridges.
For simple span bridges, the number of stress cycles per truck passage, n, used in the computation
of N in Eq. 6.6.1.2.5-3 is taken equal to 1.0 (Table 6.6.1.2.5-2). For multi-span continuous steel
plate girder or rolled-beam bridges, the number of stress cycles per truck passage, n, used in the
computation of N in Eq. 6.6.1.2.5-3 is taken equal to 1.5 for details located within one-tenth of the
span length on either side of an interior support; otherwise, n is taken equal to 1.0 (Table 6.6.1.2.5-
2). The other values of n specified in Table 6.6.1.2.5-2 are not applicable to the routine steel I-
girder bridges covered by this Guide. The necessary adjustments to the tabulated values of the 75-
year (ADTT)SL Equivalent to Infinite Life given in Table 6.6.1.2.3-2 for a value of n other than 1.0
are discussed in Article C6.6.1.2.3 (see the Discussion of Article 6.6.1.2.3 in this Guide).
For further information on fatigue design, consult the FHWA’s Steel Bridge Design Handbook,
Volume 12 - Design for Fatigue, the Reference Manual for NHI Course 130122, Design and
Evaluation of Steel Bridges for Fatigue and Fracture, and Section 6.5.5 of the Reference Manual
for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples illustrating load-induced fatigue design computations,
consult the FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous
Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous Straight
Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span Continuous Straight
Composite Steel Wide-Flange Beam Bridge.
Note that many commercial steel bridge design programs, such as NSBA's LRFD Simon line-
girder analysis and design program, can perform the fatigue evaluation of the details on the girders
of a routine steel I-girder bridge subject to a net tensile stress as specified in Article 6.6.1.2.1.
These programs will calculate the appropriate fatigue stress range in the girder flanges for either
the Fatigue I or Fatigue II limit-state load combination, as applicable, for comparison to the
nominal fatigue resistance of the detail, determined for either an infinite life or a finite life
evaluation as appropriate, based on parameters such as the detail category and the Average Daily

NSBA Guide to Navigating Routine Steel Bridge Design / 145


Truck Traffic in a single lane in one direction, (ADTT)SL, input by the user. Users should verify
the capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.6.1.3 Distortion-Induced Fatigue


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article are applicable to the routine steel I-girder bridges covered by this
Guide to check that rigid load paths are provided to adequately transmit the forces in transverse
bracing members from the beam or girder web to the flanges in order to prevent distortion-induced
fatigue, which is defined as fatigue effects due to secondary out-of-plane stresses not normally
quantified in the typical analysis and design of a bridge. The rigid load paths are provided by
attaching the various components through either welding or bolting.
The provision of Article 6.10.5.3 mentioned in this Article is applicable to the routine steel I-girder
bridges covered by this Guide (see the Discussion of Article 6.10.5.3 in this Guide).

6.6.1.3.1 Transverse Connection Plates


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
The provisions of this Article relate to the detailing necessary to prevent distortion-induced fatigue
from occurring at transverse connection plate details (i.e., vertical stiffeners attached to a beam or
girder to which a cross-frame or diaphragm is attached, including bearing stiffeners).
Except for the special case of straight, rolled-beam bridges with supports skewed less than or equal
to 10 degrees from normal as permitted later in this Article, the first bulleted item in this Article
requiring transverse connection plates to be positively attached (by welding or bolting) to the
compression and tension flanges of the beam or girder applies to the routine steel I-girder bridges
covered by this Guide. Welded connections are preferred as bolted connections possessing
sufficient stiffness are not likely to be economical. The connections for routine steel I-girder
bridges should be designed as a minimum for the larger of the calculated resultant force in those
members or the factored 20-kip lateral load specified in this Article for straight, nonskewed
bridges.
The second and third bullets in this Article are not applicable to the routine steel I-girder bridges
covered by this Guide as such bridges do not contain internal or external cross-frames or
diaphragms, floor beams, or stringers.
For the for the special case of straight, rolled-beam bridges with composite reinforced decks,
whose supports are skewed less than or equal to 10 degrees from normal, with the diaphragms
placed in contiguous lines parallel to the supports, an option is provided to allow the use of less
than full-depth end angles or connection plates bolted or welded to the web if permitted by the
Owner; the end angles or plates must satisfy the requirements specified in this Article.

NSBA Guide to Navigating Routine Steel Bridge Design / 146


Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006 Guidelines for
Design Details and G12.1-2020 Guidelines to Design for Constructability and Fabrication for
recommendations, commentary, and sample design details that allow for more economical
fabrication and erection.

6.6.1.3.2 Lateral Connection Plates


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article relate to the detailing necessary to prevent distortion-induced fatigue
from occurring at lateral connection plate details (i.e., plates used to interconnect lateral bracing
members for attachment to a flexural member). These provisions are not applicable to the routine
steel I-girder bridges covered by this Guide, which are assumed not to contain any lateral bracing.

6.6.1.3.3 Orthotropic Decks


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article relate to the detailing necessary to prevent distortion-induced fatigue
from occurring at orthotropic deck details. These provisions are not applicable to the routine steel
I-girder bridges covered by this Guide, which are assumed to have only concrete decks.

6.6.2 Fracture

6.6.2.1 Member or Component Designations and Charpy V-Notch Testing


Requirements
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides member or component designations (i.e., primary or secondary) and Charpy
V-notch testing requirements to evaluate the fracture toughness of the steel for the applicable
temperature zone. These provisions are applicable to the routine steel I-girder bridges covered by
this Guide.
Beams and girders, including the splice plates for bolted field splices, are designated as primary
members for the purposes of this Article. Cross-frame members or diaphragms, cross-frame gusset
plates, transverse connection plates, transverse intermediate web stiffeners, bearing stiffeners, and
any nonstructural components or attachments (e.g., expansion dams, drainage material, brackets,
etc.) in these bridges are designated as secondary members for the purposes of this Article (Table
6.6.2.1-1). Arbitrary or “conservative” designation of secondary members or components as
primary members or components is discouraged, as this will invoke more costly and complex
fabrication and testing requirements that do not add significant value and are not necessary. See
also the Discussions of Articles 6.8.4 and 6.9.3 for situations when cross-frame members may
be considered as “primary members” when evaluating slenderness ratios.

NSBA Guide to Navigating Routine Steel Bridge Design / 147


Primary members or components, or portions thereof, subject to a net tensile stress under the
Strength I limit state (Table 3.4.1-1) must be designated on the contract plans. Charpy V-notch
testing is required for primary members or components subject to a net tensile stress, or for
portions thereof located in designated tension zones, under Strength I. The testing is done by the
steel producers.
Specifying that Charpy V-notch testing be performed for secondary members (e.g., cross-frame or
diaphragm members) adds additional complexity and cost without providing any significant
additional value and should not be done for routine steel-girder bridges.
The minimum Charpy V-notch toughness requirements for various grades of steel for the three
temperature zones specified in Table 6.6.2.1-2 are given in Table C6.6.2.1-1. The toughness
requirements in Table C6.6.2.1-1 for fracture-critical members and for Grades HPS 50W, HPS
70W, and HPS 100W are not applicable for the routine steel I-girder bridges covered by this
Guide.
Generally, Owner-agency policy will establish the minimum service temperature used to
determine the Temperature Zone in Table 6.6.2.1-2.

6.6.2.2 Fracture-Critical Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains provisions for fracture-critical members (FCMs), which are defined as steel
primary members or portions thereof whose failure would probably cause a portion of or the entire
bridge to collapse. The routine steel I-girder bridges covered by this Guide do not contain FCMs.
A routine steel I-girder bridge, as defined in this Guide, features a cross-section with four or more
girders and a composite concrete deck, and so provides multiple redundant load paths for gravity
loads (dead loads and live loads). As such, the provisions of this Article are not applicable.

6.7 GENERAL DIMENSION AND DETAIL REQUIREMENTS

6.7.1 Effective Length of Span


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides the definition of the effective length of a span, which is typically measured
between bearing locations and is applicable to the routine steel I-girder bridges covered by this
Guide.

6.7.2 Dead Load Camber and Detailing of Structural Components


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 148


Discussion:
This Article contains provisions related to determining structure dead load deflections and cambers
and specifying them in contract documents. For routine steel I-girder bridges, deflections
accounting for the self-weight of the steel, the weight of the wet concrete deck on the noncomposite
section, and additional dead loads acting on the composite section (such as the weight of barriers
or a future wearing surface), should be determined separately, and typically can be calculated using
the same line-girder analysis model used to calculate girder force effects (such as NSBA’s LRFD
Simon line-girder analysis and design program). The specific method for specifying vertical
cambers on the contract drawings is often dictated by Owner-agency requirements.
This Article also discusses the considerations for reporting dead load deflections and vertical
cambers when staged deck placement or phased construction is specified. In the context of the
AASHTO LRFD BDS, the term “staged deck placement” refers to the placement of the deck in
discrete pours that are the full width of the bridge deck but only part of the length of the bridge.
The pours should be placed on the deck in a specific sequence identified to minimize tensile
stresses in the deck, typically by placing pours in the positive moment regions first, and negative
moment regions later. In the context of the AASHTO LRFD BDS, the term “phased construction”
refers to building the bridge in partial-width phases, where one phase is fully constructed (i.e., both
the steel superstructure and the deck are constructed), prior to the construction of the next phase.
Bridge widenings are conceptually similar to phased construction and behave similarly.
For routine steel I-girder bridges, line-girder analysis software (such as NSBA’s LRFD Simon
line-girder analysis and design program) can typically accommodate the analysis for staged deck
placement (or pours) along the length of the structure (see the Discussion of Article 6.10.3.4 in
this Guide).
For routine steel I-girder bridges as defined in this Guide, it is assumed that only single-phase
construction, simple multi-phase construction, or simple bridge widenings are being considered
(see the Definition of a “Routine Steel I-Girder Bridge”), so as to maintain the applicability of a
line-girder analysis. For more complicated situations, such as cases where a closure pour is not
provided between adjacent phases of construction, the use of a refined analysis method may be
warranted for proper calculation of load distribution, stresses, and deflections. Refer to the
AASHTO-NSBA Steel Bridge Collaboration’s Guideline G13.1-2019 Guidelines for Steel Girder
Bridge Analysis and to Sections 6.3.2.5.4 and 6.5.3.3 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures for
further guidance on staged deck placement and phased construction considerations.
This Article also presents requirements related to changes in component length to account for
cambering of structures such as trusses, arches, and cable-stayed systems. These requirements do
not apply to the routine steel I-girder bridges covered by this Guide.
Finally, this Article requires the identification of a “fit condition” in the contract documents for
certain specific types of I-girder bridges. The identification of the desired fit condition of an I-
girder bridge refers to the identification of the targeted girder dead load geometry condition (no
load, steel dead load, or total dead load) for which the cross-frames or diaphragms are detailed to
connect to the girders. The Fabricator/Detailer determines the cross-frame or diaphragm geometry

NSBA Guide to Navigating Routine Steel Bridge Design / 149


based on the vertical deflections provided in the contract documents and the specified fit condition.
I-girder bridges requiring the specification of a fit condition include the following:
• straight bridges where one or more support lines are skewed more than 20 degrees from
normal;
• horizontally-curved bridges where one or more support lines are skewed more than 20
degrees from normal and with an L/R in all spans less than or equal to 0.03; and
• horizontally-curved bridges with or without skewed supports and with a maximum L/R
greater than 0.03.
For the purposes of this Guide, routine steel I-girder bridges do not meet any of these criteria and
the requirement to identify a desired fit condition, along with requirements to consider the effect
of the selected fit condition on bearing rotations mentioned in the last two paragraphs of this Article
and extensively discussed in the Commentary, do not apply.
For interested readers, the NSBA has published both a brief guide on steel I-girder bridge fit,
Skewed and Curved Steel I-Girder Bridge Fit (Executive Summary), and a longer, more in-depth,
white paper on the topic, Skewed and Curved Steel I-Girder Bridge Fit (Full White Paper).

6.7.3 Minimum Thickness of Steel


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article specifies the minimum thickness (0.3125 in.) for structural steel, such as girder webs
and flanges, bracing members (i.e., cross-frames or diaphragm members), stiffeners, and gusset
plates/connection plates for cross-frames or diaphragms. The webs of rolled beams or channels are
sometimes thinner than 0.3125 in. and are exempted from this requirement. Additionally, filler
plates and material in barrier railings are exempted as well. The AASHTO-NSBA Steel Bridge
Collaboration Guideline G12.1-2020 Guidelines to Design for Constructability and Fabrication
suggests practical values of the minimum thickness of many of these components, which typically
significantly exceed the absolute minimum thickness value specified herein.
The additional requirements in this Article for truss gusset plates and orthotropic decks do not
apply to the routine steel I-girder bridges covered by this Guide.
Consult Owner-agency policy regarding the need to provide special protection against corrosion
or to specify a sacrificial metal thickness to account for potential section loss due to corrosion.

6.7.4 Diaphragms and Cross-Frames

6.7.4.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 150


Discussion:
This Article covers spacing and detailing requirements and other general design requirements for
cross-frames or diaphragms in various steel structure types. In the AASHTO LRFD BDS, a cross-
frame is defined as a transverse truss framework connecting adjacent longitudinal flexural
components or inside a tub section or closed box used to transfer and distribute vertical and lateral
loads and to provide stability to the compression flanges. A diaphragm is defined as a vertically
oriented solid transverse member connecting adjacent longitudinal flexural components or inside
a closed-box or tub section to transfer and distribute vertical and lateral loads and to provide
stability to the compression flanges.
Diaphragms or cross-frames should be placed at end and interior supports, with the spacing of
intermediate diaphragms or cross-frames based on an investigation of the specified stages of
construction and the final condition. At a minimum, the design should consider:
• transfer of wind loads from the bottom of the girder to the deck and from the deck into the
support bearings;
• lateral loading of flanges due to overhang falsework during deck placement;
• providing bracing that contributes to the lateral torsional buckling resistance of the girders
in regions of negative flexure at the strength limit state;
• providing bracing that contributes to the lateral torsional buckling resistance of the girders
in both regions of positive and negative flexure during critical stages of construction;
• limiting flange lateral bending moments to reasonable levels; and
• the distribution of applied dead and live loads across the width of the structure.
The Article specifically prohibits consideration of the contribution of metal stay-in-place deck
forms to the stability of the top flange of the noncomposite member in compression.
For the routine steel I-girder bridges covered by this Guide, i.e., bridges without curvature or
significant skew, the spacing of cross-frames or diaphragms is dictated primarily by the need to
limit flange lateral bending stresses due to overhang bracket or wind loads and to provide an
appropriate unbraced length to control lateral-torsional buckling.
As stated in the Commentary for this Article, the AASHTO 25.0-ft spacing limit on intermediate
cross-frames or diaphragms that had existed in Specifications prior to the AASHTO LRFD BDS
has been removed. However, for a routine I-girder bridge, this is still a reasonable upper limit to
achieve reasonably sized flanges and bracing members and to control stresses in the concrete deck.
Additionally, Owner-agency policy may dictate maximum spacing limits.
Designers may choose to vary the cross-frame or diaphragm spacing within each span, using
tighter spacings near interior supports to reduce the unbraced length when the bottom flange is in
compression. Variations in spacing should be kept to a minimum; using multiple different spacings
within a given span in response to changes in the magnitude of girder moments is neither warranted
nor recommended. During the layout of the framing plan, a cursory review of compression flange

NSBA Guide to Navigating Routine Steel Bridge Design / 151


lateral-torsional buckling resistances in Article 6.10.8.2.3 or A6.3.3, as applicable, can be made to
determine the sensitivity of the lateral-torsional buckling resistance to various unbraced lengths.
The requirements in this Article for including the bracing in the analysis model and considering
the force effects of horizontal curvature do not apply to the routine steel I-girder bridges covered
by this Guide. However, the requirement to design cross-frames or diaphragms to transfer wind
loads and meet slenderness requirements, at a minimum, does apply. In addition, although not
currently specified in the AASHTO LRFD BDS, it is recommended herein that in addition to the
minimum design requirements specified in Article 6.7.4.1 (see the Discussion of Article 6.7.4.1 in
this Guide), cross-frames or diaphragms for the routine steel I-girder bridges covered by this Guide
also be designed to satisfy the stability bracing strength and stiffness requirements specified in
AISC Appendix 6 - Article 6.3.2a. Consult Section 6.6.3.2 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures,
the FHWA’s Steel Bridge Design Handbook, Volume 13 – Bracing System Design, and Coletti,
D.A., and M.A. Grubb, “Practical Implementation of Stability Bracing Strength and Stiffness
Guidelines for Steel I-Girder Bridges,” Proceedings of the 2016 World Steel Bridge Symposium,
April 14, 2016, Orlando, FL for further information on these requirements and their application to
cross-frames and diaphragms in steel I-girder bridges.
Unless the cross-frame or diaphragm members are directly connected to the girder flanges, except
as permitted in Article 6.6.1.3.1, the connection plates are to be attached directly to the beam or
girder flanges, with the attachment designed for the larger of the calculated resultant force in those
members or the recommended minimum 20.0-kip value specified in Article 6.6.1.3.1 (see the
Discussion of Article 6.6.1.3.1 in this Guide). In addition to the wind and slenderness
requirements, the minimum factored 20.0-kip horizontal design force for the attachment between
the cross-frame or diaphragm connection plates and the beam or girder flanges may be considered
as a design limit for the cross-frame or diaphragm member(s) as well.
Finally, the Article requires that deck slab edges be supported per Article 9.4.4, whether by cross-
frame top struts (typically rolled steel beam wide-flange or channel shapes), steel diaphragms,
integral concrete diaphragms, or thickened slab ends. Typical transverse deck reinforcement is not
sized to carry vehicular wheel loads at these discontinuous locations without additional support.
The extra support also minimizes differential movement issues at expansion joint devices. Bridges
with decks continuous over the backwall are typically not subject to this requirement. The Article
also specifies similar supports anywhere the slab continuity is broken between the ends of the
bridge, such as at hinge locations in older structures (a condition that is not applicable to the design
of new routine steel I-girder bridges covered by this Guide).
For further information on cross-frames and diaphragms, consult Section 6.3.2.9 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures and various chapters of FHWA’s Steel Bridge Design Handbook.

6.7.4.2 I-Section Members


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 152


Discussion:
This Article presents design requirements for cross-frames or diaphragms in bridges containing
steel I-section members.
The Article specifies the required minimum depth of cross-frames or diaphragms, as a function of
the overall girder depth, dependent on the girder type (i.e., rolled beam or plate girder).
Additionally, end diaphragm design forces transmitted by the deck and deck joint are discussed as
well as the transfer of these forces to the girders. The provisions of this Article permit solid web
diaphragms to be designed by traditional beam theory when their span-to-depth ratio is greater
than or equal to 4.0. Otherwise, solid web diaphragms should be treated as deep beams and
designed for principal stresses. These requirements are applicable to all steel girder bridges.
For straight bridges with support skews less than or equal to 20 degrees (as is the case for the
routine steel I-girder bridges discussed in this Guide), the applicable provisions of this Article
allow intermediate cross-frames or diaphragms to be placed in contiguous lines parallel to the
skewed support lines. This is recommended, since it simplifies the design, detailing, and
fabrication of the bridge.
Other framing requirements for bridges with larger skew angles and for horizontally curved
bridges are not applicable to the routine steel I-girder bridges covered by this Guide. The
requirement to consider the effect of the tangential component of force in a skewed end cross-
frame or diaphragm on the beam or girder is also considered not applicable, as this effect is not
significant for the routine I-girder bridges covered by this Guide.
Finally, this Article requires that cross-frames or diaphragms at supports be designed to transmit
lateral forces from the superstructure to the bearings that provide lateral restraint. These lateral
forces primarily include the forces due to wind load, seismic, and centrifugal force effects. For
the routine steel I-girder bridges covered by this Guide, only lateral forces due to wind-load effects
need to be considered as these bridges are assumed located in Seismic Zone 1 and are not
horizontally curved.
The FHWA’s Steel Bridge Design Handbook, specifically Design Example 2A, Two-Span
Continuous Straight Composite Steel I-Girder Bridge, provides design examples for intermediate
and end diaphragms.
Although not currently specified in the AASHTO LRFD BDS, it is recommended herein that in
addition to the minimum design requirements specified in Article 6.7.4.1 (see the Discussion of
Article 6.7.4.1 in this Guide), cross-frames or diaphragms for the routine steel I-girder bridges
covered by this Guide also be designed to satisfy the stability bracing strength and stiffness
requirements specified in AISC Appendix 6 - Article 6.3.2a. Consult Section 6.6.3.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures, the FHWA’s Steel Bridge Design Handbook, Volume 13 –
Bracing System Design, and Coletti, D.A., and M.A. Grubb, “Practical Implementation of Stability
Bracing Strength and Stiffness Guidelines for Steel I-Girder Bridges,” Proceedings of the 2016
World Steel Bridge Symposium, April 14, 2016, Orlando, FL for further information on these
requirements and their application to cross-frames and diaphragms in steel I-girder bridges.

NSBA Guide to Navigating Routine Steel Bridge Design / 153


6.7.4.3 Composite Box-Section Members
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies design requirements for cross-frames and diaphragms in bridges containing
composite steel box-section members. The routine steel-girder bridges covered by this Guide are
not comprised of composite steel box-section members; therefore, the provisions of this Article
are not applicable.

6.7.4.4 Noncomposite Box-Section Members

6.7.4.4.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains general design requirements for diaphragms used in noncomposite steel box-
section members. The routine steel-girder bridges covered by this Guide do not utilize
noncomposite steel box-section members; therefore, the provisions of this Article are not
applicable.

6.7.4.4.2 Square and Rectangular HSS Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains design requirements for diaphragms used in noncomposite steel square and
rectangular HSS (Hollow Structural Section) members. The routine steel-girder bridges covered
by this Guide do not utilize noncomposite steel square and rectangular HSS members; therefore,
the provisions of this Article are not applicable. The use of such members in routine steel I-girder
bridges is impractical and uneconomical. There are no provisions or precedent for using these
types of sections as the main spanning elements in routine bridges. Furthermore, these types of
members are more costly to fabricate, and involve more complicated and expensive connection
details than the single-angle members typically used in cross-frames.

6.7.4.4.3 Welded and Nonwelded Built-Up Noncomposite Box-Section Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains design requirements for diaphragms and cross-frames used in welded and
nonwelded built-up noncomposite steel box-section members. The routine steel-girder bridges
covered by this Guide do not utilize welded or nonwelded built-up noncomposite steel box-section
members as either the main spanning member or as bracing members; therefore, the provisions of
this Article are not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 154


6.7.4.5 Trusses and Arches
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies design requirements for diaphragms used in truss and arch bridges. The
routine steel-girder bridges covered by this Guide do not utilize trusses or arches; therefore, the
provisions of this Article are not applicable.

6.7.5 Lateral Bracing

6.7.5.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies general design requirements for lateral bracing in steel bridges of various
types. Lateral bracing is defined in the AASHTO LRFD BDS as a truss placed in a horizontal
plane between two I-girders or two flanges of a tub girder to maintain cross-sectional geometry
and to provide additional stiffness and stability to the bridge system. Lateral bracing is typically
used to control loads and deflections due to wind in longer-span structures (particularly in the fully
erected steelwork prior to the casting of the concrete deck), to transfer superstructure seismic loads
to the supports, and to control cross-section geometry and provide global stability during
fabrication, erection and deck placement for longer-span straight or curved structures or narrow
straight or curved I-girder bridge units with three or fewer girders. Routine steel I-girder bridges,
as defined for the purposes of this Guide, are assumed to contain four or more girders in the cross-
section and feature span lengths not exceeding 200 feet, such that the use of lateral bracing is
generally not required; therefore, the provisions of this Article are not applicable.
For further information on lateral bracing, interested readers are encouraged to consult Section
6.3.2.10 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures.

6.7.5.2 I-Section Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies design requirements for lateral bracing (see the Discussion of Article 6.7.5.1
in this Guide) in bridges containing steel I-section members. Routine steel I-girder bridges, as
defined for the purposes of this Guide, typically do not need, and are assumed not to contain, lateral
bracing during construction or in the final condition as these bridges are straight and feature spans
not exceeding 200 feet; therefore, the provisions of this Article are not applicable.

6.7.5.3 Tub Section Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 155


Discussion:
This Article specifies design requirements for lateral bracing (see the Discussion of Article 6.7.5.1
in this Guide) in bridges containing composite steel tub-section members. The routine steel-girder
bridges covered by this Guide are not comprised of composite steel tub-section members;
therefore, the provisions of this Article are not applicable.

6.7.5.4 Trusses
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies design requirements for lateral bracing (see the Discussion of Article 6.7.5.1
in this Guide) in truss bridges. The routine steel-girder bridges covered by this Guide do not utilize
trusses; therefore, the provisions of this Article are not applicable.

6.7.6 Pins

6.7.6.1 Location
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides a recommendation that pins should be located to minimize the force effects
due to eccentricity. Pins are found on older existing bridges and are not used on modern steel girder
bridges. For example, older multi-span steel I-girder bridge designs sometimes used pin-and-
hanger or similar details to create hinges and impose statically determinate behavior in the
superstructure to simplify analysis of the bridge. Advancements in analytical techniques and
software have long since made it easier to design a multi-span continuous superstructure, negating
the need to force the articulation of the structure to be statically determinate. Pin and hanger details
have proven to be problematic details subject to corrosion issues and lack of redundancy. The use
of pins in routine steel I-girder bridges offers no benefits and is strongly discouraged. Therefore,
the provisions in this Article are not applicable.

6.7.6.2 Resistance

6.7.6.2.1 Combined Flexure and Shear


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article cover the resistance of pins subject to combined flexure and shear.
Pins are found on older existing bridges and are not used on modern steel girder bridges. For
example, older multi-span steel I-girder bridge designs sometimes used pin-and-hanger or similar
details to create hinges and impose statically determinate behavior in the superstructure to simplify
analysis of the bridge. Advancements in analytical techniques and software have long since made
it easier to design a multi-span continuous superstructure, negating the need to force the

NSBA Guide to Navigating Routine Steel Bridge Design / 156


articulation of the structure to be statically determinate. Pin and hanger details have proven to be
problematic details subject to corrosion issues and lack of redundancy. Pins were also used in older
steel rocker bearing designs; these types of bearings have demonstrated adverse maintenance
characteristics and poor performance. The use of pins in routine steel I-girder bridges offers no
benefits and is strongly discouraged. Therefore, the provisions in this Article are not applicable.

6.7.6.2.2 Bearing
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article covers the computation of the factored bearing resistance on pins. Pins are found on
older existing bridges and are not used on modern steel girder bridges. For example, older multi-
span steel I-girder bridge designs sometimes used pin-and-hanger or similar details to create hinges
and impose statically determinate behavior in the superstructure to simplify analysis of the bridge.
Advancements in analytical techniques and software have long since made it easier to design a
multi-span continuous superstructure, negating the need to force the articulation of the structure to
be statically determinate. Pin and hanger details have proven to be problematic details subject to
corrosion issues and lack of redundancy. Pins were also used in older steel rocker bearing designs;
these types of bearings have demonstrated adverse maintenance characteristics and poor
performance. The use of pins in routine steel I-girder bridges offers no benefits and is strongly
discouraged. Therefore, the provisions in this Article are not applicable.

6.7.6.3 Minimize Size Pin for Eyebars


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the minimum diameter of the pin used with eyebars. Eyebars are found on
older existing bridges and are not used on modern steel girder bridges. Pins are found on older
existing bridges and are not used on modern steel girder bridges. For example, older multi-span
steel I-girder bridge designs sometimes used pin-and-hanger or similar details to create hinges and
impose statically determinate behavior in the superstructure to simplify analysis of the bridge.
Advancements in analytical techniques and software have long since made it easier to design a
multi-span continuous superstructure, negating the need to force the articulation of the structure to
be statically determinate. Pin and hanger details have proven to be problematic details subject to
corrosion issues and lack of redundancy. Pins were also used in older steel rocker bearing designs;
these types of bearings have demonstrated adverse maintenance characteristics and poor
performance. The use of pins in routine steel I-girder bridges offers no benefits and is strongly
discouraged. Therefore, the provisions in this Article are not applicable.

6.7.6.4 Pins and Pin Nuts


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 157


Discussion:
The Article specifies material requirements for pins and pin nuts and the types of nuts that may be
used to secure a full bearing of all parts connected upon the turned body of a pin. Pins are found
on older existing bridges and are not used on modern steel girder bridges. For example, older multi-
span steel I-girder bridge designs sometimes used pin-and-hanger or similar details to create hinges
and impose statically determinate behavior in the superstructure to simplify analysis of the bridge.
Advancements in analytical techniques and software have long since made it easier to design a
multi-span continuous superstructure, negating the need to force the articulation of the structure to
be statically determinate. Pin and hanger details have proven to be problematic details subject to
corrosion issues and lack of redundancy. Pins were also used in older steel rocker bearing designs;
these types of bearings have demonstrated adverse maintenance characteristics and poor
performance. The use of pins in routine steel I-girder bridges offers no benefits and is strongly
discouraged. Therefore, the provisions in this Article are not applicable.

6.7.7 Heat-Curved Rolled Beams and Welded Plate Girders

6.7.7.1 Scope
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the grades of structural steel in the ASTM A709/A709M Specification that
may be used for rolled beams and constant-depth welded I-section plate girders that are heat-
curved to obtain a horizontal curvature.
The routine steel I-girder bridges covered by this Guide are not horizontally curved; therefore, the
provisions in this Article are not applicable.

6.7.7.2 Geometric Limitations


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article refers to Article 11.4.12.2.2 of the AASHTO LRFD Bridge Construction Specifications
for the cross-sectional and radius limitations on heat-curving of rolled beams and constant-depth
welded I-section plate girders to obtain a horizontal curvature. The Engineer is to indicate on the
contract documents whether heat curving is permitted according to these limitations.
The routine steel I-girder bridges covered by this Guide are not horizontally curved; therefore, the
provisions in this Article are not applicable.

6.7.8 Bent Plates


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 158


Discussion:
Bent plate diaphragms are sometimes used on routine steel I-girder bridges. The minimum bending
radius and direction of final rolling of the plate are important design parameters to be specified in
the contract documents. This article directs the Engineer to Article 11.4.3.3 of the AASHTO LRFD
Bridge Construction Specifications for fabrication requirements related to cold or hot bending and
provides Commentary with additional information (the requirements for cold bending, which is
typically used, are also summarized in the Commentary for this Article in the AASHTO LRFD
BDS). The discussion of minimum bend radii is particularly helpful to designers when detailing
bent plate diaphragms.
In addition to these specifications, the Engineer should review the AASHTO/AWS D1.5/D1.5M
Bridge Welding Code requirements and any applicable Owner-agency specifications related to the
bending of plates.

6.8 TENSION MEMBERS

6.8.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article is applicable to the design of tension members (i.e., members subject to axial tension)
and the design of connection elements subject to tension (see the Discussion of Article 6.13.5.2 in
this Guide). For the routine steel I-girder bridges covered by this Guide, these provisions are
applicable to the design of cross-frame members subject to axial tension and to the design of flange
splice plates and cross-frame gusset plates subject to tension.
The factored tensile resistance of a member or connection element at the strength limit state is to
be taken as the lesser of the resistance based on yielding on the gross section or fracture on the net
section when there are holes present (see the Discussion of Article 6.8.2.1 in this Guide). This
Article lists some important considerations when calculating the factored tensile resistance, which
are summarized as follows:
• Only holes larger than standard holes for connectors such as bolts need to be deducted from
the gross section. Such holes would include pin holes, access holes, and perforations.
• When calculating the net area, all holes are to be deducted from the section. The correction
factor for staggered holes is to be considered when deducting the area of connector holes
(see the Discussion of Article 6.8.3 in this Guide).
• The reduction factor, U, specified in Article 6.8.2.2 for tension members and Article
6.13.5.2 for connection elements is to be considered to account for the effect of shear lag
in the determination of the net section fracture resistance (see the Discussion of Articles
6.8.2.2 and 6.13.5.2 in this Guide).
• The 85-percent maximum area efficiency factor for connection elements specified in
Article 6.13.5.2 must be considered to provide reserve capacity to account for limited

NSBA Guide to Navigating Routine Steel Bridge Design / 159


inelastic deformation capabilities due to the relatively small length of the connection
elements (see the Discussion of Article 6.13.5.2 in this Guide).
• Slenderness requirements specified in Article 6.8.4 must be satisfied (see the Discussion
of Article 6.8.4 in this Guide).
• Fatigue requirements specified in Article 6.6.1 must be satisfied (see the Discussion of
Article 6.6.1 in this Guide).
• Block shear requirements specified in Article 6.13.4 must be satisfied at end connections
(see the Discussion of Article 6.13.4 in this Guide).
For further information on the design of tension members, consult Section 6.6.3.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Section 8.4.1 of Design Example 2A of the FHWA’s Steel Bridge Design
Handbook includes an example cross-frame design.
In addition, although not currently specified in the AASHTO LRFD BDS, it is recommended
herein that in addition to the minimum design requirements specified in Article 6.7.4.1 (see the
Discussion of Article 6.7.4.1 in this Guide), cross-frames or diaphragms for the routine steel I-
girder bridges covered by this Guide also be designed to satisfy the stability bracing strength and
stiffness requirements specified in AISC Appendix 6 - Article 6.3.2a. Consult Section 6.6.3.2 of
the Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures, the FHWA’s Steel Bridge Design Handbook, Volume 13 –
Bracing System Design, and Coletti, D.A., and M.A. Grubb, “Practical Implementation of Stability
Bracing Strength and Stiffness Guidelines for Steel I-Girder Bridges,” Proceedings of the 2016
World Steel Bridge Symposium, April 14, 2016, Orlando, FL for further information on these
requirements and their application to cross-frames and diaphragms in steel I-girder bridges.

6.8.2 Tensile Resistance

6.8.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article define the factored tensile resistance, Pr, of a tension member or a
connection element subject to tension at the strength limit state as the lesser of the tensile resistance
for yielding on the gross section or fracture on the net section (when there are holes present) given
by Eqs. 6.8.2.1-1 and 6.8.2.1-2, respectively. For the routine steel I-girder bridges covered by this
Guide, these provisions are applicable to the design of cross-frame members subject to axial
tension and to the design of flange splice plates and cross-frame gusset plates subject to tension.
The Commentary for this article discusses application of the shear lag reduction factor, U, (see the
Discussion of Article 6.8.2.2 in this Guide) and the bolt hole reduction factor, Rp, that accounts for
the reduced fracture resistance in the vicinity of bolt holes punched full size. Article 6.6.1.2.3
specifies that unless information is available to the contrary, bolt holes in bracing members and
their connection plates are to be assumed for design to be punched full size (see the Discussion of

NSBA Guide to Navigating Routine Steel Bridge Design / 160


Article 6.6.1.2.3 in this Guide). Bracing member connections are often punched full size, whereas
bolt holes in splice connections are typically drilled full size; the Engineer is encouraged to consult
with the Fabricator regarding the fabrication of the bolt holes. The factors U and Rp only apply
when computing the net section fracture resistance at the strength limit state.
The Commentary also contains a narrative on the rational of using both the tensile resistance for
yielding on the gross section and fracture on the net section to bound the expected behavior of the
steel to provide both reliable strength and behavior and limit excessive deformations.
For further information on the design of tension members, consult Section 6.6.3.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Section 8.4.1 of Design Example 2A of the FHWA’s Steel Bridge Design
Handbook includes an example cross-frame design.
In addition, although not currently specified in the AASHTO LRFD BDS, it is recommended
herein that in addition to the minimum design requirements specified in Article 6.7.4.1 (see the
Discussion of Article 6.7.4.1 in this Guide), cross-frames or diaphragms for the routine steel I-
girder bridges covered by this Guide also be designed to satisfy the stability bracing strength and
stiffness requirements specified in AISC Appendix 6 - Article 6.3.2a. Consult Section 6.6.3.2 of
the Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures, the FHWA’s Steel Bridge Design Handbook, Volume 13 –
Bracing System Design, and Coletti, D.A., and M.A. Grubb, “Practical Implementation of Stability
Bracing Strength and Stiffness Guidelines for Steel I-Girder Bridges,” Proceedings of the 2016
World Steel Bridge Symposium, April 14, 2016, Orlando, FL for further information on these
requirements and their application to cross-frames and diaphragms in steel I-girder bridges.

6.8.2.2 Reduction Factor, U


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article deal with the computation of the reduction factor, U, to account for
shear lag effects associated with end connection geometry when computing the net section fracture
resistance of a tension member or connection element subject to tension at the strength limit state
(see the Discussion of Articles 6.8.2.1 and 6.13.5.2 in this Guide).
Shear lag is a consideration when the tensile force in the member or element is applied
eccentrically or transmitted to some, but not all, of the connection elements; e.g., an angle having
a connection to only one leg or when the connection elements do not lie in a common plane. In
such cases, the tensile force is not uniformly distributed over the net area and the critical net section
may not be fully effective.
In lieu of a refined analysis, which is not recommended for the design of the routine steel I-girder
bridges covered by this Guide, the case figures provided in Table 6.8.2.2-1 may be used to
determine
_
the reduction factor, U, to be applied for a given scenario. Example calculations of the
terms, x and L, are given in Figure C6.8.2.2-1. In some cases, a single prescribed value of the

NSBA Guide to Navigating Routine Steel Bridge Design / 161


reduction factor, U, is listed, while in other cases, an equation for the calculation of the value of
the reduction factor, U, is provided.
Table 6.8.2.2-1 should be read very carefully when choosing the correct value of the reduction
factor, U, as the “Description of Element” can be confusing. Readers are also directed to the related
provisions and commentary in AISC’s Specifications for Structural Steel Buildings and
Commentary (specifically Article D3, Table D3.1, and the Commentary for Article D3, which has
helpful discussion and figures). Several scenarios commonly found in routine steel I-girder bridges
are listed below, with their corresponding Case number in Table 6.8.2.2-1.
• Design of bolted flange splice plates: Case 1 – The connection elements essentially lie in
a common plane and are fully connected (see the Discussion of Article 6.13.5.2 in this
Guide).
• Design of cross-frame gusset plates bolted to cross-frame connection plates (stiffeners):
Case 1 - The connection elements essentially lie in a common plane and are fully
connected.
• Design of single angle, double angle, or tee (WT) members attached to a gusset plate using
welds (a common and recommended detail for truss-type cross-frames in routine steel I-
girder bridges is to use a combination of longitudinal welds along the length of the
connection and a transverse weld across the end of the connection – see page 108 of
AASHTO-NSBA Steel Bridge Collaboration’s Guideline G1.4-2006 Guidelines for Design
Details and Figure 2.2.6.1 of AASHTO-NSBA Steel Bridge Collaboration’s Guideline
G12.1-2020 Guidelines to Design for Constructability and Fabrication): Case 2 – The
connection occurs only through one flange of the angle or only through the flanges of the
_
tee. See Figure C6.8.2.2-1 for the determination of x and L. For longitudinal welds with
unequal lengths, the average length of the longitudinal welds is to be used for L.
• Design of single angle, double angle, or tee (WT) members attached to a gusset plate or to
a cross-frame connection plate (stiffener) using a bolted connection: Case 2, or alternately
_
Case 7 or 8, as appropriate – See Figure C6.8.2.2-1 for the determination of x and L.

• Design of wide-flange or channel shapes attached to a gusset plate or to a cross-frame


connection plate (stiffener) using a bolted connection through their webs: Case 2, or
alternately Case 7 for wide-flange shapes only – See Figure C6.8.2.2-1 for the
_
determination of x and L.

• Design of gusset plates attached to cross-frame connection plates (stiffeners) using only
longitudinal welds along the length of the connection (with no transverse weld across the
end of the connection): Case 4 –For longitudinal welds with unequal lengths, the average
length of the longitudinal welds is to be used for L; the length of each weld is not to be less
_
than four times the weld size. See Figure C6.8.2.2-1 for the determination of x and L.

NSBA Guide to Navigating Routine Steel Bridge Design / 162


As noted in the Article, for open-section members, e.g., angles, tees, wide-flange shapes, and
channels, the calculated value of U from the table is not to be taken less than the ratio of the gross
area of the connected element or elements to the member gross area; for the typical cases of single
angles, double angles, or tee (WT) sections welded to gusset plates or bolted directly to the cross-
frame connection plate (stiffener), the “connected element” is taken as the angle, double angle, or
tee member (as applicable) and the “member” is taken as the gusset plate or connection plate (as
applicable). It is conservative to neglect this check of the ratio of the areas of the connected element
and the member, as they only define a lower bound to the value of the shear reduction factor, U;
using a lower value, calculated using Table 6.8.2.2-1, would be conservative.
The Commentary for this Article states the moment resulting from the eccentricity between the
member and the connection plate need not be considered in the design of the member or connection
plates for angle members and light structural tee members loaded eccentrically in axial tension;
the effect of the connection eccentricity is addressed through the use of the shear lag reduction
factor, U.
For further information on the shear lag reduction factor, U, consult Section 6.6.3.3.2.4 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. Section 8.4.1 of Design Example 2A of the FHWA’s Steel
Bridge Design Handbook includes an example cross-frame design.

6.8.2.3 Combined Axial Tension, Flexure, and Flexural and/or Torsional Shear

6.8.2.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article address the strength interaction for any combination of axial tension,
uniaxial or biaxial flexure, and flexural and/or torsional shear at the strength limit state, including
combinations where one or more of the individual actions may be zero. Eqs. 6.8.2.3.1-1 and
6.8.2.3.1-2 represent the stability and overall strength interaction effects for uniaxial or biaxial
bending combined with axial tension and general yielding under axial tension and flexure (with
the relationship represented by these equations shown in Figure C6.8.2.3.1-1). The alternative Eqs.
6.8.2.3.1-3 and 6.8.2.3.1-4 conservatively recognize that axial tension tends to have a negligible
to beneficial impact on the flexural resistances associated with compression buckling (with the
relationship represented by these equations shown in Figure C6.8.2.3.1-2). The Commentary for
this Article addresses the overall length effects associated with the lateral-torsional buckling
resistance when the flexural resistance about the x-axis is influenced by lateral-torsional buckling
and how these effects should be considered in combination with other cross-section based
resistance checks within these relationships.
This Article is not applicable to the design of I-sections used as the main spanning elements in a
routine steel I-girder bridge. Such members in routine steel I-girder bridges are not tension
members. The interaction of flange flexural shear stresses with the axial and flexural resistances
of the member is assumed to be negligible in I-section members in the AASHTO LRFD BDS. The
interaction between torsional and/or flexural shear stresses in I-section member flanges with other

NSBA Guide to Navigating Routine Steel Bridge Design / 163


members resistances is also neglected. Moment-shear strength interaction in the presence of low
(or zero) levels of axial force is also small and may be neglected in I-section members. The tensile
forces occurring in the flanges when such members are subject to major axis bending moments are
already addressed in Article 6.10 and its associated sub-Articles.
This Article is also not applicable to the design of truss-type cross-frame members subject to axial
tension in routine steel I-girder bridges. The moment resulting from the eccentricity between the
member and the connection plate need not be considered in the design of the member or connection
plates for angle members and light structural tee (WT) members loaded eccentrically in axial
tension; the effect of the connection eccentricity is addressed through the use of the shear lag
reduction factor, U (see the Discussion of Article 6.8.2.2 in this Guide).

6.8.2.3.2 Interaction with Torsional and/or Flexural Shear


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article address torsional and/or flexural shear interaction with axial tension
at the strength limit state. These provisions are applicable only for noncomposite rectangular box-
section members, including square and rectangular HSS (Hollow Structural Sections), and
noncomposite circular tubes, including round HSS, subject to fairly significant levels of torsion
(i.e., fve/ϕTFcv greater than 0.2, where the terms are defined in Article 6.9.2.2.2) in combination
with flexure and/or axial tensile force, and/or to I-section members and the preceding members
subject to flexural shear in combination with significant levels of axial tensile force (i.e., Pu/Pry
greater than 0.05, where the terms are defined in Article 6.8.2.3.1). Such members are not used in
the design of the routine steel I-girder bridges covered by this Guide. Cross-frame members in
routine steel I-girder bridges are subject to significant axial force, but generally are not subject to
any significant torsional or flexural shear effects. Therefore, the provisions of this Article are not
applicable to the routine steel I-girder bridges covered by this Guide.

6.8.2.3.3 Tension Rupture Under Axial Tension or Compression Combined with Flexure
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article address the strength interaction between flexure and axial tension or
compression pertaining to tension rupture at certain specified locations. These locations include:
1) cross-sections containing bolt holes in one or more flanges subject to tension under combined
axial tension or compression and flexure at connection or nonconnection locations; 2) cross-
sections at connection or nonconnection locations subject to combined axial tension and flexure
and containing bolt holes in other cross-section elements; and 3) cross-sections at welded
connections subject to combined axial tension and flexure. The equation in this Article focuses on
the specific axial force, tension or compression, combined with the specific moment at the cross
section under consideration. Axial compressive forces on the cross-section containing the bolt
holes result in a negative force ratio and produce a beneficial subtractive effect, whereas axial
tensile forces result in a positive force ratio and produce an additive effect.

NSBA Guide to Navigating Routine Steel Bridge Design / 164


In terms of cross-frame members subject to axial tension in the routine steel I-girder bridges
covered by this Guide, the moment resulting from the eccentricity between the member and the
connection plate need not be considered in the design of the member or connection plates for angle
members and light structural tee members loaded eccentrically in axial tension; the effect of the
connection eccentricity is addressed through the use of the shear lag reduction factor, U (see the
Discussion of Article 6.8.2.2 in this Guide).
These provisions may be applicable to tee-section or double-angle cross-frame members if they
are subject to tension under combined axial compression and flexure. It should be noted that using
rolled steel tee (WT) and double-angle sections as cross-frame members for routine steel I-girder
bridges is generally discouraged, while the use of single-angle sections is encouraged. The
magnitude of the forces in cross-frame members in routine steel I-girder bridges covered by this
Guide is generally small enough such that single-angle members are adequate. Cross-frame forces
are typically only large enough to warrant the use of rolled steel tee (WT) or double-angle sections
in curved and/or skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite
expensive to fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally
require straightening after the cutting process, which adds significant fabrication effort and cost.
Double-angle sections are often viewed as problematic from a maintenance perspective; the
surfaces between the adjacent angle flanges are difficult or impossible to paint in the field, and/or
can suffer from potentially severe pack rust.

6.8.3 Net Area


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article defines the net area, An, of a member or connected element when holes are present.
The net area of each element in a given cross-section is defined as the product of the thickness of
the member or connected element and its smallest net width. The overall net area of a given cross-
section, An, is used in the computation of the net section fracture resistance and the block shear
rupture resistance when the member or connection element is subject to tension (see the Discussion
of Articles 6.8.2.1, 6.13.4, and 6.13.5.2 in this Guide). The net width of each element in a given
cross-section is based on its gross width, reduced to account for the presence of holes in the
element. For the routine steel I-girder bridges covered by this Guide, the provisions of this Article
are applicable to the design of cross-frame members subject to axial tension and to the design of
flange splice plates and cross-frame gusset plates subject to tension.
The width of each standard bolt hole is to be taken as the nominal diameter of the hole. The width
of oversize and slotted holes, although not recommended for use in the routine steel I-girder
bridges covered by this Guide, is taken equal to the nominal diameter or width of the hole, as
applicable. Maximum hole sizes for different bolt diameters are specified in Table 6.13.2.4.2-1
(see the Discussion of Article 6.13.2.4.2 in this Guide). For example, for a 7/8-inch diameter bolt,
the maximum hole size of a standard hole equal to 15/16 inches is used in the calculation of An.
The net width is to be determined for each chain of holes extending across the connected element
along any transverse, diagonal or zigzag line. For each chain, the net width is to be determined by

NSBA Guide to Navigating Routine Steel Bridge Design / 165


subtracting the sum of all holes in the chain from the total width, and then adding back in the
quantity s2/4g for each space between consecutive holes in the chain when the holes are staggered.
The term s is the pitch of any two consecutive holes (i.e., the distance between the center of the
two holes along the line of force) and g is the gage of the same two holes (i.e., the distance between
the adjacent lines of bolts containing the two holes in the direction perpendicular to the line of
force). When holes are staggered in both legs of an angle, the gage for holes in opposite adjacent
legs is to be taken as the sum of the gages from the back of the angles less the thickness of the
angle. For welded connections, An is to be taken as the gross area less any access holes within the
connection region.
It is conservative to use the least net width in conjunction with the full tensile force to check the
member or connected element. Assuming each bolt transfers an equal share of the load whenever
the bolts are arranged symmetrically with respect to the centroidal axis of the member or connected
element, a less conservative alternative, particularly with staggered holes, is to check each possible
chain with a tensile force obtained by subtracting the force removed by each bolt ahead of that
chain from the full tensile force.
For example calculations of the net area of a tension member, consult Sections 6.6.3.3.2.3 and
6.6.4.2.5.6.1 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures.

6.8.4 Limiting Slenderness Ratio for Tension Members


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article specify limiting values for evaluating the slenderness ratio, or ratio
of the unbraced length over the radius of gyration, of primary and secondary members subject to
axial tension, or for evaluating the tension slenderness of primary and secondary compression
members subject to stress reversal.
The provisions are not applicable to the tension flanges of flexural members such as the I-shaped
sections typically used as the main girders in routine steel I-girder bridges. However, these
provisions are directly applicable to the proportioning of cross-frame members in routine steel I-
girder bridges.
Determination of whether the cross-frame members should be treated as “Primary” or “Secondary”
members should be based on a careful review of the definitions provided in Article 6.2, where it
states that a “Primary Member” is “A steel member or component that transmits gravity loads
through a necessary as-designed load path. These members are therefore subjected to more
stringent fabrication and testing requirements; considered synonymous with the term ‘main
member.’” In routine steel I-girder bridges, the cross-frame members do not meet this definition.
Recall that the definition of a routine steel I-girder bridge is limited to straight bridges with little
or no support skew. In such bridges, restoring forces are created in the cross-frames due to any
differential deflections between the girders but these restoring forces will typically be small since
the differential deflections between the girders in routine steel I-girder bridges are small. Thus, the
girders will resist the noncomposite dead loads equally in these bridges, neglecting any effect of

NSBA Guide to Navigating Routine Steel Bridge Design / 166


the deflections resulting from elastic shortening of the cross-frames, which are generally
negligible. The distribution of composite dead loads and live loads in these types of bridges occurs
primarily through the deck. As a result, the cross-frame members in a routine steel I-girder bridge
can be categorized as “Secondary Members” when addressing the provisions of this Article.
As an aside, the cross-frame members in curved and/or significantly skewed steel I-girder bridges
do in fact function to transmit gravity loads through a necessary as-designed load path due to the
significantly larger differential deflections between the girders in these bridges, and as a result the
cross-frames in those types of bridges are categorized as “Primary Members” when addressing the
provisions of this Article. While cross-frame members in straight I-girder bridges with significant
support skew brace the girders and serve as an additional transverse load path in the bridge system,
the internal cross-frame forces in these bridges are not required for equilibrium of the girders;
hence, the cross-frames in these bridges should be categorized as “Secondary Members.”
These provisions are applicable for evaluating the tension slenderness ratio of cross-frame
members in the routine steel I-girder bridges covered by this Guide; only the limiting value of the
ratio for secondary members applies.

6.8.5 Built-Up Members

6.8.5.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article contain general requirements for tension members built-up from
rolled or welded shapes connected by continuous plates with or without perforations or by tie
plates with or without lacing. Provisions are specified in this Article for the welded or bolted
connections between the plates and shapes.
Built-up tension members are not typically used in the routine steel I-girder bridges covered by
this Guide, except possibly the application of double angle members in cross-frames; the
provisions of this Article are only applicable in that case. It should be noted that using rolled steel
tee (WT) and double-angle sections as cross-frame members for routine steel I-girder bridges is
generally discouraged, while the use of single-angle sections is encouraged. The magnitude of the
forces in cross-frame members in routine steel I-girder bridges covered by this Guide is generally
small enough such that single-angle members are adequate. Cross-frame forces are typically only
large enough to warrant the use of rolled steel tee (WT) or double-angle sections in curved and/or
skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite expensive to
fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally require
straightening after the cutting process, which adds significant fabrication effort and cost. Double-
angle sections are often viewed as problematic from a maintenance perspective; the surfaces
between the adjacent angle flanges are difficult or impossible to paint in the field, and/or can suffer
from potentially severe pack rust.
For further information on built-up tension members, consult Section 6.6.3.3.4 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge

NSBA Guide to Navigating Routine Steel Bridge Design / 167


Superstructures. Design Example 2A of the FHWA’s Steel Bridge Design Handbook includes an
example cross-frame design.

6.8.5.2 Perforated Plates


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains provisions specific to the design of perforated plates used to connect rolled
or welded shapes in built-up tension members (see the Discussion of Article 6.8.5.1 in this Guide).
The routine steel I-girder bridges covered by this Guide do not utilize built-up tension members
with perforated plates; therefore, the provisions of this Article are not applicable.

6.8.6 Eyebars

6.8.6.1 Factored Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to compute the factored tensile resistance of the body of an
eyebar at the strength limit state. The resistance is based on yielding on the gross section; fracture
on the net section does not control because the net section in the head is at least 1.35 times greater
than the section in the body. The routine steel I-girder bridges covered by this Guide do not contain
eyebars; therefore, the provisions of this Article are not applicable.

6.8.6.2 Proportions
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article specify the proportioning requirements for the head and body of an
eyebar and the location and dimensioning requirements of the pin hole in the eyebar. The routine
steel I-girder bridges covered by this Guide do not contain eyebars; therefore, the provisions of
this Article are not applicable.

6.8.6.3 Packing
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article specify the detailing requirements for an eyebar assembly to prevent
corrosion-causing elements from entering the joint, lateral movement on the pin and lateral
distortion of the eyebar due to skew, and repeated eyebar contact due to vibration perpendicular to
the eyebar plane. The routine steel I-girder bridges covered by this Guide do not contain eyebars;
therefore, the provisions of this Article are not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 168


6.8.7 Pin-Connected Plates

6.8.7.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article apply to the design of pin-connected plates, which should be avoided
wherever possible. The factored tensile resistance of such plates at the strength limit state must
satisfy the provisions of Article 6.8.2.1 (see the Discussion of Article 6.8.2.1 in this Guide). The
routine steel I-girder bridges covered by this Guide do not contain pin-connected plates; therefore,
the provisions of this Article are not applicable.

6.8.7.2 Pin Plates


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article specify the factored resistance for bearing on pin-connected plates,
Pr, at the strength limit state and also provisions related to the strengthening of the main plate in
the region of the hole and attachment of the pin plates to the main plate. The routine steel I-girder
bridges covered by this Guide do not contain pin-connected plates; therefore, the provisions of this
Article are not applicable.

6.8.7.3 Proportions
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article specify the proportioning requirements for the main plate and pin
plates in a pin-connected assembly and the location and dimensioning requirements of the pin hole
in the plates. The routine steel I-girder bridges covered by this Guide do not contain pin-connected
plates; therefore, the provisions of this Article are not applicable.

6.8.7.4 Packing
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article specify the detailing requirements for a pin-connected assembly to
prevent corrosion-causing elements from entering the joints and lateral movement on the pin and
lateral distortion of the assembly due to skew. The routine steel I-girder bridges covered by this
Guide do not contain pin-connected plates; therefore, the provisions of this Article are not
applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 169


6.9 COMPRESSION MEMBERS

6.9.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article is applicable to the design of compression members (i.e., members subject to axial
compression). For the routine steel I-girder bridges covered by this Guide, these provisions are
applicable to the design of cross-frame members subject to axial compression.
Neither I-girders nor their flanges in routine steel I-girder bridges are treated as compression
members or beam columns per se. The combined effects of axial compression and flange lateral
bending in the compression flanges of girders and beams are directly addressed in Article 6.10.
Furthermore, the term “composite steel members” in this Article is not meant to imply that the
provisions of this Article apply to composite girders and beams. Instead this term is referring to
concrete-filled tubes or pipes or concrete encased steel members subject to axial compression or
to combined axial compression and flexure (see the Discussion of Articles 6.9.5 and 6.9.6 in this
Guide).
However, cross-frames in routine steel I-girder bridges often feature members subject to axial
compression or combined axial compression and flexure. As such, the provisions of Articles 6.9.1,
6.9.2, 6.9.3, and 6.9.4 are directly applicable to their design.
The language in the Commentary of this Article is short, but significant, and should be read
carefully. Examples of “significant additional eccentricity” include items such as the offset from
the centroid of a cross-frame member to the faying surface between its connection to a cross-frame
connection plate; an eccentricity such as this induces moment that should be considered in the
design of that member. However, as clearly stated in the Commentary, “imperfections and
eccentricities permissible in normal fabrication and erection” are already accounted for in
conventional steel column design formulas and do not need to be treated as sources of additional
eccentricity-induced moments.
For further information on the design of compression members, consult Section 6.6.3.4 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. Design Example 2A of the FHWA’s Steel Bridge Design
Handbook includes an example cross-frame design.

6.9.2 Compressive Resistance

6.9.2.1 Axial Compression


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
Neither I-girders nor their flanges in routine steel I-girder bridges are treated as compression
members or beam columns per se. The combined effects of axial compression and flange lateral

NSBA Guide to Navigating Routine Steel Bridge Design / 170


bending in the compression flanges of girders and beams are directly addressed in Article 6.10,
and the provisions of Article 6.9.2.1 are therefore not applicable to the design of I-sections in the
routine steel I-girder bridges covered by this Guide. However, cross-frames in routine steel I-girder
bridges often feature members subject to axial compression or combined axial compression and
flexure. As such, the provisions of this Article are directly applicable to their design.
This Article specifically outlines the basic load versus resistance equality that must be satisfied.
The definition of the nominal compressive resistance, Pn, refers to Articles 6.9.4 and 6.9.5, but
only Article 6.9.4 is applicable. Article 6.9.4 addresses the design of noncomposite members, such
as those used in typical cross-frames in routine steel I-girder bridges. Article 6.9.5 addresses
composite members; as explained in the Discussion of Article 6.9.1 in this Guide, in this context
the term “composite members” refers to concrete-filled tube or pipes or concrete encased steel
members subject to axial compression or combined axial compression and flexure (see the
Discussion of Articles 6.9.5 and 6.9.6 in this Guide).

6.9.2.2 Combined Axial Compression, Flexure, and Flexural and/or Torsional Shear

6.9.2.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article address the strength interaction for any combination of axial
compression, uniaxial or biaxial flexure, and flexural and/or torsional shear at the strength limit
state, including combinations where one or more of the individual actions may be zero. This Article
specifically presents the beam-column interaction equations which must be satisfied for certain
types of members subject to combined axial compression and flexure.
In terms of cross-frame members subject to axial compression in the routine steel I-girder bridges
covered by this Guide, consideration of the moment resulting from the eccentricity between the
member and the connection plate is dependent on the shape of the member’s cross section.
• Rolled steel single-angle members subject to combined axial compression and flexure
about one or both principal axes and meeting a short list of simple connection and loading
criteria may be designed as axially loaded compression members for flexural buckling
only, as outlined in Article 6.9.4.4 (see the Discussion of Article 6.9.4.4 in this Guide). As
a result, the provisions of Article 6.9.2.2 and the associated sub-articles (including this
Article, 6.9.2.2.1) are not applicable for the design of these members.
• Rolled steel tee (WT) and double-angle members, on the other hand, subject to combined
axial compression and flexure must be designed in accordance with the provisions of this
Article.
Channel sections are sometimes used as a top chord member in end cross-frames used to support
the edge of the deck at an expansion joint. Such members are typically connected to the deck using
shear connectors. The members in end cross-frames in routine steel I-girder bridges are typically
not subject to significant axial loads associated with their function as stability bracing for the

NSBA Guide to Navigating Routine Steel Bridge Design / 171


girders. The main sources of loading in end cross-frames in routine steel I-girder bridges are wind
(wind load on superstructure being transferred to the bearings through pier and end cross-frames)
and gravity loads from the deck (primarily the wheel loads applied at the edge of deck plus a small
amount of deck dead load). The wind loads are transferred to the bearings through the end cross-
frame diagonals, which are typically single-angle, double-angle, or tee (WT) members. Wheel
loads and dead load from the deck typically produce bending in the top chords and axial
compression in the diagonals.
Furthermore, neither I-girders nor their flanges in routine steel I-girder bridges are treated as
compression members or beam columns per se. The combined effects of axial compression and
flange lateral bending in the compression flanges of girders and beams are directly addressed in
Article 6.10.
Consequently, the remainder of this Discussion is exclusively focused on the application of the
provisions of this Article to the design of rolled steel tee (WT) and double-angle members subject
to axial compression and uniaxial or biaxial flexure.
It should be noted that using rolled steel tee (WT) and double-angle sections as cross-frame
members for routine steel I-girder bridges is generally discouraged, while the use of single-angle
sections is encouraged. The magnitude of the forces in cross-frame members in routine steel I-
girder bridges covered by this Guide is generally small enough such that single-angle members are
adequate. Cross-frame forces are typically only large enough to warrant the use of rolled steel tee
(WT) or double-angle sections in curved and/or skewed steel I-girder bridges. Rolled steel tee
(WT) sections are typically quite expensive to fabricate. Tee (WT) sections are cut from full wide-
flange (W) shapes and generally require straightening after the cutting process, which adds
significant fabrication effort and cost. Double-angle sections are often viewed as problematic from
a maintenance perspective; the surfaces between the adjacent angle flanges are difficult or
impossible to paint in the field, and/or can suffer from potentially severe pack rust.
Eqs. 6.9.2.2.1-1 and 6.9.2.2.1-2 represent the stability and overall strength interaction effects for
uniaxial or biaxial bending combined with axial compression and general yielding under axial
compression and flexure. As explained in the Commentary, these equations provide an “accurate
to conservative approximation of the resistances under combined loading for members in which
all the cross-section elements are compact” for flexure. The Commentary further explains that Eqs.
6.9.2.2.1-1 and 6.9.2.2-2 are applicable to cases where the axial and flexural stresses in the flange
of the tee or the flange legs of the double angles are in compression, such as when the connection
is made through the flange of the tee or the flange legs of the angles, which is the typical case
when tee and double-angle sections are used as cross-frame members in routine steel I-girder
bridges.
Eq. 6.9.2.2.1-3 is only applicable in cases where the toe of the stem or leg of the double-angle
section is in flexural compression; in this case the stem or leg is not considered a compact element.
It should be emphasized that this equation is only applicable when the toe of the stem or leg is in
compression when subject to bending moment only. If the toe of the stem or leg is in flexural
tension (i.e., in tension when subject to bending moment only), even if the toe is in compression
when the combined effects of axial compression and flexure are considered, then Eqs. 6.9.2.2.1-1

NSBA Guide to Navigating Routine Steel Bridge Design / 172


and 6.9.2.2.1-2 apply. Such a condition will rarely, if ever, occur in the cross-frame of a steel I-
girder bridge since in virtually all cases the connection is made through the flanges.
These interaction equations are very straightforward in and of themselves and are easy to
implement in hand calculations or design spreadsheets. The Commentary explains that the
equations presented in this Article are conservative simplifications of more exact, nonlinear
equations. These exact, nonlinear equations are rarely, if ever, used in practice. Using them for the
design of cross-frames in routine steel I-girder bridges is neither warranted nor recommended; the
simplified equations presented in the Article are more than sufficient for cross-frame design
purposes.
The Article goes on to require that the factored design moments, Mux and Muy, be determined in a
manner which accounts for the second-order magnification of moments due to geometric nonlinear
behavior. The Article allows that the effects of second-order moment magnification can be
determined either by means of a second-order elastic analysis (i.e., an iterative geometric nonlinear
analysis accounting for P- effects) or by the approximate single-step method specified in Article
4.5.3.2.2b or a comparable amplification factor-based procedure. For the design of cross-frame
members in routine steel I-girder bridges, the use of the refined second-order analysis approach is
not justified. The approximate single-step method specified in Article 4.5.3.2.2b is more than
sufficient and should be used (see the Discussion of Article 4.5.3.3.2b in this Guide).
Eq. 6.9.2.2.1-5 addresses the case where the nominal flexural resistance about the major axis of
the section is determined according to the provisions of Appendix A6. Appendix A6 addresses
optional provisions for determining the flexural capacity of composite and noncomposite I-shaped
members; Appendix A6 is intended for use in designing the main girders in a routine steel I-girder
bridge, not the design of cross-frame members in routine steel I-girder bridges, so this equation is
not applicable (see the Discussion of Appendix A6 in this Guide).
The Article further requires that interaction with torsional and/or flexural shear be considered and
directs the reader to Article 6.9.2.2.2. This language applies to the design of noncomposite box-
section members, which can be subject to significant torsional and/or flexural shear effects in their
typical applications (e.g., as arch rib members, straddle bent caps, and so forth). Cross-frame
members in routine steel I-girder bridges are not subject to any significant torsional or flexural
shear effects; as such this directive is not applicable, and as can be seen elsewhere in this Guide,
Article 6.9.2.2.2 in its entirety is considered not applicable to the design of routine steel I-girder
bridges.

6.9.2.2.2 Interaction with Torsional and/or Flexural Shear


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article is intended to apply exclusively to the design of noncomposite rectangular box-section
members, including square and rectangular hollow structural sections (HSS), and noncomposite
circular tubes, including round HSS, or I-section members subject to significant axial force
combined with torsional and/or flexural shear. The levels of torsional and/or flexural shear above
which their effects much be considered are defined in this Article.

NSBA Guide to Navigating Routine Steel Bridge Design / 173


The interaction of flange flexural shear stresses with the axial and flexural resistances of the
member is assumed to be negligible in I-section members in the AASHTO LRFD BDS. The
interaction between torsional and/or flexural shear stresses in I-section member flanges with other
members resistances is also neglected. Moment-shear strength interaction in the presence of low
(or zero) levels of axial force is also small and may be neglected in I-section members.
Cross-frame members in routine steel I-girder bridges are subject to significant axial force, but
generally are not subject to any significant torsional or flexural shear effects.
As a result, the provisions of this article are not applicable to the design of the routine steel I-girder
bridges covered by this Guide.

6.9.3 Limiting Slenderness Ratio for Compression Members


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The proportioning limits specified in this article are intended to limit the slenderness of members
subject to axial compression only, or to evaluate the compression slenderness of tension members
subject to stress reversal, so that these members can exhibit a reasonable minimum resistance to
buckling.
The provisions are not applicable to the compression flanges of flexural members such as the I-
shaped sections typically used as the main girders in routine steel I-girder bridges. However, these
provisions are directly applicable to the proportioning of cross-frame members in routine steel I-
girder bridges; only the limiting value of the ratio for secondary members applies.
Determination of whether the cross-frame members should be treated as “Primary” or “Secondary”
members should be based on a careful review of the definitions provided in Article 6.2, where it
states that a “Primary Member” is “A steel member or component that transmits gravity loads
through a necessary as-designed load path. These members are therefore subjected to more
stringent fabrication and testing requirements; considered synonymous with the term ‘main
member.’” In routine steel I-girder bridges, the cross-frame members do not meet this definition.
Recall that the definition of a routine steel I-girder bridge is limited to straight bridges with little
or no support skew. In such bridges, restoring forces are created in the cross-frames due to any
differential deflections between the girders but these restoring forces will typically be small since
the differential deflections between the girders are small. Thus, the girders will resist the
noncomposite dead loads equally in these bridges, neglecting any effect of the deflections resulting
from elastic shortening of the cross-frames which are generally negligible. The distribution of
composite dead loads and live loads in these types of bridges occurs primarily through the deck.
As a result, the cross-frame members in a routine steel I-girder bridge can be categorized as
“Secondary Members” when addressing the provisions of this Article.
As an aside, the cross-frame members in curved and/or significantly skewed steel I-girder bridges
do in fact function to transmit gravity loads through a necessary as-designed load path due to the
significantly larger differential deflections between the girders in these bridges, and as a result the
cross-frames in those types of bridges are categorized as “Primary Members” when addressing the

NSBA Guide to Navigating Routine Steel Bridge Design / 174


provisions of this Article. While cross-frame members in straight I-girder bridges with significant
support skew brace the girders and serve as an additional transverse load path in the bridge system,
the internal cross-frame forces in these bridges are not required for equilibrium of the girders;
hence, the cross-frames in these bridges should be categorized as “Secondary Members.”
The Article directs the reader to Article 4.6.2.5 for specification of appropriate effective length
factors, K. The choice of effective length factor is dependent on the type of section and its end
connections. For the single-angle members commonly used as cross-frame members in routine
steel I-girder bridges, regardless of end connection, the effective length factor, K, is specified as
1.0 (see the Discussion of Article 6.9.4.4 in this Guide for further information on the checking of
the slenderness ratio limits specified in this Article for single-angle members). For other sections,
such as rolled steel tee (WT) or double-angle sections which are sometimes used as cross-frame
members in routine steel I-girder bridges, if bolted or welded end connections are provided, the
effective length factor, K, is specified as 0.750. The case of “pinned connections” in Article 4.6.2.5
is not applicable for the types of members typically used in routine steel I-girder bridges.
The Article also offers an optional approach for determining the radius of gyration using a
“notional section.” This provision permits neglecting part of the cross-section of the compression
member to allow calculation of a more favorable (larger) radius of gyration, solely for the purposes
of satisfying the requirements of this Article. The use of this provision is typically neither
warranted nor recommended for the design of compression members in truss-type cross-frames
for routine steel I-girder bridges. Given the geometric constraints associated with the definition of
a routine steel I-girder bridge for the purposes of this Guide, it is highly unlikely that a prudent,
more economical design would result from using this provision. Instead, the cross-frame members
should be adequately sized to meet the basic slenderness requirements, which are not onerous for
typical cross-frame members in routine steel I-girder bridges.

6.9.4 Noncomposite Members

6.9.4.1 Nominal Compressive Resistance

6.9.4.1.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
Neither I-girders nor their flanges in routine steel I-girder bridges are treated as compression
members or beam columns per se. The combined effects of axial compression and flange lateral
bending in the compression flanges of girders and beams are directly addressed in Article 6.10,
and the provisions of this Article are therefore not applicable to the design of I-sections in the
routine steel I-girder bridges covered by this Guide. However, cross-frames in routine steel I-girder
bridges often feature members subject to axial load. As such, the provisions of this Article are
directly applicable to their design.
The provisions of this Article are comprehensive in terms of addressing the nominal compressive
resistance of axially loaded members; they address a variety of buckling modes and a variety of
cross-sections. The reader is cautioned to read the provisions carefully before using them.

NSBA Guide to Navigating Routine Steel Bridge Design / 175


The article specifies that three buckling modes be considered, and that the nominal compressive
resistance, Pn, is to be taken as the smallest value. Not all three modes are applicable for all cross-
sectional shapes. The three modes include:
• Flexural buckling
• Torsional buckling
• Flexural-torsional buckling
Three categories of cross-sectional shapes are considered. Some of these are not used as axially
loaded members in the design of routine steel I-girder bridges. The three categories of cross-
sectional shapes include:
• Doubly symmetric members: These are members which are symmetric about both of their
primary orthogonal axes. Examples of doubly symmetric members include rolled wide-
flange (W shapes) and other I-shaped members (H and S shapes and some plate girders)
with equal-size flanges and also closed sections such as round and hollow structural section
(HSS) steel tubes. As previously mentioned, I-shaped sections used as main girders in
routine steel I-girder bridges are not treated as compression members. Doubly symmetric
shapes are not used as cross-frame members in routine steel I-girder bridges. As a result,
the provisions in this Article related to doubly symmetric members are not applicable to
the design of the routine steel I-girder bridges covered by this Guide.
• Singly symmetric members: These are members which are symmetric about only one of
their primary orthogonal axes. Examples of singly symmetric members include I-shaped
steel plate girders with unequal-sized flanges, rolled tee (WT) sections, and double-angle
sections. As previously mentioned, I-shaped sections used as main girders in routine steel
I-girder bridges are not treated as compression members. However, rolled steel tee (WT)
and double-angle sections are sometimes used as cross-frame members in routine steel I-
girder bridges, and the provisions of this article are applicable to their design.
• Unsymmetric members: These are members which have no axis of symmetry. Examples
of unsymmetric members include single-angle sections. Single-angle sections are often
used as cross-frame members in routine steel I-girder bridges, and the provisions of this
article are applicable to their design.
The equations for nominal resistance, Pn, are grouped for applicability first based on whether the
compression member is comprised of nonslender or slender elements. The provisions for
determining the slenderness classification are found in Article 6.9.4.2.1 (see the Discussion of
Article 6.9.4.2.1 in this Guide). For compression members composed only of nonslender,
longitudinally unstiffened elements, the nominal resistance, Pn, is determined using either Eq.
6.9.4.1.1-1 or 6.9.4.1.1-2, depending on the ratio of the nominal yield resistance, Po, to the elastic
critical buckling resistance, Pe. For compression members with cross-sections containing one or
more slender elements, the nominal resistance, Pn, is determined in accordance with the provisions
of Article 6.9.4.2.2 (see the Discussion of Article 6.9.4.2.2 in this Guide).

NSBA Guide to Navigating Routine Steel Bridge Design / 176


Table 6.9.4.1.1-1 provides a handy guide for determining the nature of a given cross-sectional
shape, its potential column buckling modes, and the applicable equations, and associated Article
and Commentary, for determining the elastic critical buckling resistance, Pe.
Mention of “longitudinally stiffened elements” is not applicable to the design of the cross-frame
members used in routine steel I-girder bridges. Members with longitudinally stiffened elements,
in the context of this Article, are generally limited to noncomposite box section members, which
are not used in routine steel I-girder bridges.
It should be noted that using rolled steel tee (WT) and double-angle sections as cross-frame
members for routine steel I-girder bridges is generally discouraged, while the use of single-angle
sections is encouraged. The magnitude of the forces in cross-frame members in routine steel I-
girder bridges covered by this Guide is generally small enough such that single-angle members are
adequate. Cross-frame forces are typically only large enough to warrant the use of rolled steel tee
(WT) or double-angle sections in curved and/or skewed steel I-girder bridges. Rolled steel tee
(WT) sections are typically quite expensive to fabricate. Tee (WT) sections are cut from full wide-
flange (W) shapes and generally require straightening after the cutting process, which adds
significant fabrication effort and cost. Double-angle sections are often viewed as problematic from
a maintenance perspective; the surfaces between the adjacent angle flanges are difficult or
impossible to paint in the field, and/or can suffer from potentially severe pack rust.
For further information about column buckling theory and the nominal compressive resistance of
compression members, consult Section 6.6.3.4.2.3 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures. Design
Example 2A of the FHWA’s Steel Bridge Design Handbook includes an example cross-frame
design.

6.9.4.1.2 Elastic Flexural Buckling Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article presents the resistance equation for elastic flexural buckling of concentrically loaded
compression members. This is the classic Euler buckling equation and can be applied to a number
of members in steel structures.
Neither I-girders nor their flanges in routine steel I-girder bridges are treated as compression
members or beam columns per se. The combined effects of axial compression and flange lateral
bending in the compression flanges of girders and beams are directly addressed in Article 6.10, so
the provisions of this Article, 6.9.4.1.2, are not applicable to the design of I-sections in the routine
steel I-girder bridges covered by this Guide.
Instead, in routine steel I-girder bridges, elastic flexural buckling is considered primarily in the
design of cross-frame members. As explained in Article 6.9.4.1.1, its related Commentary, and its
related Discussion in this Guide, flexural buckling is an applicable buckling mode for the sections
typically used in cross-frames for routine steel I-girder bridges, which include single-angle
sections, tee (WT) sections, and double-angle sections.

NSBA Guide to Navigating Routine Steel Bridge Design / 177


As noted in the Commentary for this Article, Eq. 6.9.4.1.2-1 should be used to calculate the critical
flexural buckling resistance about both the x- and y-axes, and the smaller value should be taken as
Pe for use in Eq. 6.9.4.1.1-1 or 6.9.4.1.1-2, as applicable. For single-angle members, the effective
slenderness ratio, (K/r)eff, determined according to the provisions of Article 6.9.4.4, is used in
determining the elastic flexural buckling resistance (see the Discussion of Article 6.9.4.4 in this
Guide).
For further information on the elastic flexural buckling resistance, consult Section 6.6.3.4.2.3.3 of
the Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. Design Example 2A of the FHWA’s Steel Bridge Design
Handbook includes an example cross-frame design.

6.9.4.1.3 Elastic Torsional Buckling and Flexural-Torsional Buckling Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article presents resistance equations for elastic torsional buckling and flexural-torsional
buckling of concentrically loaded compression members.
Eq. 6.9.4.1.3-1 is the resistance equation for torsional buckling of open-section doubly symmetric
members, such as doubly symmetric I-shaped members. Neither I-girders nor their flanges in
routine steel I-girder bridges are treated as compression members or beam columns per se. The
combined effects of axial compression and flange lateral bending in the compression flanges of
girders and beams are directly addressed in Article 6.10, so Eq. 6.9.4.1.3-1 is generally not
applicable to the design of the routine steel I-girder bridges covered by this Guide.
Eqs. 6.9.4.1.3-2 through 6.9.4.1.3-6 address calculation of the flexural-torsional buckling
resistance of open-section singly symmetric members, such as rolled tee (WT), double-angle, and
channel members. These members are sometimes used in cross-frames or as diaphragms for
routine steel I-girder bridges and these equations must be checked for these members in such cases
when subject to axial compression.
Eqs. 6.9.4.1.3-7 through 6.9.4.1.3-9 address calculation of the flexural-torsional buckling
resistance of open-section unsymmetric members, such as single-angle members. These members
are often used in cross-frames for routine steel I-girder bridges. However, when the effective
slenderness ratio, (K/r)eff, determined according to the provisions of Article 6.9.4.4, is used in
place of (Kℓ/rs) in determining the nominal flexural resistance, Pn, of single-angle members, which
should always be done for single-angle members loaded in combined axial compression and
flexure, flexural-torsional buckling need not be checked (see the Discussion of Article 6.9.4.4 in
this Guide).
The equations are lengthy but are relatively straightforward and are a function of simple material
and geometric parameters. Many of the geometric parameters are section property variables; the
section properties for common rolled steel tee (WT), double-angle, and channel sections can be
found in AISC’s Database of Rolled Steel Shape Section Properties. Designers can easily program

NSBA Guide to Navigating Routine Steel Bridge Design / 178


the resistance equations and integrate the database into a spreadsheet to automate the calculation
of the flexural-torsional buckling resistance of tee (WT), double-angle, or channel sections.
The Commentary for this Article provides a thorough explanation of flexural-torsional buckling,
including helpful tips for simplifying the calculations and for determining when certain buckling
modes may not control.
It should be noted that using rolled steel tee (WT) and double-angle sections as cross-frame
members for routine steel I-girder bridges is generally discouraged, while the use of single-angle
sections is encouraged. The magnitude of the forces in cross-frame members in routine steel I-
girder bridges covered by this Guide is generally small enough such that single-angle members are
adequate. Cross-frame forces are typically only large enough to warrant the use of rolled steel tee
(WT) or double-angle sections in curved and/or skewed steel I-girder bridges. Rolled steel tee
(WT) sections are typically quite expensive to fabricate. Tee (WT) sections are cut from full wide-
flange (W) shapes and generally require straightening after the cutting process, which adds
significant fabrication effort and cost. Double-angle sections are often viewed as problematic from
a maintenance perspective; the surfaces between the adjacent angle flanges are difficult or
impossible to paint in the field, and/or can suffer from potentially severe pack rust.
For further information about the elastic torsional buckling and flexural-torsional buckling
resistance, and for more information about tee (WT), double-angle, and channel sections, consult
Sections 6.6.3.4.2.3.4, 6.6.3.5.3, and 6.6.3.5.4, respectively, of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.
Note that this Manual is based on AASHTO LRFD BDS 7th Edition (2014, with Interim Revisions
through 2015) provisions for double-angle and tee (WT) sections, but the explanations of general
concepts are still valid. Design Example 2A of the FHWA’s Steel Bridge Design Handbook
includes an example cross-frame design.

6.9.4.2 Effects of Local Buckling on the Nominal Compressive Resistance

6.9.4.2.1 Classification of Cross-Section Elements


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article presents the requirements for determining whether longitudinally unstiffened cross-
section elements of members subject to axial compression can be considered nonslender. The
provisions are applicable to a wide variety of cross-sectional elements in both open and closed
sections. For the routine steel I-girder bridges covered by this Guide, the provisions of this Article
are applicable to those members which are commonly used in their design.
The Commentary for this Article explains that, “Compression members with cross-sections
composed only of nonslender longitudinally unstiffened elements are able to develop their full
yield strength under uniform axial compression without any significant impact from local
buckling.” In other words, local buckling will not adversely affect the overall compressive
resistance of those types of sections. Table 6.9.4.2.1-1 provides a handy summary of the applicable
width-to-thickness or slenderness ratio limits defining a nonslender element for a wide variety of

NSBA Guide to Navigating Routine Steel Bridge Design / 179


longitudinally unstiffened cross-sectional elements. The element widths, b, to be applied in
checking these limits are also provided in the table.
The width-to-thickness and slenderness ratio limits are based on simple equations which are easily
calculated by hand or programmed in spreadsheets. Many AISC rolled shapes are proportioned
such that their cross-section elements meet these slenderness limits and they can be classified as
nonslender elements, but designers should check the cross-sectional elements of any member they
are designing as a standard practice.
For further information on slender and nonslender member elements, consult Section 6.6.3.4.2.4
of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. Design Example 2A of the FHWA’s Steel Bridge Design
Handbook includes an example cross-frame design.

6.9.4.2.2 Slender Longitudinally Unstiffened Cross-Section Elements

6.9.4.2.2a General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article presents the equation for the nominal compressive resistance, Pn, of compression
members whose cross-sections are composed of one or more longitudinally unstiffened slender
elements. The classification of a cross-sectional element as slender or nonslender is covered in
Article 6.9.4.2.1 (see the Discussion of Article 6.9.4.2.1 in this Guide).
Neither I-girders nor their flanges in routine steel I-girder bridges are treated as compression
members or beam columns per se. The combined effects of axial compression and flange lateral
bending in the compression flanges of girders and beams are directly addressed in Article 6.10, so
the provisions of this Article are not applicable to the design of I-sections used as main spanning
elements in the routine steel I-girder bridges covered by this Guide.
Cross-frame members may have a cross-section composed of one or more slender longitudinally
unstiffened elements, in which case the provisions of this Article would be applicable. The stems
of a significant number of rolled tee sections and one or both legs of many rolled angle sections
are typically classified as slender elements. In such cases, the provisions of Article 6.9.4.2.2b may
also be applicable (see the Discussion of Article 6.9.4.2.2b in this Guide).
Eqs. 6.9.4.2.2a-1 and 6.9.4.2.2a-2 define the nominal compressive resistance, Pn, as a function of
the critical (or smallest) buckling resistance based on flexural, torsional, or flexural-torsional
buckling of the overall member, reduced to account for the adverse impacts of local buckling of
any longitudinally unstiffened slender elements in the cross-section. The reduction factor is
essentially the ratio of the effective area to the gross area (Aeff /Ag) of the cross-section. The
effective area of the cross-section elements, Aeff, generally reflects reductions in the effective width
of any slender elements in the cross-section, as defined in Article 6.9.4.2.2b for all sections except
circular tubes and round Hollow Structural Shapes (HSS); the effective area, Aeff, for those types
of shapes is addressed in Article 6.9.4.2.2c. Where necessary, Eq. 6.9.4.2.2a-3 should be used to

NSBA Guide to Navigating Routine Steel Bridge Design / 180


calculate Aeff for the rolled sections used as cross-frame members in the routine steel I-girder
bridges covered by this Guide.
For further information on slender member elements, consult Section 6.6.3.4.2.4.3 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. Note that this Manual is based on AASHTO LRFD BDS 7th
Edition (2014, with Interim Revisions through 2015) provisions for slender and nonslender
elements, but the explanations of general concepts are still valid. Design Example 2A of the
FHWA’s Steel Bridge Design Handbook includes an example cross-frame design.

6.9.4.2.2b Effective Width of Slender Elements


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article defines the effective width of slender longitudinally unstiffened elements in members
subject to axial compression. See the Discussion of Article 6.9.4.2.2a in this Guide for further
information on slender longitudinally unstiffened elements. A wide variety of elements are
addressed, many of which are not (and should not be) used in the routine steel I-girder bridges
covered by this Guide.
Neither I-girders nor their flanges in routine steel I-girder bridges are treated as compression
members or beam columns per se. The combined effects of axial compression and flange lateral
bending in the compression flanges of girders and beams are directly addressed in Article 6.10, so
the provisions of this Article are not applicable to the design of I-sections used as main spanning
elements in the routine steel I-girder bridges covered by this Guide.
Cross-frame members may have a cross-section composed of one or more slender longitudinally
unstiffened elements, in which case the provisions of this Article would be applicable. The stems
of a significant number of rolled tee sections and one or both legs of many rolled angle sections
are typically classified as slender elements. In such cases, the provisions of this Article are
applicable. The effective width, be, of a given slender element is used in Article 6.9.4.2.2a to
calculate the effective area, Aeff, used in the calculation of the nominal compressive resistance, Pn,
of the overall member, accounting for the adverse effects of local buckling of any slender elements
in the cross-section (see the Discussion of Article 6.9.4.2.2a in this Guide)

6.9.4.2.2c Effective Area of Circular Tubes and Round HSS


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article specify the effective area, Aeff, of circular tubes and round hollow
structural sections (HSS) to be used in the provisions of Article 6.9.4.2.2a to compute the nominal
compressive resistance, Pn, of these members (see the Discussion of Article 6.9.4.2.2a in this
Guide). Circular tubes and round HSS are not used as compression members in the routine steel I-
girder bridges covered by this Guide, and so the provisions of this Article are not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 181


The connection details associated with circular tubes and round HSS members are typically
expensive to fabricate.

6.9.4.3 Built-Up Members

6.9.4.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article contain general requirements for compression members built-up
from rolled or welded shapes connected by continuous plates with or without perforations or by
tie plates with or without lacing. Provisions are specified in this Article for the welded or bolted
connections between the plates and shapes.
Built-up compression members are not typically used in the routine steel I-girder bridges covered
by this Guide, except possibly the application of double-angle members in cross-frames; the
provisions of this Article are only applicable in that case. It should be noted that using rolled steel
tee (WT) and double-angle sections as cross-frame members for routine steel I-girder bridges is
generally discouraged, while the use of single-angle sections is encouraged. The magnitude of the
forces in cross-frame members in routine steel I-girder bridges covered by this Guide is generally
small enough such that single-angle members are adequate. Cross-frame forces are typically only
large enough to warrant the use of rolled steel tee (WT) or double-angle sections in curved and/or
skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite expensive to
fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally require
straightening after the cutting process, which adds significant fabrication effort and cost. Double-
angle sections are often viewed as problematic from a maintenance perspective; the surfaces
between the adjacent angle flanges are difficult or impossible to paint in the field, and/or can suffer
from potentially severe pack rust.
For further information on built-up compression members, consult Section 6.6.3.4.4 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. Design Example 2A of the FHWA’s Steel Bridge Design
Handbook includes an example cross-frame design.

6.9.4.3.2 Perforated Plates


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains provisions specific to the design of perforated plates used to connect rolled
or welded shapes in built-up compression members (see the Discussion of Article 6.9.4.3.1 in this
Guide). The routine steel I-girder bridges covered by this Guide do not utilize built-up compression
members with perforated plates; therefore, the provisions of this Article are not applicable.

6.9.4.4 Single-Angle Members


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 182


Discussion:
This Article provides a significantly simplified method for determining the nominal compressive
resistance, Pn, of single-angle members subject to combined axial compression and flexure about
one or both principal axes. Single angles provide a very economical solution for, and are
commonly used as, chord and diagonal members in truss-type cross-frames in routine steel I-girder
bridges. Single angles used as cross-frame members in routine steel I-girder bridges must satisfy
the following conditions for the provisions of this Article to be applicable:
1. End connections are to a single leg of the angle, and are welded or use a minimum of two
bolts
2. The angle is loaded at the ends in compression through the same leg
3. The angle is not subjected to any intermediate transverse loads:
Cross-frame members in routine steel I-girder bridges are typically configured as truss-type
structures (K-frames, X-frames, etc.) in a single plane, and are routinely attached either to gusset
plates or directly to cross-frame connection plates (stiffeners) using either welded or bolted
connections. This is common and economical connection detailing, as can be seen in AASHTO-
NSBA Steel Bridge Collaboration’s Guidelines G1.4-2006 Guidelines for Design Details and
G12.1-2020 Guidelines to Design for Constructability and Fabrication. Cross-frames detailed in
accordance with the recommendations in these guidelines can be designed using the provisions of
this Article.
As explained in the Commentary: “In essence, these provisions permit the effect of the
eccentricities to be neglected when these members are evaluated as axially loaded compression
members for flexural buckling only using an appropriate specified effective slenderness ratio,
(Kℓ/r)eff, in place of (Kℓ/rs) in Eq. 6.9.4.1.2-1.” See the Discussion of Article 6.9.4.1.2 in this Guide
for more information. Eqs. 6.9.4.4-1 or 6.9.4.4-2, as applicable, are typically used to calculate
(Kℓ/r)eff. It is important to note that the length, ℓ, to be used in calculating (Kℓ/r)eff is to be taken as
the distance between the work points of the end connections measured along the length of the
angle, rather than the physical length of the angle member. The resulting value of Pn may need to
be adjusted according to the provisions of Article 6.9.4.2.2a if one or both legs of the rolled angle
section are classified as slender elements (see the Discussion of Article 6.9.4.2.2a in this Guide).
The effective slenderness ratio indirectly accounts for the bending in the angles due to the
eccentricity of the loading allowing the member to be proportioned according to the provisions of
Article 6.9.2.1 as if it were a pinned-end concentrically loaded compression member. Furthermore,
when the effective slenderness ratio is used, single angles need not be checked for flexural–
torsional buckling. (see the Discussion of Article 6.9.4.1.3 in this Guide). It is also important to
note that the actual maximum slenderness ratio of the angle member, Kℓ/rz, is to be used in
checking the limiting slenderness ratio for compression members specified in Article 6.9.3 (see
the Discussion of Article 6.9.3 in this Guide) rather than (Kℓ/r)eff, where K is taken equal to 1.0 for
single-angle members (Article 4.6.2.5),  is the actual physical length of the angle member, and rz
is the radius of gyration taken about the minor principal axis of the member.

NSBA Guide to Navigating Routine Steel Bridge Design / 183


The Commentary goes on to further explain the basis, and the associated limitations and criteria,
for this methodology. This methodology greatly simplifies calculations of the resistance of single-
angle members subject to combined axial compression and flexure about one or both principal
axes. The calculations can be performed by hand or easily programmed in a spreadsheet.
For further information on single-angle compression members, consult Section 6.6.3.4.5 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. Design Example 2A of the FHWA’s Steel Bridge Design
Handbook includes an example cross-frame design.

6.9.4.5 Plate Buckling Under Service and Construction Loads


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
These provisions do not apply to the design of the routine steel I-girder bridges covered by this
Guide.
Post-buckling resistance is assumed at the strength limit state in computing the nominal axial
compressive and flexural resistance of noncomposite box-section members with slender webs
and/or noncompact or slender flanges or plate panels. Under the provisions of Article 6.9.4.5, these
particular elements are investigated to check that plate local buckling does not occur under
conditions producing maximum longitudinal compressive stress acting at one or both longitudinal
edges of the element under consideration at the service limit state or during construction.
As explained in the Article itself, composite and noncomposite I-section members subject to
flexure only, such as are used as main girders in routine steel I-girder bridges, are checked for web
bend buckling at various limit states according to the provisions of Article 6.10. Furthermore, the
provisions of Article 6.9.4.5 are not applicable to the design of plate elements supported only along
one longitudinal edge (such as the flanges of I-shaped members); other provisions of Section 6 of
the BDS address the design of flanges in compression for I-shaped members.

6.9.5 Composite Members

6.9.5.1 Nominal Compressive Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article apply to the design of composite columns (i.e., concrete-filled tubes
or pipes and concrete-encased shapes) without flexure and are used to determine the nominal
compressive resistance, Pn, of these columns. For columns subject to combined axial compression
and flexure, the calculation of the nominal flexural resistance, Mn, of concrete-filled tubes or pipes
is covered in Article 6.12.2.3.2 and the calculation of Mn of concrete-encased shapes is covered in
Article 6.12.2.3.1 (see the Discussion of Articles 6.12.2.3.1 and 6.12.2.3.2 in this Guide). As such,
these provisions are considered beyond the scope of superstructure design. These members could

NSBA Guide to Navigating Routine Steel Bridge Design / 184


potentially be used in substructures for the routine steel I-girder bridges covered by this Guide,
although they are not commonly utilized.

6.9.5.2 Limitations

6.9.5.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article apply to the design of composite columns (i.e., concrete-filled tubes
or pipes and concrete-encased shapes) without flexure and specify limitations that must be met in
order to use the nominal compressive resistance equations for these columns specified in Article
6.9.5.1 (see the Discussion of Article 6.9.5.1 in this Guide). As such, these provisions are
considered beyond the scope of superstructure design. These members could potentially be used
in substructures for the routine steel I-girder bridges covered by this Guide, although they are not
commonly utilized.

6.9.5.2.2 Concrete-Filled Tubes


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
This Article applies to the design of composite columns without flexure utilizing concrete-filled
tubes or pipes where full composite action is not deemed necessary. These members could
potentially be used in substructures for the routine steel I-girder bridges covered by this Guide,
although they are not commonly utilized. As such, this provision is considered beyond the scope
of superstructure design.
This Article specifies that the wall thickness requirements for unfilled tubes given in Article 6.9.4.2
are to apply to concrete-filled tubes. For applications where full composite action is deemed
necessary under combined axial compression and flexure, the provisions of Articles 6.9.6 and
6.12.2.3.3 should be employed instead (see the Discussion of Articles 6.9.6 and 6.12.2.3.3 in this
Guide).

6.9.5.2.3 Concrete-Encased Shapes


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article apply to the design of composite columns without flexure utilizing
concrete-encased steel shapes and specifies reinforcement requirements for these columns. These
members could potentially be used in substructures for the routine steel I-girder bridges covered

NSBA Guide to Navigating Routine Steel Bridge Design / 185


by this Guide, although they are not commonly utilized. As such, these provisions are considered
beyond the scope of superstructure design.

6.9.6 Composite Concrete-Filled Steel Tubes (CFSTs)

6.9.6.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article apply to the design of composite columns utilizing larger-diameter
composite concrete-filled steel tubes, or CFSTs, with or without internal reinforcement subject to
axial compression or combined axial compression and flexure for non-seismic applications. The
computation of the nominal flexural resistance of these members, Mn, as a function of the nominal
axial resistance, Pn, is covered in Article 6.12.2.3.3 (see the Discussion of Article 6.12.2.3.3 in this
Guide). These members could potentially be used in substructures for the routine steel I-girder
bridges covered by this Guide, although they are not commonly utilized. As such, these provisions
are considered beyond the scope of superstructure design.

6.9.6.2 Limitations
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article apply to the design of composite columns utilizing larger-diameter
composite concrete-filled steel tubes, or CFSTs, with or without internal reinforcement subject to
axial compression or combined axial compression and flexure for non-seismic applications. This
Article gives specific limitations for the use of these members, including a slenderness ratio limit
for the steel tube. These members could potentially be used in substructures for the routine steel I-
girder bridges covered by this Guide, although they are not commonly utilized. As such, these
provisions are considered beyond the scope of superstructure design.

6.9.6.3 Combined Axial Compression and Flexure

6.9.6.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article apply to the design of composite columns utilizing larger-diameter
composite concrete-filled steel tubes, or CFSTs, with or without internal reinforcement subject to
combined axial compression and flexure for non-seismic applications. This Article deals with the
development of a factored stability-based P-M interaction resistance curve for these members.
These members could potentially be used in substructures for the routine steel I-girder bridges

NSBA Guide to Navigating Routine Steel Bridge Design / 186


covered by this Guide, although they are not commonly utilized. As such, these provisions are
considered beyond the scope of superstructure design.

6.9.6.3.2 Axial Compressive Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article apply to the design of composite columns utilizing larger-diameter
composite concrete-filled steel tubes, or CFSTs, with or without internal reinforcement subject to
combined axial compression and flexure for non-seismic applications. This Article covers the
calculation of the factored axial compressive resistance, Pr, of these members. These members
could potentially be used in substructures for the routine steel I-girder bridges covered by this
Guide, although they are not commonly utilized. As such, these provisions are considered beyond
the scope of superstructure design.

6.9.6.3.3 Nominal Flexural Composite Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provision in this Article applies to the design of composite columns utilizing larger-diameter
composite concrete-filled steel tubes, or CFSTs, with or without internal reinforcement subject to
combined axial compression and flexure for non-seismic applications. The provision points to
Article 6.12.2.3.3 for the computation of the nominal flexural resistance of these members, Mn, as
a function of the nominal axial resistance, Pn (see the Discussion of Article 6.12.2.3.3 in this
Guide). These members could potentially be used in substructures for the routine steel I-girder
bridges covered by this Guide, although they are not commonly utilized. As such, this provision
is considered beyond the scope of superstructure design.

6.9.6.3.4 Nominal Stability-Based Interaction Curve


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article apply to the design of composite columns utilizing larger-diameter
composite concrete-filled steel tubes, or CFSTs, with or without internal reinforcement subject to
combined axial compression and flexure for non-seismic applications. The provisions of this
Article outline the specific steps necessary to modify the material-based interaction curve
developed according to the provisions of Article 6.12.2.3.3 (see the Discussion of Article
6.12.2.3.3 in this Guide) to include stability effects based on the buckling load determined in
Article 6.9.6.3.2 to create a nominal stability-based interaction curve, which is then multiplied by
the appropriate resistance factor specified in Article 6.5.4.2 to determine the final factored

NSBA Guide to Navigating Routine Steel Bridge Design / 187


resistance of the CFST for combined axial compression and flexure for all load conditions. As
such, these provisions are considered beyond the scope of superstructure design. These members
could potentially be used in substructures for the routine steel I-girder bridges covered by this
Guide, although they are not commonly utilized.

6.10 I-SECTION FLEXURAL MEMBERS

6.10.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article contains general requirements and provisions for I-section flexural design, including
a listing of the Articles in Article 6.10 containing the beam or girder proportioning requirements
and the design requirements at each limit state, which are typically applicable or at least partially
applicable to the girders or beams used as the primary load-carrying members in routine steel I-
girder bridges, as described further in the various Discussions of related Articles provided in this
Guide.
These provisions would also be applicable (or partially applicable) to an I-section used as the top
chord of an end cross-frame, or as an end diaphragm, that is designed as a flexural member to
support the wheel loads coming onto the end of the deck.
For routine multi-span continuous rolled-beam bridges, strong consideration should be given to
applying the applicable design provisions of the optional Appendix A6 for constructibility and at
the strength limit state (see the Discussion of Appendix A6 in this Guide). These provisions
account for the ability of some I-sections to develop flexural resistances significantly greater than
the yield moment, My, when certain proportioning requirements are met; taking advantage of this
ability could potentially lead to a much more economical design.
The definition of a routine steel I-girder bridge specifically excludes the use of moment
redistribution methods and so the optional provisions of Appendix B6 are considered not
applicable to the routine steel I-girder bridges covered by this Guide (see the Discussion of Article
4.6.4.1 in this Guide).
The discussions in the Commentary for this Article on bridges containing both straight and curved
segments, kinked (chorded) girders, the consideration of flange lateral bending effects when cross-
frames or diaphragms are placed in discontinuous lines in skewed bridges, and the consideration
of flange lateral bending effects in horizontally curved bridges are not applicable to the routine
steel I-girder bridges covered by this Guide. For these bridges, flange lateral bending effects tend
to be most significant during construction and tend to be insignificant in the final constructed
condition.
Flowcharts for flexural design of I-section members according to the provisions of Article 6.10 are
provided in Appendix C6. These flowcharts are helpful in guiding the designers through the design
provisions at each limit state (see the Discussion of Appendix C6 in this Guide). Fundamental
calculations for flexural members (e.g., section property calculations, calculation of the depth of the

NSBA Guide to Navigating Routine Steel Bridge Design / 188


web in compression, web crippling and web local yielding checks, etc.) are provided in Appendix
D6 (see the Discussion of Appendix D6 in this Guide).
For design examples illustrating the flexural design of steel I-girders at each limit state, consult
the FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous Straight
Composite Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous Straight Composite
Steel I-Girder Bridge, and Design Example 2B, Two-Span Continuous Straight Composite Steel
Wide-Flange Beam Bridge. The calculations in Examples 2A and 2B illustrating the application of
the optional moment redistribution methods specified in Appendix B6 are considered not applicable
to the routine steel I-girder bridges covered by this Guide.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. Other
commercial software packages with the ability to analyze and design routine steel I-girder bridges
are also available. Users should verify the capabilities, assumptions, and general correctness of any
program’s calculations prior to initial use.
See also NSBA’s Continuous Span Standards, NSBA’s Span-to-Weight Curves, and the Short
Span Steel Bridge Alliance’s Technical Design Resources for Short Span Steel Bridges, which
provide handy benchmark data for routine steel I-girder bridge designs.

6.10.1.1 Composite Sections


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The Article provides a basic definition of a composite section as a section where the top flange of
the steel section is connected to the concrete deck with shear connectors (Article 6.10.10) in the
bridge’s final, completed condition. The shear connectors transfer horizontal shear from the deck
to the girder and prevent slip parallel to the girder between the concrete and the steel. Under these
conditions, a linear strain distribution from the top of the deck to the bottom of the girder can be
assumed, with a single location of the neutral axis of the section. Under these conditions, the
composite concrete deck can also be assumed to provide continuous lateral support to the top
flange. This Article is applicable to the routine steel I-girder bridges covered by this Guide (except
as noted in the Discussion for Article 6.10.1.2 in this Guide).
Note that composite action is not present during construction, prior to the placement and hardening
of the deck.

6.10.1.1.1 Stresses

6.10.1.1.1a Sequence of Loading


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article describes for a composite section the necessary accumulation of the elastic stresses
due to the applied dead and live loads acting on different sections; that is, the noncomposite

NSBA Guide to Navigating Routine Steel Bridge Design / 189


component dead loads (referred to herein as DC1 loads) acting on the bare steel section, the
composite component dead loads (referred to herein as DC2 loads) acting on the long-term (3n)
transformed composite section to account for the effects of concrete creep, the wearing surface
and utility loads (referred to as DW loads) acting on the long-term (3n) transformed composite
section, and the live loads plus the dynamic load allowance (LL+IM) acting on the short-term (n)
transformed composite section. The calculation of the long-term and short-term transformed
composite sections is described in Articles 6.10.1.1.1b and 6.10.1.1.1c (see the Discussion of
Articles 6.10.1.1.1b and 6.10.1.1.1c in this Guide). The accumulation of the elastic stresses must
be accounted for in the design of steel I-girder bridges at the service and strength limit states (and
in some cases involving the dead loads at the fatigue limit state).
This accumulation of the elastic stresses reflects the assumption that the routine steel I-girder
bridges covered by this Guide are built using unshored construction, in which no support of the
steel beams or girders (other than at permanent support points) is provided during the concrete
deck construction, including no temporary supports. As a result, the bare steel beams or girders
resist the permanent load applied before the concrete deck hardens and the composite girder section
(steel girder alone, steel girder plus the composite concrete deck, or steel girder plus the
longitudinal deck reinforcement, as applicable – see the Discussion of Articles 6.10.1.1.1b and
6.10.1.1.1c in this Guide) resists the permanent and transient loads applied after the concrete deck
hardens. This reflects the common practices used throughout the United States for the construction
of virtually all steel-girder bridges, including the routine steel I-girder bridges covered in this
Guide, in which temporary shoring is provided only during steel erection, if at all. Temporary
shoring is not provided during the placement of the concrete deck.
Shored construction, in which the steel beams or girders would be theoretically supported along
their entire length during the concrete deck construction so that the composite girder would resist
both permanent and transient loads, is permitted in this Article but is not recommended and
essentially never used since little is currently known about the effects of concrete creep on
composite steel girders under large dead loads; therefore, shored construction is considered not
applicable to the routine steel I-girder bridges covered by this Guide.
Since plane sections are assumed to remain plane, the calculated elastic stresses at any given point
on the cross-section due to the various loadings acting on their associated noncomposite, short-
term composite, or long-term composite sections may be summed. However, at elastic stress
levels, the principle of superposition does not apply to the bending moments due to the various
loadings, as these moments are each applied to different sections (i.e., the noncomposite, short-
term composite, and long-term composite sections). The girder stiffness is changing as each
moment is applied. Therefore, at elastic stress levels, individual bending moments may not be
summed.
Once the yield stress is exceeded however, it is considered acceptable to sum the individual
bending moments. This approach would be valid only if the optional provisions of Appendix A6
are being implemented (see the Discussion of Appendix A6 in this Guide).
Several guideline documents, such as the Reference Manual for NHI Course 130081, Load and
Resistance Factor Design (LRFD) for Highway Bridge Superstructures and various volumes of the
FHWA’s Steel Bridge Design Handbook, provide good discussions of this topic.

NSBA Guide to Navigating Routine Steel Bridge Design / 190


The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the stresses in the section in accordance with the provisions of the AASHTO LRFD BDS, greatly
reducing the time and effort required of the designer. Other commercial software packages with the
ability to analyze and design routine steel I-girder bridges are also available. Users should verify the
capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.10.1.1.1b Stresses for Sections in Positive Flexure


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article contains provisions for the calculation of flexural stresses for sections in regions of
positive flexure and is applicable to the routine steel I-girder bridges covered by this Guide.
For calculating flexural stresses within sections subjected to positive flexure for both long-term
and short-term moments applied to the composite section at all limit states, the composite section
consists of the steel section and the appropriate transformed area of the effective width of the
concrete deck; i.e., the concrete deck is transformed into equivalent steel by dividing the effective
width of the deck (see the Discussion of Article 4.6.2.6.1 in this Guide) by the modular ratio. The
modular ratio is equal to the ratio of the modulus of elasticity of steel to the modulus of elasticity
of concrete (Eq. 6.10.1.1.1b-1). The deck width is reduced rather than the deck thickness to have
a less significant effect on the computed moment of inertia.
“Transient loads” consist of live load (LL) and dynamic load amplification, or “impact” (IM). For
transient loads (i.e., LL + IM), which are applied to the short-term composite section, the concrete
deck area is transformed by using the short-term modular ratio, n.
“Permanent loads” consist of dead loads which are applied to the bridge after the composite deck
has been placed and hardened. These loads are grouped into one of two categories. The first
category, designated DC2, represents the weight of structural component and nonstructural
attachments which are added to the bridge after deck construction, including such items as barrier
rails, medians, attached signs or lights, etc. The second category, designated DW, represents the
weight of wearing surfaces that may be applied to the deck (either at the time of initial construction
of the bridge or in the future) and the weight of utilities that may be attached to the bridge (either
utilities attached at the time of initial construction of the bridge, or a load allowance for utilities
that may be attached to the bridge in the future). For DC2 and DW loads, which are applied to the
long-term composite section, the concrete deck is transformed by using the long-term modular
ratio, 3n, to account for the effects of concrete creep. When concrete is placed under a sustained
long-term stress, there is an instantaneous elastic strain, followed by a time-dependent increase in
strain known as creep. In a composite girder, due to the effects of creep, the strain in the steel
girder increases and the steel stresses become larger, while the strains and stresses in the concrete
deck are reduced. The short-term modular ratio is based on the initial tangent modulus, Ec, of the
concrete, while the long-term modular ratio is based on an effective apparent modulus, Ec/3, to

NSBA Guide to Navigating Routine Steel Bridge Design / 191


account for the effects of concrete creep in an approximate fashion. The effects of creep are usually
conservatively ignored in the computation of the stresses in the concrete deck.
The longitudinal deck reinforcement is not considered effective in compression at the strength
limit state because it is not tied; therefore, its contribution is typically neglected when computing
the composite section properties in regions of positive flexure for strength limit state checks.
Consideration may be given to including the longitudinal deck reinforcement within the effective
width of the deck in the composite section properties when computing the stresses at the fatigue
and service limit states. Typically, the area of the concrete deck haunch is not considered in the
computation of the composite section properties; the haunch depth may be considered, however,
if permitted by the Owner-agency. Note that many Owner-agencies design policies neglect the
depth of the haunch in the calculation of section properties since variations from the girder’s
anticipated camber are typically taken up in the haunch. Also note that the haunch dimension will
typically not be a constant value when rolled beam sections are used.
For further discussion and sample calculations of the section properties and elastic stresses for
composite sections in regions of positive flexure, consult Sections 6.4.2.3.2 and 6.4.2.4.1 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures, as well as FHWA’s Steel Bridge Design Handbook, Design
Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge and Design Example
2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge.
For routine rolled-beam bridges, it is possible that the elastic neutral axis of the transformed
composite section may fall within the deck. Concrete on the tension side of the neutral axis is not
to be considered effective at the strength limit state; the concrete below the neutral axis is assumed
cracked in tension and therefore ineffective. In such cases, the effective transformed area of the
concrete becomes a function of the neutral-axis position. Consult pp. 6.194 and 6.195 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures and Design Example 2B, Two-Span Continuous Straight
Composite Steel Wide-Flange Beam Bridge for an example calculation of the composite section
properties for such a case.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the stresses in the section in accordance with the provisions of the AASHTO LRFD BDS, greatly
reducing the time and effort required of the designer. Other commercial software packages with the
ability to analyze and design routine steel I-girder bridges are also available. Users should verify the
capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.10.1.1.1c Stresses for Sections in Negative Flexure


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 192


Discussion:
Simple Span Bridges:
This Article does not apply for simple span routine steel I-girder bridges as such bridges are only
subject to positive moments.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This Article contains provisions for the calculation of flexural stresses for sections in regions of
negative flexure and is applicable to the multi-span continuous routine steel I-girder bridges
covered by this Guide.
For calculating flexural stresses within sections subjected to negative flexure at the strength limit
state for both long-term and short-term moments applied to the composite section, the composite
section consists of either the steel section alone, or the steel section plus the longitudinal
reinforcement within the effective width of the concrete deck if permitted by the Owner-agency
policy. The section properties and strength of the concrete deck itself in tension are always ignored
at the strength limit state. The AASHTO LRFD BDS permits the longitudinal reinforcement
(including the minimum area of longitudinal reinforcement specified for control of deck cracking
– Article 6.10.1.7) within the effective width of the deck to be included in the section properties if
stud shear connectors are present in regions of negative flexure. Including consideration of the
longitudinal reinforcing in the deck can facilitate the use of a smaller top flange than bottom flange
in regions of negative flexure and is recommended for the routine steel multi-span continuous I-
girder bridges with stud shear connectors in regions of negative flexure if permitted by the Owner-
agency.
Articles 6.6.1.2.1 and/or 6.10.4.2.1 permit the concrete deck to be considered effective in tension
when computing the composite section properties for negative flexure at the fatigue and/or service
limit states, respectively, when certain specified requirements are satisfied (and if permitted by the
Owner-agency); otherwise, the concrete deck in tension is ignored (see the Discussion of Articles
6.6.1.2.1 and 6.10.4.2.1 in this Guide). In that case, the properties of the long-term 3n composite
section (including the transformed area of the concrete deck) would be used for permanent (dead)
loads applied after the concrete deck has hardened (i.e., DC2 and DW loads). The properties of the
short-term n composite section (including the transformed area of the concrete deck) would be
used for transient (live) loads applied after the concrete deck has hardened (i.e., LL+ IM loads).
Consideration may be given to including the longitudinal deck reinforcement within the effective
width of the concrete deck in the composite section properties in such cases.
For further discussion and sample calculations of the section properties and elastic stresses for
sections in regions of negative flexure, consult Sections 6.4.2.3.3 and 6.4.2.4.1 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures, as well as FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-
Span Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span
Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span
Continuous Straight Composite Steel Wide-Flange Beam Bridge.

NSBA Guide to Navigating Routine Steel Bridge Design / 193


The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the stresses in the section in accordance with the provisions of the AASHTO LRFD BDS, greatly
reducing the time and effort required of the designer. Other commercial software packages with the
ability to analyze and design routine steel I-girder bridges are also available. Users should verify the
capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.10.1.1.1d Concrete Deck Stresses


Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
Simple Span Bridges:
This Article contains provisions for the calculation of longitudinal flexural stresses in the concrete
deck and is conditionally applicable to the routine steel simple span I-girder bridges as described
below.
For simple span routine steel I-girder bridges, only the longitudinal compressive stresses in the
concrete deck are needed if the section is treated as a noncompact section at the strength limit state
(Article 6.10.7.2.1); the concrete deck is not subject to tension in a simple span bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the stresses in the section in accordance with the provisions of the AASHTO LRFD BDS, greatly
reducing the time and effort required of the designer. Other commercial software packages with the
ability to analyze and design routine steel I-girder bridges are also available. Users should verify the
capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This Article contains provisions for the calculation of longitudinal flexural stresses in the concrete
deck and is applicable to the multi-span continuous routine steel I-girder bridges covered by this
Guide.
The longitudinal flexural tensile stresses in the concrete deck are needed to determine the cut-off
points for the minimum one-percent longitudinal reinforcement in the deck (see the Discussion of
Article 6.10.1.7 in this Guide) and to determine if the concrete deck may be considered effective
in tension at the service limit state (see the Discussion of Article 6.10.4.2.1 in this Guide) in routine
steel multi-span continuous I-girder bridges in regions of negative flexure. The longitudinal
flexural compressive stresses in the concrete deck are needed to check the specified deck-stress
limit for noncompact sections in regions of positive flexure at the strength limit state for routine

NSBA Guide to Navigating Routine Steel Bridge Design / 194


steel I-girder bridges in which the sections are treated as noncompact sections in these regions (see
the Discussion of Article 6.10.7.2.1 in this Guide).
In a composite girder, longitudinal flexural stresses in the concrete deck are assumed to result only
from the permanent loads and transient loads applied after the concrete deck has hardened; i.e., the
DC2, DW, and (LL+IM) loads. For reasons described in the Commentary for this Article, the short-
term n-composite section (including the transformed area of the concrete deck) is to be used to
calculate the deck stresses due to these loads. It is important to remember that the calculated stress
at the top of the transformed concrete deck must be divided by the modular ratio, n, to obtain the
maximum stress in the concrete.
For further discussion and a sample calculation of the concrete deck stresses, consult Section
6.4.2.4.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures, as well as FHWA’s Steel Bridge Design Handbook,
Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge, Design
Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the stresses in the section in accordance with the provisions of the AASHTO LRFD BDS, greatly
reducing the time and effort required of the designer. Other commercial software packages with the
ability to analyze and design routine steel I-girder bridges are also available. Users should verify the
capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.10.1.1.1e Effective Width of Concrete Deck


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article indicates where the provisions are located for the calculation of the effective width of
the concrete deck (i.e., Article 4.6.2.6.1), which is used in the calculation of the composite section
properties for sections in which the concrete deck is considered effective in compression (or
tension, as applicable) for the routine steel I-girder bridges covered by this Guide. Only the
tributary width provisions in the first paragraph of Article 4.6.2.6.1 should be considered
applicable to the computation of the effective width of the concrete deck for the routine steel I-
girder bridges covered by this Guide (see the Discussion of Article 4.6.2.6.1 in this Guide).

6.10.1.2 Noncomposite Sections


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 195


Discussion:
Simple Span Bridges:
This Article provides the basic definition of a noncomposite section as a section where the top
flange of the steel section is not connected to the concrete deck by shear connectors. Thus, slip is
assumed to occur between the steel section and the concrete deck. Although unintended composite
action does occur, it is conservatively ignored. Therefore, in such cases, flexural stresses in the
section due to all permanent and transient loads are computed based on the section properties of
the steel section only.
The definition of a noncomposite section is not applicable to the simple span routine I-girder
bridges covered by this Guide. Providing no shear connectors along the entire length of the
girders/beams of a steel girder bridge is not recommended, and such designs would not represent
a routine steel I-girder bridge as covered by this Guide.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This Article provides the basic definition of a noncomposite section as a section where the top
flange of the steel section is not connected to the concrete deck by shear connectors. Thus, slip is
assumed to occur between the steel section and the concrete deck. Although unintended composite
action does occur, it is conservatively ignored. Therefore, in such cases, flexural stresses in the
section due to all permanent and transient loads are computed based on the section properties of
the steel section only.
This Article is only conditionally applicable to multi-span continuous routine steel I-girder bridges
covered by this Guide; the provisions of this Article would only apply to sections in which shear
connectors are intentionally omitted in regions of negative flexure in multi-span continuous routine
steel I-girder bridges. The decision to omit shear connectors in regions of negative flexure is
typically dependent on the preferences of the Owner-agency. However, unless specifically
required by Owner-agency policy, this practice is not recommended.

6.10.1.3 Hybrid Sections


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article covers the design of hybrid-section members, in which a hybrid section is defined as
a fabricated steel section with a web that has a specified minimum yield strength less than one or
both flanges. Rolled-beam sections are rolled from a single billet of steel and thus cannot be hybrid
sections. The routine steel I-girder bridges covered by this Guide are assumed to contain only
homogeneous-section members and do not contain hybrid-section members; therefore, the
provisions of this Article are not applicable to their design.

6.10.1.4 Variable Web Depth Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 196


Discussion:
This Article covers the design of variable web depth members, including haunched girders and
girders with a linearly varying depth. Rolled-beams sections are always constant depth. The routine
steel I-girder bridges covered by this Guide are assumed to contain only constant-depth members;
therefore, the provisions of this Article are not applicable to their design.

6.10.1.5 Stiffness
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article contains a description of the stiffness properties for various load types to be assumed
in the analysis. The provisions in the first paragraph of this Article are considered applicable to
the routine steel I-girder bridges covered by this Guide.
For permanent loads applied to the noncomposite section, the stiffness properties of the steel
section alone are to be used in the analysis. For permanent loads and transient loads applied to
composite flexural members at all limit states, the stiffness properties of the full composite section
are to be used in the analysis, with the stiffness properties of the long-term composite section used
for the permanent loads and the stiffness properties of the short-term composite section used for
the transient loads. See the Discussions of Article 6.10.1.1.1 and its associated sub-Articles in this
Guide for more information on the short-term and long-term composite sections. At sections where
the composite stiffness properties are used, the concrete is to be assumed effective in tension and
compression for the analysis (i.e., along the entire span length).
In multi-span continuous bridges, it could be theorized that the composite section in negative
moment regions would typically have a different stiffness for design calculations at the strength
limit state because the concrete deck in tension is assumed cracked for design and not participating.
However, moments and deflections computed assuming full composite action agree much better
with field measurements than those computed with a assuming no composite action. Consequently,
the Article specifies that the concrete deck must be assumed to be effective over the entire span
length for the analysis. Assuming the composite stiffness to be effective over the entire span length
gives greater girder moments at the pier and slightly smaller mid-span moments compared to
analyses based on assuming composite action in the so-called positive moment regions only. The
increase in negative girder moments occurs over a relatively short length of what is typically a
larger cross-section, while the reduction in moment occurs over a much longer positive moment
region.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the section properties for the stiffness analysis in accordance with the provisions of the AASHTO
LRFD BDS, greatly reducing the time and effort required of the designer. Other commercial software
packages with the ability to analyze and design routine steel I-girder bridges are also available. Users
should verify the capabilities, assumptions, and general correctness of any program’s calculations
prior to initial use.

NSBA Guide to Navigating Routine Steel Bridge Design / 197


The provisions in second paragraph of this Article dealing with the modeling of girder torsional
stiffness in skewed and/or curved I-girder bridges are not applicable to the routine steel I-girder
bridges covered by this Guide.

6.10.1.6 Flange Stresses and Member Bending Moments


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
General
The first three paragraphs of this Article contain provisions defining the stresses or moments, as
applicable, to be used when specific flexural resistance design checks are made. The provisions in
these paragraphs are considered applicable to the routine steel I-girder bridges covered by this
Guide. For the checking of lateral-torsional buckling resistance (see the Discussion of Articles
6.10.8.2.3 and A6.3.3 in this Guide), the correct value of the factored flexural stress, fbu, or factored
bending moment, Mu, to be used is generally the largest value causing compression in the flange
under consideration throughout the unbraced length. For design checks where the flexural
resistance is based on yielding, flange local buckling or web bend-buckling, fbu and Mu may be
determined as the corresponding values at the section under consideration.
This Article also contains provisions related to flange lateral bending stresses due to torsion and
the amplification of such stresses in discretely braced compression flanges. A discretely braced
flange is defined as a flange supported at discrete intervals by bracing sufficient to restrain lateral
deflection of the flange and twisting of the entire cross-section at the brace points (cross-frames or
diaphragms, see the Discussion of Article 6.7.4 in this Guide).
Sources of Flange Lateral Bending
Flange lateral bending stresses due to primary dead load and live load effects are not significant in
routine steel I-girder bridges as defined for the purposes of this Guide, since their geometric
parameters and framing do not lend themselves to the development of significant torsion in the
girders. Instead, for the routine steel I-girder bridges covered by this Guide, flange lateral bending
stresses generally arise from the following two sources:
• Torsion due to deck overhang loads acting on the discretely braced flanges of the bare steel
exterior (fascia) girders in regions of positive flexure during the concrete deck
construction; and
• Flange lateral bending stresses due to wind loads, both during construction and in the final
condition.
For more information on deck overhang loads during construction and the torsion they cause in
girders, please see the Discussion of Article 6.10.3.4.1 in this Guide.
Wind loading is a consideration both during construction and in the final condition and should be
addressed in the girder constructibility checks. During construction, wind loading contributes to
flange lateral bending stresses potentially both before and during deck placement. Typically, a

NSBA Guide to Navigating Routine Steel Bridge Design / 198


higher wind velocity (and thus greater wind load) is assumed during “construction inactive”
conditions (i.e., when the structural steel is erected, but prior to deck placement operations, such
as for example during overnight or weekend times, or during storm events, when the contractor is
not working) than is assumed during “construction active” conditions (i.e., during the time when
the contractor is actively working to place the concrete deck). Appropriate load combinations
should be developed and investigated, considering both permanent and temporary dead loads, wind
load, and the presence or absence of construction live loads (see the Discussion of Article 3.4.2.1
in this Guide). Many Owners have specific policies regarding required loads and load combinations
to investigate during construction. The Reference Manual for NHI Course 130102, Engineering for
Structural Stability in Bridge Construction also provides good discussion of wind loading during
construction. Under final conditions (i.e., when construction is complete and the bridge is in service),
wind loads acting on discretely braced bottom flanges cause flange lateral bending (i.e., at the
strength limit state). In lieu of a more refined analysis, Article C6.10.3.4.1 gives approximate
equations for calculation of the maximum flange lateral bending moments due to eccentric concrete
deck overhang loads (see the Discussion of Article 6.10.3.4.1 in this Guide). Determination of flange
wind moments is addressed in Article 4.6.2.7 (see the Discussion of Article 4.6.2.7 in this Guide).
Treatment of Flange Lateral Bending Stresses
It should be noted that most line girder analysis programs do not directly calculate flange lateral
bending stresses. Line girder analysis programs use one-dimensional analysis models which cannot
directly address the torsional loading effects that produce flange lateral bending. Some programs
include the ability for the user to manually input flange lateral bending moments or stresses,
calculated by the user outside the program; other programs may not have this ability. In many cases,
it may be easiest to evaluate the effects of flange lateral bending on the design outside of a line girder
analysis program using hand calculations, perhaps facilitated by programming these calculations in
a spreadsheet. This allows the designer flexibility, particularly in addressing the various
constructibility checks which must be performed. If it turns out that consideration of the effects of
flange lateral bending, whether in a constructibility check, or in the supplemental checks of the
bottom flange under final conditions, results in a controlling design condition, the designer may have
to rerun the line girder analysis with resized girders to update major-axis bending stress calculations,
deflection calculations, or other calculations.
For discretely braced compression flanges, the largest lateral bending stress, fℓ, throughout the
unbraced length of the flange must be used in combination with fbu or Mu when the flexural
resistance is based on lateral-torsional buckling; otherwise, fℓ may be determined as the
corresponding value at the section under consideration. When the maximum values of fℓ and fbu
or Mu occur at different locations within the unbraced length, it is conservative to use the maximum
values in a single application of the yielding and flange local buckling resistance equations. Flange
lateral bending is not a consideration in the web bend-buckling resistance equations because the
flange lateral bending stress is zero at the web. Top flange lateral bending stresses are ignored once
the flange is continuously braced by the hardened concrete deck. The resistance of the composite
concrete deck is adequate to compensate for the neglect of these initial lateral bending stresses.
The upper limit on the lateral bending stresses in discretely braced flanges given by Eq. 6.10.1.6-
1 applies to the lateral bending stresses in routine steel I-girder bridges. The limit applies after any

NSBA Guide to Navigating Routine Steel Bridge Design / 199


necessary amplification is applied to the first-order lateral bending stresses (see below). For cases
in which the total elastically-computed flange lateral bending stress is larger than the limit of
0.6Fyf, the reduction in the major-axis bending resistance due to flange lateral bending tends to be
greater than that determined based on the approximate one-third rule flexural resistance equations
utilized throughout Section 6 to combine the effects of major-axis bending stresses and flange
lateral bending stresses due to torsion. For further information on the one-third rule flexural
resistance equations and their development, consult Section 6.5.2 of the Reference Manual for NHI
Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.
Amplification of First-Order Flange Lateral Bending Stresses
To determine whether amplification of the first-order lateral bending stress, fℓ1, in a discretely
braced compression flange is required in routine steel I-girder bridges, Eqs. 6.10.1.6-3 and
6.10.1.6-5 apply only when checking bottom-flange lateral bending stresses due to wind load in
the final condition (i.e., at the strength limit state) in regions of negative flexure in beams or girders
designed using the provisions of the optional Appendix A6 (see the Discussion of Appendix A6 in
this Guide); otherwise, Eqs. 6.10.1.6-2 and 6.10.1.6-4 apply.
If the unbraced length, Lb, or distance between the brace points exceeds the limit given by the
applicable Eq. 6.10.1.6-2 or 6.10.1.6-3, which is typically the case, the second-order flange lateral
bending stress may be approximated by amplifying the first-order value by the amplification factor
given by either Eq. 6.10.1.6-4 or 6.10.1.6-5, as applicable. Note that the elastic lateral-torsional
buckling stress, Fcr, is not limited to RbRhFyc in the computation of the amplification factor. Also,
the largest values of fbu or Mu, as applicable, within the unbraced length under consideration are to
be used within each of the above equations. The largest value of fℓ1 within the unbraced length under
consideration is to be used in the computation of the amplification factor when the flexural
resistance is based on lateral-torsional buckling; otherwise, fℓ1 may be determined as the
corresponding value at the section under consideration.
When the amplification factor is large (e.g., greater than about 2.5), the flange is likely too narrow;
increasing the width of the top flange should reduce the amplification factor to a more reasonable
value. Amplification of the first-order lateral bending stresses in discretely braced tension flanges is
not required.
The language in the Commentary for this Article dealing with the amplification of first-order flange
lateral bending stresses in horizontally curved I-girders (along with Figure C6.10.1.6-1) does not
apply to the routine steel I-girder bridges covered by this Guide, which use straight (tangent) girders
by definition.

6.10.1.7 Minimum Negative Flexure Concrete Deck Reinforcement


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 200


Discussion:
Simple Span Bridges:
The minimum one-percent longitudinal reinforcement in the deck specified in this Article is not
required for crack control in simple span routine I-girder bridges because the concrete deck is in
compression; temperature and shrinkage reinforcement is still required however.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This Article deals with the minimum required longitudinal concrete deck reinforcement in regions
of negative flexure for control of deck cracking and is applicable to the routine steel multi-span
continuous I-girder bridges covered by this Guide.
The Article prescribes the total cross-sectional area of the longitudinal reinforcement (including
the temperature and shrinkage reinforcement) that is provided in these regions, the maximum
reinforcing bar size and required yield strength, and the layout, distribution, and maximum spacing
of the reinforcing. The reinforcement is to be distributed across the entire concrete section since
the effective width of concrete deck is close to the entire deck width in most girder bridges. The
use of small bars at relatively close spacing is intended to result in closely spaced cracks of small
width. Applicable Owner-agency standards should always be consulted for any Owner-specific
size, placement, and spacing requirements.
For the computation of girder section properties for the strength limit state checks in regions of
negative flexure in which the longitudinal reinforcement is considered to act with the steel section,
the area of the reinforcement in the two layers within the effective width of the deck over the beam
or girder under consideration can be combined into a single layer placed at the centroid of the two
layers, if desired.
The minimum one-percent longitudinal reinforcement in the deck is to be placed along the span
wherever the longitudinal tensile stress in the concrete deck (see the Discussion of Article
6.10.1.1.1d in this Guide) due to either the factored construction loads (i.e., during the sequential
deck placement) or due to load combination Service II (Table 3.4.1-1) exceeds 0.9fr, where fr is
the modulus rupture of the concrete taken equal to 0.24 fc' for the routine steel I-girder bridges
covered by this Guide. f'c may be taken as the 28-day compressive strength of the concrete or else
a more accurate estimate of the concrete strength at the time the deck casts are made can be used
to compute fr and the modular ratio for this check, if desired.
Many routine steel I-girder bridges will have their decks placed following a sequential deck
placement scheme. In this scenario, the deck is placed in series of longitudinal sections, with the
maximum size of a section typically limited by the maximum volume of concrete that can be placed
at a time. The specific dimensions and limits of each section are also determined to try to
approximately correspond to the points of permanent load contraflexure. The order of placement
of concrete in each section should generally be specified so that positive moment regions are
placed first and negative moment regions are placed last; this helps to minimize the introduction
of tensile stresses in the deck in the negative moment regions. The load factor for investigation of
the deck stress during a sequential deck placement is taken equal to 1.4 (see the Discussion of
Article 3.4.2.1 in this Guide). During the sequential deck placement, when the concrete deck is

NSBA Guide to Navigating Routine Steel Bridge Design / 201


placed in a span adjacent to a span where the concrete has already been placed, negative moment
in the adjacent span causes tensile stresses in the previously placed concrete although these regions
may be subjected primarily to positive flexure in the final condition. This requirement is intended
to help control cracking in the previously placed concrete.
After the deck hardens, the deck can experience significant tensile stresses outside the points of
permanent load contraflexure under moving live loads; the Service II requirement is intended to
help control the deck cracking in such cases under expected severe traffic loadings (see the
Discussion of Article 6.10.4.2 in this Guide for further discussion on the Service II load
combination).
Satisfaction of the above provisions, along with the provision of shear connectors along the entire
length of the span, allows the concrete deck to be considered effective in tension when computing
the composite section properties for negative flexure at the fatigue and/or service limit states if
permitted by the Owner-agency (see the Discussion of Articles 6.6.1.2.1 and 6.10.4.2.1 in this
Guide).
The discussion related to the checking of nominal yielding in the longitudinal reinforcement under
load combination Service II given in the Commentary for this Article does not apply to the routine
steel I-girder bridges covered by this Guide.
The requirement to provide sufficient development length of the longitudinal reinforcement given
in the last paragraph of the specification in this Article is only applicable in cases where shear
connectors are intentionally omitted in regions of negative flexure, which is dependent on the
preferences of the Owner-agency but is not recommended.
Beyond the limits of where the minimum one-percent longitudinal reinforcement in the deck is
required (i.e., in regions of the deck where the tensile stress in the deck is below the above-cited
limits), temperature and shrinkage reinforcement is still required in the deck.
For further discussion and sample calculations of the minimum negative flexure concrete deck
reinforcement, consult Section 6.4.2.3.3.2 of the Reference Manual for NHI Course 130081, Load
and Resistance Factor Design (LRFD) for Highway Bridge Superstructures, as well as FHWA’s
Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous Straight Composite
Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous Straight Composite Steel I-
Girder Bridge, and Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-
Flange Beam Bridge.

6.10.1.8 Tension Flanges with Holes


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides a limit on the maximum factored major-axis bending stress permitted on the
gross section of the beam or girder at the strength limit state or during construction, neglecting the
loss of area due to holes in the tension flange.

NSBA Guide to Navigating Routine Steel Bridge Design / 202


This Article is applicable to routine steel I-girder bridges at bolted field splices. Eq. 6.10.1.8-1 will
prevent a bolted field splice from being located at a section where the factored flexural resistance
of the section at the strength limit state exceeds the moment at first yield, My, unless the factored
stress in the tension flange at that section is limited to the value given by the equation; this may
dictate the location of the bolted field splices in some simple span routine steel bridges and in some
longer-span multi-span continuous bridges, or it may force an increase in sizing of the flanges of
a steel plate girder or an increase in overall sizing of a rolled beam in order to reduce the stresses.
However, this equation typically does not control the design for most multi-span continuous
bridges with reasonably well-balanced spans where the bolted field splices can be located at or
near the points of permanent load contraflexure.
This equation can also control where cross-frame or diaphragm connection plates are attached to
the tension flange by bolting (instead of welding), although the use of a bolted connection in this
situation is not recommended (see the Discussion of Articles 6.6.1.2.3 and 6.6.1.3.1 in this Guide),
and also where lateral bracing members or lateral connection plates are bolted directly to the
tension flange, which is a recommended detail for such members to provide improved fatigue
performance. However, the routine steel I-girder bridges covered by this Guide are assumed not
to contain any lateral bracing.

6.10.1.9 Web Bend-Buckling Resistance

6.10.1.9.1 Webs without Longitudinal Stiffeners


Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.
Discussion:
Simple Span Bridges and Multi-span Continuous Plate Girder Bridges:
The computation of the theoretical web bend-buckling resistance, Fcrw, is only applicable during
construction for simple span routine I-girder bridges utilizing slender web sections in the
noncomposite condition. Therefore, these provisions are conditionally applicable for the simple
span routine bridges covered by this Guide; that is, the provisions are applicable during
construction in the noncomposite condition for simple span bridges utilizing steel plate girders
with slender webs, but are not applicable for simple span bridges utilizing steel plate girders with
nonslender webs or for simple span bridges utilizing rolled beams (assuming all sections qualify
as compact web or noncompact web sections in the noncomposite condition during construction);
the reasons for this are discussed further below. See the Commentary for Article 6.10.6.2.3 and
the Discussion of Article 6.10.6.2.3 in this Guide for further discussion on the definition and
categorization of compact web, noncompact web, and slender web sections.
The computation of Fcrw is only applicable for routine multi-span continuous plate girder bridges
utilizing slender web sections in the noncomposite condition during construction, and/or utilizing
slender web sections in regions of negative flexure at the service limit state. In the computation of

NSBA Guide to Navigating Routine Steel Bridge Design / 203


Fcrw for composite sections in negative flexure at the service limit state, when the necessary
conditions are satisfied such that the concrete deck may be considered effective in tension (see the
Discussion of Article 6.10.4.2.1 in this Guide), the depth of the web in compression in the elastic
range, Dc, must be computed using Eq. D6.3.1-1 in Appendix D6 (see the Discussion of Article
D6.3.1 in this Guide). Otherwise, Dc is to be computed using the steel section alone or the steel
section plus the longitudinal reinforcement depending on the preferences of the Owner-agency.
Therefore, these provisions are conditionally applicable for the routine multi-span continuous plate
girder bridges covered by this Guide.
This Article presents the provisions necessary to compute Fcrw for webs without longitudinal
stiffeners, which is used as a simple index to control strains and transverse displacements in the
compression zone of slender-web girders during construction (see the Discussion of Article
6.10.3.2.1 in this Guide) and in regions of negative flexure at the service limit state (see the
Discussion of Article 6.10.4.2.2 in this Guide). Fcrw is not needed at the fatigue limit state because
the web bend-buckling check at the service limit state will always control.
The advent of composite design has led to a significant reduction in the size of compression flanges
in regions of positive flexure. As a result, more than half of the web of the noncomposite section
will be in compression in these regions during the construction condition before the concrete deck
has hardened. Slender web sections are more susceptible to bend-buckling in this condition. At the
service limit state, a control on the amount of transverse web displacement is also desirable. In a
multi-span continuous girder at the service limit state, slender web sections in regions of negative
flexure are most susceptible to web bend-buckling, especially for composite sections when the
necessary conditions are satisfied such that the concrete deck may be considered effective in
tension (see the Discussion of Article 6.10.4.2.1 in this Guide).
For a compact web section (e.g., a rolled-beam section) or a noncompact web section, Fcrw will
always equal or exceed Fyc, where Fyc is the specified minimum yield strength of the compression
flange; therefore, theoretical web bend-buckling in these sections will not occur for elastic stress
levels, computed according to beam theory, smaller than the limit of their flexural resistance.
Therefore, for these sections, the web bend-buckling checks described above need not be made.
The web bend-buckling checks also need not be made for sections in regions of positive flexure at
the service limit state in the routine I-girder bridges covered by this Guide (see the Discussion of
Article 6.10.4.2.2 in this Guide); however, these regions must be checked for web bend-buckling
during construction in routine plate-girder bridges utilizing slender web sections in the
noncomposite condition.
Fcrw is to be checked against the maximum factored compression flange major-axis bending stress.
Utilizing the maximum compressive stress in the web rather than the stress in the compression
flange to obtain greater precision is not warranted for this check.
The bend-bucking coefficient of 7.2 mentioned at the end of this Article is only applicable around
points of permanent-load contraflexure in multi-span continuous bridges and applies to rare cases
where both the top and bottom edges of the web are subject to small accumulated compressive
stresses.

NSBA Guide to Navigating Routine Steel Bridge Design / 204


For further discussion and sample calculations of Fcrw, consult Section 6.4.5.5 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures.
Multi-span Continuous Rolled Beam Bridges: the computation of Fcrw is not applicable for routine
multi-span continuous rolled beam bridges during construction or at the service limit state
(assuming all sections qualify as compact web or noncompact web sections in the noncomposite
condition during construction and in regions of negative flexure at the service limit state, which
should typically be the case).

6.10.1.9.2 Webs with Longitudinal Stiffeners


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article presents the provisions necessary to compute the theoretical web bend-buckling
resistance, Fcrw, for webs with longitudinal stiffeners. These provisions are not applicable to the
routine steel I-girder bridges covered by this Guide, which do not utilize web longitudinal
stiffeners.

6.10.1.10 Flange-Strength Reduction Factors

6.10.1.10.1 Hybrid Factor, Rh


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article presents provisions to compute the hybrid factor, Rh, which is a flange-stress reduction
factor that accounts for the redistribution of stress from the web to both flanges resulting from
local yielding of the web in hybrid-section members. This Article is not applicable to the routine
steel I-girder bridges covered by this Guide, which do not utilize hybrid-section members;
therefore, Rh should always be taken equal to 1.0 for these bridges.

6.10.1.10.2 Web Load-Shedding Factor, Rb


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.
Discussion:
The Article presents provisions to compute the web load-shedding factor, Rb, which is a post-
buckling flange-stress reduction factor that accounts for the nonlinear variation of stresses
subsequent to local bend-buckling of slender webs at the strength limit state. The factor accounts
for the reduction in the section flexural resistance caused by the shedding of the compressive
stresses in the web resulting from local bend-buckling of a slender web at the strength limit state

NSBA Guide to Navigating Routine Steel Bridge Design / 205


and the corresponding increase in the flexural stress within the compression flange. The Rb factor
is not applied in determining the nominal flexural resistance of the tension flange at the strength
limit state since the tension flange stress is not increased significantly by the shedding of the web
compressive stresses.
Simple Span Bridges:
Web bend-buckling is not a consideration for compact web sections (e.g., rolled beam sections) or
noncompact web sections at any limit state (see the Discussion of Article 6.10.1.9.1 in this Guide).
Web bend-buckling is explicitly prevented during construction for plate girder sections with
slender webs (see the Discussion of Article 6.10.3.2.1 in this Guide) and is not a consideration for
composite plate girder sections in regions of positive flexure without web longitudinal stiffeners
at the service or strength limit states. Therefore, load-shedding from the web to the compression
flange does not theoretically occur in either case and these provisions are not applicable for the
routine simple span I-girder bridges covered by this Guide (i.e., Rb is always taken equal to 1.0).
See the Commentary for Article 6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in this Guide
for further discussion on the definition and categorization of compact web, noncompact web, and
slender web sections.
Multi-span Continuous Rolled Beam Bridges:
Web bend-buckling is not a consideration for compact web sections (e.g., rolled beam sections) in
any regions of the beam at any limit state (see the Discussion of Article 6.10.1.9.1 in this Guide);
therefore, load-shedding from the web to the compression flange does not theoretically occur and
these provisions are not applicable for the routine multi-span continuous rolled beam bridges
covered by this Guide (i.e., Rb is always taken equal to 1.0).
See the Commentary for Article 6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in this Guide
for further discussion on the definition and categorization of compact web, noncompact web, and
slender web sections.
Multi-span Continuous Plate Girder Bridges:
Web bend-buckling is not a consideration for compact web plate girder sections or noncompact
web plate girder sections at any limit state (see the Discussion of Article 6.10.1.9.1 in this Guide).
Web bend-buckling is explicitly prevented during construction for plate girder sections with
slender webs (see the Discussion of Article 6.10.3.2.1 in this Guide) and is not a consideration for
composite plate girder sections in regions of positive flexure without web longitudinal stiffeners
at the service or strength limit states. Thus, the Rb factor will be less than 1.0 only for slender web
sections in negative flexure at the strength limit state and therefore these provisions are
conditionally applicable for the routine multi-span continuous plate girder bridges covered by this
Guide.
See the Commentary for Article 6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in this Guide
for further discussion on the definition and categorization of compact web, noncompact web, and
slender web sections.

NSBA Guide to Navigating Routine Steel Bridge Design / 206


For the routine multi-span continuous plate girder bridges covered by this Guide, in cases where
the Rb factor is less than 1.0 and must be computed (indicating that web bend-buckling has
theoretically occurred at the strength limit state), only Eqs. 6.10.1.10.2-3, 6.10.1.10.2-5, and
6.10.1.10.2-8 are applicable. The other equations in this Article are only applicable to sections
with web longitudinal stiffeners and the routine steel I-girder bridges covered by this Guide do not
contain web longitudinal stiffeners. Also, the Rb factor must only be computed at the strength limit
state for sections in regions of negative flexure that do not satisfy Eq. 6.10.1.10.2-1 (i.e., slender
web sections); otherwise, Rb is taken equal to 1.0.
For further discussion and sample calculations of the Rb factor, consult Section 6.4.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures.

6.10.2 Cross-Section Proportion Limits

6.10.2.1 Web Proportions

6.10.2.1.1 Webs without Longitudinal Stiffeners


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides web proportioning limits for sections without web longitudinal stiffeners and
is applicable to the routine steel I-girder bridges covered by this Guide.
By limiting the web slenderness of girders without longitudinal stiffeners to 150 as specified in
Eq. 6.10.2.1.1-1, transverse stiffeners need only be provided for shear in routine plate-girder
bridges and can potentially be spaced up to the maximum specified limit of 3D (i.e., three times
the web depth).
The AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to Design for
Constructability and Fabrication provides practical guidance for economical proportioning of
girder webs, such as the general preference of fabricators for a minimum web thickness of ½" to
reduce the deformation of the web and the potential for weld defects during fabrication.
Changes in the web thickness along the girder in plate-girder bridges preferably should be made
at field splices. In field sections over interior piers in continuous spans, the web thickness may
have to be increased (typically in 1/16-inch increments) over the thickness provided in adjacent
regions in positive flexure in some instances; e.g., if the web bend-buckling resistance is exceeded
in regions of negative flexure at the service limit state (see the Discussion of Articles 6.10.1.9.1
and 6.10.4.2.2 in this Guide).
A useful guideline for determining the trade-off between adding more transverse stiffeners versus
increasing the thickness of the web material in routine plate-girder bridges is that approximately 4
to 5 pounds of web material should be saved for every 1 pound of stiffener material added. This
higher unit cost reflects that additional fabrication effort is required per pound of stiffener still than
per pound of girder web steel. Generally, an unstiffened web is not the most economical alternative
for a plate-girder bridge. The best solution usually includes some transverse stiffeners over the

NSBA Guide to Navigating Routine Steel Bridge Design / 207


piers and near the abutments; a so-called partially stiffened web. Transverse stiffeners (other than
connection plates for cross-frames or diaphragms) are typically not required in routine rolled-beam
bridges.
The web depth of a plate girder dictates the flange sizes. In the absence of depth restrictions or
significantly unbalanced spans, the web depth that is selected should be near the optimum web
depth for the largest span in the unit, which is the web depth that provides the minimum cost girder.
In multi-span continuous bridges, the optimum web depth for the regions of negative flexure is
often not the same as that for the regions in positive flexure. Usually the optimum depth for the
regions of positive flexure is a better choice when combined with heavier flanges in the shorter
regions of negative flexure. Where a deeper web in the regions of positive flexure requires smaller
flanges, this may lead to stability issues during shipping and erection. A compromise depth is
usually necessary.
The optimum web depth for a composite girder is elusive since loads are applied to different
sections; there is no single algorithm that gives the optimum web depth. Instead, the optimum web
depth is best established by preparing a series of designs with different web depths to arrive at an
optimum cost-effective depth based on weight and/or cost. The NSBA's LRFD Simon line-girder
analysis and design program contains a useful web-depth optimization option that automatically
generates a series of trial-design input files from an acceptable starting design input file; the
generated input files differ from the starting input only in the vertical web depth. Each of these
input files is processed automatically by the analysis engine, and a table is prepared listing the
depth, weight, and cost (based on user-input cost factors) for the depths that have acceptable
designs. Users should verify the capabilities, assumptions, and general correctness of any
program’s calculations prior to initial use.
For further information on girder depth and I-girder sizing and proportioning, consult Sections
6.3.3.2 and 6.3.4 of the Reference Manual for NHI Course 130081, Load and Resistance Factor
Design (LRFD) for Highway Bridge Superstructures.

6.10.2.1.2 Webs with Longitudinal Stiffeners


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides web proportioning limits for sections with web longitudinal stiffeners and is
not applicable to the routine steel I-girder bridges as defined for the purposes of this Guide since
these bridges do not utilize web longitudinal stiffeners.

6.10.2.2 Flange Proportions


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
Sizing of flanges is one of the most important issues in obtaining an economical steel plate-girder
bridge. This Article provides four separate flange proportioning limits that are applicable to the
routine steel I-girder bridges covered by this Guide. The reasoning behind each of these limits is

NSBA Guide to Navigating Routine Steel Bridge Design / 208


discussed in the Commentary. The limits apply to both the compression and tension flanges. Eq.
6.10.2.2-1 in particular provides an upper bound limit for flange slenderness, and should not
necessarily be used as an indication of an economical design. Eq. 6.10.2.2-2 provides a lower
bound limit on the flange width; a larger flange width will often be required.
Fabricators prefer that plate-girder flange widths not be less than 12 inches to avoid distortion and
cupping of the flanges during welding, which sets a practical lower limit on the width. According
to the AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to Design
for Constructability and Fabrication, fabricators prefer that flange thicknesses not be less than ¾
inches.
Meeting the top flange Lfs/85 minimum width guideline presented in Eq. C6.10.2.2-1 tends to result
in individual girder field sections that are more stable and easier to handle during lifting, erection,
and shipping without the need for special stiffening trusses or falsework. This parameter should
be checked in conjunction with the flange proportioning requirements of this Article to establish a
minimum required top-flange width for each individual unspliced girder field section in the bridge.
Efficient location of flange thickness transitions at shop-welded splices in plate-girder flanges is a
matter of plate length availability and the economics of welding and inspecting a splice compared
to the cost of extending a thicker plate. The design plans should consider allowing the option to
eliminate or move a shop splice by extending a thicker flange plate, subject to the approval of the
Engineer. When evaluating such a request, the Engineer should consider the effect of the thicker
plate on the girder deflections and stresses. Usually, a change in plate length does not significantly
affect the deflections as much as the removal of a welded splice.
Parameters affecting the cost of shop-welded splices vary from shop to shop. Table 1.5.4-1 in the
AASHTO-NSBA Steel Bridge Collaboration G12.1-2020 Guidelines to Design for
Constructability and Fabrication gives suggested weight savings per inch of flange width to help
evaluate placement of shop splices. Usually, somewhere between 800 to 1,200 pounds of material
must be saved to justify the introduction of a welded shop splice. This may vary so consider
consulting the fabricator regarding this issue whenever possible. Pricing considerations dictate that
optimal ordered plate lengths are usually less than or equal to 80 feet. This is also the length that
usually fits on a single railroad flat car. Longer plates may of course be used as necessary. Table
1.4.1-1 of the AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to
Design for Constructability and Fabrication provides a table of maximum plate length availabilities
for ASTM A 709 Grades 36, 50, and 50W plates. Table 1.4.2-1 provides maximum wide flange
beam length availability.
In typical cases, providing more than two shop splices (i.e., three different flange thicknesses) in
any one field section of a plate girder is not economical, except in relatively rare cases where the
girders are unusually heavy or plate length availability limits dictate the need for additional splices
with or without a thickness change. At flange shop splices, the area of the thinner plate should not
be less than one-half the area of the thicker plate to reduce the stress concentration and produce a
smoother transition of stress across the splice.
As a practical matter, fabricators typically order plate for flange material from the mills in widths
60 inches and above; typically, the most economical plate size to buy from a mill is between 72

NSBA Guide to Navigating Routine Steel Bridge Design / 209


and 96 inches. Thus, consider sizing flanges so that as many pieces as possible can be obtained
from a wide plate with minimal waste. To minimize waste, it is also important to limit the number
of different flange plate thicknesses specified for a given project. Larger order quantities of plate
cost less and minimizing the number of different thicknesses simplifies fabrication and inspection
and reduces mill quantity extras. Also, it is preferred to select flange thicknesses in at least 1/8
inch increments up to 2 ½ inches and in ¼ inch increments over 2 ½ inches, and to limit flange
thicknesses to 3 inches or below if possible.
Flange widths for an individual plate girder should be kept constant within each field section; i.e.,
avoid changing flange widths at welded shop splices. Change the flange widths at a bolted field
splice instead. There is little need to maintain a constant flange width between individual field
sections. However, some Owners may prefer a constant-width bottom flange along the entire
length of the girder for aesthetic reasons if pedestrians or vehicles are expected to pass underneath
the bridge. Note that top and bottom flange widths within a field section can be, and often are,
different.
For further information on sizing flanges for efficient fabrication, consult Section 1.5 of the
AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to Design for
Constructability and Fabrication and Section 6.3.4.4.5 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.

6.10.3 Constructibility

6.10.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article cover the required design checks for constructibility, which apply to
the routine steel I-girder bridges covered by this Guide.
For routine steel I-girder design, the constructibility checks apply to the timeframe prior to when
the concrete deck is cured (i.e., the final structural condition).
The standard of care in many jurisdictions is that the designer need only perform a non-structural
review of the conceptual erection sequence for the structural steel framing, primarily to
demonstrate that a viable erection scheme exists (i.e., an erection sequence that is feasible given
the known site conditions and constraints, specified maintenance-of-traffic sequence and
requirements, etc.), including consideration of the location of shoring towers, lifting and holding
cranes, etc. Owner-agencies in these jurisdictions expect detailed erection engineering to be
performed by the Contractor’s engineer, not by the bridge’s designer. However, several Owner-
agencies do require that the designer perform some level of detailed erection engineering. Review
local Owner-agency design policies and construction specifications and the local standard of care
to determine the requirements in any given specific jurisdiction. Note that the performance of
detailed erection engineering is beyond the scope of this Guide.
However, once the structural steel framing system is fully erected, the designer clearly has
responsibility for checking that the structural steel has sufficient strength and stiffness to resist

NSBA Guide to Navigating Routine Steel Bridge Design / 210


construction loads. The provisions outlined in this Article generally apply to the checking of the
combined effects of the component dead loads (DC1 loads) acting on the bare steel girders during
construction, including consideration of the deck-placement sequence for multi-span continuous
bridges, the deck overhang loads, as applicable (see the Discussion of Article 6.10.3.4.1 in this
Guide for more information on deck overhang loads), construction live loads (due to the presence
of construction workers, a deck screed machine, and possibly other equipment), and wind loads.
These checks will typically govern the size of the top flange and may influence the cross-frame or
diaphragm spacing, in regions of positive flexure. The basic philosophy behind these provisions is
that no nominal yielding and no reliance on post-buckling resistance is permitted for main load-
carrying members during construction. These checks are necessary since girders in the positive
moment region have significantly less load-carrying capacity in their noncomposite condition, and
many of the loading conditions which occur during construction are not considered in the
evaluation of the bridge in its final, completed condition.
It should be noted that some of the loads which occur during construction, such as deck overhang
loads and wind loads, induce flange lateral bending stresses. Most line girder analysis programs do
not directly calculate flange lateral bending stresses. Line girder analysis programs use one-
dimensional analysis models which cannot directly address the torsional loading effects which
produce flange lateral bending. Some programs will allow the user to manually input flange lateral
bending moments or stresses, calculated by the user outside the program; other programs may not
have this ability. In many cases, it may be easiest to evaluate the effects of flange lateral bending on
the design outside of a line girder analysis program using hand calculations, perhaps facilitated by
programming these calculations in a spreadsheet. This allows the designer flexibility, particularly in
addressing the various constructibility checks which must be performed. If it turns out that
consideration of the effects of flange lateral bending, whether in a constructibility check, or in the
supplemental checks of the bottom flange under final conditions, results in a controlling design
condition, the designer may have to rerun the line girder analysis with resized girders to update
major-axis bending stress calculations, deflection calculations, or other calculations.
Load combinations for the constructibility design checks include Strength I, Strength III, and a
“Special” load combination for primary steel superstructure components specified in Article
3.4.2.1. Owners may have explicit or implicit policies which prescribe specific loads, load
combinations, and other assumptions for constructibility checks which supplement or take the
place of those presented in the AASHTO LRFD BDS.
In general, for Strength I, Article 3.4.2.1 prescribes various load factors, including a load factor of
1.5 to be applied to construction loads. Wind is not included in the Strength I load combination.
Strength III is for dead load in combination with wind load; often this limit state is used to represent
a case of inactive construction, where the structural steel is erected and in place but no construction
activity is occurring. This might model an overnight or weekend condition when construction is
not occurring. For this load combination, dead load includes the self-weight of the structural steel
and perhaps some various construction dead loads in place, depending on Owner-agency policy
and what might be permitted in the Owner-agencies standard specifications or in the contract
documents. Typically, this load combination includes no construction live load, but does include
a high wind load. Article 3.4.2.1 states that the load factor on the wind load for the Strength III

NSBA Guide to Navigating Routine Steel Bridge Design / 211


load combination during construction is to be specified by the Owner-agency. If the Owner-agency
does not provide guidance, the AASHTO Guide Specifications for Wind Loads on Bridges During
Construction can be consulted. The load factor on any construction dead loads that are included in
this load combination with the self-weight of the structural steel is not to be less than 1.25.
The “Special” load combination prescribed in Article 3.4.2.1 typically represents a case of active
construction. For this load combination the dead load includes the self-weight of the structural
steel, the dead load of the wet concrete deck (including consideration of the prescribed deck-pour
sequence), and construction dead loads such as formwork, falsework, etc. The live load consists
of construction live load, including consideration of both construction equipment and construction
personnel using a load factor of 1.4 for DC (component) dead loads and any construction loads
acting on the fully erected steelwork, including dynamic load effects (if applicable). This load
combination typically includes a reduced wind load or no wind load, depending on Owner-agency
policy.
Construction loads or loads that act on the structure only during construction loads include, but are
not limited to, the weight of materials, removable forms, personnel, and equipment such as deck
finishing machines or loads applied to the structure through falsework or other temporary supports.
The weight of the wet concrete deck and any stay-in-place forms should be considered as DC
loads.
Failure modes of concern during construction of flexural members in the routine steel I-girder
bridges covered by this Guide include nominal yielding, local buckling, lateral torsional buckling,
web bend buckling, and/or shear buckling. All constructibility design checks are to be made on
the noncomposite steel girder and are typically critical only in regions of positive flexure.
The potential for uplift at bearings during construction should be considered, but is generally not
a concern for the routine steel I-girder bridges covered by this Guide, except possibly for multi-
span continuous bridges with short end spans or other issues related to poor span balance. Also,
the provisions of Article D6.5 need only be checked for routine rolled-beam bridges at bearings.
For additional information on the design checks for constructibility, consult the FHWA’s Steel
Bridge Design Handbook, Volume 11 - Design for Constructability. For design examples
illustrating constructibility design computations, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge, and
Sections 3.3 and 6.5.3 of the Reference Manual for NHI Course 130081, Load and Resistance
Factor Design (LRFD) for Highway Bridge Superstructures. Also see the Discussion of Article
3.4.2 in this Guide for more detailed information on the required loads and load combinations to
consider for the constructibility checks. The Reference Manual for NHI Course 130102,
Engineering for Structural Stability in Bridge Construction provides further discussion of
constructibility checks and the associated loads.

NSBA Guide to Navigating Routine Steel Bridge Design / 212


6.10.3.2 Flexure

6.10.3.2.1 Discretely Braced Flanges in Compression


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article contains three equations to check the flexural resistance of the discretely braced top
(compression) flange of the bare steel section in regions of positive flexure for the combined
effects of either the DC1 loads (for simple span bridges only) or the deck-placement sequence (for
multi-span continuous bridges only) and the deck overhang loads, as applicable, during the
construction of the concrete deck. fbu is taken as the factored major-axis compressive stress in the
top flange due to the moment acting on the noncomposite section at the section under
consideration. When investigating the deck-placement sequence (see the Discussion of Article
6.10.3.4.1 in this Guide for more information on the deck placement sequence), the moment used
to calculate fbu should be the maximum accumulated moment occurring on the noncomposite
section only during the placement sequence. For the exterior (fascia) girders, fℓ is taken as the
factored flange lateral bending stress in the top flange due to the deck overhang loads at the section
under consideration (see the Discussion of Article 6.10.3.4.1 in this Guide for more information
on deck overhang loads); the flange lateral bending stress in this case will often be subject to
amplification (see the Discussion of Article 6.10.1.6 in this Guide). Each of the three equations in
this Article are applicable to the routine steel I-girder bridges covered by this Guide.
When checking these equations with the section in its noncomposite condition, the categorization of
the web is to be based on the properties of the noncomposite section. See the Commentary for Article
6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in this Guide for further discussion on the
definition and categorization of the compact web, noncompact web, and slender web sections
mentioned below.
Eq. 6.10.3.2.1-1 is a yielding limit state check. This equation does not need to be checked for
interior girders utilizing slender web sections (where fℓ due to the deck overhang brackets is equal
to zero) as this equation will not control.
Eq. 6.10.3.2.1-2 checks that the member has sufficient strength with respect to flange local
buckling (see the Discussion of Article 6.10.8.2.2 or A6.3.2 in this Guide, as applicable) and
lateral-torsional buckling (see the Discussion of Article 6.10.8.2.3 or A6.3.3 in this Guide, as
applicable). In computing the flange local buckling and lateral-torsional buckling resistances, the
web load-shedding factor, Rb, (see the Discussion of Article 6.10.1.10.2 in this Guide) is taken
equal to 1.0 since web bend-buckling is explicitly prevented during construction via Eq. 6.10.3.2.1-
3 (see below). When lateral-torsional buckling controls, the largest values of fbu and fℓ, as
applicable, within the unbraced length must be used to check Eq. 6.10.3.2.1-2 (see the Discussion
of Article 6.10.1.6 in this Guide). For sections with compact or noncompact webs, the lateral-
torsional buckling resistance for use in Eq. 6.10.3.2.1-2 may optionally be computed according to
the provisions of Article A6.3.3 (to include the beneficial contribution of the St. Venant torsional
stiffness, J), which is recommended in particular for routine rolled-beam bridges with larger
unbraced lengths during construction (see the Discussion of Articles 6.10.6.2.3 and A6.3.3 in this

NSBA Guide to Navigating Routine Steel Bridge Design / 213


Guide). The resulting lateral-torsional buckling resistance, Mnc, in this case is divided by Sxc as
defined in this Article to express the resistance in terms of stress for direct application in Eq.
6.10.3.2.1-2. In some cases, the calculated resistance may exceed Fyc. However, Eq. 6.10.3.2.1-1
will control in such cases, or perhaps Eq. 6.10.1.8-1 if there are holes in the tension flange at the
section under consideration (see the Discussion of Article 6.10.1.8 in this Guide), thus ensuring
that the combined factored stress in the flange will not exceed Fyc during construction.
Eq. 6.10.3.2.1-3 checks that theoretical web bend-buckling (Article 6.10.1.9.1) will not occur
during construction. This equation does not need to be checked for sections with compact or
noncompact webs (e.g., sections in routine rolled-beam bridges). Options to consider should the
web bend-buckling stress be exceeded are discussed in the last paragraph of the Commentary for
this Article.
For routine multi-span continuous bridges, these three equations should also be used to check the
discretely braced bottom (compression) flange of the bare steel section in regions of negative
flexure; however, the checks given by the first two equations typically do not control.
As explained in the Discussion of Article 6.10.3.1 in this Guide, designers are reminded that most
commercial steel bridge line girder analysis and design software packages are not capable of
calculating all of the design stresses associated with the constructibility checks. Generally, the
constructibility checks are performed outside of the line girder analysis and design program, using
hand calculations, perhaps facilitated by programming these calculations in a spreadsheet.

6.10.3.2.2 Discretely Braced Flanges in Tension


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article contains an equation to check the flexural resistance of the discretely braced bottom
(tension) flange of the bare steel section in regions of positive flexure for the combined effects of
either the DC1 loads (for simple span bridges only) or the deck-placement sequence (for multi-
span continuous bridges only) and the deck overhang loads, as applicable, during the construction
of the concrete deck. fbu is taken as the factored major-axis compressive stress in the bottom flange
due to the moment acting on the noncomposite section at the section under consideration. When
investigating the deck-placement sequence (see the Discussion of Article 6.10.3.4.1 in this Guide
for more information on the deck-placement sequence), the moment used to calculate fbu should
be the maximum accumulated moment occurring on the noncomposite section only during the
placement sequence. For the exterior (fascia) girders, fℓ is taken as the factored flange lateral
bending stress in the bottom flange due to wind loads and the deck overhang loads at the section
under consideration (see the Discussion of Article 6.10.3.4.1 in this Guide for more information
on deck overhang loads); the flange lateral bending stress in this case is not subject to amplification
(see the Discussion of Article 6.10.1.6 in this Guide) since the flange is in tension. The equation
in this Article, which is a yielding limit state check, is applicable to the routine steel I-girder
bridges covered by this Guide but typically does not control.

NSBA Guide to Navigating Routine Steel Bridge Design / 214


For routine multi-span continuous bridges, this equation should also be checked for the discretely
braced top (tension) flange of the bare steel section in regions of negative flexure; however, this
check typically does not control.
As explained in the Discussion of Article 6.10.3.1 in this Guide, designers are reminded that most
commercial steel bridge line girder analysis and design software packages are not capable of
calculating all of the design stresses associated with the constructibility checks. Generally, the
constructibility checks are performed outside of the line girder analysis and design program, using
hand calculations, perhaps facilitated by programming these calculations in a spreadsheet.

6.10.3.2.3 Continuously Braced Flanges in Tension or Compression


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article contains an equation to check the flexural resistance of top flange of the beam or
girder subject to tension or compression for any factored construction loads that may be applied
after the deck has hardened and is continuously braced. A continuously braced flange is defined
as a flange that is encased in concrete or anchored to the concrete deck by shear connectors
satisfying the provisions of Article 6.10.10. In some multi-span continuous routine steel I-girder
bridges, Owner-agency policy may prescribe that shear connectors not be provided in regions of
negative flexure; in those regions the provisions of this Article would not apply.
Flange lateral bending need not be considered in a continuously braced flange. The lateral
resistance of the concrete deck is generally adequate to compensate for the neglect of any initial
lateral bending stresses in the steel prior to placement of the deck and any additional lateral bending
stresses induced after the deck has been placed. Flange local buckling and lateral-torsional
buckling also need not be considered when a continuously braced flange is subject to compression.
As explained in the Discussion of Article 6.10.3.1 in this Guide, designers are reminded that most
commercial steel bridge line girder analysis and design software packages are not capable of
calculating all of the design stresses associated with the constructibility checks. Generally, the
constructibility checks are performed outside of the line girder analysis and design program, using
hand calculations, perhaps facilitated by programming these calculations in a spreadsheet.

6.10.3.2.4 Concrete Deck


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
This Article requires that a minimum area of longitudinal reinforcement equal to at least one-
percent of the concrete deck cross-sectional area is provided in multi-span continuous bridges
wherever the factored tensile stress in the deck (see the Discussion of Article 6.10.1.1.1d in this
Guide) during the deck-placement sequence exceeds 0.9fr, where fr is the modulus rupture of the

NSBA Guide to Navigating Routine Steel Bridge Design / 215


concrete taken equal to 0.24 fc' for the routine steel I-girder bridges covered by this Guide. See the
Discussion of Article 6.10.1.7 in this Guide for further discussion of this provision.
Simple Span Bridges:
The minimum one-percent longitudinal reinforcement in the concrete deck specified in this Article
is not required for crack control in routine simple span I-girder bridges because the deck is in
compression; hence, this Article is not applicable for the design of simple span routine steel I-
girder bridges. Temperature and shrinkage reinforcement is still required in the decks of these
types of bridges however.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
The provisions of this Article are applicable to the routine steel multi-span continuous I-girder
bridges covered by this Guide.

6.10.3.3 Shear
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides an equation to check the web for the sum of the factored permanent loads
and factored construction loads applied to the noncomposite section during construction. The
equation in this Article is applicable to the routine steel I-girder bridges covered by this Guide.
The factored shear resistance for this check is limited to the shear-yielding or shear-buckling
resistance (see the Discussion of Article 6.10.9 in this Guide). The use of post-buckling tension-
field action is not permitted to resist construction loads. Use of tension-field action is permitted at
the strength limit state after the deck has hardened or is made composite (if the section along the
entire web panel is proportioned according to the requirements for developing tension-field action
discussed in Article 6.10.9).
The shear in the end panels of stiffened webs is already limited to either the shear-yielding or
shear-buckling resistance, as is the shear in unstiffened webs. Therefore, this requirement typically
does not need to be checked for unstiffened webs or for the end panels of stiffened webs.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Since the shear in the web is typically not affected by lateral and
torsional loading effects such as wind load and overhang bracket loading, these types of programs
may be able to directly perform the shear constructibility checks specified by this Article. Users
should verify the capabilities, assumptions, and general correctness of any program’s calculations
prior to initial use.

NSBA Guide to Navigating Routine Steel Bridge Design / 216


6.10.3.4 Deck Placement

6.10.3.4.1 General
Determination of applicability, Simple Span Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
The provisions of this Article are applicable to the multi-span continuous routine steel I-girder
bridges covered by this Guide, while some of the provisions specifically related to the effects of
sequential deck placement on the behavior of multi-span continuous bridges may not be directly
applicable to simple-span routine steel I-girder bridges.
Consideration of Deck Placement Sequence
A sequential deck placement analysis need not be performed for simple-span bridges as it will not
control over the case assuming the entire deck is placed at once. For multi-span continuous bridges,
on the other hand, the effects of the deck placement sequence must be considered.
The provisions require that sections in positive flexure in multi-span continuous bridges that are
composite in the final condition, but noncomposite during construction, be investigated during the
various stages of the deck placement for a specified deck placement sequence shown in the contract
documents. This Article refers to Article 6.10.3.2 for the checking of the bare steel girder in regions
of positive flexure only for the effects of the placement sequence, and for the exterior girder, the
effect of the deck overhang loads (see the Discussion of Article 6.10.3.2 in this Guide). The bare
steel girder should be checked for the maximum accumulated moment acting on the noncomposite
section only during the placement sequence.
Changes in load, stiffness and bracing during the various stages of the deck placement in multi-
span continuous bridges must be considered. During deck placement, the actual composite
stiffness depends on the amount of time that the concrete has had to cure before the next portion
is cast, but such refinements are usually not considered in the analysis. Unless a retarder is used,
concrete usually obtains composite action in a matter of hours after placement. Thus, the full
composite stiffness is often used for the previously placed concrete.
Common practice when casts include both positive and negative moment regions is to cast the slab
in the positive moment regions first, and then cast the slab in the negative bending region over the
support in order to minimize cracking at the top of the slab. However, when concrete is cast in a
span adjacent to a span that already has a hardened deck, induced negative moments in the adjacent
spans will cause tensile stresses in the cured concrete that may result in transverse deck cracking.
Provision of the minimum required one-percent longitudinal reinforcement in the deck at these
sections can help control the cracking (see the Discussion of Articles 6.10.3.2.4 and 6.10.1.7 in
this Guide). In a long cast, e.g. extending from one end of the bridge over an interior support into
an adjacent span, it is possible that the concrete in the negative moment region over the support
will harden and be subject to tensile stresses during the remainder of the cast, which may result in

NSBA Guide to Navigating Routine Steel Bridge Design / 217


early age cracking of the deck. A retarder admixture may be required in the casts over the piers to
reduce the potential for early age deck cracking. In such cases, the end span must still be checked
for the critical instantaneous unbalanced case where wet concrete exists over the entire end span,
with no concrete yet on the remaining spans.
Temporary dead load deflections during the sequential deck placement can also be different from
the final noncomposite dead load deflections. If the differences are deemed significant, this should
be considered when establishing camber requirements.
When computing deflections considering staged deck placement, the stiffness of previously cast
portions of the concrete deck can potentially be based on a modular ratio closer to the short-term
modular ratio, n, since the concrete does not have enough time to creep significantly between casts.
Considering this effect increases the complexity of the analysis and should only be undertaken
when required by Owner-agency policy or when the designer deems that the nature of the
deflections truly warrants this level of refinement. At least one State DOT has found the use of a
concrete modulus of elasticity equal to 70 percent of the modulus of elasticity at 28 days (which
results in a modular ratio of approximately 1.4n for transforming the section) to be appropriate in
computing the stiffness.
The NSBA's LRFD Simon line-girder analysis and design program, along with other available
commercial software, generally have options available to perform a sequential deck placement
analysis as described above. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.
Investigation of Uplift During Deck Placement
Potential uplift at the bearings should also be investigated during the deck placement, although
uplift is generally not a concern for the routine steel I-girder bridges covered by this Guide, except
possibly for multi-span continuous bridges with short end spans or other issues related to poor
span balance. In cases where the potential for uplift may be a concern. It is suggested in the absence
of other Owner-agency guidance that a load factor of 1.0 be applied to all downward support
reactions caused by component dead loads contributing to uplift, and a load factor of 0.9 be applied
to all upward support reactions caused by component dead loads resisting uplift. This investigation
of potential for uplift can usually focus on critical construction stages and can typically be
accomplished using hand calculated modifications of the bearing reactions reported by the line
girder analysis program.
Torsion Caused by Deck Overhang Brackets
The effect of the torsion due to the forces from deck overhang brackets acting on the exterior
(fascia) girders must also be considered for both the routine multi-span continuous and simple-
span I-girder bridges covered by this Guide. The deck overhang weight is resisted by the brackets.
If the bracket is assumed to extend near the edge of the deck overhang, it can be assumed that half
of the deck overhang weight is placed on the fascia girder and half is placed on the overhang
brackets. One-half of the deck haunch weights can also conservatively be included in the total
overhang weight. Besides the weight of the deck, typical loads that may act on the overhang only
during construction include the overhang deck forms, the screed rail, the railing, the walkway, the
overhang brackets and the finishing machine. Designers should consider talking with local

NSBA Guide to Navigating Routine Steel Bridge Design / 218


contractors to obtain reasonable estimates of the load values. The vertical load on the overhang
can then be resolved into lateral forces, F, acting on the flanges. The lateral forces are dependent
on the assumed angle that the bracket makes with the web; overhang brackets bearing near the
bottom flange is the preferred configuration. The deck overhang brackets may bear on the girder
web only if means such as blocking, bracing, or other stiffening are provided so that the web is not
subject to excessive out-of-plane bending stresses or otherwise damaged, and so that the resulting
deflections of the overhang falsework do not adversely affect proper placement and screeding of
the bridge deck concrete. The lateral forces on the top flange and web increase when the bracket
bears directly on the web.
The approximate equations given in the Commentary for this Article may be used to estimate the
flange lateral bending moments at the cross-frames or diaphragms due to the lateral flange forces.
Eq. C6.10.3.4.1-1 applies if a statically equivalent uniformly distributed bracket force, Fℓ, is
assumed; bracket dead loads are typically assumed applied uniformly. Eq. C6.10.3.4.1-2 may be
used if the finishing machine load is assumed applied as a single concentrated load.
For a discretely braced compression flange, the lateral bending stress due to the overhang bracket
load will often be subject to amplification (see the Discussion of Article 6.10.1.6 in this Guide).
The first-order flange lateral bending stress, fℓ1, is determined by dividing the lateral bending
moment by the lateral section modulus of the flange (i.e., tfbf2/6). The amplified lateral bending
stress is subject to the specified limit on lateral bending stress of 0.6Fyf (Article 6.10.1.6). Major-
axis bending moments due to the overhang construction loads (e.g., formwork, walkways, and
finishing machine loads) are typically not considered because these loads are usually much smaller
in magnitude relative to other design loads on the bridge. Also, overhang construction loads
represent a temporary loading condition that is not present in the finished structure. Lateral bending
moments due to these loads are usually much more critical. The magnitude and application of the
overhang loads assumed in the design should also be shown on the contract documents.
In some cases, the flange lateral bending effects resulting from overhang bracket loads can be
significant, particularly when shallow depth girders are being used. In such cases, reducing the
width of the permanent bridge deck overhang or reducing the width or eccentricity of the
temporary overhang brackets, construction access, and construction appurtenances (such as the
deck screed rail) may be beneficial.
Finally, it should be noted that the overhang brackets often support the rails on which the deck
screed runs when finishing the deck; torsional deformation (twisting) of the girders due to the
eccentric loading applied by the overhang brackets contributes to vertical deflections of the screed
rail. These deflections affect the profile to which the deck is finished, and excessive deflection can
result in improper deck thickness. In such situations, additional temporary bracing between the
exterior and first interior girder is often used to reduce the torsional deformation of the exterior
girder. The evaluation of this effect is typically the responsibility of the Contractor’s specialty
engineer and is usually addressed in the design of the overhang falsework system, but occasionally
Owner-agencies require the designer to evaluate this effect; see local Owner-agency policy.
Software such as the TAEG program is sometimes used to automate these calculations
(https://kart.ksdot.org/).
Analysis of Deck Placement during Phased Construction or Bridge Widening

NSBA Guide to Navigating Routine Steel Bridge Design / 219


The analysis of the effects of deck placement during phased construction or bridge widening
warrants special discussion. In these situations, there will be instances where some or all of the
weight of a wet concrete deck placement will represent loading on a portion of a routine steel I-
girder bridge which already has a composite concrete deck in place. In those cases, the loading on
the composite portion of the bridge should be treated as a composite dead load, applied to the
composite section for calculating girder and deck stresses.
For example, consider a bridge built in two phases. Assume Phase 1 has five girders (numbered
Girder 1 through Girder 5), spaced 9’-0” center to center, and that at Girder 5 the deck overhang
in the temporary Phase 1 condition is 2’-0”. Later Phase 2 is built. Assume Phase 2 also has five
girders (numbered Girder 6 through Girder 10), spaced 9’-0” center to center, and that at Girder 6
the deck overhang is 2’-0”. Assume a 5’-0” closure pour will be placed, connecting Phase 1 to
Phase 2, after the five Phase 2 girders are built and their deck is cast. In this case, the weight of the
wet concrete deck for the Phase 1 deck construction represents a noncomposite dead load on
Girders 1 through 5, and the weight of wet concrete deck for the Phase 2 deck construction
represents a noncomposite dead load on Girders 6 through 10. The weight of the wet concrete
deck of the closure pour including any associated deck forms and any cross-frames added between
the two phases, on the other hand, represent composite dead loads on the two previously completed
portions of the bridge. Since each of those previous phases of construction are fairly wide, and
the closure pour load is concentrated between them, the majority of the associated closure pour
loads would be carried by Girders 5 and 6, with some of the loads carried by Girders 4 and 7 and
possibly some by Girders 3 and 8, etc.
The actual load distribution, and the application of loading to composite or noncomposite sections,
would need to be evaluated on a case by case basis, depending on the geometry and structural
configuration of the bridge during each phase of construction. Consultation with senior bridge
engineers experienced in the design of steel girder bridges built using phased construction methods
is strongly encouraged when determining how to approach the analysis and design of these types
of structures.
Further discussion of analysis and design issues associated with phased construction and widening
of steel girder bridges can be found in Section 3.17 of AASHTO-NSBA Steel Bridge
Collaboration’s Guideline G13.1-2019 Guidelines for Steel Girder Bridge Analysis, and in
Sections 1.6.2 and 2.2.2 of AASHTO-NSBA Steel Bridge Collaboration’s Guideline G12.1-2020
Guidelines to Design for Constructability and Fabrication.
Deck Placement Sequence and Overhang Bracket Design Example References
For design examples illustrating the sequential deck placement and deck overhang calculations,
consult the FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous
Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous Straight
Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span Continuous Straight
Composite Steel Wide-Flange Beam Bridge.

6.10.3.4.2 Global Displacement Amplification in Narrow I-Girder Bridge Units


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 220


Discussion:
This Article provides specific guidelines for checking the global stability of spans of slender
unsupported straight or horizontally curved multiple I-girder bridge units interconnected by cross-
frames or diaphragms when in their noncomposite condition during the deck placement. The
provisions apply only to spans of I-girder bridge units with three or fewer girders (e.g., very narrow
bridges, or partial-width stages of bridges that may occur during a phased construction or bridge
widening scenario). Therefore, these provisions are not applicable to the routine steel I-girder
bridges covered by this Guide since these bridges consist of four or more girders.
As noted in the “definition of a routine steel I-girder bridge” for the purposes of this Guide, all
phases of phased construction projects or bridge widenings should also meet the definition of a
routine steel I-girder bridge (including the requirements related to framing plan geometry and
minimum numbers of girders, etc.); if a widening or a given phase of construction meets the
definition of a routine steel I-girder bridge, then the provisions of this Article should not be
applicable. If, however, the widening or a given phase of construction is narrow and has three or
fewer girders, the provisions of this Article should be considered and global stability should be
investigated.

6.10.3.5 Dead Load Deflections


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article is considered self-explanatory and applicable to the routine steel I-girder bridges
covered by this Guide. The Article simply refers to the provisions of Article 6.7.2 for further
information on establishing the required vertical camber to compensate for the computed dead load
deflections (see the Discussion of Article 6.7.2 in this Guide).

6.10.4 Service Limit State

6.10.4.1 Elastic Deformations


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article points to optional span-to-depth ratios and live-load deflection requirements in Article
2.5.2.6 to control elastic deformations so that the bridge will perform satisfactorily over its regular
service life. This Article is applicable to the routine steel I-girder bridges covered by this Guide.
Most Owners choose to enforce a live-load deflection requirement at the service limit state; consult
the applicable Owner-agency standard. See the Discussion of Article 2.5.2.6.2 in this Guide for a
full explanation of these provisions, their applicability, and how to implement them.

6.10.4.2 Permanent Deformations

6.10.4.2.1 General
Determination of applicability, Simple Span Bridges: Partially applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 221


Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
The provisions of Article 6.10.4.2 are intended to prevent objectionable permanent deformations
of the steel beams or girders, caused by localized yielding and potential web bend-buckling under
expected severe traffic loadings, which might impair rideability. The requirements in this Article
must be satisfied under the Service II load combination (Table 3.4.1-1), which can be characterized
in a simplified manner as being approximately halfway between the load combination used for
Service I and Strength I. The load level for this limit state can reasonably be expected to be
exceeded less than once every six months on average. The live load used for Service II load
combination is HL-93 placed in one or more lanes. Although not covered in this Article, slip in
bolted slip-critical connections is also to be checked at the service limit state under the Service II
load combination. The provisions of Article 6.10.4.2 are applicable to the routine steel multi-span
bridges, and partially applicable to the routine simple-span bridges, covered by this Guide. The
checks in this Article can often control the size of the bottom flange of these bridges in regions of
positive flexure, and the web thickness in regions of negative flexure in multi-span continuous
bridges.
The provisions of Article 6.10.4.2.1 are intended to determine whether the concrete deck may be
considered effective in tension for loads applied to the composite section in regions of negative
flexure for the service limit state checks specified in Article 6.10.4.2.2 in routine multi-span
continuous bridges. It is recommended that the deck be considered effective in tension if the
following three requirements are satisfied:
1. shear connectors are provided along the entire length of the member;
2. the minimum one-percent longitudinal reinforcement in the concrete deck is provided in
accordance with Article 6.10.1.7 (see the Discussion of Article 6.10.1.7 in this Guide); and
3. the maximum longitudinal tensile stress in the concrete deck under the Service II load
combination is less than 2fr, where fr is the modulus rupture of the concrete taken equal to
0.24 f c' for the routine steel I-girder bridges covered by this Guide.

Otherwise, the steel section alone or the steel section plus the longitudinal reinforcement within
the effective width of the concrete deck is to be used (if shear connectors are provided) depending
on the preferences of the Owner-agency.
Note that the check to determine whether the concrete deck may be considered effective in tension
for the service limit state checks is not applicable for a simple span because the concrete deck is
in compression over the entire length of the span.

6.10.4.2.2 Flexure
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 222


Discussion:
The provisions of this Article are partially applicable to the routine steel I-girder bridges covered
by this Guide as described in the following.
This Article provides flange stress limits to indirectly control permanent deformations in steel
flexural members at the service limit state. Only Eqs. 6.10.4.2.2-1 and 6.10.4.2.2-2 are applicable
to the routine steel I-girder bridges covered by this Guide because shear studs are assumed
provided along the full length (or portions) of the beam or girder. Should shear connectors be
discontinued in regions of negative flexure in a multi-span continuous bridge, which is permitted
but not recommended, Eqs. 6.10.4.2.2-1 and 6.10.4.2.2-2 are to be applied, as applicable, and not
Eq. 6.10.4.2.2-3. Flange lateral bending stresses at the service limit state are not a concern for the
routine I-girder bridges covered by this Guide as the effects of torsion due to deck overhang bracket
loads and wind loads are not considered at the service limit state. Eqs. 6.10.4.2.2-1 and 6.10.4.2.2-
2 do not control and need not be checked for sections in negative flexure in multi-span continuous
bridges designed according to the provisions of Article 6.10.8 at the strength limit state (i.e.,
bridges utilizing slender web sections or compact or noncompact web sections treated as slender
web sections in regions of negative flexure), and for sections in positive flexure treated as
noncompact sections at the strength limit state (see the Discussion of Article 6.10.7 in this Guide).
A web bend buckling check is also specified in this Article at the service limit state to control
bending deformations and transverse displacements in the compression zone of the web (Eq.
6.10.2.2.2-4). Regions in negative flexure in multi-span continuous bridges utilizing slender web
sections are particularly susceptible to web bend buckling in composite girders at the service limit
state, especially when the concrete deck is considered to be effective in tension as permitted for
composite sections in Article 6.10.4.2.1 when certain specified conditions are satisfied. When the
concrete deck is considered effective in tension in regions of negative flexure, more than half of
the web is likely to be in compression increasing the susceptibility of the web to bend buckling.
As a result, the check in this case may often end up governing the web thickness of the girder in
these regions when the concrete is assumed effective. Should the necessary requirements be met
such that the concrete may be considered effective in tension as permitted in Article 6.10.4.2.1, it
is required that the elastic depth of the web in compression, Dc, be computed according to the
provisions of Article D6.3.1 for the web bend-buckling check given by Eq. 6.10.4.2.2-4 (see the
Discussion of Article D6.3.1 in this Guide). The provisions of Article D6.3.1 account for the fact
that part of the load is applied to the noncomposite section, and thus provide a more accurate
location of the neutral axis based on the total factored stresses. For composite sections in regions
of positive flexure at the service limit state, web bend-buckling will typically not control because
Dc is small in these regions in the routine steel I-girder bridges covered by this Guide, and so the
check is waived. Options to consider should Eq. 6.10.4.2.2-4 be violated include providing a larger
compression flange or a smaller tension flange to reduce Dc or providing a thicker web.Because
an explicit web bend-buckling check is specified at the service limit state, the web load-shedding
factor, Rb, is not included in the equations of this Article.
The web bend-buckling check given by Eq. 6.10.4.2.2-4 need not be checked for simple-span
routine steel I-girder bridges covered by this Guide since the beam or girder is subject to positive

NSBA Guide to Navigating Routine Steel Bridge Design / 223


flexure only and in those cases Dc is small and thus the potential for web bend-buckling is
negligible.
The web bend-buckling check given by Eq. 6.10.4.2.2-4 also need not be checked for routine multi-
span continuous rolled beam bridges covered by this Guide. In regions of negative flexure,
theoretical web bend-buckling will not occur for elastic stress levels, computed according to beam
theory, at or below Fyc for a compact or noncompact web section. In regions of positive flexure,
Dc is small and thus the potential for web bend-buckling is negligible.
The definition of a routine steel I-girder bridge specifically excludes the use of moment
redistribution methods and so the optional provisions of Appendix B6 mentioned in the
Commentary for this Article are considered not applicable to the routine steel I-girder bridges
covered by this Guide.
The provision in this Article related to compact composite sections in positive flexure utilized in
shored construction is not applicable; shored construction is not recommended for routine steel I-
girder bridges.
For further information on service limit state design, consult Section 6.5.4 of the Reference Manual
for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples illustrating service limit state design checks to limit
permanent deformations, consult the FHWA’s Steel Bridge Design Handbook, Design Example
1, Three-Span Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-
Span Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span
Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.5 Fatigue and Fracture Limit State

6.10.5.1 Fatigue
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article refers to the provisions of Article 6.6 for checking fatigue of details using the fatigue
live load (see the Discussion of Articles 6.6 and 3.6.1.4 in this Guide) and the appropriate Fatigue
load combination (Table 3.4.1-1), and to the provisions of Article 6.10.10.2 or 6.10.10.3 (as
applicable) for determining the nominal fatigue resistance of shear connectors (see the Discussion
of Articles 6.10.10.2 and 6.10.10.3 in this Guide). These provisions are applicable to the routine
steel I-girder bridges covered by this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 224


The provision for checking the fatigue stress range due to major-axis bending plus flange lateral
bending in horizontally curved I-girder bridges is not applicable to the routine steel I-girder bridges
covered by this Guide.
For further information on fatigue limit state design of shear connectors, consult the Discussion of
Article 6.10.10.1.2 in this Guide and Section 6.6.2 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures. For
design examples illustrating fatigue limit state design of shear connectors, consult the FHWA’s
Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous Straight Composite
Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous Straight Composite Steel I-
Girder Bridge, and Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-
Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.5.2 Fracture
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article refers to Article 6.6.2.1 for provisions related to the fracture toughness requirements
(i.e., Charpy V-Notch toughness requirements) specified in the contract documents. This Article
is applicable to the routine steel I-girder bridges covered by this Guide. For further explanation,
see the Discussion of Article 6.6.2.1 in this Guide.

6.10.5.3 Special Fatigue Requirement for Webs


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provision in this Article is intended to control significant elastic flexing of the web under
repeated live loads by limiting the factored shear for this check, Vu, to the shear buckling resistance
of the web, Vcr (see the Discussion of Article 6.10.9.1 in this Guide). Vu for this check is taken
equal to the unfactored permanent load shear plus the factored Fatigue I shear due to the fatigue
live load plus the applicable dynamic load allowance (see the Discussion of Articles 3.6.1.4 and
3.6.2.1 in this Guide); that is, this check is to be done for the heaviest truck expected to cross the
bridge in 75 years. This provision is applicable to the routine steel I-girder bridges covered by this
Guide.
If post-buckling tension-field action were permitted under this load condition, the principal tensile
stress range acting along the buckle would result in significant out-of-plane flexing of the web

NSBA Guide to Navigating Routine Steel Bridge Design / 225


under repeated live loads, which would be undesirable. By limiting the factored shear under this
load condition to the shear buckling resistance, Vcr, the member is assumed to be able to sustain
an infinite number of smaller loadings without fatigue cracking due to this effect.
Eq. 6.10.5.3-1 need not be checked for unstiffened webs or end panels of stiffened webs because
the shear in these cases is already limited to Vcr at the strength limit state (see the Discussion of
Article 6.10.9 in this Guide).
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.6 Strength Limit State

6.10.6.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article refers to the applicable Strength load combinations given in Table 3.4.1-1, which are
utilized in the design checks at the strength limit state for the routine steel I-girder bridges covered
by this Guide.
Note that the Commentary for this Article contains the following useful information:
• Explanation of why flexural resistances at the strength limit state are expressed in terms of
stress or moment in different parts of the specification;
• Guidance on correctly interpreting and applying the results from refined analyses at the
strength limit state (although refined methods of analysis are not necessary or
recommended for the design of routine steel I-girder bridges);
• Discussion about continuously braced flanges; and
• Discussion about the level of axial force at which a member can be solely designed as a
flexural member.

6.10.6.2 Flexure

6.10.6.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 226


Discussion:
This Article refers to the provisions Article 6.10.1.8 if there are holes in the tension flange of an I-
section flexural member; e.g., at a bolted splice (see the Discussion of Article 6.10.1.8 in this
Guide). These provisions are applicable to the routine steel I-girder bridges covered by this Guide.
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples illustrating strength limit state design
flexure checks, consult the as well as FHWA’s Steel Bridge Design Handbook, Design Example
1, Three-Span Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-
Span Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span
Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.6.2.2 Composite Sections in Positive Flexure


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article defines the requirements for a composite I-section in regions of positive flexure to
qualify as a compact or a noncompact section at the strength limit state, and where the provisions
to design each type of section are located in Article 6.10.7 (see the Discussion of Article 6.10.7 in
this Guide). The provisions of this Article are partially applicable to the routine steel I-girder
bridges covered by this Guide as described further below.
Most composite sections in regions of positive flexure in routine steel I-girder bridges will qualify
as compact sections; that is, these bridges are straight, the specified minimum yield strength of the
flanges does not exceed 70 ksi, web longitudinal stiffeners are not required, and sections will
satisfy the web-slenderness limit given by Eq. 6.10.2.2.2-1, where Dcp is the depth of the web in
compression at the plastic moment, Mp (see the Discussion of Articles D6.1 and D6.3.2 in this
Guide). For composite sections in positive flexure in routine steel I-girder bridges, the plastic
neutral axis (PNA) will almost always be located either in the top flange or in the concrete deck.
In either case, Dcp will equal zero and Eq. 6.10.2.2.2-1 may be considered automatically satisfied.
For compact sections in regions of positive flexure, the nominal flexural resistance is permitted to
exceed the moment at first yield assuming there are no holes in the tension flange at the section
under consideration, but for noncompact sections and for compact sections with holes in the
tension flange it is not. The moment at first yield, My, is defined as the moment at which an outer
fiber first attains the yield stress (see the Discussion of Article D6.2.2 in this Guide). The nominal

NSBA Guide to Navigating Routine Steel Bridge Design / 227


flexural resistance is not permitted to exceed the plastic moment, Mp, for both noncompact and
compact sections. Mp is defined as the resisting moment of a fully yielded cross-section (see the
Discussion of Article D6.1 in this Guide). Another primary difference is that for noncompact
sections, the nominal flexural resistance is expressed in terms of the elastically computed flange
stress. However, for compact sections, the nominal flexural resistance is expressed in terms of
moment. The reasons for this are discussed further in the Commentary for Article 6.10.6.1.
Sections that qualify as compact sections without holes in the tension flange may conservatively
be treated as noncompact sections, if desired. For compact sections, the strength limit state criteria
are unlikely to control the size of the bottom flange (or the size of the rolled-beam section) in
regions of positive flexure; service or fatigue limit state criteria are more likely to govern. For a
noncompact section, strength limit state criteria are more likely to control the size of the bottom
flange (or the size of the rolled-beam section).
This Article also refers to the ductility requirement given in Article 6.10.7.3 to provide a ductile
mode of failure, which must be checked for both compact and noncompact sections (see the
Discussion of Article 6.10.7.3 in this Guide).
The language at the beginning of this Article relative to a horizontally curved bridge or a kinked
(chorded) continuous bridge is not applicable to the routine steel I-girder bridges covered by this
Guide.
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples illustrating strength limit state design
flexure checks, consult the FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-
Span Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span
Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span
Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.6.2.3 Composite Sections in Negative Flexure and Noncomposite Sections


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article defines the requirements for a composite I-section in regions of negative flexure and
a noncomposite I-section in regions of positive or negative flexure to qualify as a slender web
section or a nonslender web section, and the provisions that may be used to define each type of
section at the strength limit state. These provisions are partially applicable to the routine steel I-
girder bridges covered by this Guide as described further below. These criteria, when applied at

NSBA Guide to Navigating Routine Steel Bridge Design / 228


the strength limit state, typically control the size of the top and bottom flange (or the size of the
rolled-beam section) and the cross-frame or diaphragm spacing in regions of negative flexure in
the routine steel multi-span continuous I-girder bridges covered by this Guide. These criteria can
also control the size of the noncomposite section and the cross-frame or diaphragm spacing in
regions of positive flexure during construction.
A slender web section contains a web with a slenderness at or above the limit at which the
theoretical elastic bend-buckling stress of the web in flexure (see the Discussion of Article
6.10.1.9.1 in this Guide) is reached prior to reaching the yield strength of the compression flange,
Fyc. Otherwise, the section is a nonslender web section and is not subject to web bend-bucking
prior to reaching its limit of flexural resistance. Sections with slender webs rely on significant web
post bend-buckling resistance at the strength limit state. The limiting web slenderness ratio
delineating a slender web section from a nonslender web section is called λrw and is given by Eq.
6.10.6.2.3-3 (see also Table 6.10.1.10.2-2). λrw is also referred to as the noncompact web
slenderness limit. The web slenderness based on the depth of the web in compression in the elastic
range, Dc (see the Discussion of Article D6.3.1 in this Guide), is checked against this limiting ratio
(Eq. 6.10.6.2.3-1). For sections with a specified minimum yield strength of the compression
flange, Fyc, equal to 50 ksi, the upper and lower limits of λrw are 111 and 137. Sections with a
web slenderness exceeding λrw are termed slender web sections.
In the routine steel multi-span continuous I-girder bridges covered by this Guide, nonslender web
sections in regions of negative flexure that also satisfy the provisions of Eq. 6.10.6.2.3-2, and do
not have holes in the tension flange at the section under consideration, may optionally be designed
at the strength limit state according to the provisions of Appendix A6 (see the Discussion of
Appendix A6 in this Guide); otherwise, the section is treated as a slender web section and the
provisions of Article 6.10.8 must be used to design the section (see the Discussion of Article 6.10.8
in this Guide). The provisions of Article 6.10.8 assume that the section is a slender web section
regardless of whether it is or not. As such, the nominal flexural resistance of the section computed
according to these provisions is not permitted to exceed the moment at first yield, and as a result,
the resistance equations in Article 6.10.8 are expressed in terms of stress for the reasons discussed
in the Commentary for Article 6.10.6.1. The nominal flexural resistance of nonslender web
sections designed according to the provisions of Appendix A6 is permitted to exceed the moment
at first yield. Hence, the resistance equations in Appendix A6 are expressed in terms of bending
moment, again for the reasons discussed in the Commentary for Article 6.10.6.1. For
noncomposite nonslender web sections during construction, the lateral-torsional buckling
resistance for use in Eq. 6.10.3.2.1-2 may optionally be computed according to the provisions of
Article A6.3.3 (to include the beneficial contribution of the St. Venant torsional stiffness, J).
Nonslender web sections are further categorized as either compact web sections or noncompact
web sections in Appendix A6. A compact web section is one that satisfies the web slenderness limit
based on the depth of the web in compression at the plastic moment, Dcp (see the Discussion of Article
D6.3.2 in this Guide) given by Eq. A6.2.1-1. For sections with Fyc equal to 50 ksi, the limiting
web slenderness from Eq. A6.2.1-1 based on Dcp is 91 for a shape factor, Mp/My, of 1.12 and
64 for a shape factor of 1.30. These limits would generally be satisfied by rolled shapes or
plate girders with proportions similar to those of a rolled shape, which would typically be used
in a shorter-span bridge (i.e., spans of about 120 ft or less). Sections with compact webs and

NSBA Guide to Navigating Routine Steel Bridge Design / 229


Iyc/Iyt greater than or equal to 0.3 (Eq. 6.10.6.2.3-2) are able to develop their full plastic moment
capacity, Mp, at the strength limit state provided that other steel grade, ductility, flange slenderness
and/or lateral bracing requirements are satisfied. Note that Iyc and Iyt in Eq. 6.10.6.2.3-2 are to be
computed about the vertical axis in the plane of the web (or about the strong axis of each flange).
Sections exceeding the web slenderness limit given by Eq. A6.2.1-1, but with a web slenderness based
on the elastic depth of the web in compression, Dc (see the Discussion of Article D6.3.1 in this Guide),
less than or equal to λrw, are termed noncompact web sections, which have a nominal flexural resistance
at the strength limit state that linearly transitions from Mp to the moment at first yield, My (see the
Discussions of Articles D6.1 and D6.2 in this Guide) as a function of the web slenderness. The majority
of routine steel-bridge plate-girder sections utilize either slender web sections or noncompact web
sections that approach λrw.
The definition of a routine steel I-girder bridge specifically excludes the use of moment
redistribution methods and so the optional provisions of Appendix B6 mentioned in this Article
are considered not applicable to the routine steel I-girder bridges covered by this Guide.
The language at the beginning of this Article relative to a horizontally curved bridge or a kinked
(chorded) continuous bridge is not applicable to the routine steel I-girder bridges covered by this
Guide.
Some specific guidance on the applicability of the provisions in this Article to various types of
routine steel I-girder bridges is provided below.
The provisions of this Article are not applicable at the strength limit state for the routine simple-
span bridges covered by this Guide, including the provisions of the optional Appendix A6, because
simple-span bridges are subject to positive flexure only. However, the provisions of Article A6.3.3
may optionally be applicable for the noncomposite section in a simple span bridge during
construction if the noncomposite section qualifies as a compact web or noncompact web section.
Sections in regions of negative flexure in the routine multi-span continuous rolled beam bridges
covered by this Guide typically qualify as compact web sections and should be designed at the
strength limit state using the applicable provisions of Appendix A6 instead of the more
conservative provisions of Article 6.10.8. The provisions of Article A6.3.3 are also optionally
applicable for the noncomposite section during construction in all regions of the beam.
For sections in regions of negative flexure in the routine multi-span continuous plate girder bridges
covered by this Guide that satisfy the restrictions specified in this Article, and do not have holes
in the tension flange, the optional provisions of Appendix A6 may be employed to design the
section at the strength limit state; otherwise, the section is treated as a slender web section and the
provisions of Article 6.10.8 must be used to design the section. Compact web sections should be
designed using the provisions of Appendix A6 at the strength limit state instead of the more
conservative provisions of Article 6.10.8. The provisions to use for the design of a noncompact
web section in regions of negative flexure at the strength limit state are up to the judgment of the
Engineer. The provisions of Article A6.3.3 may optionally be applicable for the noncomposite
section during construction if the noncomposite section qualifies as a compact web or noncompact
web section.

NSBA Guide to Navigating Routine Steel Bridge Design / 230


For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples illustrating strength limit state design
flexure checks, consult the FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-
Span Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span
Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span
Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.6.3 Shear
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article simply points to the provisions of Article 6.10.9 for determining the factored shear
resistance of the beam or girder at the strength limit state (see the Discussion of Article 6.10.9 in
this Guide), which are applicable to the routine steel I-girder bridges covered by this Guide.
For further information on strength limit state design for shear, consult Section 6.5.7 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples illustrating strength limit state design shear
checks, consult the FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-Span
Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous
Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span Continuous
Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.6.4 Shear Connectors


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 231


Discussion:
This Article simply points to the provisions of Article 6.10.10.4 for determining the factored shear
resistance of shear connectors for I-section flexural members at the strength limit state (see the
Discussion of Article 6.10.10.4 in this Guide), which are applicable to the routine steel I-girder
bridges covered by this Guide.
For further information on strength limit state design of shear connectors, consult the Discussion
of Article 6.10.10.4 in this Guide and Section 6.6.2 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures. For
design examples illustrating strength limit state design of shear connectors, consult the FHWA’s
Steel Bridge Design Handbook, Design Example 1, Three-Span Continuous Straight Composite
Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous Straight Composite Steel I-
Girder Bridge, and Design Example 2B, Two-Span Continuous Straight Composite Steel Wide-
Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.7 Flexural Resistance—Composite Sections in Positive Flexure

6.10.7.1 Compact Sections

6.10.7.1.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article prescribes the relationship that must be satisfied at the strength limit state for compact
composite I-sections in regions of positive flexure (i.e., the moment form of the one-third rule
flexural resistance equation). Most composite sections in regions of positive flexure in routine steel
I-girder bridges without holes in the tension flange will qualify as compact sections (see the
Discussion of Article 6.10.6.2.2 in this Guide); therefore, this Article is applicable to the routine
steel I-girder bridges covered by this Guide. As mentioned in the Discussion of Article 6.10.6.2.2
in this Guide, sections that qualify as compact sections may conservatively be treated as
noncompact sections (see the Discussion of Article 6.10.7.2 in this Guide), if desired.
The relationship given by Eq. 6.10.7.1.1-1 includes the bottom (tension) flange lateral bending
stress, fℓ. For the routine steel I-girder bridges covered by this Guide, the only source of flange
lateral bending stress to be considered at the strength limit state is wind loading occurring under
Strength load combinations that include wind load effects. fℓ cannot exceed 0.6Fyf. Amplification
of fℓ in the tension flange is not required. Lateral bending does not need to be considered in the top

NSBA Guide to Navigating Routine Steel Bridge Design / 232


(compression) flange in regions of positive flexure because the flange is continuously supported
by the concrete deck. Mu and fℓ are always taken as positive in sign in Eq. 6.10.7.1.1-1. However,
when summing dead and live load moments to obtain the total factored major-axis moment, Mu,
and total factored lateral bending stresses, f, to apply in the equations, the signs of the individual
stresses or moments must be considered. If there is no flange lateral bending considered, the fℓ
term drops out of the equation.
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.7.1.2 Nominal Flexural Resistance


Determination of applicability, Simple Span Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
This Article provides the equations for computing the nominal flexural resistance, Mn, of compact
composite I-sections in regions of positive flexure at the strength limit state for use in checking
Eq. 6.10.7.1.1-1. These equations are applicable to the multi-span continuous routine steel I-girder
bridges, and partially applicable to the simple-span routine I-girder bridges, covered by this Guide
for which the sections in regions of positive flexure qualify as compact (see the Discussion of
Article 6.10.6.2.2 in this Guide). The nominal flexural resistance of compact sections without holes
in the tension flange is permitted to exceed the moment at first yield.
There are two different equations for the nominal flexural resistance of compact composite
sections in regions of positive flexure (Eq. 6.10.7.1.2-1 or Eq. 6.10.7.1.2-2), depending on the
value of Dp compared with Dt. Dp is the distance from the top of the concrete deck to the neutral
axis of the composite section at the plastic moment (PNA), and Dt is the total depth of the
composite section, including the structural concrete deck. The location of the PNA can be
determined using the provisions of Article D6.1 (see the Discussion of Article D6.1 in this Guide);
for the routine steel I-girder bridges covered by this Guide, the PNA will most always be located

NSBA Guide to Navigating Routine Steel Bridge Design / 233


either in the top flange or in the concrete deck. The haunch depth may be considered in the
computation of Dt if permitted by the Owner-agency.
For simple span routine steel I-girder bridges, Mn is calculated from either Eq. 6.10.7.1.2-1 or
6.10.7.1.2-2, as applicable. The limitation given by Eq. 6.10.7.1.2-3 does not apply.
For multi-span continuous bridges, Mn computed from either Eq. 6.10.7.1.2-1 or 6.10.7.1.2-2, as
applicable, cannot exceed 1.3RhMy (Eq. 6.10.7.1.2-3), where My is the moment at first yield
determined from the provisions of Article D6.2.2 (see the Discussion of Article D6.2.2 in this
Guide). For typical composite sections in positive flexure, a considerable amount of yielding and
inelastic curvature is required to reach Mp. The resulting shedding of moment to adjacent interior-
pier sections that do not have additional capacity to sustain these larger moments as the positive-
moment section yields and loses its effective stiffness could potentially result in incremental
collapse under repeated live load applications. Thus, unless the specific steps below are taken, the
resistance of the section in positive flexure is conservatively limited.
The limitation given by Eq. 6.10.7.1.2-3 may be waived if the adjacent interior-pier sections satisfy
the requirements given in the two bulleted items in this Article that refer to provisions given in
Appendix B6 (see the Discussion of Appendix B6 in this Guide), which provide the necessary
ductile moment-rotation characteristics at the interior-pier sections to sustain the larger moments.
Interior-pier sections in multi-span continuous rolled beam bridges are more likely to satisfy these
requirements. However, in most cases for the routine steel multi-span continuous bridges covered
by this Guide, the excess flexural resistance above 1.3RhMy should not be necessary for compact
composite sections in regions of positive flexure at the strength limit state, as other limit state
criteria will typically control the design of the section.
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.7.2 Noncompact Sections

6.10.7.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 234


Discussion:
This Article prescribes the relationships that must be satisfied at the strength limit state for the
compression and tension flanges of noncompact composite I-sections in regions of positive
flexure. These provisions may conditionally apply for the routine steel I-girder bridges covered by
this Guide should the section not qualify as a compact section at the strength limit state, which is
atypical for a routine steel I-girder bridge, or should the section qualify as a compact section and
instead be conservatively designed as a noncompact section as mentioned in the Discussion of
Article 6.10.6.2.2 in this Guide.
For composite I-sections in regions of positive flexure, lateral bending does not need to be
considered in the top (compression) flange at the strength limit state because the flange is
continuously supported by the concrete deck. However, since the bottom (tension) flange is not
continuously supported, lateral bending must be considered in flexural resistance computations for
the tension flange (using the stress form of the one-third rule flexural resistance equation). For the
routine steel I-girder bridges covered by this Guide, the only source of flange lateral bending stress
to be considered at the strength limit state is wind loading occurring under the Strength load
combinations that include wind load effects. fℓ cannot exceed 0.6Fyf. Amplification of fℓ in the
tension flange is not required. fbu and fℓ are always taken as positive in sign in Eq. 6.10.7.2.1-2.
However, when summing dead and live load stresses to obtain the total factored major-axis stress,
fbu, and total factored lateral bending stresses, f, to apply in the equations, the signs of the
individual stresses must be considered. If there is no flange lateral bending considered, the fℓ term
drops out of the equation.
This Article further limits the maximum longitudinal compressive stress in the concrete deck at
the strength limit state for a noncompact I-section to 0.6f'c to provide for linear behavior of the
concrete, which is assumed in the calculation of the steel flange stresses. This requirement is
unlikely to control for the routine steel I-girder bridges covered by this Guide.
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.7.2.2 Nominal Flexural Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 235


Discussion:
This Article provides the equations for computing the nominal flexural resistance of the
compression flange, Fnc, and the nominal flexural resistance of the tension flange, Fnt, for
noncompact composite I-sections in regions of positive flexure at the strength limit state for use in
checking Eqs. 6.10.7.2.1-1 and 6.10.7.2.1-2, respectively. These provisions may conditionally
apply for the routine steel I-girder bridges covered by this Guide should the section not qualify as
a compact section at the strength limit state, which is atypical for a routine steel I-girder bridge, or
should the section qualify as a compact section and instead be conservatively designed as a
noncompact section as mentioned in the Discussion of Article 6.10.6.2.2 in this Guide. The
nominal flexural resistance of noncompact composite I-sections in positive flexure is limited to
the moment at first yield. Therefore, the nominal flexural resistance for each flange is best
expressed in terms of the flange stress for reasons discussed in the Commentary for Article 6.10.6.1
and is taken equal to the yield stress of the respective flange times the appropriate flange-strength
reduction factors.
For both the compression and tension flanges, the hybrid factor, Rh (see the Discussion of Article
6.10.1.10.1 in this Guide), is applied to the flange yield stress. However, routine steel I-girder
bridges as defined for the purposes of this Guide do not utilize hybrid girder designs; as a result,
the hybrid factor, Rh, is to be taken equal to 1.0 and does not affect the calculation of the nominal
flexural resistance.
In addition, for the compression flange, the web load-shedding factor, Rb (Article 6.10.1.10.2), is
also applied to the flange yield stress. However, for the routine steel I-girder bridges covered by
this Guide, Rb is to be taken equal to 1.0 and does not affect the calculation of the nominal flexural
resistance because web bend buckling and subsequent load shedding are not a consideration for
composite sections in regions of positive flexure in these bridges at the strength limit state (see the
Discussion of Article 6.10.1.10.2 in this Guide).
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.7.3 Ductility Requirement


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 236


Discussion:
The provision in this Article provides a ductility requirement that is intended to protect the concrete
deck in regions of positive flexure from premature crushing. The Dp/Dt ratio is limited to 0.42 so
that significant yielding of the bottom flange can be expected prior to the top of the concrete deck
reaching its the crushing strain (see the Discussion of Article 6.10.7.1.2 in this Guide for the
definitions of Dp and Dt). The requirement is applicable to the routine steel I-girder bridges covered
by this Guide and must be checked for both compact and noncompact sections in regions of
positive flexure. This requirement is unlikely to control for the routine steel I-girder bridges
covered by this Guide. If it is found that this requirement is affecting the design that probably
indicates a fundamental flaw in the layout of the superstructure cross-section and/or the
proportioning of the girders.

6.10.8 Flexural Resistance—Composite Sections in Negative Flexure and


Noncomposite Sections

6.10.8.1 General

6.10.8.1.1 Discretely Braced Flanges in Compression


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
Simple Span Bridges:
This Article is not applicable to the simple span routine steel I-girder bridges covered by this Guide
at the strength limit state because simple span bridges are subject to positive flexure only. The
design of the discretely braced top (compression) flange of the noncomposite section during
construction in these bridges is separately covered in Article 6.10.3.2.1 (see the Discussion of
Article 6.10.3.2.1 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This Article provides the relationship that must be satisfied at the strength limit state for discretely
braced bottom (compression) flanges of slender web sections, or compact or noncompact web
sections treated as slender web sections and not designed according to the optional provisions of
Appendix A6 (see the Discussion of Appendix A6 in this Guide), in regions of negative flexure in
multi-span continuous steel I-girder bridges. See the Commentary for Article 6.10.6.2.3 and the
Discussion of Article 6.10.6.2.3 in this Guide for further discussion on the definition and
categorization of compact web, noncompact web, and slender web sections. Therefore, this Article
is conditionally applicable at the strength limit state to the design of the bottom flange in regions
of negative flexure in routine steel multi-span continuous I-girder bridges covered by this Guide.
For composite I-sections in regions of negative flexure, lateral bending does not need to be
considered in the top (tension) flange at the strength limit state because the flange is continuously
supported by the concrete deck. However, since the bottom (compression) flange is discretely

NSBA Guide to Navigating Routine Steel Bridge Design / 237


braced, lateral bending must be considered in flexural resistance computations for the compression
flange (using the stress form of the one-third rule flexural resistance equation). The one-third rule
equation when applied to compression flanges is effectively a beam-column interaction equation,
expressed in terms of the flange stresses computed from elastic analysis. The terms fbu and fℓ are
analogous to the axial and bending moment terms of the beam-column interaction equation. For
the routine steel I-girder bridges covered by this Guide, the only source of flange lateral bending
stress to be considered at the strength limit state is wind loading occurring under the Strength load
combinations that include wind load effects. Amplification of fℓ in the discretely braced
compression flange will likely be required (see the Discussion of Article 6.10.1.6 in this Guide).
fℓ cannot exceed 0.6Fyf after amplification. fbu and fℓ are always taken as positive in sign in Eq.
6.10.8.1.1-1. However, when summing dead and live load stresses to obtain the total factored
major-axis stress, fbu, and total factored lateral bending stresses, f, to apply in the equations, the
signs of the individual stresses must be considered. If there is no flange lateral bending considered,
the fℓ term drops out of the equation. For a discretely braced compression flange, the one-third rule
equation must be checked separately for both flange local buckling and lateral-torsional buckling.
When lateral-torsional buckling controls, the largest values of fbu and fℓ, as applicable, within the
unbraced length must be used to check Eq. 6.10.8.1.1-1 (see the Discussion of Article 6.10.1.6 in
this Guide).
For the routine multi-span continuous rolled beam bridges covered by this Guide, this Article is
only applicable to the discretely braced bottom flanges in regions of negative flexure at the strength
limit state if the provisions of the optional Appendix A6 are not used, which is not recommended
for rolled beam bridges; otherwise, this Article is not applicable.
For the routine multi-span continuous plate girder bridges covered by this Guide, this Article is
applicable to the discretely braced bottom flanges in regions of negative flexure at the strength
limit state if the section is a slender web section, or if the section satisfies the restrictions specified
in Article 6.10.6.2.3 (see the Discussion of Article 6.10.6.2.3 in this Guide) and the provisions of
the optional Appendix A6 are not used to design the section, which is not recommended for
compact web sections in particular; if the provisions of Appendix A6 are used in the latter case,
this Article is not applicable.
The design of the discretely braced top (compression) flange of the noncomposite section in
regions of positive flexure and the discretely braced bottom (compression) flange of the
noncomposite section in regions of negative flexure during construction in these bridges is covered
in Article 6.10.3.2.1 (see the Discussion of Article 6.10.3.2.1 in this Guide).
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the

NSBA Guide to Navigating Routine Steel Bridge Design / 238


provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.8.1.2 Discretely Braced Flanges in Tension


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
Simple Span Bridges:
This Article is not applicable to the simple span routine steel I-girder bridges covered by this Guide
at the strength limit state as simple span bridges are subject to positive flexure only. The design of
the discretely braced bottom (tension) flange of the noncomposite section during construction in
these bridges is covered in Article 6.10.3.2.2.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This Article provides the relationship that must be satisfied at the strength limit state for discretely
braced top (tension) flanges of slender web sections, or compact or noncompact web sections
treated as slender web sections and not designed according to the optional provisions of Appendix
A6 (see the Discussion of Appendix A6 in this Guide), in regions of negative flexure in multi-span
continuous steel I-girder bridges. See the Commentary for Article 6.10.6.2.3 and the Discussion
of Article 6.10.6.2.3 in this Guide for further discussion on the definition and categorization of
compact web, noncompact web, and slender web sections. The top flange in regions of negative
flexure would only be discretely braced if the shear connectors are intentionally omitted in regions
of negative flexure, which is dependent on the preferences of the Owner-agency but not
recommended, and if the Engineer deems that the top flange is not sufficiently encased by the
concrete deck. Therefore, this Article is conditionally applicable at the strength limit state to the
design of the top flange in regions of negative flexure in routine steel multi-span continuous I-
girder bridges covered by this Guide.
The nominal flexural resistance, Fnt, in Eq. 6.10.8.1.2-1 is determined as specified in Article
6.10.8.3 (see the Discussion of Article 6.10.8.3 in this Guide) and is based on tension flange
yielding. Since the top (tension) flange is discretely braced if the conditions stated above are met,
lateral bending must be considered in flexural resistance computations for the tension flange (using
the stress form of the one-third rule flexural resistance equation). For the routine steel I-girder
bridges covered by this Guide, the only source of flange lateral bending stress to be considered at
the strength limit state is wind loading occurring under the Strength load combinations that include
wind load effects. fℓ cannot exceed 0.6Fyf. Amplification of fℓ in the tension flange is not required.
fbu and fℓ are always taken as positive in sign in Eq. 6.10.8.1.2-1. However, when summing dead
and live load stresses to obtain the total factored major-axis stress, fbu, and total factored lateral

NSBA Guide to Navigating Routine Steel Bridge Design / 239


bending stresses, f, to apply in the equations, the signs of the individual stresses must be
considered. If there is no flange lateral bending considered, the fℓ term drops out of the equation.
For the routine multi-span continuous rolled beam bridges covered by this Guide, this Article is
only applicable to the discretely braced top flanges in regions of negative flexure at the strength
limit state if the conditions stated above are met and the provisions of the optional Appendix A6
are not used, which is not recommended for rolled beam bridges; otherwise, this Article is not
applicable.
For the routine multi-span continuous plate girder bridges covered by this Guide, this Article is
applicable to the discretely braced top flanges in regions of negative flexure at the strength limit
state if the conditions stated above are met and the section is a slender web section, or if the section
satisfies the restrictions specified in Article 6.10.6.2.3 (see the Discussion of Article 6.10.6.2.3 in
this Guide) and the provisions of the optional Appendix A6 are not used to design the section,
which is not recommended for compact web sections in particular; if the provisions of Appendix
A6 are used in the latter case, this Article is not applicable.
The design of the discretely braced bottom (tension) flange of the noncomposite section in regions
of positive flexure and the discretely braced top (tension) flange of the noncomposite section in
regions of negative flexure during construction in these bridges is covered in Article 6.10.3.2.2
(see the Discussion of Article 6.10.3.2.2 in this Guide).
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.8.1.3 Continuously Braced Flanges in Tension or Compression


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Partially applicable.
Discussion:
This Article provides the relationship that must be satisfied at the strength limit state for
continuously braced top (tension) flanges of slender web sections, or compact or noncompact web
sections treated as slender web sections and not designed according to the optional provisions of

NSBA Guide to Navigating Routine Steel Bridge Design / 240


Appendix A6 (see the Discussion of Appendix A6 in this Guide), in regions of negative flexure in
multi-span continuous steel I-girder bridges. See the Commentary for Article 6.10.6.2.3 and the
Discussion of Article 6.10.6.2.3 in this Guide for further discussion on the definition and
categorization of compact web, noncompact web, and slender web sections. A continuously braced
flange is anchored to the concrete deck by shear connectors or encased in concrete.
The provision for a continuously braced flange in compression applies only to the design of the
top flange of a noncomposite section at the strength limit state in regions of positive flexure (i.e.,
with no shear connectors) in which the Engineer deems that the flange is sufficiently encased by
the concrete deck, which is not applicable to the routine steel I-girder bridges covered by this
Guide.
Simple Span Bridges:
This Article is not applicable to the routine simple span I-girder bridges covered by this Guide at
the strength limit state as simple span bridges are subject to positive flexure only, and such
structures are defined as having shear connectors so as to achieve a composite design in their final
fully-constructed condition.
The design of the continuously braced top flanges in these bridges at the strength limit state is
covered in Article 6.10.7 (see the Discussion of Article 6.10.7 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
For the routine multi-span continuous rolled beam bridges covered by this Guide, this Article is
only applicable to a continuously braced top flange in regions of negative flexure at the strength
limit state if the provisions of the optional Appendix A6 are not used, which is not recommended
for rolled beam bridges; otherwise, this Article is not applicable.
For the routine multi-span continuous plate girder bridges covered by this Guide, this Article is
only applicable to a continuously braced top flange in regions of negative flexure at the strength
limit state if the section is a slender web section, or if the section satisfies the restrictions specified
in Article 6.10.6.2.3 (see the Discussion of Article 6.10.6.2.3 in this Guide) and the provisions of
the optional Appendix A6 are not used to design the section, which is not recommended for
compact web sections in particular; if the provisions of Appendix A6 are used in the latter case,
this Article is not applicable.
Since the flange is continuously braced, only yielding of the flange is a concern and any flange
lateral bending stresses need not be considered.
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates

NSBA Guide to Navigating Routine Steel Bridge Design / 241


the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.8.2 Compression-Flange Flexural Resistance

6.10.8.2.1 General
Determination of applicability, Simple Span Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
This Article directs the Engineer to the Articles containing the equations necessary to compute the
nominal flexural resistance, Fnc, of a discretely braced compression flange based on flange local
buckling (see the Discussion of Article 6.10.8.2.2 in this Guide) or lateral-torsional buckling (see
the Discussion of Article 6.10.8.2.3 in this Guide) for use in Eq. 6.10.8.1.1-1 at the strength limit
state or in Eq. 6.10.3.2.1-2 for the noncomposite section during construction. The equations in
these Articles assume the section is a slender web section whether it is or not, and the equations
must be satisfied for both flange local buckling and lateral-torsional buckling. See the Commentary
for Article 6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in this Guide for further discussion
on the definition and categorization of compact web, noncompact web, and slender web sections.
The Commentary for this Article includes presentation of the “basic form of all I-section
compression-flange flexural resistance equations.” Designers are strongly encouraged to
familiarize themselves with the concepts presented in this Commentary and the associated graph
in Figure C6.10.8.2.1-1. Possessing a clear understanding of these fundamental concepts is
invaluable for understanding the associated provisions of the AASHTO LRFD BDS.
Simple Span Bridges:
This Article is applicable for the simple span steel I-girder bridges covered by this Guide for
determining Fnc for the discretely braced top (compression) flange of the noncomposite section
during construction for use in Eq. 6.10.3.2.1-2; the one exception being if the section qualifies as
a compact web or noncompact web section and the provisions of Article A6.3.3 are used to
compute Mnc for lateral-torsional buckling to account for the beneficial effect of the St. Venant
torsional constant, J. This Article is not applicable to these bridges at the strength limit state as
simple spans are subject to positive flexure only and the top (compression) flange is continuously
braced by the concrete deck.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
For the routine multi-span continuous rolled beam bridges covered by this Guide, this Article is
only applicable for determining Fnc for the discretely braced bottom (compression) flange in
regions of negative flexure at the strength limit state for use in Eq. 6.10.8.1.1-1 if the provisions

NSBA Guide to Navigating Routine Steel Bridge Design / 242


of the optional Appendix A6 (see the Discussion of Appendix A6 in this Guide) are not used,
which is not recommended for rolled beam bridges; otherwise, this Article is not applicable at the
strength limit state.
For the routine multi-span continuous plate girder bridges covered by this Guide, this Article is
applicable for determining Fnc for the discretely braced bottom (compression) flange in regions of
negative flexure at the strength limit state for use in Eq. 6.10.8.1.1-1 if the section is a slender web
section, or if the section satisfies the restrictions specified in Article 6.10.6.2.3 (see the Discussion
of Article 6.10.6.2.3 in this Guide) and the provisions of the optional Appendix A6 are not used to
design the section, which is not recommended for compact web sections in particular; if the
provisions of Appendix A6 are used in the latter case, this Article is not applicable at the strength
limit state.
The Article is applicable for determining Fnc for the discretely braced top (compression) flange of
the noncomposite section in regions of positive flexure and for the discretely braced bottom
(compression) flange of the noncomposite section in regions of negative flexure during
construction for use in Eq. 6.10.3.2.1-2; the one exception being if the section qualifies as a
compact web or noncompact web section (essentially true for rolled beam sections, but not
necessarily the case for plate girder sections) and the provisions of Article A6.3.3 are used to
compute Mnc for lateral-torsional buckling to account for the beneficial effect of the St. Venant
torsional constant, J.
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.8.2.2 Local Buckling Resistance


Determination of applicability, Simple Span Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 243


Discussion:
The provisions of this Article are used to compute the nominal flexural resistance, Fnc, of a
discretely braced compression flange based on flange local buckling (FLB) for use in Eq.
6.10.8.1.1-2 or 6.10.3.2.1-2, as applicable.
FLB is a limit state of buckling of a compression flange within a cross-section and is a function of
the compression-flange slenderness, λf = bfc/2tfc. For determining the FLB resistance (refer to
Figure C6.10.8.2.1-1), λpf locates Anchor Point 1 that separates sections with compact flanges from
sections with noncompact flanges (see Table C6.10.8.2.2-1). A member with a compression-flange
slenderness at or below the compact flange limit, λpf, is able to achieve the so-called “plateau
strength” or maximum potential FLB resistance (Fmax in Figure C6.10.8.2.1-1) of RbRhFyc, which
is independent of the compression-flange slenderness. λrf locates Anchor Point 2 that separates
sections with noncompact flanges from sections with slender flanges and is the point where the
inelastic and elastic FLB resistances are the same (with the resistance at this point assumed to be
RbFyr). The inelastic FLB resistance of a noncompact flange is treated as a linear function of the
compression-flange slenderness. An elastic FLB equation for slender flanges is not provided in the
specifications because for most practical bridge-girder sections, including sections used in the
routine I-girder bridges covered by this Guide (i.e., with Fyc ≤ 90 ksi), elastic FLB will not control
as λf is limited to a practical maximum value of 12.0 (see the Discussion of Article 6.10.2.2 in this
Guide). The equations in this Article assume the section is a slender web section whether it is or
not. See the Commentary for Article 6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in this
Guide for further discussion on the definition and categorization of compact web, noncompact
web, and slender web sections. The FLB resistance for moment gradient cases is treated the same
as that for the case of uniform major-axis bending; i.e., the relatively minor influence of moment-
gradient effects on the FLB resistance is neglected.
For design checks where the flexural resistance is based on FLB, fbu and fℓ in Eqs. 6.10.8.1.1-1 and
6.10.3.2.1-2 are determined as the stress at the section under consideration (see the Discussion
Article 6.10.1.6 in this Guide).
This Article is conditionally applicable for routine steel multi-span continuous I-girder bridges,
and only partially applicable to routine simple span I-girder bridges covered by this Guide, as
described further in the Discussion of Article 6.10.8.2.1 in this Guide (disregarding the language
pointing to Article A6.3.3).
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the

NSBA Guide to Navigating Routine Steel Bridge Design / 244


designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.8.2.3 Lateral Torsional Buckling Resistance


Determination of applicability, Simple Span Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article are used to compute the nominal flexural resistance, Fnc, of a
discretely braced compression flange based on lateral-torsional buckling (LTB) for use in Eq.
6.10.8.1.1-2 or 6.10.3.2.1-2, as applicable.
LTB is a limit state of buckling of an unbraced length involving lateral deflection and twist and is
a function of the unbraced length, Lb. For determining the LTB resistance (refer to Figure
C6.10.8.2.1-1), Lp locates Anchor Point 1 that separates compact unbraced lengths from
noncompact unbraced lengths. A member braced at or below the compact unbraced length limit is
able to achieve the so-called “plateau strength” or maximum potential LTB resistance (Fmax in
Figure C6.10.8.2.1-1) of RbRhFyc under uniform vertical bending, which is independent of the
unbraced length. Note that in many cases, it will not be economical to brace the girder at a distance
equal to Lp or below in order to reach Fmax, particularly under uniform bending conditions for
which Cb is equal to 1.0. Lr locates Anchor Point 2 that separates sections with noncompact
unbraced lengths from sections with slender unbraced lengths and is the point where the inelastic
and elastic LTB resistances are the same (with the resistance at this point assumed to be RbFyr).
The inelastic LTB resistance of a noncompact unbraced length is treated as a linear function of the
unbraced length. Unbraced lengths greater than Lr are termed slender unbraced lengths and their
resistance is controlled by elastic LTB. LTB in the elastic range is of primary importance for
relatively slender girders braced at longer than normal intervals, which most typically occurs
during a temporary construction condition. Eq. 6.10.8.2.3-8 for the elastic LTB stress, Fcr, is a
conservative approximation of the exact beam-theory solution for the elastic LTB resistance of a
doubly symmetric I-section (assuming load-height effects are not considered and that the St.
Venant torsional constant, J, is equal to zero). The equations in this Article assume the section is
a slender web section whether it is or not. See the Commentary for Article 6.10.6.2.3 and the
Discussion of Article 6.10.6.2.3 in this Guide for further discussion on the definition and
categorization of compact web, noncompact web, and slender web sections. Further details on
these LTB resistance equations are provided in the Commentary for this Article.
The provisions of this Article are conditionally applicable for routine steel multi-span continuous
I-girder bridges, and only partially applicable for routine simple span I-girder bridges covered by
this Guide, as described further in the Discussion of Article 6.10.8.2.1 in this Guide.
The solid curve in Figure C6.10.8.2.1-1 is for uniform bending represented by the equations given
in this Article. The dashed curve in Figure C6.10.8.2.1-1 shows the solid curve scaled by the
moment-gradient modifier, Cb, under moment-gradient conditions, which can result in the plateau

NSBA Guide to Navigating Routine Steel Bridge Design / 245


strength (Fmax) for lateral-torsional buckling to be reached at significantly larger unbraced lengths
under moment-gradient conditions when the effects of the moment gradient are included in
determining the limits on the unbraced length, Lb. Refer to Article D6.4.1 for the appropriate
equations to use under these conditions (i.e., the equations representing the dashed curve when Cb
is calculated and is greater than 1.0), which is strongly encouraged (see the Discussion of Article
D6.4.1 in this Guide). The calculation of Cb is discussed further below.
For design checks where the flexural resistance is based on LTB, fbu and fℓ in Eqs. 6.10.8.1.1-1 and
6.10.3.2.1-2 are to be taken as the largest values throughout the unbraced length in the flange under
consideration (see the Discussion of Article 6.10.1.6 in this Guide).
For unbraced lengths containing a transition to a smaller section at a distance less than or equal to
20 percent of the unbraced length from the brace point with the smaller moment, Article 6.10.8.2.3
permits the LTB resistance to be determined assuming the transition to the smaller section does
not exist. This is provided the lateral moment of inertia of the flange or flanges of the smaller
section is also equal to or larger than one-half of the corresponding value in the larger section. The
moment gradient modifier, Cb, may be applied in this case. Otherwise, the LTB resistance is to be
taken as the smallest resistance within the unbraced length and Cb must be taken equal to 1.0. The
resulting resistance must not be exceeded anywhere along the unbraced length. This conservative
approximation is based on replacing the nonprismatic member with an equivalent prismatic
member. To avoid significant reductions in the LTB resistance, it is desirable to consider locating
flange transitions (satisfying the preceding moment of inertia requirement) within 20 percent of
the unbraced length from the brace point with the smaller moment, where practical.
Simple Span Bridges:
For simple span bridges, the moment-gradient modifier, Cb, may conservatively be taken equal to
1.0 when checking LTB of the critical noncomposite section in regions of positive flexure during
construction; otherwise, consult Figure C6.4.10 for the appropriate calculation of Cb in these
regions (see the Discussion of Article C6.4.10 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
For multi-span continuous bridges, it is strongly recommended that as a minimum the moment-
gradient modifier, Cb, be calculated when checking LTB of the first unbraced length on either side
of the interior piers as described below, specifically for prismatic unbraced lengths or for
nonprismatic unbraced lengths satisfying the 20 percent rule described above. The unbraced
lengths on either side of the pier should be checked to determine which side will yield the lower
value of Cb. Cb may conservatively be taken equal to 1.0 when checking LTB of the critical
noncomposite section in regions of positive flexure during construction; otherwise, consult Figure
C6.4.10 for the appropriate calculation of Cb in these regions (see the Discussion of Article C6.4.10
in this Guide).
For multi-span continuous rolled beam bridges, if the provisions of the optional Appendix A6 are
not used, which is not recommended for rolled beam bridges, Cb should be calculated for the first
unbraced length on either side of the interior piers using Eq. 6.10.8.2.3-7 with f1 taken equal to fo
(Eq. 6.10.8.2.3-10). Major-axis bending stresses are used to calculate Cb since dead and live load
bending moments are applied to different sections in composite girders, which is significant when

NSBA Guide to Navigating Routine Steel Bridge Design / 246


the nominal flexural resistance is not permitted to exceed the moment at first yield. The factored
stresses, f2 and f0, at each end of the unbraced length are taken as positive in compression and
negative in tension in Eq. 6.10.8.2.3-7. The provisions of Article D6.4.1 should then be employed
to determine the shift in the anchor point, Lp, and the corresponding nominal LTB resistance (see
the Discussion of Article D6.4.1 in this Guide).
For multi-span continuous steel plate girder bridges, if the section is a slender web section or if the
section satisfies the restrictions specified in Article 6.10.6.2.3 (see the Discussion of Article
6.10.6.2.3 in this Guide) and the provisions of the optional Appendix A6 are not used to design
the section, which is not recommended for compact web sections in particular, Cb should be
calculated for the first unbraced length on either side of the interior piers using Eq. 6.10.8.2.3-7
with f1 taken equal to fo (Eq. 6.10.8.2.3-10). Major-axis bending stresses are used to calculate Cb
since dead and live load bending moments are applied to different sections in composite girders,
which is significant when the nominal flexural resistance is not permitted to exceed the moment
at first yield. The factored stresses, f2 and f0, at each end of the unbraced length are taken as positive
in compression and negative in tension in Eq. 6.10.8.2.3-7. The provisions of Article D6.4.1 should
then be employed to determine the shift in the anchor point, Lp, and the corresponding nominal
LTB resistance (see the Discussion of Article D6.4.1 in this Guide).
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.8.3 Flexural Resistance Based on Tension Flange Yielding


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
Simple Span Bridges:
This Article is not applicable to the routine simple span I-girder bridges covered by this Guide at
the strength limit state because simple span bridges are subject to positive flexure only.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:

NSBA Guide to Navigating Routine Steel Bridge Design / 247


The provisions of this Article are used to compute the nominal flexural resistance, Fnt, of a
discretely braced top (tension) flange at the strength limit state in regions of negative flexure in
multi-span continuous steel I-girder bridges for use in Eq. 6.10.8.1.2-1 (see the Discussion of
Article 6.10.8.1.2 in this Guide). The nominal flexural resistance is based only on nominal yielding
because flange local buckling and lateral-torsional buckling are not a consideration for flanges in
tension. The equation in this Article assumes the section is a slender web section whether it is or
not. See the Commentary for Article 6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in this
Guide for further discussion on the definition and categorization of compact web, noncompact
web, and slender web sections.
The top flange in regions of negative flexure would only be discretely braced if the shear
connectors are intentionally omitted in regions of negative flexure, which is dependent on the
preferences of the Owner-agency but not recommended, and if the Engineer deems that the top
flange is not sufficiently encased by the concrete deck. Therefore, this Article is conditionally
applicable at the strength limit state to the design of the top flange in regions of negative flexure
in routine steel multi-span continuous I-girder bridges, as described further in the Discussion of
Article 6.10.8.1.2 in this Guide.
For sections in which Myt > Myc, Eq. 6.10.8.3-1 does not control and tension flange yielding need
not be checked, where Myc and Myt are the yield moments with respect to the compression and
tension flange, respectively, determined as specified in Article D6.2 (see the Discussion of Article
D6.2 in this Guide).
For further information on strength limit state design for flexure, consult Section 6.5.6 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download
from the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It
calculates the design loads and resulting stresses, and the corresponding resistances, in accordance
with the provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of
the designer. Other commercial software packages with the ability to analyze and design routine
steel I-girder bridges are also available. Users should verify the capabilities, assumptions, and
general correctness of any program’s calculations prior to initial use.

6.10.9 Shear Resistance

6.10.9.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:

NSBA Guide to Navigating Routine Steel Bridge Design / 248


This Article provides general provisions related to the determination of the factored shear
resistance of a beam or girder, which are partially applicable to the routine steel I-girder bridges
covered by this Guide.
The provisions in this Article dealing with hybrid I-shaped members and web longitudinal
stiffeners are not applicable to the routine steel I-girder bridges covered by this Guide.
The general design equation for shear is given by Eq. 6.10.9.1-1. A helpful flowchart to guide
designers through the design provisions for the shear design of I-sections is given in Figure
C6.10.9.1-1.
This Article also provides the maximum stiffener spacing requirements for interior and end web
panels of beams or girders containing web transverse stiffeners. Interior web panels with transverse
stiffener spacings exceeding 3D, where D is the web depth, are to be considered unstiffened. End
web panels with transverse stiffener spacings exceed 1.5D are to be considered unstiffened. An
end panel is defined as a web panel adjacent to a discontinuous end of a girder. Web panels over
the interior piers of a continuous span are classified as interior web panels and should not be
classified as end panels. For the routine steel rolled-beam bridges covered by this Guide, web
transverse stiffeners are typically not needed, except for use as cross-frame or diaphragm
connection plates. When web transverse stiffeners are required, they are to be designed according
to the provisions of Article 6.10.11.1.
To find a suitable stiffener spacing, the designer should select a stiffener spacing that is less than
the specified maximum spacing described above and that satisfies the general shear equation (Eq.
6.10.9.1-1) using the maximum factored shear, Vu, in the web panel under consideration at the
strength limit state and the applicable nominal shear resistance, Vn, (Article 6.10.9.2 or 6.10.9.3).
This can sometimes involve a trial-and-error process. It is often helpful to use constant stiffener
spacings over specified ranges. Since the design shear varies over the length of the girder, this
procedure may have to be repeated for several ranges along the length of the girder. Cross-frame
or diaphragm connection plates can be considered to act as transverse stiffeners. Therefore, ranges
between connection plates are typically investigated in laying out the stiffeners.
The AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to Design for
Constructability and Fabrication provides practical guidance for economical proportioning of
girder webs, such as the general preference of fabricators for a minimum web thickness of ½" to
reduce the deformation of the web and the potential for weld defects during fabrication.
Changes in the web thickness along the girder in plate-girder bridges preferably should be made
at field splices. In field sections over interior piers in continuous spans, the web thickness may
have to be increased (typically in 1/16-inch increments) over the thickness provided in adjacent
regions in positive flexure in some instances; e.g., if the web bend-buckling resistance is exceeded
in regions of negative flexure at the service limit state (see the Discussion of Articles 6.10.1.9.1
and 6.10.4.2.2 in this Guide).
A useful guideline for determining the trade-off between adding more transverse stiffeners versus
increasing the thickness of the web material in routine plate-girder bridges is that approximately 4
to 5 pounds of web material should be saved for every 1 pound of stiffener material added. This
higher unit cost reflects that additional fabrication effort is required per pound of stiffener still than

NSBA Guide to Navigating Routine Steel Bridge Design / 249


per pound of girder web steel. Generally, an unstiffened web is not the most economical alternative
for a plate-girder bridge. The best solution usually includes some transverse stiffeners over the
piers and near the abutments; a so-called partially stiffened web. However, it should be pointed
out that a partially stiffened web can, and probably will, include a number of unstiffened panels.
This is a natural outcome of an economical web design.
For further information on strength limit state design for shear, consult Section 6.5.7 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.9.2 Nominal Resistance of Unstiffened Webs


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article are used to calculate the nominal shear resistance, Vn, of unstiffened
web panels (i.e., interior web panels with transverse stiffener spacings exceeding 3D, or end web
panels with transverse stiffener spacings exceeding 1.5D), at the strength limit state. The
provisions are applicable to the routine steel I-girder bridges covered by this Guide; in particular,
they are typically used to determine the nominal shear resistance for a rolled-beam web, and to
determine the regions where transverse stiffeners are required in a stiffened plate-girder web.
For an unstiffened web panel, the nominal shear resistance is based on either the shear yield or
shear buckling resistance (Eq. 6.10.9.2-1). C in Eq. 6.10.9.2-1 is the ratio of the shear-buckling
resistance to the shear yield strength. Its value depends on the web slenderness, and the equations
used to calculate C are given in Article 6.10.9.3.2. The plastic shear force, Vp (Eq. 6.10.9.2-2), is
equal to the web area times the assumed shear yield strength. The assumed shear yield strength is
the web yield strength divided by the square root of 3, or 0.58 (based on the von Mises yield
criterion). When C equals 1.0, the nominal shear resistance is based on shear yielding. Otherwise,
the nominal shear resistance is based on shear buckling. For unstiffened web panels, the shear-
buckling coefficient, k, is always taken equal to 5.0 in the computation of C.
A helpful flowchart to guide designers through the design provisions for the shear design of I-
sections is given in Figure C6.10.9.1-1.

NSBA Guide to Navigating Routine Steel Bridge Design / 250


The AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to Design for
Constructability and Fabrication provides practical guidance for economical proportioning of
girder webs, such as the general preference of fabricators for a minimum web thickness of ½" to
reduce the deformation of the web and the potential for weld defects during fabrication.
Changes in the web thickness along the girder in plate-girder bridges preferably should be made
at field splices. In field sections over interior piers in continuous spans, the web thickness may
have to be increased (typically in 1/16-inch increments) over the thickness provided in adjacent
regions in positive flexure in some instances; e.g., if the web bend-buckling resistance is exceeded
in regions of negative flexure at the service limit state (see the Discussion of Articles 6.10.1.9.1
and 6.10.4.2.2 in this Guide).
A useful guideline for determining the trade-off between adding more transverse stiffeners versus
increasing the thickness of the web material in routine plate-girder bridges is that approximately 4
to 5 pounds of web material should be saved for every 1 pound of stiffener material added. This
higher unit cost reflects that additional fabrication effort is required per pound of stiffener still than
per pound of girder web steel. Generally, an unstiffened web is not the most economical alternative
for a plate-girder bridge. The best solution usually includes some transverse stiffeners over the
piers and near the abutments; a so-called partially stiffened web. Transverse stiffeners (other than
connection plates for cross-frames or diaphragms) are typically not required in routine rolled-beam
bridges. However, it should be pointed out that a partially stiffened web can, and probably will,
include a number of unstiffened panels. This is a natural outcome of an economical web design.
For further information on strength limit state design for shear, consult Section 6.5.7 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.9.3 Nominal Resistance of Stiffened Webs

6.10.9.3.1 General
Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:

NSBA Guide to Navigating Routine Steel Bridge Design / 251


This Article points to the Articles containing the equations used to calculate the nominal shear
resistance, Vn, of stiffened web panels (i.e., interior web panels with transverse stiffener spacings
not exceeding 3D, or end web panels with transverse stiffener spacings not exceeding 1.5D), at the
strength limit state. The provisions are applicable to the routine steel plate girder bridges covered
by this Guide (simple span and multi-span continuous) and are not applicable to the routine steel
rolled beam bridges covered by this Guide (simple span or multi-span continuous), as web
transverse stiffeners are typically not needed on rolled beam bridges except for use as cross-frame
or diaphragm connection plates.
The nominal shear resistance of transversely stiffened interior web panels is taken as the shear-
yielding resistance or the sum of the shear-buckling resistance and the post-buckling shear
resistance due to tension-field action (Article 6.10.9.3.2). The nominal shear resistance of
transversely stiffened end web panels is taken as the shear-yielding or shear-buckling resistance
(Article 6.10.9.3.3). Web longitudinal stiffeners are not applicable to the routine steel I-girder
bridges covered by this Guide.
A helpful flowchart to guide designers through the design provisions for the shear design of I-
sections is given in Figure C6.10.9.1-1.
The AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to Design for
Constructability and Fabrication provides practical guidance for economical proportioning of
girder webs, such as the general preference of fabricators for a minimum web thickness of ½" to
reduce the deformation of the web and the potential for weld defects during fabrication.
Changes in the web thickness along the girder in plate-girder bridges preferably should be made
at field splices. In field sections over interior piers in continuous spans, the web thickness may
have to be increased (typically in 1/16-inch increments) over the thickness provided in adjacent
regions in positive flexure in some instances; e.g., if the web bend-buckling resistance is exceeded
in regions of negative flexure at the service limit state (see the Discussion of Articles 6.10.1.9.1
and 6.10.4.2.2 in this Guide).
A useful guideline for determining the trade-off between adding more transverse stiffeners versus
increasing the thickness of the web material in routine plate-girder bridges is that approximately 4
to 5 pounds of web material should be saved for every 1 pound of stiffener material added. This
higher unit cost reflects that additional fabrication effort is required per pound of stiffener still than
per pound of girder web steel. Generally, an unstiffened web is not the most economical alternative
for a plate-girder bridge. The best solution usually includes some transverse stiffeners over the
piers and near the abutments; a so-called partially stiffened web. However, it should be pointed
out that a partially stiffened web can, and probably will, include a number of unstiffened panels.
This is a natural outcome of an economical web design.
It should be noted that cross-frame and diaphragm connection plates also function as transverse
stiffeners. Strategic determination of the cross-frame spacing can sometimes make a significant
difference in the web design of a steel plate girder, particularly in longer spans where the addition
of one more line of cross-frames or diaphragms may result in an effective stiffener spacing less
than 3D without having to provide additional transverse stiffeners. Other strategies may include
using a tighter cross-frame or diaphragm spacing nearer to the interior supports, and wider spacing

NSBA Guide to Navigating Routine Steel Bridge Design / 252


near mid-span, of multi-span continuous steel plate girder bridges, again with the goal of providing
an effective stiffener spacing less than 3D without having to provide additional transverse
stiffeners in higher shear regions. These strategies should be used judiciously, recognizing the
impact of such changes on the flexural resistance of the girder and on the overall cost of the steel
superstructure.
For further information on strength limit state design for shear, consult Section 6.5.7 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.9.3.2 Interior Panels


Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
The provisions of this Article are used to calculate the nominal shear resistance, Vn, of a stiffened
interior web panel, or a web panel not adjacent to a discontinuous end of a girder with a transverse
stiffener spacing not exceeding 3D. Web panels over the interior piers of a continuous span are
classified as interior web panels and should not be classified as end panels. The provisions are
applicable to the routine steel plate girder bridges covered by this Guide (simple span and multi-
span continuous). The provisions are partially applicable to the routine steel rolled beam bridges
covered by this Guide (simple span or multi-span continuous), because in those cases the
provisions in this Article are only used for these bridges to calculate the constant C to be used in
determining Vn of the unstiffened web (Article 6.10.9.2) using a shear-buckling coefficient, k,
equal to 5.0.
For an interior panel of a stiffened plate-girder web, the equation to use to calculate Vn depends on
the ratio of the average area of the flanges within the panel relative to the area of the web (Eq.
6.10.9.3.2-1). If the inequality in this equation is satisfied, the panel can develop the full post-
buckling shear resistance due to tension-field action (i.e., Eq. 6.10.9.3.2-2 is used).
Eq. 6.10.9.3.2-2 includes two components; the shear yield or shear buckling resistance, and the
post-buckling tension-field resistance. The shear yield or shear buckling resistance component is

NSBA Guide to Navigating Routine Steel Bridge Design / 253


identical to the equation used previously for unstiffened webs (Eq. 6.10.9.2-1); shear in this case
is assumed to be carried by “beam action”. The second component is the post-buckling tension-
field resistance. This component is analogous to the tension diagonals of a Pratt truss; that is, after
buckling the tension forces are resisted by membrane action of the web while the compression
forces are resisted by the transverse stiffeners in combination with the adjacent portions of the
web. The total shear resistance is either the shear yield resistance (i.e. the second term in the
equation goes to zero when C = 1.0), or the sum of the shear buckling resistance and the post-
buckling resistance due to tension-field action. The constant C for use in Eq. 6.10.9.3.2-2 is
calculated from Eq. 6.10.9.3.2-4, 6.10.9.3.2-5, or 6.10.9.3.2-6, as applicable, using the shear
buckling coefficient, k, given by Eq. 6.10.9.3.2-7 which is dependent on the stiffener spacing, do.
When Eq. 6.10.9.3.2-1 is not satisfied, the average area of the flanges within the panel is small
relative to the area of the web and the full post-buckling resistance generally cannot be developed.
In this case, Vn is calculated from Eq. 6.10.9.3.2-8. This shear resistance equation is based on a
lesser level of the post-buckling resistance, which neglects the increase in stress within the wedges
of the web panel outside of the tension band implicitly included in the tension-field model. The
equation is similar in form to Eq. 6.10.9.3.2-2, except that this equation has an extra do/D term in
the denominator.
See also the Discussion of Article 6.10.9.3.1 in this Guide for further explanation of the basic
concepts behind the shear design provisions, helpful design tips, handy references to other
guideline documents and design examples, and comments on design software available to help
automate the shear design checks.

6.10.9.3.3 End Panels


Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
The provisions of this Article are used to calculate the nominal shear resistance, Vn, of a stiffened
end web panel, or a web panel adjacent to a discontinuous end of a girder with a transverse stiffener
spacing not exceeding 1.5D. Web panels over the interior piers of a continuous span are classified
as interior web panels and should not be classified as end panels. The provisions are applicable to
the routine steel plate girder bridges covered by this Guide (simple span and multi-span
continuous) and are partially applicable to the routine steel rolled beam bridges covered by this
Guide (simple span or multi-span continuous), as the provisions in this Article are only used for
these bridges to calculate the constant C to be used in determining Vn of the unstiffened web end
panel (Article 6.10.9.2) using a shear-buckling coefficient, k, equal to 5.0.
For the end panels of a stiffened web, Vn is based on either shear yield or shear buckling. The same
equation is used to calculate Vn that is used for an unstiffened web panel (Eq. 6.10.9.2-1). However,
in this case, the shear-buckling coefficient, k (Eq. 6.10.9.3.2-7), used in determining C is based on
the spacing from the support to the first stiffener adjacent to the support, which cannot exceed

NSBA Guide to Navigating Routine Steel Bridge Design / 254


1.5D. The shear in stiffened end panels is limited to either the shear-yield or shear-buckling
resistance to provide an anchor for the tension field in adjacent stiffened interior panels. In other
words, it absorbs any imbalance of the computed horizontal component of the diagonal tensile
stress in the adjacent panels.
It may initially seem counterintuitive to recognize tension field action in the calculation of the
shear resistance of interior panels but not in end panels, as if the full resistance of the web is being
discounted at the point where the applied shear is greatest. However, keep in mind that by limiting
the transverse stiffener spacing to 1.5D in end panels, those panels can develop a higher initial
shear buckling resistance, comparable to or greater than the sum of the shear buckling and post-
buckling resistance of the interior panels.
See also the Discussion of Article 6.10.9.3.1 in this Guide for further explanation of the basic
concepts behind the shear design provisions, helpful design tips, handy references to other
guideline documents and design examples, and comments on design software available to help
automate the shear design checks.

6.10.10 Shear Connectors

6.10.10.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides general discussion on the purpose of shear connectors and the provision of
shear connectors throughout different regions of the span length. Only stud shear connectors and
the provisions dealing with straight continuous composite bridges are applicable for the routine
steel I-girder bridges covered by this Guide.
In general, in the absence of contradicting Owner-agency policy, it is recommended for the routine
steel I-girder bridges covered by this Guide that shear connectors be provided throughout the
length of the bridge, including in regions of negative flexure in multi-span continuous bridges,
because doing so helps to better control cracking of the deck in regions of negative flexure. Shear
connectors must be provided in these regions where the longitudinal deck reinforcement is
considered in the computation of the composite section properties, which is recommended to allow
for the use of a slightly smaller top flange than bottom flange in these regions. The provision of
shear connectors in these regions also allows the concrete deck to be considered effective in tension
at the fatigue and service limit states if other requirements are satisfied (see the Discussion of
Articles 6.6.1.2.1 and 6.10.4.2.1 in this Guide). If shear connectors are omitted in these regions,
which depends on the preferences of the Owner-agency but is not recommended, other provisions
related to the shear connectors (see the Discussion of Article 6.10.10.3 in this Guide) and
longitudinal reinforcing in the deck (see the Discussion of Article 6.10.1.7 in this Guide) apply.

6.10.10.1.1 Types
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:

NSBA Guide to Navigating Routine Steel Bridge Design / 255


This Articles deals with the types of shear connectors permitted and the permissible proportions
of a stud-type shear connector. Only the provisions in this Article related to stud-type shear
connectors are applicable for the routine steel I-girder bridges covered by this Guide. Channel-
type shear connectors are rarely, if ever, used in modern steel bridge design in the U.S.

6.10.10.1.2 Pitch
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
The provisions of this Article are used to determine the pitch, p, of shear connectors to satisfy the
fatigue limit state and are partially applicable to the routine steel I-girder bridges covered by this
Guide as described further below. The resulting number of shear connectors is then checked
against the number required to satisfy the strength limit state (see the Discussion of Article
6.10.10.4 in this Guide). The fatigue limit state will typically govern the number of shear
connectors.
Only the longitudinal fatigue shear range, Vfat, needs to be considered in determining the horizontal
fatigue shear range, Vsr, when designing shear connectors for the routine steel I-girder bridges
covered in this Guide. The radial fatigue shear range, Ffat, is intended to reflect the effects of torsion
in the girders due to curvature, significant skew, or discontinuous cross-frame or diaphragm lines;
for the purposes of this Guide, routine steel I-girder bridges are defined as not having any of these
characteristics, and so it is not necessary to consider the radial fatigue shear range, Ffat.
The vertical shear range, Vf, used to calculate Vfat is determined using the fatigue live load (see the
Discussion of Article 3.6.1.4 in this Guide) shears. The fatigue live load is placed in a single lane
with a dynamic load allowance of 15 percent applied (Table 3.6.2.1-1). The shear range is the
algebraic difference of the maximum and minimum live load plus impact shears; for a simple span,
the minimum live load plus impact shear is zero.
The shears are factored for the Fatigue I load combination (Table 3.4.1-1) when the 75-year single
lane Average Daily Truck Traffic (ADTT)SL (see the Discussion of Article 3.6.1.4.2 in this Guide)
is greater than or equal to 1,090 trucks per day. Otherwise, the shears are factored for the Fatigue
II load combination. For a fatigue design life other than 75 years and/or a number of stress cycles
per truck passage (n from Table 6.6.1.2.5-2) other than 1.0, see the Commentary for Article
6.10.10.2. It is recommended that the moment of inertia, I, and first moment of the deck area, Q,
used to calculate Vfat be computed using the short-term composite section (see the Discussion of
Article 6.10.1.1.1b in this Guide). See the Discussion of Article 6.10.10.2 in this Guide for the
calculation of the shear fatigue resistance, Zr, of an individual stud shear connector.
The pitch can, and should, vary along the length of the girder. Typically, the calculation of the
required pitch is performed at uniformly spaced points along the length of the girder (e.g., at 1/10th
points or at 1/20th points, etc.), and then the specified pitch is determined over various regions of
the girder such that the specified pitch within the region is less than or equal to the required pitch
at the point (or points) under consideration within that region. The lengths of each region do not
need to be equal; in general, smaller regions are used where the required pitch is tighter and is
changing at an increased rate along the length of the girder, typically in areas near supports where

NSBA Guide to Navigating Routine Steel Bridge Design / 256


the shear is largest. The specified pitch (in even-inch increments) within each region is generally
conservatively adjusted to produce a number of spaces that sums to the length of the region. The
number of regions can be determined at the discretion of the designer. The pitch in each region
must not be less than six stud diameters and must not exceed 48.0 inches (or 24.0 inches if the web
depth is less than 24.0 inches). In general, the use of an excessive number of regions is discouraged
as it rarely saves a significant number of shear connectors. Likewise, using only one or a very few
regions is also discouraged as it would result in providing an excessive number of shear connectors.
For more explanation and examples of the determination of the design of shear connectors at the
fatigue and strength limit states, see Section 6.3.6.3 of Reference Manual for NHI Course 130122,
Design and Evaluation of Steel Bridges for Fatigue and Fracture, Section 6.6.2 of Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures as well as FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-
Span Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span
Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span
Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It performs
design calculations addressing the demand on, and resistance of, shear connectors at the fatigue and
strength limit states in accordance with the provisions of the AASHTO LRFD BDS, greatly reducing
the time and effort required of the designer. Other commercial software packages with the ability to
analyze and design routine steel I-girder bridges are also available. Users should verify the
capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.10.10.1.3 Transverse Spacing


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides the transverse spacing requirements (i.e., across the width of the top flange)
and the clear edge distance requirements for shear connectors. The provisions of this Article are
applicable to the routine steel I-girder bridges covered by this Guide. Typically, the transverse
spacing of shear connectors is held constant along the entire length of a girder for simplicity of
detailing and fabrication.
See also the Discussion of Articles 6.10.10.1.2 and 6.10.10.4.2 in this Guide for related
explanations of shear connector design.

6.10.10.1.4 Cover and Penetration


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides the cover and penetration requirements for shear connectors and is applicable
to the routine steel I-girder bridges covered by this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 257


These provisions only require 2 in. of penetration into the concrete deck. Most designers try to set
the height of shear connectors so that they penetrate above the bottom mat of deck reinforcing.
The girders in the routine steel I-girder bridges covered by this Guide are typically plumb after
deck placement, but the deck will likely have a cross-slope. This should be accounted for when
determining the specified height of the shear connectors; shear connectors closer to one edge of
the top flange will likely have a different penetration than those closer to the other edge. In
addition, for plate-girder bridges, the haunch dimension is typically measured from the top of the
web (or bottom of the top flange) and the top-flange thickness may vary along the length of the
girder. This should also be accounted for when determining the specified height of the shear
connectors.
Regardless of these variations in penetration due to deck cross slope or top-flange thickness
changes, the use of different shear connector heights along the length of a girder or across the
width of the top flange is rarely necessary and is discouraged because it adds unnecessary cost and
complexity to the fabrication of the girders. Instead, designers should choose a shear connector
height that provides at least the minimum 2 in. penetration into the deck and provides at least the
minimum required cover to the top of the deck, and allow the penetration into the deck to vary as
needed within the range defined by those two criteria.

6.10.10.2 Fatigue Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
The provisions of this Article are used to determine the fatigue resistance of an individual shear
connector, Zr. Zr is used in the calculation of the required pitch, p, at the fatigue limit state (see the
Discussion of Article 6.10.10.1.2 in this Guide). Only the provisions for stud shear connectors
should be considered applicable to the routine I-girder bridges covered by this Guide.
When the 75-year single lane Average Daily Truck Traffic (ADTT)SL (see the Discussion of Article
3.6.1.4.2 in this Guide) is greater than or equal to 1,090 trucks per day, the fatigue shear resistance
for infinite life determined from Eq. 6.10.10.2-1 is used for Zr. Otherwise, the fatigue shear
resistance for finite life determined from Eq. 6.10.10.2-2 is used for Zr. For a fatigue design life
other than 75 years and/or a number of stress cycles per truck passage (n from Table 6.6.1.2.5-2)
other than 1.0, see the Commentary for this Article.
For more explanation and examples of the determination of the design of shear connectors at the
fatigue and strength limit states, see Section 6.3.6.3 of Reference Manual for NHI Course 130122,
Design and Evaluation of Steel Bridges for Fatigue and Fracture, Section 6.6.2 of Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures as well as FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-
Span Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span
Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span
Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It performs

NSBA Guide to Navigating Routine Steel Bridge Design / 258


design calculations addressing the demand on, and resistance of, shear connectors at the fatigue and
strength limit states in accordance with the provisions of the AASHTO LRFD BDS, greatly reducing
the time and effort required of the designer. Other commercial software packages with the ability to
analyze and design routine steel I-girder bridges are also available. Users should verify the
capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.10.10.3 Special Requirements for Points of Permanent Load Contraflexure


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
Simple Span Bridges:
These provisions are not applicable to simple span bridges, which are only subject to positive
flexure.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
The provisions of this Article are to be used only when shear connectors are omitted in regions of
negative flexure, which depends on the preferences of the Owner-agency but is not recommended.
The provisions are used to determine the number of additional shear connectors that need to be
provided on each side of the points of permanent load contraflexure to develop the fatigue force
in the longitudinal reinforcement in the deck due to the negative factored fatigue live load moment
at the interior support.

6.10.10.4 Strength Limit State

6.10.10.4.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article are used to determine the total number of shear connectors required
at the strength limit state within specified regions of the span and are applicable to the routine steel
I-girder bridges covered by this Guide. The provisions are used to the determine the factored shear
resistance of a single shear connector at the strength limit state, Qr (Eq. 6.10.10.4.1-1), and the
minimum number of shear connectors, n (Eq. 6.10.10.4.1-2), that are required over the region of
the span under consideration at the strength limit state. For the calculation of the nominal shear
resistance, Qn, used in the determination of Qr, see the Discussion of Article 6.10.10.4.3 in this
Guide. n is determined as the nominal shear force, P, for the region under consideration (see the
Discussion of Article 6.10.10.4.2 in this Guide) divided by Qr.
For more explanation and examples of the determination of the design of shear connectors at the
fatigue and strength limit states, see Section 6.3.6.3 of Reference Manual for NHI Course 130122,

NSBA Guide to Navigating Routine Steel Bridge Design / 259


Design and Evaluation of Steel Bridges for Fatigue and Fracture, Section 6.6.2 of Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures as well as FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-
Span Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span
Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span
Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It performs
design calculations addressing the demand on, and resistance of, shear connectors at the fatigue and
strength limit states in accordance with the provisions of the AASHTO LRFD BDS, greatly reducing
the time and effort required of the designer. Other commercial software packages with the ability to
analyze and design routine steel I-girder bridges are also available. Users should verify the
capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.10.10.4.2 Nominal Shear Force


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
The provisions of this Article are used to compute the nominal shear force, P, at the strength limit
state within the region of the span under consideration. P is used in the calculation of the number
of studs, n, required within the region under consideration (see the Discussion of Article
6.10.10.4.1 in this Guide).
The number of rows required in the region under consideration at the strength limit state, NRows,
can be determined as n divided by the number of shear connectors per row. The required pitch, p,
in the region at the strength limit state (for comparison with the required pitch, p, determined at
the fatigue limit state – see the Discussion of Article 6.10.10.1.2 in this Guide) can then be
calculated as the total length of the region (in inches) divided by (NRows -1).
The total radial forces in the concrete deck, Fp and FT, in Eqs. 6.10.10.4.2-1 and 6.10.10.4.2-5,
respectively, are not applicable when designing shear connectors for the routine steel I-girder
bridges covered by this Guide and may be taken as zero.
For simple span bridges, there is only one region to consider in the calculation of P and n
encompassing the entire span. Use Eq. 6.10.10.4.2-1 to calculate P.
For multi-span continuous bridges (rolled beam and plate girder bridges), there are two regions of
each span to consider in the calculation of P and n. For continuous spans with shear connectors
provided along the entire span length, which is recommended for the routine steel I-girder bridges
covered by this Guide, the regions are as follows:
• For end spans:
o Region between the point of maximum live load moment and the end support - use
Eq. 6.10.10.4.2-1 to calculate P.

NSBA Guide to Navigating Routine Steel Bridge Design / 260


o Region between point of maximum live load moment and the adjacent interior
support – use Eq. 6.10.10.4.2-5 to calculate P.
• For interior spans:
o Region between the point of maximum live load moment and the left interior
support - use Eq. 6.10.10.4.2-5 to calculate P.
o Region between point of maximum live load moment and the right interior support
– use Eq. 6.10.10.4.2-5 to calculate P.

For continuous spans with shear connectors omitted in regions of negative flexure, which depends
on the preferences of the Owner-agency but is not recommended, the regions are as follows:
• For end spans:
o Region between the point of maximum live load moment and the end support - use
Eq. 6.10.10.4.2-1 to calculate P.
o Region between the point of maximum live load moment and the adjacent point of
steel dead load contraflexure – use Eq. 6.10.10.4.2-1 to calculate P.
• For interior spans:
o Region between the point of maximum live load moment and the left point of steel
dead load contraflexure - use Eq. 6.10.10.4.2-1 to calculate P.
o Region between point of maximum live load moment and the right point of steel
dead load contraflexure – use Eq. 6.10.10.4.2-1 to calculate P.

For more explanation and examples of the determination of the design of shear connectors at the
fatigue and strength limit states, see Section 6.3.6.3 of Reference Manual for NHI Course 130122,
Design and Evaluation of Steel Bridges for Fatigue and Fracture, Section 6.6.2 of Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures as well as FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-
Span Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span
Continuous Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span
Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It performs
design calculations addressing the demand on, and resistance of, shear connectors at the fatigue and
strength limit states in accordance with the provisions of the AASHTO LRFD BDS, greatly reducing
the time and effort required of the designer. Other commercial software packages with the ability to
analyze and design routine steel I-girder bridges are also available. Users should verify the
capabilities, assumptions, and general correctness of any program’s calculations prior to initial
use.

6.10.10.4.3 Nominal Shear Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:

NSBA Guide to Navigating Routine Steel Bridge Design / 261


The provisions of this Article are used to determine the nominal shear resistance, Qn, of a single
shear connector at the strength limit state. Only the equation for stud shear connectors (Eq.
6.10.10.4.3-1) should be considered applicable to the routine I-girder bridges covered by this
Guide.
See the Discussion of Article 6.10.10.4.2 in this Guide for further explanation of how to determine
the demand on shear connectors at the strength limit state.

6.10.11 Web Stiffeners

6.10.11.1 Web Transverse Stiffeners

6.10.11.1.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides general design requirements for web transverse stiffeners. Web transverse
stiffeners on routine plate-girder bridges are typically plates welded to only one side of the web,
except for cross-frame or diaphragm connection plates on interior girders, which are typically
placed on both sides of the web. Intermediate web transverse stiffeners should be kept the same
size along the length of the girder; avoiding multiple plate sizes facilitates the use of repetitive
manufacturing techniques and reduces the possibility of placement errors. The minimum thickness
used for web stiffeners and connection plates should be ½ inch to facilitate welding (see discussion
of stiffener welding in Section 9.2.5 of the FHWA Bridge Welding Reference Manual). Section
C1.3 of the AASHTO-NSBA Steel Bridge Collaboration Guideline G12.1-2020 Guidelines to
Design for Constructability and Fabrication provides recommended dimensions for these members
that will allow fabricators to use either steel plate material or flat steel bar stock for stiffeners and
connection plates.
For the routine steel rolled-beam bridges covered by this Guide, web transverse stiffeners are
typically not needed, except for use as cross-frame or diaphragm connection plates which do not
serve as web transverse stiffeners for shear. A possible exception to the requirement in the 3rd
paragraph of this Article for connection plates on rolled-beam bridges is provided in Article
6.6.1.3.1 (see the Discussion of Article 6.6.1.3.1 in this Guide).
The provisions related to web transverse stiffeners on horizontally curved girders and
longitudinally stiffened web panels are not applicable to the routine steel I-girder bridges covered
by this Guide.
For further information on the design of web transverse stiffeners, consult Section 6.6.6.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructureshttps://www.fhwa.dot.gov/bridge/pubs/nhi15047.pdf. For design
examples, consult the FHWA’s Steel Bridge Design Handbook, Design Example 1, Three-Span
Continuous Straight Composite Steel I-Girder Bridge, Design Example 2A, Two-Span Continuous
Straight Composite Steel I-Girder Bridge, and Design Example 2B, Two-Span Continuous
Straight Composite Steel Wide-Flange Beam Bridge.

NSBA Guide to Navigating Routine Steel Bridge Design / 262


For helpful practical guidance on the design and detailing of web transverse stiffeners for
economical fabrication, consult AASHTO-NSBA Steel Bridge Collaboration Guidelines G12.1-
2020 Guidelines to Design for Constructability and Fabrication and G1.4-2006 Guidelines for
Design Details.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. It calculates
the design loads and resulting stresses, and the corresponding resistances, in accordance with the
provisions of the AASHTO LRFD BDS, greatly reducing the time and effort required of the
designer. Other commercial software packages with the ability to analyze and design routine steel I-
girder bridges are also available. Users should verify the capabilities, assumptions, and general
correctness of any program’s calculations prior to initial use.

6.10.11.1.2 Projecting Width


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides two projecting width requirements for web transverse stiffeners that set limits
for the stiffener width and thickness in terms of the web depth and flange width. bf in Eq.
6.10.11.1.2-2 is to be taken as the full width of the widest compression flange within the field
section under consideration; the requirement for tub girders in the computation of bf is not
applicable. For the routine rolled steel beam bridges covered by this Guide, the requirements of
this Article apply only to the design of the cross-frame or diaphragm connection plates which do
not serve as web transverse stiffeners for shear, because rolled beams typically do not require
transverse stiffeners to meet shear design requirements for their webs.
For further information on the design of web transverse stiffeners, consult Section 6.6.6.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
For helpful practical guidance on the design and detailing of web transverse stiffeners for
economical fabrication, consult AASHTO-NSBA Steel Bridge Collaboration Guidelines G12.1-
2020 Guidelines to Design for Constructability and Fabrication and G1.4-2006 Guidelines for
Design Details.

6.10.11.1.3 Moment of Inertia


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article contains requirements for the minimum required moment of inertia, It, of a web
transverse stiffener, where It is taken about the edge in contact with the web for single stiffeners
and about the mid-thickness of the web for stiffener pairs. This provides the web transverse

NSBA Guide to Navigating Routine Steel Bridge Design / 263


stiffener with sufficient rigidity to maintain a vertical line of near zero lateral deflection of the web
along the line of the stiffener so that the web can adequately develop the shear-buckling resistance
or the combined shear-buckling and post-buckling tension-field resistance, as applicable (see the
Discussion of Article 6.10.9 in this Guide).
There are two general conditions, each with a different set of equations, for determining It. The
first condition is for web transverse stiffeners adjacent to web panels in which neither panel is
subject to post-buckling tension-field action; e.g., unstiffened web panels adjacent to each other or
adjacent to an end panel (Eqs. 6.10.11.1.3-1 and 6.10.11.1.3-2). In this case, the moment of inertia
requirement is taken as the smaller of the two limits given by Eqs. 6.10.11.1.3-1 or 6.10.11.1.3-2.
This condition would typically apply for the design of cross-frame or diaphragm connection plates
on rolled-beam bridges, which do not serve as web transverse stiffeners for shear.
The second condition is for web transverse stiffeners adjacent to web panels subject to post-
buckling tension-field action; e.g., stiffened interior web panels (Eqs. 6.10.11.1.3-7 and
6.10.11.1.3-8). Note that the moment of inertia requirement given by Eq. 6.10.11.1.3-7 depends
on the ratio ρw. The value to use for ρw depends on whether both web panels adjacent to the
transverse stiffener are subject to post-buckling tension-field action, in which case the equation
given in the first bulleted item underneath Eq. 6.10.11.1.3-8 is used to compute ρw. Otherwise, the
equation given in the second bulleted item is used.
Eq. 6.10.11.1.3-11 provides a minimum moment of inertia requirement for transverse stiffeners
used in web panels with longitudinal stiffeners. This requirement is not applicable for the routine
steel I-girder bridges covered by this Guide, which do not contain web longitudinal stiffeners.
For further information on the design of web transverse stiffeners, consult Section 6.6.6.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
For helpful practical guidance on the design and detailing of web transverse stiffeners for
economical fabrication, consult AASHTO-NSBA Steel Bridge Collaboration Guidelines G12.1-
2020 Guidelines to Design for Constructability and Fabrication and G1.4-2006 Guidelines for
Design Details.

6.10.11.2 Bearing Stiffeners

6.10.11.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article provides general design requirements for bearing stiffeners. Plates welded to both
sides of the web are typically used for bearing stiffeners in the routine I-girder bridges covered by
this Guide. The stiffeners must extend the full depth of the web. In plate-girder bridges, bearing
stiffeners must be provided at support locations. For rolled-beam bridges, the provisions of Article

NSBA Guide to Navigating Routine Steel Bridge Design / 264


D6.5 related to the limit states of web local yielding and web crippling must be checked to
determine if bearing stiffeners are necessary at support locations or if they can be omitted (see the
Discussion of Article D6.5 in this Guide).
The provisions of Article D6.5 are also used to determine if bearing stiffeners are necessary at
other locations in plate-girder or rolled-beam bridges subjected to concentrated loads, where the
loads are not transmitted through a deck or deck system. A common example of this situation is a
jacking point on a girder or on a diaphragm at an end and/or interior support, where a jack may be
placed under the girder or diaphragm and used to lift the superstructure to facilitate bearing
replacement or other maintenance activities.
Bearing stiffeners serving as connection plates for cross-frames or diaphragms must be attached
to both flanges of the cross-section. The use of fillet welds to attach the stiffeners to the flanges is
recommended. The use of complete joint penetration groove welds to attach the stiffeners to the
flange through which it receives its load is permitted but is discouraged in order to significantly
reduce the welding deformation of the flange associated with large, complete joint penetration
welds. The method of attachment that is used is dependent on the preferences of the Owner-agency.
In particularly long bridges supported by bearings with sliding surfaces, it may be necessary to
consider providing so-called “auxiliary bearing stiffeners.” This is done in situations where the
longitudinal movement of the girder under thermal expansion or contraction is sufficiently large
enough that the bearing stiffener may not be located above the bearing at the limits of the design
temperature range; in these cases, additional bearing stiffeners are sometimes provided. Care
should be taken to detail the location and size of the primary and auxiliary bearing stiffeners such
that there is sufficient access for welding to attach the stiffeners to the girder flanges and web (see
the Discussion of Article 6.10.11.2.4b in this Guide). Consideration should also be given to the
specific finishing and welding details; the flange may undergo deformations when the first stiffener
is welded such that a “mill to bear” or “finish to bear” condition may be difficult to achieve in the
adjacent stiffener. Calling out complete joint penetration welds in conjunction with “mill to bear”
or “finish to bear” requirements is discouraged.
For further information on the design of bearing stiffeners, consult Section 6.6.6.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge, Design
Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
For helpful practical guidance on the design and detailing of bearing stiffeners for economical
fabrication, consult AASHTO-NSBA Steel Bridge Collaboration Guidelines G12.1-2020
Guidelines to Design for Constructability and Fabrication and G1.4-2006 Guidelines for Design
Details. Section 9.3.2 of the FHWA Bridge Welding Reference Manual also has good guidance
on detailing of bearing and jacking stiffeners.

6.10.11.2.2 Minimum Thickness


Determination of applicability, Simple Span Bridges: Conditionally applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 265


Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
This Article specifies a required minimum thickness, tp, for bearing stiffeners in terms of the
projecting stiffener width, bt, and the specified minimum yield strength of the stiffeners, Fys (Eq.
6.10.11.2.2-1). This requirement is intended to prevent local buckling of the stiffener plates. The
projecting width of the bearing stiffeners should extend as closely as practical to the outer edges
of the flanges.
This requirement is applicable for bearing stiffeners used in the routine plate-girder bridges
covered by this Guide and is only applicable for routine rolled-beam bridges if bearing stiffeners
are necessary at support locations or elsewhere.
For further information on the design of bearing stiffeners, consult Section 6.6.6.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge, Design
Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
For helpful practical guidance on the design and detailing of bearing stiffeners for economical
fabrication, consult AASHTO-NSBA Steel Bridge Collaboration Guidelines G12.1-2020
Guidelines to Design for Constructability and Fabrication and G1.4-2006 Guidelines for Design
Details.

6.10.11.2.3 Bearing Resistance


Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
This Article specifies a bearing resistance requirement for bearing stiffeners (Eq. 6.10.11.2.3-1).
The factored bearing resistance given by Eq. 6.10.11.2.3-1 must equal or exceed the factored
bearing reaction at the strength limit state. Each stiffener should be finished to bear against the
flange through which it receives its load. Note that the bearing stiffener area to be used in this
check is not the gross stiffener area. Rather, it is the bearing area, Apn, which excludes the portions
of the stiffeners that must be clipped to facilitate the web-to-flange fillet weld and any of the
stiffener area extending beyond the edges of the flange.

NSBA Guide to Navigating Routine Steel Bridge Design / 266


This requirement is applicable for bearing stiffeners used in the routine plate-girder bridges
covered by this Guide and is only applicable for routine rolled-beam bridges if bearing stiffeners
are necessary at support locations or elsewhere.
For further information on the design of bearing stiffeners, consult Section 6.6.6.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge, Design
Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
For helpful practical guidance on the design and detailing of bearing stiffeners for economical
fabrication, consult AASHTO-NSBA Steel Bridge Collaboration Guidelines G12.1-2020
Guidelines to Design for Constructability and Fabrication and G1.4-2006 Guidelines for Design
Details.

6.10.11.2.4 Axial Resistance of Bearing Stiffeners

6.10.11.2.4a General
Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
The provisions of this Article are used to determine the factored axial resistance of an effective
column section consisting of the bearing stiffeners and a portion of the web (see the Discussion of
Article 6.10.11.2.4b in this Guide), which may be included when welded stiffeners are used. The
factored axial resistance is determined according to the provisions of Articles 6.9.2.1 and 6.9.4.1.1
using the specified minimum yield strength of the stiffener plates, Fys (see the Discussion of
Articles 6.9.2.1 and 6.9.4.1.1 in this Guide). The effective length of the column is taken as 0.75
times the web depth, which assumes some level of fixity of each end of the stiffener plates. The
radius of gyration of the column is taken about the mid-thickness of the web. The factored axial
resistance of the effective column section must equal or exceed the factored bearing reaction at the
strength limit state.
This requirement is applicable for bearing stiffeners used in the routine plate-girder bridges
covered by this Guide and is only applicable for routine rolled-beam bridges if bearing stiffeners
are necessary at support locations or elsewhere.
For further information on the design of bearing stiffeners, consult Section 6.6.6.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge, Design

NSBA Guide to Navigating Routine Steel Bridge Design / 267


Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
For helpful practical guidance on the design and detailing of bearing stiffeners for economical
fabrication, consult AASHTO-NSBA Steel Bridge Collaboration Guidelines G12.1-2020
Guidelines to Design for Constructability and Fabrication and G1.4-2006 Guidelines for Design
Details.

6.10.11.2.4b Effective Section


Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Partially applicable.
Discussion:
The provisions of this Article are used to determine the effective column section of the bearing
stiffeners and a portion of the web for computing the factored axial resistance in Article
6.10.11.2.4a (see the Discussion of Article 6.10.11.2.4a in this Guide). For stiffeners consisting of
two plates welded to the web, the effective column section is to consist of the two stiffener
elements, plus a centrally located strip of web extending not more than 9tw on each side of the
stiffeners. If more than one pair of stiffeners is used, the effective column section is to consist of
the stiffener elements, plus a centrally located strip of web extending not more than 9tw on each
side of the outer projecting elements of the group. If more than one pair of stiffeners is used, it is
recommended that a minimum spacing of 8.0 inches of 1.5 times the stiffener width be provided
between the stiffeners for welding access. The maximum spacing between the stiffeners should
not exceed 1.09tw E Fys to prevent an effective width reduction on the web of the section between
the stiffeners, where tw is the thickness of the web.
These provisions are applicable for bearing stiffeners used in the routine plate-girder bridges
covered by this Guide and is only applicable for routine rolled-beam bridges if bearing stiffeners
are necessary at support locations or elsewhere. The provisions specified for stiffeners bolted to
the web and for bearing stiffeners at interior supports on continuous-span hybrid members do not
apply to the routine steel I-girder bridges covered by this Guide.
For further information on the design of bearing stiffeners, consult Section 6.6.6.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. For design examples, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 1, Three-Span Continuous Straight Composite Steel I-Girder Bridge, Design
Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
For helpful practical guidance on the design and detailing of bearing stiffeners for economical
fabrication, consult AASHTO-NSBA Steel Bridge Collaboration Guidelines G12.1-2020
Guidelines to Design for Constructability and Fabrication and G1.4-2006 Guidelines for Design
Details.

NSBA Guide to Navigating Routine Steel Bridge Design / 268


6.10.11.3 Web Longitudinal Stiffeners

6.10.11.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article cover the general provisions related to the design of web longitudinal
stiffeners. The routine steel I-girder bridges covered by this Guide do not have web longitudinal
stiffeners and therefore these provisions are not applicable.

6.10.11.3.2 Projecting Width


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article cover the specific design provisions related to the projecting width
of web longitudinal stiffeners. The routine steel I-girder bridges covered by this Guide do not have
web longitudinal stiffeners and therefore these provisions are not applicable.

6.10.11.3.3 Moment of Inertia and Radius of Gyration


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article cover the specific design provisions related to the moment of inertia
and radius of gyration of web longitudinal stiffeners. The routine steel I-girder bridges covered by
this Guide do not have web longitudinal stiffeners and therefore these provisions are not
applicable.

6.10.12 Cover Plates

6.10.12.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article cover the general provisions related to the design of cover plates for
I-section flexural members. The routine steel I-girder bridges covered by this Guide do not have
cover plates and therefore these provisions are not applicable.

6.10.12.2 End Requirements

6.10.12.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:

NSBA Guide to Navigating Routine Steel Bridge Design / 269


The provisions of this Article cover the general design provisions related to the end requirements
for cover plates for I-section flexural members. The routine steel I-girder bridges covered by this
Guide do not have cover plates and therefore these provisions are not applicable.

6.10.12.2.2 Welded Ends


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article cover the design provisions specific to welded ends of cover plates
for I-section flexural members. The routine steel I-girder bridges covered by this Guide do not
have cover plates and therefore these provisions are not applicable.

6.10.12.2.3 Bolted Ends


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article cover the design provisions specific to bolted ends of cover plates
for I-section flexural members. The routine steel I-girder bridges covered by this Guide do not
have cover plates and therefore these provisions are not applicable.

6.11 COMPOSITE BOX-SECTION FLEXURAL MEMBERS

6.11.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

Discussion:
This Article specifies general design requirements for composite steel box-section flexural
members used in straight or horizontally curved bridges. The provisions are applicable to both
composite closed-box and tub-section members. Tub sections have an open top with two separate
top flanges laced together with lateral bracing to form a pseudo-box to resist the torsion prior to
the hardening of the deck. Tub sections are by far the most commonly used cross-section type for
composite box-section flexural members and typically have inclined webs to allow for the use of
a narrower and more economical bottom flange plate while enjoying the advantage of a wider
spacing of the webs supporting the deck. Closed-box sections enclosed at the top with a steel plate
that is composite with the concrete deck are rarely, if ever, used for these members since OSHA
regulations make it very expensive and impractical to work inside a closed box. Box-girder cross-
sections can consist of multiple single-cell steel boxes (most common), one single-cell steel box,
or a single multi-cell steel box. The latter type is rarely employed and is not covered in the
AASHTO LRFD BDS.

NSBA Guide to Navigating Routine Steel Bridge Design / 270


The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.
For further information on the design of composite steel box-section flexural members, consult the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures. For design examples, consult the FHWA’s Steel Bridge Design
Handbook, Design Example 4, Three-Span Continuous Straight Composite Steel Tub-Girder
Bridge and Design Example 5, Three-Span Continuous Horizontally Curved Composite Steel Tub-
Girder Bridge.

6.11.1.1 Stress Determinations


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This article contains provisions related to the effective width of a box flange in a composite box-
section member subject to flexure (see the Discussion of Article 6.11.1 in this Guide) to account
for the effects of shear lag, where a box flange is defined in the AASHTO LRFD BDS as a flange
that is connected to two webs. This article also contains provisions related to the stress
determinations in a composite box-section flexural member, including the determination of the
live-load distribution to the individual boxes in the cross-section, the section of an exterior girder
assumed to resist the factored wind loading, the types of box sections for which St. Venant
torsional shear stresses and transverse bending and longitudinal warping stresses due to cross-
section distortion must be considered, and a specified limit on the factored torsional shear stress
in such boxes at the strength limit state.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.1.2 Bearings
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains provisions related to the use of a single or double bearing arrangement to
support a composite box-section flexural member (see the Discussion of Article 6.11.1 in this
Guide). The potential use of tie-down bearings is also discussed.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.1.3 Flange-to-Web Connections


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains provisions related to the size and placement of welded flange-to-web
connections in a composite box-section flexural member (see the Discussion of Article 6.11.1 in

NSBA Guide to Navigating Routine Steel Bridge Design / 271


this Guide), and the minimum number of intermediate internal cross-frames or diaphragms that
must be provided within each span in order to use fillet welds for these connections.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.1.4 Access and Drainage


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article covers the provision of access holes in the bottom flange of composite box-section
flexural members (see the Discussion of Article 6.11.1 in this Guide) for inspection. Provisions
are specified for the placement of the holes, the need for reinforcement of the holes, and the
checking of local buckling of the remaining flange on each side of the hole at access holes in
bottom flanges subject to compression. Ventilation and drainage of the interior of the box section
is also discussed.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.2 Cross-Section Proportion Limits

6.11.2.1 Web Proportions

6.11.2.1.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies general design requirements for webs of composite box-section flexural
members (see the Discussion of Article 6.11.1 in this Guide), including the preferred slope of
inclined webs.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.2.1.2 Webs without Longitudinal Stiffeners


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the limiting web slenderness for webs of composite box-section flexural
members (see the Discussion of Article 6.11.1 in this Guide) without longitudinal stiffeners
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 272


6.11.2.1.3 Webs with Longitudinal Stiffeners
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the limiting web slenderness for webs of composite box-section flexural
members (see the Discussion of Article 6.11.1 in this Guide) with longitudinal stiffeners
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.2.2 Flange Proportions


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the flange proportioning requirements for top flanges of composite tub-
section flexural members (see the Discussion of Article 6.11.1 in this Guide).
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.2.3 Special Restrictions on Use of Live Load Distribution Factor for Multiple Box
Sections
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies restrictions for straight bridges utilizing multiple composite box-section
flexural members (see the Discussion of Article 6.11.1 in this Guide) that must be met in order to
employ the lateral live-load distribution factor given in Article 4.6.2.2.2b for straight multiple steel
box sections (see the Discussion of Article 4.6.2.2.2b in this Guide). Otherwise, a refined analysis
must be used to determine the live-load distribution. Furthermore, for bridges satisfying these
restrictions and with an effective box-flange width not exceeding one-fifth of the effective span
defined in Article 6.11.1.1 (see the Discussion of Article 6.11.1.1 in this Guide), shear due to St.
Venant torsion and secondary distortional bending stress effects may be neglected.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.3 Constructibility

6.11.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:

NSBA Guide to Navigating Routine Steel Bridge Design / 273


This Article refers to Article 6.10.3 for the general provisions related to the constructibility design
of composite steel box-section flexural members (see the Discussion of Article 6.10.3 in this
Guide). The provisions also require that the need for bracing to maintain individual box-section
geometry throughout all stages of construction be investigated.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.3.2 Flexure
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article refers to the provisions of Articles 6.10.3.2.1 through 6.10.3.2.3 for the checking of
the top flanges of composite steel tub-section flexural members for constructibility (see the
Discussion of Article 6.10.3.2.1 in this Guide) and defines the unbraced lengths for such flanges.
This Article also indicates the provisions of Article A6.3.3 may not be used to determine the
lateral-torsional buckling resistance of tub-section members with compact or noncompact webs,
as the provisions of Appendix A6 (see the Discussion of Appendix A6 in this Guide) do not apply
to tub-section members (see the Discussion of Article 6.10.6.2.3 in this Guide for the definitions
of compact and noncompact webs).
Provisions are also provided to check noncomposite box flanges and continuously braced box
flanges subject to compression or tension during construction and also to check composite box
flanges before the concrete deck has hardened or is made composite, where a box flange is defined
in the AASHTO LRFD BDS as a flange that is connected to two webs. The resistance equations
for box flanges in this Article include the consideration of the St. Venant torsional shear stress in
the flange due to the torque applied to the noncomposite section for the specific cases in which the
torsional shear must be considered (see the Discussion of Articles 6.11.1.1 and 6.11.2.3 in this
Guide). Flange lateral bending is not a consideration for box flanges. St. Venant torsional shears
are typically neglected in the top flanges of tub sections.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.3.3 Shear
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article indicates that when checking the shear requirement specified in Article 6.10.3.3 for
composite steel box-section flexural members during construction (see the Discussion of Article
6.10.3.3 in this Guide), the provisions of Article 6.11.9 also apply (see the Discussion of Article
6.11.9 in this Guide).
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 274


6.11.4 Service Limit State
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article refers to Article 6.10.4 for the checking of the service limit state design provisions for
composite steel box-section flexural members (see the Discussion of Article 6.10.4 in this Guide),
with certain exceptions as listed. This Article also indicates the optional service limit state moment
redistribution procedures given in Appendix B6 are not to be applied to composite box-section
flexural members (see the Discussion of Appendix B6 in this Guide) because the applicability of
these provisions to box sections has not been demonstrated. The routine steel I-girder bridges
covered by this Guide are not comprised of composite steel box-section flexural members;
therefore, the provisions of this Article are not applicable.

6.11.5 Fatigue and Fracture Limit State


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article refers to Article 6.10.5 for the checking of the fatigue limit state design provisions for
composite steel box-section flexural members (see the Discussion of Article 6.10.5 in this Guide).
Specific requirements related to the checking of fatigue of shear connectors and the special fatigue
requirement for webs given in Article 6.10.5.3 (see the Discussion of Article 6.10.5.3 in this Guide)
for such members are also provided. Situations for which longitudinal warping stresses and
transverse bending stresses due to cross-section distortion may be of concern for fatigue are also
discussed along with suggested approaches to calculate and control these stresses where necessary.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.6 Strength Limit State

6.11.6.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article refers to the applicable Strength load combinations given in Table 3.4.1-1, which are
utilized in the design checks at the strength limit state for composite steel box-section flexural
members.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 275


6.11.6.2 Flexure

6.11.6.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article refers to the provisions Article 6.10.1.8 if there are holes in the tension flange of a
composite steel box-section flexural member; e.g., at a bolted splice (see the Discussion of Article
6.10.1.8 in this Guide).
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.6.2.2 Sections in Positive Flexure


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article defines the requirements for a composite box section in regions of positive flexure to
qualify as a compact or a noncompact section at the strength limit state, and where the provisions
to design each type of section are located in Article 6.11.7 (see the Discussion of Article 6.11.7 in
this Guide). This Article also refers to the ductility requirement given in Article 6.10.7.3 to provide
a ductile mode of failure, which must be checked for both compact and noncompact sections (see
the Discussion of Article 6.10.7.3 in this Guide).

The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.6.2.3 Sections in Negative Flexure


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article indicates that the provisions of Article 6.11.8 are to apply for a composite box section
in regions of negative flexure at the strength limit state (see the Discussion of Article 6.11.8 in this
Guide). This Article also indicates the optional strength limit state moment redistribution
procedures given in Appendix B6 are not to be applied to composite box-section flexural members
(see the Discussion of Appendix B6 in this Guide) because the applicability of these provisions to
box sections has not been demonstrated.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.6.3 Shear
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 276


Discussion:
This Article simply points to the provisions of Article 6.11.9 for determining the factored shear
resistance of a composite box-section flexural member at the strength limit state (see the
Discussion of Article 6.11.9 in this Guide).
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.6.4 Shear Connectors


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article simply points to the provisions of Article 6.10.10.4 for determining the factored shear
resistance of shear connectors for composite box-section flexural members at the strength limit
state (see the Discussion of Article 6.10.10.4 in this Guide). These provisions further refer to the
provisions of Article 6.11.10, as applicable (see the Discussion of Article 6.11.10 in this Guide).
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.7 Flexural Resistance—Sections in Positive Flexure

6.11.7.1 Compact Sections

6.11.7.1.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides the relationship that must be satisfied at the strength limit state for compact
composite box sections in regions of positive flexure. Most composite sections in regions of
positive flexure in straight steel box girder bridges without holes in the tension flange will qualify
as compact sections. Sections that qualify as compact sections may conservatively be treated as
noncompact sections (see the Discussion of Article 6.11.7.2 in this Guide), if desired.
For compact sections, the nominal flexural resistance is permitted to exceed the moment at first
yield assuming there are no holes in the tension flange at the section under consideration. The
moment at first yield, My, is defined as the moment at which an outer fiber first attains the yield
stress (see the Discussion of Article D6.2.2 in this Guide). The nominal flexural resistance is not
permitted to exceed the plastic moment, Mp. Mp is defined as the resisting moment of a fully yielded
cross-section (see the Discussion of Article D6.1 in this Guide). For compact sections, the nominal
flexural resistance is expressed in terms of moment for reasons discussed in the Commentary for
Article 6.10.6.1.
Flange lateral bending is not a consideration for top flanges because the flanges are continuously
braced by the concrete deck. Flange lateral bending is also not a consideration for bottom box

NSBA Guide to Navigating Routine Steel Bridge Design / 277


flanges, where a box flange is defined in the AASHTO LRFD BDS as a flange that is connected
to two webs.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.7.1.2 Nominal Flexural Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article refers to the provisions of Article 6.10.7.1.2 for determining the nominal flexural
resistance, Mn, of compact composite box sections in regions of positive flexure at the strength
limit state (see the Discussion of Article 6.10.7.1.2 in this Guide). The single exception is that for
continuous spans, Mn must always be subject to the limitation of 1.3RhMy given in Eq. 6.10.7.1-2-
3, as the provisions of Appendix B6 described in the two bulleted items in Article 6.10.7.1.2 are
not applicable to box sections. The reasons for this limitation are described further in the
Discussion of Article 6.10.7.1.2 in this Guide.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.7.2 Noncompact Sections

6.11.7.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides the relationships that must be satisfied at the strength limit state for the
compression and tension flanges of noncompact composite box sections in regions of positive
flexure. Composite sections in regions of positive flexure in horizontally curved steel box girder
bridges must be treated as noncompact sections. For noncompact sections (and for compact
sections with holes in the tension flange), the nominal flexural resistance is not to exceed the
moment at first yield. The moment at first yield, My, is defined as the moment at which an outer
fiber first attains the yield stress (see the Discussion of Article D6.2.2 in this Guide). For
noncompact sections, the nominal flexural resistance is expressed in terms of the elastically
computed flange stress for reasons discussed in the Commentary for Article 6.10.6.1.
This Article further limits the maximum longitudinal compressive stress in the concrete deck at
the strength limit state for a noncompact box section to 0.6f'c to maintain linear behavior of the
concrete, which is assumed in the calculation of the steel flange stresses.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 278


6.11.7.2.2 Nominal Flexural Resistance
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides the equations for computing the nominal flexural resistance of the
compression flange, Fnc, and the nominal flexural resistance of the tension flange, Fnt, for
noncompact composite box sections in regions of positive flexure at the strength limit state. For
noncompact sections, the elastically computed factored stress in each flange, determined in
accordance with Article 6.10.1.1.1a, is compared with the yield stress of the flange times the
appropriate flange-stress reduction factors, Rb and Rh (see the Discussion of Articles 6.10.1.1.1a,
6.10.1.10.1 and 6.10.1.10.2 in this Guide). Separate resistance equations are provided for the top
flanges of tub sections and closed-box sections subject to compression and for the bottom flanges
of tub sections and closed-box sections subject to tension. The resistance equations for box flanges
in this Article include the consideration of the St. Venant torsional shear stress in the flange for
the specific cases in which the torsional shear must be considered (see the Discussion of Articles
6.11.1.1 and 6.11.2.3 in this Guide), where a box flange is defined in the AASHTO LRFD BDS
as a flange that is connected to two webs. The Commentary for this Article discusses the
computation of the flange torsional shear stress.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.8 Flexural Resistance—Sections in Negative Flexure

6.11.8.1 General

6.11.8.1.1 Box Flanges in Compression


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides the relationship that must be satisfied at the strength limit state for box
flanges subject to compression in regions of negative flexure, where a box flange is defined in the
AASHTO LRFD BDS as a flange that is connected to two webs. This relationship is intended to
check that box flanges subject to compression in composite box-section flexural members will
have sufficient strength to resist flange local buckling. Flange lateral bending and lateral-torsional
buckling are not a consideration for box flanges. Equations are also provided in the Commentary
for this Article to check the complex combined stress state in bottom box flanges at interior-pier
sections in cases where the internal diaphragm shear stresses and/or bending of the internal
diaphragm over the bearing sole plate are deemed significant (e.g., boxes supported on single
bearings).
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 279


6.11.8.1.2 Continuously Braced Flanges in Tension
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides the relationship that must be satisfied at the strength limit state for
continuously braced flanges subjected to tension in regions of negative flexure in composite box-
section flexural members. A continuously braced flange is defined as a flange that is encased in
concrete or anchored to the concrete deck by shear connectors satisfying the provisions of Article
6.10.10. For continuously braced top flanges of tub sections, lateral flange bending stresses and
St. Venant torsional shear stresses are neglected. The torsional shear stresses are not to be
neglected in continuously braced box flanges, where a box flange is defined in the AASHTO
LRFD BDS as a flange that is connected to two webs.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.8.2 Flexural Resistance of Box Flanges in Compression

6.11.8.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article point to the appropriate articles for determining the nominal flexural
resistance, Fnc, of box flanges subject to compression both with (see the Discussion of Article
6.11.8.2.2 in this Guide) and without (see the Discussion of Article 6.11.8.2.3 in this Guide) flange
longitudinal stiffeners in composite box-section flexural members at the strength limit state, where
a box flange is defined in the AASHTO LRFD BDS as a flange that is connected to two webs.
Flange lateral bending is not a consideration for box flanges.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.8.2.2 Unstiffened Flanges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides the equations for calculating the nominal local buckling resistance, Fnc, of
unstiffened box flanges subject to compression (i.e., flanges without longitudinal stiffeners) in
composite box-section flexural members at the strength limit state, where a box flange is defined
in the AASHTO LRFD BDS as a flange that is connected to two webs. The resistance equations
include the consideration of the St. Venant torsional shear stress in the flange for the specific cases
in which the torsional shear must be considered (see the Discussion of Articles 6.11.1.1 and
6.11.2.3 in this Guide). The Commentary for this Article discusses the computation of the flange

NSBA Guide to Navigating Routine Steel Bridge Design / 280


torsional shear stress and the calculation of the flange-stress reduction factors, Rb and Rh (see also
the Discussion of Articles 6.10.1.10.1 and 6.10.1.10.2 in this Guide).
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.8.2.3 Longitudinally Stiffened Flanges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides the equations for calculating the nominal local buckling resistance, Fnc, of
longitudinally stiffened box flanges subject to compression in composite box-section flexural
members at the strength limit state, where a box flange is defined in the AASHTO LRFD BDS as
a flange that is connected to two webs. The equations for unstiffened box flanges in compression
specified in Article 6.11.8.2.2 are used (see the Discussion of Article 6.11.8.2.2 in this Guide),
with the appropriate substitutions made to account for the presence of one or two flange
longitudinal stiffeners as listed in this Article. As discussed further in Article C6.11.11.2 (see the
Discussion of Article 6.11.11.2 in this Guide), as the number of stiffeners is increased beyond one,
the required moment of inertia from Eq. 6.11.11.2-2 to achieve the desired k value (i.e., the plate
buckling coefficient for uniform stress typically ranging from a value of 2.0 to 4.0) begins to
increase dramatically and eventually becomes nearly impractical. Therefore, for boxes of typical
proportions, it is strongly recommended that the number of longitudinal flange stiffeners not exceed
one for maximum economy.

The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.8.3 Flexural Resistance Based on Tension Flange Yielding


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to compute the nominal flexural resistance, Fnt, based on
tension flange yielding of continuously braced flanges subjected to tension in regions of negative
flexure in composite tub-section flexural members at the strength limit state. For continuously
braced flanges subjected to tension in regions of negative flexure in composite closed-box section
flexural members, Eq. 6.11.7.2.2-5 is used instead. A continuously braced flange is defined as a
flange that is encased in concrete or anchored to the concrete deck by shear connectors satisfying
the provisions of Article 6.10.10.
For sections in which Myt > Myc, Eq. 6.10.8.3-1 or 6.11.7.2.2-5, as applicable, does not control and
tension flange yielding need not be checked, where Myc and Myt are the yield moments with respect
to the compression and tension flange, respectively, determined as specified in Article D6.2 (see
the Discussion of Article D6.2 in this Guide).

NSBA Guide to Navigating Routine Steel Bridge Design / 281


The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.9 Shear Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article point to the provisions of Article 6.10.9 for the determination of the
factored shear resistance of composite box-section flexural members (see the Discussion of Article
6.10.9 in this Guide), with the exceptions noted below.
For sections with inclined webs, the depth, D, is to be taken as the depth of the web plate measured
along the slope and the web is to be designed for the component of the vertical shear in the plane
of the inclined web given by Eq. 6.11.9-1. For the specific cases in which the torsional shear must
be considered (see the Discussion of Articles 6.11.1.1 and 6.11.2.3 in this Guide), the factored
vertical shear, Vu, is to be taken as the sum of the flexural and St. Venant torsional shears. Both
webs can be conservatively designed for the critical shear or the shear in the web subject to additive
flexural and torsional shear. This Article also specifies the width of box flanges to be used in
checking Eq. 6.10.9.3.2-1 (see the Discussion of Article 6.10.9.3.2 in this Guide), where a box
flange is defined in the AASHTO LRFD BDS as a flange that is connected to two webs.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.10 Shear Connectors


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article point to the provisions of Article 6.10.10 for the design of shear
connectors for composite box-section flexural members (see the Discussion of Article 6.10.10 in
this Guide), with the exceptions noted below. In multi-span continuous box girder bridges, shear
connectors must be provided in regions of negative flexure to resist the torsional shear that exists
along the entire span in all types of composite box sections.
For the specific cases in which the torsional shear must be considered (see the Discussion of
Articles 6.11.1.1 and 6.11.2.3 in this Guide), shear connectors are to be designed for the sum of
the flexural and St. Venant torsional shears. For tub sections, the longitudinal fatigue shear range,
Vfat, is to be computed for the web subject to additive flexural and torsional shear with the resulting
pitch also used for the other top flange. The radial fatigue shear range due to curvature, Ffat1, is to
be ignored for box-section members (see the Discussion of Article 6.10.10.1.2 in this Guide).
Provisions for the design of shear connectors at the strength limit state and for the design of shear
connectors on composite box flanges in regions of positive flexure in closed-box sections are also
provided, where a box flange is defined in the AASHTO LRFD BDS as a flange that is connected
to two webs.

NSBA Guide to Navigating Routine Steel Bridge Design / 282


The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.11 Stiffeners

6.11.11.1 Web Stiffeners


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article point to the provisions of Articles 6.10.11.1 and 6.10.11.3 for the
design of web transverse stiffeners and web longitudinal stiffeners, respectively, for composite
box-section flexural members (see the Discussion of Articles 6.10.11.1 and 6.10.11.3 in this
Guide). For the design of bearing stiffeners for such members, the provisions point to the
provisions of Article 6.10.11.2 (see the Discussion of Article 6.10.11.2 in this Guide), with the
exceptions noted which relate to the fact that bearing stiffeners are typically attached to the internal
diaphragms rather than to the inclined webs at supports in these members.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

6.11.11.2 Longitudinal Compression-Flange Stiffeners


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article specify projecting width and moment of inertia requirements for
longitudinal stiffeners on box flanges subject to compression in composite box-section flexural
members, where a box flange is defined in the AASHTO LRFD BDS as a flange that is connected
to two webs. The projecting width requirement is intended to prevent local bucking of the
projecting stiffener elements. The minimum required moment of inertia for each stiffener provides
the assumed value of the plate buckling coefficient for uniform stress, k, (typically assumed to be
between 2.0 and 4.0) that is used in the computation of the nominal flexural resistance, Fnc, of the
stiffened flange (see the Discussion of Article 6.10.8.2.3 in this Guide). The actual moment of inertia
of the stiffener, I, is taken about an axis parallel to the flange and at the base of the stiffener.
The number of flange longitudinal stiffeners preferably should not exceed one for maximum
economy in boxes of typical proportions. When the number of stiffeners exceeds one, the required
moment of inertia from Eq. 6.11.11.2-2 begins to increase dramatically. For rare cases where an
exceptionally wide box flange is required and the number of stiffeners may need to exceed 2, it is
suggested that transverse flange stiffeners be considered to reduce the required size of the
longitudinal flange stiffeners to a more practical value. The Commentary for this Article provides
provisions for the design of transverse flange stiffeners in such cases.
The routine steel I-girder bridges covered by this Guide are not comprised of composite steel box-
section flexural members; therefore, the provisions of this Article are not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 283


6.12 MISCELLANEOUS FLEXURAL MEMBERS

6.12.1 General

6.12.1.1 Scope
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article cover miscellaneous rolled or built-up noncomposite or composite
members subject to flexure, often in combination with axial loads; that is, flexural members which
are not covered by the provisions of Article 6.10 or 6.11. The nominal flexural resistance of these
members is often needed for application in the interaction relationships of Articles 6.8.2.3.1,
6.9.2.2.1, and 6.9.6.3.4, as applicable (see the Discussion of Articles 6.8.2.3.1, 6.9.2.2.1, and
6.9.6.3.4 in this Guide). The specific types of members covered by these provisions are listed in
this Article. These types of members are often used in trusses, frames, or arches, or as cross-frame,
diaphragm, or lateral bracing members.
Most of the member types discussed in Article 6.12 and its associated sub-Articles are not widely
used, or are not used at all, in the routine steel I-girder bridges covered by this Guide. Some are
used, but only in limited applications, particularly as cross-frame or diaphragm members.
Tees (WT) and double angles are sometimes be used as cross-frame members, generally only when
the axial loads in these members exceed the capacity of single-angle sections. It should be noted
that using rolled steel tee (WT) and double-angle sections as cross-frame members for routine steel
I-girder bridges is generally discouraged, while the use of single-angle sections is encouraged. The
magnitude of the forces in cross-frame members in routine steel I-girder bridges covered by this
Guide is generally small enough such that single-angle members are adequate. Cross-frame forces
are typically only large enough to warrant the use of rolled steel tee (WT) or double-angle sections
in curved and/or skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite
expensive to fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally
require straightening after the cutting process, which adds significant fabrication effort and cost.
Double-angle sections are often viewed as problematic from a maintenance perspective; the
surfaces between the adjacent angle flanges are difficult or impossible to paint in the field, and/or
can suffer from potentially severe pack rust.
Channels may be used as a top chord in an end diaphragm, in which case they need to be designed
as a flexural member to support the wheel loads coming onto the end of the deck, or as a diaphragm
for a shallow-depth rolled beam structure.
Other member types mentioned in Article 6.12 and its associated sub-Articles could potentially be
used in substructures for routine steel I-girder bridges; in those cases, the determination of
applicability of the provisions related to those particular members is designated as beyond the
scope of superstructure design.

NSBA Guide to Navigating Routine Steel Bridge Design / 284


6.12.1.2 Strength Limit State

6.12.1.2.1 Flexure
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article provide the basic relationship that must be satisfied at the strength
limit state by the miscellaneous flexural members listed in Article 6.12.1.1 (see the Discussion of
Article 6.12.1.1 in this Guide) assuming low or zero levels of axial force in the member and
uniaxial flexure. This Article also defines the factored flexural resistance, Mr, of these members to
be used in the preceding relationship.
As discussed in the Commentary for Article 6.12.1.2.2, for members subject to flexure in
combination with a factored concentrically-applied axial force, Pu, in excess of 5 percent of the
factored axial resistance of the member, Pr or Pry, as applicable (defined in Articles 6.9.2.2.1 and
6.8.2.3.1, respectively) at the strength limit state, and/or if the member is subject to biaxial bending,
the member should instead be checked using the interaction relationships specified in Article
6.8.2.3 or 6.9.2.2, as applicable (see the Discussion of Articles 6.8.2.3 and 6.9.2.2 in this Guide).
For the routine steel I-girder bridges covered by this Guide, the interaction relationships specified
in Article 6.9.2.2.1 are likely to be applicable when tees (WT) or double angles are used as cross-
frame members and the members are subject to eccentric axial compression. It should be noted
that using rolled steel tee (WT) and double-angle sections as cross-frame members for routine steel
I-girder bridges is generally discouraged, while the use of single-angle sections is encouraged. The
magnitude of the forces in cross-frame members in routine steel I-girder bridges covered by this
Guide is generally small enough such that single-angle members are adequate. Cross-frame forces
are typically only large enough to warrant the use of rolled steel tee (WT) or double-angle sections
in curved and/or skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite
expensive to fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally
require straightening after the cutting process, which adds significant fabrication effort and cost.
Double-angle sections are often viewed as problematic from a maintenance perspective; the
surfaces between the adjacent angle flanges are difficult or impossible to paint in the field, and/or
can suffer from potentially severe pack rust.
These provisions may also be applicable to the design of channel sections when they are used as
the top chord of end cross-frames or as diaphragms in bridges with shallow-depth beams or girders
Some of the other member types could potentially be used in the substructure of a routine I-girder
bridge and would likely be subject to the interaction relationships, but the determination of
applicability of the provisions related to those particular members is designated as beyond the
scope of superstructure design.

6.12.1.2.2 Combined Flexure, Axial Load, and Flexural and/or Torsional Shear
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 285


Discussion:
This Article points to the provisions of Article 6.8.2.3 for combined axial tension, flexure, and
flexural and/or torsional shear (see the Discussion of Article 6.8.2.3 in this Guide, or the provisions
of Article 6.9.2.2 for combined axial compression, flexure, and flexural and/or torsional shear (see
the Discussion of Article 6.9.2.2 in this Guide, when designing a member listed in Article 6.12.1.1
(see the Discussion of Article 6.12.1.1 in this Guide) for the corresponding combined loading
conditions.
See the Discussion of Article 6.12.1.2.1 in this Guide regarding the basic conditions involving
axial force and/or flexure for which the interaction relationships given in these articles should be
used. For the routine steel I-girder bridges covered by this Guide, the interaction relationships
specified in Article 6.9.2.2.1 are likely to be applicable when tees (WT) or double angles are used
as cross-frame members and the members are subject to eccentric axial compression. It should be
noted that using rolled steel tee (WT) and double-angle sections as cross-frame members for
routine steel I-girder bridges is generally discouraged, while the use of single-angle sections is
encouraged. The magnitude of the forces in cross-frame members in routine steel I-girder bridges
covered by this Guide is generally small enough such that single-angle members are adequate.
Cross-frame forces are typically only large enough to warrant the use of rolled steel tee (WT) or
double-angle sections in curved and/or skewed steel I-girder bridges. Rolled steel tee (WT)
sections are typically quite expensive to fabricate. Tee (WT) sections are cut from full wide-flange
(W) shapes and generally require straightening after the cutting process, which adds significant
fabrication effort and cost. Double-angle sections are often viewed as problematic from a
maintenance perspective; the surfaces between the adjacent angle flanges are difficult or
impossible to paint in the field, and/or can suffer from potentially severe pack rust.
The provisions of this Article may also be applicable to the design of channel sections when they
are used as the top chord of end cross-frames or as diaphragms in bridges with shallow-depth
beams or girders. Designers are reminded that loads applied to channel sections in a direction
parallel to the plane of the web, but offset from the shear center of the section, will induce a torsion
in the section. The shear center of a singly-symmetric channel section (such as a typical AISC C
or MC channel shape) is generally located at mid-depth of the section, but offset from the web in
a direction opposite to the direction in which the flanges are pointed; in other words, the locations
of the center of gravity and the shear center are not coincident in a channel shape.
Some of the other member types could potentially be used in the substructure of a routine I-girder
bridge and would likely be subject to the interaction relationships, but the provisions related to
those particular member types are designated as beyond the scope and not applicable to
superstructure design. The last paragraph of this Article related to noncomposite circular tubes,
including round HSS (Hollow Structural Sections), under combined loading conditions is not
applicable to the routine steel I-girder bridges covered by this Guide as these members are not used
in these bridges, unless they are used in the substructure or foundations for the bridge which is
considered beyond the scope of superstructure design.

NSBA Guide to Navigating Routine Steel Bridge Design / 286


6.12.1.2.3 Flexural Shear and/or Torsion

6.12.1.2.3a General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article are used to calculate and check the factored shear resistance, Vr, at
the strength limit state of the miscellaneous flexural members listed in Article 6.12.1.1 (see the
Discussion of Article 6.12.1.1 in this Guide). The provisions are also used to calculate and check
the factored torsional resistance, Tr, at the strength limit state of noncomposite circular tubes,
including round HSS (Hollow Structural Sections) subject to torsion only or subject to combined
flexural shear and torsion. The basic relationships that must be satisfied for each case are provided,
along with the definitions of Vr and Tr to be used in these relationships. The Article also points to
the appropriate provisions for the calculation of the nominal shear resistance, Vn, and nominal
torsional resistance, Tn, as applicable, for each type of member.
For the routine steel I-girder bridges covered by this Guide, the provisions of this Article are only
applicable to check the factored shear resistance of a channel if a channel is used as a top chord in
an end diaphragm and it is designed as a flexural member to support the wheel loads coming onto
the end of the deck, or as a diaphragm for a shallow-depth rolled beam structure. The provisions
of Article 6.10.9 would be used to compute Vn in this case (see the Discussion of Article 6.10.9 in
this Guide). Designers are reminded that loads applied to channel sections in a direction parallel
to the plane of the web, but offset from the shear center of the section, will induce a torsion in the
section. The shear center of a singly-symmetric channel section (such as a typical AISC C or MC
channel shape) is generally located at mid-depth of the section, but offset from the web in a
direction opposite to the direction in which the flanges are pointed; in other words, the locations
of the center of gravity and the shear center are not coincident in a channel shape.
Some of the other member types could potentially be used in the substructure of a routine I-girder
bridge and the provisions of this Article would likely be applicable in such cases, but the provisions
related to those particular member types are designated as beyond the scope of superstructure
design. Otherwise, the provisions of this Article are not applicable.

6.12.1.2.3b Circular Tubes and Round HSS


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to calculate the nominal flexural shear resistance, Vn, and
the nominal torsional resistance, Tn, at the strength limit state of circular tubes, including round
HSS (Hollow Structural Sections). Vn and Tn are used in the basic relationships provided in Article
6.12.1.2.3a to check the factored shear and torsional resistance of these members at the strength
limit state (see the Discussion of Article 6.12.1.2.3a in this Guide).
The provisions of this Article are not applicable to the routine steel I-girder bridges covered by
this Guide as circular tubes and round HSS are not used in these bridges, including as cross-frame

NSBA Guide to Navigating Routine Steel Bridge Design / 287


members. These member types could potentially be used in the substructure of a routine I-girder
bridge and the provisions of this Article would likely be applicable in such cases, but the provisions
related to those particular member types are designated as beyond the scope and not applicable to
superstructure design.
These members are not used in routine I-girder bridges, including as cross-frame or diaphragm
members.

6.12.1.2.4 Special Provisions for HSS Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to define the web depth, D, and inside flange width, bfi, for
the design of square and rectangular HSS (Hollow Structural Section) members, and the design
wall thickness, t, for the design of square, rectangular, and round HSS members. This Article is
referred to in Articles 6.9.4.2, 6.12.1.2.3, 6.12.2.2.2, and 6.12.2.2.3.
The provisions of this Article are not applicable to the routine steel I-girder bridges covered by
this Guide as HSS members are not used in these bridges, including not being used as cross-frame
members.

6.12.2 Nominal Flexural Resistance

6.12.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article simply states that provisions for lateral-torsional buckling need not be considered
when determining the nominal flexural resistance of the following miscellaneous member types
covered in Article 6.12: composite members (i.e., concrete-encased shapes and circular concrete-
filled steel tubes); noncomposite I- and H-shaped members bent about their weak axis (i.e., y-axis);
and noncomposite circular tubes.
The provisions of this Article are not applicable to the routine steel I-girder bridges covered by
this Guide as these miscellaneous member types are not used in these bridges. These member types
could potentially be used in the substructure of a routine I-girder bridge and the provisions of this
Article would likely be applicable in such cases, but the provisions related to those particular
member types are designated as beyond the scope of superstructure design.

6.12.2.2 Noncomposite Members

6.12.2.2.1 I- and H-Shaped Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 288


Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of noncomposite I- and
H-shaped members subject to flexure about their weak axis (i.e., their y-axis). These provisions
are also applicable to channels (see the Discussion of Article 6.12.2.2.5 in this Guide) and members
consisting of two channel flanges connected by a web plate subject to flexure about their weak
axis. For cases where I- and H-shaped members are subject to flexure about their strong axis (the
axis perpendicular to the web), this Article states that the provisions of Article 6.10 shall apply.
The routine steel I-girder bridges covered by this Guide may utilize I-shaped members or channels
as a top chord in an end diaphragm or as a diaphragm for a shallow-depth rolled beam structure,
but these members are not typically subjected to loading producing flexure about their weak axis
(nor should they be). As such, the provisions of this Article do not apply to their design.
These provisions may be applicable to the design of I- or H-shaped piles that could potentially be
used in the substructure for a routine steel I-girder bridge, but the design of piles is designated as
beyond the scope of superstructure design. Therefore, the provisions of this Article are not
applicable.

6.12.2.2.2 Rectangular Box-Section Members

6.12.2.2.2a General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article points to subsequent articles for the provisions related to the flexural design of
homogeneous and hybrid doubly and singly symmetric single-cell rectangular noncomposite box-
section members with or without longitudinal stiffeners bent about either principal axis in which
the cross-section principal axes are parallel to the cross-section component plates. The provisions
also apply for the flexural design of square and rectangular HSS (Hollow Structural Sections).
The routine steel I-girder bridges covered by this Guide are not comprised of noncomposite
rectangular box-section flexural members, including HSS members; therefore, the provisions of
this Article are not applicable.

6.12.2.2.2b Cross-Section Proportion Limits


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article specify cross-section proportion limits for webs and compression
flanges with and without longitudinal stiffeners, tension flanges, the outside width of the section,
and flange extensions on compression flanges for noncomposite rectangular box-section members
subject to flexure.

NSBA Guide to Navigating Routine Steel Bridge Design / 289


The routine steel I-girder bridges covered by this Guide are not comprised of noncomposite
rectangular box-section flexural members; therefore, the provisions of this Article are not
applicable.

6.12.2.2.2c Classification of Sections with a Longitudinally Unstiffened Compression Flange


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article provide the necessary classification of noncomposite rectangular
box-section members with a longitudinally unstiffened compression flange subject to flexure based
on the web and compression-flange slenderness, and define several important cross-section based
parameters for each classification employed in the calculation of the member flexural resistance
in Article 6.12.2.2.2e (see the Discussion of Article 6.12.2.2.2e in this Guide).
The routine steel I-girder bridges covered by this Guide are not comprised of noncomposite
rectangular box-section flexural members; therefore, the provisions of this Article are not
applicable.

6.12.2.2.2d Classification of Sections with a Longitudinally Stiffened Compression Flange


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article provide the necessary classification of noncomposite rectangular
box-section members with a longitudinally stiffened compression flange subject to flexure, define
an effective area, Aeff, representing the longitudinally stiffened compression flange used to
compute the yield moment of the effective section with respect to the compression flange and other
effective section properties, and define several important cross-section based parameters for each
classification employed in the calculation of the member flexural resistance in Article 6.12.2.2.2e
(see the Discussion of Article 6.12.2.2.2e in this Guide).
The routine steel I-girder bridges covered by this Guide are not comprised of noncomposite
rectangular box-section flexural members; therefore, the provisions of this Article are not
applicable.

6.12.2.2.2e General Yielding, Compression Flange Local Buckling and Lateral Torsional
Buckling
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of noncomposite
rectangular box-section members considering the combined effects of general yielding,
compression-flange local buckling, and lateral-torsional buckling as a function of the unbraced
length, the effective section properties, and the cross-section based parameters for each cross-

NSBA Guide to Navigating Routine Steel Bridge Design / 290


section classification determined in Article 6.12.2.2.2c or 6.12.2.2.2d, as applicable (see the
Discussion of Articles 6.12.2.2.2c and 6.12.2.2.2d in this Guide).
The routine steel I-girder bridges covered by this Guide are not comprised of noncomposite
rectangular box-section flexural members; therefore, the provisions of this Article are not
applicable.

6.12.2.2.2f Service and Fatigue Limit States and Constructibility


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains provisions that must be satisfied at the service and fatigue limit states and for
constructibility for noncomposite rectangular box-section members subject to flexure. Shear in the
webs is limited to the shear-yield or shear-bucking resistance, Vcr, under the factored load for
constructibility specified in Article 3.4.2.1 and also under the unfactored permanent load plus the
Fatigue I load combination to alleviate any significant flexing of the web due to shearing action.
Post-buckling resistance is assumed at the strength limit state in computing the nominal flexural
resistance of noncomposite box-section members with slender webs and/or noncompact or slender
flanges or plate panels. Under the provisions of Article 6.9.4.5, these particular elements are
investigated to check that plate local buckling does not occur under conditions producing
maximum longitudinal compressive stress acting at one or both longitudinal edges of the element
under consideration at the service limit state or during construction (see the Discussion of Article
6.9.4.5 in this Guide). The flange flexural stresses in the box-section member are also limited under
the Service II load combination to help control permanent deformations of the member.

The routine steel I-girder bridges covered by this Guide are not comprised of noncomposite
rectangular box-section flexural members; therefore, the provisions of this Article are not
applicable.

6.12.2.2.2g Flange Effective Width or Area Accounting for Shear Lag Effects
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article provide simplified rules regarding the consideration of shear lag
effects, where required, in noncomposite rectangular box-section members subject to flexure in
lieu of a more refined analysis. Reductions to the effective compression-flange area and gross
tension flange area are specified to account for shear lag effects in the computation of the flexural
resistance of the member at the strength limit state and in the computation of the elastic flexural
stresses at the service and fatigue limit states and for constructibility. The specified reductions are
not intended to be applied within the bridge structural analysis.
The routine steel I-girder bridges covered by this Guide are not comprised of noncomposite
rectangular box-section flexural members; therefore, the provisions of this Article are not
applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 291


6.12.2.2.3 Circular Tubes and Round HSS
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of noncomposite
circular tubes, including round HSS (Hollow Structural Sections). The flexural resistance is taken
as the smaller resistance based on yielding and local buckling, as applicable. A maximum D/t ratio
is also specified for circular tubes used as flexural members, where D is the outside diameter of
the tube and t is the thickness of the tube.
The routine steel I-girder bridges covered by this Guide do not utilize noncomposite circular tubes
or round HSS; therefore, the provisions of this Article are not applicable.

6.12.2.2.4 Tees and Double Angles

6.12.2.2.4a General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of tees (WT) and double
angles loaded in the plane of symmetry formed by their y-axis. The flexural resistance is taken as
the smaller resistance based on yielding, lateral-torsional buckling, flange local buckling, and local
buckling of tee stems and double-angle web legs determined in subsequent articles, as applicable
(see the Discussion of Articles 6.12.2.2.4b through 6.12.2.2.4e in this Guide). Legs of double
angles in continuous contact or with separators may together be assumed as double-angle web legs
in applying these provisions. Flexure of these members about the y-axis is not a consideration for
the routine steel I-girder bridges covered by this Guide.
The computation of Mn for tees and double angles is typically required for application in the
appropriate interaction equations of Article 6.9.2.2.1 when these members are subject to eccentric
axial compression (see the Discussion of Article 6.9.2.2.1 in this Guide).
For the routine steel I-girder bridges covered by this Guide, these provisions are applicable if a tee
or double angle is used as a cross-frame member. It should be noted that using rolled steel tee
(WT) and double-angle sections as cross-frame members for routine steel I-girder bridges is
generally discouraged, while the use of single-angle sections is encouraged. The magnitude of the
forces in cross-frame members in routine steel I-girder bridges covered by this Guide is generally
small enough such that single-angle members are adequate. Cross-frame forces are typically only
large enough to warrant the use of rolled steel tee (WT) or double-angle sections in curved and/or
skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite expensive to
fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally require
straightening after the cutting process, which adds significant fabrication effort and cost. Double-
angle sections are often viewed as problematic from a maintenance perspective; the surfaces
between the adjacent angle flanges are difficult or impossible to paint in the field, and/or can suffer
from potentially severe pack rust.

NSBA Guide to Navigating Routine Steel Bridge Design / 292


6.12.2.2.4b Yielding
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of tee (WT) and double-
angle members based on yielding for cases where the tee stems or double-angle web legs are
subject to tension or compression. Upper limits are specified on Mn in each case to indirectly
control situations where significant yielding of the stem or web legs may occur at service load
levels.
For the routine steel I-girder bridges covered by this Guide, these provisions are applicable if a tee
or double angle is used as a cross-frame member. It should be noted that using rolled steel tee
(WT) and double-angle sections as cross-frame members for routine steel I-girder bridges is
generally discouraged, while the use of single-angle sections is encouraged. The magnitude of the
forces in cross-frame members in routine steel I-girder bridges covered by this Guide is generally
small enough such that single-angle members are adequate. Cross-frame forces are typically only
large enough to warrant the use of rolled steel tee (WT) or double-angle sections in curved and/or
skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite expensive to
fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally require
straightening after the cutting process, which adds significant fabrication effort and cost. Double-
angle sections are often viewed as problematic from a maintenance perspective; the surfaces
between the adjacent angle flanges are difficult or impossible to paint in the field, and/or can suffer
from potentially severe pack rust.

6.12.2.2.4c Lateral-Torsional Buckling


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of tee (WT) and double-
angle members based on lateral-torsional buckling for cases where the tee stems or double-angle
web legs are subject to tension or compression. For all cases, the moment-gradient modifier, Cb,
is not included in the equations for reasons discussed in the Commentary for this Article. The
lateral-torsional buckling capacity is significantly greater for the case where the tee stem or double-
angle web legs are in tension than for the case where the stem or web legs are in compression.
Additionally, detailing of end connections for cases where the tee stem or double-angle web legs
are in tension should be done in a manner to minimize the potential for fixed-end moments that
induce compression in the stem or web legs.
For the routine steel I-girder bridges covered by this Guide, these provisions are applicable if a tee
or double angle is used as a cross-frame member. It should be noted that using rolled steel tee
(WT) and double-angle sections as cross-frame members for routine steel I-girder bridges is
generally discouraged, while the use of single-angle sections is encouraged. The magnitude of the
forces in cross-frame members in routine steel I-girder bridges covered by this Guide is generally
small enough such that single-angle members are adequate. Cross-frame forces are typically only
large enough to warrant the use of rolled steel tee (WT) or double-angle sections in curved and/or

NSBA Guide to Navigating Routine Steel Bridge Design / 293


skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite expensive to
fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally require
straightening after the cutting process, which adds significant fabrication effort and cost. Double-
angle sections are often viewed as problematic from a maintenance perspective; the surfaces
between the adjacent angle flanges are difficult or impossible to paint in the field, and/or can suffer
from potentially severe pack rust.

6.12.2.2.4d Flange Local Buckling


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of tee (WT) and double-
angle members based on flange local buckling for cases where the tee flanges or double-angle
flange legs are subject to compression. The width of double-angle flange legs is to be taken as the
sum of the widths of the outstanding legs, not including any gap in-between the angles, in applying
these provisions. The elastic flange local buckling resistance equation for tee sections, i.e., Eq.
6.12.2.2.4d-2, need only be considered for fabricated tee sections as the flanges of rolled tee
sections given in the AISC Manual of Steel Construction qualify as nonslender (see the Discussion
of Article 6.9.4.2.1 in this Guide).
For the routine steel I-girder bridges covered by this Guide, these provisions are applicable if a tee
or double angle is used as a cross-frame member. It should be noted that using rolled steel tee
(WT) and double-angle sections as cross-frame members for routine steel I-girder bridges is
generally discouraged, while the use of single-angle sections is encouraged. The magnitude of the
forces in cross-frame members in routine steel I-girder bridges covered by this Guide is generally
small enough such that single-angle members are adequate. Cross-frame forces are typically only
large enough to warrant the use of rolled steel tee (WT) or double-angle sections in curved and/or
skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite expensive to
fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally require
straightening after the cutting process, which adds significant fabrication effort and cost. Double-
angle sections are often viewed as problematic from a maintenance perspective; the surfaces
between the adjacent angle flanges are difficult or impossible to paint in the field, and/or can suffer
from potentially severe pack rust.

6.12.2.2.4e Local Buckling of Tee Stems and Double Angle Web Legs
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of tee (WT) and double-
angle members based on local buckling of the tee stems or double-angle web legs for cases where
the stem or web legs are subject to compression.
For the routine steel I-girder bridges covered by this Guide, these provisions are applicable if a tee
or double angle is used as a cross-frame member. It should be noted that using rolled steel tee
(WT) and double-angle sections as cross-frame members for routine steel I-girder bridges is

NSBA Guide to Navigating Routine Steel Bridge Design / 294


generally discouraged, while the use of single-angle sections is encouraged. The magnitude of the
forces in cross-frame members in routine steel I-girder bridges covered by this Guide is generally
small enough such that single-angle members are adequate. Cross-frame forces are typically only
large enough to warrant the use of rolled steel tee (WT) or double-angle sections in curved and/or
skewed steel I-girder bridges. Rolled steel tee (WT) sections are typically quite expensive to
fabricate. Tee (WT) sections are cut from full wide-flange (W) shapes and generally require
straightening after the cutting process, which adds significant fabrication effort and cost. Double-
angle sections are often viewed as problematic from a maintenance perspective; the surfaces
between the adjacent angle flanges are difficult or impossible to paint in the field, and/or can suffer
from potentially severe pack rust.

6.12.2.2.5 Channels
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of channels. The
flexural resistance about the strong axis (i.e, the x-axis) is taken as the smaller resistance based on
yielding or lateral-torsional buckling, as applicable. For channels in flexure about the weak axis
(i.e., the y-axis), the provisions of Article 6.12.2.2.1 are to be applied with Mn limited to 1.6FySy
to indirectly prevent substantial yielding of the member at service load levels, where Sy is the
elastic section modulus about the y-axis (see the Discussion of Article 6.12.2.2.1 in this Guide).
The equations specified for the lateral-torsional buckling resistance in this Article assume the
channel is sufficiently braced at support locations to prevent twisting of the section at those points.
This Article also specifies flange and web slenderness limits for fabricated or bent-plate channels
such that flange and web local buckling need not be checked. Rolled channels given in the AISC
Manual of Steel Construction have compact flanges and webs for yield strengths not exceeding 65
ksi; therefore, these limits need not be checked for rolled channels, but can be checked for
completeness.
For the routine steel I-girder bridges covered by this Guide, these provisions are applicable when
a channel is used as a top chord in an end diaphragm and it is designed as a flexural member to
support the wheel loads coming onto the end of the deck, or when a channel is used as a diaphragm
for a shallow-depth rolled beam structure.

6.12.2.2.6 Single Angles


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Single angles are not used as pure flexural members in the routine steel I-girder bridges covered
by this Guide. The effect of the moments due to eccentricities on a single angle member may be
ignored by designing the angle for eccentric axial compression using the effective slenderness ratio
defined in Article 6.9.4.4 and by designing the angle for eccentric axial tension using the
appropriate shear lag reduction factor, U, defined in Article 6.8.2.2, as applicable (see the

NSBA Guide to Navigating Routine Steel Bridge Design / 295


Discussion of Articles 6.9.4.4 and 6.8.2.2 in this Guide). Therefore, the provisions of this Article
are not applicable to the routine steel I-girder bridges covered by this Guide.

6.12.2.2.7 Rectangular Bars and Solid Rounds


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of rectangular bars and
solid rounds. The flexural resistance is taken as the smaller resistance based on yielding or lateral-
torsional buckling, as applicable.
The routine steel I-girder bridges covered by this Guide do not utilize rectangular bars or solid
rounds; therefore, the provisions of this Article are not applicable.

6.12.2.3 Composite Members

6.12.2.3.1 Concrete-Encased Shapes


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of concrete-encased
steel shapes subject to flexure only or to flexure in combination with axial compression. The
calculation of the nominal compressive resistance, Pn, of these members is covered in Article
6.9.5.1 (see the Discussion of Article 6.9.5.1 in this Guide). The members must satisfy the
applicable limitations specified in Article 6.9.5.2 (see the Discussion of Article 6.9.5.2 in this
Guide).
These members could potentially be used in substructures for the routine steel I-girder bridges
covered by this Guide, although they are not commonly utilized. Therefore, these provisions are
considered beyond the scope and not applicable to superstructure design.

6.12.2.3.2 Concrete-Filled Tubes


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article quantify the nominal flexural resistance, Mn, of circular concrete-
filled tubes or pipes where full composite action is not deemed necessary under combined axial
compression and flexure. The calculation of the nominal compressive resistance, Pn, of these
members is covered in Article 6.9.5.1 (see the Discussion of Article 6.9.5.1 in this Guide). For
applications where full composite action is deemed necessary, the provisions of Article 6.12.2.3.3
should be employed instead (see the Discussion of Article 6.12.2.3.3 in this Guide). The members

NSBA Guide to Navigating Routine Steel Bridge Design / 296


must satisfy the applicable limitations specified in Article 6.9.5.2 (see the Discussion of Article
6.9.5.2 in this Guide).
These members could potentially be used in substructures for the routine steel I-girder bridges
covered by this Guide, although they are not commonly utilized. Therefore, these provisions are
considered beyond the scope and not applicable to superstructure design.

6.12.2.3.3 Composite Concrete-Filled Steel Tubes (CFSTs)


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article are used to develop the nominal flexural composite resistance, Mn,
of composite circular concrete-filled steel tubes, or CFSTs, subject to flexure in combination with
axial compression as a function of the nominal axial resistance, Pn. These larger-diameter
members, which may be used for piers, columns, piles and drilled shafts, must satisfy the
limitations specified in Article 6.9.6.2 (see the Discussion of Article 6.9.6.2 in this Guide). CFSTs
should not be used as pure flexural members.
The nominal material-based nominal axial force/moment interaction diagram for these members
is to be developed using either the Plastic Stress Distribution Method (PDSM) or the Strain
Compatibility Method (SCM). The application of the PDSM to these members, which is a cross-
section analysis using the constituent materials based on equilibrium at full plastification of the
section, is described in detail in the Commentary for this Article. The resulting material-based
interaction curve is modified as specified in Article 6.9.6.3.4 to include stability effects based on
the buckling load determined in Article 6.9.6.3.2 to create a nominal stability-based interaction
curve, which is then multiplied by the appropriate resistance factor specified in Article 6.5.4.2 to
determine the final factored resistance of the CFST for combined axial compression and flexure
for all load conditions (see the Discussion of Articles 6.9.6.3.2 and 6.9.6.3.4 in this Guide).
These members could potentially be used in substructures for the routine steel I-girder bridges
covered by this Guide, although they are not commonly utilized. Therefore, these provisions are
considered beyond the scope and not applicable to superstructure design.

6.12.3 Nominal Shear Resistance of Composite Members

6.12.3.1 Concrete-Encased Shapes


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article quantify the nominal shear resistance, Vn, of concrete-encased steel
shapes.

NSBA Guide to Navigating Routine Steel Bridge Design / 297


These members could potentially be used in substructures for the routine steel I-girder bridges
covered by this Guide, although they are not commonly utilized. Therefore, these provisions are
considered beyond the scope and not applicable to superstructure design.

6.12.3.2 Concrete-Filled Tubes

6.12.3.2.1 Rectangular Tubes


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article quantify the nominal shear resistance, Vn, of rectangular concrete-
filled tubes.
These members could potentially be used in substructures for the routine steel I-girder bridges
covered by this Guide, although they are not commonly utilized. Therefore, these provisions are
considered beyond the scope and not applicable to superstructure design.

6.12.3.2.2 Composite Concrete Filled Tubes


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of this Article quantify the nominal shear resistance, Vn, of circular concrete-filled
steel tubes, including composite CFSTs both with and without longitudinal reinforcement.
These members could potentially be used in substructures for the routine steel I-girder bridges
covered by this Guide, although they are not commonly utilized. Therefore, these provisions are
considered beyond the scope and not applicable to superstructure design.

6.13 CONNECTIONS AND SPLICES

6.13.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
The provisions of this Article cover several general considerations related to connection design.
Connections should be made symmetrical about the axis of the members, where practical.
Members, including bracing, should be connected so that their gravity axes intersect at a point.
Eccentric connections should be avoided, however, where this is not possible, the members and
connections must be designed for the combined effects of the shear and moment due to the
eccentricity.
Bolted connections, except for connections on lacing and handrails, are to contain not less than
two bolts. In the case of connections that transfer total member end shear, the gross section is to

NSBA Guide to Navigating Routine Steel Bridge Design / 298


be taken as the gross section of the connected elements. The calculation of individual bolt shear
loads in cases where there is an in-plane eccentricity between the axis of loading and the centroid
of the bolt group should consider the effects of both the direct shear force and the eccentric
moment; methods for evaluating eccentrically-loaded bolted connections can be found in most
steel design textbooks. When there is an out-of-plane eccentricity between the axis of loading and
the plane of the bolted connection, the need to consider the effects of the eccentric moment should
be considered on a case-by-case basis (see the Discussions of Articles 6.13.2.8 and 6.13.2.11 in
this Guide).
The configuration of some welded end connections, such as the connection of a cross-frame
member to a gusset plate, may result in an unbalanced weld condition, where the centroid of the
weld group is offset from the centroid of the member, producing an eccentricity in the plane of the
connection. In the case of single angles, the member itself is unsymmetrical and an eccentricity is
created unless the weld group is balanced to align with the member centroid. To create a balanced
condition for the welded connection of a single-angle member, the weld along the edge of the
angle nearest the centroid of the angle needs to be longer than the weld along the opposite edge of
the connected flange of the angle. Determining the appropriate weld geometry for a balanced
connection involves a relatively simple exercise in statics. Descriptions of this procedure can be
found in many steel design textbooks.
When the centroid of the member is not aligned with the centroid of the weld group, the weld
group should be designed for the additional moment generated by the eccentricity, which causes
additional shear loading of the welds. There are two methods commonly recognized for analyzing
eccentrically loaded welded connections: elastic (vector) analysis and the instantaneous center of
rotation (strength) method. A brief explanation of each method is provided below. More detailed
discussion of these methods can be found in many steel design textbooks.
Elastic analysis of an eccentrically loaded weld group assumes that each segment of weld resists a
concentrically applied load with an equal force, and rotation is assumed to occur about the centroid
of the weld configuration. The design moment (sometimes called the “torsional moment”) is
calculated by multiplying the axial force in the member by the eccentricity between the centroid
of the member and the centroid of the welded connection. The load on each weld segment caused
by the torsional moment is assumed to be proportional to the distance from the centroid. The
direction of the force caused by torsion is assumed to be perpendicular to the radial distance from
the centroid of the weld configuration. The components of the forces caused by direct load and
torsion are combined vectorially to obtain a resultant force. This method is relatively simple to
apply, but will provide conservative results, and can be excessively conservative for some
conditions.
The instantaneous center of rotation method considers the combined effect of rotation and
translation of one connection element with respect to the other. This combined effect is equivalent
to a rotation about a point defined as the instantaneous center of rotation (IC). The location of the
IC depends upon the geometry of the weld group as well as the direction and the point of
application of the load. The individual resistance of each weld segment is assumed to be
perpendicular to the radial distance from the IC. If the IC has been correctly located, the sum of
the applied forces and the weld resistances about the IC will be zero. Locating the IC can be

NSBA Guide to Navigating Routine Steel Bridge Design / 299


computationally cumbersome, and application of this method generally requires the use of
tabulated values or an iterative solution. However, for connections where torsional effects are
significant, this method provides more accurate and less conservative results than elastic analysis.
The effect of unbalanced welds on the strength and fatigue resistance of single-angle members can
typically be neglected. The Commentary to Article J1.7 of AISC’s Specifications for Structural
Steel Buildings and Commentary states that tests have shown that the effect of slight eccentricities
between the centroid of a welded connection and the centroid of the connected member have
negligible effect on the static strength of the member. This same Commentary states, “However,
the fatigue life of eccentrically loaded welded angles has been shown to be very short. Notches at
the roots of fillet welds are harmful when alternating tensile stresses are normal to the axis of the
weld, as could occur due to bending when axial cyclic loading is applied to angles with end welds
not balanced about the neutral axis. Accordingly, balanced welds are required when such members
are subjected to cyclic loading (see Figure C-J1.3).” However, this statement warrants some
context. Research has shown that single-angle members connected to gusset plates with balanced
welds generally exhibited slightly better fatigue performance than those with unbalanced, equal
length welds, but all of the tested specimens, which consisted of angles with both balanced and
unbalanced welds, “… failed within the E or E′ categories delineated by the AASHTO
specification… using the stress range calculated on the gross section.” The current AASHTO
LRFD BDS treats the welded connection of angle and tee members to gusset plates as Category E′
details, and thus already captures the effect of an unbalanced welded connection on the fatigue
resistance of the angle member.
Ideally, welded connections should be detailed to be balanced. For singly-symmetric members
(such as tee sections connected through the flange) or doubly-symmetric members (rarely used for
cross-frame members, but including I-shaped sections, HSS sections, etc.), a balanced condition
is easily achieved by specifying equal-length longitudinal welds. For unsymmetrical members
such as single-angle members, a balanced weld condition requires unequal length longitudinal
welds. It can sometimes be challenging to achieve a balanced configuration for the welded
connections of such members given the specific geometry of a particular cross-frame. In such
cases, designers are encouraged to be practical. Detailing excessively oversized or odd-shaped
gusset plates solely to achieve a balanced weld configuration is discouraged. Instead, evaluate the
weld stresses in the unbalanced weld using one of the methods described above.
Also note that historically it has been a common practice in a number of jurisdictions to specify
only a total length of weld required for the connection of cross-frame members, and allow the steel
detailer to determine the specific weld geometry based on the orientation of the member and the
gusset plate geometry. This practice is discouraged, as it may result in weld geometry which does
not provide sufficient resistance for the anticipated design loads if an unbalanced condition, not
anticipated by the designer, is detailed for the welds.
The effects of out-of-plane eccentricities on welded end connections should also be considered. In
some cases, the effect of the out-of-plane eccentric moment on the weld stresses may be
significant, in which case the total vector sum weld stress should be compared to the resistance of
the weld. In other cases, similar to the Discussion of Article 6.13.2.1.1 for bolted connections, the
out-of-plane eccentric moment may have only a negligible effect on the weld stresses in welded

NSBA Guide to Navigating Routine Steel Bridge Design / 300


end connections if the length of the welds perpendicular to the eccentricity is large compared to
the eccentric distance.
End connection angles used to connect stringers to floorbeams and/or floorbeams to girders should
be connected with high-strength bolts. Where bolting is not practical, welded connections may be
used, but they must be designed for the vertical loads and any bending moment resulting from
restraint against end rotation. End connections for the stringers, floorbeams and girders should be
made with two angles; doing so eliminates the need to account for the eccentricity of the stringer,
floorbeam, or girder end reaction in the design of the perpendicular bolted connection. The bolted
connection should however be designed for combined shear and tension (see the Discussion of
Article 6.13.2.11 in this Guide). The thickness of end connection angles of stringers, floorbeams
and girders is not to be less than 3/8 in. These provisions are applicable only to the design of girder-
stringer-floorbeam systems and are not applicable to the routine steel I-girder bridges covered by
this Guide.
The provisions in this Article related to the application of the 75 percent of the factored resistance
or the average of the calculated factored axial force effect and the factored axial resistance apply
solely to primary members subject only to axial tension or compression. In routine steel I-girder
bridges, the only primary members are the I-shaped main spanning elements, which are considered
“flexural members”. There are separate provisions (mentioned later in the Discussion of this
Article) addressing connections and splices of primary flexural members. The other typical
members in a routine steel I-girder bridge, such as the cross-frames or diaphragms, the bearing
stiffeners, etc., are considered secondary members. As a result, the provisions related to the
application of the 75 percent of the factored resistance or the average of the calculated factored
axial force effect and the factored axial resistance are not applicable to the design of routine steel
I-girder bridges. The provisions of this Article related to connections and splices for primary
members subjected to combined force effects are similarly not applicable to the routine steel I-
girder bridges covered by this Guide.
Meanwhile, the provisions of this Article refer to the appropriate Articles for the design of bolted
splices (Article 6.13.6.1.3) and welded splices (Article 6.13.6.2) for flexural members at the
strength limit state, which are applicable to the routine steel I-girder bridges covered by this Guide
(see the Discussion of Articles 6.13.6.1.3 and 6.13.6.2 in this Guide).
The provisions of this Article also specify that where cross-frames/diaphragms, lateral bracing,
stringers or floorbeams for straight or horizontally curved members are included in a structural
model used to determine force effects, or are designed for explicitly calculated force effects from
the results of a separate investigation (e.g., an approximate wind load analysis), the end
connections for those members are to be designed for the calculated factored member force effects.
Otherwise, the end connections for these members are to be designed for 75 percent of the factored
resistance corresponding to the force effect under consideration. Refined methods of analysis are
not required, nor are they recommended, for the routine steel I-girder bridges covered by this
Guide. However, it is recommended that reasonable cross-frame and diaphragm force effects be
calculated using a “separate investigation.” For the routine steel I-girder bridges which are the
subject of this Guide (i.e., bridges which are straight, with little or no skew and limited span

NSBA Guide to Navigating Routine Steel Bridge Design / 301


lengths), a reasonable “separate investigation” should include calculation of the following force
effects:
• The member loads in the cross-frames or diaphragms resulting from the transfer of wind
loads applied to the girders up to the deck
• The minimum strength requirements associated with the cross-frames or diaphragms
functioning as stability bracing for the girders
As has been mentioned in the Discussion of Articles 6.7.4.1, 6.7.4.2, 6.8.1, and 6.8.2.1 in this
Guide, although not currently specified in the AASHTO LRFD BDS, it is recommended herein
that in addition to the minimum design requirements specified in Article 6.7.4.1 (see the
Discussion of Article 6.7.4.1 in this Guide), cross-frames or diaphragms for the routine steel I-
girder bridges covered by this Guide also be designed to satisfy the stability bracing strength and
stiffness requirements specified in AISC Appendix 6 - Article 6.3.2a. Consult Section 6.6.3.2 of
the Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures, the FHWA’s Steel Bridge Design Handbook, Volume 13 –
Bracing System Design, and Coletti, D.A., and M.A. Grubb, “Practical Implementation of Stability
Bracing Strength and Stiffness Guidelines for Steel I-Girder Bridges,” Proceedings of the 2016
World Steel Bridge Symposium, April 14, 2016, Orlando, FL for further information on these
requirements and their application to cross-frames and diaphragms in steel I-girder bridges.
The provisions in this Article related to the use of standard-size bolt holes in connections in
horizontally curved bridges and timber stringers framing into steel floorbeams are not applicable
to the routine steel I-girder bridges covered by this Guide.

6.13.2 Bolted Connections

6.13.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article contains some general considerations related to the design of bolted connections.
Bolted steel parts must fit solidly together after the bolts are tightened. The bolted parts may be
coated or uncoated. It must be specified in the contract documents that all joint surfaces, including
surfaces adjacent to the bolt head and nut, be free of scale (except for tight mill scale), dirt or other
foreign material. All material within the grip of the bolt must be steel.
High-strength bolts are to be installed to have a specified initial tension, which results in an initial
precompression between the joined parts. At service load levels, the transfer of the loads between
the joined parts may then occur entirely via friction with no bearing of the bolt shank against the
side of the hole. Until the friction force is overcome, the shear resistance of the bolt and the bearing
resistance of the bolt hole will not affect the ability to transfer the load across the shear plane
between the joined parts.
The AASHTO LRFD BDS recognizes two types of high-strength bolted connections; slip-critical
connections (see the Discussion of Article 6.13.2.1.1 in this Guide) and bearing-type connections

NSBA Guide to Navigating Routine Steel Bridge Design / 302


(see the Discussion of Article 6.13.2.1.2 in this Guide). The resistance (or “strength”) of high-
strength bolted connections in transmitting shear across a shear plane between bolted steel parts is
the same whether the connection is a slip-critical or bearing-type connection. The slip-critical
connection has an additional requirement that sufficient frictional resistance be provided so that
slip will not occur between the joined parts at service load levels.
For further information on high-strength bolts, including installation provisions and verification
procedures, consult the AISC Design Guide 17 High Strength Bolts - A Primer for Engineers, the
RCSC Specifications for Structural Joints Using High-Strength Bolts available from the Research
Council on Structural Connections (RCSC), the AASHTO-NSBA Steel Bridge Collaboration
Guide Specification S2.1-2018 Steel Bridge Fabrication Guide Specification, the AASHTO LRFD
Bridge Construction Specifications, and the applicable Owner-agency standards.

6.13.2.1.1 Slip-Critical Connections


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article deal with the design of high-strength bolted slip-critical connections.
When these connections are subject to shear, the load is transferred between the joined parts by
friction up to a level of force that is dependent upon the clamping force and the coefficient of
friction of the faying surfaces. The coefficient of friction depends on the faying surface condition,
with mill scale, paint or other surface treatments determining the value of the friction coefficient.
Prior to joint slip, the bolts are not subject to shear nor are the joined parts subject to bearing stress.
Once the load exceeds the frictional resistance between the faying surfaces, slip occurs; that is, the
friction bond is broken and the two surfaces slip with respect to one another by a relatively large
amount. Generally, a rupture failure does not occur when the shear loading in the connection
exceeds the slip resistance; instead, the connection is able to continue resisting an even greater
load through the shear resistance of the bolts and the bearing resistance against the connected
material. Under this scenario, final failure of the connection is by shear failure of the bolts, yielding
or tear-out of the connected material or by an unacceptable deformation around the holes. In other
words, the ultimate resistance of the connection is not related to the slip load.
The slip and bearing resistances are computed separately for application at different load
combinations, as described further below. Because a high tensile preloading of the bolt (developed
by properly installing the bolt to a prescribed torque or rotation) is required to develop a significant
resisting friction force, only bolts with a high tensile yield strength (i.e., ASTM F3125 Grade A325
and A490 high-strength bolts) can be used in slip-critical connections.
This Article specifies that slip-critical connections be proportioned to prevent slip under Load
Combination Service II (Article 3.4.1), and to provide bearing and shear and tensile resistance
under the applicable strength load combinations. Slip is to be prevented under Load Combination
Service II to control permanent deformations caused by slip in bolted joints that could adversely
affect the serviceability of the structure. Slip of bolted field splices in flexural members is also to
be checked during the deck casting (see the Discussion of Article 6.13.6.1.3 in this Guide). It is
further assumed that under the strength load combinations, slip between the bolted parts occurs at

NSBA Guide to Navigating Routine Steel Bridge Design / 303


the higher loads and that the bolts have gone into bearing against the connected material. Thus, the
shear resistance of the bolts and bearing resistance of the bolt holes must be checked under the
appropriate strength load combination. In addition, the resistance of the connected material must
be checked at the strength limit state (See the Discussion of Articles 6.13.2.8, 6.13.2.7, 6.13.2.9,
and 6.13.5 in this Guide for information on the calculation of the slip, shear, and bearing resistances
of bolted connections and the resistance of the connected material, respectively).
According to the provisions of this Article, bolted joints subjected to stress reversal, heavy impact
loads, or severe vibration, or joints located where stress or strain due to joint slippage would be
detrimental to the serviceability of the structure, are to be designated as “slip-critical” (the
Engineer is referred to this Article for the complete list of joints that should be designated as slip-
critical). Repeated loading may introduce fatigue concerns if slip occurs in these cases, particularly
when oversize or slotted holes are used. Bolted field splices in routine steel I-girder bridges should
always be designed as slip-critical as these connections are typically subjected to significant stress
reversal under live loads. Strictly speaking, bracing members in routine steel I-girder bridges
(which are specifically identified as secondary members, and which are typically not subject to
“significant load reversal”) can be designed as bearing-type connections, as discussed in Article
6.13.2.1.2, unless Owner-agency policy requires otherwise.
For further information on high-strength bolts, including installation provisions and verification
procedures, consult the AISC Design Guide 17 High Strength Bolts - A Primer for Engineers, the
RCSC Specifications for Structural Joints Using High-Strength Bolts available from the Research
Council on Structural Connections (RCSC), the AASHTO-NSBA Steel Bridge Collaboration
Guide Specification S2.1-2018 Steel Bridge Fabrication Guide Specification, the AASHTO LRFD
Bridge Construction Specifications, and the applicable Owner-agency standards.

6.13.2.1.2 Bearing-Type Connections


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article deal with the design of high-strength bolted bearing-type
connections. In high-strength bolted bearing-type connections, the load is resisted by a
combination of the shear resistance of the bolt, the bearing resistance of the connected material
and an unknown amount of friction between the faying surfaces. The failure of a bearing-type
connection will be by shear failure of the bolts, yielding or tear-out of the connected material or
by an unacceptable deformation around the holes, with the final failure load independent of the
clamping force provided by the bolts.
This Article specifies that bearing-type connections are only permitted for joints subject to axial
compression or joints in bracing members (e.g., the connections of cross-frame or diaphragm
members). Connections utilizing ASTM A307 bolts are also to be designed as bearing-type
connections. Such connections are to be designed to provide the required factored resistance in
shear and bearing at the strength limit state. Faying surfaces of bearing-type connections need not
satisfy the surface condition preparation specified in Article 6.13.2.8 for slip-critical connections

NSBA Guide to Navigating Routine Steel Bridge Design / 304


(see the Discussion of Article 6.13.2.8 in this Guide); coatings are permitted in bearing-type
connections.
Typically, of the situations outlined above, only joints in bracing members are found on the routine
steel I-girder bridges covered by this Guide. Compression-only connections are unlikely, and A307
non-pretensioned bolts are rarely, if ever, used in structural connections in steel girder bridges.
The typical practice has been so to use pretensioned high strength bolts (Grade ASTM A325 bolts)
in cross-frame and diaphragm connections on routine steel I-girder bridges, and such practice is
recommended by this Guide.
For further information on high-strength bolts, including installation provisions and verification
procedures, consult the AISC Design Guide 17 High Strength Bolts - A Primer for Engineers, the
RCSC Specifications for Structural Joints Using High-Strength Bolts available from the Research
Council on Structural Connections (RCSC), the AASHTO-NSBA Steel Bridge Collaboration
Guide Specification S2.1-2018 Steel Bridge Fabrication Guide Specification, the AASHTO LRFD
Bridge Construction Specifications, and the applicable Owner-agency standards.

6.13.2.2 Factored Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article are used to determine the factored resistance of a bolted connection
at the strength limit state (Eq. 6.13.2.2-2) and the factored resistance of an individual bolt under
the Service II load combination (Eq. 6.13.2.2-1). The resistance factor is implicitly taken equal to
1.0 in the AASHTO LRFD BDS at the service limit state and therefore is not shown in the
resistance equation for the Service II load combination. Although not stated, Eq. 6.13.2.2-1 should
also be used to compute the factored resistance of an individual bolt in a bolted field splice under
the factored loading used to investigate slip during the deck casting.
These provisions are generally applicable to the bolted connections in the routine steel I-girder
bridges covered by this Guide. However, in most situations routine steel I-girder bridges are not
subject to externally-applied tensile forces, at least not greater than their specified pretension, and
so the provisions related to bolts subject to axial tension or combined axial tension and shear (i.e.,
Eq. 6.13.2.2-3) are only conditionally applicable.
An example of a situation where bolts might be subject to axial tension (and in fact would likely
be subject to combined axial tension and shear) would be the connection of a diaphragm to the
web of a rolled steel beam using a partial-depth bolted angle connection. Such a detail is permitted
in Article 6.6.1.3.1, which allows the connection of intermediate diaphragms on rolled beams in
straight bridges with composite reinforced concrete decks whose supports are normal or are
skewed less than 10 degrees from normal, and where those diaphragms are placed in contiguous
lines parallel to the supports (see the Discussion of Article 6.6.1.3.1 in this Guide). In such a case,
the bolted connection to the web would be subject to an out-of-plane moment inducing tension in
some of the bolts connecting the angle to the web.

NSBA Guide to Navigating Routine Steel Bridge Design / 305


Alternately, in most cases, the bolted connections in routine steel I-girder bridges are subject
primarily to shear, with only minor moments due to eccentricities between the connection and the
centroid of the connected member. In most cases, the dimensions of the connection are large
enough and the eccentricity is small enough that it is unlikely that the bolts will be subject to a net
externally-applied tensile force larger than their pretension. Typical cases where checking
combined axial tension and shear in bolts is not necessary include bolted field splice connections
and truss-type cross-frame connections.
In general, high-strength bolted connections designed according to the these provisions will have
a higher reliability at the strength limit state than the connected parts because the resistance factors
for the design of bolted connections were selected to provide a higher level of reliability than those
chosen for member design. Also, the controlling strength limit state in the connected part, e.g.,
yielding or deflection, is typically reached well before the controlling strength limit state in the
connection, e.g., the bolt shear resistance or the bearing resistance of the connected material.

6.13.2.3 Bolts, Nuts, and Washers

6.13.2.3.1 Bolts and Nuts


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article simply refer to the provisions of Article 6.4.3 for the material
specifications for high-strength bolts to be used in the bolted connections of the routine steel I-
girder bridges covered by this Guide (see the Discussion of Article 6.4.3 in this Guide).

6.13.2.3.2 Washers
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article simply refer to the provisions of Article 11.5.5.4.3 of the AASHTO
LRFD Bridge Construction Specifications and to the provisions of Article 6.4.3.1.3 for the material
specifications and other requirements for hardened washers to be used in the bolted connections
of the routine steel I-girder bridges covered by this Guide (see the Discussion of Article 6.4.3.1.3
in this Guide).
The provision in this Article related to direct tension indicators (DTIs) installed over an oversize
or slotted hole in an outer ply applies only if DTIs are used in such connections. DTIs are washers
which include mechanical features (typically small arch-shaped protrusions) which compress in
response to the pretension developed in the bolt. When correctly calibrated, the amount of
pretension can be determined by measuring the gap remaining between the washer and the
connected element. The use of DTIs is subject to Owner-agency preferences; check Owner-agency
policies and specifications before requiring or allowing their use.

NSBA Guide to Navigating Routine Steel Bridge Design / 306


6.13.2.4 Holes

6.13.2.4.1 Type

6.13.2.4.1a General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article indicates that standard-size bolt holes (see the Discussion of Article 6.13.2.4.2 in this
Guide) are to be used in high-strength bolted connections, unless otherwise specified. This
provision is applicable to the bolted connections in the routine steel I-girder bridges covered by
this Guide.

6.13.2.4.1b Oversize Holes


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article indicates that oversize bolt holes (see the Discussion of Article 6.13.2.4.2 in this
Guide) may be used in any or all plies of slip-critical connections but are not to be used in bearing-
type connections (see the Discussion of Articles 6.13.2.1.1 and 6.13.2.1.2 in this Guide).
For bolted field splices, the use of slip-critical connections is required, and the use of oversize
holes is not permitted (see the Discussion of Article 6.13.6.1.3a in this Guide).
For cross-frame and diaphragm connections in the routine steel I-girder bridges covered by this
Guide, the use of oversize holes is neither required nor prohibited by the AASHTO LRFD BDS.
Article 6.13.2.1.1 requires that, “Joints subject to stress reversal, heavy impact loads, severe
vibration or located where stress and strain due to joint slippage would be detrimental to the
serviceability of the structure shall be designated as slip-critical.” That Article also requires that
“joints in shear with bolts installed in oversize holes” are to be designated as slip-critical.
Designation of a joint as slip-critical triggers fabrication requirements which specify the required
conditions for the faying surfaces of the connection and the use of pretensioned high-strength bolts.
Also, as stated in Article 6.13.2.1.1, designation of a joint as slip-critical triggers the requirement
that the connection be proportioned to prevent slip under the Service II load combination.
However, the Service II load combination, when considered in the context of cross-frames or
diaphragms in routine steel I-girder bridges, includes only dead load and live load force effects.
Live load force effects are generally negligible in cross-frames or diaphragms of the routine steel
I-girder bridges covered by this Guide; in fact, since line girder analysis is recommended for the
design of routine steel I-girder bridges, live load force effects are not usually calculated for the
cross-frames or diaphragms. Consequently, bolted cross-frame or diaphragm connections in
routine steel I-girder bridges are typically not checked for slip resistance, regardless of whether
standard size or oversize holes are used. Designers should consult Owner-agency policy regarding
the use of standard or oversize holes for cross-frame or diaphragm connections in routine steel I-
girder bridges. In the absence of such policy, the use of standard-size holes is recommended.

NSBA Guide to Navigating Routine Steel Bridge Design / 307


The use of oversize holes in other bolted connections in routine steel I-girder bridges should be
evaluated on a case-by-case basis, comparing the value of facilitating easy fit-up of the structure
versus the value of maintaining tighter control of the constructed geometry of the structure. .

6.13.2.4.1c Short-Slotted Holes


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article indicates that short-slotted holes (see the Discussion of Article 6.13.2.4.2 in this
Guide) may be used in any or all plies of either slip-critical or bearing-type connections (see the
Discussion of Articles 6.13.2.1.1 and 6.13.2.1.2 in this Guide). In slip-critical connections, the
slots may be used without regard to the direction of loading. However, in bearing-type connections,
the length of the slot must be normal to the direction of the load.
Short-slotted holes are not typically used, nor are they recommended for use, on bolted connections
in the routine steel I-girder bridges covered by this Guide. However, they may occasionally be
used or necessary in special situations, such as phased construction or widening projects.
Vertical slotted holes for cross-frame connections have sometimes been used in straight skewed I-
girder bridges in an attempt to minimize the twist of the girders and reduce the cross-frame forces.
Such an approach is not recommended, however, as it becomes difficult to control the vertical
deflections during the deck placement.

6.13.2.4.1d Long-Slotted Holes


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article indicates that long-slotted holes (see the Discussion of Article 6.13.2.4.2 in this Guide)
may be used in only one ply of either slip-critical or bearing-type connections (see the Discussion
of Articles 6.13.2.1.1 and 6.13.2.1.2 in this Guide). In slip-critical connections, the slots may be
used without regard to the direction of loading. However, in bearing-type connections, the length
of the slot must be normal to the direction of the load.
Long-slotted holes are not typically used, nor are they recommended for use, on bolted connections
in the routine steel I-girder bridges covered by this Guide. However, they may occasionally be
used or necessary in special situations, such as phased construction or widening projects.
Vertical slotted holes for cross-frame connections have sometimes been used in straight skewed I-
girder bridges in an attempt to minimize the twist of the girders and reduce the cross-frame forces.
Such an approach is not recommended, however, as it becomes difficult to control the vertical
deflections during the deck placement.

6.13.2.4.2 Size
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 308


Discussion:
This Article provides the maximum permitted size of standard, oversize, short-slotted and long-
slotted holes in bolted connections (Table 6.13.2.4.2-1). This Article is applicable to the routine
steel I-girder bridges covered by this Guide.
In Table 6.13.2.4.2-1, d is the diameter of the bolt. Standard-size holes for bolts with diameters
less than 1 in. are 1/16-in. larger than the d. Oversize holes for bolts with diameters less than 1 in.
are 3/16-in. larger than d. Standard-size holes for bolts with diameters greater than or equal to 1
in. are 1/8-in. larger than d, which avoids the need to field ream holes to fit large-diameter hot-
forged bolts, which often have a longitudinal forging seam that interferes with holes 1/16 in. larger
than the bolt diameter.
Article 6.8.3 specifies that for the calculation of the net area in design calculations, the width of
standard bolt holes is to be taken as the nominal diameter of the hole (see the Discussion of Article
6.8.3 in this Guide). The width of oversize and slotted holes is to be taken as the nominal diameter
or width of the hole, as applicable, given in Table 6.13.2.4.2-1.

6.13.2.5 Size of Bolts


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article gives some specific requirements regarding the size of bolts in bolted connections.
These requirements are applicable to the routine steel I-girder bridges covered by this Guide. Bolts
are not to be less than 0.625 in. in diameter. Bolts 0.625 in. in diameter are not to be used in primary
members, except for 2.5-in. legs of angles and in flanges of sections whose dimensions require
0.625-in. bolts to satisfy other detailing provisions given in the specifications. Structural shapes
that do not permit the use of 0.625-in. bolts are to be limited to use in handrails.

6.13.2.6 Spacing of Bolts

6.13.2.6.1 Minimum Spacing and Clear Distance


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article provides minimum spacing and clear distance requirements for bolted connections,
which are applicable to the routine steel I-girder bridges covered by this Guide.
The minimum spacing between centers of bolts in standard holes in any direction is not to be less
than 3.0d, where d is the diameter of the bolt. The minimum clear distance, Lc, between the edges
of adjacent bolt holes in the direction of the force and transverse to the direction of the force is not
to be less than 2.0d when oversize or slotted holes are used.

6.13.2.6.2 Maximum Spacing for Sealing Bolts


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 309


Discussion:
This Article provides maximum spacing requirements for sealing bolts in bolted connections,
which are applicable to the routine steel I-girder bridges covered by this Guide. These requirements
are intended to help seal against the penetration of moisture in joints and apply to the lines of bolts
in the connection adjacent to a free edge of an outside plate or shape.
Maximum spacing requirements are specified for a single line of bolts (Eq. 6.13.2.6.2-1), and for
the staggered spacing in two lines considered together where there is a second line of bolts
uniformly staggered with the line adjacent to the free edge at a gage distance less than 1.5 + 4.0t
(Eq. 6.13.2.6.2-2). The staggered spacing need not be less than one-half the requirement for a
single line. The thickness, t, in these requirements is the thickness of the thinner outside plate or
shape. The absolute maximum spacing in both instances is 7.0 in.
In uncoated weathering steel structures, it is critical that the bolt spacing be such that the
connection joint is tight and excess moisture cannot enter between the plies of material. If sufficient
moisture enters the joint, the resulting corrosion may cause prying, or pack-out, of the joint or bolt
failure. The maximum bolt spacing requirements for sealing bolts have been shown to provide
proper tightness and stiffness of uncoated weathering steel bolted joints to avoid joint prying and
corrosion pack-out.

6.13.2.6.3 Maximum Pitch for Stitch Bolts


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article presents maximum pitch requirements for stitch bolts, which fasten together built-up
compression or tension members at intervals along their length, such that the separate components
will act as a unit to resist buckling of compression members. Requirements are given for a single
line of bolts (12t for compression members and 24t for tension members) and for two adjacent
lines of staggered bolts (Eq. 6.13.2.6.3-1 for compression members and twice the value from that
equation for tension members). The gage between adjacent lines of bolts is not to exceed 24t for
compression and tension members. The thickness, t, in these requirements is the thickness of the
thinner outside plate or shape. The pitch is not to exceed the maximum pitch specified for sealing
bolts.
These provisions are only applicable to mechanically fastened built-up members subject to axial
compression or tension. These types of members are not used in routine steel I-girder bridges;
therefore, the provisions of this Article are not applicable.

6.13.2.6.4 Maximum Pitch for Stitch Bolts at the End of Compression Members
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article provides more stringent maximum pitch requirements for stitch bolts at the ends of
mechanically fastened built-up members subject to axial compression. The pitch, p, of the stitch
bolts must not exceed 4.0d for a length equal to 1.5 times the maximum width of the member,

NSBA Guide to Navigating Routine Steel Bridge Design / 310


where d is the diameter of the bolt. Beyond this length, p may be increased gradually over a length
equal to 1.5 times the maximum width of the member until the maximum pitch given by either
12.0t or Eq. 6.13.2.6.3-1, as applicable, is reached (see the Discussion of Article 6.13.2.6.3 in this
Guide). The thickness, t, in these requirements is the thickness of the thinner outside plate or shape.
These provisions are only applicable to mechanically fastened built-up members subject to axial
compression. These types of members are not used in routine steel I-girder bridges; therefore, the
provisions of this Article are not applicable.

6.13.2.6.5 End Distance


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies the minimum and maximum end distance requirements for all types of holes
in bolted connections and are applicable to the bolted connections in the routine steel I-girder
bridges covered by this Guide.
The end distance of bolts is defined in the AASHTO LRFD BDS as the distance along the line of
force between the center of a hole and the end of the component. This Article specifies that the
end distance for all types of holes is not to less than the appropriate minimum edge distance
specified in Table 6.13.2.6.6-1 (see the Discussion of Article 6.13.2.6.6 in this Guide). When
oversize or slotted holes are used, the minimum clear end distance, which is defined as the distance
between the edge of the bolt hole and the end of the member, must not be less than the bolt
diameter.
The maximum end distance is to be taken the same as the maximum edge distance, or the lesser of
eight times the thickness of the thinnest outside plate and 5.0 in.

6.13.2.6.6 Edge Distances


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies the minimum and maximum edge distance requirements for all types of holes
in bolted connections and are applicable to the bolted connections in the routine steel I-girder
bridges covered by this Guide.
The edge distance of bolts is defined in the AASHTO LRFD BDS as the distance perpendicular to
the line of force between the center of a hole and the edge of the component. The minimum edge
distance is a function of the diameter of the bolt. This Article specifies that the minimum edge
distance is to be taken as specified in Table 6.13.2.6.6-1.
The maximum edge distance is not to be more than the lesser of eight times the thickness of the
thinnest outside plate and 5.0 in.
The minimum edge distances are based on standard fabrication practices and workmanship
tolerances. Providing edge (and end) distances larger than the specified minimum edge distances,

NSBA Guide to Navigating Routine Steel Bridge Design / 311


but not larger than the specified maximum edge distances, is encouraged as it allows for the
occurrence of minor fabrication errors (due to unavoidable workmanship tolerances) without
violating the specified minimum distances. Also, satisfying the provisions of Article 6.13.2.9
related to the bearing resistance of bolt holes (see the Discussion of Article 6.13.2.9 in this Guide)
provides sufficient end distances such that bearing and tear-out limits are not exceeded for bolts
adjacent to all types of edges.
The maximum edge distance limits are intended to help seal the faying surfaces. In uncoated
weathering steel structures, it is critical that the connection joint is tight and excess moisture cannot
enter between the plies of material. If sufficient moisture enters the joint, the resulting corrosion
may cause prying, or pack-out, of the joint or bolt failure. The maximum edge distance limits have
been shown to provide proper tightness and stiffness of uncoated weathering steel bolted joints to
avoid joint prying and corrosion pack-out.

6.13.2.7 Shear Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article are used to calculate the nominal shear resistance, Rn, of bolts at the
strength limit state and are applicable to the connections in the routine steel I-girder bridges
covered by this Guide. Note that the nominal shear resistance of anchor rods is covered separately
in Article 6.13.2.12 (see the Discussion of Article 6.13.2.12 in this Guide). In slip-critical
connections (see the Discussion of Article 6.13.2.1.1 in this Guide), it is assumed that the bolts in
the connection have slipped and gone into bearing at the strength limit state such that the shear
resistance of the bolts and bearing resistance at the bolt holes must be checked.
The nominal shear resistance depends on whether the threads are excluded from (Eq. 6.13.2.7-1)
or included in (Eq. 6.13.2.7-2) the shear plane. Eqs. 6.13.2.7-1 and 6.13.2.7-2 apply for
determining the nominal shear resistance of a high-strength bolt (ASTM F3125 bolt) or an ASTM
A307 bolt at the strength limit state in joints whose length between extreme fasteners measured
parallel to the line of action of the force is less than or equal to 38.0 in. For a bolt in a lap-splice
tension connection greater than 38.0 in. in length, the nominal shear resistance is to be taken as
0.83 times the value given by Eq. 6.13.2.7-1 or 6.13.2.7-2, as applicable. The nominal shear
resistance is based on the observation that the shear strength of a single high-strength bolt is about
0.625 times the tensile strength Fub of the bolt (see Table 6.4.3.1.1-1 and the Discussion of Article
6.4.3.1.1 in this Guide for the determination of Fub).
The shear resistance is not affected by the pretension in the bolts provided the connected material
is in contact at the faying surfaces. In shear connections with more than two bolts in the line of
force, the average bolt strength decreases as the joint length increases due to the nonuniform bolt
shear force distribution caused by deformation of the connected material. For joints up to 38.0 in.
in length, a single reduction factor of 0.90 is implicitly applied to the 0.625 multiplier described
above rather than providing a function that reflects the decrease in average bolt strength with joint
length (0.90 times 0.625 equals the 0.56 multiplier given in Eq. 6.13.2.7-1). This was felt not to
adversely affect the economy of very short joints. For bolts in joints longer than 38.0 in., the

NSBA Guide to Navigating Routine Steel Bridge Design / 312


nominal shear resistance must be reduced by an overall reduction factor of 0.75. Therefore, the
nominal shear resistance of bolts in joints longer than 38.0 in. must be further reduced by an
additional factor of 0.83 or 0.75/0.90. For bolted flange splices, note that the 38.0 in. length is to
be measured between the extreme bolts on only one side of the connection. The greater than 38.0
in. length reduction does not apply to a bolted web splice because a web splice is not a lap-splice
tension connection.
When bolts are positioned so that they cross two planes of contact (i.e., Ns = 2), this is referred to
as double shear. Double shear is a symmetrical loading situation with regard to the shear planes
and direction of shear transfer. When there is a single plane of contact involved in the load transfer
(i.e., Ns = 1), this is referred to as single shear, which is an unsymmetrical loading situation.
The average ratio of the nominal shear resistance for bolts with threads included in the shear plane
to the nominal shear resistance for bolts with threads excluded from the shear plane is 0.83 with a
standard deviation of 0.03. Therefore, a reduction factor of 0.80 is conservatively used to account
for the nominal shear resistance when threads are included in the shear plane but calculated with
the area corresponding to the nominal bolt diameter (0.56 * 0.80 equals the 0.45 multiplier given
in Eq. 6.13.2.7-2).
In determining whether the threads are excluded from the shear planes, the thread length of the
bolt is to be determined as two thread pitches greater than the specified thread length. If the threads
of a bolt are included in a shear plane of a joint, the nominal shear resistance of the bolts in all
shear planes of the joint is to conservatively be taken from Eq. 6.13.2.7-2. That is, for bolts in
double shear with a non-threaded shank in one shear plane and a threaded section in the other shear
plane, the sharing of the load between the two dissimilar shear areas is uncertain. Also, knowledge
about the specific bolt placement, which might result in both shear planes being in the threaded
section, is not ordinarily available to the Engineer. A handy summary of typical cases is provided
in a footnote in Section 2.1.1.3 of NSBA’s Bolted Field Splices for Steel Bridge Flexural Members
– Overview and Design Examples. The footnote is excerpted below for convenience:
“In determining the factored shear resistance of the bolts, if the flange splice-plate
thickness closest to the nut is greater than or equal to 0.5-in. thick, the nominal shear
resistance of the bolts should be determined assuming the threads are excluded from the
shear planes for bolts less than 1.0 in. in diameter. For bolts greater than or equal to 1.0 in.
in diameter, the nominal shear resistance of the bolts should be determined assuming the
threads are excluded from the shear planes if the flange splice-plate thickness closest to the
nut is greater than 0.75 in. in thickness. Otherwise, the threads should be assumed included
in the shear planes. The preceding assumes there is one washer under the nut, and that there
is no stick-out beyond the nut, which represents the worst case for this determination. Web
splice connections will generally have threads included in the shear plane due to the
relatively thin splice plates in the web connections.”
Since the threaded length of an ASTM A307 bolt is not as predictable as that of a high-strength
bolt, the nominal shear resistance of an A307 bolt must always be based on Eq. 6.13.2.7-2. Also,
A307 bolts with a long grip (i.e., the total thickness of the plies of a joint through which the bolt
passes exclusive of any washers or load-indicating devices) tend to bend, reducing their shear
resistance. Therefore, this Article requires that when the grip length of an A307 bolt exceeds 5.0

NSBA Guide to Navigating Routine Steel Bridge Design / 313


bolt diameters, the nominal shear resistance must be lowered 1.0 percent for each 1/16 in. of grip
in excess of 5.0 bolt diameters.
Bolted cross-frame gusset plate connections are often subject to eccentric shear; that is, when one
of the resultant member forces is applied on a line of action that does not pass through the center
of gravity of the bolt group. The resultant action may be represented as a net moment (torque)
equal to the resultant force times its eccentricity from the center of gravity, and a concentric force
acting on the connection. Since both the moment and concentric force cause shears on the bolt
group, this particular situation is referred to as eccentric shear.
For further discussion and an example design calculation of eccentric shear, consult Section
6.6.4.2.6 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures; this topic is also addressed in most steel design
textbooks.
For an example design calculation of the shear resistance of a high-strength bolt, consult Section
6.6.4.2.5.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures.

6.13.2.8 Slip Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article are used to calculate the nominal slip resistance, Rn, of a bolt in a
high-strength bolted slip-critical connection under Load Combination Service II, and in some
instances, during the deck casting (see the Discussion of Article 6.13.2.1.1 in this Guide). These
provisions are applicable to the slip-critical connections in the routine steel I-girder bridges
covered by this Guide.
The bolt pretension and surface condition of the faying surface (i.e., coefficient of friction) have
the greatest effect on the slip resistance of high-strength bolted connections. The minimum
required bolt pretension during bolt installation, Pt, to be used in Eq. 6.13.2.8-1 is given for ASTM
F3125 high-strength bolts in Table 6.13.2.8-1.
The surface condition factor, Ks, (i.e., coefficient of friction) to be used in Eq. 6.13.2.8-1 is
provided in Table 6.13.2.8-3 and is a function of the class of the surface. Four different classes of
surfaces (Classes A through D) are defined based on the mean value of slip coefficients from many
tests of clean mill scale, uncoated and coated blast-cleaned steel surfaces, unsealed pure zinc or
85/15 zinc-aluminum thermal-sprayed (i.e., metalized) surfaces with coating thicknesses not
exceeding 16 mils, and hot-dip galvanized surfaces. The surface condition is typically defined,
either directly or indirectly, in the Owner-agencies standard specifications.
It has been found that if tightly adherent mill scale is on the faying surface of a bolted connection
on uncoated weathering steel, the connection slips into bearing at a lower shear stress than on a
carbon steel with mill scale. However, if the faying surface is blast-cleaned to the Society for
Protective Coatings (SSPC) standard SP-6 (SSPC SP-6) or better, slip-critical connections on
uncoated weathering steel can be designed using a Class B surface condition. Otherwise, a Class

NSBA Guide to Navigating Routine Steel Bridge Design / 314


A surface condition, which is appropriate for clean mill-scale surfaces, must be used. The slip
resistance of bolted joints is not affected by the weathering of uncoated steel surfaces prior to
erection, but any loose rust on the connection or faying surfaces must be removed. Pre-construction
primers may be used for the cleaned bolted surfaces. Hot-dip galvanized faying surfaces need not
be treated after galvanizing. Consult the Commentary for this Article for information on the effects
of paint overspray and the required testing to qualify a particular coating to be used under these
specifications.
Since faying surfaces (that are not galvanized) are typically blast-cleaned as a minimum, a Class
A surface condition should only be used to compute the slip resistance when Class A coatings are
applied or when unpainted mill scale is left on the faying surface. Most commercially available
primers will qualify as Class B or Class D coatings.
Since all locations must develop the slip resistance before a total joint slip can occur at that plane,
the assumption is made that the slip resistance at each bolt is equal and additive with the slip
resistance at the other bolts in the connection. It is also assumed that the full slip resistances must
be mobilized at each slip plane before full joint slip can occur, although the forces at each slip
plane do not necessarily develop simultaneously. Eq. 6.13.2.8-1 is formulated for the case of a
single slip plane. Therefore, the total slip resistance of a joint with multiple slip planes can be taken
equal to the resistance of a single slip plane multiplied by the number of slip planes, Ns.
Hole size factors, Kh, in Eq. 6.13.2.8-1 less than 1.0 are provided for bolts in oversize or slotted
holes (Table 6.13.2.8-2) because of the greater possibility of significant deformation occurring in
joints with oversize or slotted holes. For long-slotted holes, even though the slip load is the same
for bolts loaded transverse or parallel to the axis of the slot, the hole size factor for loading parallel
to the axis has been reduced, based upon judgment, because of the greater consequences of slip in
this case.
In a slip-critical connection subject to the combined effects of a net overall axial tension and shear,
the tensile force reduces the contact pressure between the connected plates thereby reducing the
slip resistance to the shear forces. The reduction in slip resistance is approximately proportional to
the ratio of the applied tensile force to the bolt installation tension. The reduction factor is given
by Eq. 6.13.2.11-3 (see the Discussion of Article 6.13.2.11 in this Guide). The local reduction in
contact pressure due to an overturning moment causing a local tension in one part of a connection
does not reduce the slip resistance since there is a corresponding increase in contact pressure in
other parts of the connection.
For an example design calculation of the slip resistance of a high-strength bolt, consult Section
6.6.4.2.4.2 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures.

6.13.2.9 Bearing Resistance at Bolt Holes


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 315


Discussion:
The provisions of this Article are used to calculate the nominal bearing resistance at bolt holes, Rn,
in a bolted connection at the strength limit state and are applicable to the connections in the routine
steel I-girder bridges covered by this Guide. In slip-critical connections (see the Discussion of
Article 6.13.2.1.1 in this Guide), it is assumed that the bolts in the connection have slipped and
gone into bearing at the strength limit state such that the shear resistance of the bolts and bearing
resistance at the bolt holes must be checked.
After a major slip has occurred in a slip-critical connection, one or more bolts are in bearing against
the side of the hole. A bearing failure relates generally to either deformation of the bolt or
deformation around a bolt hole. Tests have consistently shown that the bearing stress on the bolt
itself is not critical. The contact pressure between the bolt and connected material can be expressed
as the bearing stress on the connected material. For simplicity, the bearing stress is assumed to be
a uniform stress distribution equal to the load transmitted by the bolt divided by the bearing area
taken as the bolt diameter times the thickness of the connected material. The actual failure mode
depends on the clear end distance or clear distance between the bolts, the bolt diameter, and the
thickness of the connected material. Either the bolt will split out through the end of the plate
because of insufficient end distance, or else excessive deformations are developed in the connected
material adjacent to the bolt hole because of insufficient clear distance between the bolts.
The nominal bearing resistance of an interior hole is based on the clear distance, Lc, between the
edge of the hole and the edge of the adjacent hole in the direction of the bearing force. The nominal
bearing resistance of an end hole is based on the clear distance, Lc, between the edge of the hole
and the end of the member. Eq. 6.13.2.9-1 or 6.13.2.9-2 is typically applicable for the routine steel
I-girder bridges covered by this Guide as these equations apply to standard-size holes. These two
equations can be combined and written more simply for design as Rn = 1.2 LctFu  2.4dtFu . Eqs.
6.13.2.9-3 and 6.13.2.9-4 apply only for the case of long-slotted holes perpendicular to the applied
bearing force and can be combined similarly.
The design bearing resistance is expressed in terms of a single bolt, although it is truly for the
connected material adjacent to the bolt. Since bearing failure is generally related to either
deformation of the bolt or deformation around a bolt hole, rather than a fracture-type failure, the
connection will continue to carry its failure load if subject a force greater than the bearing
resistance. Therefore, in calculating the nominal bearing resistance for the connected part, the total
bearing resistance may be taken as the sum of the bearing resistances of the individual bolts (holes)
parallel to the line of the applied force. Also, if the nominal bearing resistance of a bolt hole
exceeds the nominal shear resistance of the bolt determined as specified in Article 6.13.2.7 (see
the Discussion of Article 6.13.2.7 in this Guide), the nominal bearing resistance of the bolt hole is
to be limited to the nominal shear resistance of the bolt.
For an example design calculation of the bearing resistance of a connected part in a high-strength
bolted connection, consult Section 6.6.4.2.5.3 of the Reference Manual for NHI Course 130081,
Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.

NSBA Guide to Navigating Routine Steel Bridge Design / 316


6.13.2.10 Tensile Resistance

6.13.2.10.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains general provisions related to the tensile resistance of high-strength bolts. The
Article specifies that high-strength bolts subject to axial tension must be pretensioned to the level
given in Table 6.13.2.8-1 (see the Discussion of Article 6.13.2.8 in this Guide) regardless of
whether the design is for a slip-critical or a bearing-type connection.
Axial tension occurring without simultaneous shear occurs in bolts for tension members such as
hangers or other members whose line of action is perpendicular to the member to which it is
fastened. The applied tensile force must be taken as the force due to externally applied loads plus
any tension resulting from prying action produced by deformation of the connected parts as
specified in Article 6.13.2.10.4 (see the Discussion of Article 6.13.2.10.4 in this Guide).
Bolted connections in the routine steel I-girder bridges covered by this Guide generally are not
subject to axial tension occurring without simultaneous shear, and so the provisions of this Article
generally are not applicable. Of course, unique, specialty details may involve tension-only
connections, but these types of details would be the exception rather than the rule in routine steel
I-girder bridges.

6.13.2.10.2 Nominal Tensile Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article specifies the nominal tensile resistance of a bolt, Tn, independent of any initial
tightening force. The tensile resistance of a bolt is the product of the tensile strength of the bolt
and the tensile stress area through the threaded portion of the bolt. The tensile stress area is
approximately 76 percent of the nominal cross-sectional area of the bolt for the usual sizes of a
structural bolt. Hence, the nominal tensile resistance per unit area, based on the nominal area of
the bolt, is taken as 76 percent of the tensile strength of the bolt.
The specified nominal tensile resistance is approximately equal to the initial tightening force
specified in 6.13.2.8-1. Thus, when a tensile force is applied to a high-strength bolt that has been
properly pretensioned, the increase in the bolt tension is generally much smaller than the applied
load. There will be little increase in bolt force above the pretension load at service load levels.
After the parts separate, the bolt will act as a tension member with the applied force equaling the
bolt tension. As a result, bolts in connections subject to axial tension are required to be fully
pretensioned.
Bolted connections in the routine steel I-girder bridges covered by this Guide generally are not
subject to axial tension occurring without simultaneous shear, and so the provisions of this Article
generally are not applicable. Of course, unique, specialty details may involve tension-only

NSBA Guide to Navigating Routine Steel Bridge Design / 317


connections, but these types of details would be the exception rather than the rule in routine steel
I-girder bridges.

6.13.2.10.3 Fatigue Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article deal with the fatigue resistance of high-strength bolts subject to axial
tension. Properly pretensioned high-strength bolts subject to fatigue in axial tension must satisfy
Eq. 6.6.1.2.2-1 (see the Discussion of Article 6.6.1.2.2 in this Guide). The stress range (f) in the
equation is to be taken as the stress range in the bolt due to the passage of the 72-kip fatigue design
load (plus the 15 percent dynamic load allowance) specified in Article 3.6.1.4 (see the Discussion
of Article 3.6.1.4 in this Guide), plus any prying force resulting from the cyclic application of the
fatigue load (see the Discussion of Article 6.13.2.10.4 in this Guide); the initial tension in the bolts
is not to be included. The stress range is to be computed using the nominal diameter of the bolt.
In calculating the nominal fatigue resistance (F)n from Eq. 6.6.1.2.5-1 or 6.6.1.2.5-2, as
applicable, the detail category constant A and the constant-amplitude fatigue threshold (F)TH for
ASTM F3125 Grade A325 and A490 bolts in axial tension are to be taken directly from Tables
6.6.1.2.5-1 and 6.6.1.2.5-3, respectively, when the bolts are fully pretensioned. Otherwise,
Condition 9.2 in Table 6.6.1.2.3-1 applies (see the Discussion of Articles 6.6.1.2.3 and 6.6.1.2.5
in this Guide).
This Article limits the calculated prying force to 30 percent of the externally applied load when
bolts are subject to tensile fatigue loading. This limit is based on limited investigations of prying
effects under fatigue loading. Since low carbon ASTM A307 bolts are of lower strength and are
not pretensioned, this Article prohibits their use in connections subjected to fatigue loading.
Bolted connections in the routine steel I-girder bridges covered by this Guide generally are not
subject to axial tension occurring without simultaneous shear, and so the provisions of this Article
generally are not applicable. Of course, unique, specialty details may involve tension-only
connections, but these types of details would be the exception rather than the rule in routine steel
I-girder bridges.

6.13.2.10.4 Prying Action


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article contains a conservative simplified expression to calculate the tensile force due to
prying action in bolted connections subjected to axial tension. Prying action is defined as a lever
action that exists in connections subjected to axial tension in which the line of application of the
applied load is eccentric to the axis of the bolt, causing deformation of the fitting and an
amplification of the axial force in the bolt. Part 9 of the AISC Manual of Steel Construction
provides a more comprehensive treatment of prying action.

NSBA Guide to Navigating Routine Steel Bridge Design / 318


Bolted connections in the routine steel I-girder bridges covered by this Guide generally are not
subject to pure axial tension and prying action, but in some cases bolted connections may be
subject to combined axial tension and shear, and depending on the particular circumstances of the
detail, consideration of prying action may be warranted. As a result, the provisions of this Article
are considered conditionally applicable.
An example of a situation where bolts might be subject to axial tension (and in fact would likely
be subject to combined axial tension and shear) would be the connection of a diaphragm to the
web of a rolled steel beam using a partial-depth bolted angle connection. Such a detail is permitted
in Article 6.6.1.3.1, which allows the connection of intermediate diaphragms on rolled beams in
straight bridges with composite reinforced concrete decks whose supports are normal or are
skewed less than 10 degrees from normal, and where those diaphragms are placed in contiguous
lines parallel to the supports (see the Discussion of Article 6.6.1.3.1 in this Guide). In such a case,
the bolted connection to the web would be subject to an out-of-plane moment inducing tension in
some of the bolts connecting the angle to the web. Furthermore, depending on the specific detailing
used, the bolts may also be subject to prying action.

6.13.2.11 Combined Tension and Shear


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article deals with the resistance of high-strength bolts in connections subjected to combined
axial tension and shear; e.g., less than full-depth diaphragm end-angle bolted connections to webs
and end-angle bolted connections of stringers to floorbeams and/or floorbeams to girders.
Eqs. 6.13.2.11-1 and 6.13.2.11-2 are simplifications of elliptical interaction curves that provide the
nominal tensile resistance of bolts subjected to combined axial tension and shear. No reduction in
the nominal tensile resistance is required when the applied factored shear force on the bolt is less
than or equal to 33 percent of the nominal shear resistance of the bolt (see the Discussion of Article
6.13.2.7 in this Guide).
In a slip-critical connection subject to the combined effects of a net overall axial tension and shear,
the tensile force reduces the contact pressure between the connected plates thereby reducing the
slip resistance to the shear forces (see the Discussion of Article 6.13.2.8 in this Guide). The
reduction in slip resistance given by Eq. 6.13.2.11-3 is approximately proportional to the ratio of
the applied tensile force to the bolt installation tension. The local reduction in contact pressure due
to an overturning moment causing a local tension in one part of a connection does not reduce the
slip resistance since there is a corresponding increase in contact pressure in other parts of the
connection.
In routine steel I-girder bridges, there are only a few limited cases where bolted connections may
be subject to combined axial tension and shear. As a result, the provisions of this Article are
considered conditionally applicable.
An example of a situation where bolts might be subject to axial tension (and in fact would likely
be subject to combined axial tension and shear) would be the connection of a diaphragm to the

NSBA Guide to Navigating Routine Steel Bridge Design / 319


web of a rolled steel beam using a partial-depth bolted angle connection. Such a detail is permitted
in Article 6.6.1.3.1, which allows the connection of intermediate diaphragms on rolled beams in
straight bridges with composite reinforced concrete decks whose supports are normal or are
skewed less than 10 degrees from normal, and where those diaphragms are placed in contiguous
lines parallel to the supports (see the Discussion of Article 6.6.1.3.1 in this Guide). In such a case,
the bolted connection to the web would be subject to an out-of-plane moment inducing tension in
some of the bolts connecting the angle to the web. Furthermore, depending on the specific detailing
used, the bolts may also be subject to prying action (see the Discussion of Article 6.13.2.10.4 in
this Guide).
A converse example of a situation where consideration of axial tension in a bolted connection can
generally be safely neglected might be the bolted connection of a gusset plate to a cross-frame
connection plate (stiffener). Although some of the bolts in this connection may be subjected to
combined tension and shear due to the out-of-plane moment arising from the eccentricity in the
connection, the eccentricities and cross-frame forces in a routine steel I-girder bridge are small
enough that this effect is unlikely to be significant and may be ignored in these bridges.

6.13.2.12 Shear Resistance of Anchor Rods


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article are used to calculate the nominal shear resistance, Rn, of an ASTM
F1554 anchor rod at the strength limit state. These provisions are applicable if anchor rods are
used to connect bearing sole plates or masonry plates to the substructure.
Since the thread length of anchor rods is not limited by the specifications, threads are assumed
included in the shear plane. The implicit joint length reduction factor of 0.90 is also not applicable
to anchor rods (see the Discussion of Article 6.13.2.7 in this Guide). Therefore, the coefficient in
the expression for Rn given by Eq. 6.13.2.12-1 is equal to the shear strength-to-tensile strength
ratio multiplier of 0.625 times the reduction factor of 0.80 for threads included in the shear plane,
or 0.50 (see the Discussion of Article 6.13.2.7 in this Guide).
The provisions do not cover the global design of the anchorages to the concrete, which is beyond
the scope of this Guide.

6.13.3 Welded Connections

6.13.3.1 General
Determination of applicability, Simple Span Bridges: Applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 320


Discussion:
The provisions for welded connections in the AASHTO LRFD BDS are certainly applicable to the
design of routine steel I-girder bridges where welded steel plate girders are used as the main
spanning elements, since these girders necessarily use welds to connect the flanges to the webs, as
well as other details such as stiffeners and so on. The provisions may also be applicable to the
design of routine steel I-girder bridges where rolled steel beams are used as the main spanning
elements, if welded stiffeners or other welded connection details are involved with the diaphragms
or other features.
Welding is the process of joining two pieces of material, usually metals, by heating the pieces to a
suitable temperature such that the materials are soft enough to coalesce or fuse into one material.
The pieces are held in position for welding and may or may not be pressed together depending on
the process that is used. Arc welding, in which electrical energy in the form of an electric arc is
introduced to generate the heat necessary for welding, is the most commonly used process in the
steel-bridge construction industry. The heat of the electric arc as the current passes through the
system simultaneously melts a consumable electrode (deposited as filler material) and the parts of
the material being joined, with the joint resulting from the cooling and solidification of the fused
material.
To protect the molten region from impurities, the zone to be welded is typically blanketed in an
atmosphere supplied by a flux, which may be a fusible coating on the welding rod, a fusible powder
spread over the line of the weld or a gas sprayed over the weld. To produce a weld of the desired
quality, the properties of the electrode must be carefully controlled. Proper control of the current
and voltage along with a skilled welder are also required to produce a quality weld.
The provisions of this Article refer to the AASHTO/AWS D1.5M/D1.5 Bridge Welding Code
(BWC) for requirements related to base metal, weld metal, and welding design details. The
provisions also refer to AWS Publication A2.4 for information on welding symbols.
These provisions also require the use of matching weld metal in groove and fillet welds, with the
exception that undermatching weld metal may be used in fillet welds if the welding procedure and
weld metal are selected to facilitate the production of sound welds.
Weld metal strength may be classified as matching or undermatching. Matching weld (filler) metal
has a specified minimum tensile strength that is the same as or higher than the lower-strength base
metal. For example, matching weld metal for ASTM A709 Grade 50 steel would be E70 filler
material, where the specified minimum weld/base metal properties tensile strength are 70/65 ksi.
Although the weld metal has slightly higher properties than the base metal in this case, this is
considered to be a matching combination. Matching strengths for various weld and base metal
combinations are specified in the BWC.
According to these provisions, the use of undermatching weld metal is strongly encouraged for
fillet welds connecting steels with specified minimum yield strength greater than 50 ksi. Lower
strength weld metal will generally be more ductile than higher strength weld metal. Since the
residual stresses in a welded connection are assumed to be on the order of the yield point of the
weaker material in the connection, using lower strength weld metal will lower the level of residual
stresses in the base metal at the connection reducing the cracking tendencies. Therefore,

NSBA Guide to Navigating Routine Steel Bridge Design / 321


undermatching welds will be much less sensitive to delayed hydrogen cracking and are more likely
to consistently produce sound welds. In such cases, the Engineer should indicate where
undermatching welds are acceptable on the contract drawings. The use of undermatching weld
metal is not applicable to the routine steel I-girder bridges covered by this Guide, as steels in these
bridges are assumed to be either Grade 36 or Grade 50 steels.
These provisions are applicable to the simple span and multi-span continuous rolled beam bridges
covered by this Guide if welded connections are used for the bracing connections or for bearing
stiffeners (if required – see the Discussion of Article D6.5 in this Guide).
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on weld metal, welding processes, welding design details, and the appropriate designations of
welding symbols and consumables. Extensive guidance on the design of welded connections for
steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.

6.13.3.2 Factored Resistance

6.13.3.2.1 General
Determination of applicability, Simple Span Bridges: Applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.Conditionally applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
The provisions for welded connections in the AASHTO LRFD BDS are certainly applicable to the
design of routine steel I-girder bridges where welded steel plate girders are used as the main
spanning elements, since these girders necessarily use welds to connect the flanges to the webs, as
well as other details such as stiffeners and so on. The provisions may also be applicable to the
design of routine steel I-girder bridges where rolled steel beams are used as the main spanning
elements, if welded stiffeners or other welded connection details are involved with the diaphragms
or other features.
The provisions of this Article point the Engineer to the appropriate Articles for determining the
factored resistance, Rr, of the welded connections (see the Discussions of Articles 6.13.3.2.2
through 6.13.3.2.4 of this Guide) and the connected material at the strength limit state (see the
Discussion of Article 6.13.5 in this Guide. The provisions also point to the Article for determining
the effective area of the weld (see the Discussion of Article 6.13.3.3 in this Guide).
The factored resistance of a welded connection is based on either the factored resistance of the
base metal, or the product of the deposited weld metal strength and the effective area of the weld
that resists the load. The weld metal strength is the capacity of the weld metal itself, typically given
in units of ksi. The effective area of the weld that resists the load is the product of the effective
length and the effective throat of the weld (see the Discussion of Article 6.13.3.3 in this Guide).

NSBA Guide to Navigating Routine Steel Bridge Design / 322


The factored resistance of the connected material is governed by the thickness of the connected
parts.
The classification strength of the weld metal, Fexx, is taken as the specified minimum tensile
strength of the weld metal in ksi, which is reflected in the classification designation of the
electrode. For example, the ′70′ in E70XX (SMAW), ER70S (solid-wire GMAW), and E70C
(metal-cored GMAW); and the ′7′ in E71XX (FCAW) and F7XX (SAW) in the classification
designation of the electrodes for the indicated welding processes indicate a specified minimum
tensile strength of 70.0 ksi.
These provisions are applicable to the simple span and multi-span continuous rolled beam bridges
covered by this Guide if welded connections are used for the bracing connections or for bearing
stiffeners (if required – see the Discussion of Article D6.5 in this Guide).
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on weld metal, welding processes, welded design details, and the appropriate designations of
welding symbols and consumables. Extensive guidance on the design of welded connections for
steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.
Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006 Guidelines for
Design Details and G12.1-2020 Guidelines to Design for Constructability and Fabrication for
recommendations, commentary, and sample welded design details that allow for more economical
fabrication and erection.

6.13.3.2.2 Complete Penetration Groove-Welded Connections

6.13.3.2.2a Tension and Compression


Determination of applicability, Simple Span Bridges: Applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
The provisions for welded connections in the AASHTO LRFD BDS are certainly applicable to the
design of routine steel I-girder bridges where welded steel plate girders are used as the main
spanning elements, since these girders necessarily use welds to connect the flanges to the webs, as
well as other details such as stiffeners and so on. The provisions may also be applicable to the
design of routine steel I-girder bridges where rolled steel beams are used as the main spanning
elements, if welded stiffeners or other welded connection details are involved with the diaphragms
or other features.
Complete joint penetration (CJP) groove welds are most often used to connect structural members
aligned in the same plane (i.e., butt joints such as flange and web shop splices). They can also be
used in tee and corner joints, although CJP groove welds are not recommended for use in these
joints because of the relatively high cost and resulting welding deformations in tee joints, and the
fact that backing bars must typically be left in place in corner joints. CJP groove welds have the

NSBA Guide to Navigating Routine Steel Bridge Design / 323


same resistance as the pieces joined and are intended to transmit the full load of the members
joined.
CJP groove welds may be single- or double-sided welds. Double-sided welds, which require access
to both sides of the joint, may require less weld metal and result in less distortion and are of
particular importance when joining thick members. Groove welds are classified according to their
shape. Most groove welds require a specific edge preparation and are named accordingly. The
selection of the proper groove weld is dependent on the cost of the edge preparations, the welding
process used, and the cost of making the weld. The decision as to which groove type to use is
usually left to the Fabricator/Detailer, who will select the type of groove that will generate the
required quality at a reasonable cost. The Engineer need only indicate on the contract drawings
that a CJP groove weld is required at a particular joint.
In groove welds, the maximum forces are usually tension or compression. According to the
provisions of this Article, the factored resistance, Rr, of CJP groove-welded connections subjected
to tension or compression normal to the effective area or parallel to the axis of the weld at the
strength limit state is to be taken as the factored resistance of the base metal. Tests have shown
that groove welds of the same thickness as the connected parts are adequate to develop the factored
resistance of the connected parts.
These provisions are not applicable to the simple span and multi-span continuous rolled beam
bridges covered by this Guide as CJP groove-welded connections (e.g., flange and web welded
butt (shop) splices) are not typically used in these bridges.
Consult the FHWA Bridge Welding Reference Manual is an excellent resource for additional
information on the types of welded connections and their design. Extensive guidance on the design
of welded connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference
Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006
Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample welded design details that allow for
more economical fabrication and erection.

6.13.3.2.2b Shear
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
Complete joint penetration (CJP) groove welds are most often used to connect structural members
aligned in the same plane (i.e., butt joints such as flange and web shop splices). They can also be
used in tee and corner joints, although CJP groove welds are not recommended for use in these
joints because of the relatively high cost and the resulting welding deformations in tee joints, and
the fact that backing bars must typically be left in place in corner joints. CJP groove welds have
the same resistance as the pieces joined and are intended to transmit the full load of the members
joined.

NSBA Guide to Navigating Routine Steel Bridge Design / 324


CJP groove welds may be single- or double-sided welds. Double-sided welds, which require access
to both sides of the joint, may require less weld metal and result in less distortion and are of
particular importance when joining thick members. Groove welds are classified according to their
shape. Most groove welds require a specific edge preparation and are named accordingly. The
selection of the proper groove weld is dependent on the cost of the edge preparations, the welding
process used, and the cost of making the weld. The decision as to which groove type to use is
usually left to the Fabricator/Detailer, who will select the type of groove that will generate the
required quality at a reasonable cost. The Engineer need only indicate on the contract drawings
that a CJP groove weld is required at a particular joint.
The provisions of this Article deal with the computation of the factored resistance, Rr, of CJP
groove-welded connections subjected to shear on the effective area of the weld. In groove welds,
the maximum forces are usually tension or compression. CJP groove-welded connections are
generally not subjected to shear in the routine steel I-girder bridges covered by this Guide.
However, some Owner-agency policies or standard details use CJP welds to connect bearing
stiffeners, that are also being used as cross-frame or diaphragm connection plates, to the girder
flange; in those cases, the provisions of this Article may be applicable. Generally, a detailed
evaluation of this type of CJP weld application is not necessary though, since the other connection
plates on the bridge should be connected to the flanges using fillet welds and those connections
would typically control. Note that the use of CJP welds to connect bearing stiffeners to flanges is
generally not required and is not recommended. A much more practical and economical detail is
to provide a finish-to-bear condition alone (if the bearing stiffener does not additionally serve as a
cross-frame or diaphragm connection plate) or a combination of finish-to-bear condition with fillet
wells (if the bearing stiffener does additionally serve as a cross-frame or diaphragm connetion
plate).
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on the types of welded connections and their design. Extensive guidance on the design of welded
connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006
Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample welded design details that allow for
more economical fabrication and erection.

6.13.3.2.3 Partial Penetration Groove-Welded Connections

6.13.3.2.3a Tension or Compression


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Partial joint penetration (PJP) groove welds do not extend completely through the thickness of the
pieces being joined and are subject to special design requirements. PJP groove welds are
sometimes used when stresses are low and there is no need to develop the complete strength of the
base material. PJP groove welds can be used to connect structural members aligned in the same

NSBA Guide to Navigating Routine Steel Bridge Design / 325


plane when the joints are subject to compression or shear only, provided adequate throats can be
developed. They can also be used in tee and corner joints, particularly as larger throats are required,
and are sometimes used in combination with fillet welds in these joints.
PJP groove welds may be single- or double-sided welds. Double-sided welds, which require access
to both sides of the joint, may require less weld metal and result in less distortion and are of
particular importance when joining thick members. Groove welds are classified according to their
shape. Most groove welds require a specific edge preparation and are named accordingly. The
selection of the proper groove weld is dependent on the cost of the edge preparations, the welding
process used, and the cost of making the weld. When designing PJP welds, engineers are
encouraged to select from the series of standard joints for PJP welds in Figure 2.5 of the BWC;
such joints can be used on the Weld Process Specification (WPS) without the need for further
testing to demonstrate the suitability of the joint itself. See Section 4.4, Standard Joints, in the
FHWA Bridge Welding Reference Manual for further discussion of standard joints.
In groove welds, the maximum forces are usually tension or compression. The provisions of this
Article deal with the computation of the factored resistance, Rr, of PJP groove-welded connections
subjected to tension or compression parallel to the axis of the weld, and connections subjected to
tension or compression normal to the effective area of the weld, at the strength limit state. Eq.
6.6.1.2.5-4 should also be considered in the fatigue design of PJP groove-welded connections
subject to tension normal to the effective area of the weld (see the Discussion of Article 6.6.1.2.5
in this Guide).
PJP groove-welded connections are not used, nor are they recommended for use, in the routine
steel I-girder bridges covered by this Guide; therefore, these provisions are not applicable. In many
cases, it is more feasible and economical to use fillet welds instead of PJP groove welds.
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on the types of welded connections and their design. Extensive guidance on the design of welded
connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006
Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample welded design details that allow for
more economical fabrication and erection.

6.13.3.2.3b Shear
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Partial joint penetration (PJP) groove welds do not extend completely through the thickness of the
pieces being joined and are subject to special design requirements. PJP groove welds are
sometimes used when stresses are low and there is no need to develop the complete strength of the
base material. PJP groove welds can be used to connect structural members aligned in the same
plane when the joints are subject to compression or shear only, provided adequate throats can be

NSBA Guide to Navigating Routine Steel Bridge Design / 326


developed. They can also be used in tee and corner joints, particularly as larger throats are required,
and are sometimes used in combination with fillet welds in these joints.
PJP groove welds may be single- or double-sided welds. Double-sided welds, which require access
to both sides of the joint, may require less weld metal and result in less distortion and are of
particular importance when joining thick members. Groove welds are classified according to their
shape. Most groove welds require a specific edge preparation and are named accordingly. The
selection of the proper groove weld is dependent on the cost of the edge preparations, the welding
process used, and the cost of making the weld. When designing PJP welds, engineers are
encouraged to select from the series of standard joints for PJP welds in Figure 2.5 of the BWC;
such joints can be used on the Weld Process Specification (WPS) without the need for further
testing to demonstrate the suitability of the joint itself. See Section 4.4, Standard Joints, in the
FHWA Bridge Welding Reference Manual for further discussion of standard joints.
The provisions of this Article deal with the computation of the factored resistance, Rr, of PJP
groove-welded connections subjected to shear parallel to the axis of the weld. PJP groove-welded
connections are not used, nor are they recommended for use, in the routine steel I-girder bridges
covered by this Guide; therefore, these provisions are not applicable. In many cases, it is more
feasible and economical to use fillet welds instead of PJP groove welds.
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on the types of welded connections and their design. Extensive guidance on the design of welded
connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006
Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample welded design details that allow for
more economical fabrication and erection.

6.13.3.2.4 Fillet-Welded Connections


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
Fillet welds are the most widely used welds due to their ease of fabrication and overall economy
and are certainly the most widely used welds for bracing member, connection plate, stiffener, and
flange-to-web connections; therefore, the provisions of this Article are applicable to the routine
steel I-girder bridges covered by this Guide.
Fillet welds generally require less precision during fit-up and the edges of the joined pieces seldom
need special preparation such as beveling or squaring. Fillet welds have a triangular cross-section
and do not fully fuse the cross-sectional area of the parts they join, although full-strength
connections can be developed with fillet welds. Lap joints utilize fillet welds and are the primary
joint type for bracing member connections (e.g., bracing member-to-gusset plate joints). They are
also used for flange-to-web welds in built-up sections and for connection plate and stiffener welds
to webs and flanges. Fillet welds can also be used in tee and corner joints subject to shear.

NSBA Guide to Navigating Routine Steel Bridge Design / 327


The factored resistance of fillet welds is based on the assumption that failure of such welds is by
shear on the effective area whether the shear transfer is parallel or perpendicular to the axis of the
line of the weld. In fact, the resistance is greater for shear transfer perpendicular to the weld axis;
however, for simplicity the situations are treated the same. Therefore, the factored resistance of
fillet welds may be controlled by the shear resistance of the weld metal or by the shear rupture
resistance of the connected material. Shear yielding is not critical in welds because the material
strain hardens without large overall deformations occurring.
According to the provisions of this Article, the factored resistance, Rr, of fillet welds which are
made with matched or undermatched weld metal and which have typical weld profiles at the
strength limit state is to be taken as the smaller of the factored shear rupture resistance of the
connected material adjacent to the weld leg determined as specified in Article 6.13.5.3 (see the
Discussion of Article 6.13.5.3 in this Guide), and the product of the effective area specified in
Article 6.13.3.3 (see the Discussion of Article 6.13.3.3 in this Guide) and the factored shear
resistance of the weld metal given by Eq. 6.13.3.2.4-1, which depends on the classification
strength, Fexx, of the weld metal (see the Discussion of Article 6.13.3.2.1 in this Guide). The
factored shear rupture resistance of the base metal adjacent to the weld leg will seldom control
since the effective area of the base metal at the weld leg is typically about 30 percent greater than
the weld throat.
If fillet welds are subjected to eccentric loads that produce a combination of shear and bending
stresses, they should be proportioned on the basis of a direct vector addition of the shear forces on
the weld (consult Section 6.6.4.3.7.2.3 of the Reference Manual for NHI Course 130081, Load and
Resistance Factor Design (LRFD) for Highway Bridge Superstructures). Most of the welded
connection details commonly used in routine steel I-girder bridges feature direct load paths which
do not include significant eccentricity. An example of an unusual situation where a welded
connection in a routine steel I-girder bridge might be subject to eccentric loading is the case of a
cross-frame gusset plate welded to a cross-frame connection plate (stiffener). Although this type
of connection detail is uncommon, at least one large Owner-agency prefers its use.
Also, see the Discussion of Article 6.13.1 in this Guide for a discussion of evaluating welded end
connections for unbalanced weld conditions. Fillet-welded end connections of cross-frame
members in routine steel I-girder bridges may exhibit unbalanced weld conditions; if so, the effects
of the unbalanced geometry should be considered in the design of the connection.
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on the types of welded connections and their design. Extensive guidance on the design of welded
connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006
Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample welded design details that allow for
more economical fabrication and erection.
For specific design examples of connections utilizing fillet welds, consult Section 6.6.4.3.7.2 of
the Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures.

NSBA Guide to Navigating Routine Steel Bridge Design / 328


6.13.3.3 Effective Area
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The factored resistance of welds is based on the effective area of the weld, defined in this Article
as the effective length of the weld times the effective throat. The effective throat is defined
nominally as the shortest distance from the joint root to the face of the weld neglecting any weld
reinforcement, or the minimum width of the expected failure plane.
The effective length of a groove weld is the width of the part joined perpendicular to the direction
of stress. By definition, the effective throat of a CJP groove weld is equal to the thickness of the
thinner part joined, with no increase allowed for any weld reinforcement. To achieve fusion
throughout the thickness of the part being joined, backing is usually required if the CJP groove
weld is made from one side, and back gouging is usually required from the second side if the CJP
groove weld is made from both sides. Otherwise, qualification testing is required to show that the
full throat can be developed.
The effective throat of PJP groove welds is defined in Article 2.3 of the AASHTO/AWS
D1.5M/D1.5 Bridge Welding Code (BWC). The effective throat of PJP groove welds depends on
the probable depth of fusion that will be achieved; that is, the depth of groove preparation and
depth of penetration that can be achieved by the selected welding process and welding position.
Minimum effective throat thickness requirements for PJP groove welds are also given in the BWC.
PJP groove-welded connections are not used, nor are they recommended for use, in the routine
steel I-girder bridges covered by this Guide.
The effective length of a fillet weld is to be taken as the overall length of the full-size fillet. The
effective throat dimension of a fillet weld for a typical fillet weld with equal legs of nominal size,
a, is taken equal to 0.707a, or nominally the shortest distance from the joint root to the weld face,
neglecting any reinforcement.
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on the types of welded connections and their design. Extensive guidance on the design of welded
connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006
Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample design details that allow for more
economical fabrication and erection.
For specific design examples of connections utilizing fillet welds, consult Section 6.6.4.3.7.2 of
the Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures.

6.13.3.4 Size of Fillet Welds


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 329


Discussion:
The provisions of this Article state that the size of a fillet weld that may be assumed in
the design of a connection is to be such that the forces due to the factored loadings do not exceed
the factored resistance, Rr, of the connection specified in Article 6.13.3.2 (see the Discussion of
Article 6.13.3.2 in this Guide). The size of a fillet weld is given as the leg size of the fillet.
These provisions also specify the maximum and minimum size requirements for fillet welds.
Maximum thickness (size) requirements for fillet welds along edges of connected parts depend on
the thickness of the parts being connected, unless the weld is specifically designated on the contract
documents to be built out to obtain full throat thickness. The requirements prevent melting of the
base metal where the fillet would meet the corner of the plate if the fillet were made the full plate
thickness.
The minimum thickness (size) of a fillet weld is not to be less than that required to transmit the
required forces, nor the minimum thickness specified in Table 6.13.3.4-1. The minimum weld size
need not exceed the thickness of the thinner part joined. Note that the specified minimum weld
sizes assume that the required preheats and interpass temperatures are provided (consult the
FHWA Bridge Welding Reference Manual for additional information on preheats and interpass
temperatures). Smaller welds than the minimum size welds may be approved by the Design
Engineer if they are shown to be adequate for the applied stress and if the appropriate additional
preheat is applied.
Minimum thickness requirements for fillet welds are based on preventing too rapid a rate of
cooling in order to prevent a loss of ductility (i.e., the formation of a brittle microstructure) or a
lack of fusion. The thicker the plate joined, the faster the heat is removed from the welding area.
As a minimum, a weld of sufficient size is needed to prevent the thicker plate from removing heat
at a faster rate than it is being supplied to cause the base metal to become molten. Thus, the
minimum weld sizes implicitly imply a specified minimum heat input. In addition, restraint to
weld metal shrinkage may result in weld cracking if the welds are too small. Minimum weld sizes
are frequently used for the case of longitudinal fillet welds that resist shear (e.g., girder flange-to-
web welds). Reducing the amount of weld metal will decrease the amount of distortion in welded
assemblies; thus, the smallest acceptable weld size that will provide the required factored
resistance should be used.
Since the minimum size requirements for fillet welds imply a minimum level of heat input, the
minimum size welds must be made in a single pass, as multiple passes to make the minimum size
weld would not provide the assumed minimum level of heat input, essentially defeating the
purpose of the requirement. The largest single-pass fillet weld that can be made with the manual
SMAW process is typically 5/16 in. Single-pass fillet welds up to about ½ in. can be made with
the SAW process.
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on the types of welded connections and their design. Extensive guidance on the design of welded
connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006

NSBA Guide to Navigating Routine Steel Bridge Design / 330


Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample design details that allow for more
economical fabrication and erection.
For specific design examples of connections utilizing fillet welds, consult Section 6.6.4.3.7.2 of
the Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures.

6.13.3.5 Minimum Effective Length of Fillet Welds


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article specify the minimum effective length of fillet welds. When placing
a fillet weld, the welder builds up the weld to the full dimension as near to the beginning of the
weld as possible. However, there is always a slight tapering off of the weld where the weld starts
and ends. Therefore, a minimum effective length of the weld is required. As specified in this
Article, the minimum effective length of a fillet weld is to be taken as four times its leg size, but
not less than 1.5 inches.
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on the types of welded connections and their design. Extensive guidance on the design of welded
connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006
Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample design details that allow for more
economical fabrication and erection.
For specific design examples of connections utilizing fillet welds, consult Section 6.6.4.3.7.2 of
the Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures.

6.13.3.6 Fillet Weld End Returns


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article deal with end returns on fillet-welded connections.
Fillet welds that resist a tensile force not parallel to the axis of the weld, or are not proportioned to
withstand repeated stress, are not to terminate at corners of parts or members. Where such returns
can be made in the same plane, they are to be returned continuously, full size, around the corner,
for a length equal to twice the weld size and are to be indicated in the contract documents. Also,
fillet welds deposited on the opposite sides of a common plane of contact between two parts are to
be interrupted at a corner common to both welds.

NSBA Guide to Navigating Routine Steel Bridge Design / 331


These situations are not typically encountered in the fillet-welded connections used in the routine
steel I-girder bridges covered by this Guide. An example of this condition might be the welded
connection of a cross-frame member to a gusset plate if it were proposed to only provide a weld
along the end of the member; such a weld would be prohibited by this Article unless returns were
provided, effectively resulting in a three-sided connection (the end weld, then the weld wrapped
around to the sides of the member). However, such a detail would likely be subject to numerous
other design problems; a more robust three-sided welded connection with longer welds along the
sides of the member would be a better approach and is in fact the much more common detail used
in this situation.
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on the types of welded connections and their design. Extensive guidance on the design of welded
connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006
Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample design details that allow for more
economical fabrication and erection.

6.13.3.7 Fillet Welds for Sealing


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article cover fillet welds to be used for sealing of connections to help
prevent penetration of moisture.
Seal welds should be continuous welds combining the functions of sealing and strength. The
portion of a return sealing fillet weld wrapped around the ends of a transverse, longitudinal or
bearing stiffener, connection plate, or a lap splice connection, if permitted by the Owner-agency,
is to be exempt from the minimum size requirements specified in Article 6.13.3.4 (see the
Discussion of Article 6.13.3.4 in this Guide), and is not to be considered in determining the
resistance of the connection. Research has shown that when wrapping the fillet weld around the
ends of stiffeners or connection plates, the undercutting of the corner of the stiffener or connection
plate that occurs, even when severe, does not reduce the fatigue performance of the weld, which
is controlled by the toe of the transverse fillet weld connecting the stiffener or connection plate to
the flange.
As described in the Commentary for this Article, the ends of fillet-welded connections in
galvanized structures, in particular, should be sealed to prevent the acids used to prepare the steel
for galvanizing from being trapped in-between the components and then leaching out. Vent holes
to allow the trapped air and moisture to escape and prevent destructive pressures from developing
between the surfaces may be required if the overlap area exceeds 16.0 in.2 The recommended size
of the vent holes in such cases is given in Tables 1 and 2 of the ASTM A385 specification.
The FHWA Bridge Welding Reference Manual is an excellent resource for additional information
on the types of welded connections and their design. Extensive guidance on the design of welded

NSBA Guide to Navigating Routine Steel Bridge Design / 332


connections for steel girder bridges can be found in Section 6.6.4.3 of the Reference Manual for
NHI Course 130081, Load and Resistance Factor Design (LRFD) for Highway Bridge
Superstructures. Consult the AASHTO-NSBA Steel Bridge Collaboration Guidelines G1.4-2006
Guidelines for Design Details and G12.1-2020 Guidelines to Design for Constructability and
Fabrication for recommendations, commentary, and sample design details that allow for more
economical fabrication and erection.

6.13.4 Block Shear Rupture Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
A connected element subject to tension must be checked for a tearing limit state known as block
shear rupture. This Article specifies that block shear rupture is to be checked for the web
connection of coped beams and for tension connections, including connection plates, splice plates
and gusset plates. Tests on coped beams have indicated that a block shear failure can occur around
the perimeter of the bolt holes. Tension member connections are also susceptible, and the block
shear rupture mode should also be checked around the periphery of welded connections.
The factored tensile resistance of the member, Pr (see the Discussion of Article 6.8.2.1 in this
Guide), must not exceed the factored block shear rupture resistance, Rr, of the connected elements,
which must be investigated at the end connections. The connections are to be investigated by
considering all possible failure planes in the connected elements, including those parallel and
perpendicular to the applied forces, and determining the most critical set of planes. For a bolted
connection, the failure path is defined by the centerlines of the bolt holes. Planes parallel to the
applied force are to be considered to resist only shear stresses and planes perpendicular to the
applied force are to be considered to resist only tensile stresses.
The factored block shear rupture resistance, Rr, is determined from Eq. 6.13.4-1. In determining
Rr, the resistance to rupture along the shear plane is added to the resistance to rupture on the tensile
plane. Block shear rupture is a rupture or tearing phenomenon and not a yielding phenomenon.
However, gross yielding along the shear place can occur when tearing on the tensile plane
commences if 0.58FuAvn exceeds 0.58FyAvg. Therefore, Eq. 6.13.4-1 limits 0.58FuAvn to not exceed
0.58FyAvg.
The reduction factor, Ubs, has been included in Eq. 6.13.4-1 to approximate the effect of a non-
uniform stress distribution on the tensile plane in certain cases; e.g., coped beam connections with
multiple rows of bolts. In such cases, the tensile stress on the end plane is non-uniform because
the rows of bolts nearest the beam end pick up most of the shear. For the majority of connections
encountered in the routine steel I-girder bridges covered by this Guide, Ubs will equal 1.0.
The reduction factor, Rp, in Eq. 6.13.4-1 conservatively accounts for the reduced rupture resistance
in the vicinity of bolt holes punched full size (see the Discussion of Article 6.8.2.1 in this Guide).
Article 6.6.1.2.3 specifies that unless information is available to the contrary, bolt holes in bracing
members and their connection plates are to be assumed for design to be punched full size (see the
Discussion of Article 6.6.1.2.3 in this Guide). Bracing member connections are often punched full
size, whereas bolt holes in splice connections are typically drilled full size, but unless this is

NSBA Guide to Navigating Routine Steel Bridge Design / 333


explicitly directed in the Owner-agency’s specifications, the Engineer should assume the holes are
punched full size.
In determining the net area of cuts carrying shear stress in bolted connections with staggered holes,
the full effective diameter of the staggered holes centered within two hole diameters of the cut is
to be deducted; holes further removed are to be disregarded. In determining the net area of cuts
carrying tension stress, the effect of staggered holes adjacent to the cuts is to be determined using
the s2/4g correction described in Article 6.8.3 (see the Discussion of Article 6.8.3 in this Guide).
Block shear rupture is most likely to control in the design of bolted end connections to thin webs
of girders (e.g., coped beams) and in the design of short compact bolted connections. Although it
must also be checked for splice connections, it is unlikely to control in the design of bolted flange
and web splices of typical proportions (see the Discussion of Article 6.13.6 in this Guide).
For further information on the block shear rupture resistance and design examples illustrating the
block shear rupture resistance checks, consult Sections 6.6.3.3.2.5 and 6.6.4.2.5.6.1 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures.

6.13.5 Connection Elements

6.13.5.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This Article indicates that the provisions of Article 6.13.5 are to be applied to the design of
connection elements such as splice plates, gusset plates, corner angles, brackets, and lateral
connection plates in tension or shear, as applicable. For the routine steel I-girder bridges covered
by this Guide, these provisions are to be applied to the design of splice plates and cross-frame
gusset plates only, as applicable.

6.13.5.2 Tension
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies that the factored tensile resistance, Rr, of a connected element is to be taken
as the smallest of the resistances based on yielding, net section fracture or block shear rupture. For
the routine steel I-girder bridges covered by this Guide, these provisions are to be applied to
determine the factored tensile resistance of flange splice plates and cross-frame gusset plates, as
applicable.
A connected element subject to tension must be checked for yielding on the gross section.
Excessive elongation due to uncontrolled yielding of the gross area can limit the structural
usefulness of the connected element so that it no longer serves its intended purpose. The factored
yield resistance, Rr, of a connected element in tension is to be computed from Eq. 6.8.2.1-1 (see
the Discussion of Article 6.8.2.1 in this Guide).

NSBA Guide to Navigating Routine Steel Bridge Design / 334


A connected element subject to tension must also be checked for fracture on the net section. The
connected element can fracture by failure of the net area at a load smaller than that required to
yield the gross area depending on the ratio of net to gross area, the properties of the steel (i.e., the
ratio of Fu/Fy), and the end connection geometry. Holes in a member cause stress concentrations
at service loads, with the tensile stress adjacent to the hole typically about three times the average
stress on the net area. As the load increases and the deformation continues, all fibers across the
section will achieve or eventually exceed the yield strain. Failure occurs when the localized
yielding results in a fracture through the net area. Typically, a higher margin of safety is used when
considering the net section fracture resistance versus the yield resistance. The factored net section
fracture resistance, Rr, of a connected element in tension is to be computed from Eq. 6.8.2.1-2 (see
the Discussion of Article 6.8.2.1 in this Guide).
The calculation of the net area, An, in Eq. 6.8.2.1-2 is discussed in Article 6.8.3 (see the Discussion
of Article 6.8.3 in this Guide). According to the provisions of this Article, for flange splice plates
and cross-frame gusset plates, An is not to be taken greater than 85 percent of the gross area, Ag, of
the plate in checking Eq. 6.8.2.1-2. Should An equal or exceed 0.85Ag, then 0.85Ag is substituted
for An when checking Eq. 6.8.2.1-2. Because the length of these elements is small compared to the
overall member length, inelastic deformation of the gross area is limited. Tests have shown that
when holes are present in such short elements where general yielding on the gross section cannot
occur, there will be at least a 15 percent reduction in tensile capacity from that obtained based on
yielding of the gross section.
The reduction factor, U, in Eq. 6.8.2.1-2 accounts for the effect of shear lag in connections (see
the Discussion of Article 6.8.2.2 in this Guide). According to the provisions of this Article, for
short connection elements such as flange splice plates and cross-frame gusset plates, where the
elements of the cross-section essentially lie in a common plane and are connected by bolts or
welds, U is to be taken equal to 1.0 (except for the rare case where Case 4 in Table 6.8.2.2-1
applies; i.e., when gusset plates are attached to cross-frame connection plates using only
longitudinal welds along the length of the connection with no transverse weld across the end of
the connection – see the Discussion of Article 6.8.2.2 in this Guide).
The reduction factor, Rp, in Eq. 6.8.2.1-2 conservatively accounts for the reduced rupture resistance
in the vicinity of bolt holes punched full size (see the Discussion of Article 6.8.2.1 in this Guide).
Article 6.6.1.2.3 specifies that unless information is available to the contrary, bolt holes in
connection plates are to be assumed for design to be punched full size (see the Discussion of Article
6.6.1.2.3 in this Guide). Bolt holes in flange splice plates are typically drilled full size, whereas
bolt holes in cross-frame gusset plates are often punched full size, but unless this is explicitly
directed in the Owner-agency’s specifications, the Engineer should assume the holes are punched
full size.
Lastly, a connected element subject to tension must also be checked for block shear rupture. The
factored block shear rupture resistance, Rr, of the connected element is calculated according to Eq.
6.13.4-1 (see the Discussion of Article 6.13.4 in this Guide).
For further information on the factored tensile resistance of connection elements and design
examples illustrating the factored tensile resistance checks, consult Section 6.6.4.2.5.6.1 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for

NSBA Guide to Navigating Routine Steel Bridge Design / 335


Highway Bridge Superstructures. NSBA’s Bolted Field Splices for Steel Bridge Flexural Members
– Overview and Design Examples also provides illustrations of these calculations in the context
of bolted field splice design.

6.13.5.3 Shear
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
This Article specifies that the factored shear resistance, Rr, of a connected element is to be taken
as the smallest of the resistances based on shear yielding or shear rupture. For the routine steel I-
girder bridges covered by this Guide, these provisions are to be applied to determine the factored
shear resistance of web splice plates and cross-frame gusset plates, as applicable.
For shear yielding, the factored shear resistance of the connected element, Rr, given by Eq.
6.13.5.3-1 is conservatively based on the shear yield stress (i.e., Fy 3 = 0.58Fy).
For shear rupture, the factored shear resistance of the connected element, Rr, is given by Eq.
6.13.5.3-2. A higher margin of safety is used when considering the shear rupture resistance versus
the shear yield resistance.
The reduction factor, Rp, in Eq. 6.13.5.3-2 conservatively accounts for the reduced rupture
resistance in the vicinity of bolt holes punched full size (see the Discussion of Article 6.8.2.1 in
this Guide). Article 6.6.1.2.3 specifies that unless information is available to the contrary, bolt
holes in connection plates are to be assumed for design to be punched full size (see the Discussion
of Article 6.6.1.2.3 in this Guide). Bolt holes in web splice plates are typically drilled full size,
whereas bolt holes in cross-frame gusset plates are often punched full size, but unless this is
explicitly directed in the Owner-agency’s specifications, the Engineer should assume the holes are
punched full size.
For further information on the factored shear resistance of connection elements and design
examples illustrating the factored shear resistance checks, consult Section 6.6.4.2.5.6.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures.

6.13.6 Splices

6.13.6.1 Bolted Splices

6.13.6.1.1 Tension Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A splice is defined as a group of bolted connections (or a welded connection) sufficient to transfer
the moment, shear, axial force or torque between two structural elements joined at their ends to
form a single, longer element. Bolted splices are typically used to connect member sections

NSBA Guide to Navigating Routine Steel Bridge Design / 336


together in the field; hence, the term “field splice” is often used. The provisions of this Article
cover the design of bolted splices for members subject to axial tension.
Bolted splices for tension members are to be designed using slip-critical connections (see the
Discussion of Article 6.13.2.1.1 in this Guide), and are to satisfy the tensile resistance requirements
for connected elements specified in Article 6.13.5.2 (see the Discussion of Article 6.13.5.2 in this
Guide). The splices are to be designed at the strength limit state for the load as determined by the
requirements of Article 6.13.1 for splices of primary members subject to axial tension (see the
Discussion of Article 6.13.1 in this Guide). Where the section changes at the splice, the smaller of
the two connected sections is to be used in the design.
These provisions are not applicable to the routine steel I-girder bridges covered by this Guide as
the members in these bridges which may be subject to axial tension (such as cross-frame members)
are not typically spliced, nor should they be.
Provisions for the design of bolted field splices are addressed in Article 6.13.6.1.3, Flexural
Members, and its associated sub-Articles.
For further information on field splice design, consult Section 6.6.5 and especially 6.6.5.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures (note that it refers to the previous AASHTO provisions for bolted
field splice design). For in-depth information on the current AASHTO provisions for the design
of bolted field splices for flexural members and examples illustrating the design of such splices,
consult NSBA’s Bolted Field Splices for Steel Bridge Flexural Members – Overview and Design
Examples.

6.13.6.1.2 Compression Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
A splice is defined as a group of bolted connections (or a welded connection) sufficient to transfer
the moment, shear, axial force or torque between two structural elements joined at their ends to
form a single, longer element. Bolted splices are typically used to connect member sections
together in the field; hence, the term “field splice” is often used. The provisions of this Article
cover the design of bolted splices for members subject to axial compression (e.g., arch members,
truss chords, and columns).
Splices for compression members may either be designed at the strength limit state as: 1) open
joints (i.e., no contact between adjoining parts) with enough bolts provided in the splice to carry
100 percent of the load as determined by the requirements of Article 6.13.1 for splices of primary
members subject to axial compression (see the Discussion of Article 6.13.1 in this Guide), or 2)
milled joints in full contact bearing with the bolts designed to carry no less than 50 percent of the
lower factored resistance of the sections spliced. If the latter option is chosen, Article 6.13.6.1.2
requires that the contract documents call for inspection of the joint during fabrication and erection.
Fabricators generally prefer the first option because it is less expensive and has the potential for
fewer problems in the field.

NSBA Guide to Navigating Routine Steel Bridge Design / 337


The splices in these members are to be located as near as practicable to the panel points and usually
on the side of the panel point where the smaller force effect occurs. The arrangement of splice
elements must make provision for the various force effects in the component parts of the spliced
members.
These provisions are not applicable to the routine steel I-girder bridges covered by this Guide as
the members in these bridges which may be subject to axial compression (such as cross-frame
members) are not typically spliced, nor should they be.
Provisions for the design of bolted field splices are addressed in Article 6.13.6.1.3, Flexural
Members, and its associated sub-Articles.
For further information on field splice design, consult Section 6.6.5 and especially 6.6.5.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures (note that it refers to the previous AASHTO provisions for bolted
field splice design). For in-depth information on the current AASHTO provisions for the design
of bolted field splices for flexural members and examples illustrating the design of such splices,
consult NSBA’s Bolted Field Splices for Steel Bridge Flexural Members – Overview and Design
Examples.

6.13.6.1.3 Flexural Members

6.13.6.1.3a General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
A splice is defined as a group of bolted connections (or a welded connection) sufficient to transfer
the moment, shear, axial force or torque between two structural elements joined at their ends to
form a single, longer element. Bolted splices are typically used to connect member sections
together in the field; hence, the term “field splice” is often used. The provisions of this Article
cover general provisions for the design of bolted field splices for members subject to flexure, and
hence, are applicable to the routine steel I-girder bridges covered by this Guide.
Bolted beam or girder field splices generally include top flange splice plates, web splice plates and
bottom flange splice plates. In addition, if the plate thicknesses on one side of the joint are different
than those on the other side, filler plates are used to match the thicknesses within the splice (see
the Discussion of Article 6.13.6.1.4 in this Guide). For the flange splice plates, there is typically
one plate on the outside of the flange and two smaller plates on the inside of the flange; one on
each side of the web. For the web splice plates, there are two plates; one on each side of the web,
with at least two rows of high-strength bolts over the depth of the web used to connect the splice
plates to the member.
As required by Articles 6.13.6.1.3b and 6.13.6.1.3c, bolted flange and web splice connections are
designed at a minimum for 100 percent of the individual design resistances of the flange and web;
that is, the individual flange splices are designed for the smaller design yield resistance of the
corresponding flanges on either side of the splice (see the Discussion of Article 6.13.6.1.3b in this
Guide), and the web splice is designed for the smaller factored shear resistance of the web on either

NSBA Guide to Navigating Routine Steel Bridge Design / 338


side of the splice (see the Discussion of Article 6.13.6.1.3c in this Guide). Additional forces in the
web connection may need to be considered if the flanges are not adequate to develop the factored
design moment at the point of splice.
Bolted splices in continuous spans should be made in regions of lower moment at or near points
of permanent load contraflexure (inflection points) to reduce the size of the splice. This may not
always be possible in certain situations, such as in longer-span bridges or in cases where additional
field splices may be needed to reduce the size of a shipping piece to practical limits to better
accommodate shipping (e.g., shipping of a sharply curved member).
In cases where bolted splices are located away from points of permanent load contraflexure, the
Engineer should check the girder flanges subject to tension for net section fracture since the flanges
will have holes for the splice plate bolts. Eq. 6.10.1.8-1 provides a limit on the maximum factored
major-axis bending stress permitted on the gross section of the girder, neglecting the loss of area
due to holes in the tension flange at the bolted splice (see the Discussion of Article 6.10.1.8 in this
Guide). Eq. 6.10.1.8-1 will prevent a bolted splice from being located at a section where the
factored flexural resistance of the section at the strength limit state exceeds the moment at first
yield, My, unless the factored stress in the tension flange at that section is limited to the value given
by the equation.
Splices located in areas of stress reversal near points of permanent load contraflexure are to be
investigated for both positive and negative flexure to determine the governing condition. Web and
flange splices must have at least two rows of bolts on each side of the joint to facilitate proper
alignment and stability of the girder during construction. Also, oversize or slotted holes are not to
be used in either the member or the splice plates at bolted splices to provide geometry control
during erection before the bolts are tightened. Bolted splice connections for flexural members are
to be designed as slip-critical connections (see the Discussion of Article 6.13.2.1.1 in this Guide).
Bolted connections for flange and web splices are to be proportioned to prevent slip under Load
Combination Service II and during the casting of the concrete deck to provide geometry control.
The nominal fatigue resistance of base metal at the gross section adjacent to slip-critical bolted
connections is based on fatigue detail Category B assuming the bolts are installed in holes drilled
full size or subpunched and reamed to size (Table 6.6.1.2.3-1 – Condition 2.1), which is required
for bolted beam or girder splices. However, fatigue will not control the design of the bolted splice
plates for flexural members. The combined areas of the flange and web splice plates typically will
exceed the areas of the smaller flanges and web to which they are attached, and the flanges and
web are usually checked separately for either equivalent or more critical fatigue category details.
Therefore, fatigue of the splice plates will not need to be checked.
For further information on field splice design, consult Section 6.6.5 and especially 6.6.5.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures – note that this Reference Manual refers to the previous
AASHTO provisions for bolted field splice design. The provisions addressing the design of bolted
field splices changed significantly in 2017 with the publication of the 8th Edition of the AASHTO
LRFD BDS. For in-depth information on the current AASHTO provisions for the design of bolted
field splices for flexural members and examples illustrating the design of such splices, consult

NSBA Guide to Navigating Routine Steel Bridge Design / 339


NSBA’s Bolted Field Splices for Steel Bridge Flexural Members – Overview and Design
Examples.
In addition, NSBA’s NSBA Splice Microsoft Excel-based bolted field splice design spreadsheet
takes the time-consuming task of designing and checking a bolted splice connection and rewrites
the process with a simple input page and output form. NSBA Splice can be incorporated as a
design tool on plate girder bridges allowing the designer to quickly analyze various bolted splice
connections to determine the most efficient bolt quantity and configuration. NSBA Splice allows
the user to explore the effects of bolt spacing, bolt size, strength, and connection dimensions on
the overall splice design.

6.13.6.1.3b Flange Splices


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article cover the design of flange splice plates and their connections for
bolted field splices in flexural members and are applicable to the routine steel I-girder bridges
covered by this Guide.
At the strength limit state, each flange splice must develop the smaller full design yield resistance
of the flanges, Pfy, on either side of the splice. Pfy of each flange based on the effective area of the
flange, Ae, is calculated using Eq. 6.13.6.1.3b-1. Ae is calculated using Eq. 6.13.6.1.3b-2 and cannot
exceed the gross area of the flange, Ag. In most cases, Ae will be less than Ag. Ae is used rather than
Ag to account for the loss in section causing a reduction in the fracture resistance of the net section
at the connection for loading conditions in which the flange is subject to tension; that is, the flange
splices are designed for a lower design force such that net section fracture will not occur at the
connection. Ae only applies to tension flanges but is conservatively applied to both flanges in the
design method since each flange may potentially be subject to tension and the resulting design
force will conservatively provide a lower moment resistance of the flanges at the strength limit
state, as described further below.
The moment resistance provided by the flanges (neglecting the web and considering only the
flange force) is next checked against the factored moment at the strength limit state at the point of
splice. Should the factored moment exceed the moment resistance provided by the flanges, the
additional moment is to be resisted by the web (see the Discussion of Article 6.13.6.1.3c in this
Guide). For composite sections subject to positive flexure, the moment resistance provided by the
flanges is computed as shown in Figure C6.13.6.1.3b-1. For composite sections subject to negative
flexure and for noncomposite sections subject to positive or negative flexure, the moment
resistance provided by the flanges is computed as shown in Figure C6.13.6.1.3b-2. The moment
resistance provided by the flanges can potentially be increased by staggering the flange bolts.
The number of bolts required on one side of the flange splice at the strength limit state is found by
dividing Pfy by the factored shear resistance of the bolts (see the Discussion of Articles 6.13.2.2
and 6.13.2.7 in this Guide), including the reduction in the shear resistance of the bolts due to any
needed fillers (see the Discussion of Article 6.13.6.1.4 in this Guide). The threads typically are
excluded in flange splices (see the Discussion of Article 6.13.2.7 in this Guide for the conditions

NSBA Guide to Navigating Routine Steel Bridge Design / 340


under which the threads in flange splices may be assumed excluded from the shear planes). The
number of shear planes, Ns, is determined as described further below. For a bolt in a lap-splice
tension connection greater than 38.0 in. in length, the nominal shear resistance, Rn, is reduced by
a factor of 0.83 (see the Discussion of Article 6.13.2.7 in this Guide). For bolted flange splices,
the 38.0 in. length is measured between the extreme bolts on only one side of the splice and is
normally not exceeded.

The bearing resistance of the flange splice bolt holes is to be checked at the strength limit state
(see the Discussion of Article 6.13.2.9 in this Guide). The bearing resistance of the connection is
taken as the sum of the smaller of the shear resistance of the individual bolts and the bearing
resistance of the individual bolt holes parallel to the line of the design force. If the bearing
resistance of a bolt hole exceeds the shear resistance of the bolt, the bolt resistance is limited to
the shear resistance. Assuming the sum of the flange splice-plate thicknesses exceeds the thickness
of the thinner flange at the point of splice, and the splice plate areas satisfy the 10 percent rule
described below, the bearing resistance of the connection will be governed by the flange on either
side of the splice with the smaller product of the thickness and specified minimum tensile strength,
Fu, of the flange. Otherwise, the bearing resistance of each individual component should be
checked to determine the component governing the bearing resistance of the connection.
At the strength limit state, Pfy may be assumed equally divided to the inner and outer flange splice
plates when the areas of the inner and outer plates do not differ by more than 10 percent. In this
case, the shear resistance of the bolted connection should be checked for Pfy acting in double shear
(i.e., Ns = 2). Should the areas of the inner and outer splice plates differ by more than 10 percent,
the force in each plate should be determined by multiplying Pfy by the ratio of the area of the splice
plate under consideration to the total area of the inner and outer plates. In this case, the shear
resistance of the bolted connection should be checked for the larger of the calculated splice-plate
forces acting on a single shear plane (i.e., Ns = 1). The width of the outside splice plate should be
at least as wide as the width of the narrowest flange at the splice. The thickness of the outside
splice plate should be at least one-half the thickness of the thinner flange at the splice plus 1/16 of
an inch. The width of the inner splice plates should be chosen to allow a clearance distance of at
least 1/8-inch between the edge of each splice plate and the adjacent flange-to-web weld.
The design force in flange splice plates subject to tension at the strength limit state is not to exceed
the factored resistance of the splice plates in tension; that is, the splice plates are to be checked for
yielding on the gross section, fracture on the net section, and for block shear rupture (see the
Discussion of Articles 6.13.5.2 and 6.13.4 in this Guide). Block shear rupture will not typically
control the design of flange splice plates of typical proportion. Furthermore, the design net area of
the splice plates, An, must not exceed 0.85Ag, where Ag is the gross area of the splice plates. Should
An equal or exceed 0.85Ag, then 0.85Ag is substituted for An when checking fracture on the net
section of the splice plates; otherwise, An is used. The factors, U, Rp, and Ubs, are to be taken equal
to 1.0 for splice plates in the net section fracture and block shear rupture checks. The factored yield
resistance of the splice plates in compression, Rr, is the same as the factored yield resistance of the
splice plates in tension, and therefore, need not be checked. Buckling of the splice plates in
compression is not a concern since the unsupported length of the plates is limited by the maximum

NSBA Guide to Navigating Routine Steel Bridge Design / 341


bolt spacing and end distance requirements (see the Discussion of Articles 6.13.2.6.2 and
6.13.2.6.5 in this Guide).
The moment resistance provided by the nominal slip resistance of the flange splice bolts (see the
Discussion of Article 6.13.2.8 in this Guide) is to also be checked against the factored moment for
checking slip. The factored moments for checking slip are taken as the moment at the point of
splice under Load Combination Service II (see the Discussion of Article 3.4.1 in this Guide), and
the factored moment at the point of splice due to the deck casting sequence (see the Discussion of
Article 3.4.2.1 in this Guide). When checking the slip resistance of the bolts for a typical flange
splice with inner and outer splice plates, the flange slip force is assumed divided equally to the two
slip planes regardless of the ratio of the splice plate areas (i.e., Ns = 2). Unless slip occurs on both
planes, slip of the connection cannot occur. The moment resistance provided by the nominal slip
resistance of the flange splice bolts is computed as described above for the strength limit state,
with the nominal slip resistance of the bolts substituted for Pfy. For checking slip due to the factored
deck casting moment, the moment resistance of the noncomposite section is used. Should the
flange bolts not be sufficient to resist the factored moment for checking slip, the additional moment
is to be resisted by the web (see the Discussion of Article 6.13.6.1.3c in this Guide). Should the
web bolts not be sufficient to resist the additional moment, only then should consideration be given
to adding additional bolts to the flange splices.
The portions of this Article and the associated Commentary dealing with box sections and
horizontally curved girders are not applicable to the routine steel I-girder bridges covered by this
Guide.
For further information on field splice design, consult Section 6.6.5 and especially 6.6.5.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures – note that this Reference Manual refers to the previous
AASHTO provisions for bolted field splice design. The provisions addressing the design of bolted
field splices changed significantly in 2017 with the publication of the 8th Edition of the AASHTO
LRFD BDS. For in-depth information on the current AASHTO provisions for the design of bolted
field splices for flexural members and examples illustrating the design of such splices, consult
NSBA’s Bolted Field Splices for Steel Bridge Flexural Members – Overview and Design
Examples.
In addition, NSBA’s NSBA Splice Microsoft Excel-based bolted field splice design spreadsheet
takes the time-consuming task of designing and checking a bolted splice connection and rewrites
the process with a simple input page and output form. NSBA Splice can be incorporated as a
design tool on plate girder bridges allowing the designer to quickly analyze various bolted splice
connections to determine the most efficient bolt quantity and configuration. NSBA Splice allows
the user to explore the effects of bolt spacing, bolt size, strength, and connection dimensions on
the overall splice design.

6.13.6.1.3c Web Splices


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 342


Discussion:
The provisions of this Article cover the design of web splice plates and their connections for bolted
field splices in flexural members and are applicable to the routine steel I-girder bridges covered
by this Guide.
Web splices must be designed at a minimum at the strength limit state for a design web force taken
equal to the smaller factored shear resistance of the web, Vr, on either side of the splice (see the
Discussion of Article 6.10.9 in this Guide). Since the web splice is being designed to develop the
full factored shear resistance of the web at a minimum and the eccentricity of the shear is small
relative to the depth of the connection, the effect of the small moment introduced by the eccentricity
of the web connection is ignored at all limit states.
If the moment resistance provided by the flanges (see the Discussion of Article 6.13.6.1.3b in this
Guide) is not sufficient to resist the factored moment at the strength limit state at the point of splice,
the web splice connections are to be designed for a resultant design web force, R, taken equal to
the vector sum of the smaller factored shear resistance, Vr, and a horizontal force, Hw, in the web
that provides the required moment resistance in conjunction with the flange splices. For composite
sections subject to positive flexure, Hw is computed as shown in Figure C6.13.6.1.3c-1. For
composite sections subject to negative flexure and for noncomposite sections subject to positive
or negative flexure, Hw is computed as shown in Figure C6.13.6.1.3c-2. The web moment in these
figures is the portion of the factored moment at the strength limit state at the point of splice that
exceeds the moment resistance provided by the flanges. Because the resultant web force is assumed
divided equally to all the bolts, the traditional vector analysis for connections subject to eccentric
shear is not applied.
The number of bolts required on one side of the web splice at the strength limit state is found by
dividing design web force, Vr or R as applicable, by the factored shear resistance of the bolts (see
the Discussion of Articles 6.13.2.2 and 6.13.2.7 in this Guide). The threads should be assumed
included in the shear planes for most common web splices since the web splice-plate thicknesses
are normally less than or equal to ½ in. The number of shear planes, Ns, is equal to two for web
splices. The joint length reduction factor of 0.83 applies only to lap-splice tension connections
greater than 38.0 in. in length and does not apply to web splices. At a minimum, two vertical rows
of bolts spaced at the maximum spacing for sealing bolts should be provided (see the Discussion of
Article 6.13.2.6.2 in this Guide) with a closer spacing and/or additional rows provided only as
needed.
The bearing resistance of the web splice bolt holes is to be checked at the strength limit state (see
the Discussion of Article 6.13.2.9 in this Guide). The bearing resistance may be calculated as the
sum of the smaller of the shear resistance of the individual bolts and the bearing resistance of the
individual bolt holes parallel to the line of the design force. If the bearing resistance of a bolt hole
exceeds the shear resistance of the bolt, the bolt resistance is limited to the shear resistance. When
a moment contribution from the web is required at the strength limit state, the resultant forces
causing bearing on the web bolt holes are inclined. The bearing resistance of each bolt hole in the
web can conservatively be calculated in this case using the clear edge distance, as shown on the
left-hand side of Figure 6.13.6.1.3c. This calculation is conservative since the resultant forces act
in the direction of inclined distances larger than the clear edge distance. This calculation is also

NSBA Guide to Navigating Routine Steel Bridge Design / 343


likely to be a conservative calculation for the bolt holes in the adjacent rows. Should the bearing
resistance be exceeded, it is recommended that the clear edge distance be increased slightly in lieu
of increasing the number of bolts or thickening the web. Other possible options are discussed in
the Commentary for this Article. In rare cases where the bearing resistance is controlled by the
web splice plates, the smaller of the clear edge or end distance on the splice plates can
conservatively be used to compute the bearing resistances of each hole, as shown on the right-hand
side of Figure 6.13.6.1.3c.
Webs are to be spliced symmetrically by plates on each side. The splice plates are to extend as
near as practical for the full depth between flanges without impinging on bolt assembly clearances.
Recommended bolt assembly clearances are given in Tables 7-15 and 7-16, as applicable, of the
AISC Manual of Steel Construction. The thickness of each web splice plate should be at least one-
half the thickness of the thinner web at the splice plus 1/16 of an inch, and must be equal to or
greater than the minimum permitted thickness of 5/16 in. for structural steel (see the Discussion of
Article 6.7.3 in this Guide). For bolted web splices with thickness differences of 1/16 in. or less,
filler plates should not be provided. For a web thickness change of 1/8 in., use a 1/8 in. filler plate
on one side of the web rather than 1/16 in. filler plates on each side; filler plates 1/16 in. or less create
difficulties in fabrication and handling. A minimum gap of ½ in. between the girder sections at the
splice should be provided to provide drainage and allow for fit-up. The factored shear resistance
of the web, Vr, is not to exceed the factored shear resistance of the web splice plates; that is, the
smaller value based on shear yielding or shear rupture (see the Discussion of Article 6.13.5.3 in
this Guide). Vr is also not to exceed the factored block shear rupture resistance of the web splice
plates, which is unlikely to control (see the Discussion of Article 6.13.4 in this Guide). The factors,
Rp and Ubs, are to be taken equal to 1.0 for splice plates in the shear rupture and block shear rupture
checks, respectively.

At a minimum, bolted connections for web splices are to be checked for slip under a web slip force
taken equal to the factored shear in the web at the point of splice. The factored shear for checking
slip is taken as the shear in the web at the point of splice under Load Combination Service II (see
the Discussion of Article 3.4.1 in this Guide), or the factored shear at the point of splice due to the
deck casting sequence (see the Discussion of Article 3.4.2.1 in this Guide), whichever governs.
Should the flange bolts not be sufficient to resist the factored moment for checking slip at the point
of splice (see the Discussion of Article 6.13.6.1.3b in this Guide), the web splice bolts should
instead be checked for slip under a resultant web slip force taken equal to the vector sum of the
governing factored shear and a horizontal force, Hw, located in the web that provides the necessary
slip resistance in conjunction with the flange splices. Hw is computed as the portion of the factored
moment for checking slip at the point of splice that exceeds the moment resistance provided by
the nominal slip resistance of the flange splice bolts divided by the appropriate moment arm, Aw.
For composite sections subject to positive flexure, Aw is computed as shown in Figure
C6.13.6.1.3c-1. For composite sections subject to negative flexure and for noncomposite sections
subject to positive or negative flexure, Aw is computed as shown in Figure C6.13.6.1.3c-2. The
computed web slip force is then divided by the nominal slip resistance of the bolts (see the
Discussion of Article 6.13.2.8 in this Guide) to determine the number of web splice bolts required
on one side of the splice to resist slip. In cases where the moment resistance provided by the flange

NSBA Guide to Navigating Routine Steel Bridge Design / 344


splice bolts is sufficient at the strength limit state, but a moment contribution from the web is
required to resist slip, the number of flange splice bolts may be increased to increase the moment
resistance provided by the nominal slip resistance of the flange splice bolts in order to prevent
having to add an additional row of web splice bolts to resist the resultant web slip force.
The portions of this Article and the associated Commentary dealing with box sections are not
applicable to the routine steel I-girder bridges covered by this Guide.
For further information on field splice design, consult Section 6.6.5 and especially 6.6.5.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures – note that this Reference Manual refers to the previous
AASHTO provisions for bolted field splice design. The provisions addressing the design of bolted
field splices changed significantly in 2017 with the publication of the 8th Edition of the AASHTO
LRFD BDS. For in-depth information on the current AASHTO provisions for the design of bolted
field splices for flexural members and examples illustrating the design of such splices, consult
NSBA’s Bolted Field Splices for Steel Bridge Flexural Members – Overview and Design
Examples.
In addition, NSBA’s NSBA Splice Microsoft Excel-based bolted field splice design spreadsheet
takes the time-consuming task of designing and checking a bolted splice connection and rewrites
the process with a simple input page and output form. NSBA Splice can be incorporated as a
design tool on plate girder bridges allowing the designer to quickly analyze various bolted splice
connections to determine the most efficient bolt quantity and configuration. NSBA Splice allows
the user to explore the effects of bolt spacing, bolt size, strength, and connection dimensions on
the overall splice design.

6.13.6.1.4 Fillers
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article cover the design of fillers in axially loaded connections, including
truss gusset plate chord splices and bolted field splices for flexural members. As such, the
provisions related to fillers in bolted field splices for flexural members are applicable to the routine
steel I-girder bridges covered by this Guide when such fillers occur and their thickness is ¼ inch
or greater.
Fillers are typically used on bolted flange splices of flexural members (and sometimes on web
splices) when the thicknesses of the adjoining plates at the point of splice are different. At bolted
flange splices, it is often advantageous to transition one or more of the flange thicknesses down
adjacent to the point of splice, if possible, so as to reduce the required size of the filler plate, or
possibly change the width of the flanges and keep the thickness constant in order to eliminate the
need for a filler plate altogether. Fillers must be secured by additional bolts such that the fillers are
an integral part of the connection at the strength limit state; that is, such that the shear planes are
well-defined and that no reduction in the factored shear resistance of the bolts results due to
bending of the bolts.

NSBA Guide to Navigating Routine Steel Bridge Design / 345


Note that changing the girder web depth at bolted field splices to minimize the thickness of fillers
is not recommended. The girder web depth should be held constant at field splices in routine steel
I-girder bridges.
These provisions provide two choices of methods for developing fillers 0.25 in. or more in
thickness. The choices are to either:
1. Extend the fillers beyond the gusset or splice plate with the filler extension secured by
enough additional bolts to distribute the total stress uniformly over the combined section
of the member or filler; or
2. In lieu of extending and developing the fillers, reduce the factored shear resistance of the
bolts (see the Discussion of Article 6.13.2.7 in this Guide) by the factor, R, given by Eq.
6.13.6.1.4-1.
In general, the second method is more commonly used, and the first method is rarely, if ever, used
in routine steel I-girder bridges. R accounts for the reduction in the nominal shear resistance of the
bolts due to bending in the bolts and will likely result in having to provide additional bolts in the
connection to develop the filler(s). Note that the reduction factor is only to be applied on the side
of the connections with the filler(s). For practical reasons, consideration should be given to using
the same number of bolts on either side of the splice. Normally, the splice plate widths, filler plate
widths, and flange widths will be equal in the splice, and the connected plate area will be smaller
than the sum of the splice plate areas, such that the area ratio, γ, in the equation for R may simply
be taken as the ratio of the thickness of the filler to the thickness of the connected plate.
Note that fillers 0.25-in. or more in thickness are not to consist of more than two plates, unless
approved by the Engineer. As discussed further in the Commentary for this Article, the actual total
filler thickness may exceed the total filler thickness shown in the contract documents by up to a
maximum of 0.25 in.
These provisions also require that the specified minimum yield strength of fillers 0.25 in. or more
in thickness not be less than the larger of 70 percent of the specified minimum yield strength of
the connected plate and 36.0 ksi. This provision is primarily applicable to designs utilizing steels
with a specified minimum yield strength greater than 50 ksi, in which case there are likely to be
thinner filler-plate material availability issues. To control the potential for bolt bending and
excessive deformation of the connection in such cases, a lower limit on the specified minimum
yield strength of the fillers is specified. The effects of yielding of the fillers and connection
deformation are not considered to be significant for connections with fillers less than 0.25 in. in
thickness. Note that for connections involving the use of weathering steels, a weathering grade
product should be specified for the filler-plate material.
The resistance to slip between the filler and either connected part at the service limit state is
comparable to the slip resistance that would exist between the connected parts if the filler were not
present. Therefore, for slip-critical connections (see the Discussion of Article 6.13.2.1.1 in this
Guide), the factored slip resistance of the bolts (see the Discussion of Article 6.13.2.8 in this Guide)
is not to be adjusted for the effect of the fillers.

NSBA Guide to Navigating Routine Steel Bridge Design / 346


For further information on field splice design, consult Section 6.6.5 and especially 6.6.5.2 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures – note that this Reference Manual refers to the previous
AASHTO provisions for bolted field splice design. The provisions addressing the design of bolted
field splices changed significantly in 2017 with the publication of the 8th Edition of the AASHTO
LRFD BDS. For in-depth information on the current AASHTO provisions for the design of bolted
field splices for flexural members and examples illustrating the design of such splices, consult
NSBA’s Bolted Field Splices for Steel Bridge Flexural Members – Overview and Design
Examples.
In addition, NSBA’s NSBA Splice Microsoft Excel-based bolted field splice design spreadsheet
takes the time-consuming task of designing and checking a bolted splice connection and rewrites
the process with a simple input page and output form. NSBA Splice can be incorporated as a
design tool on plate girder bridges allowing the designer to quickly analyze various bolted splice
connections to determine the most efficient bolt quantity and configuration. NSBA Splice allows
the user to explore the effects of bolt spacing, bolt size, strength, and connection dimensions on
the overall splice design.

6.13.6.2 Welded Splices


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
A splice is defined as a welded connection (or a group of bolted connections) sufficient to transfer
the moment, shear, axial force or torque between two structural elements joined at their ends to
form a single, longer element. Welded splices are typically used to connect members and/or their
components together in the fabrication shop; hence, the term “shop splice” is often used. Welded
field splices are less commonly used than bolted field splices; if used, they should be arranged to
minimize overhead welding. The provisions of this Article cover the design of welded splices,
which must conform to the requirements of the AASHTO/AWS D1.5M/D1.5 Bridge Welding
Code.
Welded splices for tension, and compression, and flexural members are to be designed using
complete joint penetration (CJP) groove welds (see the Discussion of Article 6.13.3.2.2a in this
Guide). Only the provisions of this Article relevant to flexural members are applicable to the
routine steel I-girder bridges covered by this Guide. The use of splice plates should be avoided at
welded splices.
Fatigue should be checked at welded splices subject to an applied net tensile stress (determined as
specified in Article 6.6.1.2.1) based on the appropriate fatigue detail category for the splice
configuration given in Table 6.6.1.2.3-1 (see the Discussion of Articles 6.6.1.2.1 and 6.6.1.2.3 in
this Guide).
Changing flange widths at welded shop splices in plate girders should be avoided if possible;
flange widths are best changed at bolted field splices. This facilitates “slabbing” of flanges during
fabrication, as explained and illustrated in Section 1.5.2 of the AASHTO-NSBA Steel Bridge
Collaboration’s Guideline G12.1-2020 Guidelines to Design for Constructability and Fabrication.

NSBA Guide to Navigating Routine Steel Bridge Design / 347


Should it become necessary to splice material of different widths using welded butt joints,
symmetric transitions must be used that conform to one of the details shown in Figure 6.13.2.6.2-
1. The transition often starts at the butt splice. However, note that Figure 6.13.2.6.2-1 shows a
preferred detail in which the butt splice is located a minimum of 3.0 in. from the transition for
greater ease in fitting the run-off tabs. At welded butt splices joining material of different
thicknesses, the transition (including the weld) must be ground to a uniform slope between the
offset surfaces of not more than 1 in 2.5, and must be indicated as such in the contract documents.
Efficiently locating thickness transitions in plate girder flanges is a matter of plate length
availability and the economics of welding and inspecting a splice compared to the cost of extending
a thicker plate. A shop-welded butt splice should be introduced in a beam or girder flange when
the savings in flange material and plate-length limitation or special circumstances dictate. Table
1.5.4-1 in the AASHTO-NSBA Steel Bridge Collaboration’s Guideline G12.1-2020 Guidelines to
Design for Constructability and Fabrication shows a weight savings per inch of flange width that
may be used to evaluate placement of shop splices. The criteria vary so Fabricators should be
consulted whenever possible. In the contract plans or specifications, provide criteria the Fabricator
may follow to eliminate shop-welded flange splices by extending the thicker plate, subject to the
approval of the Engineer. When evaluating the request, the Engineer should review the percent
change in deflections and stresses resulting from the extension of the thicker plate.
The provisions of this Article are applicable to the welded butt splices (and welded field splices if
used) in the routine steel plate girder bridges covered by this Guide. Welded butt splices are not
typically used on the routine steel rolled beam bridges covered by this Guide and therefore these
provisions are not applicable to those bridges unless welded field splices are used.
The AASHTO-NSBA Steel Bridge Collaboration’s Guideline G12.1-2020 Guidelines to Design
for Constructability and Fabrication provides practical guidance on the design and detailing of
flanges and flange shop splices to facilitate economical fabrication.

6.13.7 Rigid Frame Connections

6.13.7.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article deals with the connections for a rigid frame, which is defined as a structure in which
the connections maintain the angular relationship between the beam and column members under
load. This Article states that rigid frame connections are to be designed to resist the factored
moments, shear, and axial forces at the strength limit state.
These provisions for the design of rigid frame connections are not applicable to the routine steel
I-girder bridges covered by this Guide.

6.13.7.2 Webs
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 348


Discussion:
This Article requires that rigid frame connections have sufficient strength and stiffness to resist
the factored loading effects under the strength limit state. In some cases, stiffening of the web in
the connection regions can be beneficial in meeting this requirement. A rigid frame is defined as a
structure in which the connections maintain their angular relationship between the beam and
column members under load.
These provisions for the design of the beam or connection web in a rigid frame are not applicable
to the routine steel I-girder bridges covered by this Guide.

6.14 PROVISIONS FOR STRUCTURE TYPES

6.14.1 Through-Girder Spans


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article apply to the design of through-girder spans. Through-girder spans
are bridges in which the top flanges of the main girders are located above the top of the deck. This
type of design is not used for the routine steel I-girder bridges covered by this Guide, and as such
the provisions are not applicable to their design.

6.14.2 Trusses

6.14.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of Article 6.14.2 and its related sub-Articles are primarily intended for the design
of truss bridges, where the truss is the main spanning element. These provisions are not intended
for the design of truss-type cross-frames, except for the specific application of the provisions of
Article 6.14.2.8 to the design of gusset plates used to connect truss-type cross-frame members to
cross-frame connection plates (stiffeners) (see the Discussion of Article 6.14.2.8 in this Guide).
As a result, Article 6.14.2 and its related sub-Articles, except for Article 6.14.2.8, are not
applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.14.2.2 Truss Members


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of Article 6.14.2 and its related sub-Articles are primarily intended for the design
of truss bridges, where the truss is the main spanning element. These provisions are not intended
for the design of truss-type cross-frames, except for the specific application of the provisions of
Article 6.14.2.8 to the design of gusset plates used to connect truss-type cross-frame members to
cross-frame connection plates (stiffeners) (see the Discussion of Article 6.14.2.8 in this Guide).

NSBA Guide to Navigating Routine Steel Bridge Design / 349


As a result, Article 6.14.2 and its related sub-Articles, except for Article 6.14.2.8, are not
applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.14.2.3 Secondary Stresses


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of Article 6.14.2 and its related sub-Articles are primarily intended for the design
of truss bridges, where the truss is the main spanning element. These provisions are not intended
for the design of truss-type cross-frames, except for the specific application of the provisions of
Article 6.14.2.8 to the design of gusset plates used to connect truss-type cross-frame members to
cross-frame connection plates (stiffeners) (see the Discussion of Article 6.14.2.8 in this Guide).
As a result, Article 6.14.2 and its related sub-Articles, except for Article 6.14.2.8, are not
applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.14.2.4 Diaphragms
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of Article 6.14.2 and its related sub-Articles are primarily intended for the design
of truss bridges, where the truss is the main spanning element. These provisions are not intended
for the design of truss-type cross-frames, except for the specific application of the provisions of
Article 6.14.2.8 to the design of gusset plates used to connect truss-type cross-frame members to
cross-frame connection plates (stiffeners) (see the Discussion of Article 6.14.2.8 in this Guide).
As a result, Article 6.14.2 and its related sub-Articles, except for Article 6.14.2.8, are not
applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.14.2.5 Camber
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of Article 6.14.2 and its related sub-Articles are primarily intended for the design
of truss bridges, where the truss is the main spanning element. These provisions are not intended
for the design of truss-type cross-frames, except for the specific application of the provisions of
Article 6.14.2.8 to the design of gusset plates used to connect truss-type cross-frame members to
cross-frame connection plates (stiffeners) (see the Discussion of Article 6.14.2.8 in this Guide).
As a result, Article 6.14.2 and its related sub-Articles, except for Article 6.14.2.8, are not
applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.14.2.6 Working Lines and Gravity Axes


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 350


Discussion:
The provisions of Article 6.14.2 and its related sub-Articles are primarily intended for the design
of truss bridges, where the truss is the main spanning element. These provisions are not intended
for the design of truss-type cross-frames, except for the specific application of the provisions of
Article 6.14.2.8 to the design of gusset plates used to connect truss-type cross-frame members to
cross-frame connection plates (stiffeners) (see the Discussion of Article 6.14.2.8 in this Guide).
As a result, Article 6.14.2 and its related sub-Articles, except for Article 6.14.2.8, are not
applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.14.2.7 Portal and Sway Bracing

6.14.2.7.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of Article 6.14.2 and its related sub-Articles are primarily intended for the design
of truss bridges, where the truss is the main spanning element. These provisions are not intended
for the design of truss-type cross-frames, except for the specific application of the provisions of
Article 6.14.2.8 to the design of gusset plates used to connect truss-type cross-frame members to
cross-frame connection plates (stiffeners) (see the Discussion of Article 6.14.2.8 in this Guide).
As a result, Article 6.14.2 and its related sub-Articles, except for Article 6.14.2.8, are not
applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.14.2.7.2 Through-Truss Spans


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of Article 6.14.2 and its related sub-Articles are primarily intended for the design
of truss bridges, where the truss is the main spanning element. These provisions are not intended
for the design of truss-type cross-frames, except for the specific application of the provisions of
Article 6.14.2.8 to the design of gusset plates used to connect truss-type cross-frame members to
cross-frame connection plates (stiffeners) (see the Discussion of Article 6.14.2.8 in this Guide).
As a result, Article 6.14.2 and its related sub-Articles, except for Article 6.14.2.8, are not
applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.14.2.7.3 Deck Truss Spans


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of Article 6.14.2 and its related sub-Articles are primarily intended for the design
of truss bridges, where the truss is the main spanning element. These provisions are not intended
for the design of truss-type cross-frames, except for the specific application of the provisions of
Article 6.14.2.8 to the design of gusset plates used to connect truss-type cross-frame members to

NSBA Guide to Navigating Routine Steel Bridge Design / 351


cross-frame connection plates (stiffeners) (see the Discussion of Article 6.14.2.8 in this Guide).
As a result, Article 6.14.2 and its related sub-Articles, except for Article 6.14.2.8, are not
applicable to the design of the routine steel I-girder bridges covered by this Guide.

6.14.2.8 Gusset Plates

6.14.2.8.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article addresses general requirements for the configuration and detailing of gusset plates.
The provisions of Article 6.14.2.8 and its associated sub-Articles address the design of gusset
plates; the provisions are primarily intended for use in the design of truss bridges, where the truss
is the main spanning element, but can be applied to the design of gusset plates used to connect the
members of truss-type cross-frames to cross-frame connection plates (stiffeners).
The provision that fasteners connecting each member be symmetrical with the axis of the member,
so far as practicable, is intended for the bolted connection of truss members to gusset plates; in
most cases for routine steel I-girder bridges, the members of truss-type cross-frames are welded to
the gusset plates. It is not required that the bolted connection of the gusset plate itself to the cross-
frame connection plate (stiffener) be symmetrical to any given cross-frame member.
Gusset plates for truss-type cross-frames in routine steel I-girder bridges are not considered “chord
splices” and are not multilayered and so the related provisions in the following sub-articles are not
applicable.
The remaining requirements of this provision are applicable to the design of gusset plates used in
truss-type cross-frames of the routine steel I-girder bridges covered by this Guide.

6.14.2.8.2 Multilayered Gusset and Splice Plates


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Multilayered gusset and splice plates are not used in truss-type cross-frames for routine steel I-
girder bridges. The combination of a cross-frame gusset plate and a cross-frame connection plate
(stiffener) is not a “multilayered gusset plate.” As a result, this Article is not applicable to the
design of the routine steel I-girder bridges covered by this Guide.

6.14.2.8.3 Shear Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 352


Discussion:
This Article addresses the calculation of the factored shear resistance of gusset plates for truss
bridges, but should also be applied to the design of gusset plates for truss-type cross-frames in the
routine steel I-girder bridges covered by this Guide.
The Article includes a unique equation for the factored shear resistance associated with shear
yielding of gusset plates (Eq. 6.14.2.8.3-1). Investigation of shear rupture is also required, using
Eq. 6.13.5.3-2 (see the Discussion of Article 6.13.5.3 in this Guide). The factored shear resistance
is to be taken as the smaller of the factored shear yielding or shear rupture resistance. The shear
yielding resistance is typically calculated using the gross area of the plane adjacent to a row of
bolts, while the shear rupture resistance is typically calculated using the net area of the plane
through a row of bolts (i.e., the area of the plane minus the area removed by the bolt holes). The
plane investigated is a plane parallel to the direction of the applied shear loading. The definition
of shear planes is dependent on the specific geometry of the gusset plate and the applied loads; in
some cases the load may need to be broken into orthogonal components parallel to the lines of
bolts. This Article and its associated Commentary include figures which help illustrate how to
define shear planes. Generally, it is necessary to use a to-scale drawing of the gusset plate, attached
members, and loadings to help define the controlling shear planes.
A check of block shear rupture resistance is typically also necessary, as defined in Article 6.13.4.
The block shear rupture perimeter in this check is typically defined as the perimeter of the welded
connection of a cross-frame member to the gusset plate. See the Discussion of Article 6.13.4 in
this Guide for more information.
For further information on the block shear rupture resistance and design examples illustrating the
block shear rupture resistance checks, consult Sections 6.6.3.3.2.5 and 6.6.4.2.5.6.1 of the
Reference Manual for NHI Course 130081, Load and Resistance Factor Design (LRFD) for
Highway Bridge Superstructures.

6.14.2.8.4 Compressive Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article addresses the calculation of the factored compressive resistance of gusset plates for
truss bridges but should also be applied to the design of gusset plates for truss-type cross-frames
in the routine steel I-girder bridges covered by this Guide. The resistance equations are based on
modified column buckling equations and Whitmore section analysis. This Article and its
associated Commentary include figures which help illustrate how to define the Whitmore section.
Generally, it is necessary to use a to-scale drawing of the gusset plate, attached members, and
loadings to help define the Whitmore section.

6.14.2.8.5 Tensile Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 353


Discussion:
This Article addresses the calculation of the factored tensile resistance of gusset plates for truss
bridges but should also be applied to the design of gusset plates for truss-type cross-frames in the
routine steel I-girder bridges covered by this Guide. The provisions state that the factored tensile
resistance be taken as the smallest factored resistance in tension based on yielding, fracture, or
block shear rupture determined according to the provisions of Article 6.13.5.2 (see the Discussion
of Article 6.13.5.2 in this Guide). When calculating the tensile yielding and net section fracture
resistances according to the provisions of Article 6.8.2.1 (see the Discussion of Article 6.8.2.1 in
this Guide), the Whitmore section defined in Figure 6.14.2.8.5-1 should be used. Generally, it is
necessary to use a to-scale drawing of the gusset plate, attached members, and loadings to help
define the Whitmore section.

6.14.2.8.6 Chord Splices


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article addresses the design of chord splices for truss bridges. Chord splices are the splices
of truss chords when the truss is the main spanning element of the bridge; this Article is not
applicable to the design of the routine steel I-girder bridges covered by this Guide since these types
of structures do not use chord splices.

6.14.2.8.7 Edge Slenderness


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
This Article defines a proportioning limit related to the slenderness of the unsupported edge of a
gusset plate for truss bridges, but should also be applied to the design of gusset plates for truss-
type cross-frames in the routine steel I-girder bridges covered by this Guide. The proportioning
limit is defined using a simple equation which reflects the material properties and thickness of the
gusset plate. The provision requires stiffening of the edge if the proportioning limit is not met.
However, gusset plates for truss-type cross-frames in routine steel I-girder bridges should not be
stiffened as such stiffening adds considerable fabrication expense; if the proportioning limit is not
met, increase the thickness of the gusset plate.

6.14.2.9 Half Through-Trusses


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article apply to the design of half through-trusses. Half through-trusses are
bridges in which the top chords of the trusses are located above the top of the deck while the
bottom chords are located below the deck. This type of design is not used in the routine steel I-
girder bridges covered by this Guide, and as such the provisions are not applicable to their design.

NSBA Guide to Navigating Routine Steel Bridge Design / 354


6.14.2.10 Factored Resistance
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions included in this Article are intended for use in the design of truss members when
the truss is the main spanning element of the bridge. Consequently, this Article is not applicable
to the design of the routine steel I-girder bridges covered by this Guide.

6.14.3 Orthotropic Deck Superstructures

6.14.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Orthotropic steel decks are bridge decks constructed using stiffened steel plates as the structural
system of the deck. By the definitions of this Guide, the routine steel I-girder bridges use composite
reinforced concrete decks. Thus, Article 6.14.3 and its associated sub-Articles are not applicable.

6.14.3.2 Decks in Global Compression

6.14.3.2.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Orthotropic steel decks are bridge decks constructed using stiffened steel plates as the structural
system of the deck. By the definitions of this Guide, the routine steel I-girder bridges covered by
this Guide use composite reinforced concrete decks. Thus, Article 6.14.3 and its associated sub-
Articles are not applicable.

6.14.3.2.2 Local Buckling


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Orthotropic steel decks are bridge decks constructed using stiffened steel plates as the structural
system of the deck. By the definitions of this Guide, the routine steel I-girder bridges covered by
this Guide use composite reinforced concrete decks. Thus, Article 6.14.3 and its associated sub-
Articles are not applicable.

6.14.3.2.3 Panel Buckling


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 355


Discussion:
Orthotropic steel decks are bridge decks constructed using stiffened steel plates as the structural
system of the deck. By the definitions of this Guide, the routine steel I-girder bridges covered by
this Guide use composite reinforced concrete decks. Thus, Article 6.14.3 and its associated sub-
Articles are not applicable.

6.14.3.3 Effective Width of Deck


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Orthotropic steel decks are bridge decks constructed using stiffened steel plates as the structural
system of the deck. By the definitions of this Guide, the routine steel I-girder bridges covered by
this Guide use composite reinforced concrete decks. Thus, Article 6.14.3 and its associated sub-
Articles are not applicable.

6.14.3.4 Superposition of Global and Local Effects


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Orthotropic steel decks are bridge decks constructed using stiffened steel plates as the structural
system of the deck. By the definitions of this Guide, the routine steel I-girder bridges covered by
this Guide use composite reinforced concrete decks. Thus, Article 6.14.3 and its associated sub-
Articles are not applicable.

6.14.4 Solid Web Arches

6.14.4.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Solid web arches are bridges where the steel arch is the main spanning element of the bridge. Thus,
Article 6.14.4 and its associated sub-Articles are not applicable to the design of the routine steel I-
girder bridges covered by this Guide.

6.14.4.2 Web Slenderness


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Solid web arches are bridges where the steel arch is the main spanning element of the bridge. Thus,
Article 6.14.4 and its associated sub-Articles are not applicable to the design of the routine steel I-
girder bridges covered by this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 356


6.14.4.3 Moment Amplification
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Solid web arches are bridges where the steel arch is the main spanning element of the bridge. Thus,
Article 6.14.4 and its associated sub-Articles are not applicable to the design of the routine steel I-
girder bridges covered by this Guide.

6.14.4.4 Nominal Compressive Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Solid web arches are bridges where the steel arch is the main spanning element of the bridge. Thus,
Article 6.14.4 and its associated sub-Articles are not applicable to the design of the routine steel I-
girder bridges covered by this Guide.

6.14.4.5 Nominal Flexural Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Solid web arches are bridges where the steel arch is the main spanning element of the bridge. Thus,
Article 6.14.4 and its associated sub-Articles are not applicable to the design of the routine steel I-
girder bridges covered by this Guide.

6.14.4.6 Combined Axial Compression or Tension with Flexural and Torsion


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
Solid web arches are bridges where the steel arch is the main spanning element of the bridge. Thus,
Article 6.14.4 and its associated sub-Articles are not applicable to the design of the routine steel I-
girder bridges covered by this Guide.

6.15 PILES

6.15.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of Article 6.15 and the associated sub-Articles present the requirements for the
structural design of steel piles. Other requirements for the design of piles are presented in Chapter
10 of the AASHTO LRFD BDS. Steel piles take several forms, such as steel H-piles or steel pipe
piles. The provisions of Article 6.15 and the associated sub-Articles often defer to other Articles

NSBA Guide to Navigating Routine Steel Bridge Design / 357


in Chapter 6 which directly address the calculation of resistance to axial compressive loads,
flexure, or their combined effects. In some cases, H-pile sections have been used for superstructure
elements, but in those cases the design is governed by the appropriate related Articles elsewhere
in Chapter 6. The provisions of Article 6.15 and its associated sub-Articles are specifically
intended for use in the design of these sections when used as piles in the substructure or
foundations of a bridge. As such, this Article addresses design items which are beyond the scope
of superstructure design.

6.15.2 Structural Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of Article 6.15 and the associated sub-Articles present the requirements for the
structural design of steel piles. Other requirements for the design of piles are presented in Chapter
10 of the AASHTO LRFD BDS. Steel piles take several forms, such as steel H-piles or steel pipe
piles. The provisions of Article 6.15 and the associated sub-Articles often defer to other Articles
in Chapter 6 which directly address the calculation of resistance to axial compressive loads,
flexure, or their combined effects. In some cases, H-pile sections have been used for superstructure
elements, but in those cases the design is governed by the appropriate related Articles elsewhere
in Chapter 6. The provisions of Article 6.15 and its associated sub-Articles are specifically
intended for use in the design of these sections when used as piles in the substructure or
foundations of a bridge. As such, this Article addresses design items which are beyond the scope
of superstructure design.

6.15.3 Compressive Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of Article 6.15 and the associated sub-Articles present the requirements for the
structural design of steel piles. Other requirements for the design of piles are presented in Chapter
10 of the AASHTO LRFD BDS. Steel piles take several forms, such as steel H-piles or steel pipe
piles. The provisions of Article 6.15 and the associated sub-Articles often defer to other Articles
in Chapter 6 which directly address the calculation of resistance to axial compressive loads,
flexure, or their combined effects. In some cases, H-pile sections have been used for superstructure
elements, but in those cases the design is governed by the appropriate related Articles elsewhere
in Chapter 6. The provisions of Article 6.15 and its associated sub-Articles are specifically
intended for use in the design of these sections when used as piles in the substructure or
foundations of a bridge. As such, this Article addresses design items which are beyond the scope
of superstructure design.

NSBA Guide to Navigating Routine Steel Bridge Design / 358


6.15.3.1 Axial Compression
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of Article 6.15 and the associated sub-Articles present the requirements for the
structural design of steel piles. Other requirements for the design of piles are presented in Chapter
10 of the AASHTO LRFD BDS. Steel piles take several forms, such as steel H-piles or steel pipe
piles. The provisions of Article 6.15 and the associated sub-Articles often defer to other Articles
in Chapter 6 which directly address the calculation of resistance to axial compressive loads,
flexure, or their combined effects. In some cases, H-pile sections have been used for superstructure
elements, but in those cases the design is governed by the appropriate related Articles elsewhere
in Chapter 6. The provisions of Article 6.15 and its associated sub-Articles are specifically
intended for use in the design of these sections when used as piles in the substructure or
foundations of a bridge. As such, this Article addresses design items which are beyond the scope
of superstructure design.

6.15.3.2 Combined Axial Compression and Flexure


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of Article 6.15 and the associated sub-Articles present the requirements for the
structural design of steel piles. Other requirements for the design of piles are presented in Chapter
10 of the AASHTO LRFD BDS. Steel piles take several forms, such as steel H-piles or steel pipe
piles. The provisions of Article 6.15 and the associated sub-Articles often defer to other Articles
in Chapter 6 which directly address the calculation of resistance to axial compressive loads,
flexure, or their combined effects. In some cases, H-pile sections have been used for superstructure
elements, but in those cases the design is governed by the appropriate related Articles elsewhere
in Chapter 6. The provisions of Article 6.15 and its associated sub-Articles are specifically
intended for use in the design of these sections when used as piles in the substructure or
foundations of a bridge. As such, this Article addresses design items which are beyond the scope
of superstructure design.

6.15.3.3 Buckling
Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of Article 6.15 and the associated sub-Articles present the requirements for the
structural design of steel piles. Other requirements for the design of piles are presented in Chapter
10 of the AASHTO LRFD BDS. Steel piles take several forms, such as steel H-piles or steel pipe
piles. The provisions of Article 6.15 and the associated sub-Articles often defer to other Articles

NSBA Guide to Navigating Routine Steel Bridge Design / 359


in Chapter 6 which directly address the calculation of resistance to axial compressive loads,
flexure, or their combined effects. In some cases, H-pile sections have been used for superstructure
elements, but in those cases the design is governed by the appropriate related Articles elsewhere
in Chapter 6. The provisions of Article 6.15 and its associated sub-Articles are specifically
intended for use in the design of these sections when used as piles in the substructure or
foundations of a bridge. As such, this Article addresses design items which are beyond the scope
of superstructure design.

6.15.4 Maximum Permissible Driving Stresses


Determination of applicability, All Routine Steel I-girder Bridges: Beyond scope of superstructure
design.
Discussion:
The provisions of Article 6.15 and the associated sub-Articles present the requirements for the
structural design of steel piles. Other requirements for the design of piles are presented in Chapter
10 of the AASHTO LRFD BDS. Steel piles take several forms, such as steel H-piles or steel pipe
piles. The provisions of Article 6.15 and the associated sub-Articles often defer to other Articles
in Chapter 6 which directly address the calculation of resistance to axial compressive loads,
flexure, or their combined effects. In some cases, H-pile sections have been used for superstructure
elements, but in those cases the design is governed by the appropriate related Articles elsewhere
in Chapter 6. The provisions of Article 6.15 and its associated sub-Articles are specifically
intended for use in the design of these sections when used as piles in the substructure or
foundations of a bridge. As such, this Article addresses design items which are beyond the scope
of superstructure design.

6.16 PROVISIONS FOR SEISMIC DESIGN

6.16.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The provisions of this Article apply for the seismic design of steel-girder bridge superstructures at
the extreme event limit state.
This Article discusses the need to define a clear seismic load path within the superstructure to
transmit the inertia forces to the substructure based on the stiffness characteristics of the concrete
deck, cross-frames or diaphragms, and bearings. The flow of the seismic forces is to be
accommodated through the affected components and connections of the superstructure within the
defined load path. This Article also refers to the minimum support-length requirements at
expansion bearings specified in Article 4.7.4.4 (see the Discussion of Article 4.7.4.4 in this Guide).
For the application of the seismic design provisions in the AASHTO LRFD BDS, the routine steel
I-girder bridges covered by this Guide are assumed to be located in Seismic Zone 1 (see the
Discussion of Article 6.16.3 in this Guide).

NSBA Guide to Navigating Routine Steel Bridge Design / 360


6.16.2 Materials
Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
For the application of the seismic design provisions in the AASHTO LRFD BDS, the routine steel
I-girder bridges covered by this Guide are assumed to be located in Seismic Zone 1 (see the
Discussion of Article 6.16.3 in this Guide). As specified in this Article, structural steels within the
seismic load path are to satisfy the requirements of Article 6.4.1 (see the Discussion of Article
6.4.1 in this Guide). Capacity design to protect members and connections is not to be used for
bridges located in Seismic Zone 1; thus, the expected yield strengths defined in this Article should
not be used.

6.16.3 Design Requirements for Seismic Zone 1


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
For the application of the seismic design provisions in the AASHTO LRFD BDS, the routine steel
I-girder bridges covered by this Guide are assumed to be located in Seismic Zone 1. In addition to
the minimum support-length requirements at expansion bearings specified in Article 4.7.4.4 for
bridges located in Seismic Zone 1 (see the Discussion of Article 4.7.4.4 in this Guide), the
horizontal inertial forces to be transmitted through the restrained bearings are to be calculated
based on a specified factor times the vertical reaction due to the tributary permanent load defined
in this Article and the tributary live loads assumed to exist during an earthquake, as described
further in Article 3.10.9.2 (see the Discussion of Article 3.10.9.2 in this Guide). The magnitude of
live load assumed to exist at the time of the earthquake should be consistent with the value of γeq
used in conjunction with Table 3.4.1-1 (see the Discussion of Article 3.4.1 in this Guide). These
requirements are intended to provide a clear load path for the seismic forces. Note that the design
forces discussed in Article 3.10.9.2 are only for the “connection” of the superstructure to the
substructure, and the definition of “connections” (as discussed in the Commentary for Article
3.10.7.1) includes only fixed bearings, expansion bearings with either restrainers, STUs, or
dampers, and shear keys. There is no need to apply these forces to the design of pier or end cross-
frames or other elements of the superstructure.
The primary components of a routine bridge are intended to have sufficient structural capacity
from the design due to satisfaction of the nonseismic design requirements to resist the expected
lateral seismic forces in Seismic Zone 1. However, sufficient integrity and structural connectivity
still needs to be present to resist the lateral loads at restrained bearings defined in Article 3.10.9.2
in order to mobilize the lateral resistance of the main substructure elements.
Further discussion of the impacts of seismic loading on bearing design and substructure design is
considered beyond the scope of the discussion of superstructure design topics in this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 361


6.16.4 Design Requirements for Seismic Zones 2, 3, or 4

6.16.4.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article define two potential response strategies for the seismic design of
bridges located in Seismic Zones 2, 3, or 4: 1) Type A – design an elastic superstructure with a
ductile substructure; or 2) Type B – design an elastic superstructure and substructure with a fusing
mechanism, e.g. a seismic isolation device, at the interface with the superstructure and substructure
(requires Owner-agency approval). Each of these strategies is discussed further in the Commentary
for this Article. The use of one of these strategies is required for bridges located in Seismic Zones
3 or 4 and should be considered for bridges located in Seismic Zone 2. Support cross-frame
members on bridges located in Seismic Zone 3 or 4 are to be considered primary members for
seismic design.
The routine steel I-girder bridges covered by this Guide are assumed to be located in Seismic Zone
1; therefore, the provisions of this Article are not applicable.

6.16.4.2 Deck
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains provisions for the design of the concrete deck for bridges located in Seismic
Zones 2, 3, or 4 to check that the deck can serve as a horizontal diaphragm to transfer the seismic
forces to the supports. Provisions are provided to calculate the transverse seismic shear force acting
on the deck within each span of the superstructure for designs using seismic response Strategy
Type A or Type B defined in Article 6.16.4.1 (see the Discussion of Article 6.16.4.1 in this Guide).
The routine steel I-girder bridges covered by this Guide are assumed to be located in Seismic Zone
1; therefore, the provisions of this Article are not applicable.

6.16.4.3 Shear Connectors


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
This Article contains provisions for the design of stud shear connectors along the interface between
the deck and the steel girders or along the interface between the deck and the top of the support
cross-frames or diaphragms, or both, to transfer the seismic forces in bridges located in Seismic
Zones 2, 3, or 4. At support locations, shear connectors are to be designed to resist the combination
of shear and axial forces corresponding to the transverse seismic shear force within the deck
determined according to the provisions of Article 6.16.4.2 (see the Discussion of Article 6.16.4.2
in this Guide). A tension-shear interaction equation is provided in this Article to check the
resistance of the stud shear connectors to the combined shear and axial forces.

NSBA Guide to Navigating Routine Steel Bridge Design / 362


The routine steel I-girder bridges covered by this Guide are assumed to be located in Seismic Zone
1; therefore, the provisions of this Article are not applicable.

6.16.4.4 Elastic Superstructures


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
In order to achieve an elastic superstructure, this Article specifies that the support cross-frame
members or support diaphragms in a bridge located in Seismic Zone 2, 3, or 4 are to be designed
according to the applicable provisions of Article 6.7, 6.8, or 6.9 to remain elastic during a seismic
event. No other special seismic requirements are specified for these members. The support cross-
frame members or diaphragms are to be designed for the lateral force, F, specified in Article
6.16.4.2 for bridges designed using seismic response strategy Type A or Type B, as applicable (see
the Discussion of Article 6.16.4.2 in this Guide).
The routine steel I-girder bridges covered by this Guide are assumed to be located in Seismic Zone
1; therefore, the provisions of this Article are not applicable.

6.17 REFERENCES

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
This Article provides reference citations for specifications, reports, and published papers
containing more in-depth discussion of the various provisions in Section 6, including the research
or reasoning behind the development of the provisions. These useful references are cited in the
Commentary for the individual provisions, which reinforces the benefit and importance of reading
the Commentary. Not all references in this Article are applicable to the routine steel I-girder
bridges covered by this Guide.

APPENDIX A6 FLEXURAL RESISTANCE OF COMPOSITE I-SECTIONS IN


NEGATIVE FLEXURE AND NONCOMPOSITE I-SECTIONS WITH
COMPACT OR NONCOMPACT WEBS IN STRAIGHT BRIDGES

A6.1 GENERAL

Determination of applicability, Simple Span Bridges: Partially applicable.


Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 363


Discussion:
The provisions of Appendix A6 account for the ability of certain compact and noncompact
web I-sections to develop flexural resistances significantly greater than the yield moment,
My .
If permitted by the Owner-agency, the optional provisions of Appendix A6 may be applied for the
design of composite I-sections in regions of negative flexure and noncomposite I-sections in
regions of positive or negative flexure in straight bridges without holes in the tension flange whose
supports are normal or skewed not more than 20° from normal and with intermediate cross-
frame/diaphragms placed in contiguous lines parallel to the supports, for which: 1) the specified
minimum yield strengths of the flanges do not exceed 70 ksi; 2) the web satisfies the noncompact
web slenderness limit, rw, given by Eqs. A6.1-1 and A6.1-3; and 3) the flanges satisfy the
moment-of-inertia ratio given by Eq. A6.1-4. Note that in Eq. A6.1-4, the moments of inertia of
the compression and tension flange of the steel section are both taken about the vertical axis in the
plane of the web (or about the strong axis of each flange). Sections not satisfying one or more of
the above requirements must instead be proportioned according to the provisions of Article 6.10.8
(see the Discussion of Article 6.10.8 in this Guide). The reasoning behind the above limitations on
the use of Appendix A6 is described further in the Commentary for this Article and in the
Commentary for Articles 6.10.1.8 and 6.10.6.2.3. Since the types of sections that would qualify
for the use of Appendix A6 are less commonly used, the somewhat more complex provisions for
their design have been placed in an appendix in the AASHTO LRFD BDS in order to streamline
and simplify the provisions of Article 6.10.8 within the main body of the Specifications.
Sections designed according to the provisions of Appendix A6 (i.e., nonslender web sections) are
further categorized as either compact web or noncompact web I-sections (see the Discussion of
Articles A6.2.1 and A6.2.2 in this Guide).
As stated above, the provisions of Appendix A6 account for the ability of certain compact
and noncompact web I-sections to develop flexural resistances significantly greater than the
yield moment, My (see the Discussion of Article D6.2 in this Guide). As a result, the equations
giving the nominal flexural resistance in Appendix A6 are more appropriately expressed in terms
of bending moment for reasons discussed in the Commentary for Article 6.10.6.1. The provisions
of Article 6.10.8 assume that the section is a slender web section regardless of whether it is or not;
hence, the nominal flexural resistance computed according to the provisions of Article 6.10.8 is
not permitted to exceed My. The provisions of Appendix A6 also account for the contribution of
the St. Venant torsional resistance to the lateral-torsional buckling resistance of these sections,
which may be useful for compact and noncompact web sections with larger unbraced lengths,
particularly under construction conditions.
The potential benefits of the Appendix A6 provisions tend to be smaller for I-sections with webs
that approach the noncompact web slenderness limit, rw. In such cases, the somewhat simpler and
more streamlined provisions of Article 6.10.8 are recommended for use. The potential gains in
economy from using the Appendix A6 provisions increase with decreasing web slenderness.

NSBA Guide to Navigating Routine Steel Bridge Design / 364


The Engineer is strongly encouraged to utilize the provisions of Appendix A6 for design of
straight bridges with limited skew and compact or nearly compact webs (e.g., rolled-beam
sections).
Simple Span Bridges:
The provisions of Appendix A6 are not applicable at the strength limit state for the routine simple-
span bridges covered by this Guide because simple-span bridges are subject to positive flexure
only and are composite at the strength limit state. The provisions of Article A6.3.3 may potentially
be used for these bridges to optionally compute the nominal lateral-torsional buckling resistance
of the unbraced lengths of a noncomposite section with a compact or noncompact web during
construction for use in checking Eq. 6.10.3.2.1-2 in order to include the beneficial effect of the St.
Venant torsional constant, J (see the Discussion of Article 6.10.3.2.1 in this Guide).
Multi-span Continuous Rolled Beam Bridges:
Sections in regions of negative flexure in the routine multi-span continuous rolled beam bridges
covered by this Guide typically qualify as compact web sections and should be designed at the
strength limit state using the provisions of Article A6 instead of the more conservative provisions
of Article 6.10.8. The provisions of Article A6.3.3 may potentially be used for these bridges to
optionally compute the nominal lateral-torsional buckling resistance of the unbraced lengths of the
noncomposite section during construction for use in checking Eq. 6.10.3.2.1-2 in order to include
the beneficial effect of the St. Venant torsional constant, J (see the Discussion of Article 6.10.3.2.1
in this Guide).
Multi-span Continuous Plate Girder Bridges:
Sections in regions of negative flexure in the routine multi-span continuous plate girder bridges
covered by this Guide that qualify as compact or noncompact web sections can be designed at the
strength limit state using these provisions instead of the more conservative provisions of Article
6.10.8, particularly if the sections qualify as compact web sections. The provisions of Article
A6.3.3 may potentially be used for these bridges to optionally compute the nominal lateral-
torsional buckling resistance of the unbraced lengths of a noncomposite section with a compact or
noncompact web during construction for use in checking Eq. 6.10.3.2.1-2 in order to include the
beneficial effect of the St. Venant torsional constant, J (see the Discussion of Article 6.10.3.2.1 in
this Guide).
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. However,
the software currently does not include the capability to design the girders using the provisions
of Appendix A6.

NSBA Guide to Navigating Routine Steel Bridge Design / 365


A6.1.1 Sections with Discretely Braced Compression Flanges
Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.
Discussion:
Simple Span Bridges:
This Article is not applicable to the simple span routine steel I-girder bridges covered by this Guide
at the strength limit state because simple span bridges are subject to positive flexure only and are
composite at the strength limit state (i.e., the compression flange is continuously braced by the
deck). The design of the noncomposite section with a discretely braced top (compression) flange
during construction in these bridges is separately covered in Article 6.10.3.2.1 (see the Discussion
of Article 6.10.3.2.1 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This Article provides the relationship that must be satisfied at the strength limit state for compact
or noncompact web sections with discretely braced bottom (compression) flanges designed
according to the optional provisions of Appendix A6 in regions of negative flexure in multi-span
continuous steel I-girder bridges. Therefore, this Article is conditionally applicable at the strength
limit state to the design of the section in regions of negative flexure in the routine steel multi-span
continuous I-girder bridges covered by this Guide.
For composite sections in regions of negative flexure, lateral bending does not need to be
considered in the top (tension) flange at the strength limit state because the flange is continuously
supported by the concrete deck. However, since the bottom (compression) flange is discretely
braced, lateral bending must be considered in flexural resistance computations for the section
(using the “moment form” of the one-third rule flexural resistance equation). For the routine steel
I-girder bridges covered by this Guide, the only source of flange lateral bending stress to be
considered at the strength limit state is due to wind loading when checking the Strength load
combinations that include wind load effects. Amplification of fℓ in the discretely braced
compression flange will likely be required (see the Discussion of Article 6.10.1.6 in this Guide).
Note that fℓ cannot exceed 0.6Fyf after amplification. Mu and fℓ are always taken as positive in sign
in Eq. A6.1.1-1. However, when summing dead and live load moments to obtain the total factored
major-axis moment, Mu, and total factored lateral bending stresses, fℓ, to apply in the equations,
the signs of the individual stresses or moments must be considered. If there is no flange lateral
bending considered, the term fℓ Sxc drops out of the equation. For a section with a discretely braced
compression flange, the one-third rule equation must be checked separately for both flange local
buckling and lateral-torsional buckling. When lateral-torsional buckling controls, the largest
values of Mu and fℓ, as applicable, within the unbraced length must be used to check Eq. A6.1.1-1
(see the Discussion of Article 6.10.1.6 in this Guide).

NSBA Guide to Navigating Routine Steel Bridge Design / 366


The multiplication of fℓ by Sxc in Eq. A6.1.1-1 stems from the derivation of this equation. The
equation may be expressed in a stress format by dividing both sides by the corresponding elastic
section modulus, in which case, Eq. A6.1.1-1 reduces effectively to Eqs. 6.10.3.2.1-2 and 6.10.8.1.1-
1 in the limit that the web approaches its noncompact web slenderness limit, rw. The elastic section
modulus, Sxc, in Eq. A6.1.1-1 is defined as Myc/Fyc, where Myc is the yield moment with respect to
the compression flange calculated as specified in Article D6.2 (see the Discussion of Article D6.2
in this Guide) and Fyc is the specified minimum yield strength of the compression flange, so that
for a composite section with a web proportioned exactly at rw, the flexural resistance given by
Appendix A6 will be approximately the same as the flexural resistance given by Article 6.10.8.
Slight differences between the resistance predictions may occur for reasons pointed out in the
Commentary for Article A6.1.1.
The design of the noncomposite section with a discretely braced top (compression) flange in
regions of positive flexure and a discretely braced bottom (compression) flange in regions of
negative flexure during construction in these bridges is covered in Article 6.10.3.2.1 (see the
Discussion of Article 6.10.3.2.1 in this Guide).
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. However,
the software currently does not include the capability to design the girders using the provisions
of Appendix A6.

A6.1.2 Sections with Discretely Braced Tension Flanges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.
Discussion:
Simple Span Bridges:
This Article is not applicable to the simple span routine steel I-girder bridges covered by this Guide
at the strength limit state as simple span bridges are subject to positive flexure only and are
composite at the strength limit state; as implied by the title of Appendix A6 and stated in the
Discussion of Appendix A6, the provisions of Appendix A6 do not apply to composite I-sections
in positive flexure. The design of the noncomposite section with a discretely braced bottom

NSBA Guide to Navigating Routine Steel Bridge Design / 367


(tension) flange during construction in these bridges is covered in Article 6.10.3.2.2 (see the
Discussion of Article 6.10.3.2.2 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This Article provides the relationship that must be satisfied at the strength limit state for compact
or noncompact web sections with discretely braced top (tension) flanges designed according to the
optional provisions of Appendix A6 in regions of negative flexure in multi-span continuous steel
I-girder bridges. The top flange in regions of negative flexure would only be discretely braced if
the shear connectors are intentionally omitted in regions of negative flexure, which is dependent
on the preferences of the Owner-agency but not recommended, and if the Engineer deems that the
top flange is not sufficiently encased by the concrete deck. Therefore, this Article is conditionally
applicable at the strength limit state to the design of the section in regions of negative flexure in
routine steel multi-span continuous I-girder bridges covered by this Guide.
The nominal flexural resistance, Mnt, in Eq. A6.1.2-1 based on tension-flange yielding is
determined as specified in Article A6.4 (see the Discussion of Article A6.4 in this Guide). Since
the top (tension) flange is discretely braced if the conditions stated above are met, lateral bending
must be considered in flexural resistance computations for the tension flange (using the moment
form of the one-third rule flexural resistance equation). For the routine steel I-girder bridges
covered by this Guide (which are neither curved nor significantly skewed), the only source of
flange lateral bending stress to be considered at the strength limit state is wind loading occurring
under the Strength load combinations that include wind load effects. fℓ cannot exceed 0.6Fyf.
Amplification of fℓ in the tension flange is not required. fbu and fℓ are always taken as positive in
sign in Eq. A6.1.2-1. However, when summing dead and live load moments to obtain the total
factored major-axis moment, Mu, and total factored lateral bending stresses, fℓ, to apply in the
equations, the signs of the individual stresses or moments must be considered. If there is no flange
lateral bending considered, the term fℓ Sxt drops out of the equation.
The multiplication of fℓ by Sxt in Eq. A6.1.2-1 stems from the derivation of this equation. The
equation may be expressed in a stress format by dividing both sides by the corresponding elastic
section modulus, in which case, Eq. A6.1.2-1 reduces effectively to Eqs. 6.10.7.2.1-2 and 6.10.8.1.2-
1 in the limit that the web approaches its noncompact web slenderness limit, rw. The elastic section
modulus, Sxt, in Eq. A6.1.2-1 is defined as Myt/Fyt, where Myt is the yield moment with respect to
the tension flange calculated as specified in Article D6.2 (see the Discussion of Article D6.2 in
this Guide) and Fyt is the specified minimum yield strength of the tension flange, so that for a
composite section with a web proportioned exactly at rw, the flexural resistance given by
Appendix A6 will be approximately the same as the flexural resistance given by Article 6.10.8.
Slight differences between the resistance predictions may occur for reasons pointed out in the
Commentary for Article A6.1.1.
Note that when Myc is less than or equal to Myt and f is equal to zero, the flexural resistance based
on the tension flange does not control and Eq. A6.1.2-1 need not be checked.
The design of the noncomposite section with a discretely braced bottom (tension) flange in regions
of positive flexure and a discretely braced top (tension) flange in regions of negative flexure during

NSBA Guide to Navigating Routine Steel Bridge Design / 368


construction in these bridges is covered in Article 6.10.3.2.2 (see the Discussion of Article
6.10.3.2.2 in this Guide).
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. However,
the software currently does not include the capability to design the girders using the provisions
of Appendix A6.

A6.1.3 Sections with Continuously Braced Compression Flanges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article apply only to a noncomposite section at the strength limit state in
regions of positive flexure (i.e., girders in positive flexure at the strength limit state with no shear
connectors) that is designed according to the optional provisions of Appendix A6, in which the
Engineer deems that the flange is sufficiently encased by the concrete deck; this condition is not
part of the definition of the routine steel I-girder bridges covered by this Guide.

A6.1.4 Sections with Continuously Braced Tension Flanges


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.
Discussion:
This Article provides the relationship that must be satisfied at the strength limit state for compact
or noncompact web sections with continuously braced top (tension) flanges in regions of negative
flexure in multi-span continuous steel I-girder bridges designed according to the optional
provisions of Appendix A6. A continuously braced flange is anchored to the concrete deck by
shear connectors or encased in concrete.
Since the flange is continuously braced, only yielding of the flange is a concern and flange lateral
bending stresses need not be considered.

NSBA Guide to Navigating Routine Steel Bridge Design / 369


Simple Span Bridges: This Article is not applicable to the routine simple span I-girder bridges
covered by this Guide at the strength limit state as simple span bridges are subject to positive
flexure only.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This Article is applicable to the design of multi-span continuous bridges in which shear connectors
are provided in regions of negative flexure. If shear connectors are intentionally omitted in regions
of negative flexure, which is dependent on the preferences of the Owner-agency but not
recommended, and if the Engineer deems that the top flange is not sufficiently encased by the
concrete deck, then the provisions of this Article would not be applicable.
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download
from the NSBA website is also a valuable tool for the design of routine steel I-girder bridges.
However, the software currently does not include the capability to design the girders using
the provisions of Appendix A6.

A6.2 WEB PLASTIFICATION FACTORS

A6.2.1 Compact Web Sections


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:

The provisions of this Article define the limiting slenderness ratio,  pw( D ) , for a compact web
cp

section corresponding to the web slenderness, 2Dcp/tw, where Dcp is the depth of the web in
compression at the plastic moment determined according to the provisions of Article D6.3.2 (see
the Discussion of Article D6.3.2 in this Guide). The enables the section to develop the full plastic
moment resistance, Mp (see the Discussion of Article D6.1 in this Guide), provided that other
flange slenderness and lateral torsional bracing requirements are satisfied.
The upper limit of rw(Dcp/Dc) in Eq. A6.2.1-2 is to protect against extreme cases where Dc/D is
significantly less than 0.5. In such cases, Dcp/D is typically smaller than Dc/D. As such, in certain
situations, the web slenderness associated with the elastic cross-section, 2Dc/tw, may be larger than
rw, while the slenderness associated with the plastic cross-section, 2Dcp/tw, may be smaller than
 pw( D ) . In other words, the elastic web would be classified as slender at the same time the plastic
cp

web would be classified as compact. To guard against such situations and the possibility of
theoretical bend-buckling of the web prior to reaching Mp, the upper limit of rw(Dcp/Dc) is placed
on  pw( D ) .
cp

NSBA Guide to Navigating Routine Steel Bridge Design / 370


The web slenderness limit given by Eq. A6.2.1-2 accounts for the higher demands on the web
in noncomposite singly symmetric I-sections and in composite I-sections in negative flexure
with larger shape factors, Mp/My. For sections with a specified minimum yield strength of the
compression flange, Fyc, equal to 50 ksi, the limiting web slenderness based on Dcp is 91 for a
shape factor of 1.12 and 64 for a shape factor of 1.30. These limits would generally be satisfied
by rolled shapes or plate girders with proportions similar to those of a rolled shape, which
would typically be used in a shorter-span bridge (i.e., spans of about 120 ft or less).
This Article also defines the web plastification factors, Rpc and Rpt, for the compression and tension
flange, respectively, of a compact web section (Eqs. A6.2.1-5 and A6.2.1-6). The web
plastification factors are essentially effective shape factors that define a smooth linear transition in
the maximum flexural resistance between Mp and My. For a compact web section, the web
plastification factors are equivalent to the cross-section shape factors. Thus, whenever Rpc and Rpt
are used in the appropriate flexural resistance equations, the maximum flexural resistance of a
compact web section, Mmax, will equal the plastic moment, Mp. By using Rpc and Rpt in the flexural
resistance equations, separate flexural resistance equations are not required for compact and
noncompact web sections.
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. However,
the software currently does not include the capability to design the girders using the provisions
of Appendix A6.

A6.2.2 Noncompact Web Sections


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article define the limiting slenderness ratio, λrw, for a noncompact web
section corresponding to the web slenderness, 2Dc/tw, where Dc is the elastic depth of the web in
compression determined according to the provisions of Article D6.3.1 (see the Discussion of
Article D6.3.1 in this Guide). Sections exceeding the web slenderness limit for a compact web section
given by Eq. A6.2.1-1 (see the Discussion of Article A6.2.1 in this Guide), but with a web slenderness
less than or equal to λrw, are termed noncompact web sections, which have a nominal flexural resistance
at the strength limit state that linearly transitions from the plastic moment, Mp, to the moment at first
yield, My (see the Discussions of Articles D6.1 and D6.2 in this Guide) as a function of the web
slenderness. For sections with a specified minimum yield strength of the compression flange,
Fyc, equal to 50 ksi, the upper and lower limits of λrw are 111 and 137. Plate-girder sections in a
medium-span bridge (e.g., spans in the 120-ft to 150-ft range) may potentially qualify as noncompact

NSBA Guide to Navigating Routine Steel Bridge Design / 371


web sections. Some larger rolled-beam sections may also potentially qualify as noncompact web
sections.
This Article also defines the web plastification factors, Rpc and Rpt, for the compression and tension
flange, respectively, of a noncompact web section (Eqs. A6.2.2-4 and A6.2.2-5). Eqs. A6.2.2-4
and A6.2.2-5 define the linear transition in the maximum potential flexural resistance, Mmax, of a
noncompact web section from Mp to My as a function of the web slenderness. As 2Dc/tw approaches
the noncompact web section limit, rw, the web plastification factors approach values equal to the
hybrid factor, Rh (see the Discussion of Article 6.10.1.10.1 in this Guide), and therefore, Mmax
within the appropriate flexural resistance equations approaches a limiting value of RhMyc or RhMyt,
as applicable. As 2Dcp/tw approaches the compact web section limit, pw(Dcp) (see the Discussion of
Article A6.2.1 in this Guide), the web plastification factors approach the cross-section shape
factors (Eqs. A6.2.1-5 and A6.2.1-6), and therefore, Mmax within the appropriate flexural resistance
equations approaches a limiting value of Mp. By using Rpc and Rpt in the flexural resistance
equations, separate flexural resistance equations are not required for compact and noncompact web
sections.
Upper limits of Mp/Myc and Mp/Myt are placed on Rpc and Rpt, respectively, in Eqs. A6.2.2-4 and
A6.2.2-5. These upper limits will limit the larger of the base resistances, RpcMyc or RptMyt, to Mp
for the rare case of an extremely singly symmetric section in which Myc or Myt is greater than Mp.
The flange-proportioning limit given by Eq. 6.10.2.2-4 (see the Discussion of Article 6.10.2.2 in
this Guide) will generally tend to prevent the use of such sections.

Eq. A6.2.2-6 converts the compact web section slenderness ratio,  pw( D ) , defined in terms of Dcp
cp

to a value that can be used consistently in Eqs. A6.2.2-4 and A6.2.2-5 with the web slenderness,
w, which is expressed in terms of Dc.
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. However,
the software currently does not include the capability to design the girders using the provisions
of Appendix A6.

A6.3 FLEXURAL RESISTANCE BASED ON THE COMPRESSION FLANGE

A6.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges:

NSBA Guide to Navigating Routine Steel Bridge Design / 372


Discussion:
This Article directs the Engineer to the Articles within Appendix A6 containing the equations
necessary to compute the nominal flexural resistance, Mnc, of a section with a discretely braced
compression flange based on flange local buckling (see the Discussion of Article A6.3.2 in this
Guide) or lateral-torsional buckling (see the Discussion of Article A6.3.3 in this Guide) for use in
Eq. A6.1.1-1 at the strength limit state (see the Discussion of Article A6.1.1 in this Guide), or
potentially in Eq. 6.10.3.2.1-2 for the noncomposite section during construction (see the
Discussion of Article 6.10.3.2.1 in this Guide). The equations must be satisfied for both flange
local buckling and lateral-torsional buckling.
The Commentary for Article 6.10.8.2.1 includes presentation of the “basic form of all I-section
compression-flange flexural resistance equations.” Designers are strongly encouraged to
familiarize themselves with the concepts presented in this Commentary and the associated
graph in Figure C6.10.8.2.1-1. Possessing a clear understanding of these fundamental
concepts is invaluable for understanding the associated provisions of the AASHTO LRFD
BDS.
Simple Span Bridges:
This Article is applicable for the simple span steel I-girder bridges covered by this Guide only if
the noncomposite section qualifies as a compact web or noncompact web section and the
provisions of Article A6.3.3 are used to compute Mnc for lateral-torsional buckling during
construction for use in checking Eq. 6.10.3.2.1-2 to account for the beneficial effect of the St.
Venant torsional constant, J (see the Discussion of Article 6.10.3.2.1 in this Guide).
This Article is not applicable to these bridges at the strength limit state as simple spans are subject
to positive flexure only and the top (compression) flange is continuously braced by the concrete
deck.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
For the routine multi-span continuous rolled beam and plate girder bridges covered by this Guide,
this Article is applicable for determining Mnc for sections designed using the provisions of
Appendix A6 and with a discretely braced bottom (compression) flange in regions of negative
flexure at the strength limit state for use in Eq. A6.1.1-1 (see the Discussion of Article A6.1.1 in
this Guide).
The Article is also applicable if the noncomposite section in regions of positive and negative
flexure qualifies as a compact web or noncompact web section and the provisions of Article A6.3.3
are used to compute Mnc for lateral-torsional buckling during construction for use in checking Eq.
6.10.3.2.1-2 to account for the beneficial effect of the St. Venant torsional constant, J (see the
Discussion of Article 6.10.3.2.1 in this Guide).
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,

NSBA Guide to Navigating Routine Steel Bridge Design / 373


Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. However,
the software currently does not include the capability to design the girders using the provisions
of Appendix A6.

A6.3.2 Local Buckling Resistance


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article are used to compute the nominal flexural resistance, Mnc, of a section
with a compact or noncompact web and a discretely braced compression flange based on flange
local buckling (FLB) for use in Eq. A6.1.1-1 at the strength limit state (see the Discussion of
Article A6.1.1 in this Guide).
FLB is a limit state of buckling of a compression flange within a cross-section and is a function of
the compression-flange slenderness, λf = bfc/2tfc. For determining the FLB resistance (refer to
Figure C6.10.8.2.1-1), λpf locates Anchor Point 1 that separates sections with compact flanges from
sections with noncompact flanges (see Table C6.10.8.2.2-1). A member with a compression-flange
slenderness at or below the compact flange limit, λpf, is able to achieve the so-called “plateau
strength” or maximum potential FLB resistance (Mmax in Figure C6.10.8.2.1-1) of RpMyc, which is
independent of the compression-flange slenderness. λrf locates Anchor Point 2 that separates
sections with noncompact flanges from sections with slender flanges and is the point where the
inelastic and elastic FLB resistances are the same (with the resistance at this point assumed to be
RbFyrSxc). The inelastic FLB resistance of a noncompact flange is treated as a linear function of the
compression-flange slenderness. An elastic FLB equation for slender flanges is not provided in the
specifications because for most practical bridge-girder sections, including sections used in the
routine I-girder bridges covered by this Guide (i.e., with Fyc ≤ 90 ksi), elastic FLB will not control
as λf is limited to a practical maximum value of 12.0 (see the Discussion of Article 6.10.2.2 in this
Guide). The FLB resistance for moment gradient cases is treated the same as that for the case of
uniform major-axis bending; i.e., the relatively minor influence of moment-gradient effects on the
FLB resistance is neglected.
The equation for the anchor point, λrf, includes a flange local buckling coefficient, kc. Eq. A6.3.2-
6 provides an expression for kc for built-up sections. This expression accounts for the fact that
thinner webs in built-up sections tend to offer less rotational restraint to prevent flange local
buckling. The calculated value of kc from Eq. A6.3.2-6 must fall between the range of 0.35 and
0.76. The upper-bound value of 0.76 corresponds to the traditional value that has been assumed
for rolled shapes in the AISC Specification for Structural Steel Buildings. The lower bound kc
value of 0.35 is conservatively assumed for all sections in Eq. 6.10.8.2.2-5 for λrf, which is assumed
to apply only to slender web sections.

NSBA Guide to Navigating Routine Steel Bridge Design / 374


For design checks where the flexural resistance is based on FLB, Mu and fℓ in in Eq. A6.1.1-1 are
determined as the moment and stress at the section under consideration (see the Discussion of
Article 6.10.1.6 in this Guide).
Simple Span Bridges:
The provisions of this Article are not applicable to the routine simple span bridges covered by this
Guide at the strength limit state because simple spans are subject to positive flexure and and the
top (compression) flange is continuously braced by the concrete deck. These provisions are also
not applicable to these bridges during construction because the FLB resistance of the noncomposite
section during construction is to be checked using Eq. 6.10.3.2.1-2 in conjunction with the
equations of Article 6.10.8.2.2 (see the Discussions of Articles 6.10.3.2.1 and 6.10.8.2.2 in this
Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
The provisions of this Article are conditionally applicable for the routine steel multi-span
continuous I-girder bridges covered by this Guide to calculate the nominal FLB resistance of
sections with a compact or noncompact web in regions of negative flexure at the strength limit
state with a discretely braced bottom (compression) flange for use in Eq. A6.1.1-1 (if the section
is designed using the provisions of Appendix A6).
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download
from the NSBA website is also a valuable tool for the design of routine steel I-girder bridges.
However, the software currently does not include the capability to design the girders using
the provisions of Appendix A6.

A6.3.3 Lateral Torsional Buckling Resistance


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article are used to compute the nominal flexural resistance, Mnc, of a section
with a compact or noncompact web and a discretely braced compression flange based on lateral-
torsional buckling (LTB) for use in Eq. A6.1.1-1 or 6.10.3.2.1-2, as applicable (see the Discussion
of Articles A6.1.1 and 6.10.3.2.1 in this Guide).
LTB is a limit state of buckling of an unbraced length involving lateral deflection and twist and is
a function of the unbraced length, Lb. For determining the LTB resistance (refer to Figure
C6.10.8.2.1-1), Lp locates Anchor Point 1 that separates compact unbraced lengths from
noncompact unbraced lengths. A member braced at or below the compact unbraced length limit is

NSBA Guide to Navigating Routine Steel Bridge Design / 375


able to achieve the so-called “plateau strength” or maximum potential LTB resistance (Mmax in
Figure C6.10.8.2.1-1) of RpcMyc under uniform bending, which is independent of the unbraced
length. Note that in many cases, it will not be economical to brace the girder at a distance equal to
Lp or below in order to reach Mmax, particularly under uniform bending conditions for which Cb is
equal to 1.0. Lr locates Anchor Point 2 that separates sections with noncompact unbraced lengths
from sections with slender unbraced lengths and is the point where the inelastic and elastic LTB
resistances are the same (with the resistance at this point assumed to be RbFyrSxc). The inelastic
LTB resistance of a noncompact unbraced length is treated as a linear function of the unbraced
length. Unbraced lengths greater than Lr are termed slender unbraced lengths and their resistance
is controlled by elastic LTB. LTB in the elastic range is of primary importance for relatively
slender girders braced at longer than normal intervals, which most typically occurs during a
temporary construction condition. Eq. A6.3.3-8 for the elastic LTB stress, Fcr, is the exact beam-
theory solution for the elastic LTB resistance of a doubly symmetric I-section under uniform
bending (when load-height effects are not considered). Unlike their counterparts in Article
6.10.8.2.3, which assume slender-web behavior, the equations for Fcr and Lr in Appendix A6
include the St. Venant torsional constant J, which is appropriate for stockier compact web and
noncompact web sections that are generally not subject to significant web distortion. Further
details on these LTB resistance equations are provided in the Commentary for this Article.
This Article is conditionally applicable for routine steel multi-span continuous I-girder bridges,
and only partially applicable for routine simple span I-girder bridges covered by this Guide, as
described further in the Discussion of Article A6.3.1 in this Guide.
The solid curve in Figure C6.10.8.2.1-1 is for uniform bending represented by the equations given
in this Article. The dashed curve in Figure C6.10.8.2.1-1 shows the solid curve scaled by the
moment-gradient modifier, Cb, under moment-gradient conditions, which can result in the plateau
strength (Mmax) for lateral-torsional buckling to be reached at significantly larger unbraced lengths
under moment-gradient conditions when the effects of the moment gradient are included in
determining the limits on the unbraced length, Lb. Refer to Article D6.4.2 for the appropriate
equations to use under these conditions (i.e., the equations representing the dashed curve when Cb
is calculated and is greater than 1.0), which is strongly encouraged (see the Discussion of Article
D6.4.2 in this Guide). The calculation of Cb is discussed further below.
For design checks where the flexural resistance is based on LTB, Mu, fbu and fℓ in Eqs. A6.1.1-1
and 6.10.3.2.1-2 are to be taken as the largest values throughout the unbraced length in the flange
under consideration (see the Discussion of Article 6.10.1.6 in this Guide).
For unbraced lengths containing a transition to a smaller section at a distance less than or equal to
20 percent of the unbraced length from the brace point with the smaller moment, Article A6.3.3
permits the LTB resistance to be determined assuming the transition to the smaller section does
not exist. This is provided the lateral moment of inertia of the flange or flanges of the smaller
section is also equal to or larger than one-half of the corresponding value in the larger section. The
moment gradient modifier, Cb, may be applied in this case. Otherwise, the LTB resistance is to be
taken as the smallest resistance within the unbraced length and Cb must be taken equal to 1.0. The
resulting resistance must not be exceeded anywhere along the unbraced length. This conservative
approximation is based on replacing the nonprismatic member with an equivalent prismatic

NSBA Guide to Navigating Routine Steel Bridge Design / 376


member. To avoid significant reductions in the LTB resistance, it is desirable to consider locating
flange transitions (satisfying the preceding moment of inertia requirement) within 20 percent of
the unbraced length from the brace point with the smaller moment, where practical.

For simple span bridges, the moment-gradient modifier, Cb, may conservatively be taken equal to
1.0 when checking LTB of the critical noncomposite section in regions of positive flexure during
construction; otherwise, consult Figure C6.4.10 for the appropriate calculation of Cb in these
regions (see the Discussion of Article C6.4.10 in this Guide).
For multi-span continuous bridges, it is strongly recommended that as a minimum the moment-
gradient modifier, Cb, be calculated when checking LTB of the first unbraced length on either side
of the interior piers as described below, specifically for prismatic unbraced lengths or for
nonprismatic unbraced lengths satisfying the 20 percent rule described above. The unbraced
lengths on either side of the pier should be checked to determine which side will yield the lower
value of Cb. Cb may conservatively be taken equal to 1.0 when checking LTB of the critical
noncomposite section in regions of positive flexure during construction; otherwise, consult Figure
C6.4.10 for the appropriate calculation of Cb in these regions (see the Discussion of Article C6.4.10
in this Guide).
For multi-span continuous bridges, Cb should be calculated for the first unbraced length on either
side of the interior piers using Eq. A6.3.3-7 with M1 taken equal to M0 (Eq. A6.3.3-11). In
Appendix A6, where the nominal flexural resistance is permitted to exceed the moment at first
yield for compact and noncompact web sections, the major-axis bending moments are used to
calculate Cb since the effect of applying the bending moments to different sections is less critical
in these cases. The factored moments, M2 and M0, at each end of the unbraced length are taken as
positive when they cause compression in the flange under consideration and negative when they
cause tension in Eq. A6.3.3-7. The provisions of Article D6.4.2 should then be employed to
determine the shift in the anchor point, Lp, and the corresponding nominal LTB resistance (see the
Discussion of Article D6.4.2 in this Guide).
It is convenient and always conservative to use the critical moment envelope values to calculate
Cb, particularly since concurrent moment values at the brace points are not normally tracked in the
analysis. Further information on Cb may be found in the Commentary for Article 6.10.8.2.3.
The effective length factor, K, for LTB is assumed equal to 1.0 in the equations of this Article. The
potential adjustments to K discussed in the Commentary for Article 6.10.8.2.3 should not be
necessary for the routine steel I-girder bridges covered by this Guide.
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,
Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.

NSBA Guide to Navigating Routine Steel Bridge Design / 377


The NSBA's LRFD Simon line-girder analysis and design software available for free download
from the NSBA website is also a valuable tool for the design of routine steel I-girder bridges.
However, the software currently does not include the capability to design the girders using the
provisions of Appendix A6.

A6.4 FLEXURAL RESISTANCE BASED ON TENSION FLANGE YIELDING

Determination of applicability, Simple Span Bridges: Not applicable.


Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
Simple Span Bridges:
This Article is not applicable to the routine simple span I-girder bridges covered by this Guide at
the strength limit state because simple span bridges are subject to positive flexure only. The design
of the noncomposite section with a discretely braced bottom (tension) flange during construction
in these bridges is covered in Article 6.10.3.2.2 (see the Discussion of Article 6.10.3.2.2 in this
Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
The provisions of this Article are used to compute the nominal flexural resistance, Mnt, of a section
with a discretely braced top (tension) flange at the strength limit state in regions of negative flexure
in multi-span continuous steel I-girder bridges for use in Eq. A6.1.2-1 (see the Discussion of
Article A6.1.2 in this Guide). The nominal flexural resistance is based only on nominal yielding
because flange local buckling and lateral-torsional buckling are not a consideration for flanges in
tension. The top flange in regions of negative flexure would only be discretely braced if the shear
connectors are intentionally omitted in regions of negative flexure, which is dependent on the
preferences of the Owner-agency but not recommended, and if the Engineer deems that the top
flange is not sufficiently encased by the concrete deck. Therefore, this Article is conditionally
applicable at the strength limit state to the design of the section in regions of negative flexure in
routine steel multi-span continuous I-girder bridges, as described further in the Discussion of
Article A6.1.2 in this Guide.
Eq. A6.4-1 represents a linear transition in the flexural resistance between Myt and Mp as a function
of the web slenderness, 2Dc/tw. As the web slenderness approaches the noncompact web section
limit, rw (see the Discussion of Article A6.2.2 in this Guide), Eq. A6.4-1 approaches the nominal
flexural resistance based on tension flange yielding equal to RhFyt.
For sections in which Myt > Myc, Eq. A6.4-1 does not control and need not be checked.
For further information on the provisions and application of Appendix A6, consult Section
6.5.6.2.3 of the Reference Manual for NHI Course 130081, Load and Resistance Factor Design
(LRFD) for Highway Bridge Superstructures. For design examples illustrating strength limit state
design flexure checks using Appendix A6, consult the FHWA’s Steel Bridge Design Handbook,

NSBA Guide to Navigating Routine Steel Bridge Design / 378


Design Example 2A, Two-Span Continuous Straight Composite Steel I-Girder Bridge, and Design
Example 2B, Two-Span Continuous Straight Composite Steel Wide-Flange Beam Bridge.
The NSBA's LRFD Simon line-girder analysis and design software available for free download from
the NSBA website is also a valuable tool for the design of routine steel I-girder bridges. However,
the software currently does not include the capability to design the girders using the provisions
of Appendix A6.

APPENDIX B6 MOMENT REDISTRIBUTION FROM INTERIOR-PIER I-SECTIONS


IN STRAIGHT CONTINUOUS-SPAN BRIDGES

B6.1 GENERAL

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
The provisions of this optional Appendix B6 may be used to calculate redistribution moments due
to the effects of yielding from the interior-pier sections of straight continuous-span I-section
flexural members at the service and/or strength limit states. The provisions of Appendix B6 are
not applicable to simple-span bridges since the application of these provisions necessarily implies
the formation of a plastic hinge in the superstructure at the strength limit state. The formation of a
plastic hinge in a simple-span structure would represent a collapse mechanism. Simple-span
bridges also do not have interior piers.
These provisions provide an approximate procedure and a refined method to calculate the
redistribution moments, which both utilize elastic moment envelopes and do not require the direct
use of inelastic analysis methods. The provisions may only be applied to straight continuous-span
I-section members satisfying specified limitations (see the Discussion of Article B6.2 in this
Guide). The provisions of Articles B6.3 and B6.4 are used to calculate the redistribution moments
using the approximate procedure and the provisions of Article B6.6 are used to calculate the
redistribution moments using the refined method (see the Discussion of these Articles in this
Guide).
Allowing redistribution of negative moments in multi-span continuous steel I-girder bridges can
potentially produce more economical designs, but the associated analysis and design
considerations are unfamiliar to most Engineers and most Owner-agencies currently do not permit
or encourage the use of moment redistribution methods. As a result, the use of moment
redistribution methods has been specifically excluded from the scope of this Guide for the design
of routine steel I-girder bridges.
See the Discussion of Article 4.6.4.1 in this Guide for a basic explanation of moment redistribution
methods and their associated advantages and disadvantages. For further information on the
provisions of Appendix B6, consult Section 6.5.6.6 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures, as well
as FHWA’s Steel Bridge Design Handbook, Design Example 2A, Two-Span Continuous Straight
Composite Steel I-Girder Bridge, Design Example 2B, Two-Span Continuous Straight Composite
Steel Wide-Flange Beam Bridge.

NSBA Guide to Navigating Routine Steel Bridge Design / 379


B6.2 SCOPE

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
The provisions of this Article define the scope of the straight continuous-span I-section members
to which the provisions of Appendix B6 may be applied. Specifically, the provisions of Appendix
B6 may be applied only to straight continuous I-section members whose support lines are not
skewed more than 10 degrees from normal and along which there are no staggered cross-frames.
Cross-sections throughout the unbraced lengths immediately adjacent to interior-pier sections from
which moments are redistributed must also satisfy certain specified restrictions defined in
subsequent Articles to demonstrate adequate robustness to redistribute the moments. If the
approximate procedure is used to calculate the redistribution moments, the unbraced lengths
adjacent to interior-pier sections must satisfy these restrictions. If the refined method is used to
calculate the redistribution moments, the unbraced lengths adjacent to interior-pier sections need
not satisfy these restrictions; however, moments may not be redistributed from interior-pier
sections that do not satisfy these restrictions.
Although members in routine steel I-girder bridges covered by this Guide may satisfy these
restrictions, most Owner-agencies currently do not permit or encourage the use of moment
redistribution methods. As a result, the use of moment redistribution methods has been specifically
excluded from the scope of this Guide for the design of routine steel I-girder bridges.
See the Discussion of Article 4.6.4.1 in this Guide for a basic explanation of moment redistribution
methods and their associated advantages and disadvantages. For further information on the
provisions of Appendix B6, consult Section 6.5.6.6 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures, as well
as FHWA’s Steel Bridge Design Handbook, Design Example 2A, Two-Span Continuous Straight
Composite Steel I-Girder Bridge, Design Example 2B, Two-Span Continuous Straight Composite
Steel Wide-Flange Beam Bridge.

B6.2.1 Web Proportions


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article define three web proportioning limits that must be satisfied within
the unbraced lengths immediately adjacent to interior-pier sections from which moments are
redistributed in order to apply the optional moment redistribution provisions of Appendix B6.
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.2.2 Compression Flange Proportions


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 380


Discussion:
The provisions of this Article define two compression flange proportioning limits that must be
satisfied within the unbraced lengths immediately adjacent to interior-pier sections from which
moments are redistributed in order to apply the optional moment redistribution provisions of
Appendix B6.
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.2.3 Section Transitions


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provision in this Article requires unbraced lengths immediately adjacent to interior-pier
sections from which moments are redistributed to be prismatic in order to apply the optional
moment redistribution provisions of Appendix B6.
This provision is not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.2.4 Compression Flange Bracing


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provision in this Article defines a compression flange bracing limitation that must be satisfied
within the unbraced lengths immediately adjacent to interior-pier sections from which moments
are redistributed in order to apply the optional moment redistribution provisions of Appendix B6.
This provision is not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.2.5 Shear
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provision in this Article limits the maximum factored shear within the unbraced lengths
immediately adjacent to interior-pier sections from which moments are redistributed to the shear-
yield or shear-buckling resistance in order to apply the optional moment redistribution provisions
of Appendix B6.

NSBA Guide to Navigating Routine Steel Bridge Design / 381


This provision is not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.2.6 Bearing Stiffeners


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provision in this Article requires bearing stiffeners to be placed at interior-pier sections from
which moments are redistributed in order to apply the optional moment redistribution provisions
of Appendix B6.
This provision is not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.3 SERVICE LIMIT STATE

B6.3.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provision in this Article specifies the Service II load combination is to be used to check the
service limit state requirements specified in subsequent Articles to control permanent deflections
of the member after moment redistribution.
This provision is not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.3.2 Flexure

B6.3.2.1 Adjacent to Interior-Pier Sections


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article indicate that the service limit state flange stress limits in Article
6.10.4.2 (see the Discussion of Article 6.10.4.2 in this Guide) need not be checked after moment
redistribution within the regions extending in each adjacent span from interior-pier sections
satisfying the requirements of Article B6.2 (see the Discussion of Article B6.2 in this Guide) to
the nearest flange transition or point of dead-load contraflexure, whichever is closest. The web-bend
buckling check given by Eq. 6.10.4.2.2-4 is the only check that must be satisfied within these
regions. This check is to be based on the elastic moments before moment redistribution.

NSBA Guide to Navigating Routine Steel Bridge Design / 382


These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1 in this Guide).

B6.3.2.2 At All Other Locations


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article indicate that in order to control permanent deflections of the member
after moment redistribution, the service limit state flange-stress limits in Article 6.10.4.2 (see the
Discussion of Article 6.10.4.2 in this Guide) are only to be imposed in each adjacent span at
sections outside the nearest flange transition location or point of permanent-load contraflexure,
whichever is closest to the interior support under consideration. The appropriate flexural stresses
due to the redistribution moments are to be added to the flexural stresses due to the Service II
elastic moments prior to making these checks. The redistribution moments are to be computed
according to the provisions of Article B6.3.3 (see the Discussion of Article B6.3.3 in this Guide).
The composite section properties to be used in computing the stresses in the steel section and
concrete deck due to the redistribution moments are also defined in this Article.
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.3.3 Redistribution Moments

B6.3.3.1 At Interior-Pier Sections


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to calculate the redistribution moment for the Service II
loads at each interior-pier section where the Service II flange stress limits are not checked as
permitted in Article B6.3.2.1 (see the Discussion of Article B6.3.2.1 in this Guide). In the
approximate approach, the redistribution moment is calculated utilizing a negative-flexure
effective plastic moment for the service limit state determined as specified in Article B6.5 (see the
Discussion of Article B6.5 in this Guide).
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.3.3.2 At All Other Locations


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 383


Discussion:
The provisions of this Article specify how to construct the Service II redistribution-moment
diagram using the redistribution moments at the adjacent interior-pier sections. The redistribution-
moment diagram is used to compute the Service II redistribution moments at locations other than
at the interior piers. These moments are held in equilibrium by the support reactions and remain in
the structure after the live loads are removed.
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.4 STRENGTH LIMIT STATE

B6.4.1 Flexural Resistance

B6.4.1.1 Adjacent to Interior-Pier Sections

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
The provision in this Article indicates that the flexural resistances at the strength limit state of
sections within the unbraced lengths immediately adjacent to interior-pier sections satisfying the
requirements of Article B6.2 (see the Discussion of Article B6.2 in this Guide) from which
moments are redistributed need not be checked.
This provision is not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.4.1.2 At All Other Locations

Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.


Discussion:
The provisions of this Article indicate that sections at all other locations, i.e., at locations other
than those defined in Article B6.4.1.1 (see the Discussion of Article B6.4.1.1 in this Guide) must
satisfy the strength limit state provisions of Articles 6.10.7, 6.10.8.1, or A6.1, as applicable, after
moment redistribution (see the Discussion of these Articles in this Guide). The appropriate
redistribution moments are to be added to the factored elastic moments at the strength limit state
prior to making the design checks. The redistribution moments are to be computed according to
the provisions of Article B6.4.2 (see the Discussion of Article B6.4.2 in this Guide). The composite
section properties to be used in computing the stresses in the steel section and concrete deck due
to the redistribution moments are also defined in this Article.

NSBA Guide to Navigating Routine Steel Bridge Design / 384


These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.4.2 Redistribution Moments

B6.4.2.1 At Interior-Pier Sections


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to calculate the redistribution moment for the strength limit
state at each interior-pier section where the flexural resistance is not checked as permitted in
Article B6.4.1.1 (see the Discussion of Article B6.4.1.1 in this Guide). In the approximate
approach, the redistribution moment is calculated utilizing a negative-flexure effective plastic
moment for the strength limit state determined as specified in Article B6.5 (see the Discussion of
Article B6.5 in this Guide).
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.4.2.2 At All Other Sections


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provision in this Article refers to the provisions of Article B6.3.3.2 for the procedure to
construct the strength limit state redistribution-moment diagram using the redistribution moments
at the adjacent interior-pier sections (see the Discussion of Article B6.3.3.2 in this Guide). The
redistribution-moment diagram is used to compute the strength limit state redistribution moments
at locations other than at the interior piers. These moments are held in equilibrium by the support
reactions and remain in the structure after the live loads are removed.
This provision is not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.5 EFFECTIVE PLASTIC MOMENT

B6.5.1 Interior-Pier Sections with Enhanced Moment-Rotation Characteristics


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to determine the negative-flexure effective plastic moment
for the calculation of the redistribution moments at the service and strength limit states (see the
Discussion of Articles B6.3.3.1 and B6.4.2.1 in this Guide) for interior-pier sections satisfying the

NSBA Guide to Navigating Routine Steel Bridge Design / 385


requirements of Article B6.2 (see the Discussion of Article B6.2 in this Guide) and with enhanced
moment-rotation characteristics; that is, sections with transverse stiffeners spaced at D/2 or less
over a minimum distance of D/2 on each side of the interior-pier section, or sections with so-called
“ultracompact webs”.
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.5.2 All Other Interior-Pier Sections


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to determine the negative-flexure effective plastic moment
for the calculation of the redistribution moments at the service and strength limit states (see the
Discussion of Articles B6.3.3.1 and B6.4.2.1 in this Guide) for interior-pier sections satisfying the
requirements of Article B6.2 (see the Discussion of Article B6.2 in this Guide), but not satisfying
the requirements of Article B6.5.1 that provide for enhanced moment-rotation characteristics (see
the Discussion of Article B6.5.1 in this Guide).
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.6 REFINED METHOD

B6.6.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article permit continuous-span I-section members satisfying the provisions
of Article B6.2 (see the Discussion of Article B6.2 in this Guide) to alternatively be proportioned
using a refined method in which a direct shakedown analysis is conducted at the service and/or
strength limit states. In this analysis, the redistribution moments are determined by the simultaneous
satisfaction of rotational continuity and specified inelastic moment-rotation relationships at interior-
pier sections from which moments are redistributed. The elastic moment envelope due to the factored
loads is used in the analysis. Sections adjacent to interior piers from which moments are
redistributed are to satisfy the requirements of Article B6.3.2.1 at the service limit state and
Article B6.4.1.1 at the strength limit state. Other sections are to satisfy the applicable provisions
of Articles 6.10.4.2, 6.10.7, 6.10.8.1, or A6.1 after a solution is found (see the Discussion of these
Articles in this Guide).
If software that handles this type of calculation along with the determination of the elastic moment
envelopes does not exist, significant manual work is required in conducting the analysis
calculations. The Engineer can gain some additional benefit when using a direct shakedown

NSBA Guide to Navigating Routine Steel Bridge Design / 386


analysis since the restriction that interior-pier sections within the member satisfy the requirements
of Article B6.2.1 (i.e., the web-slenderness requirements) is waived. Also, the directly calculated
inelastic rotations at the interior-pier sections will tend to be smaller than the upper-bound values
that the equations in Articles B6.3 through B6.5 are based upon (see the Discussions of Articles
B6.3 through B6.5 in this Guide).
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

B6.6.2 Nominal Moment-Rotation Curves


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions in this Article provide lower-bound nominal moment-rotation curves that may be
used in the refined method utilizing a direct shakedown analysis (see the Discussion of Article
B6.6.1 in this Guide) at interior-pier sections satisfying the requirements of Article B6.2 (see the
Discussion of Article B6.2 in this Guide). Curves are provided for interior-pier sections satisfying
the requirements of Article B6.5.1 to provide enhanced moment-rotation characteristics (see the
Discussion of Article B6.5.1 in this Guide), and for other interior-pier sections not satisfying those
requirements. Interior-pier sections not satisfying the requirements of Article B6.2 are assumed to
remain elastic in the analysis, and are to satisfy the provisions of Articles 6.10.4.2, 6.10.8.1, or
Article A6.1, as applicable, after a solution is found (see the Discussion of these Articles in this
Guide).
These provisions are not applicable since the use of moment redistribution methods has been
specifically excluded from the scope of this Guide for the design of routine steel I-girder bridges
(see the Discussion of Article B6.1in this Guide).

APPENDIX C6 BASIC STEPS FOR STEEL BRIDGE SUPERSTRUCTURES

C6.1 GENERAL

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
This Article serves as an introduction to an outline contained in Articles C6.2 and C6.3 (see the
Discussion of Articles C6.2 and C6.3 in this Guide) that provides a generic overview of the design
process for steel bridges. The outline is not fully complete and should not be used as a substitute
for a working knowledge of the provisions of Section 6 of the AASHTO LRFD BDS, but much of
this outline can still be helpful in providing the basic design steps for the routine steel I-girder
bridges covered by this Guide. For each design step, the outline refers to the specific Article(s)
within Section 6 that contain the design provisions relevant to that step. This outline was used in
the development of the Design Task Quick Links provided with this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 387


C6.2 GENERAL CONSIDERATIONS

Determination of applicability, All Routine Steel I-girder Bridges: Applicable.


Discussion:
The portions of the outline described in Article C6.1 (see the Discussion of Article C6.1 in this
Guide) dealing with general considerations such as the general design philosophy, limit states, and
design and location features are provided in this Article, and are considered applicable to the
routine steel I-girder bridges covered by this Guide.

C6.3 SUPERSTRUCTURE DESIGN

Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.


Discussion:
The portions of the outline described in Article C6.1 (see the Discussion of Article C6.1 in this
Guide) dealing with steel superstructure design are provided in this Article, and are considered
partially applicable as only the design steps pertinent to the design of routine steel I-girder bridges
covered by this Guide are applicable.
Note that the Design Tasks Quick Links provided near the beginning of this Guide are based
directly on the outline presented in this Article.

C6.4 FLOWCHARTS FOR FLEXURAL DESIGN OF I-SECTION MEMBERS

C6.4.1 Flowchart for LRFD Article 6.10.3


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Article 6.10.3 dealing with the design for constructibility (see the Discussion of Article 6.10.3 in
this Guide). This flowchart is applicable to the routine steel I-girder bridges covered by this Guide
and is strongly recommended for use in conjunction with this Guide.

C6.4.2 Flowchart for LRFD Article 6.10.4


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Article 6.10.4 dealing with the design for the service limit state (see the Discussion of Article
6.10.4 in this Guide), and is strongly recommended for use in conjunction with this Guide. The
portions of the flowchart dealing with optional moment redistribution and shored construction (see
the Discussion of Articles B6.1 and Article 6.10.1.1.1a in this Guide) are not applicable to the
routine steel I-girder bridges covered by this Guide.

NSBA Guide to Navigating Routine Steel Bridge Design / 388


C6.4.3 Flowchart for LRFD Article 6.10.5
Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Articles 6.5.3 and 6.6 dealing with the design for the fatigue and fracture limit state (see the
Discussion of Articles 6.5.3 and 6.6 in this Guide). This flowchart is applicable to the routine steel
I-girder bridges covered by this Guide and is strongly recommended for use in conjunction with
this Guide.

C6.4.4 Flowchart for LRFD Article 6.10.6


Determination of applicability, All Routine Steel I-girder Bridges: Applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Article 6.10.6 dealing with the design for the strength limit state (see the Discussion of Article
6.10.6 in this Guide), and is strongly recommended for use in conjunction with this Guide. The
portion of the flowchart dealing with optional moment redistribution (see the Discussion of Article
B6.1 in this Guide) is not applicable to the routine steel I-girder bridges covered by this Guide.
Only the portion of the flowchart dealing with the design of composite sections in positive flexure
is applicable to the design of routine simple span I-girder bridges covered by this Guide.

C6.4.5 Flowchart for LRFD Article 6.10.7


Determination of applicability, Simple Span Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Article 6.10.7 dealing with the design of composite sections in positive flexure at the strength limit
state (see the Discussion of Article 6.10.7 in this Guide). This flowchart is applicable to the routine
steel I-girder bridges covered by this Guide, except as noted below for simple span bridges, and is
strongly recommended for use in conjunction with this Guide.
Simple Span Bridges:
The portion of the flowchart dealing with continuous spans is not applicable.

C6.4.6 Flowchart for LRFD Article 6.10.8


Determination of applicability, Simple Span Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Conditionally
applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 389


Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Article 6.10.8 dealing with the design of composite sections in negative flexure at the strength
limit state and noncomposite sections subject to positive or negative flexure at the strength limit
state and during construction (see the Discussion of Articles 6.10.8 and 6.10.3 in this Guide), and
is strongly recommended for use in conjunction with this Guide. Article 6.10.8 assumes the section
under consideration is a slender web section or is conservatively treated as a slender web section
(see the Commentary for Article 6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in this Guide
for further discussion on the definition and categorization of compact web, noncompact web, and
slender web sections). The applicability of this flowchart is discussed further below.
Simple Span Bridges:
The portions of the flowchart dealing with the determination of Fnc for the discretely braced top
(compression) flange are applicable for the routine simple span I-girder bridges covered by this
Guide when checking the noncomposite section during construction using Eq. 6.10.3.2.1-2 (see
the Discussion of Article 6.10.3.2.1 in this Guide) if the section is a slender web section or is
conservatively treated as a slender web section (which is not recommended for rolled-beam
sections in particular). This flowchart is not applicable to these bridges at the strength limit state
as simple spans are subject to positive flexure only.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This flowchart is applicable for the routine multi-span continuous rolled-beam and plate girder
bridges covered by this Guide at the strength limit state if the sections in negative flexure are
slender web sections or are conservatively treated as slender web sections (which is not
recommended for rolled-beam sections, but may be reasonable for plate girder sections). The
portions of the flowchart dealing with the determination of Fnc of discretely braced compression
flanges are also applicable for these bridges when checking the noncomposite section during
construction using Eq. 6.10.3.2.1-2 (see the Discussion of Article 6.10.3.2.1 in this Guide).

C6.4.7 Flowchart for Appendix A6


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
the optional Appendix A6 dealing with the design of composite sections in negative flexure at the
strength limit state and noncomposite sections subject to positive or negative flexure at the strength
limit state, and in some cases, during construction (see the Discussion of Article Appendix A6 in
this Guide), and is strongly recommended for use in conjunction with this Guide. Appendix A6
may only be used for sections that satisfy the restrictions specified in Article 6.10.6.2.3 (see the
Discussion of Article 6.10.6.2.3 in this Guide). The applicability of this flowchart is discussed
further below.

NSBA Guide to Navigating Routine Steel Bridge Design / 390


Simple Span Bridges:
The portions of the flowchart dealing with the determination of Mnc for lateral-torsional buckling
of the discretely braced top (compression) flange are applicable for the routine simple span I-girder
bridges covered by this Guide with nonslender webs when the provisions of Article A6.3.3 are
used to compute Mnc to account for the beneficial effect of the St. Venant torsional constant, J,
when checking the noncomposite section during construction using Eq. 6.10.3.2.1-2 (see the
Discussion of Article 6.10.3.2.1 in this Guide), which is recommended for rolled-beam bridges in
particular. This Article is not applicable to these bridges at the strength limit state as simple spans
are subject to positive flexure only.
Multi-span Continuous Rolled Beam Bridges:
This flowchart is applicable for the routine multi-span continuous rolled beam bridges covered by
this Guide at the strength limit state if the sections in negative flexure satisfy the restrictions
specified in Article 6.10.6.2.3 for the use of Appendix A6, which is typically the case for rolled-
beam sections. The portions of the flowchart dealing with the determination of Mnc for lateral-
torsional buckling of discretely braced compression flanges are applicable for bridges with
nonslender webs when the provisions of Article A6.3.3 are used to compute Mnc to account for the
beneficial effect of the St. Venant torsional constant, J, when checking the noncomposite section
during construction using Eq. 6.10.3.2.1-2 (see the Discussion of Article 6.10.3.2.1 in this Guide),
which is recommended for rolled-beam bridges.
Multi-span Continuous Plate Girder Bridges:
This flowchart is applicable for the routine multi-span continuous plate girder bridges covered by
this Guide at the strength limit state if the sections in negative flexure satisfy the restrictions
specified in Article 6.10.6.2.3 for the use of Appendix A6. The portions of the flowchart dealing
with the determination of Mnc for lateral-torsional buckling of discretely braced compression
flanges are applicable for bridges with nonslender webs when the provisions of Article A6.3.3 are
used to compute Mnc to account for the beneficial effect of the St. Venant torsional constant, J,
when checking the noncomposite section during construction using Eq. 6.10.3.2.1-2 (see the
Discussion of Article 6.10.3.2.1 in this Guide).

C6.4.8 Flowchart for Article D6.4.1


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Article D6.4.1, which should always be invoked when calculating the nominal lateral-torsional
buckling resistance for uniform bending according to the provisions of Article 6.10.8.2.3, and
when a moment-gradient modifier, Cb, is calculated for the unbraced length under consideration
and is greater than 1.0 (see the Discussions of Articles 6.10.8.2.3 and D6.4.1 in this Guide). The
provisions of this Article can result in the nominal flexural resistance of the compression flange

NSBA Guide to Navigating Routine Steel Bridge Design / 391


reaching the plateau strength (Fmax) for lateral-torsional buckling at significantly larger unbraced
lengths under moment-gradient conditions when the effects of the moment-gradient are included
in determining the limits on the unbraced length, Lb (see the dashed curve in Figure C6.10.8.2.1-
1). The applicability of this flowchart is discussed further below.
Simple Span Bridges:
This flowchart is not applicable for the routine simple span bridges covered by this Guide at the
strength limit state because the top flanges are continuously braced. This flowchart may be
applicable for these bridges when computing the lateral-torsional buckling resistance of the
discretely braced top (compression) flange of the noncomposite section during construction using
the provisions of Article 6.10.8.2.3 for use in Eq. 6.10.3.2.1-2 (see the Discussion of Article
6.10.3.2.1 in this Guide), which is not recommended for rolled-beam bridges in particular, and
when Cb is calculated for the unbraced length under consideration and is greater than 1.0. Note
however that Cb is typically taken equal to 1.0 when checking lateral-torsional buckling of the
critical noncomposite section in regions of positive flexure during construction; otherwise, consult
Figure C6.4.10 for the appropriate calculation of Cb in these regions (see the Discussion of Article
C6.4.10 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This flowchart is applicable for the routine multi-span continuous rolled-beam and plate girder
bridges covered by this Guide at the strength limit state when computing the lateral-torsional
buckling resistance of the discretely braced bottom (compression) flange in regions of negative
flexure using the provisions of Article 6.10.8.2.3 (which is not recommended for rolled-beam
bridges), and when Cb is calculated for the unbraced length under consideration and is greater than
1.0. This flowchart is also applicable for these bridges when computing the lateral-torsional
buckling resistance of the discretely braced compression flanges of the noncomposite section
during construction using the provisions of Article 6.10.8.2.3 for use in Eq. 6.10.3.2.1-2 (which is
not recommended for rolled-beam bridges, see the Discussion of Article 6.10.3.2.1 in this Guide),
and when Cb is calculated for the unbraced length under consideration and is greater than 1.0. Note
however that Cb is typically taken equal to 1.0 when checking lateral-torsional buckling of the
critical noncomposite section in regions of positive flexure during construction; otherwise, consult
Figure C6.4.10 for the appropriate calculation of Cb in these regions (see the Discussion of Article
C6.4.10 in this Guide).

C6.4.9 Flowchart for Article D6.4.2


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Article D6.4.2, which should always be invoked when calculating the nominal lateral-torsional
buckling resistance for uniform bending according to the provisions of Article A6.3.3 in the

NSBA Guide to Navigating Routine Steel Bridge Design / 392


optional Appendix A6, and when a moment-gradient modifier, Cb, is calculated for the unbraced
length under consideration and is greater than 1.0 (see the Discussions of Articles A6.3.3 and
D6.4.2 in this Guide). The provisions of this Article can result in the nominal flexural resistance
of the compression flange reaching the plateau strength (Mmax) for lateral-torsional buckling at
significantly larger unbraced lengths under moment-gradient conditions when the effects of the
moment-gradient are included in determining the limits on the unbraced length, Lb (see the dashed
curve in Figure C6.10.8.2.1-1). The applicability of this flowchart is discussed further below.
Simple Span Bridges:
This flowchart is not applicable for the routine simple span bridges covered by this Guide at the
strength limit state because the top flanges are continuously braced. This flowchart may be
applicable for these bridges when computing the lateral-torsional buckling resistance of
noncomposite sections with nonslender webs during construction for use in Eq. 6.10.3.2.1-2 (see
the Discussion of Article 6.10.3.2.1 in this Guide) using the provisions of Article A6.3.3 to account
for the beneficial effect of the St. Venant torsional constant, J, which is recommended for rolled-
beam bridges in particular, and when Cb is calculated for the unbraced length under consideration
and is greater than 1.0. Note however that Cb is typically taken equal to 1.0 when checking lateral-
torsional buckling of the critical noncomposite section in regions of positive flexure during
construction; otherwise, consult Figure C6.4.10 for the appropriate calculation of Cb in these
regions (see the Discussion of Article C6.4.10 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
This flowchart is applicable for the routine multi-span continuous rolled-beam and plate girder
bridges covered by this Guide at the strength limit state when computing the lateral-torsional
buckling resistance of sections in regions of negative flexure with nonslender webs using the
provisions of Article A6.3.3 (which is recommended for rolled-beam bridges), and when Cb is
calculated for the unbraced length under consideration and is greater than 1.0. This flowchart is
also applicable for these bridges when computing the lateral-torsional buckling resistance of
noncomposite sections with nonslender webs during construction for use in Eq. 6.10.3.2.1-2 (see
the Discussion of Article 6.10.3.2.1 in this Guide) using the provisions of Article A6.3.3 to account
for the beneficial effect of the St. Venant torsional constant, J (which is recommended for rolled-
beam bridges), and when Cb is calculated for the unbraced length under consideration and is greater
than 1.0. Note however that Cb is typically taken equal to 1.0 when checking lateral-torsional
buckling of the critical noncomposite section in regions of positive flexure during construction;
otherwise, consult Figure C6.4.10 for the appropriate calculation of Cb in these regions (see the
Discussion of Article C6.4.10 in this Guide).

C6.4.10 Moment Gradient Modifier, Cb (Sample Cases)


Determination of applicability, All Routine Steel I-girder Bridges: Partially applicable.
Discussion:
This article provides helpful example illustrations of the calculation of the moment-gradient
modifier, Cb, for seven different unbraced lengths under varying moment-gradient conditions using
the equations provided for the calculation of Cb in Article 6.10.8.2.3 (see the Discussion of Article

NSBA Guide to Navigating Routine Steel Bridge Design / 393


6.10.8.2.3 in this Guide). All cases assume the member is prismatic within the unbraced length, or
the transition to a smaller section within the unbraced length is within 0.2Lb of the brace point with
the lower moment; otherwise, Cb must be taken equal to 1.0. Cb can be used to scale the lateral-
torsional buckling resistance such that the plateau strength (Fmax or Mmax, as applicable) for lateral-
torsional buckling is reached at significantly larger unbraced lengths under moment-gradient
conditions (see the dashed curve in Figure C6.10.8.2.1-1).
The first, third, and fifth sample cases are generally representative of unbraced lengths in simple
span bridges and in regions of positive flexure in multi-span continuous bridges, which represent
unbraced lengths that would need to be investigated for the noncomposite girder during
construction. Specifically, the third sample case represents a simple span braced only at its ends
and at midspan (only one-half of the member is shown). The second sample case is generally
representative of unbraced lengths in regions of negative flexure in multi-span continuous bridges.
The fourth sample case represents the rare case of a continuous span with no cross-bracing within
the span. A case (not shown) where the stress would be zero at both ends of the unbraced length
would represent a simple span with no cross-bracing within the span for which Cb would also be
taken equal to 1.0. The sixth and seventh sample cases are generally representative of unbraced
lengths subject to reverse curvature bending in multi-span continuous bridges.

C6.5 FLOWCHARTS FOR LRFD ARTICLES 6.9.4 AND 6.12.2.2.2

C6.5.1 Flowchart for LRFD Article 6.9.4


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Article 6.9.4 dealing with the determination of the nominal compressive resistance of
noncomposite I- and box-section members (see the Discussion of Article 6.9.4 in this Guide).
This flowchart is not applicable to the routine steel I-girder bridges covered by this Guide as the
members in these bridges are not fully noncomposite, are not box-section members, and are subject
to flexure only.

C6.5.2 Flowchart for LRFD Article 6.12.2.2.2


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The flowchart provided in this Article is helpful to guide the Engineer through the provisions of
Article 6.12.2.2.2 dealing with the determination of the nominal flexural resistance of rectangular
noncomposite box-section members (see the Discussion of Article 6.12.2.2.2 in this Guide). This
flowchart is not applicable to the routine steel I-girder bridges covered by this Guide as
noncomposite box-section members are not used in these bridges.

NSBA Guide to Navigating Routine Steel Bridge Design / 394


APPENDIX D6 FUNDAMENTAL CALCULATIONS FOR FLEXURAL MEMBERS

D6.1 PLASTIC MOMENT

Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.


Discussion:
The plastic moment, Mp, is defined in the AASHTO LRFD BDS as the resisting moment about the
major axis of a fully yielded cross-section. Mp is used as a theoretical measure of the maximum
potential flexural resistance at the strength limit state of a noncomposite or composite section
satisfying specific steel grade, flange and web slenderness, compression-flange bracing and
ductility requirements, as applicable. For sections that can achieve the full plastic-moment
resistance, it is assumed that the section is completely elastic up to Mp and then rotates inelastically
at Mp with no increase in the moment resistance. The effects of strain hardening are conservatively
ignored. This idealized moment-rotation behavior is termed elastic-perfectly plastic behavior. In
the AASHTO LRFD BDS, composite sections in straight bridges in regions of positive flexure
that can achieve flexural resistances at or near Mp are termed compact sections (see the Discussion
of Article 6.10.6.2.2 in this Guide). Composite sections in regions of negative flexure and
noncomposite sections subject to positive or negative flexure in straight bridges that can achieve
flexural resistances of Mp are termed compact web sections and are less commonly used (see the
Commentary for Article 6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in this Guide for
further discussion on the definition and categorization of compact web, noncompact web, and
slender web sections).
Mp is calculated as the moment of the plastic forces acting on the cross-section about the plastic
neutral axis (PNA). For sections subject to flexure only, Mp may be calculated as the moment of
the plastic forces about any axis parallel to the PNA. Plastic forces in steel portions of the cross-
section are calculated using the yield strengths of the flanges, web, and longitudinal reinforcing
steel, as appropriate. Plastic forces in concrete portions of the cross-section (in compression only)
are based on a rectangular stress block, with the magnitude of the compressive stress taken equal
to 0.85fc. Concrete in tension is neglected. Equations to calculate these plastic forces are given in
this Article. The position of the PNA is calculated based on the equilibrium condition that there is
no net axial force acting on the cross-section.
For composite sections, the stress distribution in the cross-section at Mp is assumed independent
of the manner in which the stresses are induced into the beam. Also, creep and shrinkage are
assumed to have no effect on the internal stress distribution at Mp. Thus, when checking the flexural
resistance of a composite section against Mp, the moments acting on the non-composite, long-term
composite and short-term composite sections may be directly summed for comparison to Mp. The
effect of the sequence of application of the different types of loads on the stress states and partial
yielding within the cross-section on the resistance is not considered. For composite sections in
positive flexure, the attainment of Mp is possible only if the steel girder is provided with an
adequate number of shear connectors so that the horizontal shear force from the concrete deck is
effectively transmitted to the steel girder (see the Discussion of Article 6.10.10.4 in this Guide).
The natural bond between the steel and concrete is not sufficient by itself. Mp for a composite
section in positive flexure can be determined as follows:

NSBA Guide to Navigating Routine Steel Bridge Design / 395


• Calculate the plastic forces of each individual component in the cross-section and use
them to determine whether the PNA is in the web, top flange or concrete deck;
• Calculate the location of the PNA within the element determined in Step 1; and
• Calculate Mp. Table D6.1-1 provides equations for the location of the PNA and for
calculating Mp for seven possible conditions depending on the location of the PNA.

In Table D6.1-1, d is the distance from the element plastic force to the PNA. The element forces
are assumed to act at the mid-thickness of the flanges and concrete deck, at the mid-depth of the
web and at the center of the longitudinal reinforcement. All element forces, dimensions, and
distances are to be taken as positive. The conditions should be checked in the order listed in the
table. The forces in the longitudinal reinforcement may be conservatively neglected by setting the
terms, Prb and Prt, equal to zero in the equations given in the table.
For composite sections in negative flexure, a similar procedure can be used. In this case, however,
the tensile strength of the concrete is ignored, and the contribution of the longitudinal
reinforcement should be included. Table D6.1-2 contains the equations for the two cases most
likely to occur in practice. Again, the conditions should be checked in the order listed in the table.
For homogenous doubly symmetric noncomposite sections, Mp may simply be calculated as FyZ,
where Z is the plastic section modulus calculated as the sum of the first moments of the flange and
web areas about the PNA. For rolled wide-flange sections, values of Z are tabulated in the AISC
Manual of Steel Construction. The plastic moment of a noncomposite section may also be
calculated by simply eliminating the terms pertaining to the concrete deck and longitudinal
reinforcement from the equations in Tables D6.1-1 and D6.1-2, as applicable.
For further information on the plastic moment and example calculations of the plastic moment,
consult Section 6.4.5.2 of the Reference Manual for NHI Course 130081, Load and Resistance
Factor Design (LRFD) for Highway Bridge Superstructures.
Simple Span Bridges:
These provisions are used for the routine simple span bridges covered by this Guide to compute
the plastic moment for the composite section, which is necessary to determine the nominal flexural
resistance at the strength limit state if the section qualifies and is treated as a compact section,
which is typically the case. These provisions may also be used for these bridges to compute the
plastic moment for the noncomposite section if the section has a compact or noncompact web and
the provisions of Article A6.3.3 are used to optionally determine the nominal lateral-torsional
buckling resistance for use in checking Eq. 6.10.3.2.1-2 during construction in order to include the
beneficial effect of the St. Venant torsional constant, J (see the Discussion of Article C6.10.3.2.1
in this Guide).
Multi-span Continuous Rolled Beam Bridges:
These provisions are used for the routine multi-span continuous rolled beam bridges covered by
this Guide to compute the plastic moment for the composite section in regions of positive flexure,
which is necessary to determine the nominal flexural resistance at the strength limit state if the
section qualifies and is treated as a compact section, which will typically be the case. These

NSBA Guide to Navigating Routine Steel Bridge Design / 396


provisions may also be used for these bridges to compute the plastic moment for the composite
section or for the noncomposite section (if no shear studs are used) in regions of negative flexure
if the section satisfies the restrictions specified in Article 6.10.6.2.3 (see the Discussion of Article
6.10.6.2.3 in this Guide) and the optional provisions of Appendix A6 are used to determine the
nominal flexural resistance, which will typically be the case and is strongly encouraged for rolled-
beam sections. These provisions may also be used for these bridges to compute the plastic moment
for the noncomposite section if the section has a compact or noncompact web and the provisions
of Article A6.3.3 are optionally used to determine the nominal lateral-torsional buckling resistance
for use in checking Eq. 6.10.3.2.1-2 during construction in order to include the beneficial effect of
the St. Venant torsional constant, J (see the Discussion of Article C6.10.3.2.1 in this Guide)..
Multi-span Continuous Plate Girder Bridges:
These provisions are used for the routine multi-span continuous plate girder bridges covered by
this Guide to compute the plastic moment for the composite section in regions of positive flexure,
which is necessary to determine the nominal flexural resistance at the strength limit state if the
section qualifies and is treated as a compact section, which will typically be the case. These
provisions may also be used for these bridges to compute the plastic moment for the composite
section in regions of negative flexure or for the noncomposite section (if no shear studs are used)
if the section satisfies the restrictions specified in Article 6.10.6.2.3 (see the Discussion of Article
6.10.6.2.3 in this Guide) and the optional provisions of Appendix A6 are used to determine the
nominal flexural resistance at the strength limit state. These provisions may also be used for these
bridges to compute the plastic moment for the noncomposite section if the section has a compact
or noncompact web and the provisions of Article A6.3.3 are optionally used to determine the
nominal lateral-torsional buckling resistance for use in checking Eq. 6.10.3.2.1-2 during
construction in order to include the beneficial effect of the St. Venant torsional constant, J (see the
Discussion of Article C6.10.3.2.1 in this Guide).

D6.2 YIELD MOMENT

D6.2.1 Noncomposite Sections


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The yield moment, My, is defined in the AASHTO LRFD BDS as the moment at which an outer
fiber, in a member subjected to flexure about the major-axis, attains the nominal yield stress
neglecting the effect of any residual stresses.
For a noncomposite section, this Article states that My is to be taken as the smaller of the moment
required to cause nominal first yielding in the compression flange, Myc, or the moment required to
cause nominal first yielding in the tension flange, Myt at the strength limit state.
The ratio of Mp/My is a property of the cross-sectional shape known as the shape factor. For doubly
symmetric noncomposite I-shapes bent about their major axis, the shape factor is approximately
1.12.

NSBA Guide to Navigating Routine Steel Bridge Design / 397


The computation of My for a noncomposite section may be required for the routine steel I-girder
bridges covered by this Guide if the section has a compact or noncompact web and the provisions
of Article A6.3.3 are optionally used to determine the nominal lateral-torsional buckling resistance
for use in checking Eq. 6.10.3.2.1-2 during construction in order to include the beneficial effect of
the St. Venant torsional constant, J (see the Discussion of Article C6.10.3.2.1 in this Guide). The
computation of My for a noncomposite section in regions of negative flexure (i.e., Myc and Myt) if
no shear studs are used may also be required at the strength limit state if the noncomposite section
satisfies the restrictions specified in Article 6.10.6.2.3 (see the Discussion of Article 6.10.6.2.3 in
this Guide) and the optional provisions of Appendix A6 are used to compute the nominal flexural
resistance. See the Commentary for Article 6.10.6.2.3 and the Discussion of Article 6.10.6.2.3 in
this Guide for further discussion on the definition and categorization of compact web, noncompact
web, and slender web sections.
For further information on the yield moment and example calculations of the yield moment,
consult Section 6.4.5.3 of the Reference Manual for NHI Course 130081, Load and Resistance
Factor Design (LRFD) for Highway Bridge Superstructures.

D6.2.2 Composite Sections in Positive Flexure


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Applicable.
Discussion:
The yield moment, My, is defined in the AASHTO LRFD BDS as the moment at which an outer
fiber, in a member subjected to flexure about the major-axis, attains the nominal yield stress
neglecting the effect of any residual stresses.
For composite sections in positive flexure, My is to be taken as the sum of the moments applied
separately to the steel, short-term and long-term composite sections to cause nominal first yielding
in either flange at the strength limit state. My is taken as the smaller of the moment required to
cause nominal first yielding in the compression flange, Myc, or the moment required to cause
nominal first yielding in the tension flange, Myt.
In a composite girder, moments are applied to different sections and this fact must be appropriately
accounted for in the computation of My. My for a composite section in positive flexure can therefore
be determined as follows: 1) calculate the moment, MD1, caused by the factored permanent load
applied before the concrete deck has hardened or is made composite and apply this moment to the
steel section; 2) calculate the moment, MD2, caused by the remainder of the factored permanent
load and apply this moment to the long-term composite section; 3) calculate the additional
moment, MAD, that must be applied to the short-term composite section to cause nominal yielding
in either steel flange; and 4) calculate My as the sum of the total permanent load moment and MAD
(see Eqs. D6.2.2-1 and D6.2.2-2 and the Discussion of Article 6.10.1.1.1b in this Guide). In regions
of positive flexure, the longitudinal reinforcement may be neglected in the calculation of SST and
SLT in Eq. D6.2.2-1.

NSBA Guide to Navigating Routine Steel Bridge Design / 398


The ratio of Mp/My is a property of the cross-sectional shape known as the shape factor. For singly
symmetric composite girders in regions of positive flexure, values of the shape factor on the order
of 1.4 to 1.6 are quite common.
For further information on the yield moment and example calculations of the yield moment,
consult Section 6.4.5.3 of the Reference Manual for NHI Course 130081, Load and Resistance
Factor Design (LRFD) for Highway Bridge Superstructures.
Simple Span Bridges:
The computation of My for a composite section in positive flexure is not applicable to the routine
steel simple span bridges covered by this Guide, as My is not required for the determination of the
nominal flexural resistance of a simple-span bridge at the strength limit state (see the Discussion
of Article 6.10.7.1.2 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
The computation of My for a composite section in positive flexure is applicable to the routine steel
multi-span continuous rolled beam and plate girder bridges covered by this Guide for the
determination of the nominal flexural resistance at the strength limit state if the section qualifies
and is treated as a compact section, which will typically be the case (see the Discussion of Articles
6.10.6.2.2 and 6.10.7.1.2 in this Guide).

D6.2.3 Composite Sections in Negative Flexure


Determination of applicability, Simple Span Bridges: Not applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges and Multi-span
Continuous Plate Girder Bridges: Conditionally applicable.
Discussion:
The yield moment, My, is defined in the AASHTO LRFD BDS as the moment at which an outer
fiber, in a member subjected to flexure about the major-axis, attains the nominal yield stress
neglecting the effect of any residual stresses.
For composite sections in negative flexure, the procedure specified for the calculation of My
(including Myc and Myt) for composite sections in positive flexure in Article D6.2.2 is followed,
except the composite section for both the short-term and long-term moments applied to the
composite section is to consist of the steel section and the longitudinal reinforcement within the
effective width of the concrete deck. Thus, SST and SLT in Eq. D6.2.2-1 are taken as the same value
(see Eqs. D6.2.2-1 and D6.2.2-2 and the Discussion of Articles D6.2.2 and 6.10.1.1.1c in this
Guide). Also, Myt is to be taken with respect to either the tension flange or the longitudinal
reinforcement, whichever yields first. For the calculation of Myt with respect to the longitudinal
reinforcement, MD1 is to be taken equal to zero in Eqs. D6.2.2-1 and D6.2.2-2, and Fyf in
Eq. D6.2.2-1 is to be taken equal to the specified minimum yield strength of the longitudinal
reinforcement.

NSBA Guide to Navigating Routine Steel Bridge Design / 399


For further information on the yield moment and example calculations of the yield moment,
consult Section 6.4.5.3 of the Reference Manual for NHI Course 130081, Load and Resistance
Factor Design (LRFD) for Highway Bridge Superstructures.
Simple Span Bridges:
The computation of My for a composite section in negative flexure is not applicable to the routine
steel simple span bridges covered by this Guide, as simple-span bridges are not subject to negative
flexure.
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
The computation of My for a composite section in negative flexure (i.e., Myc and Myt) is applicable
to the routine steel multi-span continuous rolled beam and plate girder bridges covered by this
Guide if the section satisfies the restrictions specified in Article 6.10.6.2.3 (see the Discussion of
Article 6.10.6.2.3 in this Guide) and the optional provisions of Appendix A6 are used to compute
the nominal flexural resistance at the strength limit state, which will typically be the case and is
strongly encouraged for rolled-beam sections, and which may be possible for plate girder sections.

D6.2.4 Sections with Cover Plates


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The yield moment, My, is defined in the AASHTO LRFD BDS as the moment at which an outer
fiber, in a member subjected to flexure about the major-axis, attains the nominal yield stress
neglecting the effect of any residual stresses.
For composite sections in positive or negative flexure, My is to be taken as the sum of the moments
applied separately to the steel, short-term and long-term composite sections to cause nominal first
yielding in either flange at the strength limit state. For both noncomposite and composite sections,
My is taken as the smaller of the moment required to cause nominal first yielding in the compression
flange, Myc, or the moment required to cause nominal first yielding in the tension flange, Myt. See
the Discussion of Articles D6.2.1 through D6.2.3 in this Guide for further information on the
computation of My for noncomposite and composite sections.
For sections containing flange cover plates, Myc or Myt is to be taken as the smallest value of
moment associated with nominal first yielding based on the stress in either the flange under
consideration or in any of the cover plates attached to that flange, whichever yields first.
The provisions of this Article are not applicable to the routine steel I-girder bridges covered by
this Guide as cover plates are not used, nor are they recommended for use, on these bridges.
For further information on the yield moment and example calculations of the yield moment,
consult Section 6.4.5.3 of the Reference Manual for NHI Course 130081, Load and Resistance
Factor Design (LRFD) for Highway Bridge Superstructures.

NSBA Guide to Navigating Routine Steel Bridge Design / 400


D6.3 DEPTH OF THE WEB IN COMPRESSION
D6.3.1 In the Elastic Range (Dc)
Determination of applicability, Simple Span Bridges: Partially applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Applicable.
Discussion:
The depth of the web in compression in the elastic range, Dc, is used primarily in computing the
web bend-buckling resistance, Fcrw, and the web load-shedding factor, Rb (see the Discussion of
Articles 6.10.1.9.1 and 6.10.1.10.2 in this Guide). For composite sections in negative flexure and
noncomposite sections subject to positive or negative flexure, Dc is also used to determine whether
the section qualifies as a slender or a nonslender web section for determining the nominal flexural
resistance (see the Discussion of Article 6.10.6.2.3 in this Guide).
Dc for composite sections is a function of the dead-to-live load stress ratio in the elastic range of
stress at the service, fatigue, and strength limit states. This is because in a composite girder, the
dead and live loads are applied to different sections. This is an especially important consideration
for composite sections since the dead-load stress has a significant effect on the location of the
elastic neutral axis. Note that when checking the section for web bend-buckling during
construction, however, while the girder is still in the noncomposite condition before the concrete
deck hardens, Dc of the steel section alone, which is a section property independent of the stress,
is always to be used in the calculations.
Dc of the composite section at sections in positive flexure increases with increasing span length
because of the increasing dead-to-live load ratio. With increasing spans, the larger noncomposite
dead load stresses acting on the steel section alone effectively cause the neutral axis to be much
lower than it would be if all loads were applied to the composite section, which obviously increases
the depth of the web in compression. Therefore, where applicable, it is important to recognize the
effect of the dead load stress on the location of the neutral axis at these sections. This Article states
that Dc for composite sections in positive flexure is to be taken as the depth over which the
algebraic sum of the stresses acting on the steel, long-term composite and short-term composite
sections due to the dead and live loads, plus impact, is compressive; Eq. D6.3.1-1 may be used to
calculate Dc at such sections. However, according to the AASHTO LRFD BDS, for composite
sections in positive flexure at the service and strength limit states, Dc only needs to be employed
in the computation of the nominal flexural resistance for sections in which longitudinal web
stiffeners are required (reasons for this are discussed further in the Commentary for Article
6.10.1.9.1). Therefore, the computation of Dc is not applicable (or necessary) at the service and
strength limit states for composite sections in positive flexure in the routine steel I-girder bridges
covered by this Guide, which do not have longitudinal web stiffeners.
The concrete deck is typically not considered to be effective in tension for composite sections in
negative flexure, except perhaps at the fatigue and service limit states as permitted when certain
conditions are satisfied (see the Discussion of Articles 6.6.1.2.1 and 6.10.4.2.1 in this Guide). The
distance between the neutral-axis locations for the steel and composite sections is small when the

NSBA Guide to Navigating Routine Steel Bridge Design / 401


concrete deck is not considered effective, as the composite section only consists of the steel section
plus the longitudinal reinforcement. As a result, the location of the neutral axis for the composite
section is essentially unaffected by the dead load stress. In fact, accounting for the effect of the
dead load stress results in a smaller value of Dc in regions of negative flexure.
Therefore, for the majority of situations involving composite sections in negative flexure, this
Article conservatively specifies the use of Dc computed for the section consisting of the steel girder
plus the longitudinal reinforcement only, without considering the algebraic sum of the stresses
acting on the noncomposite and composite sections.
A single exception to the preceding requirement occurs if the concrete deck is assumed effective
in tension in regions of negative flexure at the service limit state, as permitted for composite girders
satisfying the conditions specified in Article 6.10.4.2.1 (see the Discussion of Article 6.10.4.2.1 in
this Guide); in such cases, Eq. D6.3.1-1 must be used to compute Dc. For this case, in
Figure D6.3.1-1, the stresses fc and ft should be switched, the signs shown in the stress diagram
should be reversed, tfc should be the thickness of the bottom flange, and Dc should instead extend
from the neutral axis down to the top of the bottom flange. When calculating the web bend-
buckling resistance, Fcrw, at the service limit state (see the Discussion of Article 6.10.4.2.2 in this
Guide), a more precise calculation of Dc, accounting for the beneficial effect of the dead load stress
in this case, is required for the composite section whenever the concrete deck is permitted to be
considered effective in tension. Otherwise, the reduction in Fcrw will be too large and not reflective
of the actual potential web bend-buckling resistance at this limit state.
For further information on the elastic depth of the web in compression, Dc, and example
calculations of Dc, consult Section 6.4.5.4.1 of the Reference Manual for NHI Course 130081,
Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.
Simple Span Bridges:
For the routine simple span bridges covered by this Guide, only Dc for the noncomposite steel
section must be calculated in order to check the web bend-buckling resistance of the section during
construction (see the Discussion of Article 6.10.3.2.1 in this Guide) and to determine whether the
web of the noncomposite steel section is slender or nonslender (see the Discussion of Article
6.10.6.2.3 in this Guide). Dc need not be calculated at the service or strength limit states.
Multi-span Continuous Rolled Beam Bridges:
For the routine multi-span continuous rolled beam bridges covered by this Guide, Dc for the
noncomposite steel section must be calculated in order to check the web bend-buckling resistance
of the section during construction (see the Discussion of Article 6.10.3.2.1 in this Guide) and to
determine whether the web of the noncomposite steel section is slender or nonslender (see the
Discussion of Article 6.10.6.2.3 in this Guide). Dc need not be calculated at the service or strength
limit states for composite sections in regions of positive flexure.
For composite sections and noncomposite sections (if shear studs are not provided) in regions of
negative flexure at the strength limit state, Dc must be calculated to determine whether the section
is a slender or nonslender web section (rolled beams are typically nonslender web sections). In
this case, for composite sections, Dc is to be computed for the section consisting of the steel girder

NSBA Guide to Navigating Routine Steel Bridge Design / 402


plus the longitudinal reinforcement only, and for noncomposite sections, Dc is to be computed for
the steel section only. For composite sections and noncomposite sections in these regions at the
service limit state, Dc must be calculated in order to check the web bend-buckling resistance of the
section (see the Discussion of Article 6.10.4.2.2 in this Guide). In this case, for composite sections,
if the requirements of Article 6.10.4.2.1 are satisfied and the concrete deck is permitted to be
assumed effective in tension (see the Discussion of Article 6.10.4.2.1 in this Guide), Eq. D6.3.1-1
must be used to compute Dc. Otherwise, Dc is to be computed for the section consisting of the steel
girder plus the longitudinal reinforcement only. For noncomposite sections, Dc is always to be
computed for the steel section only.
Multi-span Continuous Plate Girder Bridges:
See the preceding Discussion for Multi-span Continuous Rolled Beam Bridges. In addition, for
slender web composite sections and noncomposite sections (if shear studs are not provided) in
regions of negative flexure at the strength limit state, Dc must be calculated in order to determine
the web load-shedding factor, Rb (see the Discussion of Article 6.10.1.10.2 in this Guide).

D6.3.2 At Plastic Moment (Dcp)


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The depth of the web in compression at the plastic moment, Dcp, is used primarily to determine if
a composite section in regions of positive flexure qualifies as a compact section at the strength
limit state (see the Discussion of Article 6.10.6.2.2 in this Guide), and to determine if a nonslender,
composite section or a nonslender, noncomposite section (if shear studs are not provided) in
regions of negative flexure qualifies as either a compact web or a noncompact web section at the
strength limit state (see the Discussion of Article A6.2.1 in this Guide).
At sections in positive flexure, Dcp is typically smaller than the elastic depth of the web in
compression, Dc, as plastic strains associated with moments larger than My are incurred. In fact,
for composite sections in positive flexure, the neutral axis at the plastic moment, Mp, will most
often be located either in the concrete deck or in the top flange of the steel girder. In such cases,
the entire web of the girder is in tension and Dcp is to be taken as zero according to Article D6.3.2.
When Dcp is equal to zero, the web-slenderness requirements in the AASHTO LRFD BDS based
on Dcp are assumed automatically satisfied. The location of the plastic neutral axis (PNA) for
composite sections in positive flexure can be determined from the conditions listed in Table D6.1-
1 (see the Discussion of Article D6.1 in this Guide). The position of the PNA is calculated based
on the equilibrium condition that there be no net axial force acting on the assumed fully yielded
cross-section.
At sections in negative flexure, Dcp is typically larger than the elastic depth of the web in
compression, Dc, as plastic strains associated with moments larger than My are incurred. The
location of the PNA for composite sections in negative flexure and for noncomposite sections can
be determined from the conditions listed in Table D6.1-2. For noncomposite sections, the terms
related to the longitudinal reinforcement in Table D6.1-2 should be set equal to zero. In calculating
Dcp in regions of negative flexure, the concrete deck is assumed not to be effective in tension.

NSBA Guide to Navigating Routine Steel Bridge Design / 403


Therefore, in most cases, the plastic neutral axis will be in the web. For rare cases in which the
plastic neutral axis is in the top flange and the entire web is in compression, Dcp is to be taken
equal to the web depth D. For composite sections in negative flexure where the plastic neutral axis
is in the web, Dcp may simply be computed using Eq. D6.3.2-2.
For further information on the depth of the web in compression at the plastic moment, Dcp, and
example calculations of Dcp, consult Section 6.4.5.4.2 of the Reference Manual for NHI Course
130081, Load and Resistance Factor Design (LRFD) for Highway Bridge Superstructures.
Simple Span Bridges:
For the routine simple span bridges covered by this Guide, Dcp for the composite section at the
strength limit state will most always be in the top flange or concrete deck and thus may be taken
equal to zero if that is the case. Dcp for the noncomposite section may be needed if the section has
a nonslender web and the provisions of Article A6.3.3 are optionally used to determine the nominal
lateral-torsional buckling resistance for use in checking Eq. 6.10.3.2.1-2 during construction in
order to include the beneficial effect of the St. Venant torsional constant, J (see the Discussion of
Article C6.10.3.2.1 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
For the routine multi-span continuous rolled beam bridges covered by this Guide, Dcp for
composite sections in regions of positive flexure at the strength limit state will most always be in
the top flange or concrete deck and thus may be taken equal to zero if that is the case. Dcp must be
calculated for these bridges if composite sections or noncomposite sections (if studs are not
provided) in regions of negative flexure satisfy the restrictions specified in Article 6.10.6.2.3 (see
the Discussion of Article 6.10.6.2.3 in this Guide) and the optional provisions of Appendix A6 are
used to determine the nominal flexural resistance at the strength limit state, which is typically the
case and is strongly recommended for rolled-beam bridges. In such cases, Dcp is used to determine
whether the section qualifies as either a compact web or a noncompact web section (see the
Discussion of Article A6.2.1 in this Guide). Dcp for the noncomposite section may be needed if the
section has a nonslender web and the provisions of Article A6.3.3 are optionally used to determine
the nominal lateral-torsional buckling resistance for use in checking Eq. 6.10.3.2.1-2 during
construction in order to include the beneficial effect of the St. Venant torsional constant, J (see the
Discussion of Article 6.10.3.2.1 in this Guide).
D6.4 LATERAL TORSIONAL BUCKLING EQUATIONS FOR CB > 1.0, WITH
EMPHASIS ON UNBRACED LENGTH REQUIREMENTS FOR
DEVELOPMENT OF THE MAXIMUM FLEXURAL RESISTANCE
D6.4.1 By the Provisions of Article 6.10.8.2.3
Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article should always be invoked when calculating the nominal lateral-
torsional buckling resistance for uniform bending according to the provisions of Article 6.10.8.2.3,
and when a moment-gradient modifier, Cb, is calculated for the unbraced length under

NSBA Guide to Navigating Routine Steel Bridge Design / 404


consideration and is greater than 1.0 (see the Discussion of Article 6.10.8.2.3 in this Guide). The
provisions of this Article can result in the nominal flexural resistance of the compression flange
reaching the plateau strength (Fmax) for lateral-torsional buckling at significantly larger unbraced
lengths under moment-gradient conditions when the effects of the moment-gradient are included
in determining the limits on the unbraced length, Lb (see the dashed curve in Figure C6.10.8.2.1-
1).
Simple Span Bridges:
The provisions of this Article are not applicable for the routine simple span bridges covered by
this Guide at the strength limit state because the top flanges are continuously braced. These
provisions may be applicable to these bridges when computing the lateral-torsional buckling
resistance of the discretely braced top (compression) flange of the noncomposite section during
construction using the provisions of Article 6.10.8.2.3 for use in Eq. 6.10.3.2.1-2 (see the
Discussion of Article 6.10.3.2.1 in this Guide), which is not recommended for rolled-beam bridges
in particular, and when Cb is calculated for the unbraced length under consideration and is greater
than 1.0. Note however that Cb is typically taken equal to 1.0 when checking lateral-torsional
buckling of the critical noncomposite section in regions of positive flexure during construction;
otherwise, consult Figure C6.4.10 for the appropriate calculation of Cb in these regions (see the
Discussion of Article C6.4.10 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
The provisions of this Article are applicable for the routine multi-span continuous rolled-beam and
plate girder bridges covered by this Guide at the strength limit state when computing the lateral-
torsional buckling resistance of the discretely braced bottom (compression) flange in regions of
negative flexure using the provisions of Article 6.10.8.2.3 (which is not recommended for rolled-
beam bridges), and when Cb is calculated for the unbraced length under consideration and is greater
than 1.0. These provisions are also applicable for these bridges when computing the lateral-
torsional buckling resistance of the discretely braced compression flanges of the noncomposite
section during construction using the provisions of Article 6.10.8.2.3 for use in Eq. 6.10.3.2.1-2
(which is not recommended for rolled-beam bridges, see the Discussion of Article 6.10.3.2.1 in
this Guide), and when Cb is calculated for the unbraced length under consideration and is greater
than 1.0. Note however that Cb is typically taken equal to 1.0 when checking lateral-torsional
buckling of the critical noncomposite section in regions of positive flexure during construction;
otherwise, consult Figure C6.4.10 for the appropriate calculation of Cb in these regions (see the
Discussion of Article C6.4.10 in this Guide).

D6.4.2 By the Provisions of Article A6.3.3


Determination of applicability, All Routine Steel I-girder Bridges: Conditionally applicable.
Discussion:
The provisions of this Article should always be invoked when calculating the nominal lateral-
torsional buckling resistance for uniform bending according to the provisions of Article A6.3.3 in
the optional Appendix A6, and when a moment-gradient modifier, Cb, is calculated for the
unbraced length under consideration and is greater than 1.0 (see the Discussion of Article A6.3.3

NSBA Guide to Navigating Routine Steel Bridge Design / 405


in this Guide). The provisions of this Article can result in the nominal flexural resistance of the
compression flange reaching the plateau strength (Mmax) for lateral-torsional buckling at
significantly larger unbraced lengths under moment-gradient conditions when the effects of the
moment-gradient are included in determining the limits on the unbraced length, Lb (see the dashed
curve in Figure C6.10.8.2.1-1).
Simple Span Bridges:
The provisions of this Article are not applicable for the routine simple span bridges covered by
this Guide at the strength limit state because the top flanges are continuously braced. These
provisions may be applicable to these bridges when computing the lateral-torsional buckling
resistance of noncomposite sections with nonslender webs during construction for use in Eq.
6.10.3.2.1-2 (see the Discussion of Article 6.10.3.2.1 in this Guide) using the provisions of Article
A6.3.3 to account for the beneficial effect of the St. Venant torsional constant, J, which is
recommended for rolled-beam bridges in particular, and when Cb is calculated for the unbraced
length under consideration and is greater than 1.0. Note however that Cb is typically taken equal to
1.0 when checking lateral-torsional buckling of the critical noncomposite section in regions of
positive flexure during construction; otherwise, consult Figure C6.4.10 for the appropriate
calculation of Cb in these regions (see the Discussion of Article C6.4.10 in this Guide).
Multi-span Continuous Rolled Beam Bridges and Multi-span Continuous Plate Girder Bridges:
The provisions of this Article are applicable for the routine multi-span continuous rolled-beam and
plate girder bridges covered by this Guide at the strength limit state when computing the lateral-
torsional buckling resistance of sections in regions of negative flexure with nonslender webs using
the provisions of Article A6.3.3 (which is recommended for rolled-beam bridges), and when Cb is
calculated for the unbraced length under consideration and is greater than 1.0. These provisions
are also applicable for these bridges when computing the lateral-torsional buckling resistance of
noncomposite sections with nonslender webs during construction for use in Eq. 6.10.3.2.1-2 (see
the Discussion of Article 6.10.3.2.1 in this Guide) using the provisions of Article A6.3.3 to account
for the beneficial effect of the St. Venant torsional constant, J (which is recommended for rolled-
beam bridges), and when Cb is calculated for the unbraced length under consideration and is greater
than 1.0. Note however that Cb is typically taken equal to 1.0 when checking lateral-torsional
buckling of the critical noncomposite section in regions of positive flexure during construction;
otherwise, consult Figure C6.4.10 for the appropriate calculation of Cb in these regions (see the
Discussion of Article C6.4.10 in this Guide).

D6.5 CONCENTRATED LOADS APPLIED TO WEBS WITHOUT BEARING


STIFFENERS

D6.5.1 General
Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 406


Discussion:
At bearing locations on rolled shapes, and at other locations on built-up sections or rolled shapes
subjected to concentrated loads where the loads are not transmitted through a deck or deck system
(e.g., a jacking point on a girder or on a diaphragm at an end and/or interior support to facilitate
bearing replacement or other maintenance activities), either bearing stiffeners must be provided or
else the web must be investigated for the limit states of web local yielding and web crippling (see
the Discussions of Articles D6.5.2 and D6.5.3 in this Guide).
The limit state of sidesway web buckling given in the AISC Specification for Structural Steel
Buildings is not included in the AASHTO LRFD BDS because it governs only for: 1) members
subjected to concentrated loads applied directly to the steel section; 2) members for which the
compression flange is braced at the load point; 3) members for which the tension flange is unbraced
at the load point; and 4) members for which the ratio of D/tw to Lb/bft is less than or equal to 1.7.
The preceding conditions do not commonly occur in bridge construction.
Simple Span Bridges:
The provisions of this Article are not applicable to the routine simple span plate-girder bridges
covered by this Guide at support reactions, at which bearing stiffeners are required (see the
Discussion of Article 6.10.11.2 in this Guide). These provisions are applicable to determine if
bearing stiffeners are required at support reactions in routine simple span rolled-beam bridges.
These provisions are also applicable for the routine simple-span plate-girder or rolled-beam
bridges covered by this Guide to determine if bearing stiffeners are required at jacking points on
the beam/girder or a diaphragm at the end supports (if provided).
Multi-span Continuous Rolled Beam Bridges:
The provisions of this Article are applicable to the routine multi-span continuous rolled beam
bridges covered by this Guide to determine if bearing stiffeners are required at support reactions.
These provisions are also applicable for these bridges to determine if bearing stiffeners are required
at jacking points on the beam or diaphragm at an end and/or interior support (if provided).
Multi-span Continuous Plate Girder Bridges:
The provisions of this Article are not applicable to the routine multi-span continuous plate-girder
bridges covered by this Guide at support reactions, at which bearing stiffeners are required (see
the Discussion of Article 6.10.11.2 in this Guide). These provisions are applicable for these bridges
to determine if bearing stiffeners are required at jacking points on the girder or diaphragm at an
end and/or interior support (if provided).

D6.5.2 Web Local Yielding


Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.

NSBA Guide to Navigating Routine Steel Bridge Design / 407


Discussion:
The limit state of web local yielding is intended to prevent localized yielding of the web at the
strength limit state due to a high compressive or tensile stress caused by a concentrated load or
bearing reaction. These provisions are also used to check for the need for a partial-or full-depth
transverse stiffener at the location where the bottom flange becomes horizontal in a variable web-
depth member. The routine steel I-girder bridges covered by this Guide are assumed to contain
only constant-depth members.
To satisfy this limit state without providing bearing stiffeners, webs subject to compressive or
tensile concentrated loads must satisfy Eq. D6.5.2-1. The nominal resistance to the concentrated
loading, Rn, in Eq. D6.5.2-1 is determined from Eq. D6.5.2-2 or D6.5.2-3. Eq. D6.5.2-2 applies if
the factored bearing reaction, Ru, is an interior-pier reaction or if the factored concentrated load,
Ru, is applied at a distance from the end of the member that is greater than d, where d is the depth
of the member. Otherwise, Eq. D6.5.2-3 applies.
The concentrated load acting on a rolled shape or built-up section is assumed critical at the toe of
the fillet, located a distance k from the outer face of the flange resisting the concentrated load or
bearing reaction. For a rolled shape, k is published in the tables in the AISC Manual of Steel
Construction giving dimensions for the shapes. For a built-up section, k may be taken as the
distance from the outer face of the flange to the web toe of the web-to-flange fillet weld.
For an interior concentrated load or interior-pier bearing reaction, the load is assumed to distribute
along the web at a slope of 2.5 to 1 and over a distance of (5k + N), where N is the length of bearing.
N must be greater than or equal to k at end bearing locations. An interior concentrated load is
assumed to be a load applied at a distance from the end of the member greater than d. For an end
concentrated load or end reaction, the load is assumed to distribute along the web at the same slope
over a distance of (2.5k + N).
Simple Span Bridges:
The provisions of this Article are not applicable to the routine simple span plate-girder bridges
covered by this Guide at support reactions, at which bearing stiffeners are required (see the
Discussion of Article 6.10.11.2 in this Guide). These provisions are applicable to determine if
bearing stiffeners are required at support reactions in routine simple span rolled-beam bridges.
These provisions are also applicable for the routine simple-span plate-girder or rolled-beam
bridges covered by this Guide to determine if bearing stiffeners are required at jacking points on
the beam/girder or a diaphragm at the end supports (if provided).
Multi-span Continuous Rolled Beam Bridges:
The provisions of this Article are applicable to the routine multi-span continuous rolled beam
bridges covered by this Guide to determine if bearing stiffeners are required at support reactions.
These provisions are also applicable for these bridges to determine if bearing stiffeners are required
at jacking points on the beam or diaphragm at an end and/or interior support (if provided).
Multi-span Continuous Plate Girder Bridges:

NSBA Guide to Navigating Routine Steel Bridge Design / 408


The provisions of this Article are not applicable to the routine multi-span continuous plate-girder
bridges covered by this Guide at support reactions, at which bearing stiffeners are required (see
the Discussion of Article 6.10.11.2 in this Guide). These provisions are applicable for these bridges
to determine if bearing stiffeners are required at jacking points on the girder or diaphragm at an
end and/or interior support (if provided).

D6.5.3 Web Crippling


Determination of applicability, Simple Span Bridges: Conditionally applicable.
Determination of applicability, Multi-span Continuous Rolled Beam Bridges: Applicable.
Determination of applicability, Multi-span Continuous Plate Girder Bridges: Conditionally
applicable.
Discussion:
The limit state of web crippling is intended to prevent local instability or crippling of the web due
to a high compressive stress caused by a concentrated load or bearing reaction.
To satisfy this limit state without providing bearing stiffeners, webs subject to compressive
concentrated loads must satisfy Eq. D6.5.3-1. The nominal resistance to the concentrated loading,
Rn, is determined from Eq. D6.5.3-2, D6.5.2-3, or D6.5.2-4. Eq. D6.5.3-2 applies if the factored
bearing reaction, Ru, is an interior-pier reaction or if the factored concentrated load, Ru, is applied
at a distance from the end of the member that is greater than or equal to d/2, where d is the depth
of the member. Otherwise, Eq. D6.5.3-3 or D.6.5.3-4 applies depending on the ratio of N/d, where
N is the length of bearing. N must be greater than or equal to k at end bearing locations. Eq. D6.5.3-
3 applies if the ratio of N/d is less than or equal to 0.2; otherwise, Eq. D6.5.3-4 applies.
Simple Span Bridges:
The provisions of this Article are not applicable to the routine simple span plate-girder bridges
covered by this Guide at support reactions, at which bearing stiffeners are required (see the
Discussion of Article 6.10.11.2 in this Guide). These provisions are applicable to determine if
bearing stiffeners are required at support reactions in routine simple span rolled-beam bridges.
These provisions are also applicable for the routine simple-span plate-girder or rolled-beam
bridges covered by this Guide to determine if bearing stiffeners are required at jacking points on
the beam/girder or a diaphragm at the end supports (if provided).
Multi-span Continuous Rolled Beam Bridges:
The provisions of this Article are applicable to the routine multi-span continuous rolled beam
bridges covered by this Guide to determine if bearing stiffeners are required at support reactions.
These provisions are also applicable for these bridges to determine if bearing stiffeners are required
at jacking points on the beam or diaphragm at an end and/or interior support (if provided).
Multi-span Continuous Plate Girder Bridges:
The provisions of this Article are not applicable to the routine multi-span continuous plate-girder
bridges covered by this Guide at support reactions, at which bearing stiffeners are required (see

NSBA Guide to Navigating Routine Steel Bridge Design / 409


the Discussion of Article 6.10.11.2 in this Guide). These provisions are applicable for these bridges
to determine if bearing stiffeners are required at jacking points on the girder or diaphragm at an
end and/or interior support (if provided).

APPENDIX E6 NOMINAL COMPRESSIVE RESISTANCE OF NONCOMPOSITE


MEMBERS CONTAINING LONGITUDINALLY STIFFENED
PLATES

E6.1 NOMINAL COMPRESSIVE RESISTANCE

E6.1.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to determine the nominal compressive resistance of a
noncomposite I-section or box-section member subject to axial compression that contains one or
more longitudinally stiffened plates.
These provisions are not applicable to the routine steel I-girder bridges covered by this Guide as
members in these bridges do not contain longitudinally stiffened plates and are subject to flexure
only.

E6.1.2 Classification of Longitudinally Stiffened Plate Panels


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to classify longitudinally stiffened plate panels as either
nonslender or slender in a noncomposite I-section or box-section member subject to axial
compression that contains one or more longitudinally stiffened plates.
These provisions are not applicable to the routine steel I-girder bridges covered by this Guide as
members in these bridges do not contain longitudinally stiffened plates and are subject to flexure
only.

E6.1.3 Nominal Compressive Resistance and Effective Area of Plates with Equally-
spaced Equal-size Longitudinal Stiffeners
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to determine the effective area of plates with equally-spaced
equal-size longitudinal stiffeners for calculating the nominal compressive resistance of a
noncomposite I-section or box-section member subject to axial compression that contains one or
more longitudinally stiffened plates.

NSBA Guide to Navigating Routine Steel Bridge Design / 410


These provisions are not applicable to the routine steel I-girder bridges covered by this Guide as
members in these bridges do not contain longitudinally stiffened plates and are subject to flexure
only.

E6.1.4 Longitudinal Stiffeners


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to design the longitudinal stiffeners in a noncomposite I-
section or box-section member subject to axial compression that contains one or more
longitudinally stiffened plates.
These provisions are not applicable to the routine steel I-girder bridges covered by this Guide as
members in these bridges do not contain longitudinally stiffened plates and are subject to flexure
only.

E6.1.5 Transverse Stiffeners

E6.1.5.1 General
Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to design transverse stiffeners when they are utilized to
enhance the resistance of a longitudinally stiffened plate in a noncomposite I-section or box-
section member subject to axial compression that contains one or more longitudinally stiffened
plates.
These provisions are not applicable to the routine steel I-girder bridges covered by this Guide as
members in these bridges do not contain longitudinally stiffened plates and are subject to flexure
only.

E6.1.5.2 Moment of Inertia


Determination of applicability, All Routine Steel I-girder Bridges: Not applicable.
Discussion:
The provisions of this Article are used to determine the moment of inertia requirements for
transverse stiffeners when they are utilized to enhance the resistance of a longitudinally stiffened
plate in a noncomposite I-section and box-section member subject to axial compression that
contains one or more longitudinally stiffened plates.
These provisions are not applicable to the routine steel I-girder bridges covered by this Guide as
members in these bridges do not contain longitudinally stiffened plates and are subject to flexure
only.

NSBA Guide to Navigating Routine Steel Bridge Design / 411


CONCLUSION
The reader is reminded that this Guide is not a substitute for the AASHTO LRFD BDS. This Guide
is meant to be used in conjunction with the AASHTO LRFD BDS, with the intent of aiding the
reader in navigating and applying its provisions. The Guide also references a number of freely
available industry practice documents that the reader is encouraged to consult for additional
information and guidance.

NSBA Guide to Navigating Routine Steel Bridge Design / 412

You might also like