LS-DYNA Theory Manual (Updated 13dec2016) PDF
LS-DYNA Theory Manual (Updated 13dec2016) PDF
LS-DYNA Theory Manual (Updated 13dec2016) PDF
Theory Manual
10/27/16 (r:8004)
LS-DYNA Dev
LSTC reserves the right to modify the material contained within this manual without prior notice.
The information and examples included herein are for illustrative purposes only and are not intended to be
exhaustive or all-inclusive. LSTC assumes no liability or responsibility whatsoever for any direct of indirect
damages or inaccuracies of any type or nature that could be deemed to have resulted from the use of this
manual.
Any reproduction, in whole or in part, of this manual is prohibited without the prior written approval of
LSTC. All requests to reproduce the contents hereof should be sent to [email protected].
AES. Copyright © 2001, Dr Brian Gladman < [email protected]>, Worcester, UK. All rights reserved.
LICENSE TERMS
The free distribution and use of this software in both source and binary form is allowed (with or without
changes) provided that:
1. distributions of this source code include the above copyright notice, this list of conditions and the
following disclaimer;
2. distributions in binary form include the above copyright notice, this list of conditions and the
following disclaimer in the documentation and/or other associated materials;
3. the copyright holder's name is not used to endorse products built using this software without
specific written permission.
DISCLAIMER
This software is provided 'as is' with no explicit or implied warranties in respect of any properties, including,
but not limited to, correctness and fitness for purpose.
Issue Date: 21/01/2002
This file contains the code for implementing the key schedule for AES (Rijndael) for block and key sizes of 16,
24, and 32 bytes.
LS-DYNA Theory Manual Table of Contents
Table of Contents
22.18 Material Model 18: Power Law Isotropic Plasticity ....................................................... 20-67
22.19 Material Model 19: Strain Rate Dependent Isotropic Plasticity..................................... 20-68
22.20 Material Model 20: Rigid ................................................................................................ 20-70
22.21 Material Model 21: Thermal Orthotropic Elastic ............................................................ 20-71
22.22 Material Model 22: Chang-Chang Composite Failure Model ........................................ 20-73
22.23 Material Model 23: Thermal Orthotropic Elastic with 12 Curves ................................... 20-75
22.24 Material Model 24: Piecewise Linear Isotropic Plasticity .............................................. 20-77
22.25 Material Model 25: Kinematic Hardening Cap Model ................................................... 20-79
22.26 Material Model 26: Crushable Foam ............................................................................. 20-84
22.27 Material Model 27: Incompressible Mooney-Rivlin Rubber .......................................... 20-88
22.27.1 Stress Update for Shell Elements ................................................................... 20-90
22.27.2 Derivation of the Continuum Tangent Stiffness.............................................. 20-92
22.27.3 The Algorithmic Tangent Stiffness for Shell Elements ................................... 20-93
22.28 Material Model 28: Resultant Plasticity ......................................................................... 20-95
22.29 Material Model 29: FORCE LIMITED Resultant Formulation......................................... 20-97
22.29.1 Internal Forces ................................................................................................ 20-99
22.29.2 Tangent Stiffness .......................................................................................... 20-102
22.30 Material Model 30: Closed-Form Update Shell Plasticity ........................................... 20-104
22.30.1 Mathematical Description of the Material Model.......................................... 20-104
22.30.2 Algorithmic Stress Update ............................................................................ 20-105
22.30.3 Tangent Stiffness Matrix ............................................................................... 20-106
22.31 Material Model 31: Slightly Compressible Rubber Model ........................................... 20-109
22.32 Material Model 32: Laminated Glass Model ................................................................ 20-110
22.33 Material Model 33: Barlat’s Anisotropic Plasticity Model ............................................ 20-111
22.34 Material Model 34: Fabric ............................................................................................ 20-112
22.35 Material Model 35: Kinematic/Isotropic Plastic Green-Naghdi Rate .......................... 20-114
22.36 Material Model 36: Barlat’s 3-Parameter Plasticity Model .......................................... 20-115
22.36.1 Material Tangent Stiffness ............................................................................ 20-116
22.36.2 Load curves in different directions ............................................................... 20-118
22.36.2.1 An introductory remark ................................................................. 20-118
22.36.2.2 The model ..................................................................................... 20-119
22.36.3 Variable Lankford parameters ...................................................................... 20-120
22.37 Material Model 37: Transversely Anisotropic Elastic-Plastic ...................................... 20-122
22.38 Material Model 38: Blatz-Ko Compressible Foam....................................................... 20-124
22.39 Material Model 39: Transversely Anisotropic Elastic-Plastic With FLD....................... 20-125
22.42 Material Model 42: Planar Anisotropic Plasticity Model .............................................. 20-126
22.51 Material Model 51: Temperature and Rate Dependent Plasticity ............................... 20-127
LS-DYNA DEV 10/27/16 (r:8004) 0-7 (Table of Contents)
Table of Contents LS-DYNA Theory Manual
22.142 Material Model 142: Transversely Anisotropic Crushable Foam ................................ 20-234
22.143 Material Model 143: Wood Model ............................................................................. 20-236
22.144 Material Model 144: Pitzer Crushable Foam ............................................................. 20-243
22.147 Material Model 147: FHWA Soil Model...................................................................... 20-244
22.154 Material Model 154: Deshpande-Fleck Foam ........................................................... 20-251
22.156 Material Model 156: Muscle ...................................................................................... 20-252
22.158 Material Model 158: Rate Sensitive Composite Fabric ............................................. 20-253
22.159 Material Model 159: Continuous Surface Cap Model ............................................... 20-254
22.161 Material Models 161 and 162: Composite MSC ....................................................... 20-260
22.163 Material Model 163: Modified Crushable Foam ........................................................ 20-266
22.164 Material Model 164: Brain Linear Viscoelastic........................................................... 20-267
22.166 Material Model 166: Moment Curvature Beam ......................................................... 20-268
22.169 Material Model 169: Arup Adhesive........................................................................... 20-270
22.170 Material Model 170: Resultant Anisotropic ................................................................ 20-271
22.175 Material Model 175: Viscoelastic Maxwell................................................................. 20-272
22.176 Material Model 176: Quasilinear Viscoelastic ............................................................ 20-273
22.177 Material Models 177 and 178: Hill Foam and Viscoelastic Hill Foam ....................... 20-274
22.177.1 Hyperelasticity Using the Principal Stretch Ratios ..................................... 20-274
22.177.2 Hill’s Strain Energy Function....................................................................... 20-275
22.177.3 Viscous Stress ............................................................................................ 20-277
22.177.4 Viscoelastic Stress Contribution................................................................. 20-277
22.177.5 Material Tangent Modulus for the Fully Integrated Brick ........................... 20-278
22.179 Material Models 179 and 180: Low Density Synthetic Foam .................................... 20-280
22.179.1 Hyperelasticity Using the Principal Stretch Ratios ..................................... 20-280
22.179.2 Strain Energy Function ............................................................................... 20-281
22.179.3 Modeling of the Hysteresis ......................................................................... 20-282
22.179.4 Viscous Stress ............................................................................................ 20-282
22.179.5 Viscoelastic Stress Contribution................................................................. 20-283
22.179.6 Stress Corresponding to First Load Cycle ................................................. 20-283
22.181 Material Model 181: Simplified Rubber/Foam ........................................................... 20-285
22.181.1 Hyperelasticity Using the Principal Stretch Ratios ..................................... 20-285
22.181.1.1 Determination of f, rubber option ............................................... 20-286
22.181.1.2 Determination of f, foam option .................................................. 20-287
22.181.2 Some Remarks ........................................................................................... 20-288
22.181.2.1 Strain rates .................................................................................. 20-288
22.181.2.2 Modeling of the Frequency Independent Damping .................... 20-288
22.187 Material Model 187: Semi-Analytical Model for the Simulation of Polymers ............ 20-289
0-10 (Table of Contents) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Table of Contents
1
Abstract
LS-DYNA is a general purpose finite element code for analyzing the large
deformation static and dynamic response of structures including structures coupled to
fluids. The main solution methodology is based on explicit time integration. An implicit
solver is currently available with somewhat limited capabilities including structural
analysis and heat transfer. A contact-impact algorithm allows difficult contact problems to
be easily treated with heat transfer included across the contact interfaces. By a
specialization of this algorithm, such interfaces can be rigidly tied to admit variable zoning
without the need of mesh transition regions. Other specializations, allow draw beads in
metal stamping applications to be easily modeled simply by defining a line of nodes along
the draw bead. Spatial discretization is achieved by the use of four node tetrahedron and
eight node solid elements, two node beam elements, three and four node shell elements,
eight node solid shell elements, truss elements, membrane elements, discrete elements, and
rigid bodies. A variety of element formulations are available for each element type.
Specialized capabilities for airbags, sensors, and seatbelts have tailored LS-DYNA for
applications in the automotive industry. Adaptive remeshing is available for shell
elements and is widely used in sheet metal stamping applications. LS-DYNA currently
contains approximately one-hundred constitutive models and ten equations-of-state to
cover a wide range of material behavior. This theoretical manual has been written to
provide users and potential users with insight into the mathematical and physical basis of
the code.
2
History of LS-DYNA
The origin of LS-DYNA dates back to the public domain software, DYNA3D, which
was developed in the mid-seventies at the Lawrence Livermore National Laboratory. The
first version of DYNA3D [Hallquist 1976a] was released in 1976 with constant stress 4- or 8-
node solid elements, 16- and 20-node solid elements with 2 × 2 × 2 Gaussian quadrature, 3,
4, and 8-node membrane elements, and a 2-node cable element. A nodal constraint contact-
impact interface algorithm [Hallquist 1977] was available. On the Control Data CDC-7600,
a supercomputer in 1976, the speed of the code varied from 36 minutes per 106 mesh cycles
with 4-8 node solids to 180 minutes per 106 mesh cycles with 16 and 20 node solids.
Without hourglass control to prevent formation of non-physical zero energy deformation
modes, constant stress solids were processed at 12 minutes per 106 mesh cycles. A
moderate number of very costly solutions were obtained with this version of DYNA3D
using 16- and 20-node solids. Hourglass modes combined with the procedure for
computing the time step size prevented us from obtaining solutions with constant stress
elements.
The next version, released in 1979, achieved the aforementioned goals. On the
CRAY the vectorized speed was 50 times faster, 0.67 minutes per million mesh cycles. A
The 1981 version [Hallquist 1981a] evolved from the 1979 version. Nine additional
material models were added to allow a much broader range of problems to be modeled
including explosive-structure and soil-structure interactions. Body force loads were
implemented for angular velocities and base accelerations. A link was also established
from the 3D Eulerian code JOY [Couch, et. al., 1983] for studying the structural response to
impacts by penetrating projectiles. An option was provided for storing element data on
disk thereby doubling the capacity of DYNA3D.
The 1982 version of DYNA3D [Hallquist 1982] accepted DYNA2D [Hallquist 1980]
material input directly. The new organization was such that equations of state and
constitutive models of any complexity could be easily added. Complete vectorization of
the material models had been nearly achieved with about a 10 percent increase in execution
speed over the 1981 version.
In the 1986 version of DYNA3D [Hallquist and Benson 1986], many new features
were added, including beams, shells, rigid bodies, single surface contact, interface friction,
discrete springs and dampers, optional hourglass treatments, optional exact volume
integration, and VAX/VMS, IBM, UNIX, COS operating systems compatibility, that greatly
expanded its range of applications. DYNA3D thus became the first code to have a general
single surface contact algorithm.
In the 1987 version of DYNA3D [Hallquist and Benson 1987] metal forming
simulations and composite analysis became a reality. This version included shell thickness
changes, the Belytschko-Tsay shell element [Belytschko and Tsay, 1981], and dynamic
relaxation. Also included were non-reflecting boundaries, user specified integration rules
for shell and beam elements, a layered composite damage model, and single point
constraints.
New capabilities added in the 1988 DYNA3D [Hallquist 1988] version included a
cost effective resultant beam element, a truss element, a C0 triangular shell, the BCIZ
triangular shell [Bazeley et al., 1965], mixing of element formulations in calculations,
composite failure modeling for solids, noniterative plane stress plasticity, contact surfaces
with spot welds, tiebreak sliding surfaces, beam surface contact, finite stonewalls,
stonewall reaction forces, energy calculations for all elements, a crushable foam
constitutive model, comment cards in the input, and one-dimensional slidelines.
In 1988 the Hallquist began working half-time at LLNL to devote more time to the
development and support of LS-DYNA for automotive applications. By the end of 1988 it
was obvious that a much more concentrated effort would be required in the development
of LS-DYNA if problems in crashworthiness were to be properly solved; therefore, at the
start of 1989 the Hallquist resigned from LLNL to continue code development full time at
Livermore Software Technology Corporation. The 1989 version introduced many
enhanced capabilities including a one-way treatment of slide surfaces with voids and
friction; cross-sectional forces for structural elements; an optional user specified minimum
time step size for shell elements using elastic and elastoplastic material models; nodal
accelerations in the time history database; a compressible Mooney-Rivlin material model; a
closed-form update shell plasticity model; a general rubber material model; unique penalty
specifications for each slide surface; external work tracking; optional time step criterion for
4-node shell elements; and internal element sorting to allow full vectorization of right-
hand-side force assembly.
• MPGS database,
• MOVIE database,
• Slideline interface file,
• automated contact input for all input types,
• automatic single surface contact without element orientation,
• constraint technique for contact,
• cut planes for resultant forces,
• crushable cellular foams,
• urethane foam model with hysteresis,
• subcycling,
• friction in the contact entities,
• strains computed and written for the 8 node thick shells,
• “good” 4 node tetrahedron solid element with nodal rotations,
• 8 node solid element with nodal rotations,
• 2 2 integration for the membrane element,
• Belytschko-Schwer integrated beam,
• thin-walled Belytschko-Schwer integrated beam,
• improved LS-DYNA database control,
• null material for beams to display springs and seatbelts in TAURUS,
• parallel implementation on Cray and SGI computers,
• coupling to rigid body codes,
• seat belt capability.
• Cable modeling,
• Airbag reference geometry,
• Multiple jet model,
• Generalized joint stiffnesses,
• Enhanced rigid body to rigid body contact,
• Orthotropic rigid walls,
• Time zero mass scaling,
• Coupling with USA (Underwater Shock Analysis),
• Layered spot welds with failure based on resultants or plastic strain,
• Fillet welds with failure,
• Butt welds with failure,
• Automatic eroding contact,
• Edge-to-edge contact,
• Automatic mesh generation with contact entities,
• Drawbead modeling,
• Shells constrained inside brick elements,
• NIKE3D coupling for springback,
• Barlat’s anisotropic plasticity,
• Superplastic forming option,
• Rigid body stoppers,
• Keyword input,
• Adaptivity,
• First MPP (Massively Parallel) version with limited capabilities.
• Built in least squares fit for rubber model constitutive constants,
• Large hystersis in hyperelastic foam,
• Bilhku/Dubois foam model,
• Generalized rubber model,
• Bucket sort frequency can be controlled by a load curve for airbag applications.
• In automatic contact each part ID in the definition may have unique:
◦ Static coefficient of friction
◦ Dynamic coefficient of friction
◦ Exponential decay coefficient
◦ Viscous friction coefficient
◦ Optional contact thickness
◦ Optional thickness scale factor
◦ Local penalty scale factor
• Automatic beam-to-beam, shell edge-to-beam, shell edge-to-shell edge and single
surface contact algorithm.
• Release criteria may be a multiple of the shell thickness in types a_3, a_5, a10, 13,
and 26 contact.
• Force transducers to obtain reaction forces in automatic contact definitions. Defined
manually via segments, or automatically via part ID’s.
• Searching depth can be defined as a function of time.
• Bucket sort frequency can be defined as a function of time.
• Interior contact for solid (foam) elements to prevent "negative volumes."
• Locking joint
• Temperature dependent heat capacity added to Wang-Nefske inflator models.
• Wang Hybrid inflator model [Wang, 1996] with jetting options and bag-to-bag
venting.
• Aspiration included in Wang’s hybrid model [Nucholtz, Wang, Wylie, 1996].
• Extended Wang’s hybrid inflator with a quadratic temperature variation for heat
capacities [Nusholtz, 1996].
• Fabric porosity added as part of the airbag constitutive model.
• Blockage of vent holes and fabric in contact with structure or itself considered in
venting with leakage of gas.
• Option to delay airbag liner with using the reference geometry until the reference
area is reached.
• Birth time for the reference geometry.
• Multi-material Euler/ALE fluids,
◦ 2nd order accurate formulations.
◦ Automatic coupling to shell, brick, or beam elements
• The reference geometry option is extended for foam and rubber materials and can
be used for stress initialization, see *INITIAL_FOAM_REFERENCE_GEOMETRY.
• A vehicle positioning option is available for setting the initial orientation and
velocities, see *INITIAL_VEHICLE_KINEMATICS.
• A boundary element method is available for incompressible fluid dynamics
problems.
• The thermal materials work with instantaneous coefficients of thermal expansion:
◦ *MAT_ELASTIC_PLASTIC_THERMAL
◦ *MAT_ORTHOTROPIC_THERMAL
◦ *MAT_TEMPERATURE_DEPENDENT_ORTHOTROPIC
◦ *MAT_ELASTIC_WITH_VISCOSITY.
• Airbag interaction flow rate versus pressure differences.
• Contact segment search option, [bricks first optional]
• A through thickness Gauss integration rule with 1-10 points is available for shell
elements. Previously, 5 were available.
• Shell element formulations can be changed in a full deck restart.
• The tied interface which is based on constraint equations, TIED_SURFACE_TO_-
SURFACE, can now fail with FAILURE option.
• A general failure criteria for solid elements is independent of the material type, see
*MAT_ADD_EROSION
• Load curve control can be based on thinning and a flow limit diagram, see *DE-
FINE_CURVE_FEEDBACK.
• An option to filter the spotweld resultant forces prior to checking for failure has
been added the option, *CONSTRAINED_SPOTWELD, by appending,_FILTERED_-
FORCE, to the keyword.
• Bulk viscosity is available for shell types 1, 2, 10, and 16.
• When defining the local coordinate system for the rigid body inertia tensor a local
coordinate system ID can be used. This simplifies dummy positioning.
• Prescribing displacements, velocities, and accelerations is now possible for rigid
body nodes.
• One-way flow is optional for segmented airbag interactions.
• Pressure time history input for airbag type, LINEAR_FLUID, can be used.
• An option is available to independently scale system damping by part ID in each of
the global directions.
• An option is available to independently scale global system damping in each of the
global directions.
2-12 (History of LS-DYNA) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual History of LS-DYNA
• Added option to constrain global DOF along lines parallel with the global axes. The
keyword is *CONSTRAINED_GLOBAL. This option is useful for adaptive remesh-
ing.
• Beam end code releases are available, see *ELEMENT_BEAM.
• An initial force can be directly defined for the cable material, *MAT_CABLE_DIS-
CRETE_BEAM. The specification of slack is not required if this option is used.
• Airbag pop pressure can be activated by accelerometers.
• Termination may now be controlled by contact, via *TERMINATION_CONTACT.
• Modified shell elements types 8, 10 and the warping stiffness option in the
Belytschko-Tsay shell to ensure orthogonality with rigid body motions in the event
that the shell is badly warped. This is optional in the Belytschko-Tsay shell and the
type 10 shell.
• A one point quadrature brick element with an exact hourglass stiffness matrix has
been implemented for implicit and explicit calculations.
• Automatic file length determination for d3plot binary database is now implement-
ed. This insures that at least a single state is contained in each d3plot file and elimi-
nates the problem with the states being split between files.
• The dump files, which can be very large, can be placed in another directory by
specifying d=/home/user /test/d3dump on the execution line.
• A print flag controls the output of data into the MATSUM and RBDOUT files by
part ID's. The option, PRINT, has been added as an option to the *PART keyword.
• Flag has been added to delete material data from the d3thdt file. See *DATABASE_-
EXTENT_BINARY and column 25 of the 19th control card in the structured input.
• After dynamic relaxation completes, a file is written giving the displaced state
which can be used for stress initialization in later runs.
◦ *MAT_GEPLASTIC_SRATE2000
◦ *MAT_ELASTIC_VISCOPLASTIC_THERMAL
◦ *MAT_COMPOSITE_LAYUP
◦ *MAT_COMPOSITE_LAYUP
◦ *MAT_COMPOSITE_direct
• for the solid elements:
◦ *MAT_JOHNSON_HOLMQUIST_CERAMICS
◦ *MAT_JOHNSON_HOLMQUIST_CONCRETE
◦ *MAT_INV_HYPERBOLIC_SIN
◦ *MAT_UNIFIED_CREEP
◦ *MAT_SOIL_BRICK
◦ *MAT_DRUCKER_PRAGER
◦ *MAT_RC_SHEAR_WALL
• and for all element options a very fast and efficient version of the Johnson-Cook
plasticity model is available:
◦ *MAT_SIMPLIFIED_JOHNSON_COOK
• A fully integrated version of the type 16 shell element is available for the resultant
constitutive models.
• A nonlocal failure theory is implemented for predicting failure in metallic materials.
The keyword *MAT_NONLOCAL activates this option for a subset of elastoplastic
constitutive models.
• A discrete Kirchhoff triangular shell element (DKT) for explicit analysis with three
in plane integration points is flagged as a type 17 shell element. This element has
much better bending behavior than the C0 triangular element.
• A discrete Kirchhoff linear triangular and quadrilaterial shell element is available as
a type 18 shell. This shell is for extracting normal modes and static analysis.
• A C0 linear 4-node quadrilaterial shell element is implemented as element type 20
with drilling stiffness for normal modes and static analysis.
• An assumed strain linear brick element is available for normal modes and statics.
• The fully integrated thick shell element has been extended for use in implicit
calculations.
• A fully integrated thick shell element based on an assumed strain formulation is
now available. This element uses a full 3D constitutive model which includes the
normal stress component and, therefore, does not use the plane stress assumption.
• The 4-node constant strain tetrahedron element has been extended for use in
implicit calculations.
• Relative damping between parts is available, see *DAMPING_RELATIVE (SMP
only).
• Preload forces are can be input for the discrete beam elements.
• Objective stress updates are implemented for the fully integrated brick shell
element.
• Acceleration time histories can be prescribed for rigid bodies.
• Prescribed motion for nodal rigid bodies is now possible.
• Generalized set definitions, i.e., SET_SHELL_GENERAL etc. provide much
flexibility in the set definitions.
• The command "sw4." will write a state into the dynamic relaxation file, D3DRLF,
during the dynamic relaxation phase if the d3drlf file is requested in the input.
• Added mass by PART ID is written into the matsum file when mass scaling is used
to maintain the time step size, (SMP version only).
• Upon termination due to a large mass increase during a mass scaled calculation a
print summary of 20 nodes with the maximum added mass is printed.
• Eigenvalue analysis of models containing rigid bodies is now available using
BCSLIB-EXT solvers from Boeing. (SMP version only).
• Second order stress updates can be activated by part ID instead of globally on the
*CONTROL_ACCURACY input.
• Interface frictional energy is optionally computed for heat generation and is output
into the interface force file (SMP version only).
• The interface force binary database now includes the distance from the contact
surface for the FORMING contact options. This distance is given after the nodes are
detected as possible contact candidates. (SMP version only).
• Type 14 acoustic brick element is implemented. This element is a fully integrated
version of type 8, the acoustic element (SMP version only).
• A flooded surface option for acoustic applications is available (SMP version only).
• Attachment nodes can be defined for rigid bodies. This option is useful for NVH
applications.
• CONSTRAINED_POINTS tie any two points together. These points must lie on a
shell element.
• Soft constraint is available for edge-to-edge contact in type 26 contact.
• CONSTAINED_INTERPOLATION option for beam to solid interfaces and for
spreading the mass and loads. (SMP version only).
• A database option has been added that allows the output of added mass for shell
elements instead of the time step size.
• A new contact option allows the inclusion of all internal shell edges in contact type
*CONTACT_GENERAL, type 26. This option is activated by adding INTERIOR
option.
• A new option allows the use deviatoric strain rates rather than total rates in material
model 24 for the Cowper-Symonds rate model.
• The CADFEM option for ASCII databases is now the default. Their option includes
more significant figures in the output files.
• When using deformable spot welds, the added mass for spot welds is now printed
for the case where global mass scaling is activated. This output is in the log file,
d3hsp file, and the messag file.
• Initial penetration warnings for edge-to-edge contact are now written into the
MESSAG file and the D3HSP file.
• Each compilation of LS-DYNA is given a unique version number.
• Finite length discrete beams with various local axes options are now available for
material types 66, 67, 68, 93, and 95. In this implementation the absolute value of
SCOOR must be set to 2 or 3 in the *SECTION_BEAM input.
• New discrete element constitutive models are available:
◦ *MAT_ELASTIC_SPRING_DISCRETE_BEAM
◦ *MAT_INELASTIC_SPRING_DISCRETE_BEAM
◦ *MAT_ELASTIC_6DOF_SPRING_DISCRETE_BEAM
◦ *MAT_INELASTIC_6DOF_SPRING_DISCRETE_BEAM
The latter two can be used as finite length beams with local coordinate systems.
• Moving SPC's are optional in that the constraints are applied in a local system that
rotates with the 3 defining nodes.
• A moving local coordinate system, CID, can be used to determine orientation of
discrete beam elements.
• Modal superposition analysis can be performed after an eigenvalue analysis. Stress
recovery is based on type 18 shell and brick (SMP only).
• Rayleigh damping input factor is now input as a fraction of critical damping, i.e.
0.10. The old method required the frequency of interest and could be highly unsta-
ble for large input values.
• Airbag option "SIMPLE_PRESSURE_VOLUME" allows for the constant CN to be
replaced by a load curve for initialization. Also, another load curve can be defined
which allows CN to vary as a function of time during dynamic relaxation. After
dynamic relaxation CN can be used as a fixed constant or load curve.
• Hybrid inflator model utilizing CHEMKIN and NIST databases is now available.
Up to ten gases can be mixed.
• Option to track initial penetrations has been added in the automatic SMP contact
types rather than moving the nodes back to the surface. This option has been avail-
able in the MPP contact for some time. This input can be defined on the fourth card
of the *CONTROL_CONTACT input and on each contact definition on the third
optional card in the *CONTACT definitions.
• If the average acceleration flag is active, the average acceleration for rigid body
nodes is now written into the d3thdt and nodout files. In previous versions of LS-
DYNA, the accelerations on rigid nodes were not averaged.
• A capability to initialize the thickness and plastic strain in the crash model is
available through the option *INCLUDE_STAMPED_PART, which takes the results
from the LS-DYNA stamping simulation and maps the thickness and strain distribu-
tion onto the same part with a different mesh pattern.
• A capability to include finite element data from other models is available through
the option, *INCLUDE_TRANSFORM. This option will take the model defined in
an INCLUDE file: offset all ID's; translate, rotate, and scale the coordinates; and
transform the constitutive constants to another set of units.
• Penetration warnings for the contact option, “ignore initial penetration,” are added
as an option. Previously, no penetration warnings were written when this contact
option was activated.
• Penetration warnings for nodes in-plane with shell mid-surface are printed for the
AUTOMATIC contact options. Previously, these nodes were ignored since it was
assumed that they belonged to a tied interface where an offset was not used; conse-
quently, they should not be treated in contact.
• For the arbitrary spot weld option, the spot welded nodes and their contact
segments are optionally written into the d3hsp file. See *CONTROL_CONTACT.
• For the arbitrary spot weld option, if a segment cannot be found for the spot welded
node, an option now exists to error terminate. See *CONTROL_CONTACT.
• Spot weld resultant forces are written into the swforc file for solid elements used as
spot welds.
• Solid materials have now been added to the failed element report and additional
information is written for the “node is deleted” messages.
• A new option for terminating a calculation is available, *TERMINATION_CURVE.
• A 10-noded tetrahedron solid element is available with either a 4 or 5 point
integration rule. This element can also be used for implicit solutions.
• A new 4 node linear shell element is available that is based on Wilson’s plate
element combined with a Pian-Sumihara membrane element. This is shell type 21.
• A shear panel element has been added for linear applications. This is shell type 22.
This element can also be used for implicit solutions.
• A null beam element for visualization is available. The keyword to define this null
beam is *ELEMENT_PLOTEL. This element is necessary for compatibility with
NASTRAN.
• A scalar node can be defined for spring-mass systems. The keyword to define this
node is *NODE_SCALAR. This node can have from 1 to 6 scalar degrees-of-
freedom.
• A thermal shell has been added for through-thickness heat conduction. Internally, 8
additional nodes are created, four above and four below the mid-surface of the shell
element. A quadratic temperature field is modeled through the shell thickness.
Internally, the thermal shell is a 12 node solid element.
• A beam OFFSET option is available for the *ELEMENT_BEAM definition to permit
the beam to be offset from its defining nodal points. This has the advantage that all
beam formulations can now be used as shell stiffeners.
• A beam ORIENTATION option for orienting the beams by a vector instead of the
third node is available in the *ELEMENT_BEAM definition for NASTRAN compati-
bility.
• Non-structural mass has been added to beam elements for modeling trim mass and
for NASTRAN compatibility.
• An optional checking of shell elements to avoid abnormal terminations is available.
See *CONTROL_SHELL. If this option is active, every shell is checked each time
step to see if the distortion is so large that the element will invert, which will result
in an abnormal termination. If a bad shell is detected, either the shell will be deleted
or the calculation will terminate. The latter is controlled by the input.
• An offset option is added to the inertia definition. See *ELEMENT_INERTIA_OFF-
SET keyword. This allows the inertia tensor to be offset from the nodal point.
• Plastic strain and thickness initialization is added to the draw bead contact option.
See *CONTACT_DRAWBEAD_INITIALIZE.
• Tied contact with offsets based on both constraint equations and beam elements for
solid elements and shell elements that have 3 and 6 degrees-of-freedom per node,
respectively. See BEAM_OFFSET and CONSTRAINED_OFFSET contact options.
These options will not cause problems for rigid body motions.
• The segment-based (SOFT = 2) contact is implemented for MPP calculations. This
enables airbags to be easily deployed on the MPP version.
• Improvements are made to segment-based contact for edge-to-edge and sliding
conditions, and for contact conditions involving warped segments.
• An improved interior contact has been implemented to handle large shear
deformations in the solid elements. A special interior contact algorithm is available
for tetrahedron elements.
• Coupling with MADYMO 6.0 uses an extended coupling that allows users to link
most MADYMO geometric entities with LS-DYNA FEM simulations. In this cou-
pling MADYMO contact algorithms are used to calculate interface forces between
the two models.
• Release flags for degrees-of-freedom for nodal points within nodal rigid bodies are
available. This makes the nodal rigid body option nearly compatible with the RBE2
option in NASTRAN.
• Fast updates of rigid bodies for metal forming applications can now be accom-
plished by ignoring the rotational degrees-of-freedom in the rigid bodies that are
typically inactive during sheet metal stamping simulations. See the keyword:
*CONTROL_RIGID.
• Center of mass constraints can be imposed on nodal rigid bodies with the SPC
option in either a local or a global coordinate system.
• Joint failure based on resultant forces and moments can now be used to simulate the
failure of joints.
• CONSTRAINED_JOINT_STIFFNESS now has a TRANSLATIONAL option for the
translational and cylindrical joints.
• Joint friction has been added using table look-up so that the frictional moment can
now be a function of the resultant translational force.
• The nodal constraint options *CONSTRAINED_INTERPOLATION and *CON-
STRAINED_LINEAR now have a local option to allow these constraints to be ap-
plied in a local coordinate system.
• Mesh coarsening can now be applied to automotive crash models at the beginning
of an analysis to reduce computation times. See the new keyword: *CONTROL_-
COARSEN.
• Force versus time seatbelt pretensioner option has been added.
• Both static and dynamic coefficients of friction are available for seat belt slip rings.
Previously, only one friction constant could be defined.
• *MAT_SPOTWELD now includes a new failure model with rate effects as well as
additional failure options.
• Constitutive models added for the discrete beam elements:
◦ *MAT_1DOF_GENERALIZED_SPRING
◦ *MAT_GENERAL_NONLINEAR_6dof_DISCRETE_BEAM
◦ *MAT_GENERAL_NONLINEAR_1dof_DISCRETE_BEAM
◦ *MAT_GENERAL_SPRING_DISCRETE_BEAM
◦ *MAT_GENERAL_JOINT_DISCRETE_BEAM
◦ *MAT_SEISMIC_ISOLATOR
• for shell and solid elements:
◦ *MAT_PLASTICITY_WITH_DAMAGE_ORTHO
◦ *MAT_SIMPLIFIED_JOHNSON_COOK_ORTHOTROPIC_DAMAGE
◦ *MAT_HILL_3R
◦ *MAT_GURSON_RCDC
• for the solid elements:
◦ *MAT_SPOTWELD
◦ *MAT_HILL_FOAM
◦ *MAT_WOOD
◦ *MAT_VISCOELASTIC_HILL_FOAM
◦ *MAT_LOW_DENSITY_SYNTHETIC_FOAM
◦ *MAT_RATE_SENSITIVE_POLYMER
◦ *MAT_QUASILINEAR VISCOELASTIC
◦ *MAT_TRANSVERSELY_ANISOTROPIC_CRUSHABLE_FOAM
◦ *MAT_VACUUM
◦ *MAT_MODIFIED_CRUSHABLE_FOAM
◦ *MAT_PITZER_CRUSHABLE FOAM
◦ *MAT_JOINTED_ROCK
◦ *MAT_SIMPLIFIED_RUBBER
◦ *MAT_FHWA_SOIL
◦ *MAT_SCHWER_MURRAY_CAP_MODEL
• Failure time added to MAT_EROSION for solid elements.
• Damping in the material models *MAT_LOW_DENSITY_FOAM and *MAT_LOW_-
DENSITY_VISCOUS_FOAM can now be a tabulated function of the smallest stretch
ratio.
• The material model *MAT_PLASTICITY_WITH_DAMAGE allows the table
definitions for strain rate.
• Improvements in the option *INCLUDE_STAMPED_PART now allow all history
data to be mapped to the crash part from the stamped part. Also, symmetry planes
can be used to allow the use of a single stamping to initialize symmetric parts.
• Extensive improvements in trimming result in much better elements after the
trimming is completed. Also, trimming can be defined in either a local or global
coordinate system. This is a new option in *DEFINE_CURVE_TRIM.
• An option to move parts close before solving the contact problem is available, see
*CONTACT_AUTO_MOVE.
• An option to add or remove discrete beams during a calculation is available with the
new keyword: *PART_SENSOR.
• Multiple jetting is now available for the Hybrid and Chemkin airbag inflator
models.
• Nearly all constraint types are now handled for implicit solutions.
• Calculation of constraint and attachment modes can be easily done by using the
option: *CONTROL_IMPLICIT_MODES.
• Penalty option, see *CONTROL_CONTACT, now applies to all *RIGIDWALL
options and is always used when solving implicit problems.
• Solid elements types 3 and 4, the 4 and 8 node elements with 6 degrees-of-freedom
per node, are available for implicit solutions.
• The warping stiffness option for the Belytschko-Tsay shell is implemented for
implicit solutions. The Belytschko-Wong-Chang shell element is now available for
implicit applications. The full projection method is implemented due to it accuracy
over the drill projection.
3
Preliminaries
along an interior boundary 𝜕b3 when 𝑥𝑖+ = 𝑥𝑖− . Here 𝛔 is the Cauchy stress, 𝜌 is the current
density, 𝐟 is the body force density, and 𝐱̈ is acceleration. The comma on 𝜎𝑖𝑗,𝑗 denotes
covariant differentiation, and 𝑛𝑗 is a unit outward normal to a boundary element on ∂b.
is integrated in time and is used for evaluating equations of state and to track the global
energy balance. In Equation (3.9), 𝑠𝑖𝑗 and 𝑝 represent the deviatoric stresses and pressure,
𝑠𝑖𝑗 = 𝜎𝑖𝑗 + (𝑝 + 𝑞)𝛿𝑖𝑗 (3.10)
1
𝑝 = − 𝜎𝑖𝑗 𝛿𝑖𝑗 − 𝑞
3
(3.11)
1
= − 𝜎𝑘𝑘 − 𝑞
3
respectively, where 𝑞 is the bulk viscosity, 𝛿𝑖𝑗 is the Kronecker delta (𝛿𝑖𝑗 = 1 if 𝑖 = 𝑗;
otherwise 𝛿𝑖𝑗 = 0), and 𝜀̇𝑖𝑗 is the strain rate tensor. The strain rates and bulk viscosity are
discussed later.
X3
x3
x2 X2
∂b
t=0 b
∂B
B0
X1
x1
Figure 3.1. Notation.
We can write:
∫ (𝜌𝑥̈𝑖 − 𝜎𝑖𝑗,𝑗 − 𝜌𝑓 )𝛿𝑥𝑖 𝑑𝜐 + ∫ (𝜎𝑖𝑗 𝑛𝑗 − 𝑡𝑖 )𝛿𝑥𝑖 𝑑𝑠 + ∫ (𝜎𝑖𝑗+ − 𝜎𝑖𝑗− )𝑛𝑗 𝛿𝑥𝑖 𝑑𝑠 = 0 (3.12)
𝑣 𝜕𝑏1 𝜕𝑏3
where 𝛿𝑥𝑖 satisfies all boundary conditions on 𝜕𝑏2 , and the integrations are over the current
geometry. Application of the divergence theorem gives
The condition 𝛿𝜋 = 0 holds for all variations, 𝛿𝑥𝑖 , and, in particular, it holds for
variations along the shape functions. In each of the 3 Cartesian directions upon setting the
variation to one of the shape functions the weak form reduces to a necessary (but not
sufficient) condition that must be satisfied by any solution so that the
number of equations = 3 × number of nodes.
At this stage it is useful to introduce a vector space having dimension ℝ(number of nodes) with
a corresponding cartesian basis {𝐞′𝑖 }number
𝑖=1
of nodes
. Since the body is discretized into 𝑛
disjoint elements, the integral in (3.15) may be separated using the spatial additively of
integration into 𝑛 terms, one for each element
𝑛
𝛿𝜋 = ∑ 𝛿𝜋𝑚 = 0. (3.17)
𝑚=1
𝛿𝜋𝑚 = ∫ 𝜌𝑥̈𝑖 𝛿𝑥𝑖 𝑑𝜐 + ∫ 𝜎𝑖𝑗 𝛿𝑥𝑖,𝑗 𝑑𝜐 − ∫ 𝜌𝑓𝑖 𝛿𝑥𝑖 𝑑𝜐 − ∫ 𝑡𝑖 𝛿𝑥𝑖 𝑑𝑠. (3.18)
𝜐𝑚 𝜐𝑚 𝜐𝑚 ∂𝑏1 ∩𝜕𝑣𝑚
In which
𝑘
𝛎 = ∑ 𝑁𝑖 𝐞𝑛′ 𝑚 (𝑖)
𝑚
(3.20)
𝑖=1
4
Solid Elements
5
8
6
Node ξ η ζ
7 1 -1 -1 -1
2 1 -1 -1
3 1 1 -1
4 -1 1 -1
5 -1 -1 1
1 4
6 1 -1 1
7 1 1 1
8 -1 1 1
2
3
is approximated by
𝑛 𝑛 𝑛
∑ ∑ ∑ 𝑔𝑗𝑘𝑙 ∣𝐽𝑗𝑘𝑙 ∣𝑤𝑗 𝑤𝑘 𝑤𝑙 , (4.11)
𝑗=1 𝑘=1 𝑙=1
Flanagan Wilkins
DYNA3D Belytschko FDM
Strain displacement matrix 94 357 843
Strain rates 87 156
Force 117 195 270
Subtotal 298 708 1113
Hourglass control 130 620 680
Total 428 1328 1793
Table 4.1. Operation counts for a constant stress hexahedron (includes adds,
subtracts, multiplies, and divides in major subroutines, and is independent of
vectorization). Material subroutines will add as little as 60 operations for the
bilinear elastic-plastic routine to ten times as much for multi-surface plasticity and
reactive flow models. Unvectorized material models will increase that share of
the cost a factor of four or more.
𝜉1 = 𝜂1 = 𝜁1 = 0,
and we can write
problems where Poisson’s ratio approaches 0.5 lock up in the constant volume bending
modes. To preclude locking, an average pressure must be used over the elements;
consequently, the zero energy modes are resisted by the deviatoric stresses. if the
deviatoric stresses are insignificant relative to the pressure or, even worse, if material
failure cause loss of this stress state component, hourglassing will still occur, but with no
means of resisting it. Sometimes, however, the cost of the fully integrated element may be
justified by increased reliability and if used sparingly may actually increase the overall
speed.
p
Figure 4.2.2. Bending mode for a fully integrated brick.
𝜕𝑥𝑖 1
𝐽𝑖𝑗 = = 𝑥𝐼𝑖 (𝜉𝑗𝐼 + 𝜉𝑗𝑘𝐼 𝜉𝑘 + 𝜉𝑗𝑙𝐼 𝜉𝑙 + 𝜉123
𝐼
𝜉𝑘 𝜉𝑙 ), (4.2.20)
𝜕𝜉𝑗 8
where 𝑘 = 1 + mod(𝑗, 3) and 𝑙 = 1 + mod(𝑗 + 1,3). For future reference let
1
𝐽𝑖𝑗0 = 𝑥𝐼𝑖 (0) 𝜉𝑗𝐼 , (4.2.21)
8
be the jacobian evaluated in the element center and in the beginning of the simulation (i.e.,
at time zero). The velocity gradient computed directly from the shape functions and
velocity components is
𝜕𝑣𝑖 −1
𝐿𝑖𝑗 = = 𝐽𝑖𝑘̇ 𝐽𝑘𝑗 = 𝐵𝑖𝑗𝐼𝑘 𝑣𝐼𝑘 , (4.2.22)
𝜕𝑥𝑗
where
𝜕𝑁𝐼 −1
𝐵𝑖𝑗𝐼𝑘 = 𝐽 𝛿 , (4.2.23)
𝜕𝜉𝑙 𝑙𝑗 𝑖𝑘
is the gradient-displacement matrix and represents the element except for the alleviation of
volumetric locking. To do just that, let 𝐵0𝑖𝑗𝐼𝑘 be defined by
𝐽𝑖𝑘̅̇ 𝐽𝑘𝑗
̅ = 𝐵0𝑖𝑗𝐼𝑘 𝑣𝐼𝑘 ,
−1
(4.2.24)
with 𝐽𝑖𝑗̅ being the element averaged jacobian matrix, and construct the gradient-
displacement matrix used for the element as
1
̅̅̅̅ 𝑖𝑗𝐼𝑘 = 𝐵𝑖𝑗𝐼𝑘 + (𝐵0𝑙𝑙𝐼𝑘 − 𝐵𝑙𝑙𝐼𝑘 )𝛿𝑖𝑗 .
𝐵 (4.2.25)
3
This is what is often called the B-bar method.
2 1
𝐿𝑖𝑗 = 𝐽𝑖𝑗̇ = 𝑣𝐼𝑖 (𝜉𝑗𝐼 + 𝜉𝑗𝑘𝐼 𝜉𝑘 + 𝜉𝑗𝑙𝐼 𝜉𝑙 + 𝜉123
𝐼
𝜉𝑘 𝜉𝑙 ), (4.2.26)
𝑙𝑗 4𝑙𝑗
where, again, 𝑘 = 1 + mod(𝑗, 3) and 𝑙 = 1 + mod(𝑗 + 1,3). Now let 𝑖 ≠ 𝑝 ≠ 𝑞 ≠ 𝑖, then a
pure bending mode in the plane with normal in direction 𝑞 and about axis 𝑝 is represented
by
𝑣𝐼𝑖 = 𝜉𝑖𝑞𝐼 ,
𝑣𝐼𝑝 = 0, (4.2.27)
𝑣𝐼𝑞 = 0,
and thus the velocity gradient is given as
1
𝐿𝑖𝑗 = (𝜉 𝐼 𝜉 𝐼 𝜉 + 𝜉𝑖𝑞𝐼 𝜉𝑗𝑙𝐼 𝜉𝑙 ),
4𝑙𝑗 𝑖𝑞 𝑗𝑘 𝑘
(4.2.28)
𝐿𝑝𝑗 = 0,
𝐿𝑞𝑗 = 0,
for 𝑗 = 1, 2, 3. The nonzero expression above amounts to
1
𝐿𝑖𝑖 = 𝜉 ,
4𝑙𝑖 𝑞
𝐿𝑖𝑝 = 0, (4.2.29)
1
𝐿𝑖𝑞 = 𝜉.
4𝑙𝑞 𝑖
Notable here is that a pure bending mode gives arise to a transverse shear strain
represented by the last expression in the above. Assuming that 𝑙𝑞 is small compared to 𝑙𝑖
this may actually lock the element.
be the aspect ratio between dimensions 𝑚 and 𝑛 at time zero. The modified jacobian is
written
1
𝐽𝑖𝑗̃ = 𝑥𝐼𝑖 (𝜉𝑗𝐼 + 𝜉𝑗𝑘𝐼 𝜉𝑘𝑖 + 𝜉𝑗𝑙𝐼 𝜉𝑙𝑖 + 𝜉123
𝐼
𝜉𝑘𝑖 𝜉𝑙𝑖 ), (4.2.31)
8
where
𝜉𝑘 𝜅𝑗𝑘 𝑖≠𝑗
𝜉𝑘𝑖 = { , (4.2.32)
𝜉𝑘 otherwise
and
𝜉𝑙 𝜅𝑗𝑙 𝑖≠𝑗
𝜉𝑙𝑖 = { . (4.2.33)
𝜉𝑙 otherwise
where 𝐵̃𝑖𝑗𝐼𝑘 is the gradient-displacement matrix used for solid element type -2 in LS-DYNA.
The B-bar method is used to prevent volumetric locking.
If we assume that this is the geometry in the beginning of the simulation and that 𝑙𝑞
is smaller than 𝑙𝑖 the transverse shear strain can be expressed as
1
𝐿𝑖𝑞 = 𝜉𝑖 , (4.2.38)
4𝑙𝑖
meaning that the transverse shear energy is not affected by poor aspect ratios, i.e., the
transverse shear strain does not grow with decreasing 𝑙𝑞 .
slight modification of the jacobian matrix will substantially reduce the computational
expense for this element. Simply substitute the expressions for 𝜉𝑘𝑖 and 𝜉𝑙𝑖 by
𝜉𝑘𝑖 = 𝜉𝑘 𝜅𝑗𝑘 , (4.2.39)
and
𝜉𝑙𝑖 = 𝜉𝑙 𝜅𝑗𝑙 . (4.2.40)
This will lead to a stiffness reduction for certain modes, in particular the out-of-
plane hourglass mode as can be seen by once again looking at the transverse shear locking
example. The velocity gradient for pure bending is now
1
𝐿𝑖𝑖 = 𝜉𝑞 𝜅𝑖𝑞 ,
4𝑙𝑖
𝐿𝑖𝑝 = 0, (4.2.41)
1
𝐿𝑖𝑞 = 𝜉𝜅 ,
4𝑙𝑞 𝑖 𝑞𝑖
and if it turns out that 𝑙𝑖 is smaller than 𝑙𝑞 , then this results in
1
𝐿𝑖𝑖 = 𝜉 . (4.2.42)
4𝑙𝑞 𝑞
That is, if 𝑖 represents the direction of the thinnest dimension, its corresponding bending
strain is inadequately reduced.
4.2.6 Example
A plate of dimensions 10 × 5 × 1 mm3 is clamped at one end and subjected to a 1 Nm
torque at the other end. The Young’s modulus is 210 GPa and the analytical solution for
the end tip deflection is 0.57143 mm. In order to study the mesh convergence for the three
fully integrated bricks the plate is discretized into 2 × 1 × 1, 4 × 2 × 2, 8 × 4 × 4, 16 × 8 × 8and
finally 32 × 16 × 16 elements, all elements having the same aspect ratio of 5 × 1. The table
below shows the results for the different fully integrated elements, and indicates an
accuracy improvement for solid elements −1 and −2.
Discretization Solid element type 2 Solid element type -2 Solid element type -1
2x1x1 0.0564 (90.1%) 0.6711 (17.4%) 0.6751 (18.1%)
4x2x2 0.1699 (70.3%) 0.5466 (4.3%) 0.5522 (3.4%)
8x4x4 0.3469 (39.3%) 0.5472 (4.2%) 0.5500 (3.8%)
16x8x8 0.4820 (15.7%) 0.5516 (3.5%) 0.5527 (3.3%)
32x16x16 0.5340 (6.6%) 0.5535 (3.1%) 0.5540 (3.1%)
component of the global deformation modes and must be admissible. One way of resisting
undesirable hourglassing is with a viscous damping or small elastic stiffness capable of
stopping the formation of the anomalous modes but having a negligible affect on the stable
global modes. Two of the early three-dimensional algorithms for controlling the hourglass
modes were developed by Kosloff and Frazier [1974] and Wilkins et al. [1974].
Since the hourglass deformation modes are orthogonal to the strain calculations,
work done by the hourglass resistance is neglected in the energy equation. This may lead
to a slight loss of energy; however, hourglass control is always recommended for the under
integrated solid elements. The energy dissipated by the hourglass forces reacting against
the formations of the hourglass modes is tracked and reported in the output files matsum
and glstat.
It is easy to understand the reasons for the formation of the hourglass modes.
Consider the following strain rate calculations for the 8-node solid element
1 8 ∂𝜙𝑘 𝑘 ∂𝜙𝑘 𝑘
𝜀̇𝑖𝑗 = (∑ 𝑥̇ + 𝑥̇ ). (4.43)
2 𝑘=1 ∂𝑥𝑖 𝑗 ∂𝑥𝑗 𝑖
Whenever diagonally opposite nodes have identical velocities, i.e.,
𝑥̇𝑖1 = 𝑥̇𝑖7 , 𝑥̇𝑖2 = 𝑥̇𝑖8 , 𝑥̇𝑖3 = 𝑥̇𝑖5 , 𝑥̇𝑖4 = 𝑥̇𝑖6 , (4.44)
the strain rates are identically zero:
𝜀̇𝑖𝑗 = 0, (4.45)
1k 2k
3k 4k
Figure 4.3. The hourglass modes of an eight-node element with one integration
point are shown [Flanagan and Belytschko 1981]. A total of twelve modes exist.
due to the asymmetries in Equations (4.15). It is easy to prove the orthogonality of the
hourglass shape vectors, which are listed in Table 4 and shown in Figure 4.3 with the
derivatives of the shape functions:
8
∂𝜙𝑘
∑ 𝛤 = 0, 𝑖 = 1, 2, 3, 𝛼 = 1, 2, 3, 4. (4.46)
𝑘=1
∂𝑥𝑖 𝛼𝑘
The product of the base vectors with the nodal velocities is zero when the element velocity
field has no hourglass component,
8
ℎ𝑖𝛼 = ∑ 𝑥̇𝑖𝑘 𝛤𝛼𝑘 = 0. (4.47)
𝑘=1
are nonzero if hourglass modes are present. The 12 hourglass-resisting force vectors, 𝑓𝑖𝛼𝑘 are
𝑓𝑖𝛼𝑘 = 𝑎ℎ ℎ𝑖𝛼 𝛤𝛼𝑘 , (4.48)
where
2⁄ 𝑐
𝑎ℎ = 𝑄HG 𝜌𝑣e 3 , (4.49)
4
in which 𝑣e is the element volume, 𝑐 is the material sound speed, and 𝑄HG is a user-defined
constant usually set to a value between .05 and .15. Equation (1.21) is hourglass control
type 1 in the LS-DYNA User’s Manual.
A shortcoming of hourglass control type 1 is that the hourglass resisting forces of Equation
(1.21) are not orthogonal to linear velocity field when elements are not in the shape of
parallelpipeds. As a consequence, such elements can generate hourglass energy with a
constant strain field or rigid body rotation. Flanagan and Belytschko [1981] developed an
hourglass control that is orthogonal to all modes except the zero energy hourglass modes.
Instead of resisting components of the bilinear velocity field that are orthogonal to the
strain calculation, Flanagan and Belytschko resist components of the velocity field that are
not part of a fully linear field. They call this field, defined below, the hourglass velocity
field
HG LIN
𝑥̇𝑖𝑘 = 𝑥̇𝑖𝑘 − 𝑥̇𝑖𝑘 , (4.50)
where
LIN
𝑥̇𝑖𝑘 ̇ (𝑥𝑗𝑘 − 𝑥̅𝑗 ),
= 𝑥̅i̇ + 𝑥̅𝑖,𝑗 (4.51)
and
1 8
𝑥̅𝑖 = ∑ 𝑥𝑖𝑘 ,
8 𝑘=1
(4.52)
1 8 𝑘
̇
𝑥̅𝑖 = ∑ 𝑥̇𝑖 .
8 𝑘=1
Flanagan and Belytschko construct geometry-dependent hourglass shape vectors that are
orthogonal to the fully linear velocity field and the rigid body field. With these vectors
they resist the hourglass velocity deformations. Defining hourglass shape vectors in terms
of the base vectors as
8
𝛾𝛼𝑘 = 𝛤𝛼𝑘 − 𝜙𝑘,𝑖 ∑ 𝑥𝑖𝑛 𝛤𝛼𝑛 , (4.53)
𝑛=1
A cost comparison in Table 4.1 shows that the default type 1 hourglass viscosity
requires approximately 130 adds or multiplies per hexahedron, compared to 620 and 680
for the algorithms of Flanagan-Belytschko and Wilkins. Therefore, for a very regular mesh,
type 1 hourglass control may provide a faster, sufficiently accurate solution, but in general,
any of the other hourglass options which are all based on the 𝛾𝛼𝑘 terms of Equation (1.26)
will be a better choice.
Type 3 hourglass control is identical to type 2, except that the shape function
derivatives in Eq. (1.26) are evaluated at the centroid of the element rather than at the
origin of the referential coordinate system. With this method, Equation (1.14) produces the
exact element volume. However, the anti-symmetry property of Equation (1.15) is not true,
so there is some increased number of computations.
Types 4 and 5 hourglass control are similar to types 2 and 3, except that they
evaluate hourglass stiffness rather than viscosity. The hourglass rates of equation (1.27) are
multiplied by the solution time step to produce increments of hourglass deformation. The
hourglass stiffness is scaled by the element’s maximum frequency so that stability can is
maintained as long as the hourglass scale factor, 𝑎ℎ , is sufficiently small.
Type 6 hourglass control improves on type 5 by scaling the stiffness such that the
hourglass forces match those generated by a fully integrated element control by doing
closed form integration over the element volume scaling the hourglass stiffness by
matching the stabilization for the 3D hexahedral element is available for both implicit and
explicit solutions. Based on material properties and element geometry, this stiffness type
stabilization is developed by an assumed strain method [Belytschko and Bindeman 1993]
such that the element does not lock with nearly incompressible material. When the user
defined hourglass constant 𝑎h is set to 1.0, accurate coarse mesh bending stiffness is
obtained for elastic material. For nonlinear material, a smaller value of 𝑎h is suggested and
the default value is set to 0.1. In the implicit form, the assumed strain stabilization matrix
is:
𝐤 𝐤12 𝐤13
stab ⎡ 11 ⎤
𝐊 = 2𝜇𝑎h ⎢𝐤21 𝐤22 𝐤23 ⎥, (4.56)
⎣𝐤31 𝐤32 𝐤33 ⎦
where the 8 × 8 submatricies are calculated by:
1 1+𝜐 1
𝐤𝑖𝑖 ≡ 𝐻𝑖𝑖 [( ) (𝛄𝑗 𝛄𝑗T + 𝛄𝑘 𝛄𝑘T ) + ( ) 𝛄4 𝛄4T ] + (𝐻𝑗𝑗 + 𝐻𝑘𝑘 )𝛄𝑖 𝛄1T ,
1−𝜐 3 2 (4.57)
𝜐 1
𝐤𝑖𝑗 ≡ ,𝐻𝑖𝑗 [( ) 𝛄 𝛄T + 𝛄𝑖 𝛄𝑗T ]
1−𝜐 𝑗 𝑖 2
with,
2
𝐻𝑖𝑖 ≡ ∫(ℎ𝑗,𝑖 ) 𝑑𝑣 = ∫(ℎ𝑘,𝑖 )2 𝑑𝑣 = 3 ∫(ℎ4,𝑖 )2 𝑑𝑣,
𝑣 𝑣 𝑣 (4.58)
𝐻𝑖𝑗 ≡ ∫ ℎ𝑖,𝑗 ℎ𝑗,𝑖 𝑑𝑣,
𝑣
where,
ℎ1 = 𝜉𝜂 ℎ2 = 𝜂𝜁 ℎ3 = 𝜁𝜉 ℎ4 = 𝜉𝜂𝜁 , (4.59)
𝑖 𝑗 𝑘
1 2 3
1 3 2
2 3 1
2 1 3
3 1 2
3 2 1
1
𝐻𝑖𝑗 = (𝚲𝑘T 𝐱𝑘 ) 𝑖 ≠ 𝑗 ≠ 𝑘, (4.61)
3
Λ𝑖 are 8 × 1 matrices of the referential coordinates of the nodes as given in Figure 4.1, and
x𝑖 are 8 × 1 matrices of the nodal coordinates in the corotational system. For each material
type, a Poisson's ratio, 𝑣, and an effective shear modulus, 𝜇, is needed.
and
𝑔̇𝑖𝑖 = 𝜇[(𝐻𝑗𝑗 + 𝐻𝑘𝑘 )𝑞𝑖𝑖̇ + 𝐻𝑖𝑗 𝑞𝑗𝑗̇ + 𝐻𝑖𝑘 𝑞𝑘𝑘
̇ ],
1
𝑔̇𝑖𝑗 = ,2𝜇 [ 𝐻 𝑞 ̇ + 𝜐𝐻𝑘𝑘 𝑞𝑖𝑖̇ ] (4.64)
1 − 𝜐 𝑖𝑖 𝑖𝑗
1+𝜈
𝑔̇𝑖4 = ,2𝜇 ( ) 𝐻𝑖𝑖 𝑞𝑖4
̇
3
where,
̇ = (𝛄𝛼T 𝐱̇𝑖 ).
𝑞𝑖𝛼 (4.65)
Subscripts 𝑖, 𝑗, and 𝑘 are permuted as per Table 44.2. As with the implicit form, calculations
are done in a corotational coordinate system in order to use the simplified equations (4.60)
and (4.61).
Type 7 hourglass control is very similar to type 6 hourglass control but with one significant
difference. As seen in Equation (1.36), type 6 obtains the current value of the generalized
stress from the previous value and the current increment. The incremental method is
nearly always sufficiently accurate, but it is possible for hourglass modes to fail to spring
back to the initial element geometry since the hourglass stiffness varies as the H terms
given by Equations (1.33) and (1.34) are recalculated in the deformed configuration each
cycle. Type 7 hourglass control eliminates this possible error by calculating the total
hourglass deformation in each cycle. For type 7 hourglass control, Equations (1.37) are
rewritten using 𝑔 and 𝑞 in place of 𝑔̇ and 𝑞 ̇, and Equation (1.38) is replaced by (1.39).
𝑞𝑖𝛼 = (𝛄𝛼T 𝐱𝑖 ) − (𝛄0𝛼
T
𝐱0𝑖 ). (4.39)
In Equation (1.39), 𝐱𝟎𝒊 and 𝛄𝟎𝜶 are evaluated using the initial, undeformed nodal coordinate
values. Type 7 hourglass control is considerably slower than type 6, so it is not generally
recommended, but may be useful when the solution involves at least several cycles of
loading and unloading that involve large element deformation of elastic or hyperelastic
material.
Both type 6 and 7 hourglass control are stiffness type methods, but may have viscosity
added through the VDC parameter on the *HOURGLASS card. The VDC parameter scales
the added viscosity, and VDC = 1.0 corresponds approximately to critical damping. The
primary motivation for damping is to reduce high frequency oscillations. A small
percentage of critical damping should be sufficient for this, but it is also possible to add
supercritical damping along with a small value of QM to simulate a very viscous material
that springs back slowly.
The element formulation is that of Puso [2000], and is essentially the mean strain
hexahedral element by Flanagan and Belytschko [1981] in which the perturbation hourglass
control is substituted for an enhanced assumed strain stabilization force.
Given the matrices
1 −1 −1 −1 1 1 1 −1
⎡1⎤ ⎡ 1 −1 −1⎤ ⎡ 1 −1 −1 1⎤
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢1⎥ ⎢ 1 1 −1⎥ ⎢−1 −1 1 −1⎥
⎢1⎥ ⎢−1 1 −1 ⎥ ⎢−1 1 −1 1⎥
𝐒=⎢ ⎥
⎢1⎥ , 𝚵=⎢ ⎢−1 −1
⎥, 𝐇 = ⎢ ⎥, (4.66)
1⎥ ⎢−1 −1 1 1⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢1⎥ ⎢ 1 −1 1⎥ ⎢−1 1 −1 −1⎥
⎢1⎥ ⎢ 1 1 1⎥ ⎢ 1 1 1 1⎥
⎣1⎦ ⎣−1 1 1⎦ ⎣ 1 −1 −1 −1⎦
we can define the vector of shape functions as
1
𝐍(𝛏) = [𝐬 + 𝚵𝛏 + 𝐇𝐡(𝛏)], (4.67)
8
where
𝜂𝜍
𝜉 ⎡ 𝜉𝜍 ⎤
𝛏 = ⎢𝜂 ⎤
⎡
⎥, 𝐡(𝛏) = ⎢
⎢ ⎥. (4.68)
⎢ 𝜉𝜂 ⎥
⎥
⎣𝜍⎦
⎣𝜉𝜂𝜍⎦
The position vector
𝑥(𝛏)
⎡𝑦(𝛏)⎤
𝐱(𝛏) = ⎢ ⎥, (4.69)
⎣ 𝑧(𝛏)⎦
is for isoparametric finite elements given as
𝐱(𝛏) = 𝐗T 𝐍(𝛏), (4.70)
where
𝑥1 𝑦1 𝑧1
⎡𝑥2 𝑦2 𝑧2 ⎤
⎢𝑥 𝑦3 𝑧3 ⎥
⎢ 3 ⎥
⎢𝑥4 𝑦4 𝑧4 ⎥
𝐗=⎢
⎢𝑥5 𝑦5
⎥
𝑧5 ⎥, (4.71)
⎢𝑥 𝑦6 𝑧6 ⎥
⎢ 6 ⎥
⎢𝑥7 𝑦7 𝑧7 ⎥
⎣𝑥8 𝑦8 𝑧8 ⎦
is the matrix of nodal coordinates. The Jacobian matrix maps the isoparametric domain to
the physical domain as
∂𝐱(𝛏)
𝐉(𝛏) = , (4.72)
∂𝛏
and we find the Jabobian matrix at the element centroid to be
1
𝐉0 = 𝐉(0) = 𝐗T 𝚵. (4.73)
8
We may use this to rewrite the vector of shape functions partially in terms of the position
vector as
T
1 T T 1⎧
{ T ∂𝐡(𝛏) −1 −T ∂𝐡(𝛏) T ⎫}
= (𝐗̇ 𝐁 + 𝐁 𝐗̇) + ⎨𝐗̇ 𝚪 𝐉(𝛏) + 𝐉(𝛏) [ ] 𝚪 𝐗̇⎬
2 2{
⎩ ∂𝛏 ∂𝛏 }
⎭
⎧ T ⎫
1 T T 1 { ∂𝐡(𝛏) ∂𝐡(𝛏) T }
= (𝐗̇ 𝐁 + 𝐁 𝐗̇) + 𝐉(𝛏)−T ⎨𝐉(𝛏)T 𝐗̇T 𝚪 +[ ] 𝚪 𝐗̇𝐉(𝛏)⎬ 𝐉(𝛏)−1 ,
2 2 {
⎩ ∂𝛏 ∂𝛏 }
⎭
where we substitute the occurrences of the jacobian matrix 𝐉(ξ)with the following
expressions
⎧ T ⎫
1 T 1 ∂𝐡(𝛏) ∂𝐡(𝛏)
T
̂0 {𝐉T0 𝐗̇T 𝚪
𝛆̇ ≈ (𝐗̇ 𝐁 + 𝐁 𝐗̇) + 𝐉−T + [ ] 𝚪 T ̇ } −1
𝐗 𝐉 ̂ (4.84)
2 2 ⎨
{ ∂𝛏 ∂𝛏 0 ⎬ 𝐉0 ,
}
⎩ ⎭
where
∥𝐣1 ∥
⎡ ⎤
𝐉0̂ = ⎢ ∥𝐣2 ∥ ⎥, (4.85)
⎣ ∥𝐣3 ∥⎦
and j𝑖 is the i:th column in the matrix 𝐉0 . This last approximation is the key to the mesh
distortion insensitivity that characterizes the element.
Changing to Voigt notation, we define the stabilization portion of the strain rate as
𝛆̇𝑠 = 𝐉−1
̂ ̃ ̃
0 𝐁𝑠 (𝛏)𝐮̇, (4.86)
where now
∥𝐣1 ∥−2
⎡ ⎤
⎢ ⎥
⎢ ∥𝐣2 ∥−2 ⎥
⎢ ⎥
⎢ ∥𝐣3 ∥−2 ⎥
𝐉−1
0 =⎢ ⎥, (4.87)
⎢ −1
∥𝐣1 ∥ ∥𝐣2 ∥ −1 ⎥
⎢ ⎥
⎢ ∥𝐣3 ∥−1 ∥𝐣2 ∥−1 ⎥
⎢ ⎥
−1 −1
⎣ ∥𝐣1 ∥ ∥𝐣3 ∥ ⎦
γ2 𝜍 + γ3 𝜂 + γ4 𝜍𝜂 0 0
⎡ ⎤
⎢ 0 γ1 𝜍 + γ3 𝜉 + γ4 𝜍𝜉 0 ⎥
⎢ 0 0 γ1 𝜂 + γ2 𝜉 + γ4 𝜉𝜂⎥
̃ 𝑠 (𝛏) = ⎢
𝐁 ⎢
⎥
⎥, (4.88)
⎢ γ1 𝜍 γ2 𝜍 0 ⎥
⎢ 0 γ2 𝜉 γ3 𝜉 ⎥
⎢ ⎥
⎣ γ1 𝜂 0 γ3 𝜂 ⎦
and 𝐮̇̃ is the vector of nodal velocities transformed to the isoparametric system according to
𝐮̇̃ = 𝕵𝐮̇, (4.89)
where 𝕵 is the 24 × 24 matrix that transforms the 8 nodal velocity vectors to the
isoparametric domain given by
𝐉T0
⎡ ⎤
𝕵 = perm ⎢
⎢ ⋱ ⎥.
⎥ (4.90)
T
⎣ 𝐉0 ⎦
T
Moreover, 𝛄𝑖 is the ith row of 𝚪 . We have deliberately neglected terms that cause
parallelepiped finite elements to lock in shear.
transformation formulae between different stress and constitutive tensors. At this point we
are only interested in how to handle the stabilization portion of the strain rate field, the
constant part is only used to update the midpoint stress as usual. Because of orthogonality
properties of the involved matrices, it turns out that we may just insert the expression for
the stabilization strain rate field to get
𝑡
𝑉
𝑠
𝛿𝑊int ≈ ∫ 𝛿𝛆̃𝑠T ⎛
⎜∫ 𝑒 𝐉−T
̂ ̂ ̃ ⎞
𝜎 −1
0 𝐂 𝐉0 𝛆̇𝑠 𝑑𝜏 ⎟ 𝑑𝑉𝑝 , (4.95a)
𝑝 ⎝0 8 ⎠
̃ (𝛏)𝑇 ⎛ 𝑡 𝑉𝑒 −T 𝜎 −1
𝚩
= [𝛿𝐮̃ 𝑇
𝛿𝜶 T] ∫ [ 𝑠
𝑇
] ⎜∫ 𝐉0̂ 𝐂 𝐉0̂ [𝚩
̃ 𝑠 (𝛏)
𝐆̃ (𝛏)] [𝐮̇̃] 𝑑𝜏 ⎞
⎟ 𝑑𝑉𝑝 . (4.95b)
𝑝
̃
𝐆 (𝛏) ⎝0 8 𝛂̇ ⎠
The stabilization contribution to the internal force vector is given by
𝑡
𝐟 T ̃ 𝑠 (𝛏)T ⎛ 𝑉𝑒 −T 𝜎 −1
0 ] ∫ [𝚩
[ 𝑢 ] = [𝕵 ] ⎜∫ 𝐉 ̂ 𝐂 𝐉0̂ [𝚩 ̃ (𝛏)][𝕵 0][𝐮̇]𝑑𝜏 ⎞
̃ 𝑠 (𝛏) 𝐆 ⎟ 𝑑𝑉𝑝 . (4.96)
𝐟𝛼 0 𝐈 𝑝 𝐆 ̃ (𝛏)T ⎝ 8 0 0 𝐈 𝛂̇ ⎠
0
The implementation of the element is very similar to the implementation of the one
point integrated mean strain hexahedral by Flanagan and Belytschko [1981]. The hourglass
forces are calculated in a different manner.
From the midpoint stress update we get a bulk and shear modulus characterizing
the material at this specific point in time. From this we form the isotropic spatial tangent
modulus 𝐂𝛔 to be used for computing the stabilization force from Equation (4.96).
∂Δ𝑢 + ∂Δ𝑤
∂Δ𝑤 𝑛+1⁄2 𝑛+1⁄2
Δ𝜀𝑧𝑧 = + 𝜙, Δ𝜀𝑧𝑥 = ∂𝑧 ∂𝑥 ,
𝑛+1⁄2 2
∂𝑧
where 𝜙 modifies the normal strains to ensure that the total volumetric strain increment at
each integration point is identical
∂Δ𝑢 + ∂Δ𝑣 + ∂Δ𝑤
𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄2 (4.98)
𝑛+1⁄2 ∂𝑥 ∂𝑦 ∂𝑧
𝜙 = Δ𝜀𝑣 − ,
3
𝑛+1⁄2
and Δ𝜀𝑣 is the average volumetric strain increment in the midpoint geometry
Δ𝑢,. Δ𝑣, and Δ𝑤 are displacement increments in the x, y, and z directions, respectively, and
(𝑧𝑛 + 𝑧𝑛+1 )
𝑛+1⁄2
𝑧 = . (4.100c)
2
To satisfy the condition that rigid body rotations cause zero straining, it is necessary to use
the geometry at the mid-step in the evaluation of the strain increments. As the default, LS-
DYNA currently uses the geometry at step 𝑛 + 1 to save operations; however, for implicit
calculations the mid-step strain calculation is always recommended, and, for explicit
calculations, which involve rotating parts, the mid-step geometry should be used especially
if the number of revolutions is large. The mid-step geometry can be activated either
globally or for a subset of parts in the model by using the options on the control card,
*CONTROL_ACCURACY.
Since the bulk modulus is constant in the plastic and viscoelastic material models,
constant pressure solid elements result. In the thermoelastoplastic material, a constant
temperature is assumed over the element. In the soil and crushable foam material, an
average relative volume is computed for the element at time step 𝑛 + 1, and the pressure
and bulk modulus associated with this relative volume is used at each integration point.
For equations of state, one pressure evaluation is done per element and is added to the
deviatoric stress tensor at each integration point.
2 r
s
Figure 4.4. Four node tetrahedron.
̅̅̅ 𝑛𝑒 and
where ̅𝑩
4
6
r
2
s
1
5
Figure 4.5. Six node Pentahedron.
8
8
15
7 7
16
14
5 20 5
13
19 6
6
4 4
17 11
3 3
12 18
10
1 1
9 2 2
DOF ui, vi, wi DOF ui, vi, wi, θxi, θyi, θzi
Figure 4.6. The 20-node solid element is transformed to an 8-node solid with 6
degrees-of-freedom per node.
θz
w
z
k
v θy
x
u
θx i
Figure 4.7. A typical element edge is shown from [Yunus, Pawlak, and Cook,
1989].
The instantaneous velocity for a midside node 𝑘 is given as a function of the corner
node velocities as (See Figure 4.7),
1 𝑦𝑗 − 𝑦𝑖 𝑧𝑗 − 𝑧𝑖
𝑢̇𝑘 = (𝑢̇𝑖 + 𝑢̇𝑗 ) + (𝜃𝑧𝑗̇ − 𝜃𝑧𝑖̇ ) + ̇ − 𝜃𝑦𝑗
(𝜃𝑦𝑖 ̇ ),
2 8 8
1 𝑧𝑗 − 𝑧𝑖 𝑥 − 𝑥𝑖
𝑣̇𝑘 = (𝑣̇𝑖 + 𝑣̇𝑗 ) + ̇ − 𝜃𝑥𝑖
(𝜃𝑥𝑗 ̇ )+ 𝑗 (𝜃𝑧𝑖̇ − 𝜃𝑧𝑗̇ ), (4.112)
2 8 8
1 𝑥𝑗 − 𝑥𝑖 𝑦 − 𝑦𝑖
𝑤̇ 𝑘 = (𝑤̇ 𝑖 + 𝑤̇ 𝑗 ) + (𝜃𝑦𝑗̇ − 𝜃𝑦𝑖 ̇ )+ 𝑗 (𝜃𝑥𝑖 ̇ − 𝜃𝑥𝑗̇ ),
2 8 8
where 𝑢, 𝑣, 𝑤, 𝜃𝑥 , 𝜃𝑦 , and 𝜃z are the translational and rotational displacements in the global
𝑥, 𝑦, and 𝑧 directions. The velocity field for the twenty-node hexahedron element in terms
of the nodal velocities is:
⎧ 𝑢̇1 ⎫
{
{ ⋮ }}
{
{ 𝑢̇ }
{ 20 }
}
𝜙 𝜙2 … 𝜙20 0 0 … 0 0 0 … 0 { { 𝑣̇1 }
{𝑢̇ ⎫
⎧ } ⎡ 1 ⎤ }
⎨ 𝑣̇ ⎬ = ⎢0 0 … 0 𝜙1 𝜙2 … 𝜙20 0 0 … 0 ⎥ ⎨ ⋮ ⎬, (4.113)
{
⎩𝑤̇ }⎭ ⎣0 0 … 0 0 0 … 0 𝜙1 𝜙2 … 𝜙20 ⎦ {
{ 𝑣̇20 }
}
{
{ 𝑤̇ 1 }
}
{
{ ⋮ }}
{ }
⎩𝑤̇ 20 ⎭
where 𝜙𝑖 are given by [Bathe and Wilson 1976] as,
DOF ui, vi, wi DOF ui, vi, wi, θxi, θyi, θzi
The standard formulation for the twenty node solid element is used with the above trans-
formations. The element is integrated with a fourteen point integration rule [Cook 1974]:
1 1 1
∫ ∫ ∫ 𝑓 (𝜉 , 𝜂 , 𝜁 )𝑑𝜉𝑑𝜂𝑑𝜁 =
−1 −1 −1
Cook reports that this rule has nearly the same accuracy as the twenty-seven point Gauss
rule, which is very costly. The difference in cost between eight point and fourteen point
integration, though significant, is necessary to eliminate the zero energy modes.
8 7 8 7 8 7
5 6 5 5
6 6
3 3 4 3
4
4
1 2 1 1
2 2
8 8
7 7
5 6 5 6
4
3 4 3
1 2 1 2
Figure 4.10a. Construction of a hexahedron element with five tetrahedrons.
8 7 8 7 8 7
5 5 5 6
6 6
4 4
3 4
3 3
1 2 1 2 1 2
8 7 8 7 8 7
5 6 5 5 6
6
3 4 4
4 3 3
1 2 1 2 1 2
Figure 4. 10b. Construction of a hexahedron element with six tetrahedrons
Figures 4.10a and 4. 10b show the construction of a hexahedron element from five
and six tetrahedron elements, respectively. When two sides of the adjacent bricks made
from five tetrahedrons are together, it is likely that four unique triangular segments exist.
This creates a problem in LS-PREPOST, which uses the numbering as a basis for
eliminating interior polygons prior to display. Consequently, the graphics in the post-
processing phase can be considerably slower with the degeneration in Figure 4.10a.
However, marginally better results may be obtained with five tetrahedrons per hexahedron
due to a better constraint count.
Section 4.11.2 through 4.11.8 describes the theory for the hexahedral CPE element,
the tetrahedron is based on the same concepts except for the volumetric correction
presented in Section 4.11.5. For more details we refer to [1] and [2]. We end with two
examples in Sections 4.11.9 and 4.11.10.
d3
D3
L
d1
D2
d2
D1
H
W
4.11.2 Geometry
The geometry of the hexahedral CPE element is characterized by the three-dimensional
directors 𝐃𝑖 and 𝐝𝑖 , 𝑖 = 0,1, … ,7, where the formers are associated with the reference
configuration and the latters with the current configuration. The reciprocal vectors 𝐃𝑖 and
𝐝𝑖 , 𝑖 = 1,2,3, are such that
𝑗
𝐝𝑖 ⋅ 𝐝 𝑗 = 𝐃 𝑖 ⋅ 𝐃 𝑗 = 𝛿 𝑖 , 𝑖, 𝑗 = 1, 2, 3, (4.11.117)
and we also have that
|𝐃𝑖 | = 1, 𝑖 = 1, 2, 3, (4.11.118)
The coordinates 𝜃𝑖 , 𝑖 = 1,2,3, ranges between
−𝐻/2 ≤ 𝜃1 ≤ 𝐻/2, − 𝑊/2 ≤ 𝜃2 ≤ 𝑊/2, − 𝐿/2 ≤ 𝜃3 ≤ 𝐿/2, (4.11.119)
and the 𝐀 matrix is given such that
7 7
𝑗 𝑗
𝐃𝑖 = ∑ 𝐴𝑖 𝐗𝑗 , 𝐝𝑖 = ∑ 𝐴𝑖 𝐱𝑗 , 𝑖 = 0, 1, . . . , 7, (4.11.120)
𝑗=0 𝑗=0
where 𝐗𝑖 and 𝐱𝑖 are the nodal coordinates in the reference and current configuration,
respectively. Hence the 𝐀 matrix represents the mapping between the nodal coordinates
and the Cosserat point directors. The reference position vector 𝐗 is expressed as
7
𝐗 = 𝐗(𝜃1 , 𝜃2 , 𝜃3 ) = ∑ 𝑁 𝑗 (𝜃1 , 𝜃2 , 𝜃3 )𝐃𝑗 , (4.11.121)
𝑗=0
where 𝐅 ̅̅̅̅ is the volume averaged deformation gradient and thus represents the
homogeneous deformations whereas 𝛃𝑖 are measures of the inhomogeneous deformations.
As for the 𝐕𝑖 we have
𝐻2 𝑊2
𝑉 = 𝐻𝑊𝐿 (𝐃1 × 𝐃2 ⋅ 𝐃3 + 𝐃 × 𝐃5 ⋅ 𝐃1 + 𝐃
12 4 12 6
𝐿2
× 𝐃4 ⋅ 𝐃2 + 𝐃5 × 𝐃6 ⋅ 𝐃3 ) ,
12
𝐻2 𝑊2
𝐕1 = 𝑉 −1 𝐻𝑊𝐿 ( 𝐃5 × 𝐃1 + 𝐃 × 𝐃6 ) ,
12 12 2 (4.11.125)
2 2
𝐻 𝐿
𝐕2 = 𝑉 −1 𝐻𝑊𝐿 ( 𝐃1 × 𝐃4 + 𝐃6 × 𝐃3 ) ,
12 12
𝑊2 𝐿2
𝐕3 = 𝑉 −1 𝐻𝑊𝐿 ( 𝐃4 × 𝐃2 + 𝐃3 × 𝐃5 ) ,
12 12
𝐕4 = 𝟎,
̅̅̅̅ is given as
The velocity gradient consistent with 𝐅
4
̅̅̅̅̇𝐅
𝐋̅ = 𝐅 ̅̅̅̅−1 = 𝐋 + ∑(𝐝̇ 𝑗+3 − 𝐋𝐝𝑗+3 ) ⊗ 𝐕𝑗 𝐅
̅̅̅̅−1 , (4.11.126)
𝑗=1
where 𝐋 = 𝐅̇ 𝐅−1 , which in turn gives the rate –of-deformation and spin tensors as
1 1
𝛆̅̇ = (𝐋̅ + 𝐋̅ 𝑇 ), ̅̅ ̅̅̅ = (𝐋̅ − 𝐋̅ 𝑇 ).
𝛚 (4.11.127)
2 2
so we can rewrite 𝐋̅ as
3 4 4
𝐋̅ = ∑ 𝐝̇ 𝑖 ⊗ 𝐝𝑖 ⎛
⎜𝐅̅̅̅̅ − ∑ 𝐝𝑗+3 ⊗ 𝐕𝑗 ⎞
⎟𝐅̅̅̅̅−1 + ∑ 𝐝̇ 𝑖+3 ⊗ 𝐕𝑖 𝐅
̅̅̅̅−1 . (4.11.129)
𝑖=1 ⎝ 𝑗=1 ⎠ 𝑖=1
4
𝐭𝜎𝑖 = 𝐽 𝑉 ̅̅̅̅−𝑇 ⎛
̅ 𝛔𝐅 ⎜𝐅̅̅̅̅𝑇 − ∑ 𝐕𝑗 ⊗ 𝐝𝑗+3 ⎞
⎟ 𝐝𝑖 , 𝐭𝜎𝑖+3 = 𝑉𝛔𝐅
̅̅̅̅−𝑇 𝐕𝑖 , 𝑖 = 1, 2, 3. (4.11.132)
⎝ 𝑗=1 ⎠
The first expression above can in turn be rewritten as
4
𝐭𝜎𝑖 = ⎛ ̅ 𝛔 − ∑ 𝐭𝜎𝑗+3 ⊗ 𝐝𝑗+3 ⎞
⎜𝐽 𝑉 ⎟ 𝐝𝑖 , 𝑖 = 1, 2, 3. (4.11.133)
⎝ 𝑗=1 ⎠
3 6 6
𝐋̃ = ∑ 𝐝̇ 𝑖 ⊗ 𝐝𝑖 ⎛
⎜𝐅̅̅̅̅ − ∑ 𝐝𝑗+3 ⊗ 𝐕𝑗 ⎞
⎟𝐅̅̅̅̅−1 + ∑ 𝐝̇ 𝑖+3 ⊗ 𝐕𝑖 𝐅
̅̅̅̅−1
𝑖=1 ⎝ 𝑗=1 ⎠ 𝑖=1
1 𝐽 6 3 ⎡ ∂𝜂 𝑘 𝑇 ̇
3
+ ∑ ∑ ⎢ 𝑘 ((𝐝 ) 𝐝𝑗+3 − ∑ 𝐝𝑖 ⋅ 𝐝𝑗+3 𝐝𝑘 ⋅ 𝐝̇ 𝑖 )⎤
⎥𝐈
3 𝐽 ̃ 𝑗=1 𝑘=1 ⎣∂𝑏𝑗 𝑖=1 ⎦ (4.11.141)
1𝐽 ⎡ ⎛ 6 3 𝑖 6
+ ⎢𝜂 ⎜ ∑ ∑ 𝐝 ⋅ 𝐝 𝐝̇
𝑗+3 𝑖 ⋅ ̅
𝐅̅̅̅ −𝑇 𝑗
𝐕 − ∑ ̅̅̅̅−𝑇 𝐕𝑖 ⎞
𝐝̇ 𝑖+3 ⋅ 𝐅 ⎟⎤⎥ 𝐈,
3 𝐽 ̃ ⎣ ⎝ 𝑗=1 𝑖=1 𝑖=1 ⎠⎦
putting back these expressions into the virtual work expression above and adding the
hourglass internal forces from below we get the same as in [2].
4.11.6 Hourglass
The hourglass resistance is based on a strain energy potential given as
4 3 4 3
𝑉𝜇 𝑗
𝜓= ∑ ∑ ∑ ∑ 𝑏𝑖 𝐵𝑖𝑘 𝑙
𝑗𝑙 𝑏𝑘 , (4.11.142)
12(1 − 𝜈) 𝑖=1 𝑗=1 𝑘=1 𝑙=1
∂𝑏𝑘𝑗 T
= −𝐝𝑖 ⋅ 𝐝𝑗+3 (𝐝𝑘 ) , 𝑖 = 1, 2, 3
∂𝐝𝑖
(4.11.145)
∂𝑏𝑘𝑗 𝑗 T
= 𝛿𝑖 (𝐝𝑘 ) , 𝑖 = 1,2,3,4
∂𝐝𝑖+3
Figure 4.11. Stress profile for tip loading in two directions for various mesh
densities and distortions
0.4
Error
0.3
0.2
0.1
0
0.2 0.4 0.6 0.8 1
Mesh size
4.11.11 References
[1]M. Jabareen and M.B. Rubin, A Generalized Cosserat Point Element (CPE) for Isotropic
Nonlinear Elastic Materials including Irregular 3-D Brick and Thin Structures,
J. Mech. Mat. Struct. 3-8, pp. 1465-1498, 2008.
[2]M. Jabareen, E. Hanukah and M.B. Rubin, A Ten Node Tetrahedral Cosserat Point
Element (CPE) for Nonlinear Isotropic Elastic Materials. Comput Mech 52, pp
257-285, 2013.
y
S
n
x i
Figure 4.11. The contour S encloses an area A.
Here, 𝐧̂ is the normal vector to 𝑆 and 𝐱̂ and 𝐲̂ are unit vectors in the x and y directions,
respectively. See Figure 4.11.
In this approach the velocity gradients which define the strain rates are element
centered, and the velocities and nodal forces are node centered. See Figure 4.12. Noting
that the normal vector 𝐧̂ is defined as:
∂𝑦 ∂𝑥
𝐧̂ = 𝐱̂ + 𝐲̂ , (4.152)
∂𝑆 ∂𝑆
and referring to Figure 4.13, we can expand the numerator in equation (4.64):
∂𝑦
∫ 𝐹(n̂ ⋅ x̂ )𝑑𝑆 = ∫ 𝐹 𝑑𝑆
𝑐
∂𝑆 (4.153)
= 𝐹23 (𝑦3 − 𝑦2 ) + 𝐹34 (𝑦4 − 𝑦3 ) + 𝐹41 (𝑦1 − 𝑦4 ) + 𝐹12 (𝑦2 − 𝑦1 ),
where 𝐹𝑖𝑗 = (𝐹𝑖 + 𝐹𝑗 )/2.
Therefore, letting 𝐴 again be the enclosed area, the following expressions are
obtained:
∂𝐹 𝐹23 (𝑦3 − 𝑦2 ) + 𝐹34 (𝑦4 − 𝑦3 ) + 𝐹41 (𝑦1 − 𝑦4 ) + 𝐹12 (𝑦2 − 𝑦1 )
=
∂𝑥 𝐴
(4.154)
(𝐹 − 𝐹4 )(𝑦3 − 𝑦1 ) + (𝑦2 − 𝑦4 )(𝐹3 − 𝐹1 )
= 2 .
2𝐴
Hence, the strain rates in the x and y directions become:
∂𝐹 𝐹23 (𝑦3 − 𝑦2 ) + 𝐹34 (𝑦4 − 𝑦3 ) + 𝐹41 (𝑦1 − 𝑦4 ) + 𝐹12 (𝑦2 − 𝑦1 ) (4.155)
=
∂𝑥 𝐴
LS-DYNA DEV 10/27/16 (r:8004) 4-41 (Solid Elements)
Solid Elements LS-DYNA Theory Manual
y 4 3
1 2
The zero energy modes, called hourglass modes, as in the three dimensional solid
elements, can be a significant problem. Consider the velocity field given by: 𝑥̇3 = 𝑥̇1 , 𝑥̇2 =
𝑥̇4 , 𝑦̇3 = 𝑦̇1 , and 𝑦̇2 = 𝑦̇4 . As can be observed from Equations (4.97) and (4.98), 𝜀𝑥𝑥 = 𝜀𝑦𝑦 =
𝜀𝑥𝑦 = 0 and the element "hourglasses" irrespective of the element geometry. In the two-
dimensional case, two modes exist versus twelve in three dimensions. The hourglass
treatment for these modes is identical to the approach used for the shell elements, which
are discussed later.
In two-dimensional planar geometries for plane stress and plane strain, the finite
element method and the integral finite difference method are identical. The velocity strains
are computed for the finite element method from the equation:
𝛆̇ = 𝐁𝐯, (4.159)
III
4 3
IV II
1 2
I
Figure 4.14. The finite difference stencil for computing nodal forces is shown.
where 𝛆̇ is the velocity strain vector, B is the strain displacement matrix, and 𝐯 is the nodal
velocity vector. Equation (4.100a) exactly computes the same velocity strains as the integral
difference method if
𝐁 = 𝐁(𝑠, 𝑡)|𝑠=𝑡=0 . (4.160)
The update of the nodal forces also turns out to be identical. The momentum
equations in two-dimensional planar problems are given by
1 ∂𝜎𝑥𝑥 ∂𝜎𝑥𝑦
( + ) = 𝑥̈,
𝜌 ∂𝑥 ∂𝑦
(4.161)
1 ∂𝜎𝑥𝑦 ∂𝜎𝑦𝑦
( + ) = 𝑦̈.
𝜌 ∂𝑥 ∂𝑦
Referring to Figure 4.14, the integral difference method gives Equation (4.113):
1 ∂𝜎𝑥𝑥 𝜎𝑥𝑥1 (𝑦𝐼 − 𝑦𝐼𝑉 ) + 𝜎𝑥𝑥2 (𝑦𝐼𝐼 − 𝑦𝐼 ) + 𝜎𝑥𝑥3 (𝑦𝐼𝐼𝐼 − 𝑦𝐼𝐼 ) + 𝜎𝑥𝑥4 (𝑦𝐼𝑉 − 𝑦𝐼𝐼𝐼 )
= . (4.162)
𝜌 ∂𝑥 1 (𝜌 𝐴 + 𝜌 𝐴 + 𝜌 𝐴 + 𝜌 𝐴 )
1 1 2 2 3 3 4 4
2
An element wise assembly of the discretized finite difference equations is possible
leading to a finite element like finite difference program. This approach is used in the
DYNA2D program by Hallquist [1980].
where again note that 𝑦 is the axis of symmetry and 𝑥 is the radial direction. The only
difference between finite element approach and the finite difference method is in the
treatment of the terms, which arise from the assumption of axisymmetry. In the finite
difference method the radial acceleration is found from the calculation:
1 𝜎𝑥𝑥 (𝑦𝐼 − 𝑦𝐼𝑉 ) + 𝜎𝑥𝑥2 (𝑦𝐼𝐼 − 𝑦𝐼 ) + 𝜎𝑥𝑥3 (𝑦𝐼𝐼𝐼 − 𝑦𝐼𝐼 ) + 𝜎𝑥𝑥4 (𝑦𝐼𝑉 − 𝑦𝐼𝐼𝐼 )
𝑥̈ = [ 1 −
2 (𝜌1 𝐴1 + 𝜌2 𝐴2 + 𝜌3 𝐴3 + 𝜌4 𝐴4 )
(4.164)
𝜎𝑥𝑦1 (𝑥𝐼 − 𝑥𝐼𝑉 ) + 𝜎𝑥𝑦2 (𝑥𝐼𝐼 − 𝑥𝐼 ) + 𝜎𝑥𝑦3 (𝑥𝐼𝐼𝐼 − 𝑥𝐼𝐼 ) + 𝜎𝑥𝑦4 (𝑥𝐼𝑉 − 𝑥𝐼𝐼𝐼 )
] + 𝛽,
(𝜌1 𝐴1 + 𝜌2 𝐴2 + 𝜌3 𝐴3 + 𝜌4 𝐴4 )
where 𝛽𝑓𝑒 is found by a summation over the four surrounding elements:
1 4 𝜎𝑥𝑥 − 𝜎𝜃𝜃𝑖
𝛽 = ∑[ 𝑖 ]. (4.165)
4 𝑖=1 (𝜌𝑥)𝑖
𝑥𝑖 is the centroid of the ith element defined as the ratio of its volume 𝑉𝑖 and area 𝐴𝑖 :
𝑉𝑖
𝑥𝑖 = , (4.166)
𝐴𝑖
𝜎𝜃𝜃𝑖 is the hoop stress, and 𝜌𝑖 is the current density.
1 𝜎𝑥𝑥 (𝑦𝐼 − 𝑦𝐼𝑉 ) + 𝜎𝑥𝑥2 (𝑦𝐼𝐼 − 𝑦𝐼 ) + 𝜎𝑥𝑥3 (𝑦𝐼𝐼𝐼 − 𝑦𝐼𝐼 ) + 𝜎𝑥𝑥4 (𝑦𝐼𝑉 − 𝑦𝐼𝐼𝐼 )
𝑥̈ = [ 1 −
2 (𝜌1 𝐴1 + 𝜌2 𝐴2 + 𝜌3 𝐴3 + 𝜌4 𝐴4 )
(4.168)
𝜎𝑥𝑦1 (𝑥𝐼 − 𝑥𝐼𝑉 ) + 𝜎𝑥𝑦2 (𝑥𝐼𝐼 − 𝑥𝐼 ) + 𝜎𝑥𝑦3 (𝑥𝐼𝐼𝐼 − 𝑥𝐼𝐼 ) + 𝜎𝑥𝑦4 (𝑥𝐼𝑉 − 𝑥𝐼𝐼𝐼 )
] + 𝛽𝑓𝑒 ,
(𝜌1 𝐴1 + 𝜌2 𝐴2 + 𝜌3 𝐴3 + 𝜌4 𝐴4 )
where 𝛽𝑓𝑒 is now area weighted.
4 (𝜎𝑥𝑥𝑖 − 𝜎𝜃𝜃𝑖 )𝐴𝑖
1
𝛽𝑓𝑒 = ∑⎡ ⎢ ⎤
⎥. (4.169)
4(𝜌1 𝐴1 + 𝜌2 𝐴2 + 𝜌3 𝐴3 + 𝜌4 𝐴4 ) 𝑖=1 ⎣ 𝑥𝑖 ⎦
Rezoning capability was added to DYNA2D in 1980 and to LS-DYNA in version 940.
In the current implementation the rezoning can be done interactively and used to relocate
the nodal locations within and on the boundary of parts. This method is sometimes
referred to as r-adaptive.
3. Initialize remeshed regions by interpolating from nodal point values of old mesh.
2 1 8
Ajusted
node
3 7
4 5 6
In the first step each variable is approximated globally by a summation over the
number of nodal points 𝑛:
𝑛
𝑔(𝑟, 𝑧) = ∑ 𝑔𝑖 Φ𝑖 (𝑟, 𝑧), (4.170)
𝑖=1
where
Φ𝑖 = set of piecewise continuous global basis functions
𝑔𝑖 = nodal point values
Given a variable to be remapped h(𝑟, 𝑧), a least squares best fit is found by
minimizing the functional
i.e.,
𝑑Π
= 0, 𝑖 = 1, 2, … , 𝑛. (4.172)
𝑑𝑔𝑖
This yields the set of matrix equations
𝐌𝐠 = 𝐟, (4.173)
where
𝐌 = ∑ 𝐌𝑒 = ∑ ∫ 𝚽𝚽T 𝑑𝐴 ,
𝑒
(4.174)
𝑒
𝐟 = ∑ 𝐟 = ∑ ∫ ℎ𝚽𝑑𝐴.
𝑒
𝑓𝑖
𝑔𝑖 = .
𝑀𝑖
In step 2, the interactive rezoning phase permits:
• Plotting of solution at current time
• Deletion of elements and slidelines
• Boundary modifications via dekinks, respacing nodes, etc.
• Mesh smoothing
A large number of interactive commands are available and are described in the Help
package. Current results can be displayed by
• Color fringes
• Contour lines
• Vectors plots
• Principal stress lines
• Deformed meshes and material outlines
• Profile plots
• Reaction forces
• Interface pressures along 2D contact interfaces
old mesh
new mesh
Figure 4.16. A four point Gauss quadrature rule over the new element is used to
determine the new element centered value.
In applying the relaxation, the new nodal positions are found and given by Equation
(1.176)
8
∑𝑖=1 𝜉𝑖 𝑥𝑖
𝑥= 8
,
∑𝑖=1 𝜉𝑖
(4.176)
8
∑𝑖=1 𝜉𝑖 𝑦𝑖
𝑦= 8
,
∑𝑖=1 𝜉𝑖
where the nodal positions relative to the node being moved are shown in the sketch in
Figure 4.15.
g4 g3
a
s
g1 g2
Figure 4.17. A four point Gauss quadrature rule over the new element is used to
determine the new element centered value.
The new element centered values, ℎ∗ , computed in Equation (1.179) are found by a 4
point Gauss Quadrature as illustrated in Figure 4.16.
∫ 𝑔𝑑𝐴
ℎ∗ = . (4.179)
∫ 𝑑𝐴
The Gauss point values are interpolated from the nodal values according to Equation
(1.180). This is also illustrated by Figure 4.17.
𝑔𝑎 = ∑ 𝜙𝑖 (𝑠𝑎 , 𝑡𝑎 )𝑔𝑖 . (4.180)
5
Cohesive elements
The cohesive elements are used for modelling cohesive interfaces between faces of solid
elements (types 19 and 21), faces of shell elements (types 20 and 22) and edges of shell
elements (type 29), typically for treating delamination.
Element 19 𝒒3 Element 20 𝒒3
𝑛4
𝑛1 𝑛4 𝑛1
𝒒1 𝒒1 𝑛3
𝑛3 𝑡
𝑛2 𝑚4 𝑛2 𝑚4
𝑚1
𝑚1
𝑚3 𝑚3
𝑚2 𝒒2 𝑚2 𝒒2
Element 29 𝒒3
𝑡
𝑛4 𝐴
Cohesive layer 𝒒3
𝑛3 𝒙t
𝑛1 𝒒1
𝒒1
𝑛2 𝒒2
𝑚4
𝒒2
𝒙b
𝑚1 𝑚3
𝑚2
As a comparison to other elements, the cohesive “strain” is the separation distance (length)
between the two surfaces, and the cohesive stress (force per area) is its conjugate with
respect to the energy surface density (energy per area). Cohesive elements 21 and 22 are
9.50
LS-DYNA Theory Manual Cohesive elements
pentahedral versions of elements 19 and 20, respectively, where the top and bottom surface
are triangles. These elements are not treated here per se, but the only difference is that the
iso-parametric interpolation functions change.
5.1 Kinematics
For this presentation we refer to Figure 5.18 for an illustration. Let
𝒅 = 𝑸𝑻 (𝒙𝑡 − 𝒙𝑏 ) − 𝒅0 , (5.181)
be the separation of the cohesive layer in the local system, where
(5.182)
𝑸 = [𝒒1 𝒒2 𝒒3 ],
is the local coordinate system and 𝒙𝑡 and 𝒙𝑏 are global coordinates on the top and bottom
surfaces for a given iso-parametric coordinate (𝜉 , 𝜂). The distance vector 𝒅0 represents the
initial gap for cases where the cohesive interface has a nonzero thickness, so 𝒅 = 𝟎 initially.
For cohesive element 19 the separation is given directly from the solid element geometries
(sum over i)
𝒙𝑡 = 𝒙𝑖𝑛 𝑁 𝑖 (𝜉 , 𝜂),
(5.183)
𝒙𝑏 = 𝒙𝑖𝑚 𝑁 𝑖 (𝜉 , 𝜂),
where
1
𝑁 𝑖 (𝜉 , 𝜂) = (1 + 𝜉 𝑖 𝜉 )(1 + 𝜂𝑖 𝜂), (5.184)
4
and
𝜉 ∗ = [−1, 1, 1, −1],
𝜂∗ = [−1, −1, 1, 1], (5.185)
are the in-plane shape functions. From here and onwards, superscripts n and m denote the
top and bottom surfaces, respectively, and thus 𝒙𝑖𝑛 and 𝒙𝑖𝑚 are the nodal coordinates
associated with the two surfaces.
For cohesive elements 20 and 29, the separation 𝒅 is updated using an incremental
formulation and we have
𝒅 ̇ = 𝑸𝑇 (𝒙̇𝑡 − 𝒙̇𝑏 ) + 𝑸̇ 𝑇 𝑸𝒅. (5.186)
Furthermore, for cohesive element 20 we have (sum over i)
𝑡
𝒙̇𝑡 = {𝒙̇𝑖𝑛 − 𝝎𝑖𝑛 × 𝒏𝑡 } 𝑁 𝑖 (𝜉 , 𝜂),
2
(5.187)
𝑡
𝒙̇𝑏 = {𝒙̇𝑖𝑚 + 𝝎𝑖𝑚 × 𝒏𝑏 } 𝑁 𝑖 (𝜉 , 𝜂),
2
where the thicknesses of the two shells adjacent to the cohesive layer are denoted t,
currently assumed to be the same for the both. In these equations, 𝒏𝑡 and 𝒏𝑏 are the top and
bottom shell normal, initially equal to 𝒒3 but they may evolve independently with time.
For cohesive element 29 we instead have
𝑡 1−𝜂 𝑡 1+𝜂
𝒙̇𝑡 = {𝒙̇1𝑛 + 𝜉 𝝎1𝑛 × 𝒏𝑡 } + {𝒙̇2𝑛 + 𝜉 𝝎2𝑛 × 𝒏𝑡 } ,
2 2 2 2
(5.188)
𝑡 1−𝜂 𝑡 1+𝜂
𝐱̇𝑏 = {𝐱̇4𝑚 + 𝜉 𝝎4𝑚 × 𝐧𝑏 } + {𝐱̇3𝑚 + 𝜉 𝝎3𝑚 × 𝐧𝑏 } .
2 2 2 2
In these expressions 𝝎𝑖𝑚 and 𝝎𝑖𝑛 denote nodal rotational velocities, and also note that for
evaluation the velocities of 𝒙𝑡 and 𝒙𝑏 we assume that the fiber pointing from assumed mid
layer coincides with that of the coordinate axes. This is in analogy to how the Belytschko-
Tsay element is treating the fiber vectors and presumably enhances robustness of the
elements. Note also that the shell normal are in this case initially equal to 𝒒1 .
For the local coordinate system, cohesive elements 19 and 20 evaluate this according to the
invariant node numbering approach for shells using the mid layer node coordinates
𝒙𝑖𝑛 + 𝒙𝑖𝑚
𝒙̅𝑖 = , 𝑖 = 1,2,3,4, (5.189)
2
as follows. First let
𝒙̅3 − 𝒙̅1
𝒆1 = ,
∣𝒙̅3 − 𝒙̅1 ∣
(5.190)
𝒙̅ − 𝒙̅2
𝒆2 = 4 ,
|𝒙̅4 − 𝒙̅2 |
and then
𝒆1 + 𝒆2
𝒒1 = − ,
|𝒆1 + 𝒆2 |
(5.191)
𝒆 − 𝒆2
𝒒2 = 1 ,
|𝒆1 − 𝒆2 |
followed by
𝒒3 = 𝒒1 × 𝒒2 . (5.192)
Cohesive element 29 starts by computing
𝒙2𝑛 + 𝒙3𝑚 − 𝒙1𝑛 − 𝒙4𝑚
𝒒2 = 𝑛 , (5.193)
∣𝒙2 + 𝒙3𝑚 − 𝒙1𝑛 − 𝒙4𝑚 ∣
followed by
𝒒 = 𝒙4𝑛 + 𝒙3𝑛 − 𝒙1𝑚 − 𝒙2𝑚 , (5.194)
𝒒 − 𝒒2 𝒒𝑇 𝒒2
𝒒3 = , (5.195)
|𝒒 − 𝒒2 𝒒𝑇 𝒒2 |
9.52
LS-DYNA Theory Manual Cohesive elements
and
𝒒1 = 𝒒2 × 𝒒3 . (5.196)
For detailed information on individual cohesive constitutive laws we refer to the materials
section.
𝜎
𝜎𝑒
𝑑
𝑑𝑒 𝑑𝑐
∫ 𝒅𝑇̇ 𝝈𝑑𝐴 = {𝒙̇𝑖𝑚 }𝑇 𝒇𝑖𝑚 + {𝒙̇𝑖𝑛 }𝑇 𝒇𝑖𝑛 + {𝝎𝑖𝑚 }𝑇 𝒓𝑖𝑚 + {𝝎𝑖𝑛 }𝑇 𝒓𝑖𝑛 , (5.197)
𝐴
𝑗 𝑗
where 𝒇𝑖 and 𝒓𝑖 is the nodal force and moment for node i on element j, respectively. The
area A represents the cohesive mid layer spanned by the iso-parametric representation and
this is used to identify the nodal forces and moments. In the following we work out the
details for cohesive element 29. Cohesive elements 19 and 20 are treated analogous.
1−𝜂 1−𝜂 𝑡
+{𝐱̇1𝑛 }𝑇 ∫ ( ) 𝐐𝛔𝑑𝐴 + {𝝎1𝑛 }𝑇 ∫ (− 𝜉 ) 𝐑𝑡𝑇 𝐐𝛔𝑑𝐴
2 2 2
𝐴 𝐴
1−𝜂 1−𝜂 𝑡
+{𝒙̇4𝑚 }𝑇 ∫ (− ) 𝑸𝝈𝑑𝐴 + {𝝎4𝑚 }𝑇 ∫ ( 𝜉 ) 𝑹𝑏𝑇 𝑸𝝈𝑑𝐴
2 2 2
𝐴 𝐴
1+𝜂 1+𝜂 𝑡
+{𝒙̇3𝑚 }𝑇 ∫ (− ) 𝑸𝝈𝑑𝐴 + {𝝎3𝑚 }𝑇 ∫ ( 𝜉 ) 𝑹𝑏𝑇 𝑸𝝈𝑑𝐴.
2 2 2
𝐴 𝐴
If the first term on the right hand side of (5.201) is neglected we can, using (5.197) and
(5.201), identify the nonzero nodal forces and moments
1−𝜂 𝑡 1−𝜂
𝒇1𝑛 = ∫ ( ) 𝑸𝝈𝑑𝐴, 𝒓1𝑛 = − ∫ ( 𝜉 ) 𝑹𝑡𝑇 𝑸𝝈𝑑𝐴,
2 2 2
𝐴 𝐴
1+𝜂 𝑡 1+𝜂
𝒇2𝑛 = ∫ ( ) 𝑸𝝈𝑑𝐴, 𝒓2𝑛 = − ∫ ( 𝜉 ) 𝑹𝑡𝑇 𝑸𝝈𝑑𝐴,
2 2 2
𝐴 𝐴
(5.202)
𝑡 1+𝜂
𝒇3𝑚 = −𝒇2𝑛 , 𝒓3𝑚 = ∫ ( 𝜉 ) 𝑹𝑏𝑇 𝑸𝝈𝑑𝐴,
2 2
𝐴
𝑡 1−𝜂
𝒇4𝑚 = −𝒇1𝑛 , 𝒓4𝑚 = ∫ ( 𝜉 ) 𝑹𝑏𝑇 𝑸𝝈𝑑𝐴.
2 2
𝐴
9.54
LS-DYNA Theory Manual Cohesive elements
In the implementation these integrals are evaluated using 4-point Gaussian quadrature,
where the integration point locations are given by
1 1 1 1
𝜉∗ = ⎡
⎢− , , ,− ⎤⎥,
⎣ √3 √3 √3 √3⎦
(5.203)
1 11 1
𝜂∗ = ⎡
⎢− ,− , , ⎤⎥.
⎣ √3 √3 √3 √3⎦
where 𝜙 is an arbitrary function of the iso-parametric coordinates, 𝐴𝑖 in the right hand side
stands for the area of the cohesive layer and 𝛔𝑖 is the cohesive interface stress, both
evaluated at and with respect to integration point i.
where
𝑑21 = ∣𝒙2𝑛 − 𝒙1𝑛 ∣,
𝑑34 = ∣𝒙3𝑚 − 𝒙4𝑚 ∣, (5.206)
This is compared to the corresponding kinetic energy for an equivalent solid type 19
cohesive layer, using that the shell type 29 translational nodal mass 𝑚𝑡 is twice that of the
solid type 19 nodal masses,
2
1 𝑚𝑡 𝑚𝑡 𝑚𝑡 𝑚𝑡 𝑚𝑡 𝑚𝑡 𝑚𝑡 𝑚𝑡 𝑡 1
𝑊= ( + + + + + + + ) ( 𝜔) = 𝑚𝑡 𝑡2 𝜔2 , (5.211)
2 2 2 2 2 2 2 2 2 2 2
which results in
1
𝑚𝑟 = 𝑚𝑡 𝑡 2 . (5.212)
4
9.56
LS-DYNA Theory Manual Cohesive elements
For a given integration point and referring to Equation (5.202) and (5.209), we collect these
nodal force triplets into complete nodal vectors 𝒇 and 𝒈, associated with the constitutive
and drilling part, respectively. That is,
𝒇1𝑛
⎡ 𝑛⎤
⎢ 𝒓1 ⎥
⎢𝒇 𝑛 ⎥
⎢ 2⎥
⎢ 𝒓𝑛 ⎥
𝒇 =⎢ 𝑚 2⎥
, (5.215)
⎢𝒇3 ⎥
⎢ 𝑚⎥
⎢ 𝒓3 ⎥
⎢ 𝑚⎥
⎢𝒇4 ⎥
⎣𝒓4𝑚 ⎦
and the same expression holds for 𝒈. Likewise we let 𝒗 be the collection of nodal velocities,
i.e.,
𝒙̇𝑛
⎡ 1𝑛 ⎤
⎢ 𝝎1 ⎥
⎢ 𝒙̇𝑛 ⎥
⎢ 2𝑛 ⎥
⎢𝝎 ⎥
𝒗 = ⎢ 𝑚2 ⎥, (5.216)
⎢ 𝒙̇3 ⎥
⎢𝝎3𝑚 ⎥
⎢ 𝑚⎥
⎢ 𝒙̇4 ⎥
⎣𝝎4𝑚 ⎦
and note that we can identify generalized strain-displacement matrices 𝑩𝑓 (3 by 24 matrix)
and 𝑩𝑔 (4 by 24 matrix) from (5.186) and (5.188) as well as (5.205) so that
𝒅 ̇ = 𝑩𝑓 𝒗,
(5.217)
𝜹 ̇ = 𝑩𝑔 𝒗,
where we have collected the drilling kinematic velocities in a vector
𝛿𝑛̇
⎡ 1𝑛 ⎤
⎢ 𝛿2̇ ⎥
𝜹̇ = ⎢
⎢𝛿𝑚
⎥
⎥ (5.218)
⎢ 3̇ ⎥
⎣𝛿4𝑚̇ ⎦
With this notation we can rewrite (5.202) and (5.209) into a more compact form
𝒇 = 𝑩𝑇𝑓 𝝈𝐴,
(5.219)
𝒈 = 𝑩𝑇𝑔 𝝇𝐴,
with 𝐴 here being the area of the integration point of interest. The stiffness matrix is then
simply the differentiation of these force vectors with respect to the nodal coordinates, and
by using (5.208), (5.213) and the chain rule of differentiation we get
𝜕𝝈
𝑲𝑓 = 𝑩𝑇𝑓 𝑩 𝐴,
𝜕𝒅 𝑓 (5.220)
𝑲𝑔 = 𝐸𝑩𝑇𝑔 𝑩𝑔 𝐴,
where 𝜕𝝈/𝜕𝒅 is the constitutive tangent from the cohesive material used.
9.58
LS-DYNA Theory Manual Belytschko Beam
6
Belytschko Beam
The Belytschko beam element formulation [Belytschko et al. 1977] is part of a family
of structural finite elements, by Belytschko and other researchers that employ a ‘co-
rotational technique’ in the element formulation for treating large rotation. This section
discusses the co-rotational formulation, since the formulation is most easily described for a
beam element, and then describes the beam theory used to formulate the co-rotational
beam element.
The choice of the reference configuration determines the type of deformations that
will be computed: total deformations result from comparing the current configuration with
the initial configuration, while incremental deformations result from comparing with the
previous configuration. In most time stepping (numerical) Lagrangian formulations,
incremental deformations are used because they result in significant simplifications of
other algorithms, chiefly constitutive models.
^
Ȳ Y
b2
e02
e01
^
b1 X
I J
X̄
The co-rotational formulation uses two types of coordinate systems: one system
associated with each element, i.e., element coordinates which deform with the element, and
another associated with each node, i.e., body coordinates embedded in the nodes. (The
term ‘body’ is used to avoid possible confusion from referring to these coordinates as
‘nodal’ coordinates. Also, in the more general formulation presented in [Belytschko et al.,
1977], the nodes could optionally be attached to rigid bodies. Thus the term ‘body
coordinates’ refers to a system of coordinates in a rigid body, of which a node is a special
case.) These two coordinate systems are shown in the upper portion of Figure 6.1.
The element coordinate system is defined to have the local x-axis 𝐱̂ originating at
node 𝐼 and terminating at node 𝐽; the local y-axis 𝐲̂ and, in three dimension, the local z-axis
𝐳̂, are constructed normal to 𝐱̂. The element coordinate system (𝐱̂, 𝐲̂ , 𝐳̂) and associated unit
vector triad (𝐞1 , 𝐞2 , 𝐞3 ) are updated at every time step by the same technique used to
construct the initial system; thus the unit vector e1 deforms with the element since it always
points from node 𝐼 to node 𝐽.
The embedded body coordinate system is initially oriented along the principal
inertial axes; either the assembled nodal mass or associated rigid body inertial tensor is
5-2 (Belytschko Beam) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Belytschko Beam
^
X
J
Ȳ
^
Y b2
e1
e2 X̄
I b1
Figure 6.2. Co-rotational coordinate system: (a) initial configuration, (b) rigid
rotational configuration and (c) deformed configuration.
used in determining the inertial principal values and directions. Although the initial
orientation of the body axes is arbitrary, the selection of a principal inertia coordinate
system simplifies the rotational equations of motion, i.e., no inertial cross product terms are
present in the rotational equations of motion. Because the body coordinates are fixed in the
node, or rigid body, they rotate and translate with the node and are updated by integrating
the rotational equations of motion, as will be described subsequently.
The unit vectors of the two coordinate systems define rotational transformations
between the global coordinate system and each respective coordinate system. These
transformations operate on vectors with global components 𝐀 = (𝐴𝑥 , 𝐴𝑦 , 𝐴𝑧 ), body
coordinates components 𝐀 ̅̅̅̅̅̅ = (𝐴𝑥̅ , 𝐴𝑦̅ , 𝐴𝑧̅ ), and element coordinate components A
̂=
(𝐴𝑥̂ , 𝐴𝑦̂ , 𝐴𝑧̂ ) which are defined as:
⎧𝐴 𝑏 𝑏2𝑥 𝑏3𝑥 ⎧𝐴 𝑥⎫
̅
{ 𝑥⎫} ⎡ 1𝑥 ⎤{ ̅ }
𝐀 = ⎨𝐴𝑦 ⎬ = ⎢
⎢𝑏1𝑦 𝑏2𝑦 𝑏3𝑦 ⎥ ̅̅̅̅̅̅},
⎥ ⎨𝐴𝑦 ⎬ = [𝛌]{𝐀 (6.1)
{ }
⎩𝐴𝑧 ⎭ ⎣𝑏1𝑧 { }
𝑏2𝑧 𝑏3𝑧 ⎦ ⎩𝐴𝑧̅ ⎭
where 𝑏𝑖𝑥 , 𝑏𝑖𝑦 , 𝑏𝑖𝑧 are the global components of the body coordinate unit vectors. Similarly
for the element coordinate system:
𝑒3𝑥 ⎧ ̂ ⎫
⎧𝐴
{ 𝑥⎫
𝑒
} ⎡ 1𝑥
𝑒2𝑥 {𝐴𝑥 }
{ }
𝐀 = ⎨ 𝑦 ⎬ = ⎢𝑒1𝑦
𝐴 𝑒2𝑦 𝑒3𝑦 ⎤
⎥ ⎨𝐴𝑦̂ ⎬ = [𝛍]{𝐀
̂ }, (6.2)
{
⎩𝐴𝑧 }
⎭ ⎣𝑒1𝑧 𝑒2𝑧 {
𝑒3𝑧 ⎦ { ̂ } }
⎩𝐴𝑧 ⎭
where 𝑒𝑖𝑥 , 𝑒𝑖𝑦 , 𝑒𝑖𝑧 are the global components of the element coordinate unit vectors. The
inverse transformations are defined by the matrix transpose, i.e.,
̅̅̅̅̅̅} = [𝛌]T {𝐀}
{𝐀
(6.3)
̂ } = [𝛍]T {𝐀},
{𝐀
since these are proper rotational transformations.
^ Ȳ
Y b2
e2 ^
e1 X
J
I e01
b1
X̄
Figure 6.3. Co-rotational coordinate system: (a) initial configuration, (b) rigid
rotational configuration and (c) deformed configuration.
The following two examples illustrate how the element and body coordinate system
are used to separate the deformations and rigid body displacements from the displace-
ments:
Rigid Rotation. First, consider a rigid body rotation of the beam element about node 𝐼, as
shown in the center of Figure 6.2, i.e., consider node 𝐼 to be a pinned connection. Because
the beam does not deform during the rigid rotation, the orientation of the unit vector 𝐞1 in
the initial and rotated configuration will be the same with respect to the body coordinates.
If the body coordinate components of the initial element unit vector 𝐞01 were stored, they
would be identical to the body coordinate components of the current element unit vector
e1 .
Thus the co-rotational formulation separates the deformation and rigid body
deformations by using:
• a coordinate system that deforms with the element, i.e., the element coordinates;
• or a coordinate system that rigidly rotates with the nodes, i.e., the body coordinates;
Then it compares the current orientation of the element coordinate system with the initial
element coordinate system, using the rigidly rotated body coordinate system, to determine
the deformations.
The superscript ^ emphasizes that these quantities are defined in the local element
coordinate system, and 𝐼 and 𝐽 are the nodes at the ends of the beam.
The beam deformations, defined above in Equation (6.5), are the usual small
displacement beam deformations (see, for example, [Przemieniecki 1986]). Indeed, one
advantage of the co-rotational formulation is the ease with which existing small
displacement element formulations can be adapted to a large displacement formulation
having small deformations in the element system. Small deformation theories can be easily
accommodated because the definition of the local element coordinate system is
independent of rigid body rotations and hence deformation displacement can be defined
directly.
orientation 𝐞1 indicates the magnitude of deformation rotations. Forming the vector cross
product between 𝐞01 and 𝐞1 :
𝐞1 × 𝐞01 = 𝜃𝑦̂ 𝐞2 + 𝜃𝑧̂ 𝐞3 , (6.8)
where
𝜃𝑦̂ = is the incremental deformation about the local 𝑦̂ axis
̅ 0 and 𝑒2𝐽
Note that the body components of 𝑒2𝐼 ̅ 0 are transformed into the current element
coordinate system before performing the indicated vector products.
1. functional forms relating the overall response of the beam, e.g., moment-curvature
relations,
Currently only the former method, as explained subsequently, is implemented; the direct
integration method is detailed in [Belytschko et al., 1977].
Axial Force. The internal axial force is calculated from the elongation of the beam 𝛿 as
given by Equation (6.6), and an axial stiffness:
𝑓𝑥𝐽̂ = 𝐾 𝑎 𝛿, (6.13)
where
𝐴𝐸
𝐾𝑎 = = is the axial stiffness
𝑙0
𝐴 = cross sectional area of the beam
𝐸 = Young's Modulus
𝑙0 = original length of the beam
Bending Moments. The bending moments are related to the deformation rotations by
𝑚
̂ 𝑦𝐼 𝐾𝑦𝑏 4 + 𝜙𝑦 2 − 𝜙𝑦 ⎧ ̂ ⎫
{𝜃𝑦𝐼 }
{ }= [ ] ⎨ ⎬, (6.14a)
𝑚
̂ 𝑦𝐽 1 + 𝜙𝑦 2 − 𝜙𝑦 4 + 𝜙𝑦 { ̂
⎩𝜃𝑦𝐽 }
⎭
𝑚
̂ 𝑧𝐼 𝐾𝑧𝑏 4 + 𝜙𝑧 2 − 𝜙𝑧 𝜃𝑧𝐼̂
{ }= [ ] { }, (6.14b)
𝑚
̂ 𝑧𝐽 1 + 𝜙𝑧 2 − 𝜙𝑧 4 + 𝜙𝑧 𝜃𝑧𝐽 ̂
where Equation (6.14a) is for bending in the 𝐱̂ − 𝐳̂ plane and Equation (6.14b) is for bending
in the 𝐱̂ − 𝐲̂ plane. The bending constants are given by
𝐸𝐼𝑦𝑦
𝐾𝑦𝑏 = (6.15a)
𝑙0
𝐸𝐼𝑧𝑧
𝐾𝑧𝑏 = (6.15b)
𝑙0
12𝐸𝐼𝑦𝑦
𝜙𝑦 = (6.15e)
𝐺𝐴𝑠 𝑙2
12𝐸𝐼𝑧𝑧
𝜙𝑧 = . (6.15f)
𝐺𝐴𝑠 𝑙2
Hence 𝜙 is the shear factor, 𝐺 the shear modulus, and 𝐴𝑠 is the effective area in shear.
Torsional Moment. The torsional moment is calculated from the torsional deformation
rotation as
𝑚 ̂ ,
̂ 𝑥𝐽 = 𝐾 𝑡 𝜃𝑥𝐽𝐼 (6.16)
where
𝐺𝐽
𝐾𝑡 = , (6.17)
𝑙0
and,
𝐽 = ∫ ∫ 𝑦̂𝑧𝑑̂ 𝑦̂𝑑𝑧̂. (6.18)
The above forces are conjugate to the deformation displacements given previously in
Equation (6.5), i.e.,
𝐝̂ T = {𝛿𝐼𝐽 , 𝜃𝑥𝐽𝐼
̂ , 𝜃𝑦𝐼
̂ , 𝜃𝑦𝐽
̂ , 𝜃𝑧𝐼
̂ , 𝜃𝑧𝐽
̂ }, (6.19)
where
𝐝̂ T 𝐟 ̂ = 𝑊 int . (6.20)
And with
𝐟 ̂T = {𝑓𝑥𝐽̂ , 𝑚
̂ 𝑥𝐽 , 𝑚
̂ 𝑦𝐼 , 𝑚
̂ 𝑦𝐽 , 𝑚
̂ 𝑧𝐼 , 𝑚
̂ 𝑧𝐽 }. (6.21)
The remaining internal force components are found from equilibrium:
𝑓𝑥𝐼̂ = −𝑓𝑥𝐽̂ 𝑚
̂ 𝑥𝐼 = −𝑚
̂ 𝑥𝐽
𝑚
̂ 𝑦𝐼 + 𝑚
̂ 𝑦𝐽
𝑓𝑧𝐼̂ = − 𝑓𝑧𝐼̂ = −𝑓𝑧𝐽̂ (6.22)
𝑙0
̂ 𝑧𝐼 + 𝑚
𝑚 ̂ 𝑧𝐽
𝑓𝑦𝐽̂ = − 𝑓𝑦𝐼̂ = −𝑓𝑦𝐽̂
𝑙0
𝑗
d2 𝐛𝑖
= 𝛚 × (𝛚 × 𝐛𝑖 ) + (𝛂𝑖 × 𝐛𝑖 ),
d𝑡2
where 𝜔 and 𝛼 are vectors of angular velocity and acceleration, respectively, obtained from
the rotational equations of motion. With the above relations substituted into Equation
(6.23), the update formula for the unit vectors becomes
𝑗+1 𝑗 Δ𝑡2
𝐛𝑖 = 𝐛𝑖 + Δ𝑡(𝛚 × 𝐛𝑖 ) + {[𝛚 × (𝛚 × 𝐛𝑖 ) + (𝛂𝑖 × 𝐛𝑖 )]}. (6.25)
2
To obtain the formulation for the updated components of the unit vectors, the body
coordinate system is temporarily considered to be fixed and then the dot product of
Equation (6.25) is formed with the unit vector to be updated. For example, to update the 𝑥̅
component of 𝐛3 , the dot product of Equation (6.25), with 𝑖 = 3, is formed with b1 , which
can be simplified to the relation
𝑗+1
Next, if it is assumed that 𝑏̅𝑥1 ≈ 1, orthogonality yields
𝑗+1 𝑗+1 𝑗+1
𝑗+1
𝑏̅𝑥3 + 𝑏̅𝑦1 𝑏̅𝑦3
𝑏̅𝑧1 =− 𝑗+1
. (6.29)
𝑏̅𝑧3
𝑗+1
The component 𝑏̅𝑥1 is then found by enforcing normality:
The updated components of 𝐛1 and 𝐛3 are defined relative to the body coordinates at time
step 𝑗. To complete the update and define the transformation matrix, Equation (6.1), at time
step 𝑗 + 1, the updated unit vectors 𝐛1 and 𝐛3 are transformed to the global coordinate
system, using Equation (6.1) with [𝛌] defined at step 𝑗, and their vector cross product is
used to form 𝐛2 .
LS-DYNA DEV 10/27/16 (r:8004) 5-9 (Belytschko Beam)
LS-DYNA Theory Manual Hughes-Liu Beam
7
Hughes-Liu Beam
The Hughes-Liu beam element formulation, based on the shell [Hughes and Liu
1981a, 1981b] discussed later, was the first beam element we implemented. It has several
desirable qualities:
• It is incrementally objective (rigid body rotations do not generate strains), allowing
for the treatment of finite strains that occur in many practical applications;
• It is simple, which usually translates into computational efficiency and robustness
• It is compatible with the brick elements, because the element is based on a
degenerated brick element formulation;
• It includes finite transverse shear strains. The added computations needed to retain
this strain component, compare to those for the assumption of no transverse shear
strain, are insignificant.
7.1 Geometry
The Hughes-Liu beam element is based on a degeneration of the isoparametric 8-
node solid element, an approach originated by Ahmad et al., [1970]. Recall the solid
element isoparametric mapping of the biunit cube
8
𝐱(𝜉 , 𝜂, 𝜁 ) = ∑ 𝑁𝑎 (𝜉 , 𝜂, 𝜁 )𝑥𝑎 , (7.1)
𝑎=1
with,
(1 + 𝜉𝑎 𝜉 )(1 + 𝜂𝑎 𝜂)(1 + 𝜁𝑎 𝜁 )
𝑁𝑎 (𝜉 , 𝜂, 𝜁 ) = , (7.2)
8
where 𝐱 is an arbitrary point in the element, (𝜉 , 𝜂, 𝜁 ) are the parametric coordinates, 𝐱𝑎 are
the global nodal coordinates of node 𝑎, and 𝑁𝑎 are the element shape functions evaluated at
node 𝑎, i.e., (𝜉𝑎 , 𝜂𝑎 , 𝜁𝑎 ) are (𝜉 , 𝜂, 𝜁 ) evaluated at node 𝑎.
In the beam geometry, 𝜉 determines the location along the axis of the beam and the
coordinate pair (𝜂, 𝜁 ) defines a point on the cross section. To degenerate the 8-node brick
geometry into the 2-node beam geometry, the four nodes at 𝜉 = −1 and at 𝜉 = 1 are
combined into a single node with three translational and three rotational degrees of
freedom. Orthogonal, inextensible nodal fibers are defined at each node for treating the
rotational degrees of freedom. Figure 7.1 shows a schematic of the biunit cube and the
beam element. The mapping of the biunit cube into the beam element is separated into
three parts:
𝐱(𝜉 , 𝜂, 𝜁 ) = 𝐱̅(𝜉 ) + 𝐗(𝜉 , 𝜂, 𝜁 ),
(7.3)
= 𝐱̅(𝜉 ) + 𝐗𝜁 (𝜉 , 𝜁 ) + 𝐗𝜂 (𝜉 , 𝜂),
where 𝐱̅ denotes a position vector to a point on the reference axis of the beam, and 𝐗𝜁 and
𝐗𝜂 are position vectors at point 𝐱̅ on the axis that define the fiber directions through that
point. In particular,
ζ
Biunit Cube
ζ
X
ξ
η
Beam Element ξ
Nodal fibers
ζ̄ x̄ζ
x^ζ
0 x^
η
xζ-
-1 -
Bottom Surface zζ
Figure 7.1. Hughes-Liu beam element.
2
𝐱̅(𝜉 ) = ∑ 𝑁𝑎 (𝜉 )𝐱̅𝑎 ,
𝑎=1
2
𝐗𝜂 (𝜉 , 𝜂) = ∑ 𝑁𝑎 (𝜉 )𝐗𝜂𝑎 (𝜂) , (7.4)
𝑎=1
2
𝐗𝜁 (𝜉 , 𝜁 ) = ∑ 𝑁𝑎 (𝜉 )𝐗𝜁𝑎 (𝜁 ).
𝑎=1
With this description, arbitrary points on the reference line 𝐱̅ are interpolated by the one-
dimensional shape function 𝑁(𝜉 ) operating on the global position of the two beam nodes
that define the reference axis, i.e., 𝐱̅a . Points off the reference axis are further interpolated
by using a one-dimensional shape function along the fiber directions, i.e., 𝐗𝜂𝑎 (𝜂) and𝐗𝜁𝑎 (𝜁 )
where
̂𝜂𝑎
𝐗𝜂𝑎 (𝜂) = 𝑧𝜂𝑎 (𝜂)𝐗 ̂𝜁𝑎
𝐗𝜁𝑎 (𝜁 ) = 𝑧𝜁𝑎 (𝜁 )𝐗
𝑧𝜂𝑎 (𝜂) = 𝑁+ (𝜂)𝑧+ −
𝜂𝑎 + 𝑁− (𝜂)𝑧𝜂𝑎 𝑧𝜁𝑎 (𝜁 ) = 𝑁+ (𝜁 )𝑧+ −
𝜁𝑎 + 𝑁− (𝜁 )𝑧𝜁𝑎
(1 + 𝜂) (1 + 𝜁 ) (7.5)
𝑁+ (𝜂) = 𝑁+ (𝜁 ) =
2 2
(1 − 𝜂) (1 − 𝜁 )
𝑁− (𝜂) = 𝑁− (𝜁 ) =
2 2
where 𝑧𝜁 (𝜁 ) and 𝑧𝜂 (𝜂) are “thickness functions”.
reference axis is located on the outer surfaces of the beam. If they are set to zero, the
reference axis is at the center.
The same parametric representation used to describe the geometry of the beam
elements is used to interpolate the beam element displacements, i.e., an isoparametric
representation. Again the displacements are separated into the reference axis displace-
ments and rotations associated with the fiber directions:
𝐮(𝜉 , 𝜂, 𝜁 ) = 𝐮
̅̅̅̅(𝜉 ) + 𝐔(𝜉 , 𝜂, 𝜁 ),
(7.7)
=𝐮
̅̅̅̅(𝜉 ) + 𝐔𝜁 (𝜉 , 𝜁 ) + 𝐔𝜂 (𝜉 , 𝜂).
The reference axis is interpolated as usual
2
̅̅̅̅(𝜉 ) = ∑ 𝑁𝑎 (𝜉 )𝐮
𝐮 ̅̅̅̅𝑎 . (7.8)
𝑎=1
Hughes and Liu introduced the notation that follows, and the associated schematic
shown in Figure 7.2, to describe the current deformed configuration with respect to the
reference configuration.
𝐲
̅̅̅̅ = 𝐱̅ + 𝐮
̅̅̅̅, 𝐘 = 𝐗 + 𝐔, ̂𝜂𝑎 = 𝐗
𝐘 ̂𝜂𝑎 + 𝐔
̂𝜂𝑎 ,
̅̅̅̅ + 𝐘,
𝐲=𝐲 (7.11)
𝐲
̅̅̅̅𝑎 = 𝐱̅𝑎 + 𝐮
̅̅̅̅𝑎 , 𝐘𝑎 = 𝐗𝑎 + 𝐔𝑎 , ̂𝜁𝑎 = 𝐗
𝐘 ̂𝜁𝑎 + 𝐔
̂𝜁𝑎 ,
In the above relations, and in Figure 7.2, the 𝐱 quantities refer to the reference configura-
tion, the 𝐲 quantities refer to the updated (deformed) configuration and the 𝐮 quantities are
the displacements. The notation consistently uses a superscript bar (⋅ ̅) to indicate reference
surface quantities, a superscript caret (⋅ ̂) to indicate unit vector quantities, lower case letter
for translational displacements, and upper case letters for fiber displacements. Thus to
update to the deformed configuration, two vector quantities are needed: the reference
surface displacement 𝐮 ̅̅̅̅ and the associated nodal fiber displacement 𝐔. The nodal fiber
displacements are defined in the fiber coordinate system, described in the next subsection.
u U
X (parallel construction)
x
Y
ū
reference axis in
undeformed
x̄ geometry
Deformed Configuration
Reference Surface
Figure 7.2. Schematic of deformed configuration displacements and position
vectors.
Equations (7.12) are used to transform the incremental fiber tip displacements to
rotational increments in the equations of motion. The second-order accurate rotational
update formulation due to Hughes and Winget [1980] is used to update the fiber vectors:
𝑌̂𝜂𝑖𝑛+1 = 𝑅𝑖𝑗 (Δ𝜃)𝑌̂𝜂𝑖𝑛 ,
(7.13)
𝑌̂𝜁𝑖𝑛+1 = 𝑅𝑖𝑗 (Δ𝜃)𝑌̂𝜁𝑖𝑛 ,
then
𝑛+1
̂𝜂𝑎
̂𝜂𝑎 = 𝐘 𝑛
̂𝜂𝑎
Δ𝐔 −𝐘 ,
(7.14)
𝑛+1
̂𝜁𝑎
̂𝜁𝑎 = 𝐘 𝑛
̂𝜁𝑎
Δ𝐔 −𝐘 ,
where
(2𝛿𝑖𝑗 + Δ𝑆𝑖𝑘 )Δ𝑆𝑖𝑘
𝑅𝑖𝑗 (Δ𝜃) = 𝛿𝑖𝑗 + ,
2𝐷
Δ𝑆𝑖𝑗 = 𝑒𝑖𝑘𝑗 Δ𝜃𝑘 , (7.15)
1
2𝐷 = 2 + (Δ𝜃12 + Δ𝜃22 + Δ𝜃32 ).
2
Here 𝛿𝑖𝑗 is the Kronecker delta and 𝑒𝑖𝑘𝑗 is the permutation tensor.
The transformation of vectors from the global to the local coordinate system can
now be defined in terms of the basis vectors as
⎧𝐴𝑥̂ ⎫ 𝑒3𝑥 T ⎧𝐴𝑥 ⎫
{
{ } } ⎡𝑒1𝑥 𝑒2𝑥
{ }
̂ ̂
𝐀 = ⎨𝐴𝑦 ⎬ = ⎢𝑒1𝑦 𝑒2𝑦 𝑒3𝑦 ⎤
⎥ ⎨𝐴𝑦 ⎬ = [𝐪]{𝐀}, (7.19)
{
{𝐴 ̂ }} ⎣𝑒1𝑧 𝑒2𝑧 𝑒3𝑧 ⎦ {
⎩𝐴𝑧 }
⎭
⎩ 𝑧⎭
̂ is a
where 𝑒𝑖𝑥 , 𝑒𝑖𝑦 , 𝑒𝑖𝑧 are the global components of the local coordinate unit vectors, 𝐀
vector in the local coordinates, and 𝐀 is the same vector in the global coordinate system.
𝑛+1
𝑙
𝜎𝑖𝑗𝑛+1 = 𝑞𝑘𝑖 𝜎𝑘𝑛 𝑞𝑛𝑗 , (7.25)
before computing the internal force vector.
where 𝐟𝑎int are the internal forces at node 𝑎 and 𝐁𝑎 is the strain-displacement matrix in the
global coordinate system associated with the displacements at node 𝑎. The 𝐁 matrix relates
six global strain components to eighteen incremental displacements [three translational
displacements per node and the six incremental fiber tip displacements of Equation (7.14)].
It is convenient to partition the 𝐁 matrix:
𝐁 = [𝐁1 , 𝐁2 ]. (7.27)
Each 𝐵𝑎 sub matrix is further partitioned into a portion due to strain and spin with the
following sub matrix definitions:
𝐵 0 0 𝐵4 0 0 𝐵7 0 0
⎡0 1 𝐵2 0 0 𝐵5 0 0 𝐵8 0 ⎤
⎢ ⎥
⎢0 0 𝐵3 0 0 𝐵6 0 0 𝐵9 ⎥
𝐁𝑎 = ⎢
⎢𝐵2
⎥, (7.28)
𝐵1 0 𝐵5 𝐵4 0 𝐵8 𝐵7 0 ⎥
⎢ ⎥
⎢0 𝐵3 𝐵2 0 𝐵6 𝐵5 0 𝐵9 𝐵8 ⎥
⎣𝐵3 0 𝐵1 𝐵6 0 𝐵4 𝐵9 0 𝐵7 ⎦
where,
⎧ 𝜕𝑁𝑎
{ 𝑁 𝑎,𝑖 = 𝑖 = 1,2,3
{ 𝜕𝑦𝑖
{
{
{ 𝜕(𝑁𝑎 𝑧𝜂𝑎 )
𝐵𝑖 = ⎨(𝑁𝑎 𝑧𝜂𝑎 ) = 𝑖 = 4,5,6. (7.29)
{ ,𝑖−3 𝜕𝑦𝑖−3
{
{ 𝜕(𝑁𝑎 𝑧𝜁𝑎 )
{ (𝑁 𝑧 )
{ 𝑎 𝜁𝑎 ,𝑖−6 = 𝑖 = 7,8,9
⎩ 𝜕𝑦𝑖−6
With respect to the strain-displacement relations, note that:
• The derivative of the shape functions are taken with respect to the global coordi-
nates;
• The 𝐁 matrix is computed on the cross-section located at the mid-point of the axis;
• The resulting 𝐁 matrix is a 6 18 matrix.
The internal force, 𝑓 , given by
𝐟′ = 𝐓T 𝐟𝑎int (7.30)
6-8 (Hughes-Liu Beam) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Hughes-Liu Beam
4 3
1 2
4 7 3
8 9 6
1 5 2
4 7 3
8 9 6
1 5 2 4 10 9 3
11 16 15 8
12 13 14 7
1 5 6 2
Figure 7.3. Integration possibilities for rectangular cross sections in the Hughes-
Liu beam element.
is assembled into the global right hand side internal force vector. 𝐓 is defined as (also see
Equation (7.12):
𝐈 𝟎
⎡𝟎 𝐡𝜂 ⎤
𝐓=⎢ ⎥, (7.31)
⎣𝟎 𝐡𝜁 ⎦
where 𝐈 the 3 3 identity matrix.
tt
A
st
Figure 7.4. Specification of the nodal thickness, 𝑠𝑡 and 𝑡𝑡 , for a beam with an
arbitrary cross-section.
For the user defined rule, it is necessary to specify the number of integration points
and the relative area for the total cross section:
𝐴
𝐴𝑟 =
𝐬𝑡 ⋅ 𝐭 𝑡
where 𝑠𝑡 and 𝑡𝑡 are the beam thickness specified on either the cross section or beam element
cards. The rectangular cross-section which contains 𝑠𝑡 and 𝑡𝑡 should completely contain the
cross-sectional geometry. Figure 7.4 illustrates this for a typical cross-section. In Figure
5.5, the area is broken into twelve integration points. For each integration point, it is
necessary to define the 𝑠 and 𝑡 parametric coordinates, (𝑠𝑖 ,𝑡𝑖 ), of the centroid of the ith
integration point and the relative area associated with the point
𝐴𝑖
𝐴𝑟𝑖 =
𝐴
A1 A2 A3 A4 A5
A6
s6 t6
A7 s
A8
Figure 7.5. A breakdown of the cross section geometry in Figure 7.4 into twelve
integration points.
where 𝐴𝑖 is the ratio of the area of the integration point and the actual area of the cross-
section, 𝐴.
8
Warped Beam Elements
With respect to the local system, the Green-Lagrange strain tensor can be written as:
𝜀𝑖𝑗 = 𝑒𝑖𝑗 + 𝜂𝑖𝑗 , (8.1)
where,
𝑒𝑖𝑗 = 0.5(𝑢𝑖,𝑗 + 𝑢𝑗,𝑖 ),
(8.2)
𝜂𝑖𝑗 = 0.5𝑢𝑘,𝑖 𝑢𝑘,𝑗 .
The geometric assumption of infinite in-plane rigidity implies 𝜀22 = 𝜀33 = 𝛾23 = 0.
Then the non-zero strain components which contribute to the strain energy are:
1
𝜀11 = 𝑢1,1 + (𝑢21,1 + 𝑢22,1 + 𝑢23,1 ),
2
2𝜀12 = 𝑢1,2 + 𝑢2,1 + 𝑢1,1, 𝑢1,2 + 𝑢2,1 𝑢2,2 + 𝑢3,1 𝑢3,2 , (8.3)
0
0
𝐱𝑃0 = 𝐱𝐶 + [𝐞 1 𝐞2 𝐞3 ] ⎡ ⎤
⎢𝑥2 ⎥, (8.4)
⎣𝑥3 ⎦
𝜛𝜙
𝐱𝑃 = 𝐱𝐶 + [𝐞′1 𝐞′2 𝐞′3 ] ⎢𝑥2 ⎤
⎡
⎥, (8.5)
⎣𝑥 3 ⎦
respectively, with
1
[𝐞′1 𝐞′2 𝐞′3 ] = [𝐈 + 𝛉 + 𝛉2 ] [𝐞1 𝐞2 𝐞3 ], (8.6)
2
where
0 −𝜃3 𝜃2
⎡𝜃 0 −𝜃1 ⎤
𝛉=⎢ 3 ⎥, (8.7)
⎣−𝜃2 𝜃1 0 ⎦
and 𝜛 is the Saint-Venant warping function about the centroid C. By the transfer theorem,
the following relation holds:
𝜛 = 𝜔 + 𝑐2 𝑥3 − 𝑐3 𝑥2 , (8.8)
where 𝜔 refers to the shear center S, and 𝑐2 and 𝑐3 are the coordinates of S.
Subtracting Equation (8.4) from Equation (8.5) and neglecting third-order terms, the
displacements vector of point P can be computed:
1 1
𝑢1 = 𝑢̅1 − 𝑥2 𝜃3 + 𝑥3 𝜃2 + 𝑥2 𝜃1 𝜃2 + 𝑥3 𝜃1 𝜃3 + 𝜛𝜙,
2 2
1 1
𝑢2 = 𝑢̅2 − 𝑥3 𝜃1 − 𝑥2 (𝜃12 + 𝜃32 ) + 𝑥3 𝜃2 θ3 + 𝜛θ3 𝜙, (8.9)
2 2
1 1
u3 = u ̅̅̅̅3 + x2 θ1 − x3 (θ21 + θ22 ) + x2 θ2 θ3 − 𝜛θ2 𝜙,
2 2
where 𝑢̅1 , 𝑢̅2 , and 𝑢̅3 are the displacements of the centroid C.
1
𝜀0 = 𝑢̅1,1 + (𝑢̅22,1 + 𝑢̅23,1 ),
2
1
𝜅1 = 𝜃1,1 + (𝜃2,1 𝜃3 − 𝜃3,1 𝜃2 ),
2
1
𝜅2 = −𝜃3,1 + (𝜃1 𝜃2,1 + 𝜃1,1 𝜃2 ) + 𝑢̅3,1 𝜃1,1 − 𝑐3 𝜙,1 ,
2
(8.11)
1
𝜅3 = 𝜃2,1 + (𝜃1 𝜃3,1 + 𝜃1,1 𝜃3 ) − 𝑢̅2,1 𝜃1,1 + 𝑐2 𝜙,1 ,
2
1
𝛾12 = 𝑢̅2,1 − 𝜃3 + 𝜃1 𝜃2 + 𝑢̅3,1 𝜃1 − 𝑢̅1,1 𝜃3 ,
2
1
𝛾13 = 𝑢̅3,1 + 𝜃2 + 𝜃1 𝜃3 − 𝑢̅2,1 𝜃1 + 𝑢̅1,1 𝜃2 .
2
Numerical testing has shown that neglecting the nonlinear terms in the curvatures 𝜅1 , 𝜅2 ,𝜅3
and bending shear strains 𝛾12 , 𝛾13 has little effect on the accuracy of the results. Therefore,
Equation (8.11) can be simplified to
𝜅1 = 𝜃1,1 , 𝜅2 = −𝜃3,1 − 𝑐3 𝜙,1 , 𝜅3 = 𝜃2,1 + 𝑐2 𝜙,1 ,
1 (8.12)
𝜀0 = 𝑢̅1,1 + (𝑢̅22,1 + 𝑢̅23,1 ), 𝛾12 = 𝑢̅2,1 − 𝜃3 , 𝛾13 = 𝑢̅3,1 + 𝜃2 .
2
Adopting Bernoulli’s assumption (𝛾12 = 𝛾13 = 0) and Vlasov’s assumption (𝜙 = θ1,1 ),
Equation (8.10) can be rewritten as:
1
𝜀11 = 𝜀0 + 𝑥2 𝜅2 + 𝑥3 𝜅3 + 𝑟2 𝜅12 + 𝜔𝜃1,11 ,
2
2𝜀12 = (𝜛,2 − 𝑥3 )𝜅1 , (8.13)
𝐿 1 1
𝑈 = ∫ ( 𝐸 ∫ 𝜀211 𝑑𝐴 + 𝐺 ∫ [(2𝜀12 )2 + (2𝜀13 )2 ]𝑑𝐴) 𝑑𝑥1 . (8.17)
0 2 𝐴 2 𝐴
∫ 𝑥2 𝑑𝐴 = 0 , ∫ 𝑥3 𝑑𝐴 = 0 , ∫ 𝑥2 𝑥3 𝑑𝐴 = 0. (8.18)
𝐴 𝐴 𝐴
1 𝐼𝑜2 4 (8.20)
+ (𝐼𝑟𝑟 − ) 𝜅1
4 𝐴
with
⎧𝑢̅
{ 2⎫ }
⎨𝑢̅3 ⎬ = 𝐍2 𝐝, (8.25)
{
⎩𝜃1 }⎭
where
𝐝T = [0 0 0 0 𝜃2𝐼 𝜃3𝐼 𝜙𝐼 𝑢̅1𝐽𝐼 0 0 𝜃1𝐽𝐼 𝜃2𝐽 𝜃3𝐽 𝜙𝐽 ]T , (8.26)
𝐍1 = [1 − 𝜉 ⋅⋅⋅⋅⋅⋅| 𝜉⋅ ⋅ ⋅ ⋅ ⋅ ⋅ ],
⋅𝑓 ⋅ ⋅ ⋅ 𝑔 ⋅ ∣ ⋅ 1−𝑓 ⋅ ⋅ ⋅ ℎ ⋅ (8.27)
⎡ ⎤
𝐍2 = ⎢⋅⋅ 𝑓 ⋅ −𝑔 ⋅ ⋅∣∣ ⋅ ⋅ 1−𝑓 ⋅ −ℎ ⋅ ⋅ ⎥,
∣
⎣⋅⋅ ⋅ 𝑓 ⋅ ⋅ 𝑔∣ ⋅ ⋅ ⋅ 1−𝑓 ⋅ ⋅ ℎ⎦
with
𝑓 = 1 − 3𝜉 2 + 2𝜉 3 𝑔 = 𝑙(𝜉 − 2𝜉 2 + 𝜉 3 )ℎ = 𝑙(𝜉 3 − 𝜉 2 ). (8.28)
where
𝐼
𝐃 = diag (1,1, 𝑜 ). (8.32)
𝐴
where
𝐼22 𝐼22 𝑐3
⎡ −𝐼33 𝑐2 ⎤
𝐇=⎢ 𝐼33 ⎥, 𝐼′𝜔 = 𝐼𝜔 + 𝐼22 𝑐32 + 𝐼33 𝑐22 . (8.34)
⎣𝐼22 𝑐3 −𝐼33 𝑐2 𝐼′𝜔 ⎦
1 1
𝑈3 = 𝐸𝑙 [∫ (𝐍3 𝐝)2 𝛖𝐍2,11 𝑑𝜉 ] 𝐝, (8.35)
2 0
where
𝛖 = (−𝐼2𝑟 − 𝐼3𝑟 𝐼′𝜔𝑟 ), 𝐼′𝜔𝑟 = 𝐼𝜔𝑟 − 𝑐3 𝐼2𝑟 + 𝑐2 𝐼3𝑟 . (8.36)
With respect to the local coordinate system, there are totally eight independent
components in the nodal force vector, in correspondence to the eight nodal displacement
components.
6.2.1 Kinematics
We introduce three coordinate systems that are mutually interrelated. The first
coordinate system is the orthogonal Cartesian coordinate system (𝑥, 𝑦, 𝑧), for which the y
and zaxes lie in the plane of the cross-section and the 𝑥-axis parallel to the longitudinal axis
of the beam. The second coordinate system is the local plate coordinate system (𝑥, 𝑠, 𝑛) as
shown in Figure 6.1, wherein the n-axis is normal to the middle surface of a plate element,
the s-axis is tangent to the middle surface and is directed along the contour line of the
cross-section. The (𝑥, 𝑠, 𝑛) and (𝑥, 𝑦, 𝑧) coordinate systems are related through an angle of
orientation 𝜃 as defined in Figure 40.1. The third coordinate set is the contour coordinate s
along the profile of the section with its origin at some point O on the profile section. Point
P is called the pole through which the axis parallel to the x-axis is called the pole axis. To
derive the analytical model for a thin-walled beam, the following two assumptions are
made:
1. The contour of the thin wall does not deform in its own plane.
2. The shear strain γsx of the middle surface is zero.
These equations apply to the whole contour. The out-of-plane displacement u can
now be found from assumption 2. On the middle surface
∂𝐮 ∂𝐯
+ = 𝟎, (8.42)
∂𝑠 ∂𝑥
which can be written
∂𝐮 ∂𝐯
=− = −𝐕′ (𝑥)cos𝜃(𝑠) − 𝐖′ (𝑥)sin𝜃(𝑠) + 𝐫(𝑠)ϕ′𝑥 (𝑥). (8.43)
∂s ∂𝑥
Integrating this relation from point O to an arbitrary point on the contour yields
(using t as a dummy for s)
P
s
q(s)
n
r(s) θ(s)
A
z
y
Figure 8.1. Definition of coordinates in thin-walled open section
s s s s
∂𝐮
∫ 𝑑𝑡 = −𝐕′ (𝑥) ∫ cos𝜃(𝑡)𝑑𝑡 − 𝐖′ (𝑥) ∫ sin𝜃(𝑡)𝑑𝑡 + ϕ′𝑥 (𝑥) ∫ 𝐫(𝑡)𝑑𝑡. (8.44)
∂𝑡
0 0 0 0
Noting that
𝑑𝑦 = cos𝜃(𝑡)𝑑𝑡,
(8.45)
𝑑𝑧 = sin𝜃(𝑡)𝑑𝑡.
we end up with
𝑢(𝑥, 𝑠) = ⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟⏟
𝑢(𝑥, 0) + V′(𝑥)𝑦(0) + W′(𝑥)𝑧(0) + ϕ′𝑥 (𝑥)ϖ
=:𝐔(𝑥)
− 𝑉′(𝑥)
⏟ 𝑦(𝑠) − W′(𝑥)
⏟ z(𝑠) + ϕ′𝑥 (𝑥)(ω(𝑠) − ϖ) (8.46)
=:𝜙𝑧 (𝑥) =:−𝜙𝑦 (𝑥)
= 𝑈(𝑥) − 𝜙𝑧 (𝑥)𝑦 + 𝜙𝑦 (𝑥)𝑧 + 𝜙′𝑥 (𝑥)(𝜔(𝑠) − ϖ).
where 𝑈 denotes the average out-of-plane displacement over the section, 𝜙𝑦 and 𝜙𝑧 denote
the rotation angle about the y and z axis1, respectively, ω is the sectorial area defined as
𝑠
𝜔(𝑠) = ∫ 𝑟(𝑡)𝑑𝑡, (8.47)
0
6.2.2 Kinetics
The kinetic energy of the beam can be written
1
𝑇= ∫ 𝜌{𝑢̇2 + 𝑣̇2 + 𝑤̇ 2 } 𝑑𝑉. (8.50)
2
𝑉
1The substitution of ′ for and ′ for can be seen as a conversion from an Euler-
Bernoulli kinematic assumption to that of Timoschenko.
7-8 (Warped Beam Elements) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Warped Beam Elements
from which the consistent mass matrix can be read out. Here A is the cross sectional area,
Izz and Iyy are the second moments of area with respect to the z and y axes, respectively,
Iωω is the sectorial second moment and Izω and Iyω are the sectorial product moments. An
approximation of this mass matrix can be made by neglecting the off diagonal components.
The diagonal components are
𝜌𝐴𝑙
𝑚TRNS = ,
3
𝜌(𝐼𝑦𝑦 + 𝐼𝑧𝑧 )𝑙
𝑚RT𝑥 = ,
3
𝜌𝐼𝑦𝑦 𝑙
𝑚RT𝑦 = , (8.52)
3
𝜌𝐼 𝑙
𝑚RT𝑧 = 𝑧𝑧 ,
3
𝜌𝐼 𝑙
𝑚TWST = 𝜔𝜔 .
3
With 𝐸 as Young’s modulus and 𝐺 as the shear modulus, the strain energy can be
written
LS-DYNA DEV 10/27/16 (r:8004) 7-9 (Warped Beam Elements)
Warped Beam Elements LS-DYNA Theory Manual
1
𝛱 = (𝐸𝜀2𝑥𝑥 + 𝐺𝛾𝑥𝑦2 2
+ 𝐺𝛾𝑥𝑧 ), (8.53)
2
where the infinitesimal strain components are (neglecting the derivatives of sectorial area)
𝜀𝑥𝑥 = 𝑈 ′ + 𝜙′𝑦 𝑧 − 𝜙′𝑧 𝑦 + 𝜗 ′ 𝜔,
𝛾𝑥𝑦 = 𝑉 ′ − 𝜙′𝑥 𝑧 − 𝜙𝑧 , (8.54)
𝛾𝑥𝑧 = 𝑊 ′ + 𝜙′𝑥 𝑦 + 𝜙𝑦 .
and the variation of the same can be written
𝛿𝜀𝑥𝑥 = 𝛿𝑈 ′ + 𝛿𝜙′𝑦 𝑧 − 𝛿𝜙′𝑧 𝑦 + 𝛿𝜗 ′ 𝜔,
𝛿𝛾𝑥𝑦 = 𝛿𝑉 ′ − 𝛿𝜙′𝑥 𝑧 − 𝛿𝜙𝑧 , (8.55)
𝛿𝛾𝑥𝑧 = 𝛿𝑊 ′ + 𝛿𝜙′𝑥 𝑦 + 𝛿𝜙𝑦 .
where the stiffness matrix can be read. Again the diagonal components are
𝐸𝐴
𝑘TRNS = ,
𝑙
𝐺𝐴
𝑘SHR = ,
𝑙
𝐺(𝐼𝑦𝑦 + 𝐼𝑧𝑧 )
𝑘RTx = ,
𝑙
(8.57)
𝐸𝐼𝑦𝑦
𝐺𝐴𝑙
𝑘RTy = + ,
𝑙 3
𝐸𝐼 𝐺𝐴𝑙
𝑘RTz = 𝑧𝑧 + ,
𝑙 3
𝐸𝐼
𝑘TWST = 𝜔𝜔 .
𝑙
From the expressions of the mass and stiffness matrix, the frequencies of the most
common modes can be estimated. These are
√3 𝐸
1. The tensile and twisting modes with frequency 𝜔 = 𝑙
√𝜌.
√3 𝐺
2. The transverse shear and torsional mode with frequency 𝜔 = 𝑙
√𝜌.
3𝐸 𝐺𝐴 3𝐸 𝐺𝐴
3. The bending modes with frequencies 𝜔 = √𝜌𝑙 2
+ 𝜌𝐼 and 𝜔 = √𝜌𝑙 2
+ 𝜌𝐼 .
𝑦𝑦 𝑧𝑧
Which one of these four that is the highest depends on the geometry of the beam
element. In LS-DYNA the first of these frequencies is used for calculating a stable time
step. We have found no reason for changing approach regarding this element.
3𝐸 𝑃𝐸𝐴
This increases the twist mode frequency to √𝜌𝑙 2
+ 𝜌𝐼 and the torsional mode to
𝜔𝜔
The nodal velocities for a beam element in the local system is written
T
𝐯 = (𝑣𝑥1 𝑣𝑦1 𝑣𝑧1 𝜔𝑥1 𝜔𝑦1 𝜔𝑧1 𝜗̇ 1 𝑣𝑥2 𝑣𝑦2 𝑣𝑧2 𝜔𝑥2 𝜔𝑦2 𝜔𝑧2 𝜗̇ 2 ) , (8.62)
where the superscript refers to the local node number. These are obtained by transforming
the translational velocities and rotational velocities using the local to global transformation
matrix qij . The strain rate – velocity matrix in the local system can be written
𝐁0 =
−𝑙−1 0 0 0 −𝑙−1 −1
0 𝑧 𝑙0 𝑦 −𝑙−1
0 𝜔 𝑙−1 0 0 0 𝑙−1 −1
0 𝑧 −𝑙0 𝑦 𝑙−1
0 𝜔
⎡ 0 0 ⎤
⎢ 1 ⎥ 1
⎢0 −𝑙−1
0 0 𝑙−1
0 𝑧 0 − 0 0 0 𝑙−1
0 0−𝑙−1
0 𝑧 0
⎥ −
⎢ 2 ⎥,(8.63) 2
⎢ 1 ⎥ 1
⎢0 0 −𝑙−1
0 −𝑙−1
0 𝑦 0 0 0 0 𝑙−1
0 0𝑙−1
0 𝑦⎥ 0
⎢ 2 ⎥ 2
⎢ −1 1 −1 1 ⎥
⎣0 0 0 −𝑙0 0 0 −
2
0 0 0 𝑙0 0 0 − ⎦
2
where 𝑙0 is the beam length in the reference configuration, i.e., beginning of the time step.
A corresponding matrix w.r.t. the current configuration is
𝐁=
−𝑙−1 0 0 0 −𝑙−1 𝑧 −𝑙−1 𝑦 −𝑙−1 𝜔 𝑙−1 0 0 0 𝑙−1 𝑧 −𝑙−1 𝑦 𝑙−1 𝜔
⎡ ⎤
⎢ −1 −1 1 −1 1 ⎥
⎢0 −𝑙 0 𝑙 𝑧 0 − 0 0 𝑙 0 −𝑙−1 𝑧 0 − 0 ⎥
⎢ 2 2 ⎥ (8.64)
⎢ 1 1 ⎥,
⎢0 0 −𝑙−1 −𝑙−1 𝑦 0 0 0 0 𝑙−1 𝑙−1 𝑦 0 0 ⎥
⎢ 2 2 ⎥
⎢ 1 1 ⎥
⎣0 0 0 −𝑙−1 0 0 − 0 0 0 𝑙−1 0 0 − ⎦
2 2
where we use the current length of the beam. These matrices are evaluated in each
integration point (𝑥, 𝑦) of the cross section. To compute the strain rate in the local system
we simply apply
𝛆̇ = 𝐁0 𝐯, (8.65)
which is then used to update the local stresses 𝛔. The internal force vector is then
assembled as
𝐟 = 𝐁T 𝛔. (8.66)
Finally the internal force is transformed to the global system using the transfor-
mation matrix.
To compute the stiffness matrix for implicit we neglect the geometric contribution
and just apply
𝐊 = 𝐁T 𝐂𝐁, (8.67)
where 𝐂 is the material tangent modulus. Again the matrix must be transformed to the
global system before used in the implicit solver.
9
Belytschko-Lin-Tsay Shell
^
y s3
4 3
e^3 r31
r42
e^2
s1 ^
x
1
e^1
r21 2
It is desired to establish the local 𝑥 axis 𝑥̂ approximately along the element edge
between nodes 1 and 2. This definition is convenient for interpreting the element stresses,
which are defined in the local 𝑥̂ − 𝑦̂ coordinate system. The procedure for constructing this
unit vector is to define a vector 𝐬1 that is nearly parallel to the vector 𝐫21 , viz.
𝐬1 = 𝐫21 − (𝐫21 ⋅ 𝐞̂3 )𝐞̂3 , (9.4)
𝐬1
𝐞̂1 = . (9.5)
‖𝐬1 ‖
The remaining unit vector is obtained from the vector cross product
𝐞̂2 = 𝐞̂3 × 𝐞̂1 . (9.6)
If the four nodes of the element are coplanar, then the unit vectors 𝐞̂1 and 𝐞̂2 are
tangent to the midplane of the shell and 𝐞̂3 is in the fiber direction. As the element
deforms, an angle may develop between the actual fiber direction and the unit normal 𝐞̂3 .
The magnitude of this angle may be characterized as
∣𝐞̂3 ⋅ 𝐟 − 1∣ < 𝛿, (9.7)
where 𝐟 is the unit vector in the fiber direction and the magnitude of 𝛿 depends on the
magnitude of the strains. According to Belytschko et al., for most engineering applications,
acceptable values of 𝛿 are on the order of 10-2 and if the condition presented in Equation
8-2 (Belytschko-Lin-Tsay Shell) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Belytschko-Lin-Tsay Shell
(9.7) is met, then the difference between the rotation of the co-rotational coordinates 𝑒 ̂ and
the material rotation should be small.
As in the Hughes-Liu shell element, the displacement of any point in the shell is
partitioned into a midsurface displacement (nodal translations) and a displacement
associated with rotations of the element’s fibers (nodal rotations). The Belytschko-Lin-Tsay
shell element uses the Mindlin [1951] theory of plates and shells to partition the velocity of
any point in the shell as:
𝐯 = 𝐯 𝑚 − 𝑧̂ 𝐞3 × 𝛉, (9.10)
where 𝐯 𝑚 is the velocity of the mid-surface, 𝛉 is the angular velocity vector, and 𝑧̂ is the
distance along the fiber direction (thickness) of the shell element. The corresponding co-
rotational components of the velocity strain (rate of deformation) are given by
1 ∂𝜐̂𝑖 ∂𝜐̂𝑗
𝑑𝑖𝑗̂ = ( + ). (9.11)
2 ∂𝑥̂𝑗 ∂𝑥̂𝑖
Substitution of Equation (9.10) into the above yields the following velocity-strain relations:
∂𝑣̂𝑥𝑚 ∂ 𝜃𝑦̂
̂
𝑑𝑥 = + 𝑧̂ , (9.12)
∂𝑥̂ ∂𝑥̂
∂𝜐̂𝑦𝑚 ∂ 𝜃𝑥̂
𝑑𝑦̂ = − 𝑧̂ , (9.13)
∂𝑦̂ ∂𝑦̂
𝑚
∂𝜐̂𝑥𝑚 ∂𝜐̂𝑦 ∂ 𝜃𝑦̂ ∂ 𝜃𝑥̂ ⎞
̂
2𝑑𝑥𝑦 = + + 𝑧̂ ⎛
⎜ − ⎟, (9.14)
∂𝑦̂ ∂𝑥̂ ⎝ ∂𝑦̂ ∂𝑥̂ ⎠
̂ = ∂𝜐̂𝑧𝑚
2𝑑𝑦𝑧 − 𝜃𝑥̂ , (9.15)
∂𝑦̂
̂ = ∂𝜐̂𝑧𝑚
2𝑑𝑥𝑧 + 𝜃𝑦̂ . (9.16)
∂𝑥̂
The velocity-strains at the center of the element, i.e., at 𝜉 = 0, and 𝜂 = 0, are obtained
by substitution of the above relations into the previously defined velocity-strain
displacement relations, Equations (9.12) and (9.16). After some algebra, this yields
𝑑𝑥̂ = 𝐵1𝐼 𝜐̂𝑥𝐼 + 𝑧̂𝐵1𝐼 𝜃𝑦𝐼
̂ , (9.22a)
∂𝑁𝐼
𝐵1𝐼 = , (9.22f)
∂𝑥̂
∂𝑁𝐼
𝐵2𝐼 = . (9.22g)
∂𝑦̂
The shape function derivatives 𝐵𝑎𝐼 are also evaluated at the center of the element, i.e., at
𝜉 = 0, and 𝜂 = 0.
̂𝑅
𝑚 𝛼𝛽 = − ∫ 𝑧̂𝜎̂𝛼𝛽 𝑑𝑧̂, (9.24)
where the superscript, 𝑅, indicates a resultant force or moment, and the Greek subscripts
emphasize the limited range of the indices for plane stress plasticity.
The above element-centered force and moment resultants are related to the local
nodal forces and moments by invoking the principle of virtual power and integrating with
a one-point quadrature. The relations obtained in this manner are
𝑓𝑥𝐼̂ = 𝐴(𝐵1𝐼 𝑓𝑥𝑥
𝑅̂ + 𝐵2𝐼 𝑓𝑥𝑦
𝑅̂ ), (9.25)
𝜅 𝑅̂
̂ 𝑥𝐼 = 𝐴 (𝐵2𝐼 𝑚
𝑚 ̂𝑅 ̂𝑅
𝑦𝑦 + 𝐵1𝐼 𝑚 𝑥𝑦 − 𝑓𝑦𝑧 ), (9.28)
4
𝜅 𝑅̂
𝑚 ̂𝑅
̂ 𝑦𝐼 = −𝐴 (𝐵1𝐼 𝑚 ̂𝑅
𝑥𝑥 + 𝐵2𝐼 𝑚 𝑥𝑦 − 𝑓𝑥𝑧 ), (9.29)
4
𝑚
̂ 𝑧𝐼 = 0, (9.30)
where 𝐴 is the area of the element, and 𝜅 is the shear factor from the Mindlin theory. In the
Belytschko-Lin-Tsay formulation, 𝜅 is used as a penalty parameter to enforce the Kirchhoff
normality condition as the shell becomes thin.
The above local nodal forces and moments are then transformed to the global
coordinate system using the transformation relations given previously as Equation (9.8).
The global nodal forces and moments are then appropriately summed over all the nodes
and the global equations of motion are solved for the next increment in nodal accelerations.
𝑟𝜃 𝐸𝑡3 𝐴
𝑄̇𝐵𝛼 = 𝐵𝛽𝐼 𝐵𝛽𝐼 𝑞𝐵𝛼̇ , (9.36)
192
𝑟𝑤 𝜅𝐺𝑡3 𝐴
𝑄̇𝐵3 = 𝐵𝛽𝐼 𝐵𝛽𝐼 𝑞𝐵3̇ , (9.37)
12
𝑟𝑚 𝐸𝑡𝐴
𝑄̇𝑀
𝛼 = 𝐵𝛽𝐼 𝐵𝛽𝐼 𝑞𝑀
𝛼̇ , (9.38)
8
where 𝑡 is the shell thickness and the parameters, 𝑟𝜃 , 𝑟𝑤 , and 𝑟𝑚 are generally assigned
values between 0.01 and 0.05.
Finally, the hourglass stresses, which are updated from the stress rates in the usual
way, i.e.,
𝐐𝑛+1 = 𝐐𝑛 + Δ𝑡𝐐̇ , (9.39)
and the hourglass resultant forces are then
̂𝐻
𝑚 𝐵
𝛼𝐼 = 𝜏𝐼 𝑄𝛼 , (9.40)
𝐻̂ = 𝜏𝐼 𝑄𝐵,
𝑓3𝐼 (9.41)
3
𝐻̂ = 𝜏𝐼 𝑄𝑀,
𝑓𝛼𝐼 𝛼 (9.42)
where the superscript 𝐻 emphasizes that these are internal force contributions from the
hourglass deformations. These hourglass forces are added directly to the previously
determined local internal forces due to deformations Equations (7.14a - f). These force
vectors are orthogonalized with respect to rigid body motion.
In the hourglass procedure, the in-plane strain field (subscript p) is decomposed into
the one point strain field plus the stabilization strain field:
𝛆̅ṗ = 𝛆̅p0̇ + 𝛆̅p𝑠̇ , (9.43)
𝑝 𝑝
𝑓 (𝜉 , 𝜂) 𝑓4 (𝜉 , 𝜂)
⎡ 1𝑝 𝑝
⎤
𝐖𝑚 = ⎢
⎢ 𝑓 2 (𝜉 , 𝜂) 𝑓 5 (𝜉 , 𝜂)⎥,
⎥ (9.45)
𝑝 𝑝
⎣𝑓3 (𝜉 , 𝜂) 𝑓6 (𝜉 , 𝜂)⎦
𝑝 𝑝
−𝑓 (𝜉 , 𝜂) 𝑓1 (𝜉 , 𝜂)
⎡ 4𝑝 𝑝
⎤
𝐖𝑏 = ⎢
⎢−𝑓 5 (𝜉 , 𝜂) 𝑓 2 (𝜉 , 𝜂)⎥,
⎥ (9.46)
𝑝 𝑝
⎣−𝑓6 (𝜉 , 𝜂) 𝑓3 (𝜉 , 𝜂)⎦
𝑝
where the terms 𝑓𝑖 (𝜉 , 𝜂) 𝑖 = 1, 2, . . . , 6, are rather complicated and the reader is referred to
the reference [Englemann and Whirley, 1991].
To obtain the transverse shear assumed strain field, the procedure given in [Bathe
and Dvorkin, 1984] is used. The transverse shear strain field can again be decomposed into
the one point strain field plus the stabilization field:
𝛆̅ṡ = 𝛆̅s0̇ + 𝛆̅s𝑠̇ , (9.47)
that is related to the hourglass velocities by
𝛆̅s𝑠̇ = 𝐖s 𝐪̇𝑠 , (9.48)
where the transverse shear stabilization strain-velocity operator 𝐖𝑠 is given by
𝑓 𝑠 (𝜉 , 𝜂) −𝑔1𝑠 𝜉 𝑔2𝑠 𝜂 𝑔3𝑠 𝜉 𝑔3𝑠 𝜂
𝐖𝑠 = [ 1𝑠 ]. (9.49)
𝑓2 (𝜉 , 𝜂) 𝑔4𝑠 𝜉 𝑔4𝑠 𝜂 −𝑔2𝑠 𝜉 𝑔1𝑠 𝜂
Again, the coefficients 𝑓1𝑠 (𝜉 , 𝜂) and 𝑔1𝑠 are defined in the reference.
In their formulation, the hourglass forces are related to the hourglass velocity field
through an incremental hourglass constitutive equation derived from an additive
decomposition of the stress into a “one-point stress,” plus a “stabilization stress.” The
integration of the stabilization stress gives a resultant constitutive equation relating
hourglass forces to hourglass velocities. The in-plane and transverse stabilization stresses
are updated according to:
𝛕s𝑠,𝑛+1 = 𝛕s𝑠,𝑛 + Δ𝑡𝑐𝑠 𝐂𝑠 𝛆̅𝑠𝑠̇ , (9.50)
The stabilization stresses can now be used to obtain the hourglass forces:
ℎ
𝐐𝑚 = ∫2ℎ ∫ 𝐖𝑚
T 𝑠
𝛕𝑝 𝑑𝐴 𝑑𝑧, (9.52)
− 𝐴
2
Figure 9.2. The twisted beam problem fails with the Belytschko-Tsay shell
element.
ℎ
𝐐𝑏 = ∫2ℎ ∫ 𝐖𝑏T 𝛕𝑝𝑠 𝑑𝐴 𝑑𝑧, (9.53)
− 𝐴
2
ℎ
𝐐𝑠 = ∫2ℎ ∫ 𝐖𝑠T 𝛕𝑠𝑠 𝑑𝐴 𝑑𝑧. (9.54)
− 𝐴
2
P3
P1
P2
h
1
̅ 𝑐𝑥𝐼 𝑣̂𝑦𝐼 + 𝑏𝑥𝐼 𝑝̇𝑦𝐼 + 𝑏𝑐𝑦𝐼 𝑣̂𝑥𝐼 + 𝑏𝑦𝐼 𝑝̇𝑥𝐼 ).
𝑑𝑥𝑦 = 𝑏𝑥𝐼 𝑣̂𝑦𝐼 + 𝑏𝑦𝐼 𝑣̂𝑥𝐼 + 𝜁 (𝑏 (9.59)
2
The coupling terms are come in through 𝑏𝑐𝑖𝐼 : which is defined in terms of the components of
the fiber vectors as:
𝑏𝑐𝑥𝐼 𝑝𝑦̂2 − 𝑝𝑦̂4 𝑝𝑦̂3 − 𝑝𝑦̂1 𝑝𝑦̂4 − 𝑝𝑦̂2 𝑝𝑦̂1 − 𝑝𝑦̂3
{ 𝑐 } = [𝑝 − 𝑝 𝑝𝑥̂3 − 𝑝𝑥̂1 𝑝𝑥̂4 − 𝑝𝑥̂2 𝑝𝑥̂1 − 𝑝𝑥̂3 ], (9.60)
𝑏𝑦𝐼 𝑥̂2 𝑥̂4
For a flat geometry the normal vectors are identical and no coupling can occur. Two
methods are used by Belytschko for computing 𝑏𝑐𝑖𝐼 and the reader is referred to his papers
for the details. Both methods have been tested in LS-DYNA and comparable results were
obtained.
The transverse shear strain components are given as
𝛾̂ 𝑥𝑧 = −𝑁𝐼 (𝜉 , 𝜂)𝜃𝑦̂̅ 𝐼 , (9.61)
^
y
K
r
e^y
Lk
enk eni
J
K
I e^x
Figure 9.4. Vector and edge definitions for computing the transverse shear strain
components.
1 𝐼 1
𝜃𝑛𝐼̅ = (𝜃𝑛𝐼 𝐼
+ 𝜃𝑛𝐽 ) + 𝐼𝐽 (𝜐̂𝑧𝐽 − 𝜐̂𝑧𝐽 ), (9.65)
2 𝐿
where the subscript n refers to the normal component of side 𝐼 as seen in Figure 7.3 and 𝐿𝐼𝐽
is the length of side 𝐼𝐽.
10
Triangular Shells
The 𝐶0 shell element due to Kennedy, Belytschko, and Lin [1986] has been
implemented as a computationally efficient triangular element complement to the
Belytschko-Lin-Tsay quadrilateral shell element ([Belytschko and Tsay 1981], [Belytschko et
al., 1984a]). For a shell element with five through-the-thickness integration points, the
element requires 649 mathematical operations (the Belytschko-Lin-Tsay quadrilateral shell
element requires 725 mathematical operations) compared to 1417 operations for the
Marchertas-Belytschko triangular shell [Marchertas and Belytschko 1974] (referred to as the
BCIZ [Bazeley, Cheung, Irons, and Zienkiewicz 1965] triangular shell element in the
DYNA3D user’s manual).
The triangular shell element’s origins are based on the work of Belytschko et al.,
[Belytschko, Stolarski, and Carpenter 1984b] where the linear performance of the shell was
demonstrated. Because the triangular shell element formulations parallels closely the
formulation of the Belytschko-Lin-Tsay quadrilateral shell element presented in the
previous section (Section 7), the following discussion is limited to items related specifically
to the triangular shell element.
^
z
^
y
3
e^2
e^3
^
x
1
e^1 2
Figure 10.1. Local element coordinate system for C0 shell element.
Figure 10.1) that deforms with the element is defined in terms of these nodal coordinates.
The procedure for constructing the co-rotational coordinate system is simpler than the
corresponding procedure for the quadrilateral, because the three nodes of the triangular
element are guaranteed coplanar.
The local x-axis 𝑥̂ is directed from node 1 to 2. The element’s normal axis 𝑧̂ is
defined by the vector cross product of a vector along 𝑥̂ with a vector constructed from node
1 to node 3. The local y-axis 𝑦̂ is defined by a unit vector cross product of 𝐞̂3 with 𝐞̂1 ,
which are the unit vectors in the 𝑧̂ directions, respectively. As in the case of the
quadrilateral element, this triad of co-rotational unit vectors defines a transformation
between the global and local element coordinate systems. See Equations (7.5 a, b).
It is convenient to partition the velocity strains and the 𝐁 matrix into membrane and
bending contributions. The membrane relations are given by
⎧𝜐̂𝑥1 ⎫
{
⎧ 𝑑𝑥̂ ⎫
M {
{ 𝜐̂𝑦1 }
}
}
{
{ ̂ } 𝑦̂3 0 𝑦̂3 0 0 0 { { }
} 1 ⎡ 𝑥2 }
𝜐̂
⎨ 𝑑 𝑦 ⎬ = ⎢ 0 𝑥̂3 − 𝑥̂2 0 −𝑥̂3 0 𝑥̂2 ⎤
⎥ ⎨𝜐̂ ⎬, (10.3)
{ 𝑥̂2 𝑦̂3
{2𝑑 ̂ }
} ⎣𝑥̂3 − 𝑥̂2 −𝑦̂3 −𝑥̂3 𝑦̂3 𝑥̂2 0⎦ { {
𝑦2 }
}
⎩ 𝑥𝑦 ⎭ { 𝑥3 }
𝜐̂
{
{𝜐̂ } }
⎩ 𝑦3 ⎭
𝐝̂ M = 𝐁M 𝐯̂ , (10.4)
⎧𝜃𝑥1 ̂ ⎫
{
{ }
{ ̂ }
𝜃𝑦1
{ } }
⎧ 𝜅̂𝑥 ⎫ 0 −𝑦̂3 0 𝑦̂3 0 0 { { ̂ }
{ } −1 ⎡ 𝜃𝑥2 }
⎨
𝜅̂𝑦 =
⎬ 𝑥̂
⎢ 3 𝑥̂2
− 0 𝑥̂3 0 −𝑥̂2 0⎤ ⎥ ⎨ ̂ ⎬, (10.5)
{2𝜅̂ } 𝑥̂2 𝑦̂3 ⎣ 𝑦̂ 𝑥̂3 − 𝑥̂2 −𝑦̂3 −𝑥̂3 0 𝑥̂2 ⎦ { 𝜃𝑦2 }
⎩ 𝑥𝑦 ⎭ 3 {
{ }
{ 𝜃𝑥3 ̂ }}
{
{𝜃 ̂ } }
⎩ 𝑦3 ⎭
or
𝛋̂ M = 𝐁M 𝛉̂ def . (10.6)
The local element velocity strains are then obtained by combining the above two
relations:
M
⎧ 𝑑𝑥̂ ⎫ ⎧ 𝑑𝑥̂ ⎫ ⎧ 𝜅̂𝑥 ⎫
{ } {
{ ̂ } { ̂ } } { }
𝜅̂𝑦 = 𝐝̂ M − 𝑧̂𝛋̂ . (10.7)
⎨ 𝑑𝑦 ⎬ = ⎨ 𝑑𝑦 ⎬ − 𝑧̂ ⎨ ⎬
{
{2𝑑 ̂ }
} {{ ̂ }} {2𝜅̂ }
⎩ 𝑥𝑦 ⎭ ⎩2𝑑𝑥𝑦 ⎭ ⎩ 𝑥𝑦 ⎭
⎧𝜃𝑥1 ̂ ⎫def
{
{ }
{ ̂ }
𝜃𝑦1 }
{
{ }
{𝜃𝑥2 ̂ } }
⎨𝜃 ̂ ⎬ ,
{
{ 𝑦2 }
}
{
{ 𝜃𝑥3 ̂ }}
{
{𝜃 ̂ } }
⎩ 𝑦3 ⎭
𝐝̂ S = 𝐁S 𝛉̂ def
. (10.9)
All of the above velocity-strain relations have been simplified by using one-point
quadrature.
In the above relations, the angular velocities 𝛉̂def are the deformation component of
the angular velocity 𝛉̂ obtained by subtracting the portion of the angular velocity due to
rigid body rotation, i.e.,
𝛉̂def = 𝛉̂ − 𝛉̂rig , (10.10)
The two components of the rigid body angular velocity are given by
rig 𝜐̂𝑧1 − 𝜐̂𝑧2
𝜃𝑦̂ = , (10.11)
𝑥̂2
^
z
3 ^
y
^
x
1 2
^
z
^
y
3
^
x
1 2
Figure 10.2. Element configurations with node 3 aligned with node 1 (left) and
node 3 aligned with node 2 (right).
Although both of these relations yield the average rigid body rotation rate, the selection of
the correct relation depends on the configuration of the element, i.e., on the location of
node 3. Since every element in the mesh could have a configuration that is different in
general from either of the two configurations shown in Figure 10.2, a more robust relation
is needed to determine the average rigid body rotation rate about the local x-axis. In most
typical grids, node 3 will be located somewhere between the two configurations shown in
Figure 10.2. Thus a linear interpolation between these two rigid body rotation rates was
devised using the distance 𝑥̂3 as the interpolant:
rig rig 𝑥̂3 rig 𝑥̂
𝜃𝑥̂ = 𝜃𝑥−left
̂ (1 − ̂
) + 𝜃𝑥−right ( 3 ). (10.15)
𝑥̂2 𝑥̂2
Substitution of Equations (10.13) and (10.14) into (10.15) and simplifying produces the
relations given previously as Equation (10.12).
̂𝑅
𝑚 𝛼𝛽 = − ∫ 𝑧̂ 𝜎̂ 𝛼𝛽 𝑑𝑧̂, (10.17)
where the superscript 𝑅 indicates a resultant force or moment and the Greek subscripts
emphasize the limited range of the indices for plane stress plasticity.
The above element-centered force and moment resultant are related to the local
nodal forces and moments by invoking the principle of virtual power and performing a
one-point quadrature. The relations obtained in this manner are
⎧𝑓𝑥1̂ ⎫
{ }
{𝑓 ̂ }
{ 𝑦1 }
{ } 𝑅
⎧𝑓𝑥𝑥 ̂ ⎫
{
{𝑓𝑥2̂ }
} { }
{
T 𝑓 ̂
𝑅
}
⎨𝑓 ̂ ⎬ = 𝐴𝐁M⎨ 𝑦𝑦 ⎬, (10.18)
{ 𝑦2 } { 𝑅̂ }
{ }
{ } ⎩𝑓𝑥𝑦 ⎭
{𝑓 ̂ }
{ 𝑥3 }
{ }
{𝑓 ̂ }
⎩ 𝑦3 ⎭
⎧𝑚̂ 𝑥1 ⎫
{
{ }
{ ̂ 𝑦1 }
𝑚 }
{ } ⎧𝑚
{ ̂𝑅𝑥𝑥 ⎫
}
{𝑚̂ 𝑥2 } { 𝑅 }
̂𝑅
T 𝑚 ̂ T 𝑓𝑥𝑧
⎨𝑚̂ 𝑦2 ⎬ = 𝐴𝐁M ⎨ 𝑦𝑦 ⎬ + 𝐴𝐁S { ̂𝑅 }, (10.19)
{ } {
{𝑚 𝑅 }
} 𝑓 𝑦𝑧
{
{ } ⎩ ̂ 𝑥𝑦 ⎭
{ ̂ 𝑥3 }
𝑚 }
{𝑚 }
⎩ ̂ 𝑦3 ⎭
where 𝐴 is the area of the element (2𝐴 = 𝑥̂2 𝑦̂3 ).
The remaining nodal forces, the 𝑧̂ component of the force (𝑓𝑧3̂ , 𝑓𝑧2̂ , 𝑓𝑧1̂ ), are
determined by successively solving the following equilibration equations
𝑚
̂ 𝑥1 + 𝑚 ̂ 𝑥3 + 𝑦̂3 𝑓𝑧3̂ = 0,
̂ 𝑥2 + 𝑚 (10.20)
𝑚
̂ 𝑦1 + 𝑚 ̂ 𝑦3 − 𝑥̂3 𝑓𝑧3̂ − 𝑥̂2 𝑓𝑧2̂ = 0,
̂ 𝑦2 + 𝑚 (10.21)
The unit normal to the shell element 𝐞3 is formed from the vector cross product
𝐞3 = 𝐥21 × 𝐥31 , (10.23)
where 𝐥21 and 𝐥31 are unit vectors originating at Node 1 and pointing towards Nodes 2 and
3 respectively, see Figure 10.3(b).
Next a unit vector g, see Figure 10.3(b), is assumed to be in the plane of the
triangular element with its origin at Node 1 and forming an angle 𝛽 with the element side
between Nodes 1 and 2, i.e., the vector 𝑙21 . The direction cosines of this unit vector are
represented by the symbols (𝑔𝑥 , 𝑔𝑦 , 𝑔𝑧 ). Since g is the unit vector, its direction cosines will
satisfy the equation
Also, since 𝐠 and 𝐞3 are orthogonal unit vectors, their vector dot product must
satisfy the equation
𝑒3𝑥 𝑔𝑥 + 𝑒3𝑦 𝑔𝑦 + 𝑒3𝑧 𝑔𝑧 = 0. (10.25)
In addition, the vector dot product of the co-planar unit vectors 𝐠 and 𝐥21 satisfies the
equation
𝐼21𝑥 𝑔𝑥 + 𝐼21𝑦𝑔𝑦 𝑔𝑦 + 𝐼21𝑧 𝑔𝑧 = cos𝛽, (10.26)
where (𝑙21𝑥 , 𝑙21𝑦 , 𝑙21𝑧 )are the direction cosines of 𝐥21 .
3
y^
x^
^
z e2
e3 e1
1 b2
b3 b1
2
(a) Element and body coordinates
3
y^
x^
^ I31
z e2
α
e3
α/2 β
1 b2
I21
b3 b1
2
Solving this system of three simultaneous equation, i.e., Equation (10.24), (10.25),
and (10.26), for the direction cosines of the unit vector g yields
𝑔𝑥 = 𝑙21𝑥 cos𝛽 + (𝑒3𝑦 𝑙21𝑧 − 𝑒3𝑧 𝑙21𝑦 )sin𝛽, (10.27)
⎧ 𝑢̂𝑥 ⎫
def 𝜙𝑚
𝑥
{
{ 𝑢̂ } } ⎡ 𝑚 ⎤ ⎧ 𝛿 ⎫
𝑦 = ⎢ 𝜙𝑦 ⎥ { }
(10.32)
⎨ − − −⎬ ⎢− − −⎥ ⎨− − −⎬,
{ } ⎢ {
⎥ ⎩ }
{
⎩ 𝑢̂𝑧 ⎭
} 𝑓 𝜃̂ ⎭
⎣ 𝜙𝑧 ⎦
𝑓
The matrices 𝜙𝑚 𝑚
𝑥 , 𝜙𝑦 and 𝜙𝑧 are the membrane and flexural interpolation functions,
respectively. The element’s membrane deformation is defined in terms of the edge
elongations. Marchertas and Belytschko adapted this idea from Argyris et al., [1964],
where incremental displacements are used, by modifying the relations for total
displacements,
2(𝑥𝑗𝑖 𝑢𝑗𝑖𝑥 + 𝑦𝑗𝑖 𝑢𝑗𝑖𝑦 + 𝑧𝑗𝑖 𝑢𝑗𝑖𝑧 ) + 𝑢2𝑗𝑖𝑥 + 𝑢2𝑗𝑖𝑦 + 𝑢2𝑗𝑖𝑧
𝛿𝑖𝑗 = , (10.35)
𝑙 0𝑖𝑗 + 𝑙𝑖𝑗
where 𝑥𝑗𝑖 = 𝑥𝑗 − 𝑥𝑖 , etc.
𝑓
[1965]; hence the LS-DYNA reference to the BCIZ element. Explicit expressions for 𝜙𝑧 are
quite tedious and are not given here. The interested reader is referred to Appendix G in
the original work of Marchertas and Belytschko [1974].
The local nodal rotations, which are interpolated by these flexural shape functions,
are defined in a manner similar to those used in the Belytschko beam element. The current
components of the original element normal are obtained from the relation
𝐞03 = 𝛍T 𝛌𝐞̅ 03 , (10.36)
where 𝛍 and 𝛌 are the current transformations between the global coordinate system and
the element (local) and body coordinate system, respectively. The vector 𝐞̅ 03 is the original
element unit normal expressed in the body coordinate system. The vector cross product
between this current-original unit normal and the current unit normal,
𝐞3 × 𝐞03 = 𝜃𝑥̂ 𝐞1 + 𝜃𝑦̂ 𝐞2 , (10.37)
define the local nodal rotations as
𝛉̂𝑥 = − 𝐞̂03𝑦 , (10.38)
∂𝑢𝑥 ∂𝑢𝑦 ∂ 2 𝑢𝑧
2𝑒𝑥𝑦 = + − 2𝑧 , (10.42)
∂𝑦 ∂𝑥 ∂𝑥 ∂𝑦
where it is understood that all quantities refer to the local element coordinate system.
∂𝜙𝑚
𝑥𝑖
⎡ ⎤
⎢ ∂𝑥⎥
⎢ ∂𝜙𝑚⎥
⎢ ⎥
𝑦𝑖
𝐄m =⎢ ⎥, (10.45)
⎢ ∂𝑦 ⎥
⎢∂𝜙𝑚 ∂𝜙𝑚 ⎥
⎢ 𝑥𝑖 + 𝑦𝑖 ⎥
⎣ ∂𝑦 ∂𝑥 ⎦
and
2 𝑓
⎡ ∂ 𝜙𝑧𝑖 ⎤
⎢ ∂𝑥2 ⎥
⎢ 2 𝑓 ⎥
⎢∂ 𝜙 ⎥
𝐄f = ⎢ 𝑧𝑖 ⎥
. (10.46)
⎢ ∂𝑦2 ⎥
⎢ ⎥
⎢ ∂2 𝜙𝑓 ⎥
⎢2 𝑧𝑖 ⎥
⎣ ∂𝑥 ∂𝑦⎦
Again, the interested reader is referred to Appendices F and G in the original work of
Marchertas and Belytschko [1974] for explicit expressions of the above two matrices.
̂ T = {𝑚
𝐦 ̂ 1𝑥 𝑚
̂ 1𝑦 𝑚
̂ 2𝑥 𝑚
̂ 2𝑦 𝑚
̂ 3𝑥 𝑚
̂ 3𝑦 }. (10.51)
The through-the-thickness integration portions of the above local force and moment
integrals are usually performed with a 3- or 5-point trapezoidal integration. A three-point
inplane integration is also used; it is inpart this three-point inplane integration that
increases the operation count for this element over the 𝐶0 shell, which used one-point
inplane integration with hourglass stabilization.
The remaining transverse nodal forces are obtained from element equilibrium
considerations. Moment equilibrium requires
𝑓̂ 1 −𝑥̂3 𝑦̂3 𝑚
̂ 1𝑥 + 𝑚
̂ 2𝑥 + 𝑚
̂ 3𝑥
{ 2𝑧 } = [ ] {𝑚
̂ 1𝑦 + 𝑚 ̂ 3𝑦 },
̂ 2𝑦 + 𝑚 (10.52)
𝑓3𝑧̂ 2𝐴 𝑥̂2 − 𝑦̂2
where 𝐴 is the area of the element. Next transverse force equilibrium provides
𝑓1𝑧̂ = −𝑓2𝑧̂ − 𝑓3𝑧̂ . (10.53)
The corresponding global components of the nodal forces are obtained from the
following transformation
⎧𝑓𝑖𝑥 ⎫ 𝑓 ⎧ 𝑥 + 𝑢𝑖𝑗𝑥 ⎫ ⎧𝑥 +𝑢 𝑒
{ } 𝑖𝑗 { 𝑖𝑗 } 𝑓
𝑦𝑖𝑗 + 𝑢𝑖𝑗𝑦 + 𝑖𝑘 {𝑦𝑖𝑘 + 𝑢𝑖𝑘𝑥 ⎫
} ̂
⎧
{𝑒3𝑥 ⎫
}
⎨𝑓𝑖𝑦 ⎬ = 𝑙 ⎨ ⎬ 𝑙 ⎨ 𝑖𝑘 𝑖𝑘𝑦 ⎬ + 𝑓𝑖𝑧 ⎨ 3𝑦 ⎬.
{
(10.54)
{𝑓 }
⎩ 𝑖𝑧 ⎭ 𝑖𝑗 {
⎩𝑧𝑖𝑗 + 𝑢𝑖𝑗𝑧 }
⎭ 𝑖𝑘
{
⎩𝑧𝑖𝑘 + 𝑢𝑖𝑘𝑧 }
⎭ ⎩𝑒3𝑧 }
⎭
Finally, the local moments are transformed to the body coordinates using the relation
⎧ 𝑚
̅̅̅̅̅ ⎧𝑚̂ 𝑖𝑥 ⎫
{ 𝑖𝑥 ⎫ } T {𝑚 }
⎨ 𝑚
̅̅̅̅̅𝑖𝑦 ⎬ = 𝛌 𝛍 ⎨ ̂ 𝑖𝑦 ⎬. (10.55)
{ ̅̅̅̅̅𝑖𝑧 } {
⎩𝑚 ⎭ ⎩𝑚̂ 𝑖𝑧 }
⎭
11
Fully Integrated Shell (Type 16)
11.1 Introduction
Shell type 16 in LS-DYNA is a fully integrated shell with assumed strain interpolants
used to alleviate locking and enhance in-plane bending behavior, see Engelmann, Whirley,
and Goudreau [1989]; Simo and Hughes [1986]; Pian and Sumihara [1985]. It uses a local
element coordinate system that rotates with the material to account for rigid body motion
and automatically satisfies frame invariance of the constitutive relations. The local element
coordinate system is similar to the one used for the Belytschko-Tsay element, where the the
first two basis vectors are tangent to the shell midsurface at the center of the element, and
the third basis vector is in the normal direction of this surface and initially coincident with
the fiber vectors.
LS-DYNA DEV 10/27/16 (r:8004) 10-1 (Fully Integrated Shell (Type 16))
Fully Integrated Shell (Type 16) LS-DYNA Theory Manual
p ̅̅̅̅ p : 𝛔p (𝐃
̅̅̅̅ )𝑑Ω,
𝛿𝑃int = ∫ 𝛿𝐃 (11.3)
Ω
s ̅̅̅̅ s : 𝛔s (𝐃
̅̅̅̅ )𝑑Ω,
𝛿𝑃int = 𝜅 ∫ 𝛿𝐃 (11.4)
Ω
𝐻 p = ∫ 𝛿[𝛔
̅̅̅̅̅p : (𝐃p (𝐯) − 𝐃
̅̅̅̅ p )]𝑑Ω,
(11.5)
Ω
𝐻 s = 𝜅 ∫ 𝛿[𝛔
̅̅̅̅̅s : (𝐃s (𝐯) − 𝐃
̅̅̅̅ s )]𝑑Ω.
(11.6)
Ω
Here κ is the shear correction factor and the superscripts mean that only the in-plane
components (p) or transverse shear (s) components are treated.
To derive the in-plane assumed strain field, the interpolants for the assumed stress
and strain rates are chosen as
𝐬
̅̅̅̅̅p = [𝐒p 𝐒p ][𝐬m ],
𝛔 (11.8)
b
𝐞m
̅̅̅̅ p = 𝐂−1 [𝐒p
𝐃 𝐒p ][𝐞 ], (11.9)
b
where
1 0 0 𝑎21 𝜂̂ 𝑏21 𝜉 ̂
p
⎡ ⎤
𝐒 =⎢
⎢0 1 0 𝑎22 𝜂̂ 𝑏22 𝜉 ̂ ⎥
⎥, (11.10)
⎣0 0 1 𝑎1 𝑎2 𝜂̂ 𝑏1 𝑏2 𝜉 ⎦̂
and
1
𝜉̂ = 𝜉 − ∫ 𝜉𝑑Ω, (11.11)
|Ω|
Ω
10-2 (Fully Integrated Shell (Type 16)) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Fully Integrated Shell (Type 16)
1
𝜂̂ = 𝜂 − ∫ 𝜂𝑑Ω. (11.12)
|Ω|
Ω
Furthermore, 𝐂 is the plane stress constitutive matrix and 𝜉 and 𝜂 are the isoparametric
coordinates. The coefficients 𝑎𝑖 and 𝑏𝑖 are defined through
1 𝑎 𝑏1
𝐉0 = [ 1 ], (11.13)
4 𝑎2 𝑏2
where 𝐉0 is the area jacobian matrix from the isoparametric to physical domain computed
at the element center.
Inserting the expressions for the strain rate and assumed stress and strain rate into
the expression for 𝐻 p and requiring 𝐻 p = 0 for arbitrary 𝐬m , 𝐬b , 𝐞m and 𝐞b , yields the
following expression for the assumed strain rate in terms of the nodal velocities
p
̅̅̅̅ p = [̅𝐁
𝐃 ̅̅̅̅m ̅̅̅̅̅b ] [𝐯p ],
𝑧𝐁 (11.14)
𝛉̇
where
̅̅̅̅m = 𝐂−1 𝐒p 𝐄̂ 𝐁
̅𝐁 ̂m, (11.15)
̅̅̅̅b = 𝐂−1 𝐒p 𝐄̂ 𝐁
̅𝐁 ̂b, (11.16)
and
T
𝐄̂ = ∫ 𝐒p 𝐂−1 𝐒p 𝑑Ω, (11.17)
𝛺
̂ m = ∫ 𝐒p T 𝐁m 𝑑Ω,
𝐁 (11.18)
𝛺
̂ b = ∫ 𝐒p T 𝐁b 𝑑Ω.
𝐁 (11.19)
𝛺
𝐯
𝐃 ̅̅̅̅𝑡 [ p𝑧 ],
̅̅̅̅ s = ̅𝐁 (11.21)
𝛉̇
where
̅̅̅̅̅𝑡 = 𝐉−T 𝐄 ∫ 𝐒𝑠 T 𝐉T 𝐁𝑡 𝑑Ω,
𝐁 (11.22)
Ω
Here 𝐉 is the area jacobian matrix from the isoparametric domain to the physical domain,
1 1−𝜉 1+𝜉 0 0
𝐄= [ ], (11.23)
2 0 0 1−𝜂 1+𝜂
δ(η)δ(1 + ξ) δ(η)δ(1 − ξ) 0 0
𝐒s = [ ], (11.24)
0 0 δ(ξ)δ(1 + η) δ(ξ)δ(1 − η)
and 𝛿 is the Dirac delta function. Defining the assumed stress as
̅̅̅̅̅p = 𝐉𝐒s 𝐬,
𝛔 (11.25)
yields 𝐻 s = 0 regardless of the choice of 𝐬 and thus a B-bar expression for the assumed
transverse strain rates is obtained as given above. The result is equivalent to defining the
isoparametric assumed shear strain rates by interpolating the corresponding strain rates
from the mid-side points A, B, C and D shown in Figure 11.1.
η
ξ B
10-4 (Fully Integrated Shell (Type 16)) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Fully Integrated Shell (Type 16)
̅̅̅̅ r = ̅𝐁
𝐃 ̅̅̅̅r 𝛉̇, (11.26)
where
̅̅̅̅̅r = 𝑤𝐁
𝐁 ̅̅̅̅̅m 𝐑. (11.27)
LS-DYNA DEV 10/27/16 (r:8004) 10-5 (Fully Integrated Shell (Type 16))
Shells with Thickness Stretch LS-DYNA Theory Manual
12
Shells with Thickness Stretch
12.1 Introduction
Thickness stretch is of considerable importance in problems involving finite thickness
strains, contact and surface loads in nonlinear shell applications. As an example, in sheet
metal forming applications, the presence of normal stresses in thickness direction improves
the accuracy of the solution and also its response on the double sided contact zone between
dies and sheet. It has also been shown that a kinematical representation of a continuous
thickness field improves the instability characteristics when compared to experimental
results, see Figure Figure 1210-2 and Björklund [2014].
There have been several attempts to account for the through-thickness deformation in the
literature, this implementation is inspired by the 7P-CYSE shell introduced in Cardoso and
Yoon [2005]. However, the formulation in this paper is rather complicated and involves for
instance assumed shear strains (ANS) to alleviate shear locking and complicated setup of
internal forces and stiffness matrices by combined analytical and numerical integration to
restore rank deficiencies. A direct implementation of this shell following the theory was
ruled out for efficiency reasons, and another approach was taken in order to presumably
get a useful shell.
The Belytschko-Tsay shell element is one of the fastest elements for thin shell simulations.
This, together with its robustness, is the reason why it is popular in finite element codes.
The implementation of shell type 25, the reduced integrated shell with thickness stretch, is
based on the formulation of the Belytschko-Tsay shell with a relaxation of the thickness
variable. This ensures that it will be efficient and hopefully also possess properties useful
for applications where through thickness deformation is important. As a fully integrated
alternative, shell type 26 is available as an analogue extension of shell type 16 (fully
integrated Belytschko-Tsay), and a triangular shell with thickness stretch is available as
type 27 mainly to allow for hybrid meshes (quadrilaterals combined with triangles) in this
context.
The theory that follows is very similar to that of the Belytschko-Tsay (type 2), Fully
Integrated Shell (type 16) and C0-shell (type 4), and here we emphasize on the parts
involving the amendments to those shell formulations.
9.6
LS-DYNA Theory Manual Shells with Thickness Stretch
25
20
Force, F [kN]
15
stress and discontinuous
thickness field
10
5
Experiment
Shell 26, transverse shear stress Plane Stress
With Normal Stress
and continuous thickness field 0
0 0.5 1 1.5 2 2.5 3
Displacement, δ [mm]
9.7
Shells with Thickness Stretch LS-DYNA Theory Manual
𝜕𝑁𝐼
In the local system one can assume a vanishing third component of both 𝝎𝐼 × 𝒏 and 𝜕𝒙
,
and the thickness strain rate is given by
𝜕𝑠 ̇ 𝑡 ̇ − 4𝜍𝑞 ̇ (12.32)
𝜀̇𝑡 = =
𝜕𝑥3 𝑡 − 4𝜍𝑞
where we used the notation 𝑠 = 𝑠𝐼 𝑁𝐼 , 𝑡 = 𝑡𝐼 𝑁𝐼 and 𝑞 = 𝑞𝐼 𝑁𝐼 , sum over 𝐼. For small strains
𝑞 = 0, and this shows that the thickness strain rate is at most linear.
For evaluating internal forces we define the strain-displacement tensor through (assuming
Voigt notation and sum over 𝐼)
𝜕𝒗 (12.33)
= 𝑩𝐼 𝒖𝐼 ,
𝜕𝒙
and the nodal vector 𝒖𝐼 is given by
(12.34)
𝒖𝐼 = (𝒗𝑇𝐼 𝝎𝐼𝑇 𝑡𝐼̇ 𝑞𝐼̇ )𝑇 ,
indicating the 8 degrees of freedom per node in these elements. The principle of virtual
work results in an internal force vector
(12.35)
𝒇𝐼 = ∫ 𝑩𝑇𝐼 𝝈 ,
where 𝝈 is the Cauchy stress and the integral is over the current element configuration.
𝑞𝐼 = ℎ𝐼
where
(12.37)
𝐼
ℎ𝐼 = (−1) /4
and all other displacement components are zero. To restrain these modes we have
included generalized strains and stresses according to the following (sum over 𝐼)
ℎ𝐼 𝑡𝐼̇ (12.38)
𝜀̇ℎ𝑡 =
𝑡
4ℎ 𝑞 ̇
𝜀̇ℎ𝑞 = − 𝐼 𝐼
𝑡
The corresponding generalized stresses are obtained through
(12.39)
𝜎̇ 𝑡 = 𝐸𝐻 𝜀̇ℎ𝑡
9.8
LS-DYNA Theory Manual Shells with Thickness Stretch
𝐸𝐻 ℎ
𝜎̇ 𝑞 = 𝜀̇
3 𝑞
and the forces are then given by
(12.40)
𝑓𝐼𝑡 = 𝐴ℎ𝐼 𝜎𝑡
𝑞
𝑓𝐼 = −4𝐴ℎ𝐼 𝜎𝑞
where 𝐴 is the area of the element and the value of 𝐸𝐻 is taken as 𝐸𝐻 = 0.05𝐸, i.e., 5% of the
Young’s modulus.
still maintaining a non-singular element we use single point integration of the thickness
strain component and an ANS approach for the transverse shear strain components
emanating from the rates of thickness components. This approach is inspired by the
methodology used for the Belytschko-Tsay shell in which the nodal fiber vectors are reset
in the beginning of each time step, and in an analogous way we reset the thickness
components in the beginning of each time step in this shell element formulation.
9.10
LS-DYNA Theory Manual Shells with Thickness Stretch
successful. That means that the same code can be executed regardless of the choice of
IDOF for this element.
−𝒇
9.11
Shells with Thickness Stretch LS-DYNA Theory Manual
𝑇
𝒇𝐼𝑐 = (𝟎 𝒇 𝑇 𝒏∙𝒇 (12.46)
𝑇
𝑡 (𝒏 × ) 0)
4 4
We can see that the total force include nodal moments caused by the frictional forces acting
at an offset from the midsurface of the shell, but also a pressure acting on the thickness
degree of freedom. In conclusion, the relaxation of the thickness in the Belytschko-Tsay
shell allows for double sided contact zones, i.e., the shell is affected by contact pressure
from both sides even though they are of equal magnitude. This is not possible in
traditional shell with zero normal stress, unless a modified option is used.
𝑛+1 (12.50)
𝜎𝑧𝑧 = 𝜎𝑐𝑛+1
where
𝛔𝑛+1 = stress in step n+1
𝛔𝑛 = stress in step n
𝐾= bulk modulus
∆𝛆 = strain increment
∆𝜀𝑣𝑜𝑙 = ∆𝛆: 𝐈 = volumetric strain increment
𝐈 = unit tensor
∆𝐬= deviatoric stress increment
9.12
LS-DYNA Theory Manual Shells with Thickness Stretch
𝑛
𝐬𝑛 = 𝛔𝑛 − 𝛔 3:𝐈 𝐈 = deviatoric stress in step n
∆𝛆𝑑𝑒𝑣 = ∆𝛆 − ∆𝛆:
3
𝐈
𝐈 = deviatoric strain increment
𝑛+1 𝑛+1 𝑛+1 𝑛+1 𝑛+1
The independent variables in (12.49) and (12.50) are 𝜎𝑥𝑥 , 𝜎𝑦𝑦 , 𝜎𝑥𝑦 , 𝜎𝑦𝑧 , 𝜎𝑥𝑧 and ∆𝜀𝑧𝑧 .
Here we assume that the stress response can be decoupled into a volumetric and deviatoric
part and the deviatoric stress increment depends only on the deviatoric part of the stress
and strain increment as indicated in the formula. We can rewrite (12.49) and (12.50) as
̃ 𝑛+1 = 𝛔
𝛔 ̃ 𝑛 + 𝐾∆𝜀̃𝑣𝑜𝑙 𝐈 + ∆𝐬(𝐬̃𝑛 , ∆𝛆̃𝑑𝑒𝑣 ) (12.51)
𝑛+1 (12.52)
𝜎̃ 𝑧𝑧 =0
by substituting
̃ 𝑛 = 𝛔𝑛 − 𝜎𝑐𝑛 𝐈
𝛔 (12.53)
(12.54)
̃ 𝑛+1 = 𝛔𝑛+1 − 𝜎𝑐𝑛+1 𝐈
𝛔
and
𝜎𝑐𝑛+1 − 𝜎𝑐𝑛 (12.55)
∆𝛆̃ = ∆𝛆 − 𝐈.
3𝐾
Since the deviatoric stress and strain increment is not changed, i.e.,
𝐬𝑛 = 𝐬̃𝑛 (12.56)
(12.57)
∆𝛆𝑑𝑒𝑣 = ∆𝛆̃𝑑𝑒𝑣
with this substitution it follows that the existing material routines can be used for solving
𝑛+1 𝑛+1 𝑛+1 𝑛+1 𝑛+1
(12.51) and (12.52) in terms of 𝜎̃ 𝑥𝑥 , 𝜎̃ 𝑦𝑦 , 𝜎̃ 𝑥𝑦 , 𝜎̃ 𝑦𝑧 , 𝜎̃ 𝑥𝑧 and ∆𝜀̃𝑧𝑧 and then use the inverse
of (12.54) and (12.55) to establish the stress and through thickness strain increment. Thus,
the algorithm is as follows
1.Given 𝛔𝑛 , ∆𝛆, 𝜎𝑐𝑛 and 𝜎𝑐𝑛+1
2.Use (12.53) and (12.55) to compute 𝛔 ̃ 𝑛 and ∆𝛆̃.
3.Do a constitutive update, (12.51) and (12.52), to get 𝛔 ̃ 𝑛+1 and ∆𝜀̃𝑧𝑧 .
4.Use (12.54) and (12.55) to compute 𝛔𝑛+1 and ∆𝜀𝑧𝑧 .
9.13
Shells with Thickness Stretch LS-DYNA Theory Manual
9.14
LS-DYNA Theory Manual Hughes-Liu Shell
13
Hughes-Liu Shell
The Hughes-Liu shell element formulation ([Hughes and Liu 1981a, b], [Hughes et
al., 1981], [Hallquist et al., 1985]) was the first shell element implemented in LS-DYNA. It
was selected from among a substantial body of shell element literature because the element
formulation has several desirable qualities:
• it is incrementally objective (rigid body rotations do not generate strains), allowing
for the treatment of finite strains that occur in many practical applications;
• it is simple, which usually translates into computational efficiency and robustness;
• it is compatible with brick elements, because the element is based on a degenerated
brick element formulation. This compatibility allows many of the efficient and
effective techniques developed for the DYNA3D brick elements to be used with this
shell element;
• it includes finite transverse shear strains;
• a through-the-thickness thinning option (see [Hughes and Carnoy 1981]) is also
available when needed in some shell element applications.
The remainder of this section reviews the Hughes-Liu shell element (referred to by
Hughes and Liu as the U1 element) which is a four-node shell with uniformly reduced
integration, and summarizes the modifications to their theory as it is implemented in LS-
DYNA. A detailed discussion of these modifications, as well as those associated with the
implementation of the Hughes-Liu shell element in NIKE3D, are presented in an article by
Hallquist and Benson [1986].
13.1 Geometry
The Hughes-Liu shell element is based on a degeneration of the standard 8-node
brick element formulation, an approach originated by Ahmad et al. [1970]. Recall from the
discussion of the solid elements the isoparametric mapping of the biunit cube:
𝐱(𝜉 , 𝜂, 𝜁 ) = 𝑁𝑎 (𝜉 , 𝜂, 𝜁 )𝐱𝑎 , (13.1)
(1 + 𝜉𝑎 𝜉 )(1 + 𝜂𝑎 𝜂)(1 + 𝜁𝑎 𝜁 )
𝑁𝑎 (𝜉 , 𝜂, 𝜁 ) = , (13.2)
8
where 𝐱 is an arbitrary point in the element, (𝜉 , 𝜂, 𝜁 ) are the parametric coordinates, 𝐱𝑎 are
the global nodal coordinates of node 𝑎, and 𝑁𝑎 are the element shape functions evaluated at
node 𝑎, i.e., (𝜉𝑎 , 𝜂𝑎 , 𝜁𝑎 ) are (𝜉 , 𝜂, 𝜁 ) evaluated at node 𝑎.
In the shell geometry, planes of constant 𝜁 will define the lamina or layers of the
shell and fibers are defined by through-the-thickness lines when both 𝜉 and 𝜂 are constant
(usually only defined at the nodes and thus referred to as ‘nodal fibers’). To degenerate the
8-node brick geometry into the 4-node shell geometry, the nodal pairs in the 𝜁 direction
(through the shell thickness) are combined into a single node, for the translation degrees of
freedom, and an inextensible nodal fiber for the rotational degrees of freedom. Figure 13.1
shows a schematic of the bi-unit cube and the shell element.
The mapping of the bi-unit cube into the shell element is separated into two parts
𝐱(𝜉 , 𝜂, 𝜁 ) = 𝐱̅(𝜉 , 𝜂) + 𝐗(𝜉 , 𝜂, 𝜁 ), (13.3)
where 𝐱̅ denotes a position vector to a point on the reference surface of the shell and X is a
position vector, based at point 𝐱̅ on the reference, that defines the fiber direction through
that point. In particular, if we consider one of the four nodes which define the reference
surface, then
𝐱̅ (𝜉 , 𝜂) = 𝑁𝑎 (𝜉 , 𝜂) 𝐱̅𝑎 , (13.4)
With this description, arbitrary points on the reference surface 𝐱̅ are interpolated by
the two-dimensional shape function 𝑁(𝜉 , 𝜂) operating on the global position of the four
shell nodes that define the reference surfaces, i.e., 𝐱̅𝑎 . Points off the reference surface are
further interpolated by using a one-dimensional shape function along the fiber direction,
i.e., 𝐗𝑎 (𝜁 ), where
̂𝑎 ,
𝐗𝑎 (𝜁 ) = 𝑧𝑎 (𝜁 ) 𝐗 (13.6)
𝑧𝑎 (𝜁 ) = 𝑁+ (𝜁 )𝑧+ −
𝑎 + 𝑁− (𝜁 )𝑧𝑎 , (13.7)
(1 + 𝜁 )
𝑁+ (𝜁 ) = , (13.8)
2
ζ
Biunit Cube
ζ
X
ξ
η
Beam Element ξ
Nodal fibers
ζ̄ x̄ζ
x^ζ
0 x^
η
xζ-
-1 -
Bottom Surface zζ
Figure 13.1. Mapping of the biunit cube into the Hughes-Liu shell element and
nodal fiber nomenclature.
(1 − 𝜁 )
𝑁− (𝜁 ) = (13.9)
2
As shown in the lower portion of Figure 13.1, 𝐗 ̂𝑎 is a unit vector in the fiber direction and
𝑧(𝜁 ) is a thickness function. (Thickness changes (see [Hughes and Carnoy 1981]) are
accounted for by explicitly adjusting the fiber lengths at the completion of a time step based
on the amount of straining in the fiber direction. Updates of the fiber lengths always lag
one time step behind other kinematical quantities.)
The reference surface may be located at the mid-surface of the shell or at either of the
shell’s outer surfaces. This capability is useful in several practical situations involving
contact surfaces, connection of shell elements to solid elements, and offsetting elements
such as stiffeners in stiffened shells. The reference surface is located within the shell
element by specifying the value of the parameter 𝜁 ̅ (see lower portion of Figure 13.1).
When 𝜁 ̅ = – 1, 0, +1, the reference surface is located at the bottom, middle, and top surface
The Hughes-Liu formulation uses two position vectors, in addition to 𝜁 ,̅ to locate the
reference surface and define the initial fiber direction. The two position vectors 𝑥𝑎+ and 𝑥𝑎−
are located on the top and bottom surfaces, respectively, at node 𝑎. From these data the
following are obtained:
1
̅ 𝑎− + (1 + 𝜁 )𝑥
𝑥̅𝑎 = (1 − 𝜁 )𝑥 ̅ 𝑎+ , (13.10)
2
(𝑥𝑎+ − 𝑥𝑎− )
𝑋̂𝑎 = , (13.11)
ℎ𝑎
1
𝑧+
𝑎 =
̅ 𝑎,
(1 − 𝜁 )ℎ (13.12)
2
1
𝑧− ̅
𝑎 = − (1 + 𝜁 )ℎ𝑎 , (13.13)
2
13.2 Kinematics
The same parametric representation used to describe the geometry of the shell
element, i.e., reference surface and fiber vector interpolation, are used to interpolate the
shell element displacement, i.e., an isoparametric representation. Again, the displacements
are separated into the reference surface displacements and rotations associated with the
fiber direction:
𝐮(𝜉 , 𝜂, 𝜁 ) = 𝐮
̅̅̅̅(𝜉 , 𝜂) + 𝐔(𝜉 , 𝜂, 𝜁 ), (13.15)
̅̅̅̅(𝜉 , 𝜂) = 𝑁𝑎 (𝜉 , 𝜂)𝐮
𝐮 ̅̅̅̅𝑎 , (13.16)
̂𝑎 ,
𝐔𝑎 (𝜁 ) = 𝑧𝑎 (𝜁 )𝐔 (13.18)
where 𝐮 is the displacement of a generic point; 𝐮 ̅̅̅̅ is the displacement of a point on the
reference surface, and 𝐔 is the ‘fiber displacement’ rotations; the motion of the fibers can be
interpreted as either displacements or rotations as will be discussed.
Hughes and Liu introduce the notation that follows, and the associated schematic
shown in Figure 13.2, to describe the current deformed configuration with respect to the
reference configuration:
𝐲=𝐲
̅̅̅̅ + 𝐘, (13.19)
𝐲
̅̅̅̅ = 𝐱̅ + 𝐮
̅̅̅̅, (13.20)
𝐲
̅̅̅̅𝑎 = 𝐱̅𝑎 + 𝐮
̅̅̅̅𝑎 , (13.21)
𝐘 = 𝐗 + 𝐔, (13.22)
𝐘𝑎 = 𝐗𝑎 + 𝐔𝑎 , (13.23)
̂𝑎 = 𝐗
𝐘 ̂𝑎 + 𝐔
̂𝑎 . (13.24)
In the above relations, and in Figure 13.2, the 𝐱 quantities refer to the reference
configuration, the 𝐲 quantities refer to the updated (deformed) configuration and the 𝐮
quantities are the displacements. The notation consistently uses a superscript bar (⋅ ̅) to
indicate reference surface quantities, a superscript caret (⋅ ̂) to indicate unit vector
quantities, lower case letters for translational displacements, and upper case letters
indicating fiber displacements. To update to the deformed configuration, two vector
quantities are needed: the reference surface displacement 𝐮
̅̅̅̅ and the associated nodal fiber
displacement 𝐔. The nodal fiber displacements are defined in the fiber coordinate system,
described in the next subsection.
u U
X (parallel construction)
x
Y
ū
reference axis in
undeformed
x̄ geometry
Deformed Configuration
Reference Surface
Figure 13.2. Schematic of deformed configuration displacements and position
vectors.
At each node a unique local Cartesian coordinate system is constructed that is used
as the reference frame for the rotation increments. The relation presented by Hughes and
Liu for the nodal fiber displacements (rotations) is an incremental relation, i.e., it relates the
current configuration to the last state, not to the initial configuration. Figure 13.3 shows
𝑓 𝑓 𝑓
two triads of unit vectors: (𝐛1 , 𝐛2 , 𝐛3 ) comprising the orthonormal fiber basis in the
reference configuration (where the fiber unit vector is now 𝐘 ̂ = 𝐛𝑓 ) and (𝐛1 , 𝐛2 , 𝐛3 )
3
indicating the incrementally updated current configuration of the fiber vectors. The
reference triad is updated by applying the incremental rotations, Δ𝜃1 and Δ𝜃2 , obtained
𝑓 𝑓
from the rotational equations of motion, to the fiber vectors (𝐛 1 and 𝐛 2 ) as shown in
Figure 13.3. The linearized relationship between the components of Δ𝑈 ̂ in the fiber system
viz, Δ𝑈̂ 𝑓 , Δ𝑈
̂ 𝑓 , Δ𝑈
̂ 𝑓 , and the incremental rotations is given by
1 2 3
̂𝑓 ⎫
⎧Δ𝑈
{
{ 1𝑓 }
} −1 0
⎡ Δ𝜃1
̂
Δ𝑈
⎨ 2⎬ = ⎢ 0 −1⎤⎥ {Δ𝜃 }. (13.25)
{
{Δ𝑈 }
𝑓} ⎣ 0 0⎦ 2
⎩ ̂ ⎭3
Although the above Hughes-Liu relation for updating the fiber vector enables a
reduction in the number of nodal degrees of freedom from six to five, it is not implemented
in LS-DYNA because it is not applicable to beam elements.
fiber
^
b3f =Y
b3
b1f b2
Δθ2
Δθ1
b1 b2f
In LS-DYNA, three rotational increments are used, defined with reference to the
global coordinate axes:
⎧ ̂
{Δ𝑈1 ⎫
} 0 𝑌̂3 −𝑌̂2 ⎧Δ𝜃1 ⎫
⎡ ⎤ { }
⎨ Δ𝑈
{ ̂ ⎬} ⎢−𝑌̂3
̂2 = ⎢ 0 𝑌̂1 ⎥
⎥ ⎨
{
Δ𝜃2 ⎬.
}
(13.26)
⎩Δ𝑈3 ⎭ ⎣ 𝑌̂2 −𝑌̂1 0 ⎦ ⎩Δ𝜃3 ⎭
Equation (13.26) is adequate for updating the stiffness matrix, but for finite rotations
the error is significant. A more accurate second-order technique is used in LS-DYNA for
updating the unit fiber vectors:
𝑌̂𝑖𝑛+1 = 𝑅𝑖𝑗 (Δ𝜃)𝑌̂𝑖𝑛 , (13.27)
1
2𝐷 = 2 + (Δ𝜃12 + Δ𝜃22 + Δ𝜃32 ). (13.30)
2
Here, 𝛿𝑖𝑗 is the Kronecker delta and 𝑒𝑖𝑗𝑘 is the permutation tensor. This rotational update is
often referred to as the Hughes-Winget formula [Hughes and Winget 1980]. An exact rota-
tional update using Euler angles or Euler parameters could easily be substituted in
Equation (13.27), but it is doubtful that the extra effort would be justified.
η = con η
stant
nt
sta
e^3
con
e^2
ξ=
ξ
e^1
layers through the thickness of the shell that correspond to the locations and associated
thicknesses of the through-the-thickness shell integration points; the analogy is that of
lamina in a fibrous composite material. The orthonormal lamina basis (Figure 13.4), with
one direction 𝑒3̂ normal to the lamina of the shell, is constructed at every integration point
in the shell.
The lamina basis is constructed by forming two unit vectors locally tangent to the
lamina:
𝐲,𝜉
𝐞̂1 = , (13.31)
∥𝐲,𝜉 ∥
𝐲,𝜂
𝐞′2 = , (13.32)
∥𝐲,𝜂 ∥
where, as before, 𝐲 is the position vector in the current configuration. The normal to the
lamina at the integration point is constructed from the vector cross product of these local
tangents:
𝐞̂3 = 𝐞̂1 × 𝐞′2 , (13.33)
The transformation of vectors from the global to lamina coordinate system can now
be defined in terms of the lamina basis vectors as
⎧𝐴𝑥̂ ⎫ 𝑒1𝑥 𝑒2𝑥 𝑒3𝑥 T ⎧𝐴𝑥 ⎫
{ }
{ } { }
̂ = 𝐴𝑦̂ = ⎡
𝐀 𝑒 𝑒 𝑒 ⎤ 𝐴𝑦 = 𝐪𝐀, (13.35)
⎨ ⎬ ⎢ 1𝑦 2𝑦 3𝑦 ⎥ ⎨ ⎬
{
{𝐴̂ }} ⎣𝑒1𝑧 𝑒2𝑧 𝑒3𝑧 ⎦ {
⎩ 𝐴 }
⎭
⎩ 𝑧⎭ 𝑧
̂ is a
where 𝑒𝑖𝑥 , 𝑒𝑖𝑦 , 𝑒𝑖𝑧 are the global components of the lamina coordinate unit vectors; 𝐀
vector in the lamina coordinates, and 𝐴 is the same vector in the global coordinate system.
𝑛+1⁄2 𝑛+1⁄2
Δ𝜀𝑖𝑗𝑙 = 𝑞𝑖𝑘 Δ𝜀𝑘𝑛 𝑞𝑗𝑛 , (13.41)
where the superscript 𝑙 indicates components in the lamina (local) coordinate system.
where 𝐟𝑎int are the internal forces at node 𝑎, 𝐁𝑎 is the strain-displacement matrix in the
lamina coordinate system associated with the displacements at node 𝑎, and 𝐓𝑎 is the
transformation matrix relating the global and lamina components of the strain-
displacement matrix. Because the B matrix relates six strain components to twenty-four
displacements (six degrees of freedom at four nodes), it is convenient to partition the B
matrix into four groups of six:
𝐁 = [𝐁1 𝐁2 𝐁3 𝐁4 ], (13.45)
Each 𝐁𝑎 submatrix is further partitioned into a portion due to strain and spin:
𝐁𝜀𝑎
𝐁𝑎 = [ ], (13.46)
𝐁𝜔𝑎
𝐵 0 0 𝐵4 0 0
⎡0 1 𝐵2 0 0 𝐵5 0 ⎤
⎢ ⎥
𝐁𝑎 = ⎢
𝜀
⎢𝐵̅̅̅̅ 2 ̅̅̅̅ 1
𝐵 0 ̅̅̅̅ 5
𝐵 ̅̅̅̅ 4
𝐵 0 ⎥ ⎥, (13.47)
⎢0 ̅̅̅̅ 3
𝐵 ̅̅̅̅ 2
𝐵 0 ̅̅̅̅ 6
𝐵 ̅̅̅̅ 5 ⎥
𝐵
⎢ ⎥
⎣𝐵̅̅̅̅ 3 0 ̅̅̅̅ 1
𝐵 ̅̅̅̅ 6
𝐵 0 ̅̅̅̅ 4 ⎦
𝐵
̅̅̅̅
𝐵 −𝐵 ̅̅̅̅ 1 0 ̅̅̅̅ 5
𝐵 −𝐵 ̅̅̅̅ 4 0
⎡ 2 ⎤
𝐁𝜔
𝑎 =⎢
⎢ 0 𝐵̅̅̅̅ 3 −𝐵 ̅̅̅̅ 2 0 ̅̅̅̅ 6
𝐵 −𝐵 ̅̅̅̅ 5 ⎥
⎥, (13.48)
⎣−𝐵 ̅̅̅̅ 3 0 ̅̅̅̅ 1
𝐵 −𝐵 ̅̅̅̅ 6 0 ̅̅̅̅ 4 ⎦
𝐵
where
⎧ ∂𝑁𝑎
{
{ 𝑁𝑎,𝑖 = for 𝑖 = 1, 2, 3
{ ∂𝑦𝑖𝑙
𝐵𝑖 = ⎨ . (13.49)
{ ∂(𝑁𝑎 𝑧𝑎 )
{
{(𝑁𝑎 𝑧𝑎 ),𝑖−3 = ∂𝑦𝑙 for 𝑖 = 4, 5, 6
⎩ 𝑖−3
• The superscript bar indicates the 𝐵’s are evaluated at the center of the lamina (0, 0,
𝜁 ). The strain-displacement matrix uses the ‘B-Bar’ (𝐵 ̅̅̅̅ )approach advocated by
Hughes [1980]. In the NIKE3D and DYNA3D implementations, this entails replac-
ing certain rows of the B matrix and the strain increments with their counterparts
evaluated at the center of the element. In particular, the strain-displacement matrix
is modified to produce constant shear and spin increments throughout the lamina.
• The resulting B-matrix is a 8 × 24 matrix. Although there are six strain and three
rotations increments, the B matrix has been modified to account for the fact that 𝜎33
will be zero in the integration of Equation (13.44).
but cannot be used effectively in explicit calculations where matrix inversions are not
feasible. In LS-DYNA only three and four-node shell elements are used with linear
interpolation functions; consequently, we compute the translational masses from the
consistent mass matrix by row summing, leading to the following mass at element node a:
The rotational masses are computed by scaling the translational mass at the node by the
factor 𝛼:
𝑀rot𝑎 = ∝ 𝑀disp𝑎 , (13.52)
∝ = max{∝1 , ∝2 }, (13.53)
1
∝1 = ⟨𝑧𝑎 ⟩2 + [𝑧𝑎 ]2 , (13.54)
12
𝑉
∝2 = , (13.55)
8ℎ
(𝑧+ −
𝑎 + 𝑧𝑎 )
⟨𝑧𝑎 ⟩ = , (13.56)
2
[𝑧𝑎 ] = 𝑧+ −
𝑎 − 𝑧𝑎 . (13.57)
Hughes and Carnoy integrate the strain tensor through the thickness of the shell in
order to determine a mean value Δ𝜀̅𝑖𝑗 :
1 1
Δ𝜀̅𝑖𝑗 = ∫ Δ𝜀𝑖𝑗 𝑑𝜁 , (13.58)
2 −1
and then project it to determine the straining in the fiber direction:
𝛆̅ 𝑓 = 𝐘
̂ T Δ𝛆̅𝐘
̂. (13.59)
Using the interpolation functions through the integration points the strains in the fiber
directions are extrapolated to the nodal points if 2 × 2 selectively reduced integration is
employed. The nodal fiber lengths can now be updated:
𝑓
ℎ𝑛+1
𝑎 = ℎ𝑛𝑎 (1 + 𝜀̅𝑎 ). (13.60)
Figure 13.5. Selectively reduced integration rule results in four inplane points
being used.
2. the operation count increases three- to fourfold. The level 3 loop is added as
shown in Figure 13.6
However, these disadvantages can be more than offset by the increased reliability and
accuracy.
We have implemented two version of the Hughes-Liu shell with selectively reduced
integration. The first closely follows the intent of the original paper, and therefore no
assumptions are made to reduce costs, which are outlined in operation counts in Table 10.1.
These operation counts can be compared with those in Table 10.2 for the Hughes-Liu shell
with uniformly reduced integration. The second formulation, which reduces the number of
operation by more than a factor of two, is referred to as the co-rotational Hughes-Liu shell
in the LS-DYNA user’s manual. This shell is considerably cheaper due to the following
simplifications:
• Strains rates are not centered. The strain displacement matrix is only computed at
time 𝑛 + 1 and not at time 𝑛 + 1 ⁄ 2.
• The stresses are stored in the local shell system following the Belytschko-Tsay shell.
The transformations of the stresses between the local and global coordinate systems
are thus avoided.
• The Jaumann rate rotation is not performed, thereby avoiding even more computa-
tions. This does not necessarily preclude the use of the shell in large deformations.
• To study the effects of these simplifying assumptions, we can compare results with
those obtained with the full Hughes-Liu shell. Thus far, we have been able to get
comparable results.
LEVEL 2 - Completion
LEVEL L1 - Completion
Figure 13.6. An inner loop, LEVEL 3, is added for the Hughes-Liu shell with
selectively reduced integration.
Table 10.1. Operation counts for the Hughes-Liu shell with selectively reduced
integration.
Table 10.2. Operation counts for the LS-DYNA implementation of the uniformly
reduced
Hughes-Liu shell.
14
Transverse Shear Treatment For
Layered Shell
The shell element formulations that include the transverse shear strain components
are based on the first order shear deformation theory, which yield constant through
thickness transverse shear strains. This violates the condition of zero traction on the top
and bottom surfaces of the shell. Normally, this is corrected by the use of a shear correction
factor. The shear correction factor is 5/6 for isotropic materials; however, this value is
incorrect for sandwich and laminated shells. Not accounting for the correct transverse
shear strain and stress could yield a very stiff behavior in sandwich and laminated shells.
This problem is addressed here by the use of the equilibrium equations without gradient in
the y-direction as described by what follows. Consider the stresses in a layered shell:
(𝑖) ∘ (𝑖) ∘ (𝑖) ∘ (𝑖) ∘ (𝑖) (𝑖)
𝜎𝑥 (𝑖) = 𝐶11 (𝜀𝑥 + 𝑧𝜒𝑥 ) + 𝐶12 (𝜀𝑦 + 𝑧𝜒𝑦 ) = 𝐶11 𝜀𝑥 + 𝐶12 𝜀𝑦 + 𝑧(𝐶11 𝜒𝑥 + 𝐶12 𝜒𝑦 ),
(𝑖) ∘ (𝑖) ∘ (𝑖) (𝑖)
𝜎𝑦(𝑖) = 𝐶12 𝜀𝑥 + 𝐶22 𝜀𝑦 + 𝑧(𝐶12 𝜒𝑥 + 𝐶22 𝜒𝑦 , (14.1)
(𝑖) (𝑖) ∘
𝜏𝑥𝑦 = 𝐶44 (𝜀𝑥𝑦 + 𝑧𝜒𝑥𝑦 ).
Assume that the bending center 𝑧̅𝑥 is known. Then
(𝑖) (𝑖) (𝑖) (𝑖)
𝜎𝑥(𝑖) = (𝑧 − 𝑧̅𝑥 )(𝐶11 𝜒𝜒 + 𝐶12 𝜒𝑦 ) + 𝐶11 𝜀𝑥 (𝑧̅𝑥 ) + 𝐶12 𝜀𝑦 (𝑧̅𝑥 ). (14.2)
The bending moment is given by the following equation:
𝑁𝐿 𝑧𝑖 𝑁𝐿 𝑧𝑖
𝑀𝑥𝑥 = 𝜒𝑥 ⎛
⎜∑ 𝐶11(𝑖)
∫ (𝑧 − 𝑧̅𝑥 )2 𝑑𝑧⎞
⎟ + 𝜒𝑦 ⎛
⎜∑ 𝐶12(𝑖)
∫ (𝑧 − 𝑧̅𝑥 )2 𝑑𝑧⎞
⎟ (14.3)
⎝ 𝑖=1 𝑧𝑖−1 ⎠ ⎝ 𝑖=1 𝑧𝑖−1 ⎠
or
1 𝑁𝐿 (𝑖)
𝑀𝑥𝑥 = [𝜒𝑥 ∑ 𝐶11 [(𝑧3𝑖 − 𝑧3𝑖−1 ) − (𝑧𝑖 − 𝑧𝑖−1 )𝑧̅2𝑥 ] (14.4)
3 𝑖=1
𝑁𝐿 (𝑖)
+ 𝜒𝑦 ∑ 𝐶12 [(𝑧3𝑖 − 𝑧3𝑖−1 ) − (𝑧𝑖 − 𝑧𝑖−1 )𝑧̅2𝑥 ]]
𝑖=1
LS-DYNA DEV 10/27/16 (r:8004) 12-1 (Transverse Shear Treatment For Layered Shell)
Transverse Shear Treatment For Layered Shell LS-DYNA Theory Manual
then
𝑧 − 𝑧̅𝑥
𝜀𝑥 = = (𝑧 − 𝑧̅𝑥 )𝜒𝑥 , (14.6)
𝜌
and
1
𝑀𝑥𝑥 = (𝜒𝑥 (𝐸𝐼)𝑥 ), (14.7)
3
3𝑀𝑥𝑥
𝜒𝑥 = . (14.8)
(𝐸𝐼)𝑥
Therefore, the stress becomes
3𝑀𝑥𝑥 𝐸(𝑖)
𝑥 (𝑧 − 𝑧̅𝑥 )
𝜎𝑥(𝑖) = . (14.9)
(𝐸𝐼)𝑥
Now considering the first equilibrium equation, one can write the following:
(𝑗)
∂𝜏𝑥𝑧 ∂𝜎 3𝑄 𝐸 (𝑧 − 𝑧̅𝑥 )
= − 𝑥 = − 𝑥𝑧 𝑥 , (14.10)
∂𝑧 ∂𝑥 (𝐸𝐼)𝑥
2
3𝑄𝑥𝑧 𝐸𝑥 (𝑧 − 𝑧𝑧̅𝑥 )
(𝑗)
(𝑗) 2 (14.11)
𝜏𝑥𝑧 =− + 𝐶𝑗 ,
(𝐸𝐼)𝑥
where 𝑄𝑥𝑧 is the shear force and 𝐶𝑗 is the constant of integration. This constant is obtained
from the transverse shear stress continuity requirement at the interface of each layer. Let
then
𝑧2𝑖−1
𝑄𝑥𝑧 𝐸(𝑖)
𝑥 ( − 𝑧𝑖−1 𝑧̅𝑥 )
2 𝑖−1 (14.12)
𝐶𝑗 = + 𝜏𝑥𝑧 ,
(𝐸𝐼)𝑥
and
𝑄𝑥𝑧 𝐸(𝑖)
𝑥 𝑧2 𝑧2
(𝑖)
𝜏𝑥𝑧 (𝑖−1)
= 𝜏𝑥𝑧 + [ 𝑖−1 − 𝑧𝑖−1 𝑧̅𝑥 − + 𝑧𝑧̅𝑥 ]. (14.13)
(𝐸𝐼)𝑥 2 2
For the first layer
(1)
3𝑄𝑥𝑧 𝐶11 𝑧2 − 𝑧2𝑜
𝜏𝑥𝑧 = − [ − 𝑧̅𝑥 (𝑧 − 𝑧𝑜 )], (14.14)
(𝐸𝐼)𝑥 2
12-2 (Transverse Shear Treatment For Layered Shell) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Transverse Shear Treatment For Layered Shell
and
ℎ𝑗 = 𝑧𝑗 − 𝑧𝑗−1 . (14.18)
To find 𝑄𝑥𝑧 , the shear force, assume that the strain energy expressed through
̅ , is equal to the strain energy expressed through the derived
average shear modules, 𝐶66
expressions as follows:
1 𝑄𝑥𝑧 1 𝜏 2
𝑈= = ∫ 𝑥𝑧 𝑑𝑧, (14.19)
̅ ℎ 2 𝐶66
2 𝐶66
2 2
1 9ℎ 𝐶11 (𝑖) (𝑧2 − 𝑧2𝑖−1 )
= ∫ [𝑓 + − 𝑧̅𝑥 (𝑧 − 𝑧𝑖−1 )] 𝑑𝑧
̅
𝐶66 (𝐸𝐼)𝑥
2 𝐶66 𝑥 2
𝑧𝑖 2
9ℎ 𝑁𝐿
(𝐶11 (𝑖) )2 𝑧2 − 𝑧𝑖−1 2
= 2
∑ (𝑖)
∫ [𝑓𝑥(𝑖) + − 𝑧̅𝑥 (𝑧 − 𝑧𝑖−1 )] 𝑑𝑧
(𝐸𝐼)𝑥 𝑖=1 𝐶66 𝑧𝑖−1
2
(14.20)
𝑁𝐿 (𝑖) 2
1 9ℎ (𝐶11 ) ℎ 𝑖
= 2
∑ 𝑖
{𝑓𝑥 [60𝑓𝑥𝑖 + 20ℎ𝑖 (𝑧𝑖 + 2𝑧𝑖−1 − 3𝑧̅𝑥 )]
60 (𝐸𝐼) 𝑖 𝐶66
𝑥
+ 𝑧̅𝑥 ℎ𝑖 [20𝑧̅𝑥 ℎ𝑖 + 35𝑧2𝑖−1 − 10𝑧𝑖−1 (𝑧𝑖 + 𝑧𝑖−1 ) − 15𝑧2𝑖 ] + 𝑧𝑖 (𝑧𝑖 + 𝑧𝑖−1 )(3𝑧2𝑖
− 7𝑧2𝑖−1 ) + 8𝑧4𝑖−1 },
then
̅ 𝛾
𝑄𝑥𝑧 = 𝜏̅𝑥𝑧 ℎ = 𝐶66 ̅̅̅̅𝑥𝑧 ℎ, (14.21)
to calculate 𝑧̅𝑥 use 𝜏𝑥𝑧 for last layer at surface 𝑧 = 0,
𝑁𝐿
(𝑖) 𝑧2𝑖 − 𝑧2𝑖−1
∑ 𝐶11 [( ) − 𝑧̅𝑥 (𝑧𝑖 − 𝑧𝑖−1 )] = 0, (14.22)
𝑖=1
2
LS-DYNA DEV 10/27/16 (r:8004) 12-3 (Transverse Shear Treatment For Layered Shell)
Transverse Shear Treatment For Layered Shell LS-DYNA Theory Manual
where
𝑁𝐿 (𝑖)
∑𝑖=1 𝐶11 ℎ𝑖 (𝑧𝑖 + 𝑧𝑖+1 )
𝑧̅𝑥 = 𝑁𝐿
. (14.23)
(𝑖)
2 ∑𝑖=1 𝐶11 ℎ𝑖
Algorithm:
The following algorithm is used in the implementation of the transverse shear treatment.
𝑁𝐿 (𝑖) 2
4. Calculate ℎ[13 ∑𝑖=1 𝐶11 (𝑧3𝑖 − 𝑧3𝑖−1 )]
6. ̅ 𝛾
Calculate 𝑄𝑥𝑧 = 𝐶66 ̅̅̅̅𝑥𝑧 ℎ
Steps 1-5 are performed at the initialization stage. Step 6 is performed in the shell
formulation subroutine, and step 7 is performed in the stress calculation inside the
constitutive subroutine.
12-4 (Transverse Shear Treatment For Layered Shell) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Shell Element
15
Eight-Node Solid Shell Element
The isoparametric eight-node brick element discussed in Section 3 forms the basis
for tshell formulation 1, a solid shell element with enhancements based on the Hughes-Liu
and the Belytschko-Lin-Tsay shells. Like the eight-node brick, the geometry is interpolated
from the nodal point coordinates as:
8
𝑗
𝑥𝑖 (𝑋𝛼 , 𝑡) = 𝑥𝑖 (𝑋𝛼 (𝜉 , 𝜂, 𝜁 ), 𝑡) = ∑ 𝜙𝑗 (𝜉 , 𝜂, 𝜁 )𝑥𝑖 (𝑡), (15.1)
𝑗=1
1
𝜙𝑗 = (1 + 𝜉 𝜉𝑗 )(1 + 𝜂𝜂𝑗 )(1 + 𝜁 𝜁𝑗 ). (15.2)
8
As with solid elements, 𝐍 is the 3 × 24 rectangular interpolation matrix:
𝜑1 0 0 𝜑2 0 … 0 0
⎡0 𝜑1 0 0 𝜑2 … 𝜑8 0 ⎤
𝐍(𝜉 , 𝜂, 𝜁 ) = ⎢ ⎥, (15.3)
⎣0 0 𝜑1 0 0 … 0 𝜑8 ⎦
𝛔 is the stress vector:
𝛔T = (𝜎𝑥𝑥 , 𝜎𝑦𝑦 , 𝜎𝑧𝑧 , 𝜎𝑥𝑦 , 𝜎𝑦𝑧 , 𝜎𝑧𝑥 ), (15.4)
and 𝐁 is the 6 × 24 strain-displacement matrix:
5 Node ξ η ζ
8
1 -1 -1 -1
6 2 1 -1 -1
3 1 1 -1
4
η 4 -1 1 -1
7 5 -1 -1 1
6 1 -1 1
7 1 1 1
2 8 -1 1 1
3
ξ
∂
⎡ 0 0 ⎤
⎢∂𝑥 ⎥
⎢ ∂ ⎥
⎢0 0 ⎥
⎢ ∂𝑦 ⎥
⎢ ⎥
⎢ ∂⎥
⎢0 0
∂𝑧 ⎥
𝐁=⎢∂ ∂
⎥ 𝐍, (15.5)
⎢ ⎥
⎢ 0 ⎥
⎢∂𝑦 ∂𝑥 ⎥
⎢ ∂ ∂⎥
⎢0 ⎥
⎢ ∂𝑧 ∂𝑦⎥
⎢ ⎥
⎢∂ ∂⎥
⎣∂𝑧 0
∂𝑥⎦
∂𝜑 ∂𝑥 ∂𝑦 ∂𝑧 ∂𝜑𝑖 ∂𝜑𝑖
⎡ 𝑖⎤ ⎡ ⎤⎡ ⎤ ⎡ ⎤
⎢ ∂𝜉 ⎥ ⎢∂𝜉 ∂𝜉 ∂𝜉 ⎥ ⎢ ∂𝑥 ⎥ ⎢ ∂𝑥 ⎥
⎢∂𝜑 ⎥ ⎢∂𝑥 ∂𝑧 ⎥
⎢ 𝑖⎥ ⎢ ∂𝑦 ⎥⎢ ∂𝜑𝑖 ⎥ ⎢∂𝜑 ⎥
⎥ = 𝐉 ⎢ 𝑖 ⎥.
⎢ ⎥=⎢ ⎥⎢ (15.7)
⎢ ∂𝜂 ⎥ ⎢∂𝜂 ∂𝜂 ∂𝜂⎥ ⎢ ∂𝑦 ⎥ ⎢ ∂𝑦 ⎥
⎢∂𝜑 ⎥ ⎢∂𝑥 ⎥ ⎢ ⎥ ⎢ ⎥
⎢ 𝑖⎥ ⎢ ∂𝑦 ∂𝑧 ⎥ ⎢∂𝜑𝑖 ⎥ ⎢∂𝜑𝑖 ⎥
⎣ ∂𝜁 ⎦ ⎣∂𝜁 ∂𝜁 ∂𝜁 ⎦ ⎣ ∂𝑧 ⎦ ⎣ ∂𝑧 ⎦
Inverting the Jacobian matrix, 𝐉, we can solve for the desired terms
∂𝜑 ∂𝜑
⎡ 𝑖⎤ ⎡ 𝑖⎤
⎢ ∂𝑥 ⎥ ⎢ ∂𝜉 ⎥
⎢∂𝜑𝑖 ⎥ ⎢∂𝜑 ⎥
⎢ ⎥ = 𝐉−1 ⎢
⎢ ⎥.
𝑖⎥
(15.8)
⎢ ∂𝑦 ⎥ ⎢ ∂𝜂 ⎥
⎢ ⎥ ⎢∂𝜑 ⎥
⎢∂𝜑𝑖 ⎥ ⎢ 𝑖⎥
⎣ ∂𝑧 ⎦ ⎣ ∂𝜁 ⎦
To obtain shell-like behavior from the solid element, it is necessary to use multiple
integration points through the shell thickness along the 𝜁 axis while employing a plane
stress constitutive subroutine. Consequently, it is necessary to construct a reference surface
within the brick shell. We locate the reference surface midway between the upper and
lower surfaces and construct a local coordinate system exactly as was done for the
Belytschko-Lin-Tsay shell element. Following the procedure outlined in Section 7,
Equations (7.1) – (7.3), the local coordinate system can be constructed as depicted in Figure
15.2. Equation (7.5a) gives the transformation matrix in terms of the local basis:
⎧ 𝐴 𝑒 𝑒 𝑒 ⎧𝐴𝑥̂ ⎫
{ 𝑥⎫} ⎡ 1𝑥 2𝑥 3𝑥 ⎤ { { } }
{𝐀} = ⎨𝐴𝑦 ⎬ = ⎢𝑒1𝑦 𝑒2𝑦 𝑒3𝑦 ⎥ ⎨𝐴𝑦̂ ⎬ = [𝛍]{𝐀
̂ } = [𝐪]T {𝐀
̂ }. (15.9)
{ }
⎩𝐴𝑧 ⎭ ⎣ 𝑒 𝑒 𝑒 ⎦ {
{ }
}
1𝑧 2𝑧 3𝑧
⎩𝐴𝑧̂ ⎭
As with the Hughes-Liu shell, the next step is to perform the Jaumann rate update:
𝜎𝑖𝑗𝑛+1 = 𝜎𝑖𝑗𝑛 + 𝜎𝑖𝑝
𝑛 𝑛
Δ𝜔𝑝𝑗 + 𝜎𝑗𝑝 Δ𝜔𝑝𝑖 , (15.10)
to account for the material rotation between time steps 𝑛 and 𝑛 + 1. The Jaumann rate
update of the stress tensor is applied in the global configuration before the constitutive
evaluation is performed. In the solid shell, as in the Hughes-Liu shell, the stresses and
history variables are stored in the global coordinate system. To evaluate the constitutive
relation, the stresses and the strain increments are rotated from the global to the lamina
coordinate system using the transformation defined previously:
𝑛+1
𝑛+1
𝜎𝑖𝑗𝑙 = 𝑞𝑖𝑘 𝜎𝑘𝑛 𝑞𝑗𝑛 , (15.11)
𝑛+1⁄2 𝑛+1⁄2
Δ𝜀𝑖𝑗𝑙 = 𝑞𝑖𝑘 Δ𝜀𝑘𝑛 𝑞𝑗𝑛 , (15.12)
^
y s3 3
4
e^3 r31
r42
e^2
s1 ^
x
1 e^1
r21 2
Figure 15.2. Construction of the reference surface in the solid shell element.
where the superscript l indicates components in the lamina (local) coordinate system. The
stress is updated incrementally:
𝑛+1 𝑛+1 𝑛+1⁄2
𝜎𝑖𝑗𝑙 = 𝜎𝑖𝑗𝑙 + Δ𝜎𝑖𝑗𝑙 . (15.13)
where 𝐸 is the elastic Young’s modulus for the material. The stress tensor of Equation
(15.13) is rotated back to the global system:
𝑙 𝑛+1
𝜎𝑖𝑗𝑛+1 = 𝑞𝑘𝑖 (𝜎𝑘𝑛 ) 𝑞𝑛𝑗 . (15.16)
A penalty stress tensor is then formed by transforming the normal penalty stress tensor (a
null tensor except for the 33 term) back to the global system:
𝑛+1
penalty n+1 penalty 𝑙
(𝜎𝑖𝑗 ) = 𝑞𝑘𝑖 [(𝜎𝑖𝑗 )] 𝑞𝑛𝑗 , (15.17)
before computing the internal force vector. The internal force vector can now be computed:
T 𝑛+1
𝐟int = ∫(𝐁𝑛+1 ) [𝝈 𝑛+1 + (𝝈 penalty ) ] 𝑑𝜐. (15.18)
The brick shell exhibits no discernible locking problems with this approach.
The treatment of the hourglass modes is identical to that described for the solid
elements in Section 3.
16
Eight-Node Solid Element for Thick
Shell Simulations
16.1 Abstract
Tshell formulation 3 is an eight-node hexahedral element incorporated into LS-
DYNA to simulate thick shell structures. The element formulation is derived in a co-
rotational coordinate system and the strain operator is calculated with a Taylor series
expansion about the center of the element. Special treatments are made on the dilatational
strain component and shear strain components to eliminate the volumetric and shear
locking. The use of consistent tangential stiffness and geometric stiffness greatly improves
the convergence rate in implicit analysis.
16.2 Introduction
Large-scale finite element analyses are extensively used in engineering designs and
process controls. For example, in automobile crashworthiness, hundreds of thousands of
unknowns are involved in the computer simulation models, and in metal forming
processing, tests in the design of new dies or new products are done by numerical
computations instead of costly experiments. The efficiency of the elements is of crucial
importance to speed up the design processes and reduce the computational costs for these
problems. Over the past ten years, considerable progress has been achieved in developing
fast and reliable elements.
DYNA, the eight-node solid thick shell element is still based on the Hughes-Liu and
Belytschko-Lin-Tsay shells [Hallquist, 1998]. A new eight-node solid element based on Liu,
1985, 1994 and 1998 is incorporated into LS-DYNA, intended for thick shell simulation.
The strain operator of this element is derived from a Taylor series expansion and special
treatments on strain components are utilized to avoid volumetric and shear locking.
The organization of this paper is as follows. The element formulations are described
in the next section. Several numerical problems are studied in the third section, followed
by the conclusions.
8
𝑣𝑖 = ∑ 𝑁𝑎 (𝜉 , 𝜂, 𝜁 )𝑣𝑖𝑎 , 𝑖 = 1, 2, 3, (16.2)
𝑎=1
1
𝑁𝑎 (𝜉 , 𝜂, 𝜁 ) = (1 + 𝜉𝑎 𝜉 )(1 + 𝜂𝑎 𝜂)(1 + 𝜁𝑎 𝜁 ), (16.3)
8
and the subscripts 𝑖 and a denote coordinate components ranging from one to three and the
element nodal numbers ranging from one to eight, respectively. The referential coordinates
𝜉 , 𝜂, and 𝜁 of node a are denoted by 𝜉𝑎 , 𝜂𝑎 , and 𝜁𝑎 , respectively.
The strain rate (or rate of deformation), 𝛆̇, is composed of six components,
𝛆̇T = [𝜀𝑥𝑥 𝜀𝑦𝑦 𝜀𝑧𝑧 𝜀𝑥𝑦 𝜀𝑦𝑧 𝜀𝑧𝑥 ], (16.4)
and is related to the nodal velocities by a strain operator, ̅𝐁
̅̅̅̅,
̅̅̅̅̅(𝜉 , 𝜂, 𝜁 )𝐯,
𝛆̇ = 𝐁 (16.5)
where
𝐯T = [vx1 vy1 vz1 ⋯ vx8 vy8 vz8 ], (16.6)
14-2 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
̅B̅̅̅
⎡ 𝑥𝑥 ⎤ 𝐵1 (1) 0 0 ⋯ 𝐵1 (8) 0 0
⎢̅B̅̅̅𝑦𝑦 ⎥ ⎡ 0 𝐵2 (1) 0 ⋯ 0 𝐵2 (8) 0 ⎤
⎢ ⎥ ⎢ ⎥
⎢̅B̅̅̅ ⎥ ⎢ 𝐵3 (8)⎥
̅̅̅̅ = ⎢ 𝑧𝑧 ⎥ = ⎢ 0
̅𝐁 0 𝐵3 (1) ⋯ 0 0 ⎥, (16.7)
⎢B ̅̅̅̅ ⎥ ⎢𝐵 (1) 𝐵1 (1) 0 ⋯ 𝐵2 (8) 𝐵1 (8) 0 ⎥
⎢ 𝑥𝑦 ⎥ ⎢ 2 ⎥
⎢̅B̅̅̅ ⎥ ⎢ 0 𝐵3 (1) 𝐵2 (1) ⋯ 0 𝐵3 (8) 𝐵2 (8)⎥
⎢ 𝑦𝑧 ⎥ ⎣𝐵 (1) 0 𝐵1 (1) ⋯ 𝐵3 (8) 0 𝐵1 (8)⎦
3
⎣̅B̅̅̅𝑧𝑥 ⎦
B1 𝑁,𝑥 (𝜉 , 𝜂, 𝜁 )
⎡B ⎤ = ⎡ ⎤
⎢ 2⎥ ⎢ ⎢𝑁,𝑦 (𝜉 , 𝜂, 𝜁 )⎥⎥. (16.8)
⎣B3 ⎦ ⎣𝑁,𝑧 (𝜉 , 𝜂, 𝜁 ) ⎦
Let
𝐱1T = 𝐱T = [𝑥1 𝑥2 𝑥3 𝑥4 𝑥5 𝑥6 𝑥7 𝑥8 ], (16.10)
The gradient vectors and their derivatives with respect to the natural coordinates at
the center of the element are given as follows,
LS-DYNA DEV 10/27/16 (r:8004) 14-3 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
1
𝐛1 = 𝐍,𝑥 (0) = [𝐷11 𝛏 + 𝐷12 𝛈 + 𝐷13 𝛇], (16.18)
8
1
𝐛2 = 𝐍,𝑦 (0) = [𝐷21 𝛏 + 𝐷22 𝛈 + 𝐷23 𝛇], (16.19)
8
1
𝐛3 = 𝐍,𝑧 (0) = [𝐷31 𝛏 + 𝐷32 𝛈 + 𝐷33 𝛇], (16.20)
8
1
𝐛1,𝜉 = 𝐍,𝑥𝜉 (0) = [𝐷12 𝜸1 + 𝐷13 𝛄2 ], (16.21)
8
1
𝐛2,𝜉 = 𝐍,𝑦𝜉 (0) = [𝐷22 𝛄1 + 𝐷23 𝛄2 ], (16.22)
8
1
𝐛3,𝜉 = 𝐍,𝑧𝜉 (0) = [𝐷32 𝛄1 + 𝐷33 𝛄2 ], (16.23)
8
1
𝐛1,𝜂 = 𝐍,𝑥𝜂 (0) = [𝐷11 𝛄1 + 𝐷13 𝛄3 ], (16.24)
8
1
𝐛2,𝜂 = 𝐍,𝑦𝜂 (0) = [𝐷21 𝛄1 + 𝐷23 𝛄3 ], (16.25)
8
1
𝐛3,𝜂 = 𝐍,𝑧𝜂 (0) = [𝐷31 𝛄1 + 𝐷33 𝛄3 ], (16.26)
8
1
𝐛1,𝜁 = 𝐍,𝑥𝜁 (0) = [𝐷11 𝛄2 + 𝐷12 𝛄3 ], (16.27)
8
1
𝐛2,𝜁 = 𝐍,𝑦𝜁 (0) = [𝐷21 𝛄2 + 𝐷22 𝛄3 ], (16.28)
8
1
𝐛3,𝜁 = 𝐍,𝑧𝜁 (0) = [𝐷31 𝛄2 + 𝐷32 𝛄3 ], (16.29)
8
1
𝐛1,𝜉𝜂 = 𝐍,𝑥𝜉𝜂 (0) = [𝐷13 𝛄4 − (𝐩T1 𝐱𝑖 )𝐛𝑖,𝜉 − (𝐫1T 𝐱𝑖 )𝐛𝑖,𝜂 ], (16.30)
8
1
𝐛2,𝜉𝜂 = 𝐍,𝑦𝜉𝜂 (0) = [𝐷23 𝛄4 − (𝐛T2 𝐱𝑖 )𝐛𝑖,𝜉 − (𝐫2T 𝐱𝑖 )𝐛𝑖,𝜂 ], (16.31)
8
1
𝐛3,𝜉𝜂 = 𝐍,𝑧𝜉𝜂 (0) = [𝐷33 𝛄4 − (𝐩T3 𝐱𝑖 )𝐛𝑖,𝜉 − (𝐫3T 𝐱𝑖 )𝐛𝑖,𝜂 ], (16.32)
8
1
𝐛1,𝜂𝜁 = 𝐍,𝑥𝜂𝜁 (0) = [𝐷11 𝛄4 − (𝐪1T 𝐱𝑖 )𝐛𝑖,𝜂 − (𝐩T1 𝐱𝑖 )𝐛𝑖,𝜁 ], (16.33)
8
14-4 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
1
𝐛2,𝜂𝜁 = 𝐍,𝑦𝜂𝜁 (0) = [𝐷21 𝛄4 − (𝐪2T 𝐱𝑖 )𝐛𝑖,𝜂 − (𝐩T2 𝐱𝑖 )𝐛𝑖,𝜁 ], (16.34)
8
1
𝐛3,𝜂𝜁 = 𝐍,𝑧𝜂𝜁 (0) = [𝐷31 𝛄4 − (𝐪3T 𝐱𝑖 )𝐛𝑖,𝜂 − (𝐩T3 𝐱𝑖 )𝐛𝑖,𝜁 ], (16.35)
8
1
𝐛1,𝜁𝜉 = 𝐍,𝑥𝜁𝜉 (0) = [𝐷12 𝛄4 − (𝐫1T 𝐱𝑖 )𝐛𝑖,𝜁 − (𝐪1T 𝐱𝑖 )𝐛𝑖,𝜉 ], (16.36)
8
1
𝐛2,𝜁𝜉 = 𝐍,𝑦𝜁𝜉 (0) = [𝐷22 𝛄4 − (𝐫2T 𝐱𝑖 )𝐛𝑖,𝜁 − (𝐪2T 𝐱𝑖 )𝐛𝑖,𝜉 ], (16.37)
8
1
𝐛3,𝜁𝜉 = 𝐍,𝑧𝜁𝜉 (0) = [𝐷32 𝛄4 − (𝐫3T 𝐱𝑖 )𝐛𝑖,𝜁 − (𝐪3T 𝐱𝑖 )𝐛𝑖,𝜉 ], (16.38)
8
where
𝐩𝑖 = 𝐷𝑖1 𝐡1 + 𝐷𝑖3 𝐡3 , (16.39)
The strain operators, 𝐁̅̅̅̅̅(𝜉 , 𝜂, 𝜁 ), can be decomposed into two parts, the dilatational
dil
̅̅̅̅̅ (𝜉 , 𝜂, 𝜁 ), and the deviatoric part, ̅𝐁 ̅̅̅̅dev (𝜉 , 𝜂, 𝜁 ), both of which can be expanded
part, 𝐁
about the element center as in Equation (16.9)
̅̅̅̅̅dil (𝛏, 𝛈, 𝛇) = ̅𝐁
𝐁 ̅̅̅̅dil
̅̅̅̅dil (0) + ̅𝐁 ̅̅̅̅̅dil ̅̅̅̅̅dil
,𝜉 (0)𝛏 + 𝐁,𝜂 (0)𝛈 + 𝐁,𝜁 (0)𝛇
(16.47)
+2[𝐁 ̅̅̅̅̅dil ̅̅̅̅̅dil ̅̅̅̅̅dil
,𝜉𝜂 (0)𝛏𝛈 + 𝐁,𝜂𝜁 (0)𝛈𝛇 + 𝐁,𝜁𝜉 (0)𝛇𝛏],
LS-DYNA DEV 10/27/16 (r:8004) 14-5 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
̅̅̅̅̅dev (𝜉 , 𝜂, 𝜁 ) = 𝐁
𝐁 ̅̅̅̅dev
̅̅̅̅̅dev (0) + ̅𝐁 ̅̅̅̅̅dev ̅̅̅̅̅dev
,𝜉 (0)𝜉 + 𝐁,𝜂 (0)𝛈 + 𝐁,𝜁 (0)𝛇
(16.48)
+2[𝐁 ̅̅̅̅̅dev ̅̅̅̅̅dev ̅̅̅̅̅dev
,𝜉𝜂 (0)𝛏𝛈 + 𝐁,𝜂𝜁 (0)𝛈𝛇 + 𝐁,𝜁𝜉 (0)𝛇𝛏],
To avoid volumetric locking, the dilatational part of the strain operators is evaluated
only at one quadrature point, the center of the element, i.e., they are constant terms
̅̅̅̅̅dil (𝝃 , 𝜼, 𝜻) = ̅𝐁
𝐁 ̅̅̅̅dil (0). (16.49)
̅̅̅̅dev
B ̅̅̅̅dev ̅̅̅̅dev ̅̅̅̅dev ̅̅̅̅dev
𝑧𝑧 (𝜉 , 𝜂, 𝜁 ) = B𝑧𝑧 (0) + B𝑧𝑧,𝜉 (0)𝜉 + B𝑧𝑧,𝜂 (0)𝜂 + B𝑧𝑧,𝜁 (0)𝜁
(16.52)
+2[B ̅̅̅̅dev ̅̅̅̅dev ̅̅̅̅dev
𝑧𝑧,𝜉𝜂 (0)𝜉𝜂 + B𝑧𝑧,𝜂𝜁 (0)𝜂𝜁 + B𝑧𝑧,𝜁𝜉 (0)𝜁𝜉 ],
̅̅̅̅dev
B ̅̅̅̅dev ̅̅̅̅dev
𝑥𝑦 (𝜉 , 𝜂, 𝜁 ) = B𝑥𝑦 (0) + B𝑥𝑦,𝜁 (0)𝜁 , (16.53)
̅̅̅̅dev
B ̅̅̅̅dev ̅̅̅̅dev
𝑧𝑥 (𝜉 , 𝜂, 𝜁 ) = B𝑧𝑥 (0) + B𝑧𝑥,𝜂 (0)𝜂. (16.55)
Here, only one linear term is left for shear strain components such that the modes
causing shear locking are removed. The normal strain components keep all non-constant
terms given in equation (16.48).
̅̅̅̅𝑦𝑦 (𝜉 , 𝜂, 𝜁 ) = B
B ̅̅̅̅𝑦𝑦 (0) + B ̅̅̅̅dev ̅̅̅̅dev ̅̅̅̅dev
𝑦𝑦,𝜉 (0)𝜉 + B𝑦𝑦,𝜂 (0)𝜂 + B𝑦𝑦,𝜁 (0)𝜁
(16.57)
+2[B ̅̅̅̅dev ̅̅̅̅dev ̅̅̅̅dev
𝑦𝑦,𝜉𝜂 (0)𝜉𝜂 + B𝑦𝑦,𝜂𝜁 (0)𝜂𝜁 + B𝑦𝑦,𝜁𝜉 (0)𝜁𝜉 ],
14-6 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
̅̅̅̅𝑥𝑦 (𝜉 , 𝜂, 𝜁 ) = B
B ̅̅̅̅dev
̅̅̅̅𝑥𝑦 (0) + B 𝑥𝑦,𝜁 (0)𝜁 , (16.59)
̅̅̅̅𝑦𝑧 (𝜉 , 𝜂, 𝜁 ) = B
B ̅̅̅̅dev
̅̅̅̅𝑦𝑧 (0) + B 𝑦𝑧,𝜉 (0)𝜉 , (16.60)
It is noted that the elements developed above cannot pass the patch test if the
elements are skewed. To remedy this drawback, the gradient vectors defined in (16.18)–
(16.20) are replaced by the uniform gradient matrices, proposed by Flanagan [1981],
b̃ B1 (𝜉 , 𝜂, 𝜁 )
⎡ 1⎤ 1 ⎡ ⎤
⎢b̃2 ⎥
⎢ ⎥ = 𝑉 ∫Ω ⎢B2 (𝜉 , 𝜂, 𝜁 )⎥ 𝑑𝑉 . (16.62)
𝑒
⎣b̃3 ⎦
𝑒
⎣B3 (𝜉 , 𝜂, 𝜁 )⎦
Where 𝑉𝑒 is the element volume and the stabilization vector are redefined as
𝛄̃ 𝛼 = 𝐡𝛼 − (𝐡𝛼T 𝐱𝑖 )𝐛̃𝑖 . (16.63)
The element using the strain submatrices (16.56)-(16.61) and uniform gradient
matrices (16.62) with four-point quadrature scheme is called HEXDS element.
g2
e^3 e^1
g1
y e^2
ζ=0
x
Figure 16.1. Definition of co-rotational coordinate system
coordinate system to the element so that the strain tensor in this local system is relevant for
LS-DYNA DEV 10/27/16 (r:8004) 14-7 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
the treatment. The co-rotational coordinate system determined here is one of the most
convenient ways to define such a local system.
∂𝐱 ∂𝑥 ∂𝑦 ∂𝑧
𝐠2 = =[ ] = [𝑁𝑎,𝜂 𝑥𝑎 𝑁𝑎,𝜂 𝑦𝑎 𝑁𝑎,𝜂 𝑧𝑎 ](𝜉,𝜂,0) . (16.65)
∂𝜂 ∂𝜂 ∂𝜂 ∂𝜂
The unit vector 𝐞̂1 of the co-rotational coordinate system is defined as the bisector of
the angle intersected by these two tangent vectors 𝐠1 and 𝐠2 ; the unit vector 𝐞̂3 is
perpendicular to the mid-surface and the other unit vector is determined by 𝐞̂1 and 𝐞̂3 , i.e.,
𝐠 𝐠 𝐠 𝐠
𝐞̂1 = ( 1 + 2 )⁄(∣ 1 + 2 ∣), (16.66)
∣𝐠1 ∣ ∣𝐠2 ∣ ∣𝐠1 ∣ ∣𝐠2 ∣
𝐠1 × 𝐠2
𝐞̂3 = , (16.67)
∣𝐠1 × 𝐠2 ∣
The rate of deformation (or velocity strain tensor), also defined in the co-rotational
coordinate system, is used as the measure of the strain rate,
T
1 ∂𝐯̂ def ∂𝐯̂ def ⎤
𝛆̇ = 𝐝̂ = ⎡
⎢ + ( ) ⎥, (16.70)
2 ∂𝐱̂ ∂𝐱̂
⎣ ⎦
where 𝐯̂ def is the deformation part of the velocity in the co-rotational system 𝐱̂. If the initial
strain 𝛆̂ (𝐗, 0) is given, the strain tensor can be expressed as,
14-8 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
𝑡
𝛆̂(X, 𝑡) = 𝛆̂(𝐗, 0) + ∫ 𝐝̂ (𝐗, 𝜏) 𝑑𝜏. (16.71)
0
The strain increment is then given by the mid-point integration of the velocity strain
tensor,
T
𝑡𝑛+1 1 ⎡∂Δ𝐮̂def ⎛∂Δ𝐮̂
⎜
def
⎞
⎟ ⎤
Δ𝛆̂ = ∫ ̂
𝐝𝑑𝜏 =̇ ⎢ ⎜
+⎜ ⎟ ⎥ (16.72)
𝑡𝑛
⎢
2 ∂𝐱̂ 1 ∂𝐱̂ 1 ⎟ ⎥,
𝑛+ ⎝ 𝑛+
⎣ 2 2⎠ ⎦
where Δ𝐮̂def is the deformation part of the displacement increment in the co-rotational
system 𝐱̂𝑛 + 1 referred to the mid-point configuration.
2
All the kinematical quantities must be computed from the last time step
configuration, Ω𝑛 , at 𝑡 = 𝑡𝑛 and the current configuration, Ω𝑛 + 1 at 𝑡 = 𝑡𝑛 + 1 since these
are the only available data. Denoting the spatial coordinates of these two configurations as
𝐱n and 𝐱n + 1 in the fixed global Cartesian coordinate system 𝑂x, as shown in Figure 16.2,
the coordinates in the corresponding co-rotational Cartesian coordinate systems, 𝑂𝐱̂𝑛 and
𝑂𝐱̂𝑛 + 1 , can be obtained by the following transformation rules:
𝐱̂𝑛 = 𝐑𝑛 𝐱𝑛 , (16.73)
LS-DYNA DEV 10/27/16 (r:8004) 14-9 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
Ωn+1
^
Xn+1
Ωn+1/2
Ωn ^
Xn+1/2
^
Xn
O
and the transformation to the co-rotational system associated with this mid-point
configuration, Ω𝑛 + 1 , is given by
2
𝐱̂ =𝐑 𝐱 . (16.76)
𝑛+1 𝑛+1 𝑛+1
2 2 2
In order to obtain the deformation part of the displacement increment referred to the
configuration at 𝑡 = 𝑡𝑛 + 1 , we need to find the rigid rotation from Ω𝑛 to Ω𝑛 + 1 provided that
2
14-10 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
the mid-point configuration, Ω𝑛 + 1 , is held still. Defining two virtual configurations, Ω′𝑛
2
and Ω′𝑛 + 1 , by rotating the element bodies Ω𝑛 and Ω𝑛 + 1 into the co-rotational system
̂
𝑂x̂𝑛 + 1 (Fig. 13.3) and denoting and 𝐱′ 𝑛 + 1 as the coordinates of Ω′𝑛 and Ω′𝑛 + 1 in the co-
2
rotational system 𝑂𝐱̂𝑛 + 1 , we have
2
̂ = 𝐱̂ ,
𝐱′ ̂
𝐱′ (16.78)
𝑛 𝑛 𝑛 + 1 = 𝐱̂𝑛 + 1 .
We can see that from Ω𝑛 to Ω′𝑛 and from Ω′𝑛 + 1 to Ω𝑛 + 1 , the body experiences two
rigid rotations and the rotation displacements are given by
T ̂ T
Δ𝐮rot
1 = 𝐱′𝑛 − 𝐱𝑛 = 𝐑 1 𝐱′𝑛 − 𝐱𝑛 = 𝐑 1 𝐱̂𝑛 − 𝐱𝑛 , (16.79)
𝑛+ 𝑛+
2 2
T ̂ T
Δ𝐮rot
2 = 𝐱𝑛+1 − 𝐱′𝑛+1 = 𝐱𝑛+1 − 𝐑 1 𝐱′𝑛+1 = 𝐱𝑛+1 − 𝐑 1 𝐱̂𝑛+1 . (16.80)
𝑛+ 𝑛+
2 2
Once the strain increment is obtained by equation (16.72), the stress increment, also
referred to the mid-point Configuration, can be calculated with the radial return algorithm.
The total strain and stress can then be updated as
𝛆̂𝑛+1 = 𝛆̂𝑛 + Δ𝛆̂, (16.84)
𝛔
̂𝑛+1 = 𝛔
̂𝑛 + Δ𝛔
̂. (16.85)
LS-DYNA DEV 10/27/16 (r:8004) 14-11 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
^
Ωn+1 Xn+1
Pn+1
,
Δu2rot Ωn+1
,
Pn+1
^
Δu Xn+1/2
Ωn+1/2
Δudef
,
Pn
,
Ωn Δu1rot
^
Ωn Xn
Pn
O
Note that the resultant stress and strain tensors are both referred to the current
configuration and defined in the current co-rotational coordinate system. By using the
tensor transformation rule we can have the strain and stress components in the global
coordinate system.
∫ 𝛿𝜀̂𝑣+1 𝑣+1
𝑖𝑗 𝜎̂ 𝑖𝑗
𝑣+1
𝑑𝑉 = 𝛿𝜋̂ext , (16.87)
̂ 𝑣+1
Ω
where 𝛿𝜋̂ext is the virtual work done by the external forces. Note that both equations are
written in the co-rotational coordinate system defined in the 𝑣th iterative configuration
𝑣
given by x𝑛+1 . The variables in this section are within the time step [𝑡𝑛 , 𝑡𝑛+1 ] and
2
superscripts indicate the number of iterations.
14-12 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
𝑣 𝑣
∫ 𝑣 𝛿𝑢̂𝑣𝑖,𝑗 𝐶𝑖𝑗𝑘𝑙
̂ Δ𝑢̂𝑘,𝑙 𝑑𝑉 + ∫ 𝛿𝑢̂𝑣𝑖,𝑗 𝑇̂𝑖𝑗𝑘𝑙 𝑣+1
Δ𝑢̂𝑘,𝑙 𝑑𝑉 = 𝛿𝜋̂ext 𝑣
− 𝛿𝜋̂ext , (16.88)
̂
Ω 𝑣 ̂
Ω
The first term on the left hand side of (16.88) denotes the material response since it is
due to pure deformation or stretching; the second term is an initial stress part resulting
from finite deformation effect.
Taking account of the residual of the previous iteration, Equation (16.87) can be
approximated as
𝑣 𝑣
∫ 𝑣 𝛿𝑢̂𝑣𝑖,𝑗 (𝐶𝑖𝑗𝑘𝑙
̂ + 𝑇̂𝑖𝑗𝑘𝑙 𝑣+1
)Δ𝑢̂𝑘,𝑙 𝑑𝑉 = 𝛿𝜋̂ext − ∫ 𝑣 𝛿𝜀̂𝑣𝑖𝑗 𝜎̂ 𝑖𝑗𝑣 𝑑𝑉 . (16.90)
̂
Ω ̂
Ω
where 𝐶𝑖𝑗𝑣̂ is the consistent tangent modulus tensor corresponding to pure deformation (see
Section 3.2.3) but expanded to a 9 by 9 matrix; 𝑇̂𝑖𝑗𝑣 is the geometric stiffness matrix which is
given as follows [Liu 1992]:
LS-DYNA DEV 10/27/16 (r:8004) 14-13 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
T =
𝜎4 𝜎6 𝜎4 𝜎6
⎡𝜎1 0 0
2
0
2 2
0 −
2 ⎤
⎢ 𝜎4 𝜎5 𝜎 𝜎5 ⎥
⎢ 𝜎2 0 0 − 4 0 ⎥
⎢ 2 2 2 2 ⎥
⎢ 𝜎5 𝜎6 𝜎 𝜎6 ⎥
⎢ 𝜎3 0 0 − 5 ⎥
⎢ 2 2 2 2 ⎥
⎢ 𝜎1 + 𝜎2 𝜎6 𝜎5 𝜎2 − 𝜎1 𝜎6 𝜎5 ⎥
⎢ − ⎥
⎢ 4 4 4 4 4 4 ⎥
⎢ 𝜎2 + 𝜎3 𝜎4 𝜎 𝜎3 − 𝜎2 𝜎4 ⎥ (16.94)
⎢ − 6 ⎥.
⎢ 4 4 4 4 4 ⎥
⎢ 𝜎1 + 𝜎3 𝜎5 𝜎 𝜎1 − 𝜎3 ⎥
⎢ − 4 ⎥
⎢ 4 4 4 4 ⎥
⎢ 𝜎1 + 𝜎2 𝜎6 𝜎5 ⎥
⎢ symm. − − ⎥
⎢ 4 4 4 ⎥
⎢ 𝜎2 + 𝜎3 𝜎 ⎥
⎢ − 4 ⎥
⎢ 4 4 ⎥
⎢ 𝜎3 + 𝜎1 ⎥
⎣ 4 ⎦
By interpolation
Δ𝐮 = 𝐍Δ𝐝, 𝛿𝐮 = 𝐍𝛿𝐝; (16.95)
̅̅̅̅̅Δ𝐝,
Δ𝛆 = 𝐁 𝛿𝛆 = ̅𝐁
̅̅̅̅𝛿𝐝, (16.96)
where 𝐍 and ̅𝐁 ̅̅̅̅ are, respectively, the shape functions and strain operators defined in
Section 2. This leads to a set of equations
̂ 𝑣 Δ𝐝̂ = 𝐫 𝑣+1
𝐊 ̂ ̂ − 𝐟int
𝑣+1
= 𝐟ext 𝑣̂
, (16.97)
𝑣̂
̂ 𝑣 , and the internal nodal force vector, 𝐟int
where the tangent stiffness matrix, 𝐊 , are
𝐊 ̂̅̅̅̅T (𝐂𝑣̂ + 𝐓
̂ 𝑣 = ∫ ̅𝐁 ̅̂̅̅̅̅dV,
̂ 𝑣 )𝐁 (16.98)
𝑣 ̂
Ω
𝑣̂
𝐟int ̂̅̅̅̅T 𝛔
= ∫ 𝑣̅𝐁 ̂𝑣 dV. (16.99)
̂
Ω
The tangent stiffness and nodal force are transformed into the global coordinate
system tensorially as
𝐊𝑣 = 𝐑𝑣T 𝐊
̂ 𝑣 𝐑𝑣 , (16.100)
𝑣
𝐫 𝑣+1 = 𝐑𝑣T 𝐫int
̂ , (16.101)
𝑣
where 𝐑𝑣 is the transformation matrix of the co-rotational system defined by 𝐱𝑛+1 . Finally,
we get a set of linear algebraic equations
14-14 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
𝑃 1
𝑢𝑦 (𝑥, 𝑦) = [3𝑣̅𝑦2 (𝐿 − 𝑥) + (4 + 5𝑣̅)𝐷2 𝑥 + (3𝐿 − 𝑥)𝑥2 ], (16.104)
̅̅̅̅𝐼
6𝐸 4
where
1 3
𝐼= 𝐷 , (16.105)
12
̅𝐸̅̅̅ = {𝐸, { 𝑣
⎧ for plane stress
𝑣̅ = ⎨ 𝑣 (16.106)
𝐸/(1 − 𝑣2 ) , {
⎩1 − 𝑣 for plane strain
The distribution of the applied load to the nodes at 𝑥 = 𝐿 is also obtained from the
closed-form stress fields.
LS-DYNA DEV 10/27/16 (r:8004) 14-15 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
M P x
P D
L
Figure 16.4. Timoshenko cantilever beam.
p/2
D/2 x
L Pt. A
y
θ
p/2
D/2 x
L/4 Pt. A
The parameters for the cantilever beam are: 𝐿 = 1.0, 𝐷 = 0.02, 𝑃 = 2.0, 𝐸 = 1 × 107 ;
and two values of Poisson’s ratio: (1)𝑣 = 0.25, (2)𝑣 = 0.4999.
Since the problem is anti-symmetric, only the top half of the beam is modeled. Plane
strain conditions are assumed in the z-direction and only one layer of elements is used in
this direction. Both regular mesh and skewed mesh are tested for this problem.
Normalized vertical displacements at point A for each case are given in Table 13.1.
Tables 13.1a and 13.1b show the normalized displacement at point A for the regular mesh.
14-16 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
There is no shear or volumetric locking for this element. For the skewed mesh, with the
skewed angle increased, we need more elements to get more accurate solution (Table
13.1c).
(a) 𝑣 = 0.25, regular mesh
Analytical solution 𝑤A = 9.3777 × 10−2
1 5 10
LS-DYNA DEV 10/27/16 (r:8004) 14-17 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
Due to symmetry only one octant of the cylinder is modeled. The computed
displacements at the loading point are compared to the analytic solutions in Table 13.2.
HEXDS element works well in both cases, indicating that this element can avoid not only
shear locking but also membrane locking; this is not unexpected since membrane locking
occurs primarily in curved elements [Stolarski, 1983].
p
R L
ric
et
m
P
m
symmetric
sy
t
Figure 16.6. Pinched cylinder and the element model
14-18 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
t
2L
R
θ
Mesh 10 10 2 16 16 4 20 20 4
Mesh 10 10 2 16 16 4 20 20 4
Due to symmetry only one quarter of the roof is modeled. The computed
displacement at the midpoint of the edge is compared to the analytic solution in Table 13.3.
In this example the HEXDS element can get good result with 100 × 2 elements.
LS-DYNA DEV 10/27/16 (r:8004) 14-19 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
R0=50.8mm
tight die
t=2mm r=2mm
Rd=54mm
R=54mm
Mesh 881 16 16 1 32 32 1 10 10 2
Because of the small corner radius of the die, the same difficulties as in the problem
of sheet stretch under the rigid cylinders lead the shell elements to failure in this problem.
Three-dimensional solid elements are needed and fine meshes should be put in the areas
near the center and the edge of the sheet.
One quarter of the sheet is modeled with 1400 2 HEXDS elements due to the
double symmetries. The mesh is shown in Fig. 16.9. Two layers of elements are used in the
thickness. Around the center and near the circular edge of the sheet, fine mesh is used.
The nodes on the edge are fixed in x- and y-directions and the bottom nodes on the edge
are prescribed in three directions. No friction is considered in this simulation. For
14-20 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Eight-Node Solid Element for Thick Shell Simulations
comparison, the axisymmetric four-node element with reduced integration (CAX4R) is also
used and the mesh for this element is the same as shown in the top of Figure 13.9.
The results presented here are after the punch has traveled down 50 mm. The
profile of the circular sheet is shown in Figure 16.10 where we can see that the sheet under
the punch experiences most of the stretching and the thickness of the sheet above the die
changes a lot. The deformation between the punch and the die is small. However, the
LS-DYNA DEV 10/27/16 (r:8004) 14-21 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
Figure 16.11. Reaction force vs. punch travel for the circular sheet
sheet thickness obtained by the CAX4R element is less than that by the HEXDS element
and there is slight difference above the die. These observations can be verified by the strain
distributions in the sheet along the radial direction (Figure 13.12). The direction of the
radial strain is the tangent of the mid-surface of the element in the rz plane and the
thickness strain is in the direction perpendicular to the mid-surface of the element. The
unit vector of the circumferential strain is defined as the cross-product of the directional
cosine vectors of the radial strain and the thickness strain. We can see that the CAX4R
element yields larger strain components in the area under the punch than the HEXDS
element. The main difference of the strain distributions in the region above the die is that
the CAX4R element gives zero circumferential strain in this area but the HEXDS element
yields non-zero strain. The value of the reaction force shown in the Figure 13.11 is only one
quarter of the total punch reaction force since only one quarter of the sheet is modeled.
From this figure we can see that the sheet begins softening after the punch travels down
about 45 mm, indicating that the sheet may have necking though this cannot be seen clearly
from Figure 16.10.
16.5 Conclusions
A new eight-node hexahedral element is implemented for the large deformation
elastic-plastic analysis. Formulated in the co-rotational coordinate system, this element is
shown to be effective and efficient and can achieve fast convergence in solving a wide
variety of nonlinear problems.
By using a co-rotational system which rotates with the element, the locking
phenomena can be suppressed by omitting certain terms in the generalized strain
operators. In addition, the integration of the constitutive equation in the co-rotational
LS-DYNA DEV 10/27/16 (r:8004) 14-23 (Eight-Node Solid Element for Thick Shell Simulations)
Eight-Node Solid Element for Thick Shell Simulations LS-DYNA Theory Manual
system takes the same simple form as small deformation theory since the stress and strain
tensors defined in this co-rotational system are objective.
Test problems studied in this paper demonstrate that the element is suitable to
continuum and structural numerical simulations. In metal sheet forming analysis, this
element has advantages over shell elements for certain problems where through the
thickness deformation and strains are significant.
14-24 (Eight-Node Solid Element for Thick Shell Simulations) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Truss Element
17
Truss Element
One of the simplest elements is the pin-jointed truss element shown in Figure 17.1.
This element has three degrees of freedom at each node and carries an axial force. The
displacements and velocities measured in the local system are interpolated along the axis
according to
𝑥
𝑢 = 𝑢1 + (𝑢2 − 𝑢1 ), (17.1)
𝐿
𝑥
𝑢̇ = 𝑢̇1 + (𝑢̇2 − 𝑢̇1 ), (17.2)
𝐿
where at 𝑥 = 0, 𝑢 = 𝑢1 and at 𝑥 = 𝐿, 𝑢 = 𝑢2 . Incremental strains are found from
(𝑢̇2 − 𝑢̇1 )
Δ𝜀 = Δ𝑡 (17.3)
𝐿
and are computed in LS-DYNA using
𝑛+1⁄2 𝑛+1⁄2
2 (𝑢̇2 − 𝑢̇1 )
𝑛+1⁄2 𝑛+1⁄2 (17.4)
Δ𝜀 = Δ𝑡
𝐿𝑛 + 𝐿𝑛+1
N1
u1
N2
u2
Figure 17.1. Truss element.
Two constitutive models are implemented for the truss element: elastic and elastic-
plastic with kinematic hardening.
18
Membrane Element
In membrane elements the rotational degrees of freedom at the nodal points may be
constrained, so that only the translational degrees-of-freedom contribute to the straining of
the membrane. A triangular membrane element may be obtained by collapsing adjacent
nodes of the quadrilateral.
The above velocity-strain relations are evaluated only at the center of the shell.
Standard bilinear nodal interpolation is used to define the mid-surface velocity, angular
velocity, and the element’s coordinates (isoparametric representation). These interpolation
relations are given by
𝑣𝑚 = 𝑁𝐼 (𝜉 , 𝜂)𝑣𝐼 , (18.2)
𝑥𝑚 = 𝑁𝐼 (𝜉 , 𝜂)𝑥𝐼 , (18.3)
where the subscript 𝐼 is summed over all the element’s nodes and the nodal velocities are
obtained by differentiating the nodal coordinates with respect to time, i.e., 𝜐𝐼 = ẋ𝐼 . The
bilinear shape functions are defined in Equations (7.10).
The velocity strains at the center of the element, i.e., at 𝜉 = 0, and 𝜂 = 0, are obtained
as in Section 7 giving:
𝑑𝑥̂ = 𝐵1𝐼 𝜐̂𝑥𝐼 , (18.4)
𝑅̂ = ℎ𝜎̂ 𝛼𝛽 ,
𝑓𝛼𝛽 (18.9)
where the superscript R indicates a resultant force and the Greek subscripts emphasize the
limited range of the indices for plane stress plasticity.
The above element centered force resultants are related to the local nodal forces by
𝑓𝑥𝐼̂ = 𝐴(𝐵1𝐼 𝑓𝑥𝑥
𝑅̂ + 𝐵2𝐼 𝑓𝑥𝑦
𝑅̂ ), (18.10)
The above local nodal forces are then transformed to the global coordinate system
using the transformation relations given in Equation (7.5a).
19
Discrete Elements and Masses
The discrete elements and masses in LS-DYNA provide a capability for modeling
simple spring-mass systems as well as the response of more complicated mechanisms.
Occasionally, the response of complicated mechanisms or materials needs to be included in
LS-DYNA models, e.g., energy absorbers used in passenger vehicle bumpers. These
mechanisms are often experimentally characterized in terms of force-displacement curves.
LS-DYNA provides a selection of discrete elements that can be used individually or in
combination to model complex force-displacement relations.
The discrete elements are assumed to be massless. However, to solve the equations
of motion at unconstrained discrete element nodes or nodes joining multiple discrete
elements, nodal masses must be specified at these nodes. LS-DYNA provides a direct
method for specifying these nodal masses in the model input.
All of the discrete elements are two-node elements, i.e., three-dimensional springs or
trusses. A discrete element may be attached to any of the other LS-DYNA continuum,
structural, or rigid body element. The force update for the discrete elements may be
written as
𝐟𝑖̂ +1 = 𝐟𝑖̂ + Δ𝐟,̂ (19.1)
where the superscript 𝑖 + 1 indicates the time increment and the superposed caret (⋅ ̂)
indicates the force in the local element coordinates, i.e., along the axis of the element. In the
default case, i.e., no orientation vector is used; the global components of the discrete
element force are obtained by using the element’s direction cosines:
⎧𝐹 ⎧Δ𝑙 ⎫ ⎧𝑛𝑥 ⎫
{ 𝑥⎫} 𝑓 ̂ { 𝑥} ̂{𝑛𝑦 } = 𝑓 𝐧̂ ,
⎨𝐹𝑦 ⎬ = ⎨Δ𝑙𝑦 ⎬ = 𝑓 ⎨ ⎬ (19.2)
{ 𝑙 { } {
⎩𝑛𝑧 }
~
⎩𝐹𝑧 }
⎭ ⎩Δ𝑙𝑧 ⎭ ⎭
where
16.1
Discrete Elements and Masses LS-DYNA Theory Manual
⎧ Δ𝑙
{ 𝑥⎫ } ⎧𝑥 −𝑥
{𝑦2 − 𝑦1 ⎫
}
Δ𝐥 = ⎨Δ𝑙𝑦 ⎬ = ⎨ 2 1 ⎬. (19.3)
{Δ𝑙 } {
⎩𝑧2 − 𝑧1 }
⎭
⎩ 𝑧⎭
𝑙 is the length
and (𝑥𝑖 , 𝑦𝑖 , 𝑧𝑖 ) are the global coordinates of the nodes of the spring element. The forces in
Equation (19.2) are added to the first node and subtracted from the second node.
For a node tied to ground we use the same approach but for the (𝑥2 , 𝑦2 , 𝑧2 )
coordinates in Equation (19.2) the initial coordinates of node 1, i.e., (𝑥0 , 𝑦0 , 𝑧0 ) are used
instead; therefore,
⎧ 𝐹 𝑥 −𝑥 𝑛𝑥 ⎫
{ 𝑥⎫} 𝑓̂⎧ {𝑦0 − 𝑦1 ⎫
} ⎧
̂ 𝑛𝑦 }.
{
⎨ 𝐹𝑦⎬ = ⎨ 0 1⎬ = 𝑓 ⎨ ⎬ (19.5)
{ 𝑙 {
⎩𝑧0 − 𝑧1 } {
⎩𝑛𝑧 }
⎩𝐹𝑧 }
⎭ ⎭ ⎭
The increment in the element force is determined from the user specified force-
displacement relation. Currently, nine types of force-displacement/velocity relationships
may be specified:
1. linear elastic;
2. linear viscous;
3. nonlinear elastic;
4. nonlinear viscous;
5. elasto-plastic with isotropic hardening;
6. general nonlinear;
7. linear viscoelastic.
8. inelastic tension and compression only.
9. muscle model.
The force-displacement relations for these models are discussed in the following
later.
We will first consider the portion of the displacement that lies in the direction of the
vector. The displacement of the spring is updated based on the change of length given by
Δ𝐼 = 𝐼 − 𝐼0 , (19.7)
where 𝐼0 is the initial length in the direction of the vector and lis the current length given
for a node to node spring by
𝐼 = 𝑚1 (𝑥2 − 𝑥1 ) + 𝑚2 (𝑦2 − 𝑦1 ) + 𝑚3 (𝑧2 − 𝑧1 ), (19.8)
and for a node to ground spring by
𝐼 = 𝑚1 (𝑥0 − 𝑥1 ) + 𝑚2 (𝑦0 − 𝑦1 ) + 𝑚3 (𝑧0 − 𝑧1 ), (19.9)
The latter case is not intuitively obvious and can affect the sign of the force in unexpected
ways if the user is not familiar with the relevant equations. The nodal forces are then given
by
⎧ 𝐹 ⎧𝑚1 ⎫
{ 𝑥⎫ } ̂{𝑚2 }.
⎨ 𝐹𝑦 ⎬ = 𝑓 ⎨ ⎬ (19.10)
{ {
⎩𝑚3 }
⎩𝐹𝑧 }⎭ ⎭
The orientation vector can be either permanently fixed in space as defined in the
input or acting in a direction determined by two moving nodes which must not be
coincident but may be independent of the nodes of the spring. In the latter case, we
recompute the direction every cycle according to:
⎧ 𝑚 ⎧ 𝑥𝑛 − 𝑥1𝑛 ⎫
{𝑚 1 ⎫
} 1 { 2𝑛 }
⎨ 2⎬ = 𝑛 ⎨ 𝑦2 − 𝑦1𝑛 ⎬. (19.11)
{
⎩𝑚 3 }
⎭ 𝑙 {
⎩𝑧𝑛2 − 𝑧𝑛1 }
⎭
In Equation (19.9) the superscript, 𝑛, refers to the orientation nodes.
For the case where we consider motion in the plane perpendicular to the orientation
vector we consider only the displacements in the plane, Δ𝑙𝑝 , given by,
Δ𝐥𝑝 = Δ𝐥 − 𝐦(𝐦 ⋅ Δ𝐥). (19.12)
We update the displacement of the spring based on the change of length in the plane given
by
𝑝
Δ𝑙𝑝 = 𝑙𝑝 − 𝑙0 , (19.13)
𝑝
where 𝑙0 is the initial length in the direction of the vector and 𝑙 is the current length given
for a node to node spring by
16.3
Discrete Elements and Masses LS-DYNA Theory Manual
𝑝 𝑝 𝑝
𝑙𝑝 = 𝑚1 (𝑥2 − 𝑥1 ) + 𝑚2 (𝑦2 − 𝑦1 ) + 𝑚3 (𝑧2 − 𝑧1 ), (19.14)
and for a node to ground spring by
𝑝 𝑝 𝑝
𝑙𝑝 = 𝑚1 (𝑥0 − 𝑥1 ) + 𝑚2 (𝑦0 − 𝑦1 ) + 𝑚3 (𝑧0 − 𝑧1 ), (19.15)
where
⎧𝑚𝑝1 ⎫ ⎧Δ𝑙 𝑝
⎫
{
{ 𝑝} } 1 {
{ 𝑥𝑝 }}
⎨𝑚 2 ⎬ = ⎨Δ𝑙𝑦 ⎬. (19.16)
{ 𝑝}
{ } 𝑝2 𝑝2 { 𝑝}
𝑝2 { }
⎩𝑚3 ⎭ √Δ𝑙𝑥 + Δ𝑙𝑦 + Δ𝑙𝑧 ⎩Δ𝑙𝑧 ⎭
After computing the displacements, the nodal forces are then given by
⎧ 𝐹 ⎧𝑚𝑝1 ⎫
{ 𝑥⎫ } {
{ }
̂ 𝑚𝑝 .
}
⎨ 𝐹𝑦 ⎬ = 𝑓 ⎨ 2⎬ (19.17)
{
⎩𝐹𝑧 }⎭ { 𝑝}
{ }
⎩𝑚3 ⎭
For example, if it is known that a component shows a dynamic crush force at 15 m/s
equal to 2.5 times the static crush force, use 𝑘𝑑 = 1.5 and 𝑉0 = 15.
16.4
LS-DYNA Theory Manual Discrete Elements and Masses
If the spring element is initially of zero length and if no orientation vectors are used
then only the tensile part of the stress strain curve needs to be defined. However, if the
16.5
Discrete Elements and Masses LS-DYNA Theory Manual
Nolinear Elastic/Viscous
Force
Displacement/Velocity
Figure 19.1. Piecewise linear force-displacement curve for nonlinear elastic
discrete element.
spring element is initially of finite length then the curve must be defined in both the
positive and negative quadrants.
16.6
LS-DYNA Theory Manual Discrete Elements and Masses
Elsto-Plastic with
Isotropic Hardening
ET
FY
Force
E
Elasto-Plastic Unloading
Displacement
Figure 19.2. Loading and unloading force-displacement curves for elasto-plastic
discrete element.
This element combines the features of the above-described nonlinear elastic and
elasto-plastic discrete elements by allowing the user to specify independent initial yield
forces in tension (FYT) and in compression (FYC). If the discrete element force remains
between these initial yield values, the element unloads along the loading path (Figure
19.3(b)). This corresponds to the nonlinear elastic discrete element.
However, if the discrete element force exceeds either of these initial yield values, the
specified unloading curve is used for subsequent unloading. Additionally, the initial
loading and unloading curves are allowed to move in the force-displacement space by
specifying a mixed hardening parameter 𝛽 where 𝛽 = 0 corresponds to kinematic
hardening (Figure 19.3(c)) and 𝛽 = 0 𝛽 = 1 corresponds to isotropic hardening (Figure
19.3(d)).
16.7
Discrete Elements and Masses LS-DYNA Theory Manual
loading
β>0 β>0
curve
options
β=0 Fyt β=0
Fyt - Fyc
δ δ
Fyc
unloading curve
force force
F2
Fyt-Fyc F1
δ δ
F2
Figure 19.3. Loading and unloading force displacement curves for general
nonlinear discrete element.
where 𝐾∞ is the long duration stiffness, 𝐾0 is the short time stiffness, and 𝛽 is a decay
parameter that controls the rate at which the stiffness transitions between the short and
long duration stiffness (Figure 16.4).
This model corresponds to a three-parameter Maxwell model (see insert in
Figure 19.4) which consists of a spring and damper in series connected to another spring in
parallel. Although this discrete element behavior could be built up using the above-
described linear elastic and linear viscous discrete elements, such a model would also
require the user to specify the nodal mass at the connection of the series spring and
damper. This mass introduces a fourth parameter which would further complicate fitting
the model to experimental data.
16.8
LS-DYNA Theory Manual Discrete Elements and Masses
Log K0
K∞ Visco-Elastic
Log K(t)
K0-K∞
1/β
Log K∞
Log t
Figure 19.4. Typical stiffness relaxation curve used for the viscoelastic discrete
element.
16.9
Discrete Elements and Masses LS-DYNA Theory Manual
a(t) LM
vM
FCE CE
M SEE
F FM
FPE
LM PE
Figure 19.5. Discrete model for muscle contraction dynamics, based on a Hill-
type representation. The total force is the sum of passive force FPE and active
force FCE. The passive element (PE) represents energy storage from muscle
elasticity, while the contractile element (CE) represents force generation by the
muscle. The series elastic element (SEE), shown in dashed lines, is often neglected
when a series tendon compliance is included. Here, a(t) is the activation level, LM
is the length of the muscle, and vM is the shortening velocity of the muscle.
The material behavior of the muscle model is adapted from the original model
proposed by Hill (1938). Reviews of this model and extensions can be found in Winters
(1990) and Zajac (1989). The most basic Hill-type muscle model consists of a contractile
element (CE) and a parallel elastic element (PE) (Figure 19.5). An additional series elastic
element (SEE) can be added to represent tendon compliance. The main assumptions of the
Hill model are that the contractile element is entirely stress free and freely distensible in the
resting state, and is described exactly by Hill’s equation (or some variation). When the
muscle is activated, the series and parallel elements are elastic, and the whole muscle is a
simple combination of identical sarcomeres in series and parallel. The main criticism of
Hill’s model is that the division of forces between the parallel elements and the division of
extensions between the series elements is arbitrary, and cannot be made without
introducing auxiliary hypotheses. However, these criticisms apply to any discrete element
model. Despite these limitations, the Hill model has become extremely useful for modeling
musculoskeletal dynamics, as illustrated by its widespread use today.
When the contractile element (CE) of the Hill model is inactive, the entire resistance
to elongation is provided by the PE element and the tendon load-elongation behavior. As
activation is increased, force then passes through the CE side of the parallel Hill model,
providing the contractile dynamics. The original Hill model accommodated only full
16.10
LS-DYNA Theory Manual Discrete Elements and Masses
Without the SEE, the total force in the muscle FM is the sum of the force in the CE
and the PE because they are in parallel:
𝐹M = 𝐹PE + 𝐹CE . (19.26)
The relationships defining the force generated by the CE and PE as a function of LM, VM
and 𝑎(𝑡) are often scaled by 𝐹max , the peak isometric force (p. 80, Winters 1990), L0, the
initial length of the muscle (p. 81, Winters 1990), and 𝑉max , the maximum unloaded CE
shortening velocity (p. 80, Winters 1990). From these, dimensionless length and velocity
can be defined:
𝐿M
𝐿= ,
𝐿o
(19.27)
𝑉M
𝑉= .
𝑉max ∗ 𝑆V (𝑎(t))
Here, 𝑆V scales the maximum CE shortening velocity 𝑉max and changes with activation
level 𝑎(𝑡). This has been suggested by several researchers, i.e. Winters and Stark [1985].
The activation level specifies the level of muscle stimulation as a function of time. Both
have values between 0 and 1. The functions 𝑆V (𝑎(𝑡)) and 𝑎(𝑡) are specified via load curves
in LS-DYNA, or default values of 𝑆V = 1 and 𝑎(𝑡) = 0 are used. Note that L is always
positive and that 𝑉 is positive for lengthening and negative for shortening.
The relationship between FCE, V and L was proposed by Bahler et al. [1967]. A
three-dimensional relationship between these quantities is now considered standard for
computer implementations of Hill-type muscle models [i.e., eqn 5.16, p. 81, Winters 1990].
It can be written in dimensionless form as:
𝐹CE = 𝑎(𝑡) ∗ 𝐹max ∗ 𝑓TL (𝐿) ∗ 𝑓TV (𝑉), (19.28)
The force in the parallel elastic element FPE is determined directly from the current length
of the muscle using an exponential relationship [eqn 5.5, p. 73, Winters 1990]:
𝐹PE
𝑓PE = =0 𝐿≤1
𝐹MAX
(19.29)
𝐹PE 1 𝐾
𝑓PE = = [exp ⎛
⎜ sh (L − 1)⎞
⎟ − 1] 𝐿>1
𝐹MAX exp(𝐾sh ) − 1 ⎝𝐿max ⎠
For computation of the total force developed in the muscle FM, the functions for the
tension-length 𝑓TL and force-velocity 𝑓TV relationships used in the Hill element must be
16.11
Discrete Elements and Masses LS-DYNA Theory Manual
defined. These relationships have been available for over 50 years, but have been refined to
allow for behavior such as active lengthening. The active tension-length curve 𝑓TL describes
the fact that isometric muscle force development is a function of length, with the maximum
force occurring at an optimal length. According to Winters, this optimal length is typically
around 𝐿 = 1.05, and the force drops off for shorter or longer lengths, approaching zero
force for 𝐿 = 0.4 and 𝐿 = 1.5. Thus the curve has a bell-shape. Because of the variability in
this curve between muscles, the user must specify the function 𝑓TL via a load curve,
specifying pairs of points representing the normalized force (with values between 0 and 1)
and normalized length 𝐿 (Figure 19.6).
The active tension-velocity relationship 𝑓TV used in the muscle model is mainly due
to the original work of Hill. Note that the dimensionless velocity V is used. When V = 0,
the normalized tension is typically chosen to have a value of 1.0. When V is greater than or
equal to 0, muscle lengthening occurs. As V increases, the function is typically designed so
that the force increases from a value of 1.0 and asymptotes towards a value near 1.4. When
V is less than zero, muscle shortening occurs and the classic Hill equation hyperbola is
used to drop the normalized tension to 0 (Figure 16.6). The user must specify the function
𝑓TV via a load curve, specifying pairs of points representing the normalized tension (with
values between 0 and 1) and normalized velocity V.
16.12
LS-DYNA Theory Manual Discrete Elements and Masses
set of belt elements. The user enters a load curve for loading, the points of which are
(Strain, Force). Strain is defined as engineering strain, i.e.
current length
Strain = − 1. (19.30)
initial length
In addition, the magnitude of the damping forces is limited to one tenth of the force
calculated from the forces-strain relationship and is zero when the belt is slack. Damping
forces are not applied to elements attached to sliprings and retractors.
The user inputs a mass per unit length that is used to calculate nodal masses on
initialization.
A ‘minimum length’ is also input. This controls the shortest length allowed in any
element and determines when an element passes through sliprings or are absorbed into the
retractors. One tenth of a typical initial element length is usually a good choice.
16.13
Discrete Elements and Masses LS-DYNA Theory Manual
positive (i.e., the current length is greater then the unstretched length), a tension force is
calculated from the material characteristics and is applied along the current axis of the
element to oppose further stretching. The unstretched length of the belt is taken as the
initial distance between the two nodes defining the position of the element plus the initial
slack length. At the beginning of the calculation the seatbelt elements can be obtained
within a retractor.
19.12 Sliprings
Sliprings are defined in the LS-DYNA input by giving a slipring ID and element ID’s
for two elements who share a node which is coincident with the slipring node. The slipring
node may not be attached to any belt elements.
Sliprings allow continuous sliding of a belt through a sharp change of angle. Two
elements (1 and 2 in Figure 19.7) meet at the slipring. Node B in the belt material remains
attached to the slipring node, but belt material (in the form of unstretched length) is passed
from element 1 to element 2 to achieve slip. The amount of slip at each timestep is
calculated from the ratio of forces in elements 1 and 2. The ratio of forces is determined by
the relative angle between elements 1 and 2 and the coefficient of friction, 𝜇. The tension in
the belts is taken as T1 and T2, where T2 is on the high-tension side and T1 is the force on
the low-tension side. Thus if T2 is sufficiently close to T1 no slip occurs; otherwise, slip is
just sufficient to reduce the ratio T2⁄T1 to 𝑒𝜇𝜃 . No slip occurs if both elements are slack.
The out-of-balance force at node B is reacted on the slipring node; the motion of node B
follows that of slipring node.
If, due to slip through the slipring, the unstretched length of an element becomes
less than the minimum length (as entered on the belt material card), the belt is remeshed
locally: the short element passes through the slipring and reappears on the other side (see
Figure 19.7). The new unstretched length of e1 is 1.1 × minimum length. Force and strain
in e2 and e3 are unchanged; force and strain in e1 are now equal to those in e2. Subsequent
slip will pass material from e3 to e1. This process can continue with several elements
passing in turn through the slipring.
To define a slipring, the user identifies the two belt elements which meet at the
slipring, the friction coefficient, and the slipring node. The two elements must have a
common node coincident with the slipring node. No attempt should be made to restrain or
constrain the common node for its motion will automatically be constrained to follow the
slipring node. Typically, the slipring node is part of the vehicle body structure and,
therefore, belt elements should not be connected to this node directly, but any other feature
can be attached, including rigid bodies.
16.14
LS-DYNA Theory Manual Discrete Elements and Masses
Slip ring
B B
Element 2 Element 1
Element 1 Element 3
Element 2
Element 3
Before After
Figure 19.7. Elements passing through slipring.
19.13 Retractors
Retractors are defined by giving a node, the “retractor node” and an element ID of
an element outside the retractor but with one node that is coincident with the retractor
node. Also sensor ID’s must be defined for up to four sensors which can activate the
seatbelt.
Retractors allow belt material to be paid out into a belt element, and they operate in
one of two regimes: unlocked when the belt material is paid out or reeled in under constant
tension and locked when a user defined force-pullout relationship applies.
The retractor is initially unlocked, and the following sequence of events must occur
for it to become locked:
• Any one of up to four sensors must be triggered. (The sensors are described below).
• Then a user-defined time delay occurs.
• Then a user-defined length of belt must be payed out (optional).
• Then the retractor locks.
and once locked, it remains locked.
In the unlocked regime, the retractor attempts to apply a constant tension to the belt.
This feature allows an initial tightening of the belt, and takes up any slack whenever it
occurs. The tension value is taken from the first point on the force-pullout load curve. The
16.15
Discrete Elements and Masses LS-DYNA Theory Manual
maximum rate of pull out or pull in is given by 0.01 × fed length per time step. Because of
this, the constant tension value is not always achieved.
In the locked regime, a user-defined curve describes the relationship between the
force in the attached element and the amount of belt material paid out. If the tension in the
belt subsequently relaxes, a different user-defined curve applies for unloading. The
unloading curve is followed until the minimum tension is reached.
The curves are defined in terms of initial length of belt. For example, if a belt is
marked at 10mm intervals and then wound onto a retractor, and the force required to make
each mark emerge from the (locked) retractor is recorded, the curves used for input would
be as follows:
Pyrotechnic pretensions may be defined which cause the retractor to pull in the belt
at a predetermined rate. This overrides the retractor force-pullout relationship from the
moment when the pretensioner activates.
If desired, belt elements may be defined which are initially inside the retractor.
These will emerge as belt material is paid out, and may return into the retractor if sufficient
material is reeled in during unloading.
Elements e2, e3 and e4 are initially inside the retractor, which is paying out material
into element e1. When the retractor has fed Lcrit into e1, where:
Lcrit = fed length − 1.1 × minimum length (19.32)
Here, minimum length is defined on belt material input, and fed length is defined on
retractor input.
element e2 emerges with an unstretched length of 1.1 × minimum length; the unstretched
length of element e1 is reduced by the same amount. The force and strain in e1 are
unchanged; in e2, they are set equal to those in e1. The retractor now pays out material into
e2.
If no elements are inside the retractor, e2 can continue to extend as more material is
fed into it.
16.16
LS-DYNA Theory Manual Discrete Elements and Masses
Element 1
Before Element 1
Element 2
Element 3 Element 2
After
Element 4
Element 3
Element 4
As the retractor pulls in the belt (for example, during initial tightening), if the
unstretched length of the mouth element becomes less than the minimum length, the
element is taken into the retractor.
To define a retractor, the user enters the retractor node, the ‘mouth’ element (into
which belt material will be fed, e1 in Figure 19.8, up to 4 sensors which can trigger
unlocking, a time delay, a payout delay (optional), load and unload curve numbers, and
the fed length. The retractor node is typically part of the vehicle stricture; belt elements
should not be connected to this node directly, but any other feature can be attached
including rigid bodies. The mouth element should have a node coincident with the
retractor but should not be inside the retractor. The fed length would typically be set either
to a typical element initial length, for the distance between painted marks on a real belt for
comparisons with high-speed film. The fed length should be at least three times the
minimum length.
If there are elements initially inside the retractor (e2, e3 and e4 in the Figure) they
should not be referred to on the retractor input, but the retractor should be identified on
the element input for these elements. Their nodes should all be coincident with the
retractor node and should not be restrained or constrained. Initial slack will automatically
be set to 1.1 × minimum length for these elements; this overrides any user-defined value.
19.14 Sensors
Sensors are used to trigger locking of retractors and activate pretensioners. Four
types of sensor are available which trigger according to the following criteria:
Type 1–When the magnitude of x-, y-, or z- acceleration of a given node has remained
above a given level continuously for a given time, the sensor triggers. This does not work
with nodes on rigid bodies.
Type 2–When the rate of belt payout from a given retractor has remained above a given
level continuously for a given time, the sensor triggers.
Type 4–The sensor triggers when the distance between two nodes exceeds a given
maximum or becomes less than a given minimum. This type of sensor is intended for use
with an explicit mas/spring representation of the sensor mechanism.
By default, the sensors are inactive during dynamic relaxation. This allows initial
tightening of the belt and positioning of the occupant on the seat without locking the
retractor or firing any pretensioners. However, a flag can be set in the sensor input to make
the sensors active during the dynamic relaxation phase.
with weblockers
without weblockers
Force
Pullout
Figure 19.9. Retractor force pull characteristics.
16.18
LS-DYNA Theory Manual Discrete Elements and Masses
19.15 Pretensioners
Pretensioners allow modeling of three types of active devices which tighten the belt
during the initial stages of a crash. The first type represents a pyrotechnic device which
spins the spool of a retractor, causing the belt to be reeled in. The user defines a pull-in
versus time curve which applies once the pretensioner activates. The remaining types
represents preloaded springs or torsion bars which move the buckle when released. The
pretensioner is associated with any type of spring element including rotational. Note that
the preloaded spring, locking spring and any restraints on the motion of the associated
nodes are defined in the normal way; the action of the pretensioner is merely to cancel the
force in one spring until (or after) it fires. With the second type, the force in the spring
element is cancelled out until the pretensioner is activated. In this case the spring in
question is normally a stiff, linear spring which acts as a locking mechanism, preventing
motion of the seat belt buckle relative to the vehicle. A preloaded spring is defined in
parallel with the locking spring. This type avoids the problem of the buckle being free to
‘drift’ before the pretensioner is activated.
19.16 Accelerometers
The accelerometer is defined by three nodes in a rigid body which defines a triad to
measure the accelerations in a local system. The presence of the accelerometer means that
the accelerations and velocities of node 1 will be output to all output files in local instead of
global coordinates.
The three nodes should all be part of the same rigid body. The local axis then rotates
with the body.
16.19
LS-DYNA Theory Manual Simplified Arbitrary Lagrangian-Eulerian
20
Simplified Arbitrary Lagrangian-
Eulerian
3. Remapping the solution from the distorted mesh to the smooth mesh.
specify the materials in each element and the data must be updated by the remap
algorithms.
The range of problems that can be solved with an ALE formulation is a direct
function of the sophistication of the algorithms for smoothing the mesh. Early ALE codes
were not very successful largely because of their primitive algorithms for smoothing the
mesh. In simplified ALE formulations, most of the difficulties with the mesh are associated
with the nodes on the material boundaries. If the material boundaries are purely
Lagrangian, i.e., the boundary nodes move with the material at all times, no smooth mesh
maybe possible and the calculation will terminate. The algorithms for maintaining a
smooth boundary mesh are therefore as important to the robustness of the calculations as
the algorithms for the mesh interior.
The cost of the advection step per element is usually much larger than the cost of the
Lagrangian step. Most of the time in the advection step is spent in calculating the material
transported between the adjacent elements, and only a small part of it is spent on
calculating how and where the mesh should be adjusted. Second order accurate monotonic
advection algorithms are used in LS-DYNA despite their high cost per element because
their superior coarse mesh accuracy which allows the calculation to be performed with far
fewer elements than would be possible with a cheaper first order accurate algorithm.
The second order transport accuracy is important since errors in the transport
calculations generally smooth out the solution and reduce the peak values in the history
variables. Monotonic advection algorithms are constructed to prevent the transport
calculations from creating new minimum or maximum values for the solution variables.
They were first developed for the solution of the Navier Stokes equations to eliminate the
spurious oscillations that appeared around the shock fronts. Although monotonic
algorithms are more diffusive than algorithms that are not monotonic, they must be used
for stability in general purpose codes. Many constitutive models have history variables
that have limited ranges, and if their values are allowed to fall outside of their allowable
ranges, the constitutive models are undefined. Examples include explosive models, which
require the burn fraction to be between zero and one, and many elastoplasticity models,
such as those with power law hardening, which require a non-negative plastic strain.
Each element solution variable must be transported. The total number of solution
variables, including the velocity, is at least six and depends on the material models. For
elements that are modeled with an equation of state, only the density, the internal energy,
and the shock viscosity are transported. When the elements have strength, the six compo-
nents of the stress tensor and the plastic strain must also be advected, for a total of ten
solution variables. Kinematic hardening, if it is used, introduces another five solution
variables, for a total of fifteen.
The nodal velocities add an extra three solution variables that must be transported,
and they must be advected separately from the other solution variables because they are
centered at the nodes and not in the elements. In addition, the momentum must be
conserved, and it is a product of the node-centered velocity and the element-centered
density. This imposes a constraint on how the momentum transport is performed that is
unique to the velocity field. A detailed consideration of the difficulties associated with the
transport of momentum is deferred until later.
Perhaps the simplest strategy for minimizing the cost of the ALE calculations is to
perform them only every few time steps. The cost of an advection step is typically two to
five times the cost of the Lagrangian time step. By performing the advection step only
every ten steps, the cost of an ALE calculation can often be reduced by a factor of three
without adversely affecting the time step size. In general, it is not worthwhile to advect an
element unless at least twenty percent of its volume will be transported because the gain in
the time step size will not offset the cost of the advection calculations.
∇2 𝜂 = 0. (1.1.1b)
We solve Equations (1.1.1b) for the coordinates 𝑥(𝜉 , 𝜂) and 𝑦(𝜉 , 𝜂) of the mesh lines:
that is, we invert them so that the geometric coordinates 𝑥, 𝑦 become the dependent
variables and the curvilinear coordinates 𝜉 , 𝜂 the independent variables. By the usual
methods of changing variables we obtain
𝛼𝑥𝜉𝜉 − 2𝛽𝑥𝜉𝜂 + 𝛾𝑥𝜂𝜂 = 0, (20.1.2a)
where a variable subscript indicates differentiation with respect to that variable. Since the
curvilinear coordinates are each assumed to satisfy Laplace’s equation, the second
summation in Equation (20.1.6) vanishes and we have
3
∇2 𝐴 = ∑ 𝑔𝑖𝑗 𝐴𝜉 𝑖 𝜉 𝑗 . (20.1.7)
𝑖,𝑗=1
Also,
𝑔 ≡ det(𝑔𝑖𝑗 ) = [𝐚1 ⋅ (𝐚2 × 𝐚3 )]2 = 𝐽 2 , (20.1.13)
where 𝑔𝑖𝑗 is the covariant metric tensor given by
𝑔𝑖𝑗 ≡ 𝐚𝑖 ⋅ 𝐚𝑗 , (20.1.14)
and 𝐽 is the Jacobian of the transformation.
𝑔𝑔𝑖𝑗 = (𝐚𝑖 ⋅ 𝐚𝑘 )(𝐚𝑗 ⋅ 𝐚𝑘 ) − (𝐚𝑖 ⋅ 𝐚𝑗 )𝐚𝑘2 = 𝑔𝑖𝑘 𝑔𝑗𝑘 − 𝑔𝑖𝑗 𝑔𝑘𝑘 . (20.1.17)
Before substituting (20.1.11) into (17.1.13a, b), we return to our original notation:
𝜉 + 𝜉 1, 𝜂 + 𝜉 2, 𝜁 + 𝜉 3. (20.1.18)
Then, using (20.1.11), we get
2
𝑔𝑔11 = 𝐫𝜂2 𝐫𝜁2 − (𝐫𝜂 ⋅ 𝐫𝜁 ) , (20.1.19)
2
𝑔𝑔22 = 𝐫𝜁2 𝐫𝜉2 − (𝐫𝜁 ⋅ 𝐫𝜉 ) , (20.1.20)
2
𝑔𝑔33 = 𝐫𝜉2 𝐫𝜂2 − (𝐫𝜉 ⋅ 𝐫𝜂 ) , (20.1.21)
for the three diagonal components, and
𝑔𝑔12 = (𝐫𝜉 ⋅ 𝐫𝜁 )(𝐫𝜂 ⋅ 𝐫𝜁 ) − (𝐫𝜉 ⋅ 𝐫𝜂 )𝐫𝜁2 , (20.1.22)
To check that these equations reproduce the two-dimensional equations when there
is no variation in one-dimension, we take 𝜁 as the invariant direction, thus reducing
(17.1.19) to
𝑔𝑔11 𝐫𝜉𝜉 + 2𝑔𝑔12 𝐫𝜉𝜂 + 𝑔𝑔22 𝐫𝜂𝜂 = 0. (20.1.33)
If we let 𝜁 = 𝑧, then the covariant base vectors become
𝐚1 = 𝑥𝜉 𝐢 + 𝑦𝜉 𝐣, (20.1.34)
𝐚2 = 𝑥𝜂 𝐢 + 𝑦𝜂 𝐣, (20.1.35)
𝐚3 = 𝐤. (20.1.36)
From (17.1.22), using (17.1.13), we get
𝑔𝑔11 = 𝑥𝜂2 + 𝑦𝜂2 , (20.1.37)
Before differencing Equations (17.1.19) we simplify the notation and write them in
the form
ζ η
ξ
Figure 20.1. Example Caption
𝛼1 𝐫𝜉𝜉 + 𝛼2 𝐫𝜂𝜂 + 𝛼3 𝐫𝜁𝜁 + 2𝛽1 𝐫𝜉𝜂 + 2𝛽2 𝐫𝜂𝜁 + 2𝛽3 𝐫𝜁𝜉 = 0, (20.1.40)
where
𝛼1 = (𝑥𝜂 𝑦𝜁 − 𝑥𝜁 𝑦𝜂 )2 + (𝑥𝜂 𝑧𝜁 − 𝑥𝜁 𝑧𝜂 )2 + (𝑦𝜂 𝑧𝜁 − 𝑦𝜁 𝑧𝜂 )2 , (20.1.41)
𝛽1 = (𝑥𝜉 𝑥𝜁 + 𝑦𝜉 𝑦𝜁 + 𝑧𝜉 𝑧𝜁 )(𝑥𝜂 𝑥𝜁 + 𝑦𝜂 𝑦𝜁 + 𝑧𝜂 𝑧𝜁 )
(20.1.44)
−(𝑥𝜉 𝑥𝜂 + 𝑦𝜉 𝑦𝜂 + 𝑧𝜉 𝑧𝜂 )(𝑥𝜁2 + 𝑦𝜁2 + 𝑧2𝜁 ),
𝛽2 = (𝑥𝜂 𝑥𝜉 + 𝑦𝜂 𝑦𝜉 + 𝑧𝜂 𝑧𝜉 )(𝑥𝜉 𝑥𝜁 + 𝑦𝜉 𝑦𝜁 + 𝑧𝜉 𝑧𝜁 )
(20.1.45)
−(𝑥𝜁 𝑥𝜂 + 𝑦𝜁 𝑦𝜂 + 𝑧𝜁 𝑧𝜂 )(𝑥𝜉2 + 𝑦𝜉2 + 𝑧2𝜉 ),
𝛽3 = (𝑥𝜂 𝑥𝜁 + 𝑦𝜂 𝑦𝜁 + 𝑧𝜂 𝑧𝜁 )(𝑥𝜉 𝑥𝜂 + 𝑦𝜉 𝑦𝜂 + 𝑧𝜉 𝑧𝜂 )
(20.1.46)
−(𝑥𝜉 𝑥𝜂 + 𝑦𝜉 𝑦𝜂 + 𝑧𝜉 𝑧𝜂 )(𝑥𝜂2 + 𝑦𝜂2 + 𝑧2𝜂 ),
1
𝐫𝜉𝜂 = [(𝐫𝑖+1,𝑗+1 + 𝐫𝑖−1,𝑗−1 ) − (𝐫𝑖+1,𝑗−1 + 𝐫𝑖−1,𝑗+1 )], (20.1.53)
4
1
𝐫𝜂𝜁 = [(𝐫𝑗+1,𝑘+1 + 𝐫𝑗−1,𝑘−1 ) − (𝐫𝑗+1,𝑘−1 + 𝐫𝑗−1,𝑘+1 )], (20.1.54)
4
1
𝐫𝜁𝜉 = [(𝐫𝐼+1,𝑘+1 + 𝐫𝐼−1,𝑘−1 ) − (𝐫𝐼+1,𝑘−1 + 𝐫𝐼−1,𝑘+1 )], (20.1.55)
4
where for brevity we have omitted subscripts 𝑖, 𝑗, or 𝑘 (e.g., 𝑘 + 1 stands for 𝑖, 𝑗, 𝑘 + 1).
Note that these difference expressions use only coordinate planes that pass through the
central point, and therefore do not include the eight corners of the cube.
where the sum is over the 18 nearest (in the transform space) neighbors of the given point.
The coefficients 𝜔𝑚 are given in Table 17.1.
expressing the position of the central point as a weighted mean of its 18 nearest neighbors.
The denominator of (20.1.57) is equal to 2(𝛼1 + 𝛼2 + 𝛼3 ) which is guaranteed to be positive
by (17.1.25). This vector equation is equivalent to the three scalar equations
∑𝑚 𝜔𝑚 x𝑚
𝑥= , (20.1.58)
∑𝑚 ω𝑚
∑ 𝑚 𝜔 𝑚 y𝑚
𝑦= , (20.1.59)
∑ 𝑚 𝜔𝑚
m Index 𝜔𝑚
1 𝑖+1 𝛼1
2 𝑖– 1 𝛼1
3 𝑗+1 𝛼2
4 𝑗– 1 𝛼2
5 𝑘+1 𝛼3
6 𝑘– 1 𝛼3
7 𝑖 + 1, 𝑗 + 1 𝛽1 /2
8 𝑖– 1, 𝑗– 1 𝛽1 /2
9 𝑖 + 1, 𝑗– 1 −𝛽1 /2
10 𝑖– 1, 𝑗 + 1 −𝛽1 /2
11 𝑗 + 1, 𝑘 + 1 𝛽2 /2
12 𝑗– 1, 𝑘– 1 𝛽2 /2
13 𝑗 + 1, 𝑘– 1 −𝛽2 /2
14 𝑗– 1, 𝑘 + 1 −𝛽2 /2
15 𝑖 + 1, 𝑘 + 1 𝛽3 /2
16 𝑖– 1, 𝑘– 1 𝛽3 /2
17 𝑖 + 1, 𝑘– 1 −𝛽3 /2
18 𝑖– 1, 𝑘 + 1 −𝛽3 /2
∑ 𝑚 𝜔𝑚 𝑧𝑚
𝑧= . (20.1.60)
∑ 𝑚 𝜔𝑚
the same weights 𝜔𝑚 appearing in each equation.
These equations are nonlinear, since the coefficients are functions of the coordinates.
Therefore we solve them by an iterative scheme, such as SOR. When applied to Equation
(20.1.58), for example, this gives for the (n+1)st iteration
∑ 𝑚 𝜔 𝑚 𝑥𝑚
𝑥𝑛+1 = (1 − f)𝑥𝑛 + f ( ), (20.1.61)
∑ 𝑚 𝜔𝑚
where the over relaxation factor 𝑓 must satisfy 0 < 𝑓 < 2. In (20.1.61) the values of 𝑥𝑚 at the
neighboring points are the latest available values. The coefficients 𝜔𝑚 are recalculated
before each iteration using Table 17.1 and Equations (17.1.25).
involve the coordinates of the central point, since the ’s and’s do not. Hence we simply
solve Equations (17.1.29) for the new coordinates (𝑥, 𝑦, 𝑧), holding the 18 neighboring
points fixed, without needing to iterate.
αtot
𝑛+1
∑α=1 𝑉a 𝑥αn
𝑥𝐾 = . (20.1.64)
αtot
∑α=1 Va
𝑛+1
𝑥𝑛+1 = 𝑤E 𝑥E𝑛+1 + 𝑤SA 𝑥SA 𝑛+1
+ 𝑤K 𝑥K . (20.1.65)
The remap step maps the solution from a distorted Lagrangian mesh on to the new
mesh. The underlying assumptions of the remap step are 1) the topology of the mesh is
fixed (a complete rezone does not have this limitation), and 2) the mesh motion during a
step is less than the characteristic lengths of the surrounding elements. Within the fluids
community, the second condition is simply stated as saying the Courant number, 𝐶, is less
than one.
𝑢Δ𝑡 𝑓
𝐶= = ≤ 1, (20.2.66)
Δ𝑥 V
Since the mesh motion does not occur over any physical time scale, t is arbitrary,
and ut is the transport volume, 𝑓 , between adjacent elements. The transport volume
calculation is purely geometrical for ALE formulations and it is not associated with any of
the physics of the problem.
The algorithms for performing the remap step are taken from the computational
fluids dynamics community, and they are referred to as “advection” algorithms after the
first order, scalar conservation equation that is frequently used as a model hyperbolic
problem.
∂φ ∂φ
+ a(𝑥) = 0. (20.2.67)
∂t ∂𝑥
A good advection algorithm for the remap step is accurate, stable, conservative and
monotonic. Although many of the solution variables, such as the stress and plastic strain,
are not governed by conservation equations like momentum and energy, it is still highly
desirable that the volume integral of all the solution variables remain unchanged by the
remap step. Monotonicity requires that the range of the solution variables does not
increase during the remap. This is particularly important with mass and energy, where
negative values would lead to physically unrealistic solutions.
The Donor Cell Algorithm. Aside from its first order accuracy, it is everything a good
𝜑
advection algorithm should be: stable, monotonic, and simple. The value of 𝑓𝑗 is
dependent on the sign of a at node 𝑗, which defines the upstream direction.
𝑛+1 𝑛 Δ𝑡 𝜑 𝜑
φ𝑗+ 1⁄ = φ𝑗+1⁄ + (𝑓𝑗 − 𝑓𝑗+1 ), (20.2.68)
2 2 Δ𝑥
𝜑 𝑎𝑗 𝑛 𝑛
|𝑎𝑗 | 𝑛 𝑛
𝑓𝑗 = (φ𝑗− 1⁄ + φ𝑗+1⁄ ) + (φ𝑗− 1⁄ − φ𝑗+1⁄ ). (20.2.69)
2 2 2 2 2 2
The donor cell algorithm is a first order Godunov method applied to the advection
equation. The initial values of 𝜙 to the left and the right of node 𝑗 are φ𝑛𝑗−1⁄ and φ𝑛𝑗+1⁄ , and
2 2
the velocity of the contact discontinuity at node 𝑗 is 𝑎𝑗 .
The Van Leer MUSCL Algorithm. Van Leer [1977] introduced a family of higher order
Godunov methods by improving the estimates of the initial values of left and right states
for the Riemann problem at the nodes. The particular advection algorithm that is
presented in this section is referred to as the MUSCL (monotone upwind schemes for
conservation laws) algorithm for brevity, although MUSCL really refers to the family of
algorithms that can be applied to systems of equations.
The donor cell algorithm assumes that the distribution of 𝜙 is constant over an
element. Van Leer replaces the piecewise constant distribution with a higher order
interpolation function, φ𝑛𝑗+1⁄ (𝑥) that is subject to an element level conservation constraint.
2
The value of 𝜙 at the element centroid is regarded in this context as the average value of 𝜙
over the element instead of the spatial value at 𝑥𝑗+1⁄ .
2
𝑥𝑗+1
𝑛 𝑛
φ𝑗+ 1⁄ = ∫ φ𝑗+ 1⁄ (𝑥)d𝑥, (20.2.70)
2 𝑥𝑗 2
The first step up from a piecewise constant function is a piecewise linear function,
where 𝑥 is now the volume coordinate. The volume coordinate of a point is simply the
volume swept along the path between the element centroid and the point. Conservation is
guaranteed by expanding the linear function about the element centroid.
𝑛 𝑛 𝑛 𝑛
φ𝑗+ 1⁄ (𝑥) = 𝑆𝑗+1⁄ (𝑥 − 𝑥𝑗+1⁄ ) + φj+1⁄ . (20.2.71)
2 2 2 2
Letting 𝑠𝑛𝑗+1⁄ be a second order approximation of the slope, the monotonicity limited
2
𝑛
value of the slope, 𝑠𝑗+ 1⁄ , according to the first limiting approach, is determined by
2
assuming the maximum permissible values at the element boundaries.
𝑛 1 L R L 𝑛 R
𝑆𝑗+ 1⁄ = (sgn(s ) + sgn(s )) × min (∣s ∣, ∣s𝑗+1⁄ ∣ , ∣s ∣), (20.2.72)
2 2 2
φ𝑛𝑗+1⁄ − φ𝑛𝑗−1⁄
2 2
𝑠L = , (20.2.73)
1 Δ𝑥
2 𝑗+1⁄2
φ𝑛𝑗+3⁄ − φ𝑛𝑗+1⁄
2 2
𝑠R = . (20.2.74)
1 Δ𝑥
2 𝑗+1⁄2
The second limiter is similar to the first, but it assumes that the maximum
permissible values occur at the centroid of the transport volumes. Note that as stated in
Equation (17.2.6), this limiter still limits the slope at the element boundary even if the
element is the downstream element at that boundary. A more compressive limiter would
not limit the slope based on the values of 𝜙 at the downstream boundaries. For example, if
𝑎𝑗 is negative, only 𝑆𝑅 would limit the value of 𝑆𝑛 in Equation (17.2.6). If the element is the
downstream element at both boundaries, then the slope in the element has no effect on the
solution.
φ𝑛𝑗+1⁄ − φ𝑛𝑗−1⁄
2 2
𝑠L = . (20.2.75)
1 Δ𝑥 − 1 max(0, 𝑎𝑗 Δ𝑡)
2 𝑗+1⁄2 2
φ𝑛𝑗+3⁄ − φ𝑛𝑗+1⁄
2 2
𝑠R = . (20.2.76)
1𝑥 + 1 min(0, 𝑎𝑗+1 Δ𝑡)
2 𝑗+1⁄2 2
𝑛 𝑛 𝑛
φ− C
𝑗 = S𝑗−1⁄ (𝑥 − 𝑥𝑗−1⁄ ) + φ𝑗−1⁄ , (20.2.79)
2 2 2
1
𝑥C = 𝑥𝑗𝑛 + 𝑎𝑗 Δ𝑡. (20.2.80)
2
The method for obtaining the higher order approximation of the slope is not unique.
Perhaps the simplest approach is to fit a parabola through the centroids of the three
𝑛
adjacent elements and evaluate its slope at 𝑥𝑗+ 1⁄ . When the value of 𝜙 at the element
2
centroids is assumed to be equal to the element average this algorithm defines a projection.
𝑠𝑛𝑗+1⁄ = 2 2 2 2
, (20.2.81)
2 Δ𝑥𝑗 Δ𝑥𝑗+1 (Δ𝑥𝑗 + Δ𝑥𝑗+1 )
𝑛 𝑛
Δ𝑥𝑗 = 𝑥𝑗+ 1⁄ − 𝑥𝑗−1⁄ . (20.2.82)
2 2
The advection in LS-DYNA is performed isotropically. The fluxes through each face
of element A are calculated simultaneously, but the values of 𝜙 in the transport volumes
are calculated using the one-dimensional expressions developed in the previous sections.
6
1 ⎛ 𝑛 𝑛
f𝑗 ⎞
φ
φ𝑛+1
A = 𝑛+1
⎜ V φ
A A + ∑ ⎟. (20.2.83)
VA ⎝ 𝑗=1 ⎠
The one-dimensional MUSCL scheme, which requires elements on either side of the
element whose transport is being calculated, cannot be used on the boundary elements in
the direction normal to the boundary. Therefore, in the boundary elements, the donor cell
algorithm is used to calculate the transport in the direction that is normal to the boundary,
while the MUSCL scheme is used in the two tangential directions.
It is implicitly assumed by the transport calculations that the solution variables are
defined per unit current volume. In LS-DYNA, some variables, such as the internal energy,
are stored in terms of the initial volume of the element. These variables must be rescaled
before transport, then the initial volume of the element is advected between the elements,
and then the variables are rescaled using the new “initial” volumes. Hyperelastic materials
are not currently advected in LS-DYNA because they require the deformation gradient,
which is calculated from the initial geometry of the mesh. If the deformation gradient is
integrated by using the midpoint rule, and it is advected with the other solution variables,
then hyperelastic materials can be advected without any difficulties.
Δ𝑡 Δ𝑡
F𝑛+1 = (I − )−1 (I + ) F𝑛 . (20.2.84)
𝑛+1⁄2 𝑛+1⁄2
2L 2L
Advection of the Nodal Velocities. Except for the Godunov schemes, the velocity is
centered at the nodes or the edges while the remaining variables are centered in the
elements. Momentum is advected instead of the velocity in most codes to guarantee that
momentum is conserved. The element-centered advection algorithms must be modified to
advect the node-centered momentum. Similar difficulties are encountered when node-
centered algorithms, such as the SUPG method [Brooks and Hughes 1982], are applied to
element-centered quantities [Liu, Chang, and Belytschko, to be published]. There are two
approaches: 1) construct a new mesh such that the nodes become the element centroids of
the new mesh and apply the element-centered advection algorithms, and 2) construct an
auxiliary set of element-centered variables from the momentum, advect them, and then
reconstruct the new velocities from the auxiliary variables. Both approaches can be made
to work well, but their efficiency is heavily dependent on the architecture of the codes. The
algorithms are presented in detail for one dimension first for clarity. Their extensions to
three dimensions, which are presented later, are straightforward even if the equations do
become lengthy. A detailed discussion of the algorithms in two dimensions is presented in
Reference [Benson 1992].
Notation. Finite difference notation is used in this section so that the relative locations of
the nodes and fluxes are clear. The algorithms are readily applied, however, to
unstructured meshes. To avoid limiting the discussion to a particular element-centered
advection algorithm, the transport volume through node 𝑖 is 𝑓 , the transported mass is 𝑓𝑖 ,̃
and the flux of 𝜙 is 𝜙𝑖 𝑓𝑖 . Most of the element-centered flux-limited advection algorithms
calculate the flux of 𝜙 directly, but the mean value of 𝜙 in the transport volumes is
calculated by dividing the 𝜙i 𝑓i , by the transport volume. A superscript “-” or “+” denotes
the value of a variable before or after the advection. Using this notation, the advection of 𝜙
in one dimension is represented by Equation (17.2.12), where the volume is V.
(φ− V−
𝑗+1⁄ 𝑗+1⁄
+ φ𝑖 𝑓𝑖 − φ𝑖+1 𝑓𝑖+1 )
+
φ𝑗+ 2 2 (20.2.85)
1⁄ = ,
2 V+
+ −
V𝑗+ 1⁄ = V𝑗+1⁄ + 𝑓𝑖 − 𝑓𝑖+1 . (20.2.86)
2 2
The Staggered Mesh Algorithm. YAQUI [Amsden and Hirt 1973] was the first code to
construct a new mesh that is staggered with respect to original mesh for momentum
advection. The new mesh is defined so that the original nodes become the centroids of the
new elements. The element-centered advection algorithms are applied to the new mesh to
advect the momentum. In theory, the momentum can be advected with the transport
volumes or the velocity can be advected with the mass.
(M𝑗− v− ̃ ̃
𝑗 + v𝑗−1⁄ 𝑓𝑗−1⁄ − v𝑗+1⁄ 𝑓𝑗+1⁄ )
2 2 2 2
v+
𝑗 = , (20.2.87)
M𝑗+
(M𝑗− v−
𝑗 + {ρv}𝑗−1⁄ 𝑓𝑗−1⁄ − {ρv}𝑗+1⁄ 𝑓𝑗−1⁄ )
2 2 2 2
v+
𝑗 = , (20.2.88)
M𝑗+
1 − −
M𝑗+ = (M𝑗− 1⁄ + 𝜌𝑗−1 𝑓𝑗−1 − 𝜌𝑗 𝑓𝑗 + M𝑗+1⁄ + 𝜌𝑗 𝑓𝑗 − 𝜌𝑗+1 𝑓𝑗+1 ), (20.2.91)
2 2 2
1
M𝑗+ = M𝑗− + [(𝜌𝑗−1 𝑓𝑗−1 − 𝜌𝑗 𝑓𝑗 ) + (𝜌𝑗 𝑓𝑗 − 𝜌𝑗+1 𝑓𝑗+1 )]. (20.2.92)
2
The staggered mass fluxes and transport volumes are defined by equating Equation
(20.2.90) and Equation (17.2.15).
1
𝜌𝑗+1⁄ 𝑓𝑗+1⁄ = 𝑓𝑗+̃ 1⁄ = (𝜌𝑗 𝑓𝑗 + 𝜌𝑗+1 𝑓𝑗+1 ). (20.2.93)
2 2 2 2
The density 𝜌𝑗+1⁄ is generally a nonlinear function of the volume 𝑓𝑗+1⁄ , hence calculating
2 2
𝑓𝑗+1⁄ from Equation (20.2.93) requires the solution of a nonlinear equation for each
2
transport volume. In contrast, the mass flux is explicitly defined by Equation (20.2.93).
Most codes, including KRAKEN [Debar 1974], CSQ [Thompson 1975], CTH [McGlaun
1989], and DYNA2D [Hallquist 1980], use mass fluxes with the staggered mesh algorithm
because of their simplicity.
random number of edges that each staggered element might have. In the ALE calculations
of DYNA2D, only the nodes that have a locally logically regular mesh surrounding them
can be moved in order to avoid these difficulties [Benson 1992]. These difficulties do not
occur in finite difference codes which process logically regular blocks of zones. Another
criticism is the staggered mesh algorithm tends to smear out shocks because not all the
advected variables are element-centered [Margolin 1989]. This is the primary reason,
according to Margolin [1989], that the element-centered algorithm was adopted in SALE
[Amdsden, Ruppel, and Hirt 1980].
The SALE Algorithm. SALE advects an element-centered momentum and redistributes its
changes to the nodes [Amdsden, Ruppel, and Hirt 1980]. The mean element velocity,
𝐯
̅̅̅̅𝑗+1⁄ , specific momentum, 𝐩𝑗+1⁄ , element momentum, 𝐏𝑗+1⁄ , and nodal momentum are
2 2 2
defined by Equation (17.2.17).
1
̅̅̅̅𝑗+1⁄ = (𝐯𝑗 + 𝐯𝑗+1 ),
𝐯 (20.2.94)
2 2
𝐩𝑗+1⁄ = ρ𝑗+1⁄ 𝐯
̅̅̅̅𝑗+1⁄ , (20.2.95)
2 2 2
𝐏j+1⁄ = M𝑗+1⁄ 𝐯
̅̅̅̅𝑗+1⁄ . (20.2.96)
2 2 2
Denoting the change in the element momentum Δ𝐏𝑗+1⁄ , the change in the velocity at a
2
node is calculated by distributing half the momentum change from the two adjacent
elements.
Δ𝐏𝑗−1⁄ = p𝑗−1 𝑓𝑗−1 − p𝑗 𝑓𝑗 , (20.2.97)
2
1
P𝑗+ = P𝑗− + (Δ𝐏𝑗−1⁄ + Δ𝐏𝑗+1⁄ ), (20.2.98)
2 2 2
𝐏𝑗+
𝐯𝑗+ = . (20.2.99)
M𝑗+
̅̅̅̅𝑗+1⁄ with
This algorithm can also be implemented by advecting the mean velocity, 𝐯
2
̅̅̅̅𝑗 𝑓𝑗 .̃
the transported mass, and the transported momentum 𝐩𝑗 𝑓𝑗 is changed to 𝐯
The HIS (Half Index Shift) Algorithm. Benson [1992] developed this algorithm based on
his analysis of other element-centered advection algorithms. It is designed to overcome the
dispersion errors of the SALE algorithm and to preserve the monotonicity of the velocity
field. The SALE algorithm is a special case of a general class of algorithms. To sketch the
idea behind the HIS algorithm, the discussion is restricted to the scalar advection equation.
Two variables, Ψ1,𝑗+1⁄ and Ψ2,𝑗+1⁄ are defined in terms of a linear transformation of 𝜙𝑗 and
2 2
𝜙𝑗+1 . The linear transformation may be a function of the element 𝑗 + 1/2.
−
{Ψ1,𝑗+1⁄2 ⎫
⎧ −
}
= [a b ] {φ𝑗 }. (20.2.100)
⎨ − ⎬ −
{
⎩
Ψ2,𝑗+ 1⁄ ⎭} c d φ𝑗+1
2
φ+ 1 ⎧Ψ+ ⎫
𝑗
{ + }= [d −b] { 1,𝑗+1⁄2 }. (20.2.101)
φ𝑗+1 ad − bc −c a ⎨ { Ψ+ ⎬
⎩ 2,𝑗+1⁄2 }
⎭
A function is monotonic over an interval if its derivative does not change sign. The
sum of two monotonic functions is monotonic, but their difference is not necessarily
− −
monotonic. As a consequence, Ψ1,𝑗+ 1⁄ and Ψ2,𝑗+1⁄ are monotonic over the same intervals
2 2
as φ−
𝑗 if all the coefficients in the linear transformation have the same sign. On the other
+ +
hand, φ+𝑗 is not necessarily monotonic even if Ψ1,𝑗+1⁄ and Ψ2,𝑗+1⁄ are monotonic because of
2 2
the appearance of the negative signs in the inverse matrix. Monotonicity can be maintained
by transforming in both directions provided that the transformation matrix is diagonal.
Symmetry in the overall algorithm is obtained by using a weighted average of the values of
𝜙𝑗 calculated in elements 𝑗 + 1/2 and 𝑗 − 1/2.
{Ψ1,𝑗+1⁄2 ⎫
⎧ −
}
= [1 0] {v𝑗 } (20.2.102)
⎨ ⎬ −
{ Ψ
⎩ 2,𝑗+1⁄2 }
⎭ 0 1 v𝑗+1
(M− 1 Ψ− −
+ Ψ𝑚,𝑗 𝑓𝑗 ̃ − Ψ𝑚,𝑗+1
− ̃ )
𝑓𝑗+1
𝑗+ 𝑚,𝑗+1
+ 2 2 (20.2.103)
Ψ𝑚,𝑗+1/2 = ,
M+ 1
𝑗+
2
1
v𝑗 = (M𝑗+1/2 Ψ1,𝑗+1/2 + M𝑗−1/2 Ψ2,𝑗−1/2 ). (20.2.104)
2M𝑗
Dispersion Errors. A von Neumann analysis [Trefethen 1982] characterizes the dispersion
errors of linear advection algorithms. Since the momentum advection algorithm modifies
the underlying element-centered advection algorithm, the momentum advection algorithm
does not necessarily have the same dispersion characteristics as the underlying algorithm.
18-20 (Simplified Arbitrary Lagrangian-Eulerian) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Simplified Arbitrary Lagrangian-Eulerian
The von Neumann analysis provides a tool to explore the changes in the dispersion
characteristics without considering a particular underlying advection algorithm.
The model problem is the linear advection equation with a constant value of c. A
class of solutions can be expressed as complex exponentials, where i is √−1 , is the
frequency, and χ is the wave number.
∂φ ∂φ
+c = 0, (20.2.105)
∂𝑡 ∂𝑥
For Equation (17.2.24), the dispersion equation is 𝜔 = 𝑐𝜒, but for discrete
approximations of the equation and for general hyperbolic equations, the relation is 𝜔 =
𝜔𝜒. The phase velocity, cp, and the group speed, cg, are defined by Equation (17.2.25).
ω
cp = , (20.2.107)
χ
∂ω
cg = . (20.2.108)
∂χ
The mesh spacing is assumed to have a constant value 𝐽, and the time step, ℎ, is also
constant. The + and - states in the previous discussions correspond to times n and n + 1 in
the dispersion analysis. An explicit linear advection method that has the form given by
Equation (1.2.109) results in a complex dispersion equation, Equation (17.2.27), where Π is
a complex polynomial.
𝜑𝑛+1
𝑗 = 𝜑𝑛𝑗 + F(c, ℎ, 𝐽, . . . , 𝜑𝑛𝑗−1 , 𝜑𝑛𝑗, 𝜑𝑛𝑗+1 , . . . ), (20.2.109)
The dispersion equation has the general form given in Equation (1.2.112), where Πr
and Πi denote the real and imaginary parts of Π, respectively.
Π𝑖
ωℎ = tan−1 ( ). (20.2.112)
1 + Πr
Recognizing that the relations in the above equations are periodic in 𝜔ℎ and χ 𝐽 , the
normalized frequency and wave number are defined to simplify the notation.
𝜔
̅̅̅̅ = ωℎ, ̅̅̅̅ = χ 𝐽 .
χ (20.2.113)
The von Neumann analysis of the SALE algorithm proceeds by first calculating the
increment in the cell momentum.
1
p𝑛𝑗+1/2 = (v𝑛𝑗 + v𝑗+1
𝑛
), (20.2.114)
2
1
p𝑛𝑗+1/2 = (1 + 𝑒−𝑖χ̅̅̅̅ )v𝑛𝑗 , (20.2.115)
2
𝑛+1 𝑛
Δp𝑛+1
𝑗+1/2 = P𝑗+1⁄ − P𝑗+1⁄ , (20.2.116)
2 2
1
Δp𝑛+1
𝑗+1/2 = (1 + e
−𝑖χ
̅̅ ̅̅
)Πv𝑛𝑗 . (20.2.117)
2
The dispersion relation for the SALE advection algorithm is given by Equation
(1.2.121).
1 (1 + cos(χ
̅̅̅̅))Π𝑖 ⎞
⎛ 2
̅̅̅̅ = tan−1 ⎜
𝜔 ⎜
⎜
⎟
⎟
⎟. (20.2.121)
1 + 1 (1 + cos(χ ̅̅̅̅))Π
⎝ 2 r⎠
By comparing Equation (20.2.112) and Equation (20.2.121), the effect of the SALE
momentum advection algorithm on the dispersion is to introduce a factor λ, equal to
1
2
(1 + cos(χ ̅̅̅̅))Π, into the spatial part of the advection stencil. For small values of χ ̅̅̅̅, λ is
close to one, and the dispersion characteristics are not changed, but when χ ̅̅̅̅ is π, the phase
and group velocity go to zero and the amplification factor is one independent of the
underlying advection algorithm. Not only is the wave not transported, it is not damped
out. The same effect is found in two dimensions, where λ, has the form 14 (1 + cos(χ ̅̅̅̅) +
cos(χ̅̅̅̅
̅̅̅̅ ) + cos(χ ̅̅̅̅
̅̅̅̅)cos(χ
̅̅̅̅ )).
In contrast, none of the other algorithms alter the dispersion characteristics of the
underlying algorithm. Benson has demonstrated for the element-centered algorithms that
the SALE inversion error and the dispersion problem are linked. Algorithms that fall into
the same general class as the SALE and HIS algorithms will, therefore, not have dispersion
problems [Benson 1992].
18-22 (Simplified Arbitrary Lagrangian-Eulerian) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Simplified Arbitrary Lagrangian-Eulerian
The SALE advection algorithm calculates the average momentum of the element
from the four velocities at the nodes and distributes 1⁄8 of the change in momentum to each
node.
𝑘+1 𝑙+1𝑗+1
1
p𝑗+1⁄ ,𝑘+1⁄ ,𝑙+1⁄ = 𝜌𝑗+1⁄ ,𝑘+1⁄ ,𝑙+1⁄ ∑ ∑ ∑ v𝐽𝐾𝐿 , (20.2.123)
2 2 2 8 2 2 2
𝐽=𝑗 𝐾=𝑘 𝐿=𝑙
2 𝑗+1⁄
2 𝑘+1⁄
2 𝑙+1⁄
1 ⎛ − − 1 ⎞
v+
𝑗,𝑘,𝑙 = + ⎜ ⎜
⎜ 𝑀 𝑣
𝑗,𝑘,𝑙 𝑗,𝑘,𝑙 + ∑ ∑ ∑ 𝛥𝑃𝐽,𝐾,𝐿 ⎟
⎟
⎟. (20.2.125)
𝑀𝑗,𝑘,𝑙 8 1 1 1
⎝ 𝐽=𝑗− ⁄2 𝐾=𝑘− ⁄2 𝐿=𝑙− ⁄2 ⎠
The HIS algorithm is also readily extended to three dimensions. The variable
definitions are given in Equation (1.2.126) and Equation (1.2.127), where the subscript A
refers to the local numbering of the nodes in the element. In an unstructured mesh, the
relative orientation of the nodal numbering within the elements may change. The subscript
̃ is the local
A is always with reference to the numbering in element 𝑗, 𝑘, 𝑙. The subscript A
node number in an adjacent element that refers to the same global node number as A.
ΨA,𝑗+1⁄ ,𝑘+1⁄ ,𝑙+1⁄ = vA,𝑗+1⁄ ,𝑘+1⁄ ,𝑙+1⁄ , (20.2.126)
2 2 2 2 2 2
times may not work at late times regardless of how the mesh is distributed over the
material domain. To circumvent this difficulty, a manual rezoning capability has been
implemented in LS-DYNA. The general procedure is to 1) interrupt the calculation when
the mesh is no longer acceptable, 2) generate a new mesh with INGRID by using the
current material boundaries from LS-DYNA (the topologies of the new and old mesh are
unrelated), 3) remap the solution from the old mesh to the new mesh, and 4) restart the
calculation.
This chapter will concentrate on the remapping algorithm since the mesh generation
capability is documented in the INGRID manual [Stillman and Hallquist 1992]. The
remapping algorithm first constructs an interpolation function on the original mesh by
using a least squares procedure, and then interpolates for the solution values on the new
mesh.
The objective function for minimization, 𝐽, is defined material by material, and each
material is remapped independently.
1
𝐽 = ∫(φA NA − φ)2 dV. (20.3.129)
2
V
The objective function is minimized by setting the derivatives of 𝐽 with respect to𝜙𝐴
equal to zero.
∂𝐽
= ∫(φB NB − φ)NA dV = 0. (20.3.130)
∂φA
V
The least square values of 𝜙A are calculated by solving the system of linear
equations, Equation (17.3.4).
The “mass matrix”, MAB , is lumped to give a diagonal matrix. This eliminates the spurious
oscillations that occur in a least squares fit around the discontinuities in the solution (e.g.,
shock waves) and facilitates an explicit solution for 𝜙A . The integral on the right hand side
of Equation (20.3.131) is evaluated using one point integration. By introducing these
simplifications, Equation (17.3.4) is reduced to Equation (1.3.133), where the summation
over a is restricted to the elements containing node A.
∑α φα Vα
φA = . (20.3.133)
∑α Vα
The value of 𝜙α is the mean value of 𝜙 in element a. From this definition, the value
of 𝜙α is calculated using Equation (1.3.134).
1
φα = ∫ φA NA dV. (20.3.134)
Vα
Vα
The integrand in Equation (20.3.134) is defined on the old mesh, so that Equation
(20.3.134) is actually performed on the region of the old mesh that overlaps element in the
new mesh, where the superscript “*” refers to elements on the old mesh.
1
φα = ∑ ∗ ∫ φA NA dV∗ . (20.3.135)
Vα β
Vα ∩V∗β
One point integration is currently used to evaluate Equation (20.3.135), although it would
be a trivial matter to add higher order integration. By introducing this simplification,
Equation (20.3.135) reduces to interpolating the value of 𝜙𝛼 from the least squares fit on the
old mesh.
𝜙a = 𝜙A NA (𝜉 ∗ , 𝜂∗ , 𝜁 ∗ ). (20.3.136)
The isoparametric coordinates in the old mesh that correspond to the spatial location
of the new element centroid must be calculated for Equation (20.3.136). The algorithm that
is described here is from Shapiro [1990], who references [Thompson and Maffeo 1985,
Maffeo 1984, Maffeo 1985] as the motivations for his current strategy, and we follow his
notation. The algorithm uses a “coarse filter” and a “fine filter” to make the search for the
correct element in the old mesh efficient.
The coarse filter calculates the minimum and maximum coordinates of each element
in the old mesh. If the new element centroid, (𝑥𝑠 , 𝑦𝑠 , 𝑧𝑠 ), lies outside of the box defined by
the maximum and minimum values of an old element, then the old element does not
contain the new element centroid.
Several elements may pass the coarse filter but only one of them contains the new
centroid. The fine filter is used to determine if an element actually contains the new
centroid. The fine filter algorithm will be explained in detail for the two-dimensional case
since it easier to visualize than the three-dimensional case, but the final equations will be
given for the three-dimensional case.
The two edges adjacent to each node in Figure 20.2 (taken from [Shapiro 1990])
define four skew coordinate systems. If the coordinates for the new centroid are positive
for each coordinate system, then the new centroid is located within the old element.
Because of the overlap of the four positive quarter spaces defined by the skew coordinate
systems, only two coordinate systems actually have to be checked. Using the first and third
coordinate systems, the coordinates, α𝑖 , are the solution of Equation (17.3.9).
𝑉𝑠 = 𝑉1 + 𝑎1 𝑉12 + 𝑎2 𝑉14 , (20.3.137)
Two sets of linear equations are generated for the α𝑖 by expanding the vector
equations.
𝑥 −𝑥 𝑥 −𝑥 𝛼 𝑥 −𝑥
[𝑦2 − 𝑦1 𝑦4 − 𝑦1 ] {𝛼1 } = {𝑦𝑠 − 𝑦1 }, (20.3.139)
2 1 4 1 2 𝑠 1
𝑥 −𝑥 𝑥4 − 𝑥3 𝛼3 𝑥𝑠 − 𝑥3
[𝑦2 − 𝑦3 𝑦4 − 𝑦3 ] {𝛼4 } = {𝑦𝑠 − 𝑦3 }. (20.3.140)
2 3
y
4
3
(xs, ys)
1 α4v34 α3v32
α1v12
v1 α2v14
vs 2
x
Figure 20.2. Skew Coordinate System
The fine filter sometimes fails to locate the correct element when the mesh is
distorted. When this occurs, the element that is closest to the new centroid is located by
finding the element for which the sum of the distances between the new centroid and the
nodes of the element is a minimum.
Once the correct element is found, the isoparametric coordinates are calculated
using the Newton-Raphson method, which usually converges in three or four iterations.
∂𝑁A ∂𝑁A ∂𝑁A
⎡𝑥A 𝑥A 𝑥A ⎤
⎢ ∂𝜉 ∂𝜂 ∂𝜁 ⎥
⎢ ∂𝑁 ∂𝑁A ⎥ ⎧Δ𝜉 ⎫ ⎧𝑥 − 𝑥 𝑁 ⎫
⎢ A ∂𝑁A ⎥ {Δ𝜂 } {𝑦s − 𝑦A 𝑁A }
⎢𝑦A 𝑦A 𝑦A ⎥ =⎨ s A A ⎬, (20.3.145)
∂𝜉 ∂𝜂 ∂𝜁 ⎥ ⎨
{ ⎬
⎢
⎢ ∂𝑁 ⎥ ⎩Δ𝜁 ⎭ ⎩𝑧s − 𝑧A 𝑁A }
} { ⎭
⎢𝑧 A ∂𝑁A ∂𝑁A ⎥
A 𝑧A 𝑧A
⎣ ∂𝜉 ∂𝜂 ∂𝜁 ⎦
𝜉 𝑖+1 = 𝜉 𝑖 + 𝛥𝜉 ,
𝜂𝑖+1 = 𝜂𝑖 + 𝛥𝜂, (20.3.146)
𝜁 𝑖+1 = 𝜁 𝑖 + 𝛥𝜁 .
21
Stress Update Overview
where
LS-DYNA DEV 10/27/16 (r:8004) 19-1 (Stress Update Overview)
Stress Update Overview LS-DYNA Theory Manual
and 𝑟𝑖𝑗𝑛 gives the rotation of the stress at time 𝑡𝑛 to the configuration at 𝑡𝑛 + 1
𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄2
𝑟𝑖𝑗𝑛 = (𝜎𝑖𝑝
𝑛
𝜔𝑝𝑗 𝑛
+ 𝜎𝑗𝑝 𝜔𝑝𝑖 ) Δ𝑡 . (21.8)
In the implicit NIKE2D/3D [Hallquist 1981b] codes, which are used for low
frequency structural response, we do a half-step rotation, apply the constitutive law, and
complete the second half-step rotation on the modified stress. This approach has also been
adopted for some element formulations in LS-DYNA when the invariant stress update is
active. An exact or second order accurate rotation is performed rather than the
approximate one represented by Equation (21.3), which is valid only for small incremental
rotations. A typical implicit time step is usually 100 to 1000 or more times larger than the
explicit time step; consequently, the direct use of Equation (21.7) could lead to very
significant errors.
𝑛+1⁄2
where 𝜀̇′𝑖𝑗 is the deviatoric strain rate tensor:
𝑛+1⁄2 1
𝜀̇′𝑖𝑗 = 𝜀̇𝑖𝑗 − 𝜀̇𝑘𝑘 𝛿. (21.11)
3
Before the equation of state, Equation (21.9), is evaluated, we compute the bulk
viscosity, 𝑞, and update the total internal energy 𝑒 of the element being processed to a trial
value 𝑒∗ :
1 𝑛−1⁄2 𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄
𝑒∗𝑛+1 = 𝑒𝑛 − Δ𝑣 (𝑝𝑛 + 𝑞 +𝑞 )+𝑣 𝑠𝑖𝑗 Δ𝜀𝑖𝑗 2 , (21.12)
2
where 𝑣 is the element volume and
1
𝑛+1⁄2 𝑛+1 1
Δ𝑣 = 𝑣𝑛+1 − 𝑣𝑛 , 𝑣 = (𝑣𝑛 + 𝑣𝑛+1 ), 𝑠𝑖𝑗 2 = (𝑠𝑛𝑖𝑗 + 𝑠𝑛+1
𝑖𝑗 ).
(21.13)
2 2
The time-centering of the viscosity is explained by Noh [1976].
Assume we have an equation of state that is linear in internal energy of the form
𝑝𝑛+1 = 𝐴𝑛+1 + 𝐵𝑛+1 𝐸𝑛+1 , (21.14)
where
𝑛+1 𝑒𝑛+1
𝐸 = , (21.15)
𝑣0
and 𝜐0 is the initial volume of the element. Noting that
1
𝑒𝑛+1 = 𝑒∗𝑛+1 − Δ𝑣𝑝𝑛+1 , (21.16)
2
pressure can be evaluated exactly by solving the implicit form
(𝐴𝑛+1 + 𝐵𝑛+1 𝐸∗𝑛+1 )
𝑝𝑛+1 = , (21.17)
1
(1 + 𝐵 𝑛+1 Δ𝑣
2 𝑣0 )
and the internal energy can be updated in Equation (21.16). If the equation of state is not
linear in internal energy, a one-step iteration is used to approximate the pressure
𝑝∗𝑛+1 = 𝑝(𝑉 𝑛+1 , 𝐸∗𝑛+1 ). (21.18)
Internal energy is updated to 𝑛 + 1 using 𝑝∗𝑛+1 in Equation (21.16) and the final pressure is
then computed:
𝑝𝑛+1 = 𝑝(𝑉 𝑛+1 , 𝐸𝑛+1 ). (21.19)
This is also the iteration procedure used in KOVEC [Woodruff 1973]. All the equations of
state in LS-DYNA are linear in energy except the ratio of polynomials.
𝐹𝑖𝑗 is the deformation gradient matrix and 𝑈𝑖𝑗 and 𝑉𝑖𝑗 are the positive definite right and left
stretch tensors:
𝜕𝑥𝑖
𝐹𝑖𝑗 = . (21.23)
𝜕𝑋𝑗
Stresses are updated for all materials by adding to the rotated Cauchy stress at time
n.
𝜏𝑖𝑗𝑛 = 𝑅𝑛𝑘𝑖 𝑅𝑛𝑙𝑗 𝜎𝑘𝑙
𝑛
, (21.24)
the stress increment obtained by an evaluation of the constitutive equations,
𝑛+1⁄2 𝑛+1⁄2
Δ𝜏𝑖𝑗 = 𝐶𝑖𝑗𝑘𝑙 Δ𝑑𝑘𝑙 , (21.25)
where
𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄2
Δ𝑑𝑖𝑗 = 𝑅𝑘𝑖 𝑅𝑙𝑗 Δ𝜀𝑘𝑙 (21.26)
At low stress levels in elastoplastic materials the stresses, 𝜎𝑖𝑗 , depends only on the
state of strain; however, above a certain stress level, called the yield stress, 𝜎𝑦 (𝑎𝑖 ),
ε ε = ln(L/L0)
elastic plastic strain σ = P/A
strain
Figure 21.1. The uniaxial tension test demonstrates plastic behavior.
σ3
yield surface σ 1 = σ2 = σ3
defined by
F(δij, k)
σ2
deviatoria plane
yield curve = intersection of the deviatoric
σ1 plane with the yield surface
Figure 21.2. The yield surface in principal stress space in pressure independent.
nonrecoverable plastic deformations are obtained. The yield stress changes with increasing
plastic deformations, which are measured by internal variables, 𝑎𝑖 .
In the uniaxial tension test, a curve like that in Figure 21.1 is generated where
logrithmic uniaxial strain is plotted against the uniaxial true stress which is defined as the
applied load 𝑃 divided by the actual cross-sectional area, 𝐴.
For the simple von Mises plasticity models the yield stress is pressure independent
and the yield surface is a cylinder in principal stress space as shown in Figure 21.2. With
isotropic hardening the diameter of the cylinder grows but the shape remains circular. In
kinematic hardening the diameter may remain constant but will translate in the plane as a
function of the plastic strain tensor, See Figure 21.3.
σ3
initial yield
curve in the current
deviatoric yield
plane surface
σ1 σ2
Figure 21.3. With kinematic hardening the yield surface may shift as a function
of plastic strain.
p
As depicted in Figure 21.5 the plastic strain increments 𝑑𝜀𝑖𝑗 are normal to the plastic
potential function. This is the normality rule of plasticity.
Post-yielding behavior from uniaxial tension tests typically show the following
behaviors illustrated in Figure 21.4:
hardening ideal softening
İ
0 0 0
İ
İ
Figure 21.4. Hardening, ideal, and softening plasticity models.
σ2
σ1
σ3
Figure 21.5. The plastic strain is normal to the yield surface.
The behavior of these hardening laws are characterized in Table 18.1. below.
Although LS-DYNA permits softening to be defined and used, such softening behavior will
result in strain localization and nonconvergence with mesh refinement.
A retangular cartesian coordinate system is used so that the covariant and contravar-
iant metric tensors in the reference (undeformed) and deformed configuration are:
𝑔𝑖𝑗 = 𝑔𝑖𝑗 = 𝛿𝑖𝑗 ,
𝜕𝑥𝑘 𝜕𝑥𝑘
𝐺𝑖𝑗 = ,
𝜕𝑋𝑖 𝜕𝑋𝑗 (21.35)
𝜕𝑋𝑖 𝜕𝑋𝑗
𝐺𝑖𝑗 = .
𝜕𝑥𝑘 𝜕𝑥𝑘
The Green-St. Venant strain tensor and the principal strain invariants are defined as
1
𝛾𝑖𝑗 = (𝐺𝑖𝑗 − 𝛿𝑖𝑗 ), (21.36)
2
𝐼1 = 𝛿𝑖𝑗 𝐺𝑖𝑗 ,
1
𝐼2 = (𝛿𝑖𝑟 𝛿𝑗𝑠 𝐺𝑟𝑖 𝐺𝑠𝑗 − 𝛿𝑖𝑟 𝛿𝑗𝑠 𝐺𝑖𝑗 𝐺𝑟𝑠 ), (21.37)
2
𝐼3 = det(𝐺𝑖𝑗 ),
To allow for an arbitrary orientation of the shell elements within the finite element
mesh, each ply in the composite has a unique orientation angle, 𝛽, which measures the
offset from some reference in the element. Each integration point through the shell
thickness, typically though not limited to one point per ply, requires the definition of 𝛽at
that point. The reference is determined by the angle 𝛹which can be defined for each
element on the element card, and is measured from the 1-2 element side. Figures 21.6 and
21.7 depict these angles.
We update the stresses in the shell in the local shell coordinate system which is
defined by the 1-2 element side and the cross product of the diagonals. Thus to transform
the stress tensor into local system determined by the fiber directions entails a transfor-
mation that takes place in the plane of the shell.
In the implementation of the material model we first transform the Cauchy stress
and velocity strain tensor 𝑑𝑖𝑗 into the coordinate system of the material denoted by the
subscript L
𝛔L = 𝐪T 𝛔𝐪,
𝐪L = 𝐪T 𝐝𝐪,
𝜎11 𝜎12 𝜎13
𝛔L = ⎡
⎢𝜎21 𝜎22 𝜎23 ⎤
⎥, (21.43)
⎣𝜎32 𝜎32 𝜎33 ⎦
𝑑 𝑑12 𝑑13
⎡ 11 ⎤
𝛆L = ⎢𝑑21 𝑑22 𝑑23 ⎥,
⎣𝑑32 𝑑32 𝑑33 ⎦
θ = ψ+β
z y
θ
x
Figure 21.7. A multi-layer laminate can be defined. The angle i is defined for
the ith lamina.
The Arabic subscripts on the stress and strain (𝛔 and 𝛆) are used to indicate the
principal material directions where 1 indicates the fiber direction and 2 indicates the
transverse fiber direction (in the plane). The orthogonal 3 × 3 transformation matrix is
given by
cos𝜃 −sin𝜃 0
𝐪=⎡
⎢ sin𝜃 cos𝜃 0⎤⎥. (21.44)
⎣ 0 0 1⎦
In shell theory we assume a plane stress condition, i.e., that the normal stress, 𝜎33 , to the
mid-surface is zero. We can now incrementally update the stress state in the material
coordinates
𝑛+1⁄2
𝛔L𝑛+1 = 𝛔L𝑛 + Δ𝛔L , (21.45)
n4 n3
3 1
2 β
ψ
n1 X
n2
Figure 21.6. Orientation of material directions relative to the 1-2 side.
Δ𝜎 𝑄 𝑄12 0 0 0 𝑑
⎡Δ𝜎11 ⎤ ⎡𝑄11 ⎡ 11 ⎤
⎢ 22 ⎥ ⎢ 12 𝑄22 0 0 0 ⎤⎥ ⎢𝑑22 ⎥
𝑛+1⁄2
Δ𝛔L = ⎢Δ𝜎12 ⎥ = ⎢ 0 0 𝑄44 0 0 ⎥ ⎢𝑑12 ⎥ Δ𝑡. (21.46)
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢Δ𝜎23 ⎥ ⎢ 0 0 0 𝑄55 0 ⎥ ⎢𝑑23 ⎥
⎣Δ𝜎31 ⎦ ⎣ 0 0 0 0 𝑄66 ⎦ ⎣𝑑31 ⎦L
The terms 𝑄𝑖𝑗 are referred to as reduced components of the lamina and are defined as
𝐸11
𝑄11 = ,
1 − 𝜈12 𝜈21
𝐸22
𝑄22 = ,
1 − 𝜈12 𝜈21
𝜈12 𝐸11 (21.47)
𝑄12 = ,
1 − 𝜈12 𝜈21
𝑄44 = 𝐺12 ,
𝑄55 = 𝐺23 ,
𝑄66 = 𝐺31 .
Because of the symmetry properties,
𝐸𝑗𝑗
𝜈𝑗𝑖 = 𝜈𝑖𝑗 , (21.48)
𝐸𝑖𝑖
where 𝜈𝑖𝑗 is Poisson’s ratio for the transverse strain in jth direction for the material
undergoing stress in the ith-direction, 𝐸𝑖𝑗 are the Young’s modulii in the ith direction, and
𝐺𝑖𝑗 are the shear modulii.
After completion of the stress update we transform the stresses back into the local
shell coordinate system.
𝛔 = 𝐪𝛔L 𝐪T . (21.49)
1 𝜐 21 𝜐 31
⎡ − − 0 0 0 ⎤
⎢ 𝐸 11 𝐸 22 𝐸 33 ⎥
⎢ 𝜐 1 𝜐 ⎥
⎢− 12 − 32 0 0 0 ⎥
⎢ 𝐸 11 𝐸 22 𝐸 33 ⎥
⎢ ⎥
⎢− 𝜐 13 𝜐
− 23
1
0 0 0 ⎥
⎢ 𝐸 𝐸 22 𝐸 33 ⎥
𝐂−1
l =⎢ 11 ⎥. (21.50)
⎢ 1 ⎥
⎢ 0 0 0 0 0 ⎥
⎢ 𝐺12 ⎥
⎢ 1 ⎥
⎢ 0 0 0 0 0 ⎥
⎢ 𝐺23 ⎥
⎢ ⎥
⎢ 0 1 ⎥
0 0 0 0
⎣ 𝐺31 ⎦
c d
a
b
Figure 21.8. Local material directions are defined by a triad which can be input
for each solid element.
The biggest concern when dealing with local material directions is that the results
are not invariant with element numbering since the orientation of the local triad is fixed
with respect to the base of the brick element, nodes 1-4, in Figure 21.9. For Hyperelastic
materials where the stress tensor is computed in the initial configuration, this is not a
problem, but for materials like the honeycomb foams, the local directions can change due
to element distortion causing relative movement of nodes 1-4. In honeycomb foams we
assume that the material directions are orthogonal in the deformed configuration since the
stress update is performed in the deformed configuration.
Several erosion criteria are available that are independent of the material models.
Each one is applied independently, and once any one of them is satisfied, the element is
deleted from the calculation. The criteria for failure are:
• 𝑃 ≥ 𝑃min where P is the pressure (positive in compression), and 𝑃min is the pressure
at failure.
• 𝜎1 ≥ 𝜎max , where 𝜎1 is the maximum principal stress, and 𝜎max is the principal
stress at failure.
These failure models apply to solid elements with one point integration in 2 and 3
5 8
6 7
1 4
2 3
Figure 21.9. The orientation of the triad for the local material directions is stored
relative to the base of the solid element. The base is defined by nodes 1-4 of the
element connectivity.
dimensions.
Material 𝜎0 (Kbar)
1020 Steel 10.0 2 12.5
OFHC Copper 3.60 2 10.0
C1008 14.0 2 0.38
HY100 15.7 2 61.0
7039-T64 8.60 2 3.00
Recall, the spin tensor and strain rate tensor, Equations (21.3) and (21.5),
respectively:
1 𝜕𝑣𝑖 𝜕𝑣𝑗
𝜔𝑖𝑗 = ( − ), (21.54)
2 𝜕𝑥𝑗 𝜕𝑥𝑖
1 𝜕𝑣𝑖 𝜕𝑣𝑗
𝜀̇𝑖𝑗 = ( + ). (21.55)
2 𝜕𝑥𝑗 𝜕𝑥𝑖
where 𝜌𝑛𝑖𝑗 gives the rotational correction that transforms the strain tensor at time 𝑡𝑛 into the
configuration at 𝑡𝑛 + 1
𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄2
𝜌𝑛𝑖𝑗 = (𝜀𝑛𝑖𝑝 𝜔𝑝𝑗 + 𝜀𝑛𝑗𝑝 𝜔𝑝𝑖 ) Δ𝑡 . (21.57)
For shell elements we integrate the strain tensor at the inner and outer integration
points and output two tensors per element. When the mid surface strains are plotted in LS-
PREPOST, these are the average values.
𝑒𝑛̇ = 𝑒 ̇̃ (21.5)
𝑛
∑𝑖=𝑛−𝑁+1 𝑒𝑖̇
𝑒̇= = rate used in material routine (21.6)
𝑁
As for the running average option, the subscript denotes stored history variables necessary
for computing the running average strain rates and n is the cycle number. This also
requires storage of 𝑁 − 1 history variables, for most materials 𝑁 = 12.
18.18
LS-DYNA Theory Manual Stress Update Overview
18.19
Material Models LS-DYNA Theory Manual
22
Material Models
LS-DYNA accepts a wide range of material and equation of state models, each with
a unique number of history variables. Approximately 150 material models are
implemented, and space has been allotted for up to 10 user-specified models.
1 Elastic
2 Orthotropic Elastic
3 Kinematic/Isotropic Elastic-Plastic
4 Thermo-Elastic-Plastic
5 Soil and Crushable/Non-crushable Foam
6 Viscoelastic
7 Blatz - Ko Rubber
8 High Explosive Burn
9 Null Hydrodynamics
10 Isotropic-Elastic-Plastic-Hydrodynamic
11 Temperature Dependent, Elastoplastic, Hydrodynamic
12 Isotropic-Elastic-Plastic
13 Elastic-Plastic with Failure Model
14 Soil and Crushable Foam with Failure Model
15 Johnson/Cook Strain and Temperature Sensitive Plasticity
16 Pseudo TENSOR Concrete/Geological Model
17 Isotropic Elastic-Plastic Oriented Crack Model
18 Power Law Isotropic Plasticity
19 Strain Rate Dependent Isotropic Plasticity
20 Rigid
21 Thermal Orthotropic Elastic
22 Composite Damage Model
23 Thermal Orthotropic Elastic with 12 Curves
24 Piecewise Linear Isotropic Plasticity
25 Inviscid Two Invariant Geologic Cap Model
26 Metallic Honeycomb
27 Compressible Mooney-Rivlin Rubber
28 Resultant Plasticity
29 Force Limited Resultant Formulation
30 Closed-Form Update Shell Plasticity
31 Slightly Compressible Rubber Model
32 Laminated Glass Model
33 Barlat’s Anisotropic Plasticity Model
34 Fabric
35 Kinematic/Isotropic Elastic-Plastic Green-Naghdi Rate
36 Barlat’s 3-Parameter Plasticity Model
37 Transversely Anisotropic Elastic-Plastic
38 Blatz-Ko Compressible Foam
39 Transversely Anisotropic Elastic-Plastic with FLD
40 Nonlinear Elastic Orthotropic Material
41-50 User Defined Material Models
42 Planar Anisotropic Plasticity Model
48 Strain Rate Dependent Plasticity with Size Dependent Failure
51 Temperature and Rate Dependent Plasticity
52 Sandia’s Damage Model
53 Low Density Closed Cell Polyurethane Foam
54-55 Composite Damage Model
57 Low Density Urethane Foam
58 Laminated Composite Fabric
59 Composite Failure
60 Elastic with Viscosity
61 Maxwell/Kelvin Viscoelastic
62 Viscous Foam
63 Isotropic Crushable Foam
64 Strain Rate Sensitive Power-Law Plasticity
65 Modified Zerilli/Armstrong
66 Linear Stiffness/Linear Viscous 3D Discrete Beam
67 Nonlinear Stiffness/Viscous 3D Discrete Beam
68 Nonlinear Plastic/Linear Viscous 3D Discrete Beam
69 Side Impact Dummy Damper, SID Damper
70 Hydraulic/Gas Damper
71 Cable
72 Concrete Damage Model
73 Low Density Viscoelastic Foam
74 Elastic Spring for the Discrete Beam
75 Bilkhu/Dubois Foam Model
76 General Viscoelastic
77 Hyperviscoelastic Rubber
78 Soil/Concrete
79 Hysteretic Soil
LS-DYNA DEV 10/27/16 (r:8004) 20-21 (Material Models)
Material Models LS-DYNA Theory Manual
80 Ramberg-Osgood Plasticity
81 Plastic with Damage
82 Isotropic Elastic-Plastic with Anisotropic Damage
83 Fu-Chang’s Foam with Rate Effects
84-85 Winfrith Concrete
84 Winfrith Concrete Reinforcement
86 Orthotropic-Viscoelastic
87 Cellular Rubber
88 MTS Model
89 Plasticity Polymer
90 Acoustic
91 Soft Tissue
93 Elastic 6DOF Spring Discrete Beam
94 Inelastic Spring Discrete Beam
95 Inelastic 6DOF Spring Discrete Beam
96 Brittle Damage Model
97 General Joint Discrete Beam
100 Spot weld
101 GE Thermoplastics
102 Hyperbolic Sin
103 Anisotropic Viscoplastic
104 Damage 1
105 Damage 2
106 Elastic Viscoplastic Thermal
110 Johnson-Holmquist Ceramic Model
111 Johnson-Holmquist Concrete Model
112 Finite Elastic Strain Plasticity
113 Transformation Induced Plasticity
114 Layered Linear Plasticity
115 Elastic Creep Model
116 Composite Lay-Up Model
117-118 Composite Matrix
119 General Spring and Damper Model
120 Gurson Dilational-Plastic Model
120 Gurson Model with Rc-Dc
121 Generalized Nonlinear 1DOF Discrete Beam
122 Hill 3RC
123 Modified Piecewise Linear Plasticity
124 Tension-Compression Plasticity
126 Metallic Honeycomb
127 Arruda-Boyce rubber
128 Anisotropic heart tissue
129 Lung tissue
130 Special Orthotropic
131 Isotropic Smeared Crack
20-22 (Material Models) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Material Models
In the table below, a list of the available material models and the applicable element types
are given. Some materials include strain rate sensitivity, failure, equations of state, and
thermal effects and this is also noted. General applicability of the materials to certain kinds
of behavior is suggested in the last column.
Notes:
Gn General
Cm Compo-
sites
Cr Ceram-
ics
Fl Fluids
Fm Foam
Gl Glass
Hy Hydro-
Strain-Rate Effects
dyn
Equation-of-State
Material Number
Thermal Effects
Material Title Mt Metal
Pl Plastic
Thick Shells
Thin Shells
Rb Rubber
Sl
Failure
Beams
Bricks
Soil/Co
nc
1 Elastic Y Y Y Y Gn, Fl
2 Orthotropic Elastic (Anisotropic - solids) Y Y Y Cm, Mt
3 Plastic Kinematic/Isotropic Y Y Y Y Y Y Cm, Mt, Pl
4 Elastic Plastic Thermal Y Y Y Y Y Mt, Pl
5 Soil and Foam Y Fm, Sl
6 Linear Viscoelastic Y Y Y Y Rb
7 Blatz-Ko Rubber Y Y Rb,
Polyurethane
8 High Explosive Burn Y Y Hy
9 Null Material Y Y Y Y Fl, Hy
10 Elastic Plastic Hydro(dynamic) Y Y Y Hy, Mt
11 Steinberg: Temp. Dependent Y Y Y Y Y Hy, Mt
Elastoplastic
12 Isotropic Elastic Plastic Y Y Y Mt
13 Isotropic Elastic Plastic with Failure Y Y Mt
14 Soil and Foam with Failure Y Y Fm, Sl
15 Johnson/Cook Plasticity Model Y Y Y Y Y Y Hy, Mt
16 Pseudo TENSOR Geological Model Y Y Y Y Sl
17 Oriented Crack (Elastoplastic with Y Y Y Hy, Mt, Pl
Fracture)
18 Power Law Plasticity (Isotropic) Y Y Y Y Y Mt, Pl
Notes:
Gn General
Cm Com-
posites
Cr Ceram-
ics
Fl Fluids
Fm Foam
Gl Glass
Hy Hydro-
Strain-Rate Effects
dyn
Equation-of-State
Material Number
Thermal Effects
Material Title Mt Metal
Pl Plastic
Thick Shells
Thin Shells
Rb Rubber
Sl
Failure
Beams
Bricks
Soil/Co
nc
Notes:
Gn General
Cm Com-
posites
Cr Ceram-
ics
Fl Fluids
Fm Foam
Gl Glass
Hy Hydro-
Strain-Rate Effects
dyn
Equation-of-State
Material Number
Thermal Effects
Material Title Mt Metal
Pl Plastic
Thick Shells
Thin Shells
Rb Rubber
Sl
Failure
Beams
Bricks Soil/Co
nc
Notes:
Gn General
Cm Compo-
sites
Cr Ceramics
Fl Fluids
Fm Foam
Strain-Rate Effects
Gl Glass
Equation-of-State
Material Number
Thermal Effects
Hy Hydro-
dyn
Thick Shells
Thin Shells
Mt Metal
Material Title Pl Plastic
Failure
Beams
Bricks
Rb Rubber
Sl Soil/Conc
Notes:
Gn General
Cm Compo-
sites
Cr Ceramics
Fl Fluids
Fm Foam
Gl Glass
Hy Hydro-
Strain-Rate Effects
dyn
Equation-of-State
Material Number
Thermal Effects
Mt Metal
Material Title Pl Plastic
Thick Shells
Thin Shells
Rb Rubber
Sl
Failure
Beams
Bricks
Soil/Con
c
Notes:
Gn General
Cm Compo-
sites
Cr Ceramics
Fl Fluids
Fm Foam
Gl Glass
Hy Hydro-
Strain-Rate Effects
dyn
Equation-of-State
Material Number
Thermal Effects
Mt Metal
Material Title Pl Plastic
Thick Shells
Thin Shells
Rb Rubber
Sl
Failure
Beams
Bricks
Soil/Con
c
Notes:
Gn General
Cm Compo-
sites
Cr Ceramics
Fl Fluids
Fm Foam
Strain-Rate Effects
Gl Glass
Equation-of-State
Material Number
Thermal Effects
Hy Hydro-
dyn
Thick Shells
Thin Shells
Mt Metal
Material Title Pl Plastic
Failure
Beams
Bricks
Rb Rubber
Sl Soil/Conc
Notes:
Gn General
Cm Compo-
sites
Cr Ceramics
Fl Fluids
Fm Foam
Strain-Rate Effects
Gl Glass
Equation-of-State
Material Number
Thermal Effects
Hy Hydro-
dyn
Thick Shells
Thin Shells
Mt Metal
Material Title Pl Plastic
Failure
Beams
Bricks
Rb Rubber
Sl Soil/Conc
Notes:
Gn General
Cm Compo-
sites
Cr Ceramics
Fl Fluids
Fm Foam
Strain-Rate Effects
Gl Glass
Equation-of-State
Material Number
Thermal Effects
Hy Hydro-
dyn
Thick Shells
Thin Shells
Mt Metal
Material Title Pl Plastic
Failure
Beams
Bricks
Rb Rubber
Sl Soil/Conc
and pressure
𝑝𝑛+1 = −𝐾ln𝑉 𝑛+1 , (22.1.2)
where 𝐺 and 𝐾 are the elastic shear and bulk moduli, respectively, and 𝑉 is the relative
volume, i.e., the ratio of the current volume to the initial volume.
After computing 𝑆𝑖𝑗 , we use Equation (21.32) to obtain the Cauchy stress. This
model will predict realistic behavior for finite displacement and rotations as long as the
strains are small.
In isotropic hardening, the center of the yield surface is fixed but the radius is a
function of the plastic strain. In kinematic hardening, the radius of the yield surface is
fixed but the center translates in the direction of the plastic strain. Thus the yield condition
is
2
1 𝜎𝑦
𝜙 = 𝜉𝑖𝑗 𝜉𝑖𝑗 − = 0, (22.3.1)
2 3
where
𝜉𝑖𝑗 = 𝑠𝑖𝑗 − 𝛼𝑖𝑗 (22.3.2)
p
𝜎𝑦 = 𝜎0 + 𝛽𝐸p 𝜀eff . (22.3.3)
Strain rate is accounted for using the Cowper-Symonds [Jones 1983] model which
scales the yield stress by a strain rate dependent factor
1
⎡ 𝜀̇ 𝑝 ⎤ p
𝜎𝑦 = ⎢1 + ( ) ⎥ (𝜎0 + 𝛽𝐸p 𝜀eff ), (22.3.6)
𝐶
⎣ ⎦
where 𝑝 and 𝐶 are user defined input constants and 𝜀̇ is the strain rate defined as:
δx
Yield
Stress
⎛l⎛
ln ⎜ ⎜
⎝ l0 ⎝
The current radius of the yield surface, 𝜎𝑦 , is the sum of the initial yield strength, 𝜎0 ,
p
plus the growth 𝛽𝐸p 𝜀eff ,where 𝐸p is the plastic hardening modulus
𝐸t 𝐸
𝐸p = , (22.3.8)
𝐸 − 𝐸t
p
and 𝜀eff is the effective plastic strain
𝑡 1
p 2 p p ⁄2
𝜀eff = ∫ ( 𝜀̇𝑖𝑗 𝜀̇𝑖𝑗 ) 𝑑𝑡. (22.3.9)
3
0
The plastic strain rate is the difference between the total and elastic (right
superscript e) strain rates:
p
𝜀̇𝑖𝑗 = 𝜀̇𝑖𝑗 − 𝜀̇e𝑖𝑗 . (22.3.10)
In the implementation of this material model, the deviatoric stresses are updated
elastically, as described for model 1, but repeated here for the sake of clarity:
𝜎𝑖𝑗∗ = 𝜎𝑖𝑗𝑛 + 𝐶𝑖𝑗𝑘𝑙 Δ𝜀𝑘𝑙 , (22.3.11)
where
𝜎𝑖𝑗∗ is the trial stress tensor,
𝜎𝑖𝑗𝑛 is the stress tensor from the previous time step,
C𝑖𝑗𝑘𝑙 is the elastic tangent modulus matrix,
The application of the Jaumann rate to update the stress tensor allows for the
possibility that the normal stress, 𝜎33 , will not be zero. The first step in updating the stress
tensor is to compute a trial plane stress update assuming that the incremental strains are
elastic. In the above, the normal strain increment Δ𝜀33 is replaced by the elastic strain
increment
𝜎33 + 𝜆(Δ𝜀11 + Δ𝜀22 )
Δ𝜀33 = − , (22.3.18)
𝜆 + 2𝜇
where 𝜆 and 𝜇 are Lamé’s constants.
When the trial stress is within the yield surface, the strain increment is elastic and
the stress update is completed. Otherwise, for the plastic plane stress case, secant iteration
is used to solve Equation (22.3.16)for the normal strain increment (Δ𝜀33 ) required to
produce a zero normal stress:
p𝑖
𝑖 ∗
3𝐺Δ𝜀eff 𝜉33 (22.3.19)
𝜎33 = 𝜎33 − ,
𝛬
Here, the superscript 𝑖 indicates the iteration number.
The secant iteration formula for Δε33 (the superscript p is dropped for clarity) is
Δ𝜀𝑖33 − Δ𝜀𝑖−1
33
Δ𝜀𝑖+1 𝑖−1
33 = Δ𝜀33 − 𝑖 𝑖−1
𝑖−1
𝜎33 , (22.3.20)
𝜎33 − 𝜎33
where the two starting values are obtained from the initial elastic estimate and by
assuming a purely plastic increment, i.e.,
Δ𝜀133 = −(Δ𝜀11 − Δ𝜀22 ). (22.3.21)
These starting values should bound the actual values of the normal strain increment.
The iteration procedure uses the updated normal stain increment to update first the
deviatoric stress and then the other quantities needed to compute the next estimate of the
𝑖
normal stress in Equation (22.3.19). The iterations proceed until the normal stress 𝜎33 is
sufficiently small. The convergence criterion requires convergence of the normal strains:
∣Δ𝜀𝑖33 − Δ𝜀𝑖−1
33 ∣
< 10−4 . (22.3.22)
∣Δ𝜀𝑖+1
33 ∣
After convergence, the stress update is completed using the relationships given in
Equations (22.3.16) and (22.3.17)
Letting 𝑇 represent the temperature, we compute the elastic co-rotational stress rate
as
𝜎𝑖𝑗∇ = 𝐶𝑖𝑗𝑘𝑙 (𝜀̇𝑘𝑙 − ε̇T𝑘𝑙 ) + 𝜃𝑖𝑗̇ 𝑑𝑇, (22.4.1)
where
𝑑𝐶𝑖𝑗𝑘𝑙
𝜃𝑖𝑗̇ = 𝐶−1 𝜎̇ , (22.4.2)
𝑑𝑇 𝑘𝑙𝑚𝑛 𝑚𝑛
and 𝐶𝑖𝑗𝑘𝑙 is the temperature dependent elastic constitutive matrix:
⎡1 − 𝜐 𝜐 𝜐 0 0 0 ⎤
⎢ ⎥
⎢ ⎥
⎢ 𝜐 1−𝜐 𝜐 0 0 0 ⎥
⎢ ⎥
⎢ ⎥
⎢ 𝜐 𝜐 1−𝜐 0 0 0 ⎥
𝐸 ⎢ ⎥
𝐶𝑖𝑗𝑘𝑙 = ⎢ ⎥, (22.4.3)
(1 + 𝜐)(1 − 2𝜐) ⎢ 1 − 2𝜐
⎢ 0 0 0 0 0 ⎥⎥
⎢ 2 ⎥
⎢ 1 − 2𝜐 ⎥
⎢ 0 0 0 0 0 ⎥
⎢ 2 ⎥
⎢ 1 − 2𝜐⎥
⎣ 0 0 0 0 0
2 ⎦
where 𝜐 is Poisson’s ratio. The thermal strain rate can be written in terms of the coefficient
of thermal expansion 𝛼 as:
𝜀̇T𝑖𝑗 = 𝛼𝑇̇ 𝛿𝑖𝑗 , (22.4.4)
1 𝜎𝑦 (𝑇)2
𝜙 = 𝑠𝑖𝑗 𝑠𝑖𝑗 − , (22.4.5)
2 3
where
p
𝜎𝑦 (𝑇) = 𝜎𝑜 (𝑇) + 𝐸p (𝑇)𝜀eff . (22.4.6)
The initial yield, 𝜎o , and plastic hardening modulus, 𝐸p , are temperature dependent.
If the behavior is elastic we do nothing; otherwise, we scale back the stress deviators by the
factor 𝑓s :
𝑠𝑛+1
𝑖𝑗 = 𝑓s 𝑠∗𝑖𝑗 , (22.4.7)
where
𝜎𝑦
𝑓s = 1
,
⁄2 (22.4.8)
(3 𝑠∗𝑖𝑗 𝑠∗𝑖𝑗 )
2
and update the plastic strain by the increment
1
⁄2
(1 − 𝑓s )(3 𝑠∗𝑖𝑗 𝑠∗𝑖𝑗 )
p 2 (22.4.9)
Δ𝜀eff = .
𝐺 + 3𝐸p
tension compression
⎛V⎛
Volumetric Strain, ln ⎜ ⎜
⎝V0⎝
The bulk unloading modulus is used
Pressure Cutoff Value
if the volumetric crushing option is on
(VCR = 0). In thiscase the aterial's response
follows the black arrows.
Figure 22.5.1. Volumetric strain versus pressure curve for soil and crushable
foam model.
1⁄
2
⎛
⎜ 𝑎0 + 𝑎1 𝑝 + a2 𝑝2 ⎞
⎟
s𝑛+1
𝑖𝑗 =⎜
⎜ ⎟
⎟ s∗𝑖𝑗 . (22.5.2)
1s s
⎝ 2 𝑖𝑗 𝑖𝑗 ⎠
If the hydrostatic tension exceeds the cutoff value, the pressure is set to the cutoff
value and the deviatoric stress tensor is zeroed.
where
𝜙(𝑡) = G∞ + (G0 − G∞ )𝑒−𝛽𝑡 , (22.6.2)
is the shear relaxation modulus. A recursion formula is used to compute the new value of
the hereditary integral at time 𝑡𝑛+1 from its value at time 𝑡𝑛 . Elastic bulk behavior is
assumed:
𝑝 = 𝐾ln𝑉, (22.6.3)
where pressure is integrated incrementally.
where 𝐺 is the shear modulus, 𝑉 is the relative volume, 𝜐 is Poisson’s ratio, and 𝐶𝑖𝑗 is the
right Cauchy-Green strain:
𝜕𝑥𝑘 𝜕𝑥𝑘
C𝑖𝑗 = , (22.7.2)
𝜕𝑋𝑖 𝜕𝑋𝑗
after determining 𝑆𝑖𝑗 , it is transformed into the Cauchy stress tensor, 𝜎𝑖𝑗 :
𝜌 𝜕𝑥𝑖 𝜕𝑥𝑗
𝜎𝑖𝑗 = 𝑆 , (22.7.3)
𝜌0 𝜕𝑋𝑘 𝜕𝑋𝑙 𝑘𝑙
where 𝜌0 and 𝜌 are the initial and current density, respectively. The default value of υ is
0.463.
1−𝑉
𝐹2 = , (22.8.3)
1 − 𝑉CJ
where 𝑉CJ is the Chapman-Jouguet relative volume and 𝑡 is current time. If 𝐹 exceeds 1, it
is reset to 1. This calculation of the burn fraction usually requires several time steps for 𝐹 to
reach unity, thereby spreading the burn front over several elements. After reaching unity,
𝐹 is held constant. This burn fraction calculation is based on work by Wilkins [1964] and is
also discussed by Giroux [1973].
𝑠∗𝑛+1
𝑖𝑗 = 𝑠𝑛𝑖𝑗 + 𝑠𝑖𝑝 𝛺𝑝𝑗 + 𝑠𝑗𝑝 𝛺𝑝𝑖 + 2𝐺𝜀̇′𝑖𝑗 𝑑𝑡, (22.8.4)
where 𝐺 is the shear modulus, and 𝜀̇′𝑖𝑗 is the deviatoric strain rate. The von Mises yield
condition is given by:
𝜎𝑦2
𝜙 = 𝐽2 − , (22.8.5)
3
where the second stress invariant, 𝐽2 , is defined in terms of the deviatoric stress
components as
1
𝐽2 = 𝑠𝑖𝑗 𝑠𝑖𝑗 , (22.8.6)
2
and the yield stress is 𝜎𝑦 . If yielding has occurred, i.e., 𝜙 > 0, the deviatoric trial stress is
scaled to obtain the final deviatoric stress at time 𝑛 + 1:
𝜎𝑦 ∗𝑛+1
𝑠𝑛+1
𝑖𝑗 = 𝑠𝑖𝑗 , (22.8.7)
√3𝐽2
If 𝜙 ≤ 0, then
𝑠𝑛+1
𝑖𝑗 = 𝑠∗𝑛+1
𝑖𝑗 . (22.8.8)
Before detonation pressure is given by the expression
1
𝑝𝑛+1 = 𝐾 ( − 1). (22.8.9)
𝑉 𝑛+1
where K is the bulk modulus. Once the explosive material detonates:
𝑠𝑛+1
𝑖𝑗 = 0. (22.8.10)
and the material behaves like a gas.
The shadow burn option should be active when computing the lighting time if there
exist elements within the mesh for which there is no direct line of sight from the detonation
points. The shadow burn option is activated in the control section. The lighting time is
based on the shortest distance through the explosive material. If inert obstacles exist within
the explosive material, the lighting time will account for the extra time required for the
detonation wave to travel around the obstacles. The lighting times also automatically
accounts for variations in the detonation velocity if different explosives are used. No
additional input is required for the shadow option but care must be taken when setting up
the input. This option works for two and three-dimensional solid elements. It is
recommended that for best results:
1. Keep the explosive mesh as uniform as possible with elements of roughly the same
dimensions.
2. Inert obstacle such as wave shapers within the explosive must be somewhat larger
than the characteristic element dimension for the automatic tracking to function
properly. Generally, a factor of two should suffice. The characteristic element
dimension is found by checking all explosive elements for the largest diagonal
3. The detonation points should be either within or on the boundary of the explosive.
Offset points may fail to initiate the explosive.
4. Check the computed lighting times in the post processor LS-PrePost. The lighting
times may be displayed at time = 0, state 1, by plotting component 7 (a component
normally reserved for plastic strain) for the explosive material. The lighting times
are stored as negative numbers. The negative lighting time is replaced by the burn
fraction when the element ignites.
Sometimes it is advantageous to model contact surfaces via shell elements which are
not part of the structure, but are necessary to define areas of contact within nodal rigid
bodies or between nodal rigid bodies. Beams and shells that use this material type are
completely bypassed in the element processing. The Young’s modulus and Poisson’s ratio
are used only for setting the contact interface stiffnesses, and it is recommended that
reasonable values be input.
The pressure, 𝑝, deviatoric strain rate, 𝜀̇′𝑖𝑗 , deviatoric stress rate, 𝑠𝑖𝑗̇ , volumetric strain
rate, and 𝜀̇v , are defined in Equation (1.1.1):
1 1
𝑝 = − 𝜎𝑖𝑗 𝛿𝑖𝑗 𝜀̇′𝑖𝑗 = 𝜀̇𝑖𝑗 − 𝜀̇v
3 3
𝑠𝑖𝑗 = 𝜎𝑖𝑗 + 𝑝𝛿𝑖𝑗 𝜀̇v = 𝜀̇𝑖𝑗 𝛿𝑖𝑗 (22.10.1)
𝑠∇ ′ ′
𝑖𝑗 = 2𝜇𝜀̇𝑖𝑗 = 2𝐺𝜀̇𝑖𝑗 .
𝑠∇
𝑖𝑗 = 𝑠𝑖𝑗̇ − 𝑠𝑖𝑝 𝛺𝑝𝑗 − 𝑠𝑗𝑝 𝛺𝑝𝑖 . (22.10.2)
where the left superscript, *, denotes a trial stress value. The effective trial stress is defined
by
⁄2 1
3 (22.10.4)
𝑠 = ( ∗𝑠𝑛+1
∗ ∗ 𝑛+1
𝑠 ) ,
2 𝑖𝑗 𝑖𝑗
The plastic strain increment can be found by subtracting the deviatoric part of the
1 𝑛
strain increment that is elastic, 2𝐺 (𝑠𝑛+1
𝑖𝑗 − 𝑠R ′
𝑖𝑗 ), from the total deviatoric increment, Δε𝑖𝑗 , i.e.,
p 1 𝑛
Δ𝜀𝑖𝑗 = Δ𝜀′𝑖𝑗 − (𝑠𝑛+1
𝑖𝑗 − 𝑠𝑖𝑗R ). (22.10.7)
2𝐺
Recalling that,
𝑛
( ∗𝑠𝑛+1
𝑖𝑗 − 𝑠R
𝑖𝑗 )
′
Δ𝜀𝑖𝑗 = , (22.10.8)
2𝐺
and substituting Equation (22.10.8) into (22.10.7) we obtain,
p ( ∗𝑠𝑛+1
𝑖𝑗 − 𝑠𝑛+1
𝑖𝑗 )
Δ𝜀𝑖𝑗 = . (22.10.9)
2𝐺
Substituting Equation (22.10.6)
𝑠𝑛+1
𝑖𝑗 = 𝑚 ∗𝑠𝑛+1
𝑖𝑗 , (22.10.10)
into Equation (22.10.9) gives,
p 1 − 𝑚 ∗ 𝑛+1 1 − 𝑚 𝑛+1
Δ𝜀𝑖𝑗 = ( ) 𝑠𝑖𝑗 = 𝑠𝑖𝑗 = 𝑑λ𝑠𝑛+1
𝑖𝑗 . (22.10.11)
2𝐺 2𝐺𝑚
By definition an increment in effective plastic strain is
1
2 p p ⁄2
p (22.10.12)
Δ𝜀 = ( Δ𝜀𝑖𝑗 Δ𝜀𝑖𝑗 ) .
3
Squaring both sides of Equation (22.10.11) leads to:
p p 1 − 𝑚 2 ∗ 𝑛+1 ∗ 𝑛+1
Δ𝜀𝑖𝑗 Δ𝜀𝑖𝑗 = ( ) 𝑠𝑖𝑗 𝑠𝑖𝑗 (22.10.13)
2𝐺
or from Equations (22.10.4) and (22.10.12):
3 p2 1 − 𝑚 2 2 ∗2
Δ𝜀 = ( ) 𝑠 (22.10.14)
2 2𝐺 3
Hence,
∗
p1 − 𝑚 ∗ 𝑠 − 𝜎y
Δ𝜀 = 𝑠 = (22.10.15)
3𝐺 3𝐺
where we have substituted for m from Equation (22.10.6)
𝜎y
𝑚= ∗ (22.10.16)
𝑠
Thus,
p
(𝑠∗ − 𝜎y𝑛 )
Δ𝜀 = . (22.10.20)
(3𝐺 + 𝐸p )
The algorithm for plastic loading can now be outlined in five simple stress. If the
effective trial stress exceeds the yield stress then
p
(𝑠∗ − σy𝑛 )
Δ𝜀 = . (22.10.21)
(3𝐺 + 𝐸p )
Both the shear modulus 𝐺 and yield strength 𝜎y increase with pressure but decrease
with temperature. As a melt temperature is reached, these quantities approach zero. We
define the shear modulus before the material melts as
𝑓𝐸
1⁄ 𝐸 − 𝐸c −
(22.11.1)
𝐺 = 𝐺0 [1 + 𝑏𝑝𝑉 3 − ℎ( − 300)] 𝑒 𝐸m −𝐸 ,
3𝑅′
where 𝐺0 , 𝑏, ℎ, and 𝑓 are input parameters, 𝐸c is the cold compression energy:
𝑥
900𝑅′ exp(𝑎𝑥)
𝐸c (𝑋) = ∫ 𝑝𝑑𝑥 − , (22.11.2)
2(𝛾 −𝑎−1)
0 (1 − 𝑋) 𝑜 2
where,
𝑋 = 1 − 𝑉, (22.11.3)
and 𝐸m is the melting energy:
𝐸m (𝑋) = 𝐸c (𝑋) + 3𝑅′𝑇m (𝑋), (22.11.4)
which is a function of the melting temperature 𝑇m (𝑋):
𝑇mo exp(2𝑎𝑋)
𝑇m (𝑋) = , (22.11.5)
2(𝛾 −𝑎−1)
(1 − 𝑋) 𝑜 3
and the melting temperature 𝑇mo at 𝜌 = 𝜌0 . The constants 𝛾0 and a are input parameters.
In the above equation, 𝑅′ is defined by
𝑅𝜌0
, 𝑅′ = (22.11.6)
𝐴
where 𝑅 is the gas constant and A is the atomic weight. The yield strength 𝜎y is given by:
1 𝑓𝐸
𝐸 − 𝐸c −
(22.11.7)
𝜎y = 𝜎0′ [1 + 𝑏′𝑝𝑉 3 − ℎ( − 300)] 𝑒 𝐸m −𝐸 .
3𝑅′
where 𝛾1 is the initial plastic strain, and 𝑏′ and 𝜎0′ are input parameters. Where 𝜎0′ exceeds
𝜎max , the maximum permitted yield strength, 𝜎0′ is set to equal to 𝜎max . After the material
melts, 𝜎y and 𝐺 are set to zero.
Once the yield strength and shear modulus are known, the numerical treatment is
similar to that for material model 10.
p p p
where 𝑑𝜀eff = √23 𝑑𝜀𝑖𝑗 𝑑𝜀𝑖𝑗 , and the plastic tangent modulus is defined in terms of the input
tangent modulus, 𝐸t , as
𝐸𝐸t
𝐸p = . (22.12.5)
𝐸 − 𝐸t
This model is not recommended for shell elements. In the plane stress implementa-
tion, a one-step radial return approach is used to scale the Cauchy stress tensor to if the
state of stress exceeds the yield surface. This approach to plasticity leads to inaccurate shell
thickness updates and stresses after yielding. This is the only model in LS-DYNA for plane
stress that does not default to an iterative approach.
∗ 𝜀̅𝑝̇ 1
𝜀̇ = = effective plastic strain rate for 𝜀̇0 , in units of
𝜀̇0 [time]
𝑇 − 𝑇room
𝑇∗ =
𝑇melt − 𝑇room
Constants for a variety of materials are provided in Johnson and Cook [1983].
zero and no hydrostatic tension is permitted. If tensile pressures are calculated, they are
reset to 0 in the spalled material. Thus, the spalled material behaves as rubble. The
hydrostatic tension spall model detects spall if the pressure becomes more tensile than the
specified limit, 𝑝min . Once spall is detected, the deviatoric stresses are set to zero and the
pressure is required to be compressive. If hydrostatic tension is calculated then the
pressure is reset to 0 for that element.
In addition to the above failure criterion, this material model also supports a shell
element deletion criterion based on the maximum stable time step size for the element,
Δ𝑡max . Generally, Δ𝑡max goes down as the element becomes more distorted. To assure
stability of time integration, the global LS-DYNA time step is the minimum of the Δ𝑡max
values calculated for all elements in the model. Using this option allows the selective
deletion of elements whose time step Δ𝑡max has fallen below the specified minimum time
step, Δ𝑡crit . Elements which are severely distorted often indicate that material has failed
and supports little load, but these same elements may have very small time steps and
therefore control the cost of the analysis. This option allows these highly distorted
elements to be deleted from the calculation, and, therefore, the analysis can proceed at a
larger time step, and, thus, at a reduced cost. Deleted elements do not carry any load, and
are deleted from all applicable slide surface definitions. Clearly, this option must be
judiciously used to obtain accurate results at a minimum cost.
This model is well suited for implementing standard geologic models like the Mohr-
Coulomb yield surface with a Tresca limit, as shown in Figure 22.16.1. Examples of
converting conventional triaxial compression data to this type of model are found in (Desai
and Siriwardane, 1984). Note that under conventional triaxial compression conditions, the
LS-DYNA input corresponds to an ordinate of 𝜎1 − 𝜎3 rather than the more widely used
𝜎1 −𝜎3
2
, where 𝜎1 is the maximum principal stress and 𝜎3 is the minimum principal stress.
This material combined with equation-of-state type 9 (saturated) has been used very
successfully to model ground shocks and soil-structure interactions at pressures up to
100kbar.
To invoke Mode I of this model, set 𝑎0 , 𝑎1 , 𝑎2 , 𝑎0f , and 𝑎1f to zero, The tabulated
values of pressure should then be specified on cards 4 and 5, and the corresponding values
Mohr-Coulomb
Tresca
Friction Angle
Cohesion
The upper curve is best described as the maximum yield strength curve and the
lower curve is the material failure curve. There are a variety of ways of moving between
the two curves and each is discussed below.
Pressure
Figure 22.2. Two-curve concrete model with damage and failure.
curves until after 20 time steps the yield strength is defined by the failure curve.
Define 𝑎0 , 𝑎1 , 𝑎2 , 𝑎0f and 𝑎1f , and 𝑏1 . Cards 4 through 7 now give 𝜂 as a function of 𝜆
and scale the yield stress as
𝜎yield = 𝜎failed + 𝜂(𝜎max − 𝜎failed ), (22.16.4)
and then apply any tensile failure criteria.
1
2
⎛ 𝑓c′ ⎞3
𝜎cut = 1.7 ⎜ ⎟
⎝ −𝑎 0 ⎠
𝑓c′
𝑎0 =
4
1 (22.16.5)
𝑎1 =
3
1
𝑎2 = ′
3𝑓c
𝑎0f = 0
𝑎1f = 0.385
Note that these 𝑎0f and 𝑎1f defaults will be overwritten by non-zero entries on Card
3. If plastic strain or damage scaling is desired, Cards 5 through 8 and b1 should be
specified in the input. When 𝑎0 is input as a negative quantity, the equation-of-state can be
given as 0 and a trilinear EOS Type 8 model will be automatically generated from the
unconfined compressive strength and Poisson's ratio. The EOS 8 model is a simple
pressure versus volumetric strain model with no internal energy terms, and should give
reasonable results for pressures up to 5kbar (approximately 72,500 psi).
Mixture model
A reinforcement fraction, 𝑓r , can be defined along with properties of the reinforcing
material. The bulk modulus, shear modulus, and yield strength are then calculated from a
simple mixture rule, i.e., for the bulk modulus the rule gives:
𝐾 = (1 − 𝑓r )𝐾m + 𝑓r 𝐾r , (22.16.6)
where 𝐾m and 𝐾r are the bulk moduli for the geologic material and the reinforcing
material, respectively. This feature should be used with caution. It gives an isotropic effect
in the material instead of the true anisotropic material behavior. A reasonable approach
would be to use the mixture elements only where reinforcing material exists and plain
elements elsewhere. When the mixture model is being used, the strain rate multiplier for
the principal material is taken from load curve N1 and the multiplier for the reinforcement
is taken from load curve N2.
p p p
where 𝑑𝜀eff = √23 𝑑𝜀𝑖𝑗 𝑑𝜀𝑖𝑗 , and the plastic tangent modulus is defined in terms of the input
tangent modulus, 𝐸t , as
𝐸𝐸t
𝐸p = . (22.17.5)
𝐸 − 𝐸t
The oriented crack fracture model is based on a maximum principal stress criterion.
When the maximum principal stress exceeds the fracture stress, 𝜎f , the element fails on a
plane perpendicular to the direction of the maximum principal stress. The normal stress
and the two shear stresses on that plane are then reduced to zero. This stress reduction is
done according to a delay function that reduces the stresses gradually to zero over a small
number of time steps. This delay function procedure is used to reduce the ringing that may
otherwise be introduced into the system by the sudden fracture.
After a tensile fracture, the element will not support tensile stress on the fracture
plane, but in compression will support both normal and shear stresses. The orientation of
LS-DYNA DEV 10/27/16 (r:8004) 20-65 (Material Models)
Material Models LS-DYNA Theory Manual
this fracture surface is tracked throughout the deformation, and is updated to properly
model finite deformation effects. If the maximum principal stress subsequently exceeds the
fracture stress in another direction, the element fails isotropically. In this case the element
completely loses its ability to support any shear stress or hydrostatic tension, and only
compressive hydrostatic stress states are possible. Thus, once isotropic failure has
occurred, the material behaves like a fluid.
A parameter, SIGY, in the input governs how the strain to yield is identified. If
SIGY is set to zero, the strain to yield if found by solving for the intersection of the linearly
elastic loading equation with the strain hardening equation:
𝜎 = 𝐸𝜀,
(22.18.2)
𝜎 = 𝑘𝜀𝑛 ,
which gives the elastic strain at yield as:
1
𝐸 𝑛−1 (22.18.3)
𝜀yp =( ) .
𝑘
Strain rate is accounted for using the Cowper-Symonds model which scales the yield
stress with the factor
1
𝜀̇ ⁄P (22.18.5)
1+( ) ,
𝐶
where 𝜀̇ is the strain rate. A fully viscoplastic formulation is optional with this model
which incorporates the Cowper-Symonds formulation within the yield surface. An
additional cost is incurred but the improvement allows for dramatic results.
Both Young's modulus and the tangent modulus may optionally be made functions
of strain rate by specifying a load curve ID giving their values as a function of strain rate. If
these load curve ID's are input as 0, then the constant values specified in the input are used.
Note that all load curves used to define quantities as a function of strain rate must
have the same number of points at the same strain rate values. This requirement is used to
allow vectorized interpolation to enhance the execution speed of this constitutive model.
This model also contains a simple mechanism for modeling material failure. This
option is activated by specifying a load curve ID defining the effective stress at failure as a
function of strain rate. For solid elements, once the effective stress exceeds the failure stress
the element is deemed to have failed and is removed from the solution. For shell elements
the entire shell element is deemed to have failed if all integration points through the
thickness have an effective stress that exceeds the failure stress. After failure the shell
element is removed from the solution.
In addition to the above failure criterion, this material model also supports a shell
element deletion criterion based on the maximum stable time step size for the element,
Δ𝑡max . Generally, Δ𝑡max goes down as the element becomes more distorted. To assure
stability of time integration, the global LS-DYNA time step is the minimum of the Δ𝑡max
values calculated for all elements in the model. Using this option allows the selective
deletion of elements whose time step Δ𝑡max has fallen below the specified minimum time
step, Δ𝑡crit . Elements which are severely distorted often indicate that material has failed
and supports little load, but these same elements may have very small time steps and
therefore control the cost of the analysis. This option allows these highly distorted
20-68 (Material Models) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Material Models
elements to be deleted from the calculation, and, therefore, the analysis can proceed at a
larger time step, and, thus, at a reduced cost. Deleted elements do not carry any load, and
are deleted from all applicable slide surface definitions. Clearly, this option must be
judiciously used to obtain accurate results at a minimum cost.
Two unique rigid part IDs may not share common nodes unless they are merged
together using the rigid body merge option. A rigid body may be made up of disjoint finite
element meshes, however. LS-DYNA assumes this is the case since this is a common
practice in setting up tooling meshes in forming problems.
All elements which reference a given part ID corresponding to the rigid material
should be contiguous, but this is not a requirement. If two disjoint groups of elements on
opposite sides of a model are modeled as rigid, separate part ID's should be created for
each of the contiguous element groups if each group is to move independently. This
requirement arises from the fact that LS-DYNA internally computes the six rigid body
degrees-of-freedom for each rigid body (rigid material or set of merged materials), and if
disjoint groups of rigid elements use the same part ID, the disjoint groups will move
together as one rigid body.
Inertial properties for rigid materials may be defined in either of two ways. By
default, the inertial properties are calculated from the geometry of the constituent elements
of the rigid material and the density specified for the part ID. Alternatively, the inertial
properties and initial velocities for a rigid body may be directly defined, and this overrides
data calculated from the material property definition and nodal initial velocity definitions.
Young's modulus, E, and Poisson's ratio, υ are used for determining sliding interface
parameters if the rigid body interacts in a contact definition. Realistic values for these
constants should be defined since unrealistic values may contribute to numerical problem
in contact.
Since 𝐂l is symmetric
υ12 υ
= 21 , etc. (22.21.6)
𝐸11 𝐸22
After computing 𝑆𝑖𝑗 we use Equation (18.32) to obtain the Cauchy stress. This model
will predict realistic behavior for finite displacement and rotations as long as the strains are
small.
For shell elements, the stresses are integrated in time and are updated in the
corotational coordinate system. In this procedure the local material axes are assumed to
remain orthogonal in the deformed configuration. This assumption is valid if the strains
remain small.
The third equation defines the nonlinear shear stress parameter 𝛼. A fiber matrix
shearing term augments each damage mode:
2
𝜏12
+ 3 𝛼𝜏 4
2𝐺12 4 12
𝜏̅ = 2 , (22.22.2)
𝑆12 3
+ 𝛼𝑆4
2𝐺12 4 12
which is the ratio of the shear stress to the shear strength.
𝜎2 2 ⎡ 𝐶2 2 𝜎2
𝐹comp =( ) + ⎢( ) − 1⎤
⎥ + 𝜏̅ , (22.22.4)
2𝑆12 ⎣ 2𝑆12 ⎦ 𝐶2
where failure is assumed whenever 𝐹comb > 1. If 𝐹comb > 1, then the material constants 𝐸2 ,
𝜐1 , and 𝜐2 are set to zero.
Since 𝐂l is symmetric
𝜐12 𝜐21
= , etc. (22.23.6)
𝐸11 𝐸22
𝑛+1
𝜀𝑛+1 𝑛
𝑏𝑏 = 𝜀𝑏𝑏 + 𝛼𝑏 (𝑇
2 ) [𝑇 𝑛+1 − 𝑇 𝑛 ], (22.23.8)
𝑛+1
𝜀𝑛+1 𝑛
𝑐𝑐 = 𝜀𝑐𝑐 + 𝛼𝑐 (𝑇
2 ) [𝑇 𝑛+1 − 𝑇 𝑛 ].
After computing 𝑆𝑖𝑗 we use Equation (16.32) to obtain the Cauchy stress. This model
will predict realistic behavior for finite displacement and rotations as long as the strains are
small.
For shell elements, the stresses are integrated in time and are updated in the
corotational coordinate system. In this procedure the local material axes are assumed to
remain orthogonal in the deformed configuration. This assumption is valid if the strains
remain small.
In the implementation of this material model, the deviatoric stresses are updated
elastically (see material model 1), the yield function is checked, and if it is satisfied the
deviatoric stresses are accepted. If it is not, an increment in plastic strain is computed:
1⁄
(3 𝑠∗𝑖𝑗 𝑠∗𝑖𝑗 ) − 𝜎y
2
(22.24.4)
= 2
p
Δ𝜀eff ,
3𝐺 + 𝐸p
is the shear modulus and 𝐸p is the current plastic hardening modulus. The trial deviatoric
stress state 𝑠∗𝑖𝑗 is scaled back:
𝜎𝑦
𝑠𝑛+1
𝑖𝑗 = 𝑠∗ .
1⁄ 𝑖𝑗 (22.24.5)
(3 𝑠∗𝑖𝑗 𝑠∗𝑖𝑗 )
2
2
For shell elements, the above equations apply, but with the addition of an iterative
loop to solve for the normal strain increment, such that the stress component normal to the
mid surface of the shell element approaches zero.
1. Strain rate may be accounted for using the Cowper-Symonds model which scales
the yield stress with the factor
1
𝜀̇ ⁄𝑝 (22.24.6)
𝛽=1+( ) .
𝐶
where 𝜀̇ is the strain rate.
2. For complete generality a load curve, defining 𝛽, which scales the yield stress may
be input instead. In this curve the scale factor versus strain rate is defined.
3. If different stress versus strain curves can be provided for various strain rates, the
option using the reference to a table definition can be used. See Figure 19.24.1.
If a table ID is specified a curve ID is given for each strain rate, see Section 23.
Intermediate values are found by interpolating between curves. Effective plastic strain
versus yield stress is expected. If the strain rate values fall out of range, extrapolation is not
used; rather, either the first or last curve determines the yield stress depending on whether
the rate is low or high, respectively.
The cap model is formulated in terms of the invariants of the stress tensor. The
square root of the second invariant of the deviatoric stress tensor, √𝐽2D is found from the
deviatoric stresses 𝐒 as
1
√𝐽2D ≡ √ 𝑠𝑖𝑗 𝑠𝑖𝑗 , (22.25.1)
2
and is the objective scalar measure of the distortional or shearing stress. The first invariant
of the stress, 𝐽1 , is the trace of the stress tensor.
The cap model consists of three surfaces in √𝐽2D − 𝐽1 space, as shown in Figure
22.25.1. First, there is a failure envelope surface, denoted 𝑓1 in the figure. The functional
form of 𝑓1 is
𝑓1 = √𝐽2D − min(𝐹e (𝐽1 ), 𝑇mises ), (22.25.2)
where 𝐹e is given by
𝐹e (𝐽1 ) ≡ 𝛼 − 𝛾exp(−𝛽𝐽1 ) + 𝜃𝐽1 . (22.25.3)
5
4
3
2
σy 1
εpeff
Figure 22.25.1. Rate effects may be accounted for by defining a table of curves.
J 2D
J 2D = Fe
J2D = Fc
f1
f2
f3
J1
T O X( )
Figure 22.25.2. The yield surface of the two-invariant cap model in
pressure √𝐽2D − 𝐽1 space Surface 𝑓1 is the failure envelope, 𝑓2 is the cap surface,
and 𝑓3 is the tension cutoff.
and 𝑇mises ≡ |𝑋(𝜅𝑛 ) − 𝐿(𝜅𝑛 )|. This failure envelope surface is fixed in √𝐽2D − 𝐽1 space, and
therefore does not harden unless kinematic hardening is present. Next, there is a cap
surface, denoted 𝑓2 in the figure, with 𝑓2 given by
𝑓2 = √𝐽2D − 𝐹c (𝐽1 , 𝜅), (22.25.4)
where 𝐹c is defined by
1
𝐹c (𝐽1 , 𝜅) ≡ √[𝑋(𝜅) − 𝐿(𝜅)] 2 − [𝐽1 − 𝐿(𝜅)] 2 , (22.25.5)
𝑅
𝑋(𝜅) is the intersection of the cap surface with the 𝐽1 axis
𝑋(𝜅) = 𝜅 + 𝑅𝐹e (𝜅), (22.25.6)
and 𝐿(𝜅) is defined by
𝜅 𝜅>0
𝐿(𝜅) ≡ { . (22.25.7)
0 𝜅≤0
p
The hardening parameter 𝜅 is related to the plastic volume change 𝜀v through the
hardening law
p
𝜀v = W{1 − exp[−𝐷(𝑋(κ) − 𝑋0 )]}. (22.25.8)
failure envelope surface above, the tension cutoff surface on the left, and the cap surface on
the right.
An additive decomposition of the strain into elastic and plastic parts is assumed:
𝜀 = 𝜀 + 𝜀P , (22.25.10)
where 𝜀e is the elastic strain and 𝜀p is the plastic strain. Stress is found from the elastic
strain using Hooke’s law,
𝜎 = 𝐶(𝜀 − 𝜀P ), (22.25.11)
where 𝜎 is the stress and 𝐶 is the elastic constitutive tensor.
Associated plastic flow is assumed, so using Koiter’s flow rule the plastic strain rate
is given as the sum of contribution from all of the active surfaces,
3
∂𝑓𝑘
𝜀̇ = ∑ 𝜆̇ 𝑘
p
. (22.25.14)
𝑘=1
∂𝑠
One of the major advantages of the cap model over other classical pressure-
dependent plasticity models is the ability to control the amount of dilatency produced
under shear loading. Dilatency is produced under shear loading as a result of the yield
surface having a positive slope in √𝐽2D − 𝐽1 space, so the assumption of plastic flow in the
direction normal to the yield surface produces a plastic strain rate vector that has a
component in the volumetric (hydrostatic) direction (see Figure 22.25.1). In models such as
the Drucker-Prager and Mohr-Coulomb, this dilatency continues as long as shear loads are
applied, and in many cases produces far more dilatency than is experimentally observed in
material tests. In the cap model, when the failure surface is active, dilatency is produced
just as with the Drucker-Prager and Mohr-Columb models. However, the hardening law
permits the cap surface to contract until the cap intersects the failure envelope at the stress
point, and the cap remains at that point. The local normal to the yield surface is now
vertical, and therefore the normality rule assures that no further plastic volumetric strain
LS-DYNA DEV 10/27/16 (r:8004) 20-81 (Material Models)
Material Models LS-DYNA Theory Manual
(dilatency) is created. Adjustment of the parameters that control the rate of cap
contractions permits experimentally observed amounts of dilatency to be incorporated into
the cap model, thus producing a constitutive law which better represents the physics to be
modeled. Another advantage of the cap model over other models such as the Drucker-
Prager and Mohr-Coulomb is the ability to model plastic compaction. In these models all
purely volumetric response is elastic. In the cap model, volumetric response is elastic until
the stress point hits the cap surface. Therefore, plastic volumetric strain (compaction) is
generated at a rate controlled by the hardening law. Thus, in addition to controlling the
amount of dilatency, the introduction of the cap surface adds another experimentally
observed response characteristic of geological material into the model.
The cap model contains a number of parameters which must be chosen to represent
a particular material, and are generally based on experimental data. The parameters 𝛼, 𝛽,𝜃
and 𝛾 are usually evaluated by fitting a curve through failure data taken from a set of
triaxial compression tests. The parameters 𝑊, 𝐷, and X0 define the cap hardening law.
The value W represents the void fraction of the uncompressed sample and 𝐷 governs the
20-82 (Material Models) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Material Models
slope of the initial loading curve in hydrostatic compression. The value of R is the ration of
major to minor axes of the quarter ellipse defining the cap surface. Additional details and
guidelines for fitting the cap model to experimental data are found in [Chen and Baladi,
1985].
The behavior before compaction is orthotropic where the components of the stress
tensor are uncoupled, i.e., an 𝑎 component of strain will generate resistance in the local 𝑎
direction with no coupling to the local 𝑏 and 𝑐 directions. The elastic moduli vary linearly
with the relative volume from their initial values to the fully compacted values:
𝐸𝑎𝑎 = 𝐸𝑎𝑎u + 𝛽(𝐸 − 𝐸𝑎𝑎u ),
𝐸𝑏𝑏 = 𝐸𝑏𝑏u + 𝛽(𝐸 − 𝐸𝑏𝑏u ),
𝐸𝑐𝑐 = 𝐸𝑐𝑐u + 𝛽(𝐸 − 𝐸𝑐𝑐u ),
(22.26.1)
𝐺𝑎𝑏 = 𝐺𝑎𝑏u + 𝛽(𝐺 − 𝐺𝑎𝑏u ),
𝐺𝑏𝑐 = 𝐺𝑏𝑐u + 𝛽(𝐺 − 𝐺𝑏𝑐u ),
𝐺𝑐𝑎 = 𝐺𝑐𝑎u + 𝛽(𝐺 − 𝐺𝑐𝑎u ),
where
1 − 𝑉min
𝛽 = max [min ( , 1) ,0], (22.26.2)
1 − 𝑉𝑓
and 𝐺 is the elastic shear modulus for the fully compacted honeycomb material
𝐸
𝐺= . (22.26.3)
2(1 + 𝜈)
The relative volume V is defined as the ratio of the current volume over the initial
volume; typically, 𝑉 = 1 at the beginning of a calculation. The relative volume, 𝑉min , is the
minimum value reached during the calculation.
unloading and
reloading path
0 strain: -εij
Unloading is based on the interpolated Young’s
moduli which must provide an unloading
tangent that exceeds the loading tangent.
Figure 22.26.1. Stress quantity versus volumetric strain. Note that the “yield
stress” at a volumetric strain of zero is nonzero. In the load curve definition, the
“time” value is the volumetric strain and the “function” value is the yield stress.
The load curves define the magnitude of the average stress as the material changes
density (relative volume). Each curve related to this model must have the same number of
points and the same abscissa values. There are two ways to define these curves: as a
function of relative volume V, or as a function of volumetric strain defined as:
𝜀𝑉 = 1 − 𝑉. (22.26.4)
In the former, the first value in the curve should correspond to a value of relative
volume slightly less than the fully compacted value. In the latter, the first value in the
curve should be less than or equal to zero corresponding to tension and should increase to
full compaction. When defining the curves, care should be taken that the extrapolated
values do not lead to negative yield stresses.
At the beginning of the stress update we transform each element’s stresses and
strain rates into the local element coordinate system. For the uncompacted material, the
trial stress components are updated using the elastic interpolated moduli according to:
trial
𝑛+1 𝑛
𝜎𝑎𝑎 = 𝜎𝑎𝑎 + 𝐸𝑎𝑎 Δ𝜀𝑎𝑎 ,
trial
𝑛+1 𝑛
𝜎𝑏𝑏 = 𝜎𝑏𝑏 + 𝐸𝑏𝑏 Δ𝜀𝑏𝑏 ,
trial
𝑛+1 𝑛
𝜎𝑐𝑐 = 𝜎𝑐𝑐 + 𝐸𝑐𝑐 Δ𝜀𝑐𝑐 ,
(22.26.5)
𝑛+1trial 𝑛
𝜎𝑎𝑏 = 𝜎𝑎𝑏 + 2𝐺𝑎𝑏 Δ𝜀𝑎𝑏 ,
trial
𝑛+1 𝑛
𝜎𝑏𝑐 = 𝜎𝑏𝑐 + 2𝐺𝑏𝑐 Δ𝜀𝑏𝑐 ,
trial
𝑛+1 𝑛
𝜎𝑐𝑎 = 𝜎𝑐𝑎 + 2𝐺𝑐𝑎 Δ𝜀𝑐𝑎 = 1.
Then we independently check each component of the updated stresses to ensure that
they do not exceed the permissible values determined from the load curves, e.g., if
trial
∣𝜎𝑖𝑗𝑛+1 ∣ > 𝜆𝜎𝑖𝑗 (𝑉min ), (22.26.6)
then
trial
𝜆𝜎𝑖𝑗𝑛+1
𝜎𝑖𝑗𝑛+1 = 𝜎𝑖𝑗 (𝑉min ) trial
. (22.26.7)
∣𝜎𝑖𝑗𝑛+1 ∣
The parameter 𝜆 is either unity or a value taken from the load curve number, LCSR,
that defines 𝜆 as a function of strain rate. Strain rate is defined here as the Euclidean norm
of the deviatoric strain rate tensor.
For fully compacted material we assume that the material behavior is elastic-
perfectly plastic and updated the stress components according to
𝑛+1⁄2
𝑠trial
𝑖𝑗 = 𝑠𝑛𝑖𝑗 + 2𝐺Δ𝜀dev
𝑖𝑗 , (22.26.8)
We next check to see if the yield stress for the fully compacted material is exceeded
by comparing
1
3 trial trial ⁄2 (22.26.10)
𝑠trial
eff = ( 𝑠𝑖𝑗 𝑠𝑖𝑗 ) .
2
the effective trial stress, to the yield stress 𝜎y . If the effective trial stress exceeds the yield
stress, we simply scale back the stress components to the yield surface:
𝜎y
𝑠𝑛+1
𝑖𝑗 = trial 𝑠trial
𝑖𝑗 . (22.26.11)
𝑠eff
We can now update the pressure using the elastic bulk modulus, 𝐾:
𝑛+1⁄2
𝑝𝑛+1 = 𝑝𝑛 − 𝐾Δ𝜀𝑘𝑘 ,
(22.26.12)
𝐸
𝐾= ,
3(1 − 2𝜈)
and obtain the final value for the Cauchy stress
𝜎𝑖𝑗𝑛+1 = 𝑠𝑛+1
𝑖𝑗 − 𝑝𝑛+1 𝛿𝑖𝑗 . (22.26.13)
After completing the stress update, we transform the stresses back to the global
configuration.
A and B are user defined constants, whereas C and D are related to A and B as
follows
1
C = A + B,
2
(22.27.2)
A(5𝜐 − 2) + B(11𝜐 − 5)
D= .
2(1 − 2𝜐)
For small volumetric deformations the bulk modulus, 𝐾, can be defined as the ratio
of the pressure over the volumetric strain as the relative volume approaches unity:
𝑝
𝐾 = lim ( ). (22.27.7)
𝑉→1 𝑉 − 1
∂𝑊
= A,
∂𝐼1
𝜕𝑊
= B,
𝜕𝐼2
𝜕𝑊
= −2C𝐼3−3 + 2D(𝐼3 − 1) = −2C𝜆−18 + 2D(𝜆6 − 1), (22.27.8)
𝜕𝐼3
2
𝑝= 3
{A𝜆2 + 2𝜆4 B + 𝜆6 [−2C𝜆−18 + 2D(𝜆6 − 1)]}
𝜆
2
= 3 {A𝜆2 + 2𝜆4 B − 2C𝜆−12 + 2D(𝜆12 − 𝜆6 )}.
𝜆
In the limit as the stretch ratio approaches unity, the pressure must approach zero:
lim 𝑝 = 0. (22.27.9)
𝜆→1
Therefore, A + 2B − 2C = 0 and
C = 0.5A + B. (22.27.10)
We therefore obtain:
3 3 2𝐺(1 + 𝜐) 2(A + B)(1 + 𝜐)
14A + 32B + 12D = 𝐾 = ( )= . (22.27.12)
2 2 3(1 − 2𝜐) (1 − 2𝜐)
Solving for D we obtain the desired equation:
A(5𝜐 − 2) + B(11𝜐 − 5)
D= . (22.27.13)
2(1 − 2𝜐)
The second Piola-Kirchhoff stress tensor, S, is found by taking the partial derivative
of the strain energy function with respect to the Green-Lagrange strain tensor, E.
∂𝑊 ∂𝑊 ∂𝐼 ∂𝐼 2C ∂I
𝑆𝑖𝑗 = =2 = 2 [A 1 + B 2 + (2D(𝐼3 − 1) − 2 ) 3 ]. (22.27.15)
∂𝐸𝑖𝑗 ∂𝐶𝑖𝑗 ∂𝐶𝑖𝑗 ∂𝐶𝑖𝑗 𝐼3 ∂𝐶𝑖𝑗
Inserting Equation (22.27.16) into Equation (22.27.15) yields the following expression
for the second Piola-Kirchhoff stress:
1
𝑆𝑖𝑗 = 2A𝛿𝑖𝑗 + 2B(𝐼1 𝛿𝑖𝑗 − 𝐶𝑖𝑗 ) − 4C 𝐶𝑖𝑗−1 + 4D(𝐼3 − 1)I3 𝐶𝑖𝑗−1 . (22.27.17)
𝐼32
Equation (22.27.17) can be transformed into the Cauchy stress by using the push forward
operation
1
𝜎𝑖𝑗 = 𝐹𝑖𝑘 𝑆𝑘𝑙 𝐹𝑗𝑙 . (22.27.18)
𝐽
where 𝐽 = det(𝐹𝑖𝑗 ).
recapitulated in this section. When dealing with shell elements, the stress (as well as
constitutive matrix) is typically evaluated in corotational coordinates after which it is
transformed back to the standard basis according to
𝜎𝑖𝑗 = 𝑅𝑖𝑘 𝑅𝑗𝑙 𝜎̂ 𝑘𝑙 . (22.27.19)
Here 𝑅𝑖𝑗 is the rotation matrix containing the corotational basis vectors. The so-
called corotated stress 𝜎̂ 𝑖𝑗 is evaluated using Equation 19.27.21 with the exception that the
deformation gradient is expressed in the corotational coordinates, i.e.,
1
̂ 𝑆𝑘𝑙 𝐹𝑗𝑙̂ ,
𝜎̂ 𝑖𝑗 = 𝐹𝑖𝑘 (22.27.20)
𝐽
where 𝑆𝑖𝑗 is evaluated using Equation (22.27.17). The corotated deformation gradient is
incrementally updated with the aid of a time increment Δ𝑡, the corotated velocity gradient
̂𝑖𝑗 with which the embedded coordinate system is rotating.
𝐿̂𝑖𝑗 , and the angular velocity 𝛺
̂𝑖𝑘 )𝐹𝑘𝑗
𝐹𝑖𝑗̂ = (𝛿𝑖𝑘 + Δ𝑡𝐿̂𝑖𝑘 − Δ𝑡𝛺 ̂ . (22.27.21)
In LS-DYNA, α is given by
⎧ (0)
{ 𝜎̂ 33 (−1) (0)
{ (−1) (0)
∣𝜎̂ 33 − 𝜎̂ 33 ∣ ≥ 10−4
{𝜎̂ − 𝜎̂ 33
𝛼 = ⎨ 33 , (22.27.24)
{
{ −1 otherwise
{
⎩
and the stresses are determined from this value of α. Finally, to make sure that the normal
stress through the thickness vanishes, it is set to 0 (zero) before exiting the stress update
routine.
𝜕 ( 12 𝐶𝑖𝑗−1 )
𝐼3 2 −1 −1 1 −1 −1 (22.27.27)
= − 2 𝐶𝑘𝑙 𝐶𝑖𝑗 − 2 (𝐶𝑘𝑗 𝐶𝑖𝑙 + 𝐶𝑙𝑗−1 𝐶𝑖𝑘
−1
)
𝜕𝐶𝑘𝑙 𝐼3 2𝐼3
Since LS-DYNA needs the tangential modulus for the Cauchy stress, it is a good idea
to transform the terms in Equation (22.27.27) before summing them up. The push forward
pk
operation for the fourth-order tensor 𝐸𝑖𝑗𝑘𝑙 is
1
𝐸TC PK
𝑖𝑗𝑘𝑙 = 𝐹𝑖𝑎 𝐹𝑗𝑏 𝐹𝑘𝑐 𝐹𝑙𝑑 𝐸𝑎𝑏𝑐𝑑 . (22.27.29)
𝐽
Since the right Cauchy-Green tensor is 𝐂 = 𝐅T 𝐅 and the left Cauchy-Green tensor is
𝐛 = 𝐅T 𝐅, and the determinant and trace of the both stretches are equal, the transformation
is in practice carried out by interchanging
𝐶𝑖𝑗−1 → 𝛿𝑖𝑗 , (22.27.30)
1 4C
𝐽𝐸TC
𝑖𝑗𝑘𝑙 = 4B [𝑏𝑘𝑙 𝑏𝑖𝑗 − (𝑏𝑖𝑘 𝑏𝑗𝑙 + 𝑏𝑖𝑙 𝑏𝑗𝑘 )] + 2 [4𝛿𝑖𝑗 𝛿𝑘𝑙 + (𝛿𝑘𝑗 𝛿𝑖𝑙 + 𝛿𝑙𝑗 𝛿𝑖𝑚 )] +
2 I3
(22.27.32)
1
8D𝐼3 [(2𝐼3 − 1)𝛿𝑖𝑗 𝛿𝑘𝑙 − (𝐼3 − 1)(𝛿𝑘𝑗 𝛿𝑖𝑙 + 𝛿𝑙𝑗 𝛿𝑖𝑘 )].
2
Using this exact expression for the tangent stiffness matrix in the context of shell
elements is not adequate since it does not take into account that the normal stress is zero
and it must be modified appropriately. To this end, we assume that the tangent moduli in
̂ 𝑖𝑗 to the corotated rate
Equation (22.27.33) relates the corotated rate-of-deformation tensor 𝐷
•
of stress 𝜎̂ 𝑖𝑗 ,
𝜎̂ 𝑖𝑗• = 𝐸̂TC ̂
𝑖𝑗𝑘𝑙 𝐷𝑘𝑙 . (22.27.34)
Even though this is not completely true, we believe that attempting a more thorough
treatment would hardly be worth the effort. The objective can now be stated as to find a
TCalg
modified tangent stiffness matrix 𝐸̂ijkl such that
•alg TCalg ̂
𝜎̂ 𝑖𝑗 = 𝐸̂𝑖𝑗𝑘𝑙 𝐷𝑘𝑙 , (22.27.35)
alg
where 𝜎̂ 𝑖𝑗 is the stress as it is evaluated in LS-DYNA. The stress update, described in
Section 19.27.1, is performed in a rather ad hoc way which probably makes the stated
objective unachievable. Still we attempt to extract relevant information from it that enables
us to come somewhat close.
TCalg
To preclude the obvious singularity, a small positive value is assigned to 𝐸̂3333 ,
TCalg TCalg TCalg
𝐸̂3333 = 10−4 (∣𝐸̂1111 ∣ + ∣𝐸̂2222 ∣). (22.27.39)
In applying this model to shell elements the resultants are updated incrementally
using the midplane strains 𝜀m and curvatures 𝜅:
Δ𝑛 = Δ𝑡𝐶𝜀m (22.28.1)
ℎ3
Δ𝑚 = Δ𝑡 𝐶𝜅, (22.28.2)
12
where the plane stress constitutive matrix is given in terms of Young’s Modulus 𝐸 and
Poisson’s ratio 𝜈 as:
̅̅̅̅̅ = 𝑚2𝑥𝑥 − 𝑚𝑥𝑥 𝑚𝑦𝑦 + 𝑚2𝑦𝑦 + 3𝑚2𝑥𝑦 .
𝑚 (22.28.3)
Defining
𝑛̅ = 𝑛2𝑥𝑥 − 𝑛𝑥𝑥 𝑛𝑦𝑦 + 𝑛2𝑦𝑦 + 3𝑛2𝑥𝑦 , (22.28.4)
1 1
𝑚
̅̅̅̅̅𝑛̅ = 𝑚𝑥𝑥 𝑛𝑥𝑥 − 𝑚𝑥𝑥 𝑛𝑦𝑦 − 𝑛𝑥𝑥 𝑚𝑦𝑦 + 𝑚𝑦 𝑛𝑦 + 3𝑚𝑥𝑦 𝑛𝑥𝑦 , (22.28.6)
2 2
the Ilyushin yield function becomes
4|𝑚
̅̅̅̅̅𝑛̅| 16𝑚̅̅̅̅̅
𝑓 (𝑚, 𝑛) = 𝑛̅ + + 2
≤ 𝑛2y = ℎ2 𝜎y2 . (22.28.7)
ℎ √3 ℎ
LS-DYNA DEV 10/27/16 (r:8004) 20-95 (Material Models)
Material Models LS-DYNA Theory Manual
In our implementation we update the resultants elastically and check to see if the yield
condition is violated:
𝑓 (𝑚, 𝑛) > 𝑛2y . (22.28.8)
If so, the resultants are scaled by the factor 𝛼:
𝑛2y
𝛼=√ . (22.28.9)
𝑓 (𝑚, 𝑛)
We update the yield stress incrementally:
𝜎y𝑛+1 = 𝜎y𝑛 + 𝐸P Δ𝜀eff
plastic , (22.28.10)
√𝑓 (𝑚, 𝑛) − 𝑛y
Δ𝜀eff (22.28.11)
plastic = 𝑝 .
ℎ(3𝐺 + 𝐸 )
Kennedy, et. al., report that this model predicts results that may be too stiff; users of
this model should proceed cautiously.
In applying this material model to the Belytschko beam, the flow rule changes to
̂ 2𝑦 𝐴
4𝑚 ̂ 2𝑧 𝐴
4𝑚
𝑓 (𝑚, 𝑛) = 𝑓𝑥2̂ + + ≤ 𝑛2y = 𝐴2 𝜎y2 , (22.28.12)
3𝐼𝑦𝑦 3𝐼𝑧𝑧
have been updated elastically according to Equations (4.16)-(4.18). The yield condition is
checked with Equation (22.28.8), and if it is violated, the resultants are scaled as described
above.
This model is frequently applied to beams with non-rectangular cross sections. The
accuracy of the results obtained should be viewed with some healthy suspicion. No work
hardening is available with this model.
Axial collapse occurs when the compressive axial load reaches the collapse load.
The collapse load-versus-collapse deflection is specified in the form of a load curve. The
points of the load curve are (true strain, collapse force). Both quantities should be entered as
positive for all points, and will be interpreted as compressive i.e., collapse does not occur in
tension. The first point should be the pair (zero, initial collapse load).
The collapse load may vary with end moment and with deflection. In this case,
several load-deflection curves are defined, each corresponding to a different end moment.
Each load curve should have the same number of point pairs and the same deflection
values. The end moment is defined as the average of the absolute moments at each end of
the beam, and is always positive. It is not possible to make the plastic moment vary with
axial load.
M8
M7
M6
M5
M4
Axial Force
M3
M2
M1
Figure 22.29.1. The force magnitude is limited by the applied end moment. For
an intermediate value of the end moment, LS-DYNA interpolates between the
curves to determine the allowable force.
𝐸 = Young′ smodulus
𝐺 = Shear modulus
𝐴 = Cross sectional area
𝐴s = Effective area in shear
𝑙𝑛 = Reference length of beam
𝑙𝑛+1 = Current length of beam
(22.29.1)
𝐼𝑦𝑦 = Second moment of inertia about 𝑦
𝐼𝑧𝑧 = Second moment of inertia about 𝑧
𝐽 = Polar moment of inertia
𝑒𝑖 = 𝑖th local base vector in the current configuration
𝑦𝐼 = nodal vector in y direction at node I in the current configuration
𝑧𝐼 = nodal vector in z direction at node I in the current configuration
We emphasize that the local 𝑦 and 𝑧 base vectors in the reference configuration
always coincide with the corresponding nodal vectors. The nodal vectors in the current
configuration are updated using the Hughes-Winget formula while the base vectors are
computed from the current geometry of the element and the current nodal vectors.
12𝐸𝐼∗∗
𝜑∗ = (22.29.9)
𝐺𝐴s ln ln
Plastic Correction
After the elastic update the state of force is checked for yielding as follows. As a
preliminary note we emphasize that whenever yielding does not occur the elastic
stiffnesses and forces are taken as the new stiffnesses and forces.
⎛
⎜ el
⎞
⎟
el ⎜
⎜ 𝐴 𝑖(𝐼𝐼)
⎟
⎟
𝐴𝑛+1
𝑖(II) = 𝐴𝑖(𝐼𝐼) ⎜
⎜
⎜ 1 − 𝛼 Y ⎟
⎟,
⎟ (22.29.14)
⎜
⎜ ∂𝑚 𝑖𝐼 ⎟⎟
max (0.001, 𝐴el
𝑖(𝐼𝐼) + )
⎝ ∂𝜃𝑖𝐼P ⎠
where α ≤ 1 is a parameter chosen such that the moment-rotation matrix remains positive
definite.
⎛
⎜ ⎞
⎟
⎜
⎜ 𝐾tel ⎟
⎟
𝐾tn+1 = 𝐾tel ⎜
⎜1
⎜ −α Y
⎟
⎟,
⎟ (22.29.17)
⎜
⎜ ∂𝑚 t ⎟
𝐾t + P ⎟
el
⎝ ∂𝜃t ⎠
where again α ≤ 1 is chosen so that the stiffness is positive.
Axial collapse is modeled by limiting the axial force by 𝑓aY (𝜀, 𝑚), i.e., a function of the
axial strains and the magnitude of bending moments. If the axial elastic force exceeds this
value it is reduced to yield
𝑓a𝑛+1 = 𝑓aY (𝜀𝑛+1 , 𝑚𝑛+1 )sgn(𝑓ael ), (22.29.18)
and the axial stiffness is given by
Y
el ∂𝑓a
𝐾a𝑛+1 = max (0.05𝐾a , ). (22.29.19)
∂𝜀
We neglect the influence of change in bending moments when computing this parameter.
Damping
Damping is introduced by adding a viscous term to the internal force on the form
δ
𝑑⎡ 𝜃t ⎤
𝐟v = 𝐃 ⎢ ⎢ ⎥,
⎥ (22.29.20)
𝑑𝑡 ⎢𝜃𝑦 ⎥
⎣𝜃𝑧 ⎦
el
⎡𝐾a ⎤
⎢ 𝐾tel ⎥
𝐃 = 𝛾⎢
⎢
⎥,
⎥ (22.29.21)
⎢ 𝐴el
𝑦 ⎥
⎣ 𝐴el
𝑧 ⎦
Transformation
The internal force vector in the global system is obtained through the transformation
𝐟g𝑛+1 = 𝐒𝐟l𝑛+1 , (22.29.22)
where
−𝑒 0 −𝑒3 /𝑙𝑛+1 −𝑒3 /𝑙𝑛+1 𝑒2 /𝑙𝑛+1 𝑒2 /𝑙𝑛+1
⎡ 1 ⎤
⎢ 0 −𝑒1 𝑒2 0 𝑒3 0 ⎥
𝐒=⎢ 𝑛+1 ⎥
, (22.29.23)
⎢ 𝑒1 0 𝑒3 /𝑙𝑛+1 𝑒3 /𝑙𝑛+1 −𝑒2 /𝑙𝑛+1 −𝑒2 /𝑙 ⎥
⎣ 0 𝑒1 0 𝑒2 0 𝑒3 ⎦
f 𝑛+1
⎡ a 𝑛+1 ⎤
⎢ 𝑚t ⎥
𝐟l𝑛+1 =⎢
⎢𝑚𝑛+1
⎥.
⎥ (22.29.24)
⎢ y ⎥
⎣𝑚𝑛+1z ⎦
There are two contributions to the tangent stiffness, one geometrical and one
material contribution. The geometrical contribution is given (approximately) by
1
𝐊geo = 𝐑(𝐟l𝑛+1 ⊗ 𝐈)𝐖 − 𝐓𝐟l𝑛+1 𝐋, (22.29.28)
l𝑛+1 l𝑛+1
where
𝑅 0 𝑅3 /𝑙𝑛+1 𝑅3 /𝑙𝑛+1 −𝑅2 /𝑙𝑛+1 −𝑅2 /𝑙𝑛+1
⎡ 1 ⎤
⎢ 0 𝑅1 −𝑅2 0 −𝑅3 0 ⎥
𝐑=⎢ 𝑛+1 ⎥
, (22.29.29)
⎢−𝑅1 0 −𝑅3 /𝑙𝑛+1 −𝑅3 /𝑙𝑛+1 𝑅2 /𝑙𝑛+1 𝑅2 /𝑙 ⎥
⎣ 0 −𝑅1 0 −𝑅2 0 −𝑅3 ⎦
0 0 −𝑒3 −𝑒3 𝑒2 𝑒2
⎡0 0 0 0 0 0 ⎤
𝐓=⎢
⎢0
⎥, (22.29.31)
0 𝑒3 𝑒3 −𝑒2 −𝑒2 ⎥
⎣0 0 0 0 0 0 ⎦
Here 𝐾 and 𝐺 are bulk and shear modulii, 𝜃 and e are volumetric and shear
logarithmic strains and 𝛼 and 𝜀L are constant material parameters. There is an option to
define the bulk and shear modulii as functions of the martensite fraction according to
𝐾 = 𝐾A + 𝜉S (𝐾S − 𝐾A ),
(22.30.4)
𝐺 = 𝐺A + 𝜉S (𝐺S − 𝐺A ),
in case the stiffness of the martensite differs from that of the austenite. Furthermore, the
unit vector 𝐧 is defined as
𝐧 = 𝐞/(‖𝐞‖ + 10−12 ), (22.30.5)
and a loading function is introduced as
𝐹 = 2𝐺‖𝐞‖ + 3𝛼𝐾𝜃 − 𝛽𝜉S , (22.30.6)
where
𝛽 = (2𝐺 + 9𝛼2 𝐾)𝜀L . (22.30.7)
For the evolution of the martensite fraction 𝜉S in the material, the following rule is
adopted
𝐹 − 𝑅ASs > 0⎫
} 𝐹̇
𝐹̇ > 0 ⎬ ⇒ 𝜉Ṡ = −(1 − 𝜉S ) (22.30.8)
𝜉S < 1 } 𝐹 − 𝑅AS
⎭ f
𝐹 − 𝑅SAs < 0⎫
} 𝐹̇
𝐹̇ < 0 ⎬ ⇒ 𝜉Ṡ = 𝜉S . (22.30.9)
𝜉S > 0 } 𝐹 − 𝑅SA
⎭ f
AS SA
Here 𝑅AS SA
s , 𝑅f , 𝑅s and 𝑅f are constant material parameters. The Cauchy stress is
finally obtained as
𝛕
𝛔= , (22.30.10)
𝐽
where 𝐽 is the Jacobian of the deformation.
⎛𝜆𝑖 ⎞ (22.30.11)
𝑒𝑖 = log ⎜
⎜
⎜ 1⎟
⎟,
⎟
⎝𝐽 3 ⎠
where
𝐽 = 𝜆1 𝜆2 𝜆3 . (22.30.12)
is the total Jacobian of the deformation. Using Equation (22.30.6) with 𝜉S = 𝜉S𝑛 , a value𝐹trial
of the loading function can be computed. The discrete counterpart of Equation (22.30.8)
becomes
𝐹trial − 𝑅AS
s > 0⎫
}
𝐹 trial
− 𝐹 > 0 ⎬ ⇒ Δ𝜉S
n
𝜉Sn < 1 }
⎭
(22.30.13)
trial AS
𝐹 − 𝛽Δ𝜉S − min(max(𝐹 , 𝑅AS
𝑛
s ), 𝑅f )
= −(1 − 𝜉S𝑛 − Δ𝜉S )
𝐹trial − 𝛽Δ𝜉S − 𝑅AS
f
𝐹trial − 𝑅SA
s < 0⎫ 𝐹trial − 𝛽Δ𝜉S − min(max(𝐹n , 𝑅SA SA
} 𝑛 f ) , 𝑅s )
𝐹 trial n
−𝐹 <0 ⎬ ⇒ Δ𝜉S = (𝜉S + Δ𝜉𝑆 ) . (22.30.14)
n } 𝐹trial − 𝛽Δ𝜉S − 𝑅SA
f
𝜉S > 0 ⎭
If none of the two conditions to the left are satisfied, set 𝜉S𝑛+1 = 𝜉S𝑛 , 𝐹𝑛+1 = 𝐹trial and
compute the stress 𝜎 𝑛+1 using Equations (22.30.1), (22.30.2), (22.30.5), (22.30.10) and 𝜉S =
𝜉S𝑛 . When phase transformation occurs according to a condition to the left, the
corresponding equation to the right is solved for Δ𝜉S . If the bulk and shear modulii are
constant this is an easy task. Otherwise 𝐹trial as well as 𝛽 depends on this parameter and
makes things a bit more tricky. We have that
𝐸 S − 𝐸A
𝐹trial = 𝐹ntrial (1 + Δ𝜉S ) ,
𝐸n
(22.30.15)
𝐸 − 𝐸A
𝛽 = 𝛽n (1 + S Δ𝜉S ),
𝐸n
where 𝐸S and 𝐸A are Young’s modulii for martensite and austenite, respectively. The
subscript 𝑛 is introduced for constant quantities evaluated at time 𝑡𝑛 . To simplify the
upcoming expressions, these relations are written
𝐹trial = 𝐹ntrial + Δ𝐹trial Δ𝜉S ,
(22.30.16)
𝛽 = 𝛽n + Δ𝛽Δ𝜉S .
and
̃𝑛 − 𝑅SA
𝑓 (Δ𝜉S ) = Δ𝛽𝜉S𝑛 Δ𝜉S2 + (𝐹SA f + (𝛽𝑛 − Δ𝐹
trial 𝑛
)𝜉S )Δ𝜉S +
𝑛 ̃𝑛 trial
(22.30.18)
𝜉S (𝐹SA − 𝐹𝑛 ) = 0.
respectively, where we have for simplicity set
̃𝑛 = min(max(𝐹𝑛 , 𝑅AS AS
𝐹AS s ) , 𝑅f ) ,
(22.30.19)
̃𝑛 = min(max(𝐹𝑛 , 𝑅SA
𝐹SA SA
f ), 𝑅s ).
The solutions to these equations are approximated with two Newton iterations
starting in the point Δ𝜉S = 0. Now set 𝜉S𝑛+1 = min(1, max(0, 𝜉S𝑛 + Δ𝜉S )) and compute 𝜎 𝑛+1
and 𝐹𝑛+1 according to Equations (22.30.1), (22.30.2), (22.30.5), (22.30.6), (22.30.10) and 𝜉S =
𝜉S𝑛+1 .
𝐹trial − 𝑅SA
s < 0⎫}
𝐹trial − 𝐹𝑛 < 0 ⎬ ⇒ 𝐻 SA = 1, (22.30.24)
𝜉S𝑛 + Δ𝜉S ≥ 0 } ⎭
using the quantities computed in the previous stress update. For the variation of the
martensite fraction we take variations of Equations (22.30.17) and (22.30.18) with
δ𝐹ntrial = 2𝐺𝐧: δ𝐞 + 3𝛼𝐾δ𝜃, (22.30.25)
which results in
δ𝜉S = γ(2𝐺𝐧: δ𝐞 + 3𝛼𝐾δ𝜃), (22.30.26)
where
(1 − 𝜉S𝑛 )𝐻 AS 𝜉S𝑛 𝐻 SA
𝛾= + . (22.30.27)
𝑅AS ̃𝑛 trial 𝑛 ̃𝑛 − 𝑅SA + (𝛽𝑛 − Δ𝐹𝑛trial )𝜉S𝑛
f − 𝐹AS + (𝛽𝑛 − Δ𝐹𝑛 )(1 − 𝜉S ) 𝐹SA f
As can be seen, we use the value of 𝛾 obtained in the previous stress update since
this is easier to implement and will probably give a good indication of the current value of
this parameter.
δ𝜃 = 𝐢: δ𝛆, (22.30.30)
results in
𝜉S 𝜀L
δ𝛕 = {2𝐺 (1 − ) 𝐈dev
‖𝐞‖ + 10−12
𝜉S 𝜀L
+ 2𝐺 [ − 2𝐺𝛾𝜀L + 2𝛾(𝐺S − 𝐺A )(‖𝐞‖ − 𝜉S 𝜀L )] 𝐧 ⊗ 𝐧
‖𝐞‖ + 10−12
where 𝐈dev is the fourth order deviatoric identity tensor. In general this tangent is not
symmetric because of the terms on the second line in the expression above. We simply use
a symmetrization of the tangent stiffness above in the implementation. Furthermore, we
transform the tangent to a tangent closer related to the one that should be used in the LS-
DYNA implementation,
𝜉S 𝜀L
𝐂 = 𝐽 −1 {2𝐺 (1 − ) 𝐈dev
‖𝐞‖ + 10−12
𝜉S 𝜀L
+ 2G [ − 2𝐺𝛾𝜀L + 2𝛾(𝐺S − 𝐺A )(‖𝐞‖ − 𝜉S 𝜀L )] 𝐧 ⊗ 𝐧} δ𝜀.
‖𝐞‖ + 10−12
The strain energy functional, 𝑈, is defined in terms of the input constants as:
𝑈 = C100 𝐼1 + C200 𝐼12 + C300 𝐼13 + C400 𝐼14 + C110 𝐼1 𝐼2 +
(22.31.1)
C210 𝐼12 𝐼2 + C010 𝐼2 + C020 𝐼22 + 𝑓 (𝐽),
where the strain invariants can be expressed in terms of the deformation gradient matrix,
𝐹𝑖𝑗 , and the Green-St. Venant strain tensor, 𝐸𝑖𝑗 :
𝐽 = ∣𝐹𝑖𝑗 ∣
𝐼1 = 𝐸𝑖𝑖 (22.31.2)
1 𝑖𝑗
𝐼2 = 𝛿𝑝𝑞 𝐸𝑝𝑖 𝐸𝑞𝑗 .
2!
The glass layers are modeled by isotropic hardening plasticity with failure based on
exceeding a specified level of plastic strain. Glass is quite brittle and cannot withstand
large strains before failing. Plastic strain was chosen for failure since it increases
monotonically and, therefore, is insensitive to spurious numerical noise in the solution.
The material to which the glass is bonded is assumed to stretch plastically without
failure. The user defined integration rule option must be used with this material. The user
defined rule specifies the thickness of the layers making up the safety glass. Each
integration point is flagged with a zero if the layer is glass and with a one if the layer is
polymer.
An iterative plane stress plasticity algorithm is used to enforce the plane stress
condition.
This model was developed by Barlat, Lege, and Brem [1991] for modeling material
behavior in forming processes. The finite element implementation of this model is
described in detail by Chung and Shah [1992] and is used here.
If the reference configuration of the airbag is taken as the folded configuration, the
geometrical accuracy of the deployed bag will be affected by both the stretching and the
compression of elements during the folding process. Such element distortions are very
difficult to avoid in a folded bag. By reading in a reference configuration such as the final
unstretched configuration of a deployed bag, any distortions in the initial geometry of the
folded bag will have no effect on the final geometry of the inflated bag. This is because the
stresses depend only on the deformation gradient matrix:
𝜕𝑥𝑖
𝐹𝑖𝑗 = , (22.34.1)
𝜕𝑋𝑗
where the choice of 𝑋𝑗 may coincide with the folded or unfold configurations. It is this
unfolded configuration which may be specified here. When the reference geometry is used
then the no-compression option should be active. With the reference geometry it is
possible to shrink the airbag and then perform the inflation. Although the elements in the
shrunken bag are very small, the time step can be based on the reference geometry so a
very reasonable time step size is obtained. The reference geometry based time step size is
optional in the input.
The parameters fabric leakage coefficient, FLC, fabric area coefficient, FAC, and
effective leakage area, ELA, for the fabric in contact with the structure are optional for the
Wang-Nefske and hybrid inflation models. It is possible for the airbag to be constructed of
multiple fabrics having different values of porosity and permeability. The gas leakage
through the airbag fabric then requires an accurate determination of the areas by part ID
available for leakage. The leakage area may change over time due to stretching of the
airbag fabric or blockage when the outer surface of the bag is in contact with the structure.
LS-DYNA can check the interaction of the bag with the structure and split the areas into
regions that are blocked and unblocked depending on whether the regions are in contact or
not, respectively. Typically, the parameters, FLC and FAC, must be determined
experimentally and their variation with time and pressure are optional inputs that allow
for maximum modeling flexibility.
In the cited work by Johnson and Bammann, they conclude that the Green-Naghdi
stress rate is to be preferred over all other proposed stress rates, including the most widely
used Jaumann rate, because the Green-Naghdi stress rate is based on the notions of
invariance under superimposed rigid-body motions. However, implementation of the
Green-Naghdi stress rate comes at a significant computational cost compared to the
Jaumann stress rate, e.g., see the discussion in this manual in the section entitled Green-
Naghdi Stress Rate.
2 The results of a simple shear simulation, monotonically increasing shear deformation, produce sinusoidal
stress response.
20-114 (Material Models) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Material Models
𝑑𝑊
𝑑𝜀
/𝑊
𝑅= 𝑑𝑇
.
𝑑𝜀
/𝑇
∂𝑓 𝜕𝑓
⎛
⎜ (𝛔)⎞⎟ ⎛
⎜ (𝜎 , 𝜎 , 𝜎 )⎞
⎜
⎜ ∂𝜎11 ⎟
⎟ ⎜
⎜ 𝜕𝜎11 11 22 12 ⎟ ⎟
⎟
⎜
⎜ ∂𝑓 ⎟
⎟ ⎜
⎜ 𝜕𝑓 ⎟
⎟
⎜
⎜ ⎟
⎟
(𝛔)⎟ ⎜ ⎜
⎜ (𝜎11 , 𝜎22 , 𝜎12 )⎟⎟
⎜
⎜ ∂𝜎22 ⎟ ⎜ 𝜕𝜎22 ⎟
⎟
⎜
⎜ ⎟
⎟ ⎜
⎜ ⎟
⎟
∂𝑓 ⎜
⎜ ∂𝑓 ⎟
⎟ ⎜
⎜ 𝜕𝑓 ⎟
⎟
(𝛔) = ⎜
⎜
⎜ (𝛔) ⎟
⎟
⎟ = ⎜
⎜
⎜ (𝜎11 , 𝜎22 , 𝜎12 ) ⎟
⎟
⎟ , (22.36.3)
∂𝛔 ⎜ ∂𝜎12 ⎟ ⎜ 𝜕𝜎12 ⎟
⎜
⎜ ⎟
⎟ ⎜
⎜ ⎟
⎟
⎜
⎜ ⎟
⎟ ⎜
⎜ ⎟
⎟
⎜
⎜ 0 ⎟
⎟ ⎜
⎜ 0 ⎟
⎟
⎜
⎜ ⎟
⎟ ⎜
⎜ ⎟
⎟
⎜
⎜ ⎟
⎟ ⎜
⎜ ⎟
⎟
⎜ 0 ⎟ ⎜ 0 ⎟
⎝ ⎠ ⎝ ⎠
where
𝜕𝑓 ∂𝜎 𝜎 (𝜎 , 𝜎 , 𝜎 )1−𝑚
(𝜎11 , 𝜎22 , 𝜎12 ) = eff (𝜎11 , 𝜎22 , 𝜎12 ) = eff 11 22 12 ⋅
𝜕𝜎11 ∂𝜎11 2
1 𝜎 − ℎ𝜎22
{𝑎(𝐾1 − 𝐾2 )|𝐾1 − 𝐾2 |𝑚−2 ( − 11 )+
2 4𝐾2
(22.36.4)
1 𝜎 − ℎ𝜎22
𝑎(𝐾1 + 𝐾2 )|𝐾1 + 𝐾2 |𝑚−2 ( + 11 )+
2 4𝐾2
𝜎 − ℎ𝜎22
𝑐2𝑚 𝐾2𝑚−1 11 },
4𝐾2
1 𝜈
⎛
⎜ 𝜈 1 ⎞
⎟ p
𝜀̇11 − 𝜀̇11
𝜎̇ 11 ⎜
⎜ ⎟
⎟ ⎛ ⎞
⎜ 1−𝜈 ⎟ ⎜
⎜ p ⎟
⎛
⎜𝜎̇ 22 ⎞
⎟ ⎜ ⎟ ⎜ 𝜀̇22 − 𝜀̇22 ⎟ ⎟
⎜
⎜ ⎟
⎟ 𝐸 ⎜ ⎜
⎜ 2
⎟
⎟
⎟
⎜
⎜
⎜ p ⎟
⎟
⎟
𝛔̇ = ⎜
⎜
⎜𝜎̇ 12 ⎟
⎟
⎟ = ⎜
⎜ ⎟
⎟ ⎜
⎜ 2(𝜀̇12 − 𝜀̇ )
12 ⎟⎟
⎜ ⎟ 1 − ν 2 ⎜ 1−𝜈 ⎟ ⎜ p ⎟
⎜𝜎̇ 23 ⎟ ⎜
⎜ ⎟
⎟ ⎜
⎜ 2(𝜀̇ − 𝜀̇ ) ⎟ (22.36.7)
⎜
⎜ 2 ⎟
⎟ ⎜ 23 23 ⎟⎟
⎝𝜎̇ 13 ⎠ ⎜
⎜ 1 − 𝜈⎟
⎟ p
⎝2(𝜀̇13 − 𝜀̇13 )⎠
⎝ 2 ⎠
= 𝐂ps (𝛆̇ − 𝛆̇p ).
where 𝐸 is the Young’s modulus, 𝜈 is the Poisson’s ratio and 𝐂ps denotes the plane stress
elastic tangential stiffness matrix. The associative flow rule for the plastic strain can be
written
∂𝑓
𝛆̇p = λ̇ , (22.36.8)
∂𝛔
and the consistency condition results in
∂𝑓 T ∂𝑓
𝛔̇ + 𝛆̇ = 0. (22.36.9)
∂𝛔 ∂𝜀p p
For algorithmic consistency, the effective plastic strain rate is defined as 𝜀̇p = 𝜆̇ .
∂𝑓
Multiplying Equation (22.36.7) with ∂𝛔 and using Equation (22.36.8) and Equation (22.36.9)
gives
T
∂𝑓
𝐂ps 𝛆̇
λ̇ = ∂𝛔 .
T (22.36.10)
∂𝑓 ps ∂𝑓 ∂𝑓
𝐂 −
∂𝛔 ∂𝛔 ∂𝜀p
Inserting
T
∂𝑓
𝐂ps 𝛆̇ ∂𝑓
𝛆̇p = ∂𝛔 ,
T ∂𝛔 (22.36.11)
∂𝑓 ∂𝑓 ∂𝑓
𝐂ps −
∂𝛔 ∂𝛔 ∂𝜀p
into Equation (22.36.7) results in
T
⎛ ∂𝑓 ∂𝑓 ⎞
⎜
⎜
⎜ {𝐂ps } {𝐂ps } ⎟
⎟
⎟
∂𝛔 ∂𝛔
𝛔̇ = ⎜
⎜
⎜ 𝐂ps − ⎟
⎟
⎟ 𝛆̇. (22.36.12)
⎜
⎜ ∂𝑓
T
∂𝑓 ∂𝑓 ⎟
⎟
⎜ 𝐂ps − ⎟
⎝ ∂𝛔 ∂𝛔 ∂𝜀p ⎠
To get the elastic-plastic tangent stiffness tensor in the element coordinate system it
needs to be transformed back. Since the elastic tangential stiffness tensor is isotropic with
respect to the axis of rotation, the plastic tangent stiffness tensor can be written
⎛
⎜ ps ∂𝑓 ps ∂𝑓 ⎞
⎜ {𝐐𝐂 } {𝐐𝐂 }⎟
ps ⎜
⎜ ps ∂𝛔 ∂𝛔 ⎟
⎟
⎟
𝐂plastic =⎜
⎜ 𝐂 − ⎟
⎟ , (22.36.13)
⎜
⎜ ∂𝑓
T
∂𝑓 ∂𝑓 ⎟
⎟
⎜ 𝐂ps − ⎟
⎝ ∂𝛔 ∂𝛔 ∂𝜀 p ⎠
where 𝐐 is the rotation matrix in Voigt form.
𝜎22 𝒏(𝑅𝜑 ) 𝜕𝑓 𝒏(𝑅45 ) 𝜎𝑌
𝜕𝝈 𝜎𝑌90 (𝜀𝑝 )
𝜕𝑓
𝜕𝑓 𝜎𝑌45 (𝜀𝑝 )
𝜕𝝈
𝜕𝝈
𝜎𝑌00 (𝜀𝑝 )
𝒏(𝑅90 )
𝑓 (𝝈, 𝜀𝑝 ) ≤ 0
𝜎𝑌 (𝜀𝑝 )
𝛼00 , 𝛼45 , 𝛼90
changes
𝜕𝑓 𝜎11 𝜀𝑝
𝒏(𝑅00 ) 𝜕𝝈
20-1 Plastic flow direction (left) and hardening (right) illustrated for variable
R-values and hardening. Changes in 𝛼00 , 𝛼45 and 𝛼90 come from changes in
𝜎eff = 𝑘𝜑 𝜎𝜑
where 𝑘𝜑 = 1 if 𝜑 = 0 but not in general. The plastic work relation, which defines the
effective plastic strain for the current material, gives the following expression for the
effective plastic strain
𝑝
𝜀 𝜑 𝜎𝜑
𝜀𝑝 = 𝜎eff .
This means that there is a relationship with a stress-strain hardening curve using the
effective stress and strain and a corresponding stress-strain hardening curve using the
actual stress and strain values. Assume that a test reveals that the hardening is given by
the curve
𝑝
𝜎𝜑 = 𝜎𝜑 (𝜀𝜑 )
and we want to determine the hardening curve used by LS-DYNA
𝜎eff = 𝜎eff (𝜀𝑝 ),
then using the relationships above yields
𝜎eff (𝜀𝑝 ) = 𝑘𝜑 𝜎𝜑 (𝑘𝜑 𝜀𝑝 ).
Consequently a user input hardening curve must internally be transformed to an effective
hardening curve to be used in the material model to get the desired behavior. Still, the
effective plastic strain is not going to be equal to the plastic strain component in the tensile
direction and validation of the hardening behavior is not straightforward. Therefore we
introduce a new effective plastic strain 𝜀̃𝑝 with evolution given by
𝑑𝜀̃𝑝 𝑑𝜀𝑝
𝑑𝑡
= 𝑘𝜑 𝑑𝑡
that can be used to verify the hardening relationship. This is actually the von Mises plastic
strain in the work hardening sense and is output to the d3plot database as history variable
#2 for post-processing.
and let 𝜎 be the largest eigenvalue to this matrix and 𝑞𝑖 the associated eigenvector
components. Furthermore we define 𝜎𝑣 = (𝜎̂ 11 + 𝜎̂ 22 )/2 as the normalized volumetric
(𝜎 ̂22 )2
̂11 −𝜎
stress and 𝜎𝑑 = √ 4
2
+ 𝜎̂ 12 the normalized shear stress, and make a note that 0 ≤
𝜎𝑑 ≤ 1/√2 and
𝜎𝑑 = 0 → biaxial stress state
𝜎𝑑 = 1/2 → uniaxial stress state
𝜎𝑑 = 1/√2 → shear stress state.
If 𝑎 = 2𝜎𝑣2 is the fraction of stress that is volumetric and 𝑏 = 𝜎 2 an indicator of uniaxial
stress state, then 𝑐 = 𝑎(1 − 4{𝑏 − 1/2}2 ) is a normalized measure that indicates when the
stress is deviatoric/uniaxial or volumetric. That is, 𝑐 = 0 means that the stress is deviatoric
or uniaxial and 𝑐 = 1 means that it is volumetric.
Now, let 𝑞 = 4𝑞21 (1 − 𝑞21 ) be the fraction of the eigenvector 𝑞𝑖 that points in the 00 or 90
direction. That is, 𝑞 = 1 means that the eigenvector points in the 45 direction and 𝑞 = 0
means that it is pointing in either the 00 or 90 direction. Moreover, the same way of
reasoning is valid for the eigenvector associated to the smallest eigenvalue.
To determine in which direction of 00 or 90 a certain eigenvector is pointing we introduce
𝑑 = 𝑏𝑞21 + (1 − 𝑏)(1 − 𝑞21 ), and deduce that 𝑑 = 1 means that the eigenvector 𝑞𝑖 points in the
00 direction and 𝑑 = 0 means that it is pointing in the 90 direction.
We are now ready to give partial expressions for the three uniaxial convex parameters
𝛼̃00 = (1 − 𝑐)𝑑(1 − 𝑞) + 𝑐/4
𝛼̃45 = (1 − 𝑐)𝑞 + 𝑐/2
𝛼̃00 = (1 − 𝑐)(1 − 𝑑)(1 − 𝑞) + 𝑐/4
and these are completed by means of adding the biaxial and shear parts
𝛼bi = max(0,1 − 4𝜎𝑑2 )
𝛼sh = max(0,4𝜎𝑑2 − 1)
𝛼00 = 𝛼̃00 (1 − 𝛼bi − 𝛼sh )
𝛼45 = 𝛼̃45 (1 − 𝛼bi − 𝛼sh )
𝛼90 = 𝛼̃90 (1 − 𝛼bi − 𝛼sh )
This set of parameters fulfills the requirements mentioned above and allows for a decent
expression for a directional dependent yield stress.
In the consistency condition we do not consider the derivatives of the convex parameters
with respect to the stress, as we assume that these will not have a major impact on
convergence.
where 𝜎̂ 𝑖𝑗 are the normalized stress components, 𝑛𝑖 is the direction of plastic flow and
where we have suppressed the dependence of stress and plastic strain in 𝑅. By setting
𝜕𝑓 𝜕𝑓 𝜕𝑓
∆𝑛1 = 𝑛1 − 𝜕𝜎 , ∆𝑛2 = 𝑛2 − 𝜕𝜎 , ∆𝑛4 = 𝑛4 − 𝜕𝜎
11 22 12
and
∆𝑅 = 𝑅(𝜀𝑝 ) − 𝑅(0)
we can simplify this equation as
(𝜎̂ 22 + {𝜎̂ 11 + 𝜎̂ 22 }𝑅)∆𝑛1 + (𝜎̂ 11 + {𝜎̂ 11 + 𝜎̂ 22 }𝑅)∆𝑛2 − 𝜎̂ 12 ∆𝑛4 =
𝜕𝑓 𝜕𝑓
−(𝜕𝜎 + 𝜕𝜎 ){𝜎̂ 11 + 𝜎̂ 22 }∆𝑅
11 22
assuming that the relation already holds for the yield surface normal and R-value in the
reference configuration.
This equation is complemented with a consistency condition of the plastic flow
𝜎̂ 11 ∆𝑛1 + 𝜎̂ 22 ∆𝑛2 + 𝜎̂ 12 ∆𝑛4 = 0.
These two equations are linearly independent if and only if
(𝜎
̂ −𝜎̂ ) 2
{𝜎̂ 11 + 𝜎̂ 22 }√ 11 4 22 + 𝜎̂ 12
2
≠0
and then the equation
−𝜎̂ 12 ∆𝑛1 + 𝜎̂ 12 ∆𝑛2 + (𝜎̂ 11 − 𝜎̂ 22 )∆𝑛4 = 0
can be used to complement the previous two. This defines a system of equations that can
be used to solve in least square sense for the perturbation ∆𝑛𝑖 of the yield surface normal to
get the R-value of interest. To avoid numerical problems and make the perturbation
continuous with respect to the stress, the right hand side of the first equation is changed to
𝜕𝑓 𝜕𝑓 2
−(𝜕𝜎 + 𝜕𝜎 ){𝜎̂ 11 + 𝜎̂ 22 }((𝜎̂ 11 − 𝜎̂ 22 )2 + 4𝜎̂ 12 )∆𝑅.
11 22
This results in a non-associated flow rule, meaning that the plastic flow is not in the
direction of the yield surface normal. Again, we don’t take any special measures into
account for the stress return algorithm as we believe that the perturbation of the normal
is small enough not to deteriorate convergence. In figure 20-1 the plastic flow direction
is illustrated as function of the stress on the yield surface.
Consider Cartesian reference axes which are parallel to the three symmetry planes of
anisotropic behavior. Then the yield function suggested by Hill [1948] can be written
2
F(𝜎22 − 𝜎33 )2 + G(𝜎33 − 𝜎11 )2 + H(𝜎11 − 𝜎22 )2 + 2L𝜎23 2
+ 2M𝜎31 2
+ 2N𝜎12 − 1 = 0, (22.37.1)
where 𝜎y1 , 𝜎y2 , and 𝜎y3 , are the tensile yield stresses and 𝜎y12 , 𝜎y23 , and 𝜎y31 are the shear
yield stresses. The constants F, G, H, L, M, and N are related to the yield stress by
1
2L = 2
𝜎y23
1
2M = 2
𝜎y31
1
2N = 2
𝜎y12
(22.37.2)
1 1 1
2F = 2
+ 2
− 2
𝜎y2 𝜎y3 𝜎y1
1 1 1
2G = 2
+ 2
− 2
𝜎y3 𝜎y1 𝜎y2
1 1 1
2H = 2
+ 2
− 2
.
𝜎y1 𝜎y2 𝜎y3
For the particular case of transverse anisotropy, where properties do not vary in the
𝑥1 − 𝑥2 plane, the following relations hold:
20-122 (Material Models) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Material Models
1
F = 2G = 2
𝜎y3
2 1
2H = 2
− 2 (22.37.5)
𝜎y 𝜎y3
2 1 1
N= 2
− 2
,
𝜎y 2 𝜎y3
𝜎y
Letting K = 𝜎y3, the yield criterion can be written
𝑭(𝛔) = 𝜎e = 𝜎y , (22.37.6)
where
2 2
𝐹(𝛔) ≡ [𝜎11 + 𝜎22 2
+ K2 𝜎33 − K2 𝜎33 (𝜎11 + 𝜎22 ) − (2 − K2 )𝜎11 𝜎22
1 (22.37.7)
1 2 2 ⁄2
+2L𝜎y2 (𝜎23
2
+ 2
𝜎31 ) + 2 (2 − K ) 𝜎12 ] .
2
p
The rate of plastic strain is assumed to be normal to the yield surface so 𝜀̇𝑖𝑗 is found
from
p ∂𝐹
𝜀̇𝑖𝑗 = λ . (22.37.8)
∂𝜎𝑖𝑗
Now consider the case of plane stress, where 𝜎33 = 0. Also, define the anisotropy
input parameter 𝑅 as the ratio of the in-plane plastic strain rate to the out-of-plane plastic
strain rate:
p
𝜀̇22
𝑅= p . (22.37.9)
𝜀̇33
It then follows that
2
𝑅= − 1. (22.37.10)
K2
Using the plane stress assumption and the definition of 𝑅, the yield function may
now be written
1
2 2 2R 2R + 1 2 ⁄2 (22.37.11)
F(𝛔) = [𝜎11 + 𝜎22 − 𝜎11 𝜎22 + 2 𝜎 ] .
R+1 R + 1 12
𝜇 𝐼2
𝑊(𝐼1 , 𝐼2 , 𝐼3 ) = ( + 2√𝐼3 − 5), (22.38.1)
2 𝐼3
where 𝜇 is the shear modulus and 𝐼1 , 𝐼2 , and 𝐼3 are the strain invariants. Blatz and Ko
[1962] suggested this form for a 47 percent volume polyurethane foam rubber with a
Poisson’s ratio of 0.25. The second Piola-Kirchhoff stresses are given as
1 𝐼
𝑆𝑖𝑗 = 𝜇 [(𝐼𝛿𝑖𝑗 − 𝐺𝑖𝑗 ) + (√𝐼3 − 2 ) 𝐺𝑖𝑗 ], (22.38.2)
𝐼3 𝐼3
𝜕𝑥 𝜕𝑥 𝜕𝑋𝑖 𝜕𝑋𝑗
where 𝐺𝑖𝑗 = 𝜕𝑋𝑘 𝜕𝑋𝑘 , 𝐺𝑖𝑗 = 𝜕𝑥𝑘 𝜕𝑥𝑘
, after determining 𝑆𝑖𝑗 , it is transformed into the Cauchy
𝑖 𝑗
stress tensor:
𝜌 𝜕𝑥𝑖 𝜕𝑥𝑗
σ𝑖𝑗 = 𝑆 , (22.38.3)
𝜌0 𝜕𝑋𝑘 𝜕𝑋𝑙 𝑘𝑙
where 𝜌0 and 𝜌 are the initial and current density, respectively.
The strains on which these calculations are based are integrated in time from the
strain rates:
εmjr = 0
εmjr
Plane Strain
80
70
60
% Major Strain
50
40
εmnr εmjr
30 εmnr
20
Draw
εmjr
10
Stretch
∇𝑛+1⁄2 𝑛+1⁄2
𝜀𝑛+1
𝑖𝑗 = 𝜀𝑛𝑖𝑗 + 𝜀𝑖𝑗 Δ𝑡 , (22.39.2)
and are stored as history variables. The resulting strain measure is logarithmic.
The anisotropic parameters R00, R45, and R90 are defined in terms of 𝐹, 𝐺, 𝐻, and𝑁
as [Hill, 1989]:
𝐻
2𝑅00 = ,
𝐺
2𝑁
2𝑅45 = − 1, (22.42.2)
(𝐹 + 𝐺)
𝐻
2𝑅90 = .
𝐹
To avoid numerical problems the minimum strain rate, 𝜀̇min must be defined and the
initial yield stress 𝜎0 is calculated as
𝜎0 = 𝐴𝜀𝑚 𝑛
0 𝜀̇min = 𝐸𝜀0 , (22.42.4)
1
𝐸 𝑚−1
(22.42.5)
𝜀0 = ( 𝑛 ) .
𝐴𝜀̇min
This is equivalent to writing the constitutive model with respect to a set of directors
whose direction is defined by the plastic deformation [Bammann and Aifantis 1987,
Bammann and Johnson 1987]. Decomposing both the skew symmetric and symmetric
parts of the velocity gradient into elastic and plastic parts we write for the elastic stretching
𝐃e and the elastic spin 𝐖e ,
𝐃e = 𝐃 − 𝐃p − 𝐃th , 𝐖e = 𝐖 = 𝐖p . (22.51.3)
Within this structure it is now necessary to prescribe an equation for the plastic spin
𝐖 in addition to the normally prescribed flow rule for 𝐃p and the stretching due to the
p
The evolution of the internal variables 𝛼 and 𝜅 are prescribed in a hardening minus
recovery format as,
o
𝛂 = ℎ(𝑇)𝐃p − [𝑟d (T) ∣𝐃p ∣ + 𝑟s (𝑇)] |𝛂|𝛂, (22.51.7)
If we assume that 𝐖p = 0, we recover the Jaumann stress rate which results in the
prediction of an oscillatory shear stress response in simple shear when coupled with a
Prager kinematic hardening assumption [Johnson and Bammann 1984]. Alternatively we
can choose,
𝐖p = 𝐑T 𝐔̇ 𝐔−1 𝐑, (22.51.9)
which recovers the Green-Naghdi rate of Cauchy stress and has been shown to be
equivalent to Mandel’s isoclinic state [Bammann and Aifantis 1987]. The model employing
this rate allows a reasonable prediction of directional softening for some materials but in
general under-predicts the softening and does not accurately predict the axial stresses
which occur in the torsion of the thin walled tube.
The final equation necessary to complete our description of high strain rate
deformation is one which allows us to compute the temperature change during the
deformation. In the absence of a coupled thermomechanical finite element code we assume
adiabatic temperature change and follow the empirical assumption that 90 - 95% of the
plastic work is dissipated as heat. Hence,
0.9
𝑇̇ = (𝛔 ⋅ 𝐃p ), (22.51.10)
𝜌𝐶v
where 𝜌 is the density of the material and 𝐶v the specific heat.
In some respects this model is similar to the crushable honeycomb model type 26 in
that the components of the stress tensor are uncoupled until full volumetric compaction is
achieved. However, unlike the honeycomb model this material possesses no directionality
but includes the effects of confined air pressure in its overall response characteristics.
𝜎𝑖𝑗 = 𝜎𝑖𝑗sk − δ𝑖𝑗 𝜎 air , (22.53.1)
where 𝜎𝑖𝑗sk is the skeletal stress and 𝜎 air is the air pressure computed from the equation:
𝑝0 𝛾
𝜎 air = − , (22.53.2)
1+𝛾−𝜙
where 𝑝0 is the initial foam pressure usually taken as the atmospheric pressure and 𝛾
defines the volumetric strain
𝛾 = 𝑉 − 1 + 𝛾0 , (22.53.3)
where 𝑉 is the relative volume and 𝛾0 is the initial volumetric strain which is typically
zero. The yield condition is applied to the principal skeletal stresses which are updated
independently of the air pressure. We first obtain the skeletal stresses:
𝜎𝑖𝑗sk = 𝜎𝑖𝑗 + 𝜎𝑖𝑗 𝜎 air , (22.53.4)
𝛔𝑖skt
𝛔𝑖sk = min(𝜎y , ∣𝛔𝑖skt ∣) . (22.53.6)
∣𝛔𝑖skt ∣
where 𝑎, 𝑏, and 𝑐 are user defined input constants. After scaling the principal stresses they
are transformed back into the global system and the final stress state is computed
𝜎𝑖𝑗 = 𝜎𝑖𝑗sk − 𝛿𝑖𝑗 𝜎 air . (22.53.8)
The Chang/Chang criteria is given as follows: for the tensile fiber mode,
𝜎aa 2 𝜎
𝜎aa > 0 then 𝑒f2 = ( ) + 𝛽 ( ab ) − 1 {≥ 0 failed , (22.54.1)
Xt Sc < 0 elastic
Gab = 0
Eb = νba = νab = 0 ⇒ . (22.54.8)
XC = 2Yc , for 50% fiber volume
In the Tsay/Wu criteria the tensile and compressive fiber modes are treated as in the
Chang/Chang criteria. The failure criterion for the tensile and compressive matrix mode
is given as:
2 2
𝜎bb 𝜎 (Y − Yt ) 𝜎bb
2
𝑒md = + ( ab ) + c − 1 {≥ 0 failed . (22.54.9)
Yc Yt Sc Yc Yt < 0 elastic
For 𝛽 = 1 we get the original criterion of Hashin [1980] in the tensile fiber mode. For 𝛽 = 0,
we get the maximum stress criterion which is found to compare better to experiments.
2. If DFAILT is greater than zero, failure occurs if the tensile fiber strain is greater
than DFAILT or less than DFAILC.
3. If EFS is greater than zero, failure occurs if the effective strain is greater than EFS.
4. If TFAIL is greater than zero, failure occurs according to the element time step as
described in the definition of TFAIL above.
When failure has occurred in all the composite layers (through-thickness integration
points), the element is deleted. Elements which share nodes with the deleted element
become “crashfront” elements and can have their strengths reduced by using the SOFT
parameter with TFAIL greater than zero.
Information about the status in each layer (integration point) and element can be
plotted using additional integration point variables. The number of additional integration
point variables for shells written to the LS-DYNA database is input by the
*DATABASE_BINARY definition as variable NEIPS. For Models 54 and 55 these
additional variables are tabulated below (i = shell integration point):
The following components, defined by the sum of failure indicators over all
through-thickness integration points, are stored as element component 7 instead of the
effective plastic strain.:
LS-DYNA DEV 10/27/16 (r:8004) 20-133 (Material Models)
Material Models LS-DYNA Theory Manual
The model uses tabulated input data for the loading curve where the nominal
stresses are defined as a function of the elongations, 𝜀𝑖 , which are defined in terms of the
principal stretches, 𝜆𝑖 , as:
𝜀𝑖 = 𝜆 𝑖 − 1. (22.57.1)
The stretch ratios are found by solving for the eigenvalues of the left stretch tensor,
𝑉𝑖𝑗 , which is obtained via a polar decomposition of the deformation gradient matrix, 𝐹𝑖𝑗 :
𝐹𝑖𝑗 = 𝑅𝑖𝑘 𝑈𝑘𝑗 = 𝑉𝑖𝑘 𝑅𝑘𝑗 . (22.57.2)
The update of 𝑉𝑖𝑗 follows the numerically stable approach of Taylor and Flanagan
[1989]. After solving for the principal stretches, the elongations are computed and, if the
elongations are compressive, the corresponding values of the nominal stresses, 𝜏𝑖 are
interpolated. If the elongations are tensile, the nominal stresses are given by
𝜏𝑖 = 𝐸𝜀𝑖 . (22.57.3)
Unloading
curves
Strain Strain
Figure 22.57.1. Behavior of the low-density urethane foam model.
When hysteretic unloading is used, the reloading will follow the unloading curve if
the decay constant, 𝛽, is set to zero. If 𝛽 is nonzero the decay to the original loading curve
is governed by the expression:
1 − 𝑒−𝛽𝑡 . (22.57.5)
The bulk viscosity, which generates a rate dependent pressure, may cause an
unexpected volumetric response and, consequently, it is optional with this model.
Rate effects are accounted for through linear viscoelasticity by a convolution integral
of the form
𝑡 ∂𝜀𝑘𝑙
𝜎𝑖𝑗r = ∫ 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) 𝑑𝜏, (22.57.6)
0 ∂𝜏
where 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) is the relaxation function. The stress tensor, 𝜎𝑖𝑗r , augments the stresses
determined from the foam, 𝜎𝑖𝑗f ; consequently, the final stress, 𝜎𝑖𝑗 , is taken as the summation
of the two contributions:
𝜎𝑖𝑗 = 𝜎𝑖𝑗f + 𝜎𝑖𝑗r . (22.57.7)
Since we wish to include only simple rate effects, the relaxation function is
represented by one term from the Prony series:
N
𝑔(𝑡) = 𝛼0 + ∑ α𝑚 𝑒−𝛽𝑡 , (22.57.8)
𝑚=1
given by,
𝑔(𝑡) = 𝐸d 𝑒−𝛽1𝑡 . (22.57.9)
This model is effectively a Maxwell fluid which consists of a damper and spring in
series. We characterize this in the input by a Young's modulus, 𝐸d , and decay constant, 𝛽1 .
The formulation is performed in the local system of principal stretches where only the
principal values of stress are computed and triaxial coupling is avoided. Consequently, the
one-dimensional nature of this foam material is unaffected by this addition of rate effects.
The addition of rate effects necessitates twelve additional history variables per integration
point. The cost and memory overhead of this model comes primarily from the need to
“remember” the local system of principal stretches.
The stress limits are factors used to limit the stress in the softening part to a given
value,
𝜎min = SLIMxx ⋅ strength, (22.58.1)
thus, the damage value is slightly modified such that elastoplastic like behavior is
achieved with the threshold stress. As a factor for SLIMxx a number between 0.0 and 1.0 is
possible. With a factor of 1.0, the stress remains at a maximum value identical to the
strength, which is similar to ideal elastoplastic behavior. For tensile failure a small value
for SLIMTx is often reasonable; however, for compression SLIMCx = 1.0 is preferred. This
is also valid for the corresponding shear value. If SLIMxx is smaller than 1.0 then
localization can be observed depending on the total behavior of the lay-up. If the user is
intentionally using SLIMxx < 1.0, it is generally recommended to avoid a drop to zero and
set the value to something in between 0.05 and 0.10. Then elastoplastic behavior is
achieved in the limit which often leads to less numerical problems. Defaults for
SLIMXX = 1.0E-8.
The crashfront-algorithm is started if and only if a value for TSIZE (time step size,
with element elimination after the actual time step becomes smaller than TSIZE) is input .
The damage parameters can be written to the postprocessing database for each
integration point as the first three additional element variables and can be visualized.
For material model FS = 1 an interaction between normal stresses and shear stresses
is assumed for the evolution of damage in the a- and b- directions. For the shear damage is
always the maximum value of the damage from the criterion in a- or b- direction is taken.
SC
TAU1
SLIMS*SC
GAMMA1 GMS
γ
Figure 22.58.1. Stress-strain diagram for shear.
In addition a stress limiter can be used to keep the stress constant via the SLIMS
parameter. This value must be less than or equal to 1.0 and positive, which leads to an
elastoplastic behavior for the shear part. The default is 1.0E-08, assuming almost brittle
failure once the strength limit SC is reached.
The stress before the update is used for 𝛔′. For shell elements, the through-thickness
strain rate is calculated as follows
𝐺
𝑑𝑡σ′ 33 ,
𝑑𝜎33 = 0 = 𝐾(𝜀̇11 + 𝜀̇22 + 𝜀̇33 )𝑑𝑡 + 2𝐺𝜀̇′33 𝑑𝑡 − (22.60.3)
𝜐
where the subscript 𝑖𝑗 = 33 denotes the through-thickness direction and 𝐾 is the elastic bulk
modulus. This leads to:
𝜀̇33 = −a(𝜀̇11 + 𝜀̇22 ) + 𝑏𝑝, (22.60.4)
𝐾 − 2𝐺
𝑎= 3 , (22.60.5)
𝐾 + 4𝐺
3
𝐺𝑑𝑡
𝑏= , (22.60.6)
𝜐(𝐾 + 4 𝐺)
3
in which 𝑝 is the pressure defined as the negative of the hydrostatic stress.
For the Kelvin model the stress evolution equation is defined as:
1 G
𝑠𝑖𝑗̇ + 𝑠𝑖𝑗 = (1 + 𝛿𝑖𝑗 )G0 𝜀′̇ 𝑖𝑗 + (1 + 𝛿𝑖𝑗 ) ∞ 𝜀′𝑖𝑗 , (22.61.3)
𝜏 𝜏
where 𝛿𝑖𝑗 is the Kronecker delta, G0 is the instantaneous shear modulus, G∞ is the long term
shear modulus, and τ is the decay constant.
The pressure is determined from the bulk modulus and the volumetric strain:
𝑝 = −𝐾𝜀v , (22.61.4)
where
𝑉
𝜀v = ln ( ), (22.61.5)
𝑉0
defines the logarithmic volumetric strain from the relative volume.
Bandak’s [1991] calculation of the total strain tensor, 𝜀𝑖𝑗 , for output uses an
incremental update based on Jaumann rate:
𝛻𝑛+1⁄2 𝑛+1⁄2
𝜀𝑛+1
𝑖𝑗 = 𝜀𝑛𝑖𝑗 + 𝑟𝑖𝑗𝑛 + 𝜀𝑖𝑗 𝛥𝑡 , (22.61.6)
where
𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄2
𝛥𝜀𝑖𝑗 = 𝜀̇𝑖𝑗 𝛥𝑡 , (22.61.7)
and 𝑟𝑖𝑗𝑛 gives the rotation of the stain tensor at time 𝑡𝑛 to the configuration at 𝑡𝑛+1
𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄2
𝑟𝑖𝑗𝑛 = (𝜀𝑛𝑖𝑝 𝜔𝑝𝑗 𝑛
+ 𝜀𝑗𝑝 𝜔𝑝𝑖 ) 𝛥𝑡 . (22.61.8)
The model consists of a nonlinear elastic stiffness in parallel with a viscous damper.
A schematic is shown in Figure 22.62.1. The elastic stiffness is intended to limit total crush
while the viscous damper absorbs energy. The stiffness 𝐸2 prevents timestep problems.
E1
V1
E2
This isotropic foam model crushes one-dimensionally with a Poisson’s ratio that is
essentially zero. The stress versus strain behavior is depicted in Figure 22.63.1 where an
example of unloading from point a to the tension cutoff stress at b then unloading to point
c and finally reloading to point d is shown. At point the reloading will continue along the
loading curve. It is important to use nonzero values for the tension cutoff to prevent the
disintegration of the material under small tensile loads. For high values of tension cutoff
the behavior of the material will be similar in tension and compression.
In the implementation we assume that Young’s modulus is constant and update the
stress assuming elastic behavior.
𝑛+1⁄2 𝑛+1⁄2
𝜎𝑖𝑗trial = 𝜎𝑖𝑗𝑛 + 𝐸𝜀̇𝑖𝑗 Δ𝑡 . (22.63.1)
The magnitudes of the principal values, 𝜎𝑖trial , 𝑖 = 1,3 are then checked to see if the yield
stress, 𝜎y , is exceeded and if so they are scaled back to the yield surface:
𝜎𝑖trial
if 𝜎y < ∣𝜎𝑖trial ∣ then 𝜎𝑖𝑛+1 = 𝜎y . (22.63.2)
∣σ𝑖trial ∣
After the principal values are scaled, the stress tensor is transformed back into the
global system. As seen in Figure 22.63.1, the yield stress is a function of the natural
a
σij
d
E
c b
Volumetric strain - In V
Figure 22.63.1. Yield stress versus volumetric strain curve for the crushable
foam.
This model can be combined with the superplastic forming input to control the
magnitude of the pressure in the pressure boundary conditions in order to limit the
effective plastic strain rate so that it does not exceed a maximum value at any integration
point within the model.
The relationship between heat capacity (specific heat) and temperature may be
characterized by a cubic polynomial equation as follows:
Cp = G 1 + G 2 𝑇 + G 3 𝑇 2 + G 4 𝑇 3 . (22.65.4)
For null stiffness coefficients, no forces corresponding to these null values will
develop. The viscous damping coefficients are optional.
For null load curve ID’s, no forces are computed. The force resultants are found
from load curves (See Figure 22.67.1) that are defined in terms of the force resultant versus
the relative displacement in the local coordinate system for the discrete beam.
(a.) (b.)
R
E
S R
U E
L S
T U
A L
N T
T A
N
DISPLACEMENT T
|DISPLACEMENT|
Figure 22.67.1. The resultant forces and moments are determined by a table
lookup. If the origin of the load curve is at [0,0] as in (b.) and tension and
compression responses are symmetric.
After failure, the discrete element is deleted. If failure is included, either one or both
of the criteria may be used.
R
E
S
U
L
T
A
N
T
PLASTIC DISPLACEMENT
Figure 22.68.1. The resultant forces and moments are limited by the yield
definition. The initial yield point corresponds to a plastic displacement of zero.
d4
d3
d2 Piston
d1
Vp
2Ri - h
Figure 22.69.1. Mathematical model for the Side Impact Dummy damper.
F
O
R
C
E Last orifice
closes.
DISPLACEMENT
In the implementation, the orifices are assumed to be circular with partial covering
by the orifice controller. As the piston closes, the closure of the orifice is gradual. This
gradual closure is taken into account to insure a smooth response. If the piston stroke is
exceeded, the stiffness value, 𝑘, limits further movement, i.e., if the damper bottoms out in
tension or compression, the damper forces are calculated by replacing the damper by a
bottoming out spring and damper, k and c, respectively. The piston stroke must exceed the
initial length of the beam element. The time step calculation is based in part on the stiffness
value of the bottoming out spring. A typical force versus displacement curve at constant
relative velocity is shown in Figure 22.69.2. The factor, SF, which scales the force defaults
to 1.0 and is analogous to the adjusting ring on the damper.
As the damper is compressed two actions contribute to the force that develops.
First, the gas is adiabatically compressed into a smaller volume. Secondly, oil is forced
through an orifice. A profiled pin may occupy some of the cross-sectional area of the
orifice; thus, the orifice area available for the oil varies with the stroke. The force is
assumed proportional to the square of the velocity and inversely proportional to the
available area. The equation for this element is:
𝑛
𝑉 2 𝐶0
𝐹 = SCLF ⋅ {𝐾h ( ) + [𝑃0 ( ) − 𝑃a ] ⋅ 𝐴p }, (22.70.1)
𝑎0 𝐶0 − 𝑆
where 𝑆 is the element deflection and 𝑉 is the relative velocity across the element.
Orifice
The area and offset are defined on either the cross section or element cards in the LS-
DYNA input. For a slack cable the offset should be input as a negative length. For an
initial tensile force the offset should be positive. If a load curve is specified, the Young’s
modulus will be ignored and the load curve will be used instead. The points on the load
curve are defined as engineering stress versus engineering strain, i.e., the change in length
over the initial length. The unloading behavior follows the loading.
Since we wish to include only simple rate effects, the relaxation function is
represented by up to six terms of the Prony series:
𝑁
𝑔(𝑡) = 𝛼0 + ∑ 𝛼𝑚 𝑒−𝛽𝑡 . (22.73.3)
𝑚=1
This model is effectively a Maxwell fluid which consists of a dampers and springs in
series. The formulation is performed in the local system of principal stretches where only
the principal values of stress are computed and triaxial coupling is avoided. Consequently,
the one-dimensional nature of this foam material is unaffected by this addition of rate
effects. The addition of rate effects necessitates 42 additional history variables per
integration point. The cost and memory overhead of this model comes primarily from the
need to “remember” the local system of principal stretches and the evaluation of the
viscous stress components.
If the load curve ID for 𝑓 (Δ𝐿) is specified, nonlinear behavior is activated. For this
case the force is given by:
∣Δ𝐿̇∣
𝐹 = 𝐹0 + 𝐾 𝑓 (Δ𝐿) [1 + C1 ⋅ Δ𝐿̇ + C2 ⋅ sgn(Δ𝐿̇)ln (max {1. , })]
DLE (22.74.2)
+𝐷Δ𝐿̇ + 𝑔(Δ𝐿)ℎ(Δ𝐿̇),
where C1 and C2 are damping coefficients for nonlinear behavior, DLE is a factor to scale
time units, and 𝑔(Δ𝐿) is an optional load curve defining a scale factor versus deflection for
load curve ID, ℎ(𝑑Δ𝐿/𝑑𝑡).
The cross sectional area is defined on the section card for the discrete beam
elements, in *SECTION_BEAM. The square root of this area is used as the contact
thickness offset if these elements are included in the contact treatment.
In defining the curves, the stress and strain pairs should be positive values starting
with a volumetric strain value of zero.
Max Stress
Pressure Yield
Volumetric Strain
Figure 22.75.1. Behavior of crushable foam. Unloading is elastic.
where 𝑡𝑚 and 𝒔𝑚 are (strain) quantities governed by the following evolution in time
𝑔
𝒔∇
𝑚 = 𝑫dev − 𝛽𝑚 𝒔𝑚 (22.76.2)
and
∇
𝑡𝑚 = 𝐷vol − 𝛽𝑘𝑚 𝑡𝑚 (22.76.3)
𝑔
Here 𝐾𝑚 and 𝐺𝑚 are bulk and shear moduli, respectively, 𝛽𝑘𝑚 and 𝛽𝑚 are the corresponding
decay coefficients, 𝐷vol and 𝑫dev are the volumetric and deviatoric strain rates, 𝒊 is the 2nd
order identity tensor and ∇ denotes the Jaumann objective rate. It should immediately be
noted that for all decay coefficients equal to 0 (zero), the model is reduced to a time
independent elastic model,
𝝈 ∇ = 𝐾𝐷vol 𝒊 + 2𝐺𝑫dev (22.76.4)
with bulk and shear modulus given by 𝐾 = ∑𝑚 𝐾𝑚 and 𝐺 = ∑𝑚 𝐺𝑚 . For small
displacement theory, the stress can be integrated to be given by
𝑡 𝑘 𝑡 𝑔
𝝈(𝑡) = ∑ (𝐾𝑚 ∫ 𝑒−𝛽𝑚 (𝑡−𝜏) 𝐷vol (𝜏) 𝑑𝜏𝒊 + 2𝐺𝑚 ∫ 𝑒−𝛽𝑚(𝑡−𝜏) 𝑫dev (𝜏) 𝑑𝜏) (22.76.5)
𝑚 0 0
For shell elements, the same theory applies except that the objective rate ∇ is the
corotational time derivative instead of the Jaumann rate.
̇
mat
𝒇int = ∫ 𝑩𝑇 𝝈 ∇𝑇 𝑑Ω (22.76.6)
where 𝑩 is the strain displacement matrix, the integration is over the current configuration
Ω and ∇𝑇 stands for Truesdell rate. This expression should later be identified with
mat
𝜕𝒇int 𝜕𝒇 mat
̇
mat
𝒇int = 𝒖̇ + int (22.76.7)
𝜕𝒖 𝜕𝑡
in order to determine the tangent modulus. Neglecting discrepancies between the
Jaumann and Truesdell rates, we can use (22.76.2) and (22.76.3) in (22.76.6) to get
̇
mat 𝑔
𝒇int = ∫ 𝑩𝑇 ∑(𝐾𝑚 𝒊⨂𝒊 + 2𝐺𝑚 𝑰dev ) 𝑩𝑑𝛺 𝒖̇ − ∫ 𝑩𝑇 ∑(𝐾𝑚 𝛽𝑘𝑚 𝑡𝑚 𝒊 + 2𝐺𝑚 𝛽𝑚 𝒔𝑚 ) 𝑑𝛺, (22.76.8)
𝑚 𝑚
where 𝑰dev is the 4th order deviatoric identity tensor. Comparing this expression with
(22.76.7) one can conclude that
mat
𝜕𝒇int
= ∫ 𝑩𝑇 ∑(𝐾𝑚 𝒊⨂𝒊 + 2𝐺𝑚 𝑰dev ) 𝑩𝑑𝛺, (22.76.9)
𝜕𝒖 𝑚
Instead of inputting the stiffness and relaxation parameters, one can input a relaxation
curve from test according to Figure 22.76.1. The time scale is determined by BSTART and
LS-DYNA will determine all parameters as to best fit the curve.
The 𝑞𝑖𝑗 are the components of the orthogonal tensor containing the eigenvectors of
the principal basis. The Cauchy stress is then given by
𝜎𝑖𝑗 = 𝐽 −1 𝜏𝑖𝑗 , (22.77.3)
where 𝐽 = 𝜆1 𝜆2 𝜆3 is the relative volume change.
The constitutive tensor that relates the rate of deformation to the Truesdell
(convected) rate of Kirchoff stress can in the principal basis be expressed as
𝜕𝜏𝑖𝑖E
𝐷TKE
𝑖𝑖𝑗𝑗 = 𝜆𝑗 − 2𝜏𝑖𝑖E 𝛿𝑖𝑗
𝜕𝜆𝑗
𝜆2𝑗 𝜏𝑖𝑖E − 𝜆2𝑖 𝜏𝑗𝑗E
𝐷TKE
𝑖𝑗𝑖𝑗 = , 𝑖 ≠ 𝑗, 𝜆𝑖 ≠ 𝜆𝑗 (no sum). (22.77.4)
𝜆2𝑖 − 𝜆2𝑗
𝜆𝑖 𝜕𝜏𝑖𝑖E 𝜕𝜏𝑖𝑖E
𝐷TKE
𝑖𝑗𝑖𝑗 = ( − ), 𝑖 ≠ 𝑗, 𝜆𝑖 = 𝜆𝑗
2 𝜕𝜆𝑖 𝜕𝜆𝑗
𝐷TC −1 TK
𝑖𝑗𝑘𝑙 = 𝐽 𝐷𝑖𝑗𝑘𝑙 . (22.77.6)
When dealing with shell elements, the tangent moduli in the corotational
coordinates is of interest. This matrix is given by
̂ TC
𝐷 TC −1 TK −1
𝑖𝑗𝑘𝑙 = 𝑅𝑝𝑖 𝑅𝑞𝑗 𝑅𝑟𝑘 𝑅𝑠𝑙 𝐷𝑝𝑞𝑟𝑠 = 𝐽 𝑅𝑝𝑖 𝑅𝑞𝑗 𝑅𝑟𝑘 𝑅𝑠𝑙 𝐷𝑝𝑞𝑟𝑠 = 𝐽 𝑞𝑖𝑝 ̂ 𝑞𝑙𝑠̂ 𝐷TKE
̂ 𝑞𝑗𝑞̂ 𝑞𝑘𝑟 𝑝𝑞𝑟𝑠 , (22.77.7)
where 𝑅𝑖𝑗 is the matrix containing the unit basis vectors of the corotational system and𝑞𝑖𝑗̂ =
𝑅𝑘𝑖 𝑞𝑘𝑗 . The latter matrix can be determined as the eigenvectors of the co-rotated left
Cauchy-Green tensor (or the left stretch tensor). In LS-DYNA, the tangent stiffness matrix
is after assembly transformed back to the standard basis according to standard
transformation formulae.
where
′ 𝜕𝑊1 2 2 −1
𝑊1𝑖 ≔ 𝜆𝑖 = (2𝜆𝑖 − 𝐼1 ) 𝐼3 3
𝜕𝜆𝑖 3
(22.77.12)
′ 𝜕𝑊2 4 −2
𝑊2𝑖 : = 𝜆𝑖 = (2𝜆2𝑖 (𝐼1 − 𝜆2𝑖 ) − 𝐼2 ) 𝐼3 3 .
𝜕𝜆𝑖 3
Again, using only the nonzero coefficients mentioned above, Equation (22.77.14) is
reduced to
𝜕𝜏𝑖𝑖E ′ ′ ′ ′ ′ ′ ′ ′
𝜆𝑗 = 𝐶11 (𝑊1𝑖 𝑊2𝑗 + 𝑊1𝑗 𝑊2𝑖 ) + 2(𝐶20 + 3𝐶30 𝑊1 )𝑊1𝑗 𝑊1𝑖 + 2𝐶02 𝑊2𝑖 𝑊2𝑗 +
𝜕𝜆𝑗
(22.77.16)
(𝐶10 + 𝐶11 𝑊2 + 2𝐶20 𝑊1 + 3𝐶30 𝑊12 )𝑊1𝑖𝑗
′′ ′′
+ (𝐶01 + 𝐶11 𝑊1 + 2𝐶02 𝑊2 )𝑊2𝑖𝑗 +
𝐾𝐽(2𝐽 − 1).
where
𝛼 𝛼 𝛼
𝑎𝑚 = 𝜆̃ 1𝑚 + 𝜆̃ 2𝑚 + 𝜆̃ 3𝑚 . (22.77.20)
Here 𝑛 is a number less than or equal to 6, 𝜎̂ 𝑖𝑗ve is the co-rotated viscoelastic stress,
̂ dev
𝐷 𝑖𝑗 is the deviatoric co-rotated rate-of-deformation and 𝐺𝑚 and 𝛽𝑚 are material
parameters. The parameters 𝐺𝑚 can be thought of as shear moduli and 𝛽𝑚 as decay
coefficients determining the relaxation properties of the material. This rate form can be
integrated in time to form the corotated viscoelastic stress
6 𝑡
𝜎̂ 𝑖𝑗ve = ∑ 2𝐺𝑚 ∫ 𝑒−𝛽𝑚(𝑡−𝜏) 𝐷
̂ dev
𝑖𝑗 (𝜏)𝑑𝜏 . (22.77.23)
𝑚=1 0
For the constitutive matrix, we refer to Borrvall [2002] and here simply state that it is
equal to
𝑛
1 1
̂ TCve
𝐷 𝑖𝑗𝑘𝑙 = ∑ 2𝐺𝑚 ( (𝛿𝑖𝑘 𝛿𝑗𝑙 + 𝛿𝑖𝑙 𝛿𝑗𝑘 ) − 𝛿𝑖𝑗 𝛿𝑘𝑙 ). (22.77.24)
𝑚=1
2 3
or equivalently the internal force is assembled in the corotational system and then
transformed back to the standard basis according to standard transformation formulae.
Here 𝑅𝑖𝑗 is the rotation matrix containing the corotational basis vectors. The so-called
corotated stress 𝜎̂ 𝑖𝑗 is evaluated as the sum of the stresses given in Sections 19.77.1 and
19.77.4.
For the hyperelastic stress contribution, the principal stretches are needed and here
taken as the square root of the eigenvalues of the co-rotated left Cauchy-Green tensor 𝑏𝑖𝑗̂ .
The corotated left Cauchy-Green tensor is incrementally updated with the aid of a time
increment Δ𝑡, the corotated velocity gradient 𝐿̂𝑖𝑗 , and the angular velocity 𝛺̂𝑖𝑗 with which
the embedded coordinate system is rotating,
𝑏̂𝑖𝑗 = 𝑏̂𝑖𝑗 + Δ𝑡(𝐿̂𝑖𝑘 − 𝛺
̂𝑖𝑘 )𝑏̂𝑘𝑗 + Δ𝑡𝑏𝑖𝑘
̂ (𝐿̂𝑖𝑘 − 𝛺
̂𝑖𝑘 ). (22.77.26)
and the stresses are determined from this value of 𝛼. Finally, to make sure that the normal
stress through the thickness vanishes, it is set to 0 (zero) before exiting the stress update
routine.
If the load curve ID is provided as a positive number, the deviatoric perfectly plastic
pressure dependent yield function 𝜙, is described in terms of the second invariant, 𝐽2 , the
pressure, 𝑝, and the tabulated load curve, 𝐹(𝑝), as
𝜙 = √3𝐽2 − 𝐹(𝑝) = σ𝑦 − 𝐹(𝑝), (22.78.1)
where 𝐽2 is defined in terms of the deviatoric stress tensor as:
1
𝐽2 = 𝑆𝑖𝑗 𝑆𝑖𝑗 , (22.78.2)
2
assuming that if the ID is given as negative, then the yield function becomes:
𝜙 = 𝐽2 − 𝐹(𝑝), (22.78.3)
being the deviatoric stress tensor.
If cracking is invoked, the yield stress is multiplied by a factor f which reduces with
plastic stain according to a trilinear law as shown in Figure 22.78.1.
1.0
ε1 ε2 εp
ε
ε2
ε1
P
Figure 22.78.2. Cracking strain versus pressure.
𝜀1 and 𝜀2 are tabulated functions of pressure that are defined by load curves (see Figure
22.78.2). The values on the curves are pressure versus strain and should be entered in
order of increasing pressure. The strain values should always increase monotonically with
pressure.
By properly defining the load curves, it is possible to obtain the desired strength and
ductility over a range of pressures. See Figure 22.78.3.
p3
p2
Yield
stress p1
Plastic strain
Figure 22.78.3. Example Caption
The constants 𝑎0 , 𝑎1 , 𝑎2 govern the pressure sensitivity of the yield stress. Only the
ratios between these values are important - the absolute stress values are taken from the
stress-strain curve.
The stress strain pairs (𝛾1, 𝜏1), ... (𝛾5, 𝜏5) define a shear stress versus shear strain
curve. The first point on the curve is assumed by default to be (0,0) and does not need to be
entered. The slope of the curve must decrease with increasing 𝛾. Not all five points need
be to be defined. This curve applies at the reference pressure; at other pressures the curve
varies according to 𝑎0 , 𝑎1 , and 𝑎2 as in the soil and crushable foam model, Material 5.
The stress versus strain behavior may be treated by a bilinear stress strain curve by
defining the tangent modulus, ETAN. Alternately, a curve similar to that shown in Figure
22.81.1 is expected to be defined by (EPS1,ES1) - (EPS8,ES8); however, an effective stress
versus effective plastic strain curve (LCSS) may be input instead if eight points are
insufficient. The cost is roughly the same for either approach. The most general approach
yield stress versus
effective plastic strain
σyield for undamaged material
Failure Begins
nominal stress
after failure
damage increases
linearly with plastic
strain after failure rupture
Two options to account for strain rate effects are possible. Strain rate may be
accounted for using the Cowper-Symonds model which scales the yield stress with the
factor,
1
𝜀̇ ⁄𝑝 (22.81.1)
1+( ) ,
𝐶
where 𝜀̇ is the strain rate, 𝜀̇ = √𝜀̇𝑖𝑗 𝜀̇𝑖𝑗 . If the viscoplastic option is active, VP = 1.0, and if
SIGY is > 0 then the dynamic yield stress is computed from the sum of the static stress,
p
𝜎ys (𝜀eff ), which is typically given by a load curve ID, and the initial yield stress, SIGY,
multiplied by the Cowper-Symonds rate term as follows:
p 1⁄
p
p p p 𝜀̇ (22.81.2)
𝜎𝑦 (𝜀eff , 𝜀̇eff ) = 𝜎𝑦s (𝜀eff ) + SIGY ⋅ ( eff ) ,
C
where the plastic strain rate is used. With this latter approach similar results can be
obtained between this model and material model: *MAT_ANISOTROPIC_VISCOPLASTIC.
p
If SIGY = 0, the following equation is used instead where the static stress, 𝜎ys (𝜀eff ), must be
defined by a load curve:
p 1⁄
p
p p p ⎡ 𝜀̇ ⎤
𝜎y (𝜀eff , 𝜀̇eff ) = 𝜎ys (𝜀eff ) ⎢
⎢1 + ( eff ) ⎥.
⎥ (22.81.3)
C
⎣ ⎦
This latter equation is always used if the viscoplastic option is off. For complete
generality a load curve (LCSR) to scale the yield stress may be input instead. In this curve
the scale factor versus strain rate is defined.
The constitutive properties for the damaged material are obtained from the
undamaged material properties. The amount of damage evolved is represented by the
constant, 𝜔, which varies from zero if no damage has occurred to unity for complete
rupture. For uniaxial loading, the nominal stress in the damaged material is given by
𝑃
𝜎nominal =
, (22.81.4)
𝐴
where P is the applied load and A is the surface area. The true stress is given by:
𝑃
𝜎true = , (22.81.5)
𝐴 − 𝐴loss
where 𝐴loss is the void area. The damage variable can then be defined:
𝐴loss
𝜔= , 0 ≤ 𝜔 ≤ 1. (22.81.6)
𝐴
In this model damage is defined in terms of plastic strain after the failure strain is
exceeded:
p p
𝜀eff − 𝜀failure p p p
𝜔= p p if 𝜀failure ≤ 𝜀eff ≤ 𝜀rupture . (22.81.7)
𝜀rupture − 𝜀failure
After exceeding the failure strain softening begins and continues until the rupture
strain is reached.
By default, deletion of a shell element occurs when all integration points in the shell
have failed. A parameter is available, NUMINT, that defines the number of through
thickness integration points for shell element deletion. The default of all integration points
is not recommended since shells undergoing large strain are often not deleted due to nodal
fiber rotations which limit strains at active integration points after most points have failed.
Better results are obtained if NUMINT is set to 1 or a number less than one half of the
number of through thickness points. For example, if four through thickness points are
used, NUMINT should not exceed 2, even for fully integrated shells which have 16
integration points.
0 failure
ε peff- f s
Figure 22.81.2. A nonlinear damage curve is optional. Note that the origin of the
curve is at (0,0). It is permissible to input the failure strain, fs, as zero for this
option. The nonlinear damage curve is useful for controlling the softening
behavior after the failure strain is reached.
l
u 𝜎11
𝜎11 = , (22.81.9)
1 − D1t
l
u 𝜎22
𝜎22 = ,
1 − D2𝑡
l
u 2𝜎12
𝜎12 = ,
2 − 𝐷𝑡1 − 𝐷𝑡2
l
u 𝜎23
𝜎23 = ,
1 − 𝐷𝑡2
l
u 𝜎13
𝜎13 = .
1 − 𝐷𝑡1
l+ u+ 2 − 𝐷t+ t+
1 − 𝐷2
𝜎12 = 𝜎12 , (22.81.12)
2
l+ u+
𝜎23 = 𝜎23 (1 − 𝐷t+
2 ),
l+ u+
𝜎13 = 𝜎13 (1 − 𝐷t+
1 ),
which is transformed back to the local system to obtain the new global damaged stress as
𝑙+
𝜎𝑖𝑗𝑡+ = 𝑞𝑖𝑘 𝑞𝑗𝑙 𝜎𝑘𝑙 . (22.81.13)
There are options of using visco-plasticity in the current model. The details of this
part of the stress update is omitted here.
20-172 (Material Models) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Material Models
Upon entry the stress is divided by the factor 1 − 𝐹𝑡 to account for the Rc-Dc
damage. Before exiting the routine, the stress is multiplied by the new Rc-Fc (reversed)
fracture fraction 1 − 𝐹𝑡+ . An integration point is considered failed when min(1 − 𝐷1 , 1 −
𝐷2 )(1 − 𝐹) < 0.001.
The strain is divided into two parts: a linear part and a non-linear part of the strain
1-V
Figure 22.83.1. Rate effects in Fu Chang’s foam model.
∞
𝛔(𝑡) = ∫ [𝐄N
𝑡 (𝜏), 𝐒(𝑡)] 𝑑𝜏, (22.83.3)
𝜏=0
∞
where 𝐒(𝑡) is the state variable and ∫𝜏=0 is a functional of all values of 𝜏 in 𝑇𝜏 : 0 ≤ 𝜏 ≤ ∞
and
𝐄N N
𝑡 (𝜏) = 𝐄 (𝑡 − 𝜏), (22.83.4)
where 𝜏 is the history parameter:
𝐄N
𝑡 (𝜏 = ∞) ⇔ the virgin material. (22.83.5)
It is assumed that the material remembers only its immediate past, i.e., a
neighborhood about 𝜏 = 0. Therefore, an expansion of 𝐄N
𝑡 (𝜏) in a Taylor series about 𝜏 =
0 yields:
𝜕𝐄N
𝑡
𝐄N N
𝑡 (𝜏) = 𝐄 (0) + (0)𝑑𝑡. (22.83.6)
𝜕𝑡
Hence, the postulated constitutive equation becomes:
we may write
𝐄̇ N
𝑡 = 𝑓 (𝐒(𝑡), 𝐬(𝑡)), (22.83.9)
which states that the nonlinear strain rate is the function of stress and a state variable which
represents the history of loading. Therefore, the proposed kinetic equation for foam
materials is:
𝛔 tr(𝛔𝐒) 2𝑛0
𝐄̇ N = 𝐷 exp [−𝑐0 ( ) ], (22.83.10)
‖𝛔‖ 0 (‖𝛔‖)2
where 𝐷0 , 𝑐0 , and 𝑛0 are material constants, and 𝐒 is the overall state variable. If either
𝐷0 = 0 or 𝑐0 → ∞ then the nonlinear strain rate vanishes.
𝑛2
𝑆𝑖𝑗̇ = [𝑐1 (𝑎𝑖𝑗 𝑅 − 𝑐2 𝑆𝑖𝑗 )𝑃 + 𝑐3 𝑊 𝑛1 (∥𝐸̇𝑁 ∥) 𝐼𝑖𝑗 ]𝑅 (22.83.11)
𝑛3
∥𝐄̇ N ∥
⎛
𝑅 = 1 + 𝑐4 ⎜ ⎞
− 1⎟ (22.83.12)
⎝ 𝑐5 ⎠
𝑃 = tr(𝛔𝐄̇ N ) (22.83.13)
W = ∫ tr(𝛔𝑑𝐄), (22.83.14)
In the implementation by Fu Chang the model was simplified such that the input
constants 𝑎𝑖𝑗 and the state variables 𝑆𝑖𝑗 are scalars.
−1⁄3 (22.87.1)
𝐽1 = 𝐼1 𝐼3
−2⁄3
𝐽2 = 𝐼2 𝐼3
In order to prevent volumetric work from contributing to the hydrostatic work the
first and second invariants are modified as shown. This procedure is described in more
detail by Sussman and Bathe [1987].
The effects of confined air pressure in its overall response characteristics are
included by augmenting the stress state within the element by the air pressure.
𝜎𝑖𝑗 = 𝜎𝑖𝑗sk − 𝛿𝑖𝑗 𝜎 air , (22.87.2)
where 𝜎𝑖𝑗sk is the bulk skeletal stress and σair is the air pressure computed from the
equation:
𝑝0 𝛾
𝜎 air = − , (22.87.3)
1+𝛾−𝜙
where 𝑝0 is the initial foam pressure usually taken as the atmospheric pressure and 𝛾
defines the volumetric strain
𝛾 = 𝑉 − 1 + 𝛾0 , (22.87.4)
where 𝑉 is the relative volume of the voids and 𝛾0 is the initial volumetric strain which is
typically zero. The rubber skeletal material is assumed to be incompressible.
Rate effects are taken into account through linear viscoelasticity by a convolution
integral of the form:
𝑡 𝜕𝜀𝑘𝑙
𝜎𝑖𝑗 = ∫ 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) 𝑑𝜏, (22.87.5)
0 𝜕𝜏
or in terms of the second Piola-Kirchhoff stress, 𝑆𝑖𝑗 , and Green's strain tensor, 𝐸𝑖𝑗 ,
𝑡 𝜕𝐸𝑘𝑙
𝑆𝑖𝑗 = ∫ 𝐺𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) 𝑑𝜏, (22.87.6)
0 𝜕𝜏
where 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) and 𝐺𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) are the relaxation functions for the different stress
measures. This stress is added to the stress tensor determined from the strain energy
functional.
Since we wish to include only simple rate effects, the relaxation function is
represented by one term from the Prony series:
𝑁
𝑔(𝑡) = 𝛼0 + ∑ 𝛼𝑚 𝑒−𝛽𝑡 , (22.87.7)
𝑚=1
given by,
𝑔(𝑡) = 𝐸𝑑 𝑒−𝛽1𝑡 . (22.87.8)
This model is effectively a Maxwell fluid which consists of a damper and spring in
series. We characterize this in the input by a shear modulus, 𝐺, and decay constant, 𝛽1 .
Air
Figure 22.87.1. Cellular rubber with entrapped air. By setting the initial air
pressure to zero, an open cell, cellular rubber can be simulated.
LS-DYNA DEV 10/27/16 (r:8004) 20-179 (Material Models)
Material Models LS-DYNA Theory Manual
⎡ tanh (α σ̂ )⎤
∂σ̂ ⎢ 𝜎̂ εs ⎥
p ≡ Θo ⎢1 − ⎥. (22.88.2)
∂𝜀 ⎢ tanh(𝛼) ⎥
⎣ ⎦
∂𝜎
̂
The term, ∂𝜀 p , represents the hardening due to dislocation generation and the stress
𝜎
̂
ratio, 𝜎
̂εs
, represents softening due to dislocation recovery. The threshold stress at zero
strain-hardening 𝜎̂ εs is called the saturation threshold stress. Relationships for 𝛩𝑜 , 𝜎̂ εs are:
𝜀̇p 𝜀̇p
𝛩𝑜 = 𝑎𝑜 + 𝑎1 ln ( ) + 𝑎2 √ , (22.88.3)
𝜀0 𝜀0
which contains the material constants 𝑎o , 𝑎1 , and 𝑎2 . The constant, 𝜎̂ εs , is given as:
𝑘𝑇/𝐺𝑏3 𝐴
𝜀̇𝑝
𝜎̂ εs = 𝜎̂ εso ( ) , (22.88.4)
𝜀̇εso
which contains the input constants: 𝜎̂ εso , 𝜀̇εso , 𝑏, 𝐴, and 𝑘. The shear modulus 𝐺 appearing
in these equations is assumed to be a function of temperature and is given by the
correlation.
𝑏1
𝐺 = 𝐺0 − 𝑏2
, (22.88.5)
𝑒𝑇 −1
which contains the constants: 𝐺0, 𝑏1 , and 𝑏2 . For thermal-activation controlled deformation
𝑠th is evaluated via an Arrhenius rate equation of the form:
1
1 𝑝
⎡ 𝜀̇ 𝑞⎤
⎢ ⎛
⎜ 𝑘𝑇ln ( p0 )⎞ ⎥
𝜀̇ ⎟
𝑠th = ⎢1 − ⎜
⎜
⎜
⎟
⎟ ⎥ . (22.88.6)
⎢ ⎜ 𝐺𝑏3 𝑔0 ⎟ ⎟ ⎥
⎢ ⎥
⎝ ⎠
⎣ ⎦
The stress-strain curve is permitted to have sections steeper (i.e. stiffer) than the
elastic modulus. When these are encountered the elastic modulus is increased to prevent
spurious energy generation.
The derivatives of the fiber term 𝐹 are defined to capture the behavior of crimped
collagen. The fibers are assumed to be unable to resist compressive loading - thus the
model is isotropic when 𝜆 < 1. An exponential function describes the straightening of the
fibers, while a linear function describes the behavior of the fibers once they are straightened
past a critical fiber stretch level 𝜆 ≥ 𝜆∗ (input parameter XLAM):
⎧0 𝜆<1
{
{ 𝐶
∂𝐹 { 3 [exp(𝐶4 (𝜆 − 1)) − 1] 𝜆 < 𝜆∗
= 𝜆 . (22.91.2)
∂λ ⎨
{ 1
{
{ (𝐶5 𝜆 + 𝐶6 )
⎩𝜆 𝜆 ≥ 𝜆∗
Here, 𝐒e is the elastic part of the second PK stress as derived from the strain energy, and
𝐺(𝑡 − 𝑠) is the reduced relaxation function, represented by a Prony series:
6
𝑡
𝐺(t) = ∑ S𝑖 exp ( ). (22.91.4)
𝑖=1
𝑇𝑖
Puso and Weiss [1998] describe a graphical method to fit the Prony series coefficients to
relaxation data that approximates the behavior of the continuous relaxation function
proposed by Y-C. Fung, as quasilinear viscoelasticity.
The final force, which includes rate effects and damping, is given by:
∣Δ𝐿̇∣
𝐹𝑛+1 = 𝐹 ⋅ [1 + 𝐶1 ⋅ Δ𝐿̇ + 𝐶2 ⋅ sgn(Δ𝐿̇)ln (max {1. , })] + DΔ𝐿̇ + 𝑔(𝛥𝐿)ℎ(𝛥𝐿̇), (22.94.4)
DLE
where 𝐶1, 𝐶2 are damping coefficients, DLE is a factor to scale time units.
Unless the origin of the curve starts at (0,0), the negative part of the curve is used
when the spring force is negative where the negative of the plastic displacement is used to
interpolate, 𝐹y . The positive part of the curve is used whenever the force is positive. In
these equations, Δ𝐿 is the change in length
Δ𝐿 = current length-initial length. (22.94.5)
Description of properties:
1. 𝐸 is the Young's modulus of the undamaged material also known as the virgin
modulus.
2. 𝜐 is the Poisson's ratio of the undamaged material also known as the virgin
Poisson's ratio.
3. 𝑓𝑛 is the initial principal tensile strength (stress) of the material. Once this stress
has been reached at a point in the body a smeared crack is initiated there with a
normal that is co-linear with the 1st principal direction. Once initiated, the crack is
fixed at that location, though it will convect with the motion of the body. As the
loading progresses the allowed tensile traction normal to the crack plane is pro-
gressively degraded to a small machine dependent constant.
where 𝐧 is the smeared crack normal, 𝜀 is the small constant, 𝐻 is the softening
modulus, and 𝛼 is an internal variable. 𝐻 is set automatically by the program; see
𝑔c below. 𝛼 measures the crack field intensity and is output in the equivalent plas-
tic strain field, 𝜀̅p , in a normalized fashion.
4. 𝑓s is the initial shear traction that may be transmitted across a smeared crack plane.
The shear traction is limited to be less than or equal to 𝑓s (1 − 𝛽)(1 − exp[−𝐻𝛼]),
through the use of two orthogonal shear damage surfaces. Note that the shear
degradation is coupled to the tensile degradation through the internal variable
alpha which measures the intensity of the crack field. 𝛽 is the shear retention factor
defined below. The shear degradation is taken care of by reducing the material's
shear stiffness parallel to the smeared crack plane.
6. 𝛽 is the shear retention factor. As the damage progresses the shear tractions
allowed across the smeared crack plane asymptote to the product 𝛽𝑓s .
Remark: A variety of experimental data has been replicated using this model from
quasi-static to explosive situations. Reasonable properties for a standard grade concrete
would be 𝐸 = 3.15 × 106 psi, 𝑓n = 450 psi, 𝑓s = 2100 psi, 𝑣 = 0.2, 𝑔c = 0.8 lbs/in, 𝛽 = 0.03,𝜂 =
0.0 psi-sec, 𝜎y = 4200 psi. For stability, values of 𝜂 between 104 to 106 psi/sec are
recommended. Our limited experience thus far has shown that many problems require
nonzero values of 𝜂 to run to avoid error terminations. Various other internal variables
such as crack orientations and degraded stiffness tensors are internally calculated but
currently not available for output.
where
𝐴, 𝐵, 𝐶 and 𝑛 are input constants
𝜀̅p effective plastic strain
̇
𝜀̇∗ = 𝜀̇𝜀̅ effective strain rate for 𝜀̇0 = 1s−1
0
If the viscoplastic option is active, VP = 1.0, the parameters SIGMAX and SIGSAT
are ignored since these parameters make convergence of the viscoplastic strain iteration
loop difficult to achieve. The viscoplastic option replaces the plastic strain in the forgoing
equations by the viscoplastic strain and the strain rate by the viscoplastic strain rate.
Numerical noise is substantially reduced by the viscoplastic formulation.
It is advisable to include all spot welds, which provide the slave nodes, and spot
welded materials, which define the master segments, within a single type 7 tied interface.
As a constraint method, multiple type 7 interfaces are treated independently which can
lead to significant problems if such interfaces share common nodal points. The offset
option, “o 7”, should not be used with spot welds.
The DAMAGE-FAILURE option causes one additional line to be read with the
damage parameter and a flag that determines how failure is computed from the resultants.
On this line the parameter, DMG, if nonzero, invokes damage mechanics combined with
the plasticity model to achieve a smooth drop off of the resultant forces prior to the
removal of the spot weld. The parameter FOPT determines the method used in computing
resultant based failure, which is unrelated to damage.
The weld material is modeled with isotropic hardening plasticity coupled to two
failure models. The first model specifies a failure strain which fails each integration point
SPOTWELD ELEMENT r
n2
Trr
t
Mtt
Frr s
Frt
n1
Frs Mss
in the spot weld independently. The second model fails the entire weld if the resultants are
outside of the failure surface defined by:
2 2 2 2 2 2
𝑁 𝑁 𝑁 𝑀 𝑀 𝑇
( 𝑟𝑟 ) + ( 𝑟𝑠 ) + ( 𝑟𝑡 ) + ( 𝑟𝑟 ) + ( 𝑠𝑠 ) + ( 𝑟𝑟 ) − 1 = 0, (22.100.1)
𝑁𝑟𝑟F 𝑁𝑟𝑠F 𝑁𝑟𝑡F 𝑀𝑟𝑟F 𝑀𝑠𝑠F 𝑇𝑟𝑟F
where the numerators in the equation are the resultants calculated in the local coordinates
of the cross section, and the denominators are the values specified in the input. If the user
defined parameter, NF, which the number of force vectors stored for filtering, is nonzero
the resultants are filtered before failure is checked. The default value is set to zero which is
generally recommended unless oscillatory resultant forces are observed in the time history
databases. Even though these welds should not oscillate significantly, this option was
added for consistency with the other spot weld options. NF affects the storage since it is
necessary to store the resultant forces as history variables.
If the failure strain is set to zero, the failure strain model is not used. In a similar
manner, when the value of a resultant at failure is set to zero, the corresponding term in the
failure surface is ignored. For example, if only N𝑟𝑟F is nonzero, the failure surface is
reduced to |N𝑟𝑟 | = N𝑟𝑟F . None, either, or both of the failure models may be active
depending on the specified input values.
The inertias of the spot welds are scaled during the first time step so that their stable
time step size is Δ𝑡. A strong compressive load on the spot weld at a later time may reduce
the length of the spot weld so that stable time step size drops below Δ𝑡. If the value of Δ𝑡 is
zero, mass scaling is not performed, and the spot welds will probably limit the time step
size. Under most circumstances, the inertias of the spot welds are small enough that
scaling them will have a negligible effect on the structural response and the use of this
option is encouraged.
Spotweld force history data is written into the SWFORC ASCII file. In this database
the resultant moments are not available, but they are in the binary time history database.
The constitutive properties for the damaged material are obtained from the
undamaged material properties. The amount of damage evolved is represented by the
constant, ω, which varies from zero if no damage has occurred to unity for complete
rupture. For uniaxial loading, the nominal stress in the damaged material is given by
𝑃
𝜎nominal =
, (22.100.2)
𝐴
where 𝑃 is the applied load and 𝐴 is the surface area. The true stress is given by:
𝑃
𝜎true = , (22.100.3)
𝐴 − 𝐴loss
where 𝐴loss is the void area. The damage variable can then be defined:
𝐴loss (22.100.4)
𝜔= , 0 ≤ 𝜔 ≤ 1.
𝐴
LS-DYNA DEV 10/27/16 (r:8004) 20-191 (Material Models)
Material Models LS-DYNA Theory Manual
In this model damage is defined in terms of plastic strain after the failure strain is
exceeded:
p p
𝜀eff − 𝜀failure p p p
𝜔= p p if 𝜀failure ≤ 𝜀eff ≤ 𝜀rupture . (22.100.5)
𝜀rupture − 𝜀failure
After exceeding the failure strain softening begins and continues until the rupture strain is
reached.
In this material the yield stress may vary throughout the finite element model as a
function of strain rate and hydrostatic stress. Post yield stress behavior is captured in
material softening and hardening values. Finally, ductile or brittle failure measured by
plastic strain or maximum principal stress respectively is accounted for by automatic
element deletion.
The final equation necessary to complete the description of high strain rate
deformation herein is one that allows computation of the temperature change during the
deformation. In the absence of a coupled thermo-mechanical finite element code we
assume adiabatic temperature change and follow the empirical assumption that 90-95% of
the plastic work is dissipated as heat. Thus the heat generation coefficient is
0.9
HC ≈ , (22.102.3)
𝜌𝐶𝑣
where 𝜌 is the material density and 𝐶𝑣 is the specific heat.
Strain rate is accounted for using the Cowper-Symonds model which, e.g., model 3,
scales the yield stress with the factor:
1
𝜀̇ ⁄𝑝 (22.103.6)
1+( ) .
𝐶
To convert these constants set the viscoelastic constants, 𝑉𝑘 and 𝑉𝑚 , to the following
values:
1
1 𝑝
𝑉𝑘 = 𝜎 ( ) ,
𝐶
(22.103.7)
1
𝑉𝑚 = .
𝑝
This model properly treats rate effects and should provide superior results to
models 3 and 24.
The transverse plate shear stresses in the principal strain directions are assumed to
be damaged as follows:
𝐷1
𝜎13 = (1 − ) 𝜎130 ,
2
(22.104.2)
𝐷
𝜎23 = (1 − 2 ) 𝜎230 .
2
In the anisotropic damage formulation, 𝐷1 (𝜀1 ) and 𝐷2 (𝜀2 ) are anisotropic damage
functions for the loading directions 1 and 2, respectively. Stresses 𝜎110 , 𝜎220 , 𝜎120 , 𝜎130 and
𝜎230 are stresses in the principal shell strain directions as calculated from the undamaged
elastic-plastic material behavior. The strains 𝜀1 and 𝜀2 are the magnitude of the principal
strains calculated upon reaching the damage thresholds. Damage can only develop for
tensile stresses, and the damage functions 𝐷1 (𝜀1 ) and 𝐷2 (𝜀2 ) are identical to zero for
negative strains 𝜀1 and 𝜀2 . The principal strain directions are fixed within an integration
point as soon as either principal strain exceeds the initial threshold strain in tension. A
more detailed description of the damage evolution for this material model is given in the
description of material 82.
⎧ 𝑌
{
{𝑆(1 − 𝐷) 𝑟 ̇ for 𝑟 > 𝑟𝐷 and 𝜎1 > 0
𝐷̇ = ⎨ . (22.104.4)
{
{ 0 otherwise
⎩
where 𝑟𝐷 is the damage threshold, 𝑌 is a positive material constant, 𝑆 is the strain energy
release rate, and 𝜎1 is the maximal principal stress. The strain energy density release rate is
2
1 𝜎vm 𝑅𝑣
𝑌 = 𝐞e : 𝐂: 𝐞e = , (22.104.5)
2 2𝐸(1 − 𝐷)2
where 𝜎vm is the equivalent von Mises stress. The triaxiality function 𝑅𝑣 is defined as
2 𝜎 2
𝑅𝑣 = (1 + 𝜈) + 3(1 − 2𝜈) ( H ) . (22.104.6)
3 𝜎vm
3
𝐿=𝑀=𝑁= , (22.104.11)
2
𝑅00 = 𝑅45 = 𝑅90 = 1, (22.104.12)
so that isotropic behavior is obtained.
Strain rate is accounted for using the Cowper-Symonds model which scales the yield
stress with the factor:
1
𝜀̇ ⁄p (22.104.13)
1+( ) .
𝐶
To convert these constants, set the viscoelastic constants, 𝑉𝑘 and 𝑉𝑚 , to the following
values:
1
1 𝑝
𝑉𝑘 = 𝜎 ( ) ,
𝐶
(22.104.14)
1
𝑉𝑚 = .
𝑝
Viscous effects are accounted for using the Cowper-Symonds model, which scales
the yield stress with the factor:
p 1⁄
p
𝜀̇ (22.106.2)
1 + ( eff ) .
𝐶
The equivalent stress for a ceramic-type material is given in terms of the damage
parameter 𝐷 by
𝜎 ∗ = 𝜎i∗ − 𝐷(𝜎i∗ − 𝜎f∗ ). (22.110.1)
Here,
𝜎i∗ = 𝑎(𝑝∗ + 𝑡∗ )𝑛 (1 + 𝑐ln𝜀̇∗ ), (22.110.2)
represents the intact, undamaged behavior. The superscript, '*', indicates a normalized
quantity. The stresses are normalized by the equivalent stress at the Hugoniot elastic limit
(see below), the pressures are normalized by the pressure at the Hugoniot elastic limit, and
the strain rate by the reference strain rate defined in the input. In this equation 𝑎 is the
intact normalized strength parameter, 𝑐 is the strength parameter for strain rate
dependence, 𝜀̇∗ is the normalized plastic strain rate, and,
𝑇
𝑡∗ = ,
PHEL
(22.110.3)
∗ 𝑝
𝑝 = ,
PHEL
where 𝑇 is the maximum tensile pressure strength, PHEL is the pressure component at the
p
𝐷 = ∑ Δ𝜀p /𝜀f , (22.110.4)
represents the accumulated damage based upon the increase in plastic strain per
computational cycle and the plastic strain to fracture
p
𝜀f = 𝑑1 (𝑝∗ + 𝑡∗ )𝑑2 , (22.110.5)
where 𝑑1 and 𝑑2 are user defined input parameters. The equation:
𝜎f∗ = 𝑏(𝑝∗ )𝑚 (1 + 𝑐ln𝜀̇∗ ) ≤ SFMAX, (22.110.6)
represents the damaged behavior where 𝑏 is an input parameter and SFMAX is the
maximum normalized fracture strength. The parameter, 𝑑1 , controls the rate at which
damage accumulates. If it approaches 0, full damage can occur in one time step, i.e.,
instantaneously. This rate parameter is also the best parameter to vary if one attempts to
reproduce results generated by another finite element program.
Given HEL and the shear modulus, 𝑔, 𝜇hel can be found iteratively from
4 𝜇
HEL = 𝑘1 𝜇hel + 𝑘2 𝜇2hel + 𝑘3 𝜇3hel + 𝑔 ( hel ), (22.110.8)
3 1 + 𝜇hel
and, subsequently, for normalization purposes,
PHEL = 𝑘1 𝜇hel + 𝑘2 𝜇2hel + 𝑘3 𝜇3hel , (22.110.9)
and
𝜎hel = 1.5(HEL − PHEL). (22.110.10)
These are calculated automatically by LS-DYNA if PHEL is zero on input.
3
̅̅̅̅ = √ 𝜎𝑖𝑗 𝜎𝑖𝑗 .
𝜎 (22.115.2)
2
The creep strain, therefore, is only a function of the deviatoric stresses. The
volumetric behavior for this material is assumed to be elastic. By varying the time constant
𝑚 primary creep (𝑚 < 1), secondary creep (𝑚 = 1), and tertiary creep (𝑚 > 1) can be
modeled. This model is described by Whirley and Henshall (1992).
This material law is based on standard composite lay-up theory. The implementa-
tion, [Jones 1975], allows the calculation of the force, 𝐍, and moment, 𝐌, stress resultants
from:
where 𝐴𝑖𝑗 is the extensional stiffness, 𝐷𝑖𝑗 is the bending stiffness, and 𝐵𝑖𝑗 is the coupling
stiffness, which is a null matrix for symmetric lay-ups. The mid-surface strains and
curvatures are denoted by 𝜀0𝑖𝑗 and 𝜅𝑖𝑗 , respectively. Since these stiffness matrices are
symmetric, 18 terms are needed per shell element in addition to the shell resultants, which
are integrated in time. This is considerably less storage than would typically be required
with through thickness integration which requires a minimum of eight history variables
per integration point, e.g., if 100 layers are used 800 history variables would be stored. Not
only is memory much less for this model, but the CPU time required is also considerably
reduced.
The calculation of the force, 𝑁𝑖𝑗 , and moment, 𝑀𝑖𝑗 , stress resultants is given in terms
of the membrane strains, 𝜀0𝑖 , and shell curvatures, 𝜅𝑖 , as:
⎧ 𝑁𝑥 ⎫ 𝐶11 𝐶12 𝐶13 𝐶14 𝐶15 𝐶16 ⎧ { 𝜀0𝑥 ⎫
}
{ } ⎡
{
{ 𝑁 }
𝑦 }
⎢𝐶21 𝐶22 𝐶23 𝐶24 𝐶25 𝐶26 ⎤ { 𝜀
⎥{ 𝑦 }
{ 0}
{
{ 𝑁𝑥𝑦 }
} ⎢𝐶31 𝐶32 𝐶33 𝐶34 𝐶35 𝐶36 ⎥ { } }
⎢ ⎥ 𝜀0𝑧 ,
⎨ 𝑀𝑥 ⎬ = ⎢𝐶41 𝐶42 𝐶43 𝐶44 𝐶45 𝐶46 ⎥ ⎨ ⎬ (22.117.1)
{
{ } ⎢ ⎥{ 𝜅𝑥 }
{ 𝑀𝑦 }} ⎢𝐶51 𝐶52 𝐶53 𝐶54 𝐶55 𝐶56 ⎥ { 𝜅𝑦 }
{
{ }
{𝑀 } ⎣𝐶61 { } }
⎩ 𝑥𝑦 ⎭ 𝐶62 𝐶63 𝐶64 𝐶65 𝐶66 ⎦ { 𝜅
⎩ 𝑥𝑦 ⎭
}
where 𝐶𝑖𝑗 = 𝐶𝑗𝑖 . In this model this symmetric matrix is transformed into the element local
system and the coefficients are stored as element history variables.
After failure the discrete element is deleted. If failure is included either the tension
failure or the compression failure or both may be used.
Unloading curve
Unload = 0 Unload = 1
Resultant
loading-unloading Resultant
curve
Displacement Displacement
Unload = 2 Unload = 3
Resultant
Resultant
Unloading curve
Unloading curve
Displacement Displacement
𝐷 = ∫ 𝜔1 𝜔2 𝑑𝜀p , (22.120.1)
The stress-strain behavior follows one curve in compression and another in tension.
The sign of the mean stress determines the state where a positive mean stress (i.e., a
negative pressure) is indicative of tension. Two load curves, 𝑓t (𝑝) and 𝑓c (𝑝), are defined,
which give the yield stress, 𝜎y , versus effective plastic strain for both the tension and
compression regimes. The two pressure values, 𝑝t and 𝑝c , when exceeded, determine if the
tension curve or the compressive curve is followed, respectively. If the pressure, 𝑝, falls
between these two values, a weighted average of the two curves are used:
⎧ 𝑝c − 𝑝
{scale =
if − 𝑝t ≤ 𝑝 ≤ 𝑝c ⎨ 𝑝c + 𝑝t . (22.124.1)
{ 𝜎 = scale ⋅ 𝑓 (𝑝) + (1 − scale) ⋅ 𝑓 (𝑝)
⎩ y t c
Strain rate is accounted for using the Cowper and Symonds model, which scales the
yield stress with the factor
1
𝜀̇ ⁄p (22.124.2)
1+( ) ,
𝐶
where 𝜀̇ is the strain rate 𝜀̇ = √𝜀̇𝑖𝑗 𝜀̇𝑖𝑗 .
The behavior before compaction is orthotropic where the components of the stress
tensor are uncoupled, i.e., a component of strain will generate resistance in the local a-
direction with no coupling to the local 𝑏 and 𝑐 directions. The elastic modulii vary from
their initial values to the fully compacted values linearly with the relative volume:
𝐸𝑎𝑎 = 𝐸𝑎𝑎𝑢 + 𝛽𝑎𝑎 (𝐸 − 𝐸𝑎𝑎𝑢 ), 𝐺𝑎𝑏 = 𝐺𝑎𝑏𝑢 + 𝛽(𝐺 − 𝐺𝑎𝑏𝑢 ),
𝐸𝑏𝑏 = 𝐸𝑏𝑏𝑢 + 𝛽𝑏𝑏 (𝐸 − 𝐸𝑏𝑏𝑢 ), 𝐺𝑏𝑐 = 𝐺𝑏𝑐𝑢 + 𝛽(𝐺 − 𝐺𝑏𝑐𝑢 ), (22.126.1)
𝐸𝑐𝑐 = 𝐸𝑐𝑐𝑢 + 𝛽𝑐𝑐 (𝐸 − 𝐸𝑐𝑐𝑢 ), 𝐺𝑐𝑎 = 𝐺𝑐𝑎𝑢 + 𝛽(𝐺 − 𝐺𝑐𝑎𝑢 ),
where
1−𝑉
𝛽 = max [min ( , 1) , 0], (22.126.2)
1 − 𝑉𝑓
and 𝐺 is the elastic shear modulus for the fully compacted honeycomb material
𝐸
𝐺= . (22.126.3)
2(1 + 𝜈)
The relative volume, 𝜈, is defined as the ratio of the current volume over the initial
volume, and typically, 𝜈 = 1 at the beginning of a calculation.
The load curves define the magnitude of the stress as the material undergoes
deformation. The first value in the curve should be less than or equal to zero correspond-
ing to tension and increase to full compaction. Care should be taken when defining the
curves so the extrapolated values do not lead to negative yield stresses.
At the beginning of the stress update we transform each element’s stresses and
strain rates into the local element coordinate system. For the uncompacted material, the
trial stress components are updated using the elastic interpolated modulii according to:
trial trial
𝑛+1 𝑛 𝑛+1 𝑛
𝜎𝑎𝑎 = 𝜎𝑎𝑎 + 𝐸𝑎𝑎 Δ𝜀𝑎𝑎 , 𝜎𝑎𝑏 = 𝜎𝑎𝑏 + 2𝐺𝑎𝑏 Δ𝜀𝑎𝑏 ,
trial trial
𝑛+1 𝑛 𝑛+1 𝑛 (22.126.4)
𝜎𝑏𝑏 = 𝜎𝑏𝑏 + 𝐸𝑏𝑏 Δ𝜀𝑏𝑏 , 𝜎𝑏𝑐 = 𝜎𝑏𝑐 + 2𝐺𝑏𝑐 Δ𝜀𝑏𝑐 ,
trial trial
𝑛+1 𝑛 𝑛+1 𝑛
𝜎𝑐𝑐 = 𝜎𝑐𝑐 + 𝐸𝑐𝑐 Δ𝜀𝑐𝑐 , 𝜎𝑐𝑎 = 𝜎𝑐𝑎 + 2𝐺𝑐𝑎 Δ𝜀𝑐𝑎 ,
We then independently check each component of the updated stresses to ensure that
they do not exceed the permissible values determined from the load curves, e.g., if
trial
∣𝜎𝑖𝑗𝑛+1 ∣ > 𝜆𝜎𝑖𝑗 (𝜀𝑖𝑗 ), (22.126.5)
then
trial
𝜆𝜎𝑖𝑗𝑛+1
𝜎𝑖𝑗𝑛+1 = 𝜎𝑖𝑗 (𝜀𝑖𝑗 ) trial
. (22.126.6)
∣𝜎𝑖𝑗𝑛+1 ∣
The components of 𝜎𝑖𝑗 (𝜀𝑖𝑗 ) are defined by load curves. The parameter 𝜆 is either
unity or a value taken from the load curve number, LCSR, that defines 𝜆 as a function of
strain-rate. Strain-rate is defined here as the Euclidean norm of the deviatoric strain-rate
tensor.
For fully compacted material we assume that the material behavior is elastic-
perfectly plastic and updated the stress components according to:
𝑛+1⁄2
𝑠trial
𝑖𝑗 = 𝑠𝑛𝑖𝑗 + 2𝐺Δ𝜀dev
𝑖𝑗 , (22.126.7)
We now check to see if the yield stress for the fully compacted material is exceeded
by comparing
⁄21
3 (22.126.9)
𝑠trial = ( 𝑠trial 𝑠 trial
) ,
eff
2 𝑖𝑗 𝑖𝑗
the effective trial stress to the yield stress, σy . If the effective trial stress exceeds the yield
stress, we simply scale back the stress components to the yield surface
𝜎y
𝑠𝑛+1
𝑖𝑗 = trial 𝑠trial
𝑖𝑗 . (22.126.10)
𝑠eff
We can now update the pressure using the elastic bulk modulus, K
𝑛+1⁄2
𝑝𝑛+1 = 𝑝𝑛 − 𝐾Δ𝜀𝑘𝑘 ,
(22.126.11)
𝐸
𝐾= ,
3(1 − 2𝜈)
and obtain the final value for the Cauchy stress
𝜎𝑖𝑗𝑛+1 = 𝑠𝑛+1
𝑖𝑗 − 𝑝𝑛+1 𝛿𝑖𝑗 . (22.126.12)
After completing the stress update we transform the stresses back to the global
configuration.
Assume that the elastic update results in a trial stress 𝜎 trial in the material coordinate
system. This stress tensor is diagonalized to obtain the principal stresses 𝜎1trial , 𝜎2trial and
𝜎3trial and the corresponding principal directions 𝐪1 , 𝐪2 and 𝐪3 relative to the material
coordinate system. The angle that each direction makes with the strong axis of anisotropy
𝐞1 is given by
σij
unloading and
reloading path
0 Strain: -εij
Unloading is based on the interpolated Young’s
moduli which must provide an unloading
tangent that exceeds the loading tangent.
Figure 22.126.1. Stress quantity versus strain. Note that the “yield stress” at a
strain of zero is nonzero. In the load curve definition the “time” value is the
directional strain and the “function” value is the yield stress.
𝜑𝑖 = arccos∣𝐪𝑖 ⋅ 𝐞1 ∣, 𝑖 = 1, 2, 3.
(22.126.13)
⎛
⎜ 𝜎 Y (𝜎 trial ) ⎞
⎟
𝜎𝑖 = 𝜎𝑖trial min ⎜
⎜
⎜ 1, ⎟
⎟
⎟ , (22.126.15)
⎜ 3 trial trial
⎟
⎝ √∑𝑘=1 𝜎𝑘 𝜎𝑘 ⎠
and the new stress is transformed back to the material coordinate system3.
This stress update is not uniquely defined when the stress tensor possesses multiple
eigenvalues, thus the following simple set of rules is applied. If all principal stresses are
equal, one of the principal directions is chosen to coincide with the strong axis of
anisotropy. If two principal stresses are equal, then one of the directions corresponding to
this stress value is chosen perpendicular to the strong axis of anisotropy.
22.126.2 Support for Independent Shear and Hydrostatic Yield Stress Limit
The model just described turned out to be weak in shear [Okuda 2003] and there
were no means of adding shear resistance without changing the behavior in pure uniaxial
compression. We propose the following modification of the model where the user can
prescribe the shear and hydrostatic resistance in the material without changing the uniaxial
behavior.
Assume that the elastic update results in a trial stress 𝜎 trial in the material coordinate
system. This stress tensor is diagonalized to obtain the principal stresses 𝜎1trial , 𝜎2trial and
𝜎3trial and the corresponding principal directions 𝐪1 , 𝐪2 and 𝐪3 relative to the material
coordinate system. For this discussion we assume that the principal stress values are
ordered so that 𝜎1trial ≤ 𝜎2trial ≤ 𝜎3trial . Two cases need to be treated.
3 Since each component of the stress tensor is scaled by the same factor in Equation 19.126.14, the stress is in
practice not transformed back but the scaling is performed on the stress in the material coordinate system.
LS-DYNA DEV 10/27/16 (r:8004) 20-213 (Material Models)
Material Models LS-DYNA Theory Manual
If 𝜎1trial ≤ −𝜎3trial then the principal stress value with largest magnitude is 𝜎1trial , and
consequently 𝜎1trial ≤ 0. Let
𝜎𝑢 = 𝜎1trial + max(∣𝜎2trial ∣, ∣𝜎3trial ∣),
1 (22.126.16)
𝜎p = {−max(∣𝜎2trial ∣, ∣𝜎3trial ∣) + 𝜎2trial + 𝜎3trial },
3
and finally
𝜎d3 = 𝜎3trial − 𝜎p .
The total stress is the sum of a uniaxial stress represented by 𝜎u , a hydrostatic stress
represented by 𝜎p and a deviatoric stress represented by 𝜎d1 , 𝜎d2 and 𝜎d3 . The angle that the
direction of 𝜎u makes with the strong axis of anisotropy 𝐞1 is given by 𝜙 = arccos∣𝐪1 ⋅ 𝐞1 ∣.
Now a limit stress for the general multiaxial stress is determined as a convex
combination of the three stress contributions as follows
𝜎 Y (𝜎 trial )
𝜎uY (𝜙, 𝜀vol )𝜎u2 + 3√3𝜎pY (𝜀vol )𝜎p2 + √2𝜎dY (𝜀vol ){(𝜎d1 )2 + (𝜎d2 )2 + (𝜎d3 )2 } (22.126.18)
= .
𝜎u2 + 3𝜎p2 + (𝜎d1 )2 + (𝜎d2 )2 + (𝜎d3 )2
Here 𝜎uY (𝜙, εvol ) is the prescribed uniaxial stress limit, 𝜎pY (𝜀vol ) is the hydrostatic stress
limit and 𝜎dY (𝜀vol ) is the stress limit in simple shear. The input for the first of these is
exactly as for the old model. The other two functions are for now written
𝜎pY (𝜀vol ) = 𝜎pY + 𝜎 S (𝜀vol ),
(22.126.19)
𝜎dY (𝜀vol ) = 𝜎dY S
+ 𝜎 (𝜀 vol
),
where 𝜎pY and 𝜎dY are user prescribed constant stress limits and 𝜎 S is the function
describing the densification of the material when loaded in the direction of the strong
material axis. We use the keyword parameters ECCU and GCAU for the new input as
follows.
• ECCU σdY , average stress limit (yield) in simple shear.
• GCAU σpY , average stress limit (yield) in hydrostatic compression (pressure).
Both of these parameters should be positive.
⎛⎜ ⎞
𝜎 Y (𝜎 trial )
⎟
𝜎𝑖 = ⎜1,
𝜎𝑖trial min ⎜ ⎟,
⎟ (22.126.20)
⎜
⎜ ⎟
⎟
3 trial trial
⎝ √ ∑ 𝜎
𝑘=1 𝑘
𝜎𝑘 ⎠
and the new stress is transformed back to the material coordinate system.
If σ3trial ≥ −σ1trial then the principal stress value with largest magnitude is σ3trial and
consequently σ3trial ≥ 0. Let
The angle that the direction of 𝜎u makes with the strong axis of anisotropy 𝐞1 is
given by 𝜙 = arccos∣𝐪3 ⋅ 𝐞1 ∣. The rest of the treatment is the same as for the case when
𝜎1trial ≤ −𝜎3trial . To motivate the model, let us consider three states of stress.
1. For a uniaxial stress 𝜎, we have 𝜎u = 𝜎 and 𝜎p = 𝜎d1 = 𝜎d2 = 𝜎d3 = 0. This leads us
to 𝜎 Y (𝜎 trial ) = 𝜎uY (𝜙, 𝜀vol ) and hence the stress level will be limited by the user
prescribed uniaxial stress limit.
1 1 11
𝑊(𝐽1 , 𝐽2 , 𝐽) = 𝑛𝑘𝜃 [ (𝐽1 − 3) + (𝐽12 − 9) + (𝐽 3 − 27)]
2 20𝑁 1050𝑁 2 1
19 519
+ 𝑛𝑘𝜃 [ 3
(𝐽14 − 81) + 4
(𝐽15 − 243)] + 𝑊H (𝐽),
7000𝑁 673750𝑁 (22.127.1)
−1⁄3
𝐽1 = 𝐼1 𝐽 ,
𝐽2 = 𝐼2 𝐽.
The hydrostatic work term is expressed in terms of the bulk modulus, 𝐾, and 𝐽, as:
𝐾
(𝐽 − 1)2 .
𝑊H (𝐽) = (22.127.2)
2
Rate effects are taken into account through linear viscoelasticity by a convolution integral
of the form:
𝑡 𝜕𝜀𝑘𝑙
𝜎𝑖𝑗 = ∫ 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) 𝑑𝜏, (22.127.3)
0 𝜕𝜏
or in terms of the second Piola-Kirchhoff stress, 𝑆𝑖𝑗 , and Green's strain tensor, 𝐸𝑖𝑗 ,
𝑡 𝜕𝐸𝑘𝑙
𝑆𝑖𝑗 = ∫ 𝐺𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) 𝑑𝜏, (22.127.4)
0 𝜕𝜏
where 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) and 𝐺𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) are the relaxation functions for the different stress
measures. This stress is added to the stress tensor determined from the strain energy
functional.
If we wish to include only simple rate effects, the relaxation function is represented
by six terms from the Prony series:
𝑁
𝑔(𝑡) = 𝛼0 + ∑ 𝛼𝑚 𝑒−𝛽𝑡 , (22.127.5)
𝑚=1
given by,
𝑛
𝑔(𝑡) = ∑ 𝐺𝑖 𝑒−𝛽𝑖𝑡 . (22.127.6)
𝑖=1
This model is effectively a Maxwell fluid which consists of a dampers and springs in series.
We characterize this in the input by shear modulii, 𝐺𝑖 , and decay constants, 𝛽𝑖 . The
viscoelastic behavior is optional and an arbitrary number of terms may be used.
The tissue model is described in terms of the energy functional in terms of the Green
strain components, 𝐸𝑖𝑗 ,
𝐶 𝑄
𝑊(𝐸, 𝐽) = (𝑒 − 1) + 𝑊H (𝐽),
2 (22.128.1)
𝑄= 𝑏1 𝐸211 + 𝑏2 (𝐸222 + 𝐸233 + 𝐸223 + 𝐸232 ) + 𝑏3 (𝐸212 + 𝐸221 + 𝐸213 + 𝐸231 ),
where the hydrostatic work term is in terms of the bulk modulus, 𝐾, and the third invariant
𝐽, as:
𝐾
(𝐽 − 1)2 .
𝑊H (𝐽) = (22.128.2)
2
The Green components are modified to eliminate any effects of volumetric work following
the procedures of Ogden.
If we wish to include only simple rate effects, the relaxation function is represented
by six terms from the Prony series:
𝑁
𝑔(𝑡) = 𝛼0 + ∑ 𝛼𝑚 𝑒−𝛽𝑡 , (22.129.5)
𝑚=1
given by,
𝑛
𝑔(𝑡) = ∑ 𝐺𝑖 𝑒−𝛽𝑖𝑡 . (22.129.6)
𝑖=1
This model is effectively a Maxwell fluid which consists of a dampers and springs in
series. We characterize this in the input by shear moduli, 𝐺𝑖 , and decay constants, 𝛽𝑖 . The
viscoelastic behavior is optional and an arbitrary number of terms may be used.
Three methods are offered to model progressive failure. The maximum principal
stress criterion detects failure if the maximum (most tensile) principal stress exceeds 𝜎max .
Upon failure, the material can no longer carry stress.
The second failure model is the smeared crack model with linear softening stress-
strain curve using equivalent uniaxial strains. Failure is assumed to be perpendicular to
the principal strain directions. A rotational crack concept is employed in which the crack
directions are related to the current directions of principal strain. Therefore crack
directions may rotate in time. Principal stresses are expressed as
𝜎1 𝐸̅̅̅̅ 0 0 𝜀̃ ̅̅̅̅1 𝜀̃1
𝐸
⎡ 1 ⎤ ⎛ 1⎞ ⎛ ⎜ ⎞
⎛ ⎞
⎜𝜎2 ⎟ = ⎢
⎢0 ̅̅̅̅2
𝐸 0 ⎥ ⎥ ⎜𝜀̃2 ⎟ = ⎜
⎜𝐸̅̅̅̅2 𝜀̃2 ⎟
⎟
⎟, (22.131.1)
⎝𝜎3 ⎠ ⎣0 0 ̅𝐸̅̅̅3 ⎦ ⎝𝜀̃3 ⎠ ⎝𝐸 ̅̅̅̅3 𝜀̃3 ⎠
̅̅̅̅1 , 𝐸
with 𝐸 ̅̅̅̅2 and 𝐸
̅̅̅̅3 being secant stiffness in the terms that depend on internal variables.
In the model developed for DYCOSS it has been assumed that there is no interaction
between the three directions in which case stresses simply follow
⎧
{𝐸𝜀̃𝑗 if 0 ≤ 𝜀̃𝑗 ≤ 𝜀̃𝑗,ini
{
{
{ 𝜀̃𝑗 − 𝜀̃𝑗,ini
{
𝜎𝑗 (𝜀̃𝑗 ) = ⎨(1 − )𝜎
̅̅̅̅ if 𝜀̃𝑗,ini < 𝜀̃𝑗 ≤ 𝜀̃𝑗,ult , (22.131.2)
{ 𝜀̃𝑗,ult − 𝜀̃𝑗,ini
{
{
{0 if 𝜀̃𝑗 > 𝜀̃𝑗,ult
{
⎩
with 𝜎
̅̅̅̅ the ultimate stress, 𝜀̃𝑗,ini the damage threshold, and 𝜀̃𝑗,ult the ultimate strain in 𝑗-
direction. The damage threshold is defined as
̅̅̅̅
𝜎
𝜀̃𝑗,ini = . (22.131.3)
𝐸
The ultimate strain is obtained by relating the crack growth energy and the
dissipated energy
∫∫𝜎
̅̅̅̅ 𝑑𝜀̃𝑗,ult 𝑑𝑉 = 𝐺𝐴, (22.131.4)
with 𝐺 the energy release rate, 𝑉 the element volume and 𝐴 the area perpendicular to the
principal strain direction. The one-point elements in LS-DYNA have a single integration
point and the integral over the volume may be replaced by the volume. For linear
softening it follows
2𝐺𝐴
𝜀̃𝑗,ult = , (22.131.5)
𝑉𝜎̅̅̅̅
The third model is a damage model presented by Brekelmans et. al. [1991]. Here
the Cauchy stress tensor 𝜎 is expressed as
𝜎 = (1 − 𝐷)𝐸𝜀, (22.131.6)
where 𝐷 represents the current damage and the factor (1 − 𝐷) is the reduction factor
caused by damage. The scalar damage variable is expressed as a function of a so-called
damage equivalent strain 𝜀d
𝜀ini (𝜀ult − 𝜀d )
𝐷 = 𝐷(𝜀d ) = 1 − , (22.131.7)
𝜀d (𝜀ult − 𝜀ini )
with ultimate and initial strains as defined by
2
𝑘−1 1 𝑘−1 6𝑘 (22.131.8)
𝜀d = 𝐽1 + √ ( 𝐽1 ) + 𝐽 ,
2𝑘(1 − 2𝑣) 2𝑘 1 − 2𝑣 (1 + 𝑣)2 2
where the constant 𝑘 represents the ratio of the strength in tension over the strength in
compression
𝜎ult,tension
𝑘= , (22.131.9)
𝜎ult,compression
𝐽1 and 𝐽2 are the first and the second invariants of the strain tensor representing the
volumetric and the deviatoric straining, respectively
𝐽1 = 𝜀𝑖𝑗 ,
(22.131.10)
𝐽2 = 𝜀′𝑖𝑗 𝜀′𝑖𝑗 ,
where 𝜀′𝑖𝑗 denotes the deviatoric part of the strain tensor.
If the compression and tension strength are equal the dependency on the volumetric
strain vanishes in Equation (22.131.8) and failure is shear dominated. If the compressive
strength is much larger than the strength in tension, 𝑘 becomes large and the 𝐽1 terms in
Equation (22.131.8) dominate the behavior.
The 𝑋′𝑖 and 𝑋′′𝑖 are the eigenvalues of 𝑋′𝑖𝑗 and 𝑋′′𝑖𝑗 and are given by
1
𝑋′1 = (𝑋′11 + 𝑋′22 + √(𝑋′11 − 𝑋′22 )2 + 4𝑋′212 ) ,
2
(22.133.3)
1
𝑋′2 = (𝑋′11 + 𝑋′22 − √(𝑋′11 − 𝑋′22 )2 + 4𝑋′212 ),
2
and
1
𝑋 ′′1 = (𝑋 ′′11 + 𝑋 ′′22 + √(𝑋′′11 − 𝑋′′22 )2 + 4𝑋′′212 ) ,
2
(22.133.4)
1
𝑋 ′′2 = (𝑋 ′′11 + 𝑋 ′′22 − √(𝑋′′11 − 𝑋′′22 )2 + 4𝑋′′212 ),
2
respectively. The 𝑋′𝑖𝑗 and 𝑋′′𝑖𝑗 are given by4
𝑋′ 𝐿′ 𝐿′12 0 𝑠𝑥𝑥
⎛ 11 ⎞ ⎛ 11 ⎞ ⎛𝑠 ⎞
⎜𝑋′22 ⎟ = ⎜𝐿′21 𝐿′22 0 ⎟ ⎜ 𝑦𝑦 ⎟ ,
⎝𝑋′12 ⎠ ⎝0 0 𝐿′33 ⎠ ⎝𝑠𝑥𝑦 ⎠
(22.133.5)
𝑋 ′′11
𝐿 ′′11
𝐿′′12
0 𝑠𝑥𝑥
⎛𝑋 ′′22 ⎟
⎜ ⎛𝐿′′21
⎞=⎜ 𝐿′′22 0 ⎟ ⎞ ⎜𝑠𝑦𝑦 ⎞
⎛ ⎟,
⎝𝑋 ′′12 ⎠ ⎝0 ′′33 ⎠ ⎝𝑠𝑥𝑦 ⎠
0 𝐿
where
𝐿′11 2 0 0
⎛
⎜ ⎞
⎟
⎜𝐿′ ⎟ ⎛−1 0 0⎞ 𝛼1
⎜
⎜
12 ⎟
⎟ 1⎜
⎜
⎜0
⎟
⎟
⎟ ⎛ ⎞,
⎜
⎜𝐿′ 21 ⎟
⎟ = ⎜
⎜ −1 ⎟
0⎟ ⎜ 𝛼2 ⎟ (22.133.6)
⎜
⎜ ⎟ 3⎜ ⎜0 ⎟
⎟
0 ⎝𝛼7 ⎠
⎜𝐿′22 ⎟⎟ 2
⎝0 0 3⎠
⎝𝐿′33 ⎠
𝐿′′11 −2 2 8 −2 0 𝛼3
⎛
⎜ ′′12 ⎞
⎟
𝐿 ⎟ 1⎜ 1 ⎛
⎜ −4 −4 4 0⎞⎟ ⎛
⎜ 𝛼4 ⎞
⎟
⎜
⎜ ⎟= ⎜ ⎟
⎟ ⎜
⎜ ⎟
⎟.
⎜
⎜𝐿′′21 ⎟
⎟ 9⎜
⎜ 4 −4 −4 1 0⎟⎟ ⎜
⎜ 𝛼5⎟⎟
⎜
⎜𝐿′′22 ⎟ ⎜
⎜−2 ⎟ ⎜ ⎟
⎟ 8 2 −2 0⎟ ⎜𝛼6 ⎟
⎝𝐿′′33 ⎠ ⎝0 0 0 0 9⎠ ⎝𝛼8 ⎠
where 𝛼1 to 𝛼8 are the parameters that determine the shape of the yield surface.
𝜎Yt (𝜀p , 𝜀̇p , 𝛽) = 𝜎Yv (𝜀p , 𝜀̇p ) + 𝛽(𝜎0 − 𝜎Yv (𝜀p , 𝜀̇p )), (22.133.7)
where 𝛽 determines the fraction kinematic hardening and 𝜎0 is the initial yield stress. The
yield stress for purely isotropic hardening is given by
1
⎛ 𝜀̇p p ⎞
𝜎Yv (𝜀p , 𝜀̇p )⎜
= 𝜎Y (𝜀p ) ⎜1 + { } ⎟
⎜ ⎟, (22.133.8)
𝐶 ⎟
⎝ ⎠
where 𝐶 and 𝑝 are the Cowper-Symonds material parameters for strain rate effects.
For the plastic return algorithms, the gradient of effective stress is computed as
∂𝜎eff 𝜕𝜙′
⎛
⎜ ⎞
⎟ ⎛
⎜ ⎞
⎜
⎜ ∂𝑠 xx ⎟
⎟ ⎜
⎜ 𝜕𝑋′11 ⎟⎟
⎟
⎜
⎜ ⎟
⎟ 𝑎−1 L′11 L′21 0 ⎜ ⎜ ⎟
⎟
⎜ ∂𝜎 eff ⎟ 𝑎𝜎eff ⎛ ⎞ ⎜ 𝜕𝜙′ ⎟
⎜
⎜ ⎟
⎟ = ⎜ ⎟ ⎜ ⎟
⎜
⎜ ∂𝑠 ⎟
⎟ ⎜L′12 L′22 0 ⎟⎜ ⎜ ⎟
⎟
⎜ yy ⎟ 2 ⎜ 𝜕𝑋′ 22 ⎟
⎜
⎜ ⎟
⎟ ⎝ 0 0 ⎜
L′33 ⎠ ⎜ ⎟
⎟
⎜∂𝜎eff ⎟
⎜ ⎟ ⎜
⎜ 𝜕𝜙′ ⎟
⎟
⎝ ∂𝑠xy ⎠ ⎝𝜕𝑋′12 ⎠
(22.133.10)
∂𝜙′′
⎛
⎜ ⎞
⎟
⎜
⎜ ∂X′′ 11 ⎟
⎟
𝑎−1 L′′11 L′′21 0 ⎜
⎜ ⎟
⎟
𝑎𝜎eff ⎛
⎜ ⎞
⎟ ⎜
⎜ ∂𝜙′′ ⎟
⎟
+ ⎜L′′12 L′′22 0 ⎟⎜ ⎜ ⎟
⎟ ,
2 ⎜ ∂X′′ 22 ⎟
⎝ 0 0 ⎜
L′′33 ⎠ ⎜ ⎟
⎟
⎜
⎜ ∂𝜙′′ ⎟ ⎟
⎝∂X′′12 ⎠
The algorithm for the plane stress update as well as the formula for the tangent
modulus is given in detail in Section 19.36.1 and is not repeated here.
20-226 (Material Models) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Material Models
trial 1 2
𝑓3 = 𝜎33 (Δ𝜀33 ) + 2𝐺Δ𝜆(∇𝜎eff (t) + ∇𝜎eff (t)) = 0, (22.133.24)
where
𝛽
ℎ= 𝐁∇𝜎eff (𝑡), (22.133.25)
∇𝜎eff (𝑡)T 𝐁∇𝜎eff (𝑡)
in terms of the unknown variables 𝐭 (stress), Δε33 (thickness strain increment) and Δ𝜆
(plastic strain increment). In the above
1 2 1 2 1 t
∂σY
⎡
𝐃=⎢ 1 ⎤, ⎡
𝐀 = ⎢1 2 ⎤, ⎡
𝐁 = ⎢1 2 ⎤, 𝐻= . (22.133.26)
⎥ ⎥ ⎥
⎣ 0.5⎦ ⎣ 1 ⎦ ⎣ 0.5⎦ ∂εp
This system of equations is solved using a Newton method with an additional line
search for robustness. Using the notation
𝑓 t
⎡ 1⎤
𝐟 = ⎢𝑓2 ⎥ , 𝐱=⎡
⎢Δ𝜆 ⎥,
⎤ (22.133.27)
⎣𝑓3 ⎦ ⎣Δ𝜀33 ⎦
a Newton step is completed as
−1
∂𝑓
𝑥+ = 𝑥− − 𝑠 ( ) f (22.133.28)
∂𝑥
for a step size 𝑠 ≤ 1 chosen such that the norm of the objective function is decreasing. The
gradient of the objective function is given by
∇𝑓 = ∇𝑓1−β + ∇𝑓β (22.133.29)
where
I + 2𝐺Δ𝜆𝐷∇2 𝜎eff 2𝐺𝐷∇𝜎eff −𝐶3
⎡ ⎤
∇f1−β =⎢
⎢ −(∇𝜎eff )T 𝐻 0 ⎥⎥ (22.133.30)
⎣ 2𝐺Δ𝜆𝐞T ∇2 𝛔eff T
2Ge ∇σeff C33 ⎦
∂h
⎡Δ𝜆𝐻 𝐻ℎ 0⎤
∇fβ = ⎢
⎢ ∂𝑡 ⎥
⎥ (22.133.31)
⎢0T 0 0⎥
⎣0 0 0⎦
and
1 2G 4G
𝐞=⎡
⎢1⎤⎥, C3 = (K − ) e, C33 = (K + ), (22.133.32)
⎣0⎦ 3 3
𝐺 and 𝐾 stands for the shear and bulk modulus, respectively. This algorithm requires
computation of the effective stress hessian. The derivation of this is quite straightforward
but the expression for it is rather long and is hence omitted in this report.
Rate effects are taken into account through linear viscoelasticity by a convolution
integral of the form:
𝑡 𝜕𝜀𝑘𝑙
𝜎𝑖𝑗 = ∫ 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) 𝑑𝜏, (22.134.1)
0 𝜕𝜏
where 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) is the relaxation function.
If we wish to include only simple rate effects for the deviatoric stresses, the
relaxation function is represented by six terms from the Prony series:
𝑁
𝑔(𝑡) = ∑ 𝐺𝑚 𝑒−𝛽𝑚 𝑡 . (22.134.2)
𝑚=1
For volumetric relaxation, the relaxation function is also represented by the Prony
series in terms of bulk modulii:
𝑁
𝑘(𝑡) = ∑ 𝐾𝑚 𝑒−𝛽𝑘𝑚 𝑡 . (22.134.3)
𝑚=1
Axial collapse occurs when the compressive axial load reaches the collapse load.
Collapse load versus collapse deflection is specified in the form of a load curve. The points
of the load curve are either (true strain, collapse force) or (change in length, collapse force).
Both quantities should be entered as positive for all points, and will be interpreted as
compressive. The first point should be (zero, initial collapse load).
The collapse load may vary with end moment as well as with deflections. In this
case several load-deflection curves are defined, each corresponding to a different end
moment. Each load curve should have the same number of points and the same deflection
values. The end moment is defined as the average of the absolute moments at each end of
the beam and is always positive.
Moment Interaction
Plastic hinges can form due to the combined action of moments about the three axes.
This facility is activated only when yield moments are defined in the material input. A
hinge forms when the following condition is first satisfied.
2 2 2
𝑀r 𝑀s 𝑀t
( ) +( ) +( ) ≥ 1, (22.139.3)
𝑀ryield 𝑀syield 𝑀tyield
where,
𝑀r , 𝑀s , 𝑀t = current moments
𝑀ryield , 𝑀syield , 𝑀tyield = yield moments
Note that scale factors for hinge behavior defined in the input will also be applied to the
yield moments: for example, 𝑀syield in the above formula is given by the input yield
moment about the local axis times the input scale factor for the local s-axis. For strain-
softening characteristics, the yield moment should generally be set equal to the initial peak
of the moment-rotation load curve.
On forming a hinge, upper limit moments are set. These are given by
𝑀
𝑀rupper = MAX ⎜⎛𝑀r , ryield ⎟
⎞, (22.139.4)
⎝ 2 ⎠
and similarly for 𝑀s and 𝑀t . Thereafter the plastic moments will be given by:
𝑀rp = min(𝑀rcurve , 𝑀rcurve )
and similarly for 𝑠 and 𝑡, where
𝑀rp = current plastic moment
𝑀rcurve = moment taken from load curve at the current rotation scaled according to
the scale factor.
The effect of this is to provide an upper limit to the moment that can be generated; it
represents the softening effect of local buckling at a hinge site. Thus if a member is bent
about its local s-axis it will then be weaker in torsion and about its local 𝑡-axis. For
moment-softening curves, the effect is to trim off the initial peak (although if the curves
subsequently harden, the final hardening will also be trimmed off).
It is not possible to make the plastic moment vary with the current axial load, but it
is possible to make hinge formation a function of axial load and subsequent plastic moment
a function of the moment at the time the hinge formed. This is discussed in the next
section.
Upon forming a hinge, the magnitude of that component of moment will not be
permitted to exceed the current plastic moment. The current plastic moment is obtained by
interpolating between the plastic moment vs. plastic rotation curves input on cards 10, 12,
14, 16, or 18. Curves may be input for up to 8 hinge moments, where the hinge moment is
defined as the yield moment at the time that the hinge formed. Curves must be input in
order of increasing hinge moment and each curve should have the same plastic rotation
values. The first or last curve will be used if the hinge moment falls outside the range of
the curves. If no curves are defined, the plastic moment is obtained from the curves on
cards 4 through 6. The plastic moment is scaled by the scale factors on lines 4 to 6.
A hinge will form if either the independent yield moment is exceeded or if the
moment interaction equation is satisfied. If both are true, the plastic moment will be set to
the minimum of the interpolated value and 𝑀rp .
M8
M7
M6
M5
Axial Force
M4
M3
M2
M1
Figure 22.139.1. The force magnitude is limited by the applied end moment. For
an intermediate value of the end moment LS-DYNA interpolates between the
curves to determine the allowable force value.
1 𝑍2 𝑆𝑖𝑗 − 𝛺𝑖𝑗
𝜀𝑖𝑗 = 𝐷o exp [− ( o )] ⎛
⎜ ⎞
⎟, (22.141.1)
2 3𝐾2 ⎝ √𝐾2 ⎠
where 𝐷o is the maximum inelastic strain rate, 𝑍o is the isotropic initial hardness of
material, 𝛺𝑖𝑗 is the internal stress, 𝑆𝑖𝑗 is the deviatoric stress component, and 𝐾2 is defined
as follows:
1
𝐾2 = (𝑆𝑖𝑗 − 𝛺𝑖𝑗 )(𝑆𝑖𝑗 − 𝛺𝑖𝑗 ), (22.141.2)
2
and represent the second invariant of the overstress tensor. The elastic components of the
strain are added to the inelastic strain to obtain the total strain. The following relationship
defines the internal stress variable rate:
2
𝛺𝑖𝑗 = 𝑞𝛺m 𝜀̇I𝑖𝑗 − 𝑞𝛺𝑖𝑗 𝜀̇Ie , (22.141.3)
3
where 𝑞 is a material constant, 𝛺m is a material constant that represents the maximum
value of the internal stress, and 𝜀̇Ie is the effective inelastic strain.
Many polymers used for energy absorption are low density, crushable foams with
no noticeable Poisson effect. Frequently manufactured by extrusion, they are transversely
isotropic. This class of material is used to enhance automotive safety in low velocity
(bumper impact) and medium velocity (interior head impact) applications. These materials
require a transversely isotropic, elastoplastic material with a flow rule allowing for large
permanent volumetric deformations.
The MAT_HONEYCOMB model uses a local coordinate system defined by the user.
One of the axes of the local system coincides with the extrusion direction of the honeycomb
in the undeformed configuration. As an element deforms, its local coordinate system
rotates with its mean rigid body motion. Each of the six stress components is treated
independently, and each has its own law relating its flow stress to its plastic strain.
where 𝜃 is the angle of the local coordinate system relative to the global system. For
uniaxial loading along the global 1-axis, the stress will be (accounting for the sign of the
volume strain),
y y y
𝜎11 = {[cos(𝜃)]2 𝜎11 + [sin(𝜃)]2 𝜎22 + 2 ⋅ sin(𝜃)cos(𝜃)𝜎12 }sgn(𝜀V )}, (22.142.3)
assuming the strain is large enough to cause yielding in both directions.
y y
If the shear strength is neglected, 𝜎11 will vary smoothly between 𝜎11 and 𝜎22 and
never exceed the maximum of the two yield stresses. This behavior is intuitively what we
would like to see. However, if the value of shear yield stress isn’t zero, 𝜎11 will be greater
y y y y
than either 𝜎11 or 𝜎22 . To illustrate, if 𝜎11 and 𝜎22 are equal (a nominally isotropic
response) the magnitude of the stress is
y y
|𝜎11 | = 𝜎11 + 2 ⋅ sin(𝜃)cos(𝜃)𝜎12 , (22.142.4)
and achieves a maximum value at 45 degrees of
y y
|𝜎11 | = 𝜎11 + 𝜎12 . (22.142.5)
For cases where there is anisotropy, the maximum occurs at a different angle and
will have a different magnitude, but it will exceed the maximum uniaxial yield stress. In
fact, a simple calculation using Mohr’s circle shows that the maximum value will be
y 1 y y 1 y y 2 y
𝜎max = (𝜎11 + 𝜎22 ) + √(𝜎11 − 𝜎22 ) + 4𝜎12 , (22.142.6)
2 2
The general constitutive relation for an orthotropic material, written in terms of the
principal material directions [Bodig & Jayne, 1993] is:
𝜎1 𝐶 𝐶12 𝐶13 0 0 0 𝜀1
⎡𝜎2 ⎤ ⎡𝐶11 0 ⎤
⎢𝜎 ⎥ ⎢ 21 𝐶22 𝐶23 0 0 ⎥ ⎢ 𝜀2 ⎤
⎡
⎢ 3⎥ = ⎢
⎢𝐶31 𝐶32 𝐶33 0 0 0 ⎥ ⎥=⎢ 𝜀3 ⎥
⎥.
⎢𝜎4 ⎥ (22.143.1)
⎢ ⎥ ⎢ 0 0 0 2𝐶44 0 0 ⎥ ⎢ 𝜀 ⎥
⎥ ⎢
4⎥
⎢
⎢𝜎5 ⎥ ⎢ 0 0 0 0 2𝐶55 0 ⎥ ⎢ 𝜀5 ⎥
⎣𝜎6 ⎦ ⎣ 0 0 0 0 0 2𝐶66 ⎦ ⎣𝜀6 ⎦
The subscripts 1, 2, and 3 refer to the longitudinal, tangential, and radial, stresses and
strains (1 = 11 , 2 = 22 , 3 = 33 , 1 = 11 , 2 = 22 , 3 = 33 ), respectively. The
subscripts 4, 5, and 6 are in a shorthand notation that refers to the shearing stresses and
strains 4 = 12 , 5 = 13 , 6 = 23 , 4 = 12 , 5 = 13 , 6 = 23 ). As an alternative notation
for wood, it is common to substitute L (longitudinal) for 1, T (tangential) for 2, and R
(radial) for 3. The components of the constitutive matrix, 𝐶𝑖𝑗 , are listed here in terms of the
nine independent elastic constants of an orthotropic material:
E11 (1 − ν23 ν32 )
C11 = ,
Δ
E (1 − ν31 ν13 )
C22 = 22 ,
Δ
E (1 − ν12 ν21 )
C33 = 33 ,
Δ
(ν + ν31 ν23 )E11
C12 = 21 ,
Δ
(ν + ν21 ν32 )E11 (22.143.2)
C13 = 31 ,
Δ
(ν + ν12 ν31 )E22
C23 = 32 ,
Δ
C44 = G12 ,
C55 = G13 ,
C66 = G23 ,
Δ = 1 − ν12 ν21 − ν23 ν32 − ν31 ν13 − 2ν21 ν32 ν13 .
The following identity, relating the dependent (minor Poisson’s ratios ν21 , ν31 , and
ν32 ) and independent elastic constants, is obtained from symmetry considerations of the
constitutive matrix:
𝜈𝑖𝑗 𝜈𝑗𝑖
= for 𝑖, 𝑗 = 1, 2, 3. (22.143.3)
𝐸𝑖 𝐸𝑗
One common assumption is that wood materials are transversely isotropic. This
means that the properties in the tangential and radial directions are modeled the same, i.e.
𝐸22 = 𝐸33 , 𝐺12 = 𝐺13 , and 12 = 13 . This reduces the number of independent elastic
constants to five, 𝐸11 , 𝐸22 , 12 , 𝐺12 , and 𝐺23 . Further, Poisson's ratio in the isotropic plane,
23 , is not an independent quantity. It is calculated from the isotropic relation: = (𝐸 −
2𝐺)/2𝐺 where E = 𝐸22 = 𝐸33 and 𝐺 = 𝐺23 . Transverse isotropy is a reasonable assumption
because the difference between the tangential and radial properties of wood (particularly
Southern yellow pine and Douglas fir) is small in comparison with the difference between
the tangential and longitudinal properties.
The yield surfaces parallel and perpendicular to the grain are formulated from six
ultimate strength measurements obtained from uniaxial and pure-shear tests on wood
specimens:
XT Tensile strength parallel to the grain
XC Compressive strength parallel to the grain
YT Tensile strength perpendicular to the grain
YC Compressive strength perpendicular to the grain
S|| Shear strength parallel to the grain
S Shear strength perpendicular to the grain
Here 𝑋 and 𝑌 are the strengths parallel and perpendicular to the grain, respectively,
and S are the shear strengths. The formulation is based on the work of Hashin [1980].
For the parallel modes, the yield criterion is composed of two terms involving two of
the five stress invariants of a transversely isotropic material. These invariants are 𝐼1 = 𝜎11
2 2
and 𝐼4 = 𝜎12 + 𝜎13 This criterion predicts that the normal and shear stresses are mutually
weakening, i.e. the presence of shear stress reduces the strength below that measured in
uniaxial stress tests. Yielding occurs when 𝑓|| ≥ 0, where:
2 2 2
𝜎11 (𝜎12 + 𝜎13 ) 𝑋𝑡 for 𝜎11 > 0
𝑓|| = + −1 𝑋={ . (22.143.4)
𝑋2 𝑆2|| 𝑋𝑐 for 𝜎11 < 0
For the perpendicular modes, the yield criterion is also composed of two terms
involving two of the five stress invariants of a transversely isotropic material. These
invariants are 𝐼2 = 22 + 33 and 𝐼3 = 𝜎23
2
− 𝜎22 𝜎33 . Yielding occurs when 𝑓 0, where:
Figure 22.143.1. The yield criteria for wood produces smooth surfaces in stress
space.
2
(𝜎22 + 𝜎33 )2 (𝜎23 − 𝜎22 𝜎33 ) 𝑌t for 𝜎22 + 𝜎33 > 0
𝑓⊥ = + −1 𝑌={ (22.143.5)
𝑌 2 𝑆2⊥ 𝑌c for 𝜎22 + 𝜎33 < 0
Each yield criterion is plotted in 3D in Figure 22.143.1 in terms of the parallel and
perpendicular stresses. Each criterion is a smooth surface (no corners).
The plasticity algorithms limit the stress components once the yield criteria in
[Murry 2002] are satisfied. This is done by returning the trial elastic stress state back to the
yield surface. The stress and strain tensors are partitioned into elastic and plastic parts.
Partitioning is done with a return mapping algorithm which enforces the plastic
consistency condition.
Separate plasticity algorithms are formulated for the parallel and perpendicular
modes by enforcing separate consistency conditions. The solution of each consistency
condition determines the consistency parameters, and . The solutions, in turn,
determine the stress updates. No input parameters are required.
The stresses are readily updated from the total strain increments and the consistency
parameters, as follows:
𝜕𝑓
̅𝜎̅̅̅𝑖𝑗𝑛+1 = 𝜎𝑖𝑗∗𝑛+1 − 𝐶𝑖𝑗𝑘𝑙 Δ𝜆 ∣ (22.143.6)
𝜕𝜎𝑘𝑙 𝑛
20-238 (Material Models) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Material Models
For each mode (parallel and perpendicular), the user inputs two parameters: the
initial yield surface location in uniaxial compression, 𝑁, and the rate of translation, 𝑐. Say
the user wants pre-peak nonlinearity to initiate at 70% of the peak strength. The user will
input 𝑁 = 0.3 so that 1 − 𝑁 = 0.7. If the user wants to harden rapidly, then a large value of
𝑐 is input, like 𝑐 = 1 msec. If the user wants to harden gradually, then a small value of 𝑐 is
input, like 𝑐 = 0.2 msec.
The state variable that defines the translation of the yield surface is known as the
back stress, and is denoted by 𝑖𝑗 . Hardening is modeled in compression, but not shear, so
the only back stress required for the parallel modes is 11 . The value of the back stress is
11 = 0 upon initial yield and 11 = −𝑁|| 𝑋𝑐 at ultimate yield (in uniaxial compression).
The maximum back stress occurs at ultimate yield and is equal to the total translation of the
longitudinal stress
XT
Tension Ultimate Yield Strength
Xc
CII large
|longitudinal stress|
CII small
Increasing Rate of Translation NIIXc
(1-NII)Xc
Initial Yield Surface Initial Yield Strength
-(1-NII)Xc
-xc
Ultimate Yield Surace
Compression
SII
Strain
Square root of parallel shear invariant
yield surface in stress space. The back stress components required for the perpendicular
modes are 22 and 33 . The value of the backstress sum is 22 + 33 = 0 upon initial yield
and 22 + 33 = −𝑁 𝑌𝑐 at ultimate yield (biaxial compression without shear).
Separate damage formulations are modeled for the parallel and perpendicular
modes. These formulations are loosely based on the work of Simo and Ju [1987]. If failure
occurs in the parallel modes, then all six stress components are degraded uniformly. This
is because parallel failure is catastrophic, and will render the wood useless. If failure
occurs in the perpendicular modes, then only the perpendicular stress components are
degraded. This is because perpendicular failure is not catastrophic: we expect the wood to
continue to carry load in the parallel direction. Based on these assumptions, the following
degradation model is implemented:
𝑑m = max(𝑑(𝜏|| ), 𝑑(𝜏⊥ )) ,
𝑑|| = 𝑑(𝜏|| ),
𝜎11 = (1 − 𝑑|| )𝜎
̅̅̅̅11 ,
𝜎22 = (1 − 𝑑m )𝜎
̅̅̅̅22 ,
(22.143.8)
𝜎33 = (1 − 𝑑m )𝜎
̅̅̅̅33 ,
σ12 = (1 − 𝑑|| )𝜎
̅̅̅̅12 ,
𝜎13 = (1 − 𝑑|| )𝜎
̅̅̅̅13 ,
𝜎23 = (1 − 𝑑m )𝜎
̅̅̅̅23 .
Here, each scalar damage parameter, 𝑑, transforms the stress tensor associated with the
undamaged state, 𝜎 ̅̅̅̅𝑖𝑗 , into the stress tensor associated with the damaged state, 𝑖𝑗 . The
stress tensor 𝜎
̅̅̅̅𝑖𝑗 is calculated by the plasticity algorithm prior to application of the damage
model. Each damage parameter ranges from zero for no damage and approaches unity for
maximum damage. Thus 1 − 𝑑 is a reduction factor associated with the amount of damage.
Each damage parameter evolves as a function of a strain energy-type term. Mesh size
dependency is regulated via a length scale based on the element size (cube root of volume).
Damage-based softening is brittle in tension, less brittle in shear, and ductile (no softening)
in compression, as demonstrated in Figure 22.143.1.
Element erosion occurs when an element fails in the parallel mode, and the parallel
damage parameter exceeds 𝑑|| = 0.99. Elements do not automatically erode when an
element fails in the perpendicular mode. A flag is available, which, when set, allows
elements to erode when the perpendicular damage parameter exceeds 𝑑 = 0.99. Setting
this flag is not recommended unless excessive perpendicular damage is causing
computational difficulties.
Figure 19.143.3. Softening response modeled for parallel modes of Southern yellow
pine.
Data available in the literature for pine [Reid & Peng, 1997] indicates that dynamic
strength enhancement is more pronounced in the perpendicular direction than in the
parallel direction. Therefore, separate rate effects formulations are modeled for the parallel
and perpendicular modes. The formulations increase strength with increasing strain rate
by expanding each yield surface:
𝜎11 = 𝑋 + 𝐸11 𝜀̇ 𝜂|| Parallel
. (22.143.9)
𝜎22 = 𝑌 + 𝐸22 𝜀̇ 𝜂⊥ Perpendicular
Here 𝑋 and 𝑌 are the static strengths, 11 and 22 are the dynamic strengths, and 𝐸11 𝜀̇ 𝜂||
and 𝐸22 𝜀̇ 𝜂⊥ are the excess stress components. The excess stress components depend on the
value of the fluidity parameter, , as well as the stiffness and strain rate. The user inputs
two values, 0 and 𝑛, to define each fluidity parameter:
𝜂0||
𝜂|| = 𝑛|| ,
𝜀̇
(22.143.10)
𝜂0⊥
𝜂 ⊥ = 𝑛⊥ .
𝜀̇
The two parameter formulation [Murray, 1997] allows the user to model a nonlinear
variation in dynamic strength with strain rate. Setting 𝑛 = 0 allows the user to model a
linear variation in dynamic strength with strain rate.
where 𝑃 is the pressure, 𝜙 is the internal friction angle, 𝐾(𝜃) is a function of the angle in
deviatoric plane, √𝐽2 is the square root of the second invariant of the stress deviator, 𝑐 is the
amount of cohesion and
3√3𝐽3
cos3𝜃 = 3
, (22.147.3)
2𝐽22
J3 is the third invariant of the stress deviator, AHYP is a parameter for determining how
close to the standard Mohr-Coulomb yield surface the modified surface is fitted. If the user
defined parameter, AHYP, is input as zero, the standard Mohr-Coulomb surface is
recovered. The parameter aℎypshould be set close to zero, based on numerical
considerations, but always less than 𝑐 cot𝜙. It is best not to set the cohesion, 𝑐, to very small
values as this causes excessive iterations in the plasticity routines.
To generalize the shape in the deviatoric plane, we have changed the standard
Mohr- Coulomb 𝐾(𝜃) function to a function used by Klisinski [1985]
4(1 − 𝑒2 )cos2 𝜃 + (2𝑒 − 1)2
𝐾(𝜃) = 1
, (22.147.4)
2(1 − 𝑒2 )cos𝜃 + (2𝑒 − 1)[4(1 − 𝑒2 )cos2 𝜃 + 5𝑒2 − 4𝑒]2
where 𝑒 is a material parameter describing the ratio of triaxial extension strength to triaxial
compression strength. If e is set equal to 1, then a circular cone surface is formed. If 𝑒 is set
to 0.55, then a triangular surface is found, 𝐾(𝜃) is defined for 0.5 < 𝑒 ≤ 1.0.
𝜑 − 𝜑init
Δ𝜑 = 𝐸t (1 − ) Δ𝜀eff plas . (22.147.5)
𝐴𝑛 𝜑max
where 𝜀eff plas is the effective plastic strain. 𝐴𝑛 is the fraction of the peak strength internal
friction angle where nonlinear behavior begins, 0 < 𝐴𝑛 ≤ 1. The input parameter 𝐸𝑡
determines the rate of the nonlinear hardening.
To simulate the effects of moisture and air voids including excess pore water
pressure, both the elastic and plastic behaviors can be modified. The bulk modulus is
𝐾𝑖
𝐾= . (22.147.6)
1 + 𝐾𝑖 𝐷1 𝑛cur
where
𝐾𝑖 = initial bulk modulus
𝑛cur = current porosity = Max[0, (𝑤 − 𝜀v )]
𝑤 = volumetric strain corresponding to the volume of air voids = 𝑛(1 − 𝑆)
𝜀v = total volumetric strain
𝐷1 = material constant controlling the stiffness before the air voids are collapsed
𝑒
𝑛 = porosity of the soil =
1+𝑒
γsp (1 + mc )
𝑒 = void ratio = −1
ρ
u
p
𝜌𝑚c
𝑆 = degree of saturation =
𝑛(1 + 𝑚c )
and 𝜌, 𝛾, 𝑚c are the soil density, specific gravity, and moisture content, respectively.
Note that the model is following the standard practice of assuming compressive
stresses and strains are positive. If the input parameter 𝐷1 is zero, then the standard linear
elastic bulk modulus behavior is used.
To simulate the loss of shear strength due to excess pore water effects, the model
uses a standard soil mechanics technique [Holtz and Kovacs, 1981] of reducing the total
pressure, 𝑃, by the excess pore water pressure, 𝑢, to get an “ effective pressure”, 𝑃′;
therefore,
𝑃′ = 𝑃 − 𝑢. (22.147.8)
Figure 22.147.2 shows pore water pressure will affect the algorithm for the plasticity
surface. The excess pore water pressure reduces the total pressure, which will lower the
shear strength, √𝐽2 . A large excess pore water pressure can cause the effective pressure to
become zero.
To calculate the pore water pressure, 𝑢, the model uses an equation similar to the
equation used for the moisture effects on the bulk modulus.
𝐾sk
𝑢= 𝜀 , (22.147.9)
1 + 𝐾sk 𝐷2 𝑛cur 𝑣
where
𝐾sk = bulk modulus for soil without air voids (skeletal bulk modulus)
𝑛cur = current porosity = Max[0, (𝑤 − 𝜀𝑣 )]
𝑤 = volumetric strain corresponding to the volume of air voids = 𝑛(1 − 𝑆)
𝜀v = total volumetric strain
𝐷2 = material constant controlling the pore water pressure before
the air voids are collapsed to 𝐷2 ≥ 0
𝑒
𝑛 = porosity of the soil =
1+𝑒
γsp (1 + mc )
𝑒 = void ratio = −1
ρ
𝜌𝑚c
𝑆 = degree of saturation =
𝑛(1 + 𝑚c )
and 𝜌, 𝛾, 𝑚c are the soil density, specific gravity, and moisture content, respectively. The
pore water pressure will not be allowed to become negative, 𝑢 ≥ 0.
Figure 22.147.3 is a plot of the pore pressure versus volumetric strain for different
parameter values. With the 𝐷2 parameter set relatively high compared to 𝐾sk there is no
pore pressure until the volumetric strain is greater than the strains associated with the air
voids. However, as 𝐷2 is lowered, the pore pressure starts to increase before the air voids
are totally collapsed. The 𝐾sk parameter affects the slope of the post-void collapse pressure
- volumetric behavior.
The parameter 𝐷2 can be found from Skempton pore water pressure parameter 𝐵,
where 𝐵 is defined as [Holtz and Kovacs, 1981]:
1
𝐵= ,
𝐾
1 + 𝑛 sk
𝐾v (22.147.10)
1−𝐵
𝐷2 = .
𝐵𝐾sk [𝑛(1 − 𝑆)]
To simulate strain softening behavior the FHWA soil model uses a continuum
damage algorithm. The strain-based damage algorithm is based on the work of J. W. Ju
and J. C. Simo [1987, 1989]. They proposed a strain based damage criterion, which is
uncoupled from the plasticity algorithm.
Figure 22.147.3. The effects of 𝐷2 and 𝐾sk parameters on pore water pressure.
as:
𝑑𝑗+1 = 𝑑𝑗 if 𝜉𝑗+1 ≤ 𝑟𝑗
𝜉𝑗+1 − 𝜉0 , (22.147.13)
𝑑𝑗+1 = if 𝜉𝑗+1 > 𝑟𝑗
𝛼 − 𝜉0
where 𝑟t is a damage threshold surface, 𝑟𝑗+1 = max{𝑟𝑗 , 𝜉𝑗+1 ), and 𝜉0 = 𝑟0 (DINT). The mesh
sensitivity parameter, 𝛼, will be described below.
Typically, the damage, 𝑑, varies from 0 to a maximum of 1. However, some soils can
have a residual strength that is pressure dependent. The residual strength is represented
by 𝜙res , the minimum internal friction angle.
The maximum damage allowed is related to the internal friction angle of residual
strength by:
sin𝜙 − sin𝜙res
𝑑max = , (22.147.14)
sin𝜙
If 𝜙res > 0, then 𝑑max , the maximum damage, will not reach 1, and the soil will have some
residual strength.
When material models include strain softening, special techniques must be used to
prevent mesh sensitivity. Mesh sensitivity is the tendency of the finite element
model/analysis to produce significantly different results as the element size is reduced.
The mesh sensitivity occurs because the softening in the model concentrates in one
element. As the element size is reduced the failure becomes localized in smaller volumes,
which causes less energy to be dissipated by the softening leading to instabilities or at least
mesh sensitive behavior.
To eliminate or reduce the effects of strain softening mesh sensitivity, the softening
parameter, α (the strain at full damage), must be modified as the element size changes.
The FHWA soil model uses an input parameter, “void formation”, 𝐺f , that is like fracture
energy material property for metals. The void formation parameter is the area under the
softening region of the pressure volumetric strain curve times the cube root of the element
1
volume, 𝑉 3 .
1
1 𝛼 𝑃peak (𝛼 − 𝜉0 )𝑉 3 (22.147.15)
𝐺f = 𝑉 3 ∫ 𝑃 𝑑𝜀v = ,
𝜉0 2
with 𝜉0 , the volumetric strain at peak pressure (strain at initial damage, DINT). Then 𝛼 can
be found as a function of the volume of the element 𝑉:
2𝐺f
𝛼= 1⁄
+ 𝜉0 . (22.147.16)
𝐾𝜉0 𝑉 3
1⁄
If 𝐺f is made very small relative to 𝐾𝜉0 𝑉 3,
then the softening behavior will be
brittle.
Strain-rate enhanced strength is simulated by a two-parameter Devaut-Lions viscoplastic
update algorithm, developed by Murray [1997]. This algorithm interpolates between the
elastic trial stress (beyond the plasticity surface) and the inviscid stress. The inviscid
stresses (𝜎
̅̅̅̅) are on the plasticity surface.
𝜎
̅̅̅̅vp = (1 − 𝜍)𝜎
̅̅̅̅ + 𝜍𝜎
̅̅̅̅trial , (22.147.17)
𝜂 𝛾
where 𝜍 = Δ𝑡+𝜂, and 𝜂 = ( 𝜀̇r)(𝑣𝑛−1)/𝑣𝑛 .
As 𝜁 approaches 1, then the viscoplastic stress becomes the elastic trial stress.
Setting the input value 𝛾r = 0 eliminates any strain-rate enhanced strength effects.
The model allows element deletion, if needed. As the strain softening (damage)
increases, the effective stiffness of the element can get very small, causing severe element
distortion and hourglassing. The element can be “deleted” to remedy this behavior. There
are two input parameters that affect the point of element deletion. DAMLEV is the damage
threshold where element deletion will be considered. EPSPRMAX is the maximum
principal strain where element will be deleted. Therefore,
𝑑 ≥ DAMLEV and 𝜀prmax > EPSPRMAX, (22.147.18)
for element deletion to occur. If DAMLEV is set to zero, there is no element deletion. Care
must be taken when employing element deletion to assure that the internal forces are very
small (element stiffness is zero) or significant errors can be introduced into the analysis.
The keyword option, NEBRASKA, gives the soil parameters used to validate the
material model with experiments performed at University of Nebraska at Lincoln. The
units for this default inputs are milliseconds, kilograms, and millimeters. There are no
required input parameters except material id (MID). If different units are desired the unit
conversion factors that need to multiply the default parameters can be input.
𝛷 = 𝜎̂ − 𝜎Y , (22.154.1)
The equivalent stress, 𝜎̂ , is given by:
2
𝜎VM 2
+ 𝛼2 𝜎m
𝜎̂ 2 = , (22.154.2)
1 + (𝛼/3)2
where, 𝜎VM , is the von Mises effective stress,
3
𝜎VM = √ 𝛔dev : 𝛔dev , (22.154.3)
2
and, 𝜎m and 𝛔dev , is the mean and deviatoric stress
𝜎m = tr(𝛔)
(22.154.4)
𝛔dev = 𝛔 − σm 𝐈.
Rate effects are taken into account through a Maxwell model using linear
viscoelasticity by a convolution integral of the form:
𝑡 𝜕𝜀𝑘𝑙
𝜎𝑖𝑗 = ∫ 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) 𝑑𝜏 , (22.158.1)
0 𝜕𝜏
where 𝑔𝑖𝑗𝑘𝑙 (𝑡 − 𝜏) is the relaxation function for different stress measures. This stress is
added to the stress tensor determined from the strain energy functional. Since we wish to
include only simple rate effects, the relaxation function is represented by six terms from the
Prony series:
𝑁
𝑔(𝑡) = ∑ 𝐺𝑚 𝑒−𝛽𝑚 𝑡 . (22.158.2)
𝑚=1
We characterize this in the input by the shear moduli, 𝐺𝑖 , and the decay constants,
𝛽𝑖 . An arbitrary number of terms, not exceeding 6, may be used when applying the
viscoelastic model. The composite failure is not directly affected by the presence of the
viscous stress tensor.
Stress Invariants. The yield surface is formulated in terms of three stress invariants: 𝐽1 is
the first invariant of the stress tensor, 𝐽′2 is the second invariant of the deviatoric stress
tensor, and 𝐽′3 is the third invariant of the deviatoric stress tensor. The invariants are
defined in terms of the deviatoric stress tensor, 𝑆𝑖𝑗 and pressure, 𝑃, as follows:
𝐽1 = 3𝑃,
1
𝐽 ′ 2 = 𝑆𝑖𝑗 𝑆𝑖𝑗 , (22.159.1)
2
1
𝐽 ′ 3 = 𝑆𝑖𝑗 𝑆𝑗𝑘 𝑆𝑘𝑖 .
3
Plasticity Surface. The three invariant yield function is based on these three invariants,
and the cap hardening parameter, 𝑘, as follows:
𝑓 (𝐽1 , 𝐽′2 , 𝐽′3 , 𝜅) = 𝐽′2 − ℜ2 𝐹f2 𝐹c . (22.159.2)
Here 𝐹f is the shear failure surface, 𝐹c is the hardening cap, and is the Rubin three-
invariant reduction factor. The cap hardening parameter 𝜅 is the value of the pressure
invariant at the intersection of the cap and shear surfaces.
Trial elastic stress invariants are temporarily updated via the trial elastic stress
tensor, 𝜎 T These are denoted 𝐽1𝑇 , 𝐽2′𝑇 and 𝐽3′𝑇 . Elastic stress states are modeled when
Figure 22.159.1. General Shape of the concrete model yield surface in two-
dimensions.
𝑓 (𝐽1 , 𝐽′2 , 𝐽′3 , 𝜅 𝑇 ) < 0. Elastic-plastic stress states are modeled when 𝑓 (𝐽1 , 𝐽′2 , 𝐽′3 , 𝜅 𝑇 ) > 0. In
this case, the plasticity algorithm returns the stress state to the yield surface such that
𝑓 (𝐽1𝑃 , 𝐽′𝑃2 , 𝐽′𝑃3 , 𝜅 𝑃 ) = 0. This is accomplished by enforcing the plastic consistency condition
with associated flow.
Shear Failure Surface. The strength of concrete is modeled by the shear surface in the
tensile and low confining pressure regimes:
𝐹f (𝐽1 ) = 𝛼 − 𝜆𝑒−𝛽𝐽1 + 𝜃𝐽1 . (22.159.3)
Here the values of 𝛼, 𝛽, 𝜆, and 𝜃 are selected by fitting the model surface to strength
measurements from triaxial compression (txc) tests conducted on plain concrete cylinders.
Rubin Scaling Function. Concrete fails at lower values of √3𝐽2′ (principal stress difference)
for triaxial extension (txe) and torsion (tor) tests than it does for txc tests conducted at the
same pressure. The Rubin scaling function ℜ determines the strength of concrete for any
state of stress relative to the strength for txc, via ℜ𝐹𝑓 . Strength in torsion is modeled as
𝑄1 𝐹f . Strength in txe is modeled as 𝑄2 𝐹f , where:
𝑄1 = 𝛼1 − 𝜆1 𝑒−𝛽1𝐽1 + 𝜃1 𝐽1 ,
(22.159.4)
𝑄2 = 𝛼2 − 𝜆2 𝑒−𝛽2𝐽1 + 𝜃2 𝐽1 .
Cap Hardening Surface. The strength of concrete is modeled by a combination of the cap
and shear surfaces in the low to high confining pressure regimes. The cap is used to model
plastic volume change related to pore collapse (although the pores are not explicitly
modeled). The isotropic hardening cap is a two-part function that is either unity or an
ellipse:
[𝐽1 − 𝐿 (𝜅)][|𝐽1 − 𝐿(𝜅)| + 𝐽1 − 𝐿(𝜅)]
𝐹c ( 𝐽1 , 𝜅 ) = 1 − , (22.159.5)
2[𝑋(𝜅) − 𝐿 (𝜅)] 2
where 𝐿(𝜅) is defined as:
𝜅 if 𝜅 > 𝜅0
L(κ) = { . (22.159.6)
𝜅0 otherwise
The equation for 𝐹c is equal to unity for 𝐽1 𝐿(𝜅). It describes the ellipse for 𝐽1 >
𝐿(𝜅). The intersection of the shear surface and the cap is at 𝐽1 = 𝜅. 𝜅0 is the value of 𝐽1 at
the initial intersection of the cap and shear surfaces before hardening is engaged (before the
cap moves). The equation for 𝐿(𝜅) restrains the cap from retracting past its initial location
at 𝜅0 .
The intersection of the cap with the 𝐽1 axis is at 𝐽1 = 𝑋(𝜅). This intersection depends
upon the cap ellipticity ratio 𝑅, where 𝑅 is the ratio of its major to minor axes:
𝑋(𝜅) = 𝐿(𝜅) + 𝑅𝐹f (𝐿(𝜅)). (22.159.7)
The cap moves to simulate plastic volume change. The cap expands (𝑋(𝜅) and𝜅 increase)
to simulate plastic volume compaction. The cap contracts (𝑋(𝜅) and 𝜅 decrease) to
simulate plastic volume expansion, called dilation. The motion (expansion and
contraction) of the cap is based upon the hardening rule:
p 2
𝜀𝑣 = 𝑊(1 − 𝑒−𝐷1(𝑋−𝑋0)−𝐷2(𝑋−𝑋0) ). (22.159.8)
p
Here 𝜀v the plastic volume strain, 𝑊 is the maximum plastic volume strain, and 𝐷1 and 𝐷2
are model input parameters. 𝑋0 is the initial location of the cap when 𝜅 = 𝜅0 .
The five input parameters (𝑋0 , 𝑊, 𝐷1 , 𝐷2 , and 𝑅) are obtained from fits to the
pressure-volumetric strain curves in isotropic compression and uniaxial strain. 𝑋0
determines the pressure at which compaction initiates in isotropic compression. 𝑅
combined with 𝑋0 , determines the pressure at which compaction initiates in uniaxial strain.
𝐷1 and 𝐷2 determine the shape of the pressure-volumetric strain curves. 𝑊 determines the
maximum plastic volume compaction.
Damage. Concrete exhibits softening in the tensile and low to moderate compressive
regimes.
vp
𝜎𝑖𝑗𝑑 = (1 − 𝑑)𝜎𝑖𝑗 , (22.159.9)
A scalar damage parameter, 𝑑, transforms the viscoplastic stress tensor without damage,
denoted 𝜎 vp , into the stress tensor with damage, denoted 𝜎 d . Damage accumulation is
based upon two distinct formulations, which we call brittle damage and ductile damage.
The initial damage threshold is coincident with the shear plasticity surface, so the threshold
does not have to be specified by the user.
Ductile Damage. Ductile damage accumulates when the pressure (𝑃) is compressive and
an energy-type term, 𝜏c , exceeds the damage threshold, 𝜏0c . Ductile damage accumulation
depends upon the total strain components, 𝜀𝑖𝑗 , as follows:
1
𝜏c = √ 𝜎𝑖𝑗 𝜀𝑖𝑗 (22.159.10)
2
The stress components 𝜎𝑖𝑗 are the elasto-plastic stresses (with kinematic hardening)
calculated before application of damage and rate effects.
Brittle Damage. Brittle damage accumulates when the pressure is tensile and an energy-
type term, 𝜏t , exceeds the damage threshold, 𝜏0t . Brittle damage accumulation depends
upon the maximum principal strain, 𝜀max , as follows:
𝜏t = √𝐸𝜀2max . (22.159.11)
Ductile Damage:
𝑑max 1+𝐵
𝑑(𝜏1 ) = ( − 1). (22.159.13)
𝐵 1 + 𝐵exp[−𝐴(𝜏c − 𝜏0c )]
The damage parameter that is applied to the six stresses is equal to the current
maximum of the brittle or ductile damage parameter. The parameters 𝐴 and 𝐵 or 𝐶 and 𝐷
set the shape of the softening curve plotted as stress-displacement or stress-strain. The
parameter 𝑑max is the maximum damage level that can be attained. It is internally
calculated and is less than one at moderate confining pressures. The compressive softening
parameter, 𝐴, may also be reduced with confinement, using the input parameter PMOD, as
follows:
𝐴 = 𝐴(𝑑max + 0.001)PMOD . (22.159.14)
Regulating Mesh Size Sensitivity. The concrete model maintains constant fracture
energy, regardless of element size. The fracture energy is defined here as the area under
the stress-displacement curve from peak strength to zero strength. This is done by
internally formulating the softening parameters 𝐴 and 𝐶 in terms of the element length, 𝑙
(cube root of the element volume), the fracture energy, 𝐺f , the initial damage threshold,𝜏0t
or 𝜏0c , and the softening shape parameters, 𝐷 or 𝐵.
PWRT
⎛
⎜ −𝐽1 ⎞⎟
𝐺f = 𝐺fs + trans(𝐺ft − 𝐺fs ) where trans = ⎜
⎜ ⎟
⎟
′
⎝√3𝐽 2 ⎠
if the pressure is compressive
PWRC
⎛ 𝐽1 ⎞
𝐺f = 𝐺fs + trans(𝐺fc − 𝐺fs ) where trans = ⎜
⎜
⎜ ⎟
⎟
⎟
⎝ √3𝐽′ 2⎠
Viscoplastic Rate Effects. At each time step, the viscoplastic algorithm interpolates
p
between the elastic trial stress, 𝜎𝑖𝑗T , and the inviscid stress (without rate effects), 𝜎𝑖𝑗 , to set
vp
the viscoplastic stress (with rate effects), 𝜎𝑖𝑗 :
vp p
𝜎𝑖𝑗 = (1 − 𝛾)𝜎𝑖𝑗T + 𝛾𝜎𝑖𝑗 , (22.159.16)
Δ𝑡/𝜂
with 𝛾 = 1+Δ𝑡/𝜂.
This interpolation depends upon the effective fluidity coefficient, , and the time
step, 𝑡. The effective fluidity coefficient is internally calculated from five user-supplied
input parameters and interpolation equations:
if the pressure is tensile
pwrt
⎛ −𝐽1 ⎞
𝜂 = 𝜂s + trans(𝜂t − 𝜂s ) trans = ⎜
⎜
⎜ ⎟
⎟
⎟
√3𝐽 ′
⎝ 2⎠
This viscoplastic model may predict substantial rate effects at high strain rates (𝜀̇ >
100). To limit rate effects at high strain rates, the user may input overstress limits in tension
(OVERT) and compression (OVERC). These input parameters limit calculation of the
fluidity parameter, as follows:
where OVER = OVERT when the pressure is tensile, and OVER = OVERC when the
pressure is compressive.
The user has the option of increasing the fracture energy as a function of effective
strain rate via the REPOW input parameter, as follows:
REPOW
𝐸𝜀̇𝜂
𝐺rate
f = 𝐺f (1 + ) (22.159.19)
𝑓′
Here 𝐺rate
f is the fracture energy enhanced by rate effects, and 𝑓′ is the yield strength before
application of rate effects (which is calculated internally by the model). The term in
brackets is greater than, or equal to one, and is the approximate ratio of the dynamic to
static strength.
Matrix mode failures must occur without fiber failure, and hence they will be on
planes parallel to fibers. For simplicity, only two failure planes are considered: one is
perpendicular to the planes of layering and the other one is parallel to them. The matrix
failure criteria for the failure plane perpendicular and parallel to the layering planes,
respectively, have the forms:
Perpendicular matrix mode:
2
⟨𝜎𝑏 ⟩ 2 𝜏 𝜏 2
𝑓4 = ( ) + ( ′𝑏𝑐 ) + ( 𝑎𝑏 ) − 1 = 0. (22.161.4)
𝑆𝑏T 𝑆𝑏𝑐 𝑆𝑎𝑏
𝑆′𝑏𝑐 = 𝑆(0)
𝑏𝑐 + tan(𝜑)⟨−𝜎𝑏 ⟩,
(22.161.6)
𝑆𝑐𝑎 = 𝑆(0)
𝑐𝑎 + tan(𝜑)⟨−𝜎𝑐 ⟩,
𝑆"𝑏𝑐 = 𝑆(0)
𝑏𝑐 + tan(𝜑)⟨−𝜎𝑐 ⟩,
where 𝜑 is a material constant as tan(𝜑) is similar to the coefficient of friction, and 𝑆(0) (0)
𝑎𝑏 , 𝑆𝑐𝑎
and 𝑆(0)
𝑏𝑐 are the shear strength values of the corresponding tensile modes.
When a matrix failure (delamination) in the a-b plane is predicted, the strength
(0)
values for 𝑆(0)
𝑐𝑎 and 𝑆𝑏𝑐 are set to zero. This results in reducing the stress components 𝜎𝑐 ,𝜏𝑏𝑐
and 𝜏𝑐𝑎 to the fractured material strength surface. For tensile mode, 𝜎𝑐 > 0, these stress
components are reduced to zero. For compressive mode, 𝜎𝑐 < 0, the normal stress 𝜎𝑐 is
assumed to deform elastically for the closed matrix crack. Loading on the failure envelop,
the shear stresses are assumed to ‘slide’ on the fractured strength surface (frictional shear
stresses) like in an ideal plastic material, while the subsequent unloading shear stress-strain
path follows reduced shear moduli to the zero shear stress and strain state for both 𝜏𝑏𝑐 and
𝜏𝑐𝑎 components.
The post failure behavior for the matrix crack in the a-c plane due to 𝑓4 is modeled in
the same fashion as that in the a-b plane as described above. In this case, when failure
occurs, 𝑆(0) (0)
𝑎𝑏 and 𝑆𝑏𝑐 are reduced to zero instantaneously. The post fracture response is then
governed by failure criterion of f5 with 𝑆(0) (0)
𝑎𝑏 = 0 and 𝑆𝑏𝑐 = 0. For tensile mode, 𝜎𝑏 , , 𝜏𝑎𝑏 and
𝜏𝑏𝑐 are zero. For compressive mode, 𝜎𝑏 < 0, 𝜎𝑏 is assumed to be elastic, while 𝜏𝑎𝑏 and 𝜏𝑏𝑐
‘slide’ on the fracture strength surface as in an ideal plastic material, and the unloading
path follows reduced shear moduli to the zero shear stress and strain state. It should be
noted that 𝜏𝑏𝑐 is governed by both the failure functions and should lie within or on each of
these two strength surfaces.
2 2
⟨𝜎𝑏 ⟩ 2 (𝜏𝑎𝑏 + 𝜏𝑏𝑐 )
𝑓7 = ( ) + 2
− 1 = 0, (22.161.8)
𝑆𝑏T 𝑆𝑏FS
where 𝑆𝑎T and 𝑆𝑏T are the axial tensile strengths in the fill and warp directions,
respectively, and 𝑆𝑎FS and 𝑆𝑏FS are the layer shear strengths due to fiber shear failure in the
fill and warp directions. These failure criteria are applicable when the associated 𝜎𝑎 or 𝜎𝑏 is
positive. It is assumed 𝑆aFS = SFS, and
𝑆𝑏T
𝑆𝑏FS = SFS ∗ . (22.161.9)
𝑆𝑎T
2
⟨𝑝⟩ 𝜎𝑎 + 𝜎𝑏 + 𝜎𝑐
𝑓10 =( ) − 1 = 0, 𝑝=− . (22.161.12)
𝑆FC 3
A plain weave layer can fail under in-plane shear stress without the occurrence of
fiber breakage. This in-plane matrix failure mode is given by
𝜏𝑎𝑏 2
𝑓11 = ( ) − 1 = 0, (22.161.13)
𝑆𝑎𝑏
where 𝑆𝑎𝑏 is the layer shear strength due to matrix shear failure.
Another failure mode, which is due to the quadratic interaction between the
thickness stresses, is expected to be mainly a matrix failure. This through the thickness
matrix failure criterion is
When failure predicted by this criterion occurs within elements that are adjacent to
the ply interface, the failure plane is expected to be parallel to the layering planes, and,
thus, can be referred to as the delamination mode. Note that a scale factor 𝑆 is introduced
to provide better correlation of delamination area with experiments. The scale factor 𝑆 can
be determined by fitting the analytical prediction to experimental data for the delamination
area.
When the in-plane matrix shear failure is predicted by f11 the axial load carrying
capacity within a failed element is assumed unchanged, while the in-plane shear stress is
assumed to be reduced to zero.
For through the thickness matrix (delamination) failure given by equations 𝑓12 , the
in-plane load carrying capacity within the element is assumed to be elastic, while the
(0)
strength values for the tensile mode, 𝑆(0)
𝑐𝑎 and 𝑆𝑏𝑐 , are set to zero. For tensile mode, 𝜎𝑐 > 0,
the through the thickness stress components are reduced to zero. For compressive mode,
𝜎𝑐 < 0, 𝜎𝑐 is assumed to be elastic, while 𝜏𝑏𝑐 and 𝜏𝑐𝑎 ‘slide’ on the fracture strength surface
as in an ideal plastic material, and the unloading path follows reduced shear moduli to the
zero shear stress and strain state.
The effect of strain-rate on the layer strength values of the fiber failure modes is
modeled by the strain-rate dependent functions for the strength values {IRT } as
{ε̅}̇
{SRT } = {S0 } ( 1 + Crate1 ln ), (22.161.16)
ε̇0
⎧ ∣𝜀̇𝑎 ∣ ⎫
⎧ 𝑆𝑎T ⎫ { }
{ } { ∣𝜀̇𝑎 ∣ }
{ 𝑆𝑎C } { }
{ } { ∣𝜀̇ ∣ }
{ 𝑆𝑏T } { 𝑏 }
{𝑆RT } = ⎨ ⎬ and {𝜀̅
} ̇ = ⎨ ∣𝜀̇ ∣ ⎬, (22.161.17)
{ 𝑆𝑏C }
𝑏
{ ∣𝜀̇𝑐 ∣ }
{𝑆 } { }
{ FC } { }
{ } { 1}
⎩ 𝑆FS ⎭ { 2 2 }
⎩(𝜀̇𝑐𝑎 + 𝜀̇𝑏𝑐 )2 ⎭
where 𝐶rate is the strain-rate constants, and {𝑆0 }are the strength values of {𝑆RT } at the
reference strain-rate 𝜀̇0 .
Damage Model
The damage model is a generalization of the layer failure model of Material 161 by
adopting the MLT damage mechanics approach, Matzenmiller et al. [1995], for
characterizing the softening behavior after damage initiation. Complete model description
is given in Yen [2001]. The damage functions, which are expressed in terms of ply level
engineering strains, are converted from the above failure criteria of fiber and matrix failure
modes by neglecting the Poisson’s effect. Elastic moduli reduction is expressed in terms of
the associated damage parameters 𝜛𝑖 :
E′𝑖 = (1 − ϖ𝑖 )E𝑖 (22.161.18)
ϖ𝑖 = 1 − exp(−𝑟𝑖𝑚𝑖 /𝑚𝑖 ) 𝑟𝑖 ≥ 0 𝑖 = 1, . . . ,6, (22.161.19)
′
where 𝐸𝑖 are the initial elastic moduli, 𝐸𝑖 are the reduced elastic moduli, 𝑟𝑖 are the damage
thresholds computed from the associated damage functions for fiber damage, matrix
damage and delamination, and mi are material damage parameters, which are currently
assumed to be independent of strain-rate. The damage function is formulated to account
for the overall nonlinear elastic response of a lamina including the initial ‘hardening’ and
the subsequent softening beyond the ultimate strengths.
In the damage model (Material 162), the effect of strain-rate on the nonlinear stress-
strain response of a composite layer is modeled by the strain-rate dependent functions for
the elastic moduli {𝐸RT } as
{𝜀̅}̇
{𝐸RT } = {𝐸0 } ( 1 + {𝐶rate } ln ), (22.161.20)
𝜀̇0
⎧ 𝐸𝑎 ⎫ ⎧ ∣𝜀̇𝑎 ∣ ⎫ ⎧𝐶rate2 ⎫
{ { } { }
{ 𝐸𝑏 }} { ∣𝜀̇𝑏 ∣ }
{ } {𝐶rate2 }
{ }
{
{ 𝐸𝑐 }} { ∣𝜀̇𝑐 ∣ } {𝐶rate4 }
̇
{𝐸RT } = ⎨ ⎬ , {𝜀̅} = ⎨ and {𝐶rate } = ⎨ ,
{ 𝐺𝑎𝑏 } { ∣𝜀̇𝑎𝑏 ∣⎬
} {𝐶rate3 ⎬
} (22.161.21)
{
{ 𝐺 } { ∣𝜀̇ ∣} {𝐶 }
{ 𝑏𝑐 }} {
{ 𝑏𝑐 }
}
{ rate3 }
{ }
⎩𝐺𝑐𝑎 ⎭ ⎩ ∣𝜀̇𝑐𝑎 ∣⎭ ⎩𝐶rate3 ⎭
where {𝐶rate } are the strain-rate constants. {𝐸0 } are the modulus values of {𝐸RT } at the
reference strain-rate 𝜀̇0 .
To prevent high frequency oscillations in the strain rate from causing similar high
frequency oscillations in the yield stress, a modified volumetric strain rate is used when
interpolating to obtain the yield stress. The modified strain rate is obtained as follows. If
NCYCLE is > 1, then the modified strain rate is obtained by a time average of the actual
strain rate over NCYCLE solution cycles. For SRCLMT > 0, the modified strain rate is
capped so that during each cycle, the modified strain rate is not permitted to change more
than SRCLMT multiplied by the solution time step.
1-V
Figure 22.163.1. Rate effects are defined by a family of curves giving yield stress
versus volumetric strain.
Yield condition:
𝑓 = |𝑀| − 𝑀Y (𝜅̅p ) = 0. (22.166.4)
Loading and unloading conditions:
𝜆̇ ≥ 0, 𝑓 ≤ 0, 𝜆̇ 𝑓 = 0. (22.166.5)
Consistency condition:
∂𝑀Y p
𝑓 ̇ = 0 ⇒ 𝑀̇ sign(𝑀) − 𝜅̅ = 0
∂𝜅̅p
⇒ 𝜆̇ ≡ 𝜅̅ṗ
𝑀sign(𝑀) (𝐸𝐼)e p
= p = p (𝜅̇ − 𝜅̇ ) sign(𝑀)
(𝐸𝐼) (𝐸𝐼) (22.166.6)
e
(𝐸𝐼)
= [𝜅̇ − 𝜆̇ sign(𝑀)]sign(𝑀)
(𝐸𝐼)p
(𝐸𝐼)e 𝜅̇ sign(𝑀)
⇒ 𝜆̇ ≡ 𝜅̅̇ =
(𝐸𝐼)p + (𝐸𝐼)e
Moment rate is also the product of tangential bending stiffness and total curvature:
𝑀̇ = (𝐸𝐼)ep 𝜅̇. (22.166.7)
Elastic, plastic, and tangential stiffnesses are obtained from user-defined curves:
𝑑𝑀 𝑑𝑀
(𝐸𝐼)ep = , (𝐸𝐼)p = . (22.166.8)
𝑑𝜅 𝑑𝜅̅p
e (𝐸𝐼)ep (𝐸𝐼)p
(𝐸𝐼) = . (22.166.9)
(𝐸𝐼)p − (𝐸𝐼)ep
For Torsion-Twist, simply replace 𝑀 by 𝑇, 𝜅 by 𝛽, (𝐸𝐼) by (𝐺𝐽). For Force-Strain, simply
replace 𝑀 by 𝑁, 𝜅 by 𝜀, (𝐸𝐼) by (𝐸𝐴).
In-plane stresses are set to zero: it is assumed that the stiffness and strength of the
substrate is large compared with that of the adhesive, given the relative thicknesses. If the
substrate is modeled with shell elements, it is expected that these will lie at the mid-surface
of the substrate geometry. Therefore the solid elements representing the adhesive will be
thicker than the actual bond. The yield and failure surfaces are treated as a power-law
combination of direct tension and shear across the bond:(𝜎/𝜎max )PWRT + (𝜏/𝜏max )PWRS =
1.0 at yield. The stress-displacement curves for tension and shear are shown in the
diagrams below. In both cases, 𝐺c is the area under the curve. Because of the algorithm
used, yielding in tension across the bond does not require strains in the plane of the bond –
unlike the plasticity models, plastic flow is not treated as volume-conserving.
Stress Stress
SHRMAX dp = SHRP.dfs
TENMAX
Area = Gcten Area = Gcshr
Failure Failure
Displacement dp Displacement
dft dfs
Tension Shear
Figure 22.169.1.
If we wish to include only simple rate effects, the relaxation function is represented
by six terms from the Prony series:
𝑁
𝑔(𝑡) = ∑ 𝐺𝑚 𝑒−𝛽𝑚 𝑡 . (22.175.2)
𝑚=1
We characterize this in the input by shear moduli, 𝐺𝑖 , and the decay constants, 𝛽𝑖 .
An arbitrary number of terms, up to 6, may be used when applying the viscoelastic model.
For volumetric relaxation, the relaxation function is also represented by the Prony
series in terms of bulk moduli:
𝑁
𝑘(𝑡) = ∑ 𝐾𝑚 𝑒−𝛽𝑘𝑚 𝑡 . (22.175.3)
𝑚=1
The Arrhenius and Williams-Landau-Ferry (WLF) shift functions account for the
effects of the temperature on the stress relaxation. A scaled time, 𝑡′ ,
𝑡
𝑡′ = ∫ 𝛷(𝑇)𝑑𝑡, (22.175.4)
0
is used in the relaxation function instead of the physical time. The Arrhenius shift function
is
1 1
𝛷(𝑇) = exp (−𝐴 { − }), (22.175.5)
𝑇 𝑇REF
and the Williams-Landau-Ferry shift function is
𝑇 − TREF
𝛷(𝑇) = exp (−𝐴 ). (22.175.6)
𝐵 + 𝑇 − 𝑇REF
where G is the shear modulus. In place of the effective plastic strain in the D3PLOT
database, the effective strain is output:
2
𝜀effective = √ 𝜀𝑖𝑗 𝜀𝑖𝑗 . (22.176.2)
3
The polynomial for instantaneous elastic response should contain only odd terms if
symmetric tension-compression response is desired.
The 𝑞𝑖𝑗 are the components of the orthogonal tensor containing the eigenvectors of
the principal basis. The Cauchy stress is then given by
𝜎𝑖𝑗 = 𝐽 −1 𝜏𝑖𝑗 , (22.177.3)
where 𝐽 = 𝜆1 𝜆2 𝜆3 is the relative volume change.
The constitutive tensor that relates the rate of deformation to the Truesdell
(convected) rate of Kirchhoff stress in the principal basis can be expressed as
TKE 𝜕𝜏𝑖𝑖E ⎫
}
𝐶𝑖𝑖𝑗𝑗 = 𝜆𝑗 − 2𝜏𝑖𝑖E 𝛿𝑖𝑗 }
𝜕𝜆𝑗 }
}
2 E 2 E
𝜆𝑗 𝜏𝑖𝑖 − 𝜆𝑖 𝜏𝑗𝑗 }
TKE
}
𝐶𝑖𝑗𝑖𝑗 = 2 2
, 𝑖 ≠ 𝑗, 𝜆𝑖 ≠ 𝜆𝑗 ⎬ (no sum). (22.177.4)
𝜆𝑖 − 𝜆𝑗 }
}
E E }
TKE 𝜆𝑖 𝜕𝜏𝑖𝑖 𝜕𝜏𝑖𝑖 }
𝐶𝑖𝑗𝑖𝑗 = ( − ), 𝑖 ≠ 𝑗, 𝜆𝑖 = 𝜆𝑗 }
}
2 𝜕𝜆𝑖 𝜕𝜆𝑗 ⎭
and finally the constitutive tensor relating the rate of deformation to the Truesdell rate of
Cauchy stress is obtained through
TC TK
𝐶𝑖𝑗𝑘𝑙 = 𝐽 −1 𝐶𝑖𝑗𝑘𝑙 . (22.177.6)
for the viscous stress with u̇ being the nodal velocity. The Truesdell rate of the viscous
stress can be written,
𝛔𝛁𝐓 = 𝐂𝐃
̇ + 𝐂𝐃̇ + tr(𝐃)𝛔 − 𝐋𝛔 − 𝛔𝐋𝐓 , (22.177.15)
where Lis the velocity gradient. The terms on the right hand side can be treated as follows.
For the first term, we can assume that 𝑑𝑚 ∝ J1/3 and then approximate
2
𝐂̇ = − tr(𝐃)𝐂. (22.177.16)
3
Using Equation (22.177.10), Equation (22.177.13), the first term on the right hand side
of Equation (22.177.15), Equation (22.177.16) and the expression
𝐃 = 𝐁𝐮̇, (22.177.17)
a material tangent modulus contribution can be identified in Equation (22.177.14) as
2
− 𝛔𝛅T , (22.177.18)
3
where 𝛅 denotes the identity matrix in Voigt notation.
Post-poning the treatment of the first term, the second of these two terms can be
treated easily as this gives the following contribution to the material time derivative
𝛾
∫ 𝐁T 𝐂𝐁𝑑𝛺𝑚 𝑢̇ , (22.177.20)
𝛽Δ𝑡
𝛺𝑚
where γ and 𝛽 are parameters in the Newmark scheme and Δ𝑡 is the time step. From this
expression, a material tangent modulus can through Equation (22.177.14) be identified as,
𝛾
𝐂mat = 𝐂. (22.177.21)
𝛽Δ𝑡
The third term in Equation (22.177.15) contributes to the material tangent modulus
as
𝛔𝛅T (22.177.22)
resulting in a material tangent modulus given so far by
𝛾 1
𝐂 + 𝛔𝛅T . (22.177.23)
𝛽Δ𝑡 3
Since this stress contribution is viscous and proportional to the mesh size, it is our
belief that it serves as a stabilizing stress in the occurrence of a coarse mesh and/or large
deformation rates, and really has little or nothing to do with the actual material models. If
only the simulation process is slow (which it often is in an implicit analysis) and/or the
mesh is sufficiently fine, this stress should be negligible compared to the other stress(es).
With this in mind, we feel that it is not crucial to derive an exact tangent for this stress but
we can be satisfied with an approximation. Even if attempting a more thorough derivation
of the tangent stiffness, we would most certainly have to make approximations along the
way. Hence we do not see this as an attractive approach.
In the implementation we have simply neglected all terms involving stresses since
the experience from earlier work is that such terms generally have a negative effect on the
tangent if they are not absolutely correct. In addition, most of the terms involving stresses
contribute to a nonsymmetric tangent stiffness, which cannot be supported by LS-DYNA at
the moment. Hence the material tangent modulus for the viscous stress is given by
Equation (22.177.21). We are aware of that this may be a crude approximation, and if
experiments show that it is a poor one, we will take a closer look at it.
In material type 178, the viscous stress acts only in the direction of the principal
stretches and in compression. With C being an isotropic tensor, we evaluate the tangent
stiffness modulus in the principal basis according to Equation (22.177.21), modify it to
account for the mentioned conditions and then transform it back to the global frame of
reference.
12
𝜎𝑖𝑗∇ = ∑ 2𝐺𝑚 𝑠m∇
𝑖𝑗 , (22.177.24)
𝑚=1
where
𝑠m∇ m
𝑖𝑗 = 𝐷𝑖𝑗 − 𝛽𝑚 𝑠𝑖𝑗 , (22.177.25)
Here 𝐺𝑚 and 𝛽𝑚 are material constants, and 𝐷𝑖𝑗 is the rate-of-deformation tensor. Referring
to Borrvall [2002], we state that the tangent stiffness modulus for this stress contribution
can be written
12
𝐶𝑖𝑗𝑘𝑙 = ∑ 𝐺𝑚 (𝛿𝑖𝑘 𝛿𝑗𝑙 + 𝛿𝑖𝑙 𝛿𝑗𝑘 ). (22.177.26)
𝑚=1
Just as for the viscous stress, this stress acts only in the direction of the principal stretches.
Hence the tangent modulus is formed in the principal basis, modified to account for this
condition and then transformed back to the global frame of reference.
To implement this tangent, the last term is the most difficult to deal with as it
involves the time derivative (or variation) of the pressure. For certain types of material
models, for instance material type 77 in LS-DYNA, the pressure is a function of the relative
volume
𝑝 = 𝑝(𝐽), (22.177.29)
and with the approximation
𝑝̅ = 𝑝(𝐽 )̅ , (22.177.30)
the last term can be evaluated to
∫ 𝐽𝑝′(𝐽)𝐁T (𝐈 ⊗ 𝐈)𝐁𝑑𝛺𝑚 𝑢̇ − ∫ 𝐽 𝑝̅ ′ (𝐽 )̅ 𝐁
̅̅̅̅̅T (𝐈 ⊗ 𝐈)𝐁
̅̅̅̅̅𝑑𝛺𝑚 𝑢̇,
(22.177.31)
𝛺𝑚 𝛺𝑚
and a symmetric tangent stiffness can quite easily be implemented. We have here used 𝐽 ̅
and 𝐁̅̅̅̅̅ for the mean values of 𝐽 and 𝐁, respectively. For other types of material models,
such as the ones described in this document or material type 27 in LS-DYNA, the
expression for the pressure is more complicated. A characterizing feature is that a non-zero
pressure can occur under constant volume. This will in general complicate the
implementation of the last term and will also contribute to a non-symmetric tangent
stiffness that cannot be handled in LS-DYNA at the moment. For material 27, neglecting
this had a tremendous impact on the performance of the implicit solution procedure, (see
material type 27). For the current material models, it seems to be of less importance, and
we believe that this is due to the higher compressibility allowed.
The 𝑞𝑖𝑗 are the components of the orthogonal tensor containing the eigenvectors of
the principal basis. The Cauchy stress is then given by ,
𝜎𝑖𝑗 = 𝐽 −1 𝜏𝑖𝑗 , (22.179.3)
where 𝐽 = 𝜆1 𝜆2 𝜆3 is the relative volume change.
The constitutive tensor that relates the rate of deformation to the Truesdell
(convected) rate of Kirchhoff stress can in the principal basis be expressed as
TKE 𝜕τEii
𝐶𝑖𝑖𝑗𝑗 = 𝜆𝑗 − 2𝜏𝑖𝑖E 𝛿𝑖𝑗
𝜕𝜆𝑗
TKE
𝜆2𝑗 𝜏𝑖𝑖E − 𝜆2𝑖 𝜏𝑗𝑗E
𝐶𝑖𝑗𝑖𝑗 = , 𝑖 ≠ 𝑗, 𝜆𝑖 ≠ 𝜆𝑗 (no sum). (22.179.4)
𝜆2𝑖 − 𝜆2𝑗
TKE 𝜆𝑖 𝜕𝜏𝑖𝑖E 𝜕𝜏𝑖𝑖E
𝐶𝑖𝑗𝑖𝑗 = ( − ), 𝑖 ≠ 𝑗, 𝜆𝑖 = 𝜆𝑗
2 𝜕𝜆𝑖 𝜕𝜆𝑗
where
⎧𝑠 (𝜆 − 1 − 𝑠 ) if 𝜆 ≥
𝑠
+1
{
{ 2𝐸 𝐸
{
{
{𝐸 2 𝑠
𝑤(𝜆) = ⎨ 2 (𝜆 − 1) if 1 ≤ 𝜆 < + 1
𝐸 (22.179.8)
{
{ 𝜆
{
{
{ 𝑓s (1 − 𝜇)𝑑𝜇
∫ otherwise
⎩1
Here s is the nominal tensile cutoff stress and 𝐸 is the stiffness coefficient relating a change
in principal stretch to a corresponding change in nominal stress. The function 𝑓s (≤ 0) gives
the nominal compressive stress as a function of the strain in compression for the second
and all subsequent load cycles and is supplied by the user. To apply the formulas in the
previous section, we require
⎧𝑠𝜆 𝑠
{ 𝑖 if 𝜆𝑖 ≥ + 1
𝜕𝑤 { { 𝐸
E
𝜏𝑖𝑖 = 𝜆𝑖 = ⎨𝐸𝜆 (𝜆 − 1) if 1 ≤ 𝜆 < 𝑠 + 1 (22.179.9)
𝜕𝜆𝑖 { 𝑖 𝑖 𝑖
𝐸
{
{
⎩𝜆𝑖 𝑓𝑠 (1 − 𝜆𝑖 ) otherwise
⎧𝑠𝜆 𝑠
{ 𝑖 if 𝜆𝑖 ≥ + 1
𝜕𝜏 E {
{ 𝐸
𝜆𝑗 𝑖𝑖 = 𝛿𝑖𝑗 ⎨𝐸𝜆 (2𝜆 − 1) 𝑠 (22.179.10)
𝜕𝜆𝑗 { 𝑖 𝑖 if 1 ≤ 𝜆 𝑖 < +1
{ 𝐸
{
⎩𝜆𝑖 (𝑓𝑠 (1 − 𝜆𝑖 ) − 𝜆𝑖 𝑓 ′𝑠 (1 − 𝜆𝑖 )) otherwise
𝛾
𝐂mat = 𝐂, (22.179.16)
𝛽Δ𝑡
where 𝛾 and 𝛽 are parameters in the Newmark scheme and Δ𝑡 is the time step.
This stress acts only in the direction of the principal stretches. Hence the tangent
modulus is formed in the principal basis, modified to account for this condition and then
transformed back to the global frame of reference.
In the orthotropic case the principal Kirchhoff stress contribution is instead given by
1
𝜏𝑖𝑖E = 𝜆𝑖 {𝑔𝑠 (1 − 𝜆𝑖 ) − 𝑓𝑠 (1 − 𝜆𝑖 )}𝜉𝑖 . (22.179.23)
where the quantity 𝜀m in the orthotropic case is the maximum compressive principal strain
in any direction during the simulation thus far. As for the isotropic case, the material is
completely damaged after one load cycle and reloading will follow load curve 𝑓s . In
addition, the directions corresponding to no loading will remain unaffected.
The factors 𝜉 and 𝜉𝑖 are treated as constants in the determination of the tangent
stiffness so the contribution is regarded as hyperelastic and follows the exposition given in
Section 19.179.1.
The reason for not differentiating the coefficients 𝜅, 𝜉 and 𝜉𝑖 is that they are always
non-differentiable. Their changes depend on whether the material is loaded or unloaded,
i.e., the direction of the load. Even if they were differentiable their contributions would
occasionally result in a non-symmetric tangent stiffness matrix and any attempt to
symmetrize this would probably destroy its properties. After all, we believe that the one-
dimensional nature and simplicity of this foam will be enough for good convergence
properties even without differentiating these coefficients.
𝑃
𝑔(𝜆) = 𝑃/𝐴
𝐴 𝜆 = 𝜆1 = 𝑙/𝐿
𝜆2 = 𝜆3 = 𝑑/𝐷
𝑑
𝑙
𝐷
E
𝜏𝑖𝑗 = 𝑞𝑖𝑘 𝑞𝑗𝑙 𝜏𝑘𝑙 . (22.181.2)
The 𝑞𝑖𝑗 are the components of the orthogonal tensor containing the eigenvectors of
the principal basis. The Cauchy stress is then given by
𝜎𝑖𝑗 = 𝐽 −1 𝜏𝑖𝑗 , (22.181.3)
where 𝐽 = 𝜆1 𝜆2 𝜆3 is the relative volume change.
Now, the Ogden strain energy potential results in a Kirchhoff stress on the form
1 3
𝜏𝑖𝑖𝐸 = 𝑓 (𝜆̃ 𝑖 ) + 𝐾𝑚 (𝐽 − 1) − ∑ 𝑓 (𝜆̃ 𝑘 ) (22.4)
3 𝑘=1
for a (large) bulk modulus 𝐾𝑚 and where 𝜆̃ 𝑖 = 𝜆𝑖 /𝐽 1/3 are the isochoric stretches. In the
Ogden material, 𝑓 has a specific form a priori that requires a least square approximation for
fitting test data. This is of course a restriction and the idea in material 181 is to let 𝑓 be
determined directly from input data. The ansatz for the compressible foam option is to let
− 𝜈
𝜏𝑖𝑖𝐸 = 𝑓 (𝜆𝑖 ) − 𝑓 (𝐽 1−2𝜈 )
(22.5)
for a given Poisson’s ratio 𝜈, a decision that will be made clear below. So assume that 𝑔(𝜆)
is the curve providing the engineering stress as function of stretch in a uniaxial test, see
figure 20-122, then the principal Kirchhoff stresses are
𝐸
𝜏11 = 𝜆𝑔(𝜆)
𝐸 𝐸
𝜏22 = 𝜏33 =0
(22.6)
What follows is the determination of the internal function 𝑓 for the rubber and foam option.
which coincide with the isochoric counterparts. Using these expressions when equating
(22.4) and (22.6), one must determine 𝑓 from
2
𝜆𝑔(𝜆) = (𝑓 (𝜆) − 𝑓 (𝜆−1/2 )) + 𝐾𝑚 (𝐽 − 1)
3
1
0 = (𝑓 (𝜆−1/2 ) − 𝑓 (𝜆)) + 𝐾𝑚 (𝐽 − 1).
3
(22.8)
By subtracting these two equations to eliminate the influence of the pressure we get
−1
𝜆𝑔(𝜆) = 𝑓 (𝜆) − 𝑓 (𝜆 2 )
(22.9)
and by letting 𝑛 be large enough the function 𝑓 can be determined since the last term tends
to zero.
(22.13)
Note that the equation corresponding to the second of (22.6) vanishes because of the ansatz
in (22.5). The same technique as for the rubber option is used, recursive expansion gives
2 2 𝑛
𝑓 (𝜆) = 𝜆𝑔(𝜆) + 𝜆−𝜈 𝑔(𝜆−𝜈 ) + 𝜆𝜈 𝑔(𝜆𝜈 ) + ⋯ + 𝑓 (𝜆(−𝜈) )
(22.14)
where
σ𝑥𝑥
⎛
⎜ σ𝑦𝑦 ⎞⎟
⎜
⎜ ⎟
⎟
⎜ σ 𝑧𝑧 ⎟
𝛔=⎜
⎜
⎜ σ
⎟
⎟
⎟ ,
⎜ 𝑥𝑦 ⎟
⎜
⎜σ𝑦𝑧 ⎟ ⎟
σ
⎝ 𝑧𝑥 ⎠
𝐹11 𝐹12 𝐹12 0 0 0
⎛
⎜ 𝐹12 𝐹11 𝐹12 0 0 0 ⎞⎟
⎜
⎜ ⎟
⎟
⎜ 𝐹12 𝐹12 𝐹11 0 0 0 ⎟
𝐅=⎜
⎜
⎜
⎟
⎟, (22.187.2)
⎜
⎜ 0 0 0 𝐹44 0 0 ⎟⎟
⎟
⎜
⎜ 0 0 0 0 𝐹44 0 ⎟⎟
⎝ 0 0 0 0 0 𝐹44 ⎠
𝐹1 0 0 0 0 0
⎛
⎜ 0 𝐹1 0 0 0 0⎞⎟
⎜
⎜ ⎟
⎟
⎜ 0 0 𝐹1 0 0 0⎟
𝐁=⎜
⎜
⎜
⎟
⎟.
⎜
⎜ 0 0 0 0 0 0⎟⎟
⎟
⎜0 0 0 0 0 0⎟
⎝0 0 0 0 0 0⎠
Some restrictions apply to the choice of the coefficients. The existence of a stress-free state
and the equivalence of pure shear and biaxial tension/compression require respectively
𝐹0 ≤ 0 and 𝐹44 = 2(𝐹11 − 𝐹12 ). (22.187.3)
Although 4 independent coefficients remain in the expression for the isotropic yield surface
at this point, however the yield condition is not affected if all coefficients are multiplied by
a constant. Consequently only 3 coefficients can be freely chosen and 3 experiments under
different states of stress can be fitted by this formulation.
LS-DYNA DEV 10/27/16 (r:8004) 20-289 (Material Models)
Material Models LS-DYNA Theory Manual
Without loss of generality the expression for the yield surface can be reformulated in
terms of the first two stress invariants: pressure and von Mises stress:
σ𝑥𝑥 + σ𝑦𝑦 + σ𝑧𝑧
𝑝=− ,
3
. (22.187.4)
3 2
σvm = √ ((σ𝑥𝑥 + 𝑝) + (σ𝑦𝑦 + 𝑝) + (σ𝑧𝑧 + 𝑝) + 2σ𝑥𝑦 + 2σ𝑦𝑧 + 2σ𝑧𝑥 )
2 2 2 2 2
2
The expression for the yield surface then becomes
2
𝑓 = σvm − 𝐴0 − 𝐴1 𝑝 − 𝐴2 𝑝2 ≤ 0, (22.187.5)
and identification of the coefficients gives
𝐴0 = −𝐹0 , 𝐴1 = 3𝐹1 and 𝐴2 = 9(1 − 𝐹11 ), (22.187.6)
or equivalently
𝐴1 𝐴 𝐹 1 𝐴
𝐹0 = −𝐴0 , 𝐹1 = , 𝐹11 = 1 − 2 , 𝐹44 = 3 and 𝐹12 = 𝐹11 − 44 = − ( + 2 ). (22.187.7)
3 9 2 2 9
Since there is no loss of generality, the simpler formulation in invariants is adopted from
this point on. In principle the coefficients of the yield surface can now be determined from
3 experiments. Typically we would perform uniaxial tension, uniaxial compression and
simple shear tests:
Alternatively we can also compute the coefficients relating to the formulation in stress
space:
⎫ ⎧ 1 1
𝐹0 + 𝐹1 𝜎t + 𝐹11 σt2 = 0 }
} {
{ 𝐹1 = 𝐹0 ( − )
} { σc σt
}
} {
{ 𝐹0
𝐹0 − 𝐹1 σc + 𝐹11 σc2 = 0⎬ ⇒ ⎨𝐹11 = − . (22.187.9)
} { σt σc
}
} {
{ 𝐹0
𝐹0 + 𝐹44 σs2 = 0 }
} {{𝐹44 = − 2
⎭ ⎩ σs
Both are easily seen to be equivalent.
T ∂2 𝑓
𝑓 = 𝛔 𝐅𝛔 + 𝐁𝛔 + 𝐹0 → = 2𝐅 (22.187.11)
∂σ2
A sufficient condition for convexity in 6D stress space is then that the matrix F should be
positive semidefinite. This means all eigenvalues of F should be positive or zero. The
conditions for convexity will now be examined in physical terms for two cases: plane stress
and general 3D.
The 3D case
In the full 3D case, the convexity condition is generally more stringent. Again we
require the eigenvalues of F to be non-negative, where F is now the full 6 by 6 matrix:
𝐹11 + 2𝐹12 ≥ 0⎫ 2
}
𝐹11 − 𝐹12 ≥ 0 ⎬ ⇒ {3σs ≥ σt σc . (22.187.15)
𝐹44 ≥ 0 } −𝐹0 ≥ 0
⎭
Leading to
√σt σc √σt σc
σs ≥ > . (22.187.16)
√3 2
Alternatively a yield surface containing a linear rather than a quadratic term was
implemented in SAMP-1.
𝑓 = σvm − 𝐴0 − 𝐴1 𝑝 − 𝐴2 𝑝2 ≤ 0. (22.187.17)
As it will be difficult in general to guarantee a reasonable flow behaviour from three
independent measurements in shear, tension and compression, a simplified flow rule has
been implemented as the default in SAMP-1. The generally non-associated flow surface is
given as:
2
𝑔 = σvm + α𝑝2 . (22.187.18)
This flow rule is associated if:
𝐴1 = 0,
(22.187.19)
𝐴2 = −α (= cte).
And clearly leads to a constant value for the plastic Poisson ratio:
9 − 2α 9 1 − 2ν𝑝
ν𝑝 = ⇒ α= . (22.187.20)
18 + 2α 2 1 + ν𝑝
Plausible flow behaviour just means that:
9
0≤α≤ ⇒ 0 ≤ ν𝑝 ≤ 0.5. (22.187.21)
2
In SAMP-1 the value of the plastic Poisson coefficient is given by the user, either as a
constant or as a load curve in function of the uniaxial plastic strain. This allows adjusting
the flow rule of the material to measurements of transversal deformation during uniaxial
tensile or compressive testing. This can be important for plastics since often a non-
isochoric behaviour is measured.
Figure 22.2. Influence of the flow rule on the plastic Poisson ratio
The possible values for the plastic Poisson ratio and the resulting flow behaviour are
illustrated in Figure 22.2.
In SAMP-1 the formulation is slightly modified and based on a flow rule given as:
𝑔′ = √σvm
2 + α𝑝2 . (22.187.22)
A simple damage model was added to the SAMP-1 material law where the damage
parameter d is a function of plastic strain only. A load curve must be provided by the user
giving d as a function of the (true) plastic strain under uniaxial tension. The value of the
critical damage Dc leading to rupture is then the only other required additional input. The
implemented damage model is isotropic.
The implemented model then uses the notion of effective cross section, which is the
true cross section of the material minus the cracks that have developed. We will use the
following notation:
𝐴0 → undeformed cross section
We define the effective stress as the force divided by the effective cross section:
𝑓
σ= ,
𝐴
(22.25)
𝑓 𝑓 σ
σeff = = = ,
𝐴eff 𝐴(1 − 𝑑) 1 − 𝑑
which allows defining an effective yield stress:
σy
σy,eff = . (22.26)
1−𝑑
⎧ ̇ ⎫
𝐹 = 𝐹0 + 𝐾 𝑓 (Δ𝐿) ⎡ ⎛max {1. , ∣Δ𝐿∣ }⎟
⎢1 + 𝐶1 ⋅ Δ𝐿̇ + 𝐶2 ⋅ sgn(Δ𝐿̇)ln ⎜ ⎞⎤
⎥ + 𝐷Δ𝐿̇
⎨
{ 𝐷𝐿𝐸 ⎬
}
⎣ ⎝ ⎩ ⎭⎠⎦ (22.196.2)
+ 𝑔(Δ𝐿)ℎ(Δ𝐿̇).
If TYPE = 1, inelastic behavior is obtained. In this case, the yield force is taken
from the load curve:
𝐹Y = 𝐹y (Δ𝐿plastic ), (22.196.4)
The final force, which includes rate effects and damping, is given by:
⎛
⎧
{ ∣Δ𝐿̇∣ ⎫
}⎞⎤
𝐹𝑛+1 = 𝐹 ⋅ ⎡
⎢ 1 + 𝐶1 ⋅ Δ𝐿̇ + 𝐶2 ⋅ sgn(Δ𝐿̇ )ln ⎜ max ⎨ 1. , ⎬ ⎟⎥ + 𝐷Δ𝐿̇
⎣ ⎝ {
⎩ 𝐷𝐿𝐸 }
⎭⎠⎦ (22.196.7)
+ 𝑔(Δ𝐿)ℎ(Δ𝐿̇).
Unless the origin of the curve starts at (0,0), the negative part of the curve is used when
the spring force is negative where the negative of the plastic displacement is used to
interpolate, 𝐹y . The positive part of the curve is used whenever the force is positive.
The cross sectional area is defined on the section card for the discrete beam
elements, See *SECTION_BEAM. The square root of this area is used as the contact
thickness offset if these elements are included in the contact treatment.
23
Equation of State Models
LS-DYNA has 10 equation of state models which are described in this section.
1. Linear Polynomial
2. JWL High Explosive
3. Sack “Tuesday” High Explosive
4. Gruneisen
5. Ratio of Polynomials
6. Linear Polynomial With Energy Deposition
7. Ignition and Growth of Reaction in High Explosives
8. Tabulated Compaction
9. Tabulated
10. Propellant-Deflagration
The forms of the first five equations of state are given in the KOVEC user’s manual
[Woodruff 1973] as well as below.
The linear polynomial equation of state may be used to model gas with the gamma
law equation of state. This may be achieved by setting:
𝐶0 = 𝐶1 = 𝐶2 = 𝐶3 = 𝐶6 = 0, (23.1.4)
and
𝐶4 = 𝐶5 = 𝛾 − 1, (23.1.5)
where 𝛾 is the ratio of specific heats. The pressure is then given by:
𝜌
𝑝 = (𝛾 − 1) 𝐸. (23.1.6)
𝜌0
Note that the units of 𝐸 are the units of pressure.
This equation of state is used with the explosive burn (material model 8) material
model which determines the lighting time for the explosive element.
This equation of state is used with the explosive burn (material model 8) material
model which determines the lighting time for the explosive element.
𝜌
𝜇= . (23.5.14)
𝜌0 − 1
In expanded zoned 𝐹1 is replaced by 𝐹′1 = 𝐹1 + 𝛽𝜇2 Constants 𝐴𝑖𝑗 , 𝛼, and 𝛽 are user
input.
The mixture of unreacted explosive and reaction products is defined by the fraction
reacted 𝐹 (𝐹 = 0 implies no reaction, 𝐹 = 1 implies complete conversion from explosive to
products). The pressures and temperature are assumed to be in equilibrium, and the
relative volumes are assumed to be additive:
𝑉 = (1 − 𝐹)𝑉𝑒 + 𝐹𝑉𝑝 . (23.7.20)
The JWL equations of state and the reaction rates have been fitted to one- and two-
dimensional shock initiation and detonation data for four explosives: PBX-9404, RX-03-BB,
PETN, and cast TNT. The details of calculational method are described by Cochran and
Chan [1979]. The detailed one-dimensional calculations and parameters for the four
explosives are given by Lee and Tarver [1980]. Two-dimensional calculations with this
model for PBX 9404 and LX-17 are discussed by Tarver and Hallquist [1981].
C
B
Pressure P A
particle geometry, packing density, heat of reaction, and atmospheric pressure burn rate
data which allowed us to develop the numerical model presented here for their NaN3 +
Fe2 O3 driver airbag propellant. The deflagration model, its implementation, and the
results for the ICI propellant are presented in the are described by [Hallquist, et. al., 1990].
The unreacted propellant and the reaction product equations-of-state are both of the
form:
𝜔𝐶v 𝑇
𝑝 = 𝐴 𝑒−𝑅1𝑉 + 𝐵𝑒−𝑅2𝑉 + , (23.10.24)
𝑉−𝑑
where 𝑝 is pressure (in Mbars), 𝑉 is the relative specific volume (inverse of relative
density), 𝜔 is the Gruneisen coefficient, 𝐶v is heat capacity (in Mbars -cc/cc°K), 𝑇 is
temperature in °𝐾, 𝑑 is the co-volume, and 𝐴, 𝐵, 𝑅1 and 𝑅2 are constants. Setting𝐴 = 𝐵 = 0
yields the van der Waal’s co-volume equation-of-state. The JWL equation-of-state is
generally useful at pressures above several kilobars, while the van der Waal’s is useful at
pressures below that range and above the range for which the perfect gas law holds. Of
course, setting 𝐴 = 𝐵 = 𝑑 = 0 yields the perfect gas law. If accurate values of 𝜔 and 𝐶v
plus the correct distribution between “cold” compression and internal energies are used,
the calculated temperatures are very reasonable and thus can be used to check propellant
performance.
The reaction rate used for the propellant deflagration process is of the form:
∂𝐹
= 𝑍(1 − 𝐹)𝑦 𝐹𝑥 𝑝𝑤 + 𝑉(1 − 𝐹)𝑢 𝐹𝑟𝑝𝑠
∂𝑡
(23.10.25)
for 0 < 𝐹 < 𝐹limit1
for 𝐹limit2 < 𝐹 < 1
where 𝐹 is the fraction reacted (𝐹 = 0 implies no reaction, 𝐹 = 1 is complete reaction), 𝑡 is
time, and 𝑝 is pressure (in Mbars), 𝑟, 𝑠, 𝑢, 𝑤, 𝑥, 𝑦, 𝐹limit1 and 𝐹limit2 are constants used to
describe the pressure dependence and surface area dependence of the reaction rates. Two
(or more) pressure dependent reaction rates are included in case the propellant is a mixture
or exhibited a sharp change in reaction rate at some pressure or temperature. Burning
surface area dependences can be approximated using the (1 − 𝐹)𝑦 𝐹𝑥 terms. Other forms of
the reaction rate law, such as Arrhenius temperature dependent 𝑒−𝐸/𝑅𝑇 type rates, can be
used, but these require very accurate temperatures calculations. Although the theoretical
justification of pressure dependent burn rates at kilobar type pressures is not complete, a
vast amount of experimental burn rate versus pressure data does demonstrate this effect
and hydrodynamic calculations using pressure dependent burn accurately simulate such
experiments.
The deflagration reactive flow model is activated by any pressure or particle velocity
increase on one or more zone boundaries in the reactive material. Such an increase creates
pressure in those zones and the decomposition begins. If the pressure is relieved, the
reaction rate decreases and can go to zero. This feature is important for short duration,
To obtain good agreement with experimental deflagration data, the model requires
an accurate description of the unreacted propellant equation-of-state, either an analytical fit
to experimental compression data or an estimated fit based on previous experience with
similar materials. This is also true for the reaction products equation-of-state. The more
experimental burn rate, pressure production and energy delivery data available, the better
the form and constants in the reaction rate equation can be determined.
Therefore the equations used in the burn subroutine for the pressure in the
unreacted propellant
𝑅3 ⋅ 𝑇𝑢
𝑃𝑢 = 𝑅1 ⋅ 𝑒−𝑅5⋅𝑉𝑢 + 𝑅2 ⋅ 𝑒−𝑅6⋅𝑉𝑢 + , (23.10.26)
𝑉𝑢 − FRER
where 𝑉𝑢 and 𝑇𝑢 are the relative volume and temperature respectively of the unreacted
propellant. The relative density is obviously the inverse of the relative volume. The
pressure 𝑃p in the reaction products is given by:
−𝑋𝑃1⋅𝑉𝑝 −𝑋𝑃2⋅𝑉𝑝
𝐺 ⋅ 𝑇𝑝
𝑃p = 𝐴 ⋅ 𝑒 +𝐵⋅𝑒 + . (23.10.27)
𝑉𝑝 − CCRIT
As the reaction proceeds, the unreacted and product pressures and temperatures are
assumed to be equilibrated (𝑇𝑢 = 𝑇𝑝 = 𝑇, 𝑝 = 𝑃𝑢 = 𝑃𝑝 ) and the relative volumes are
additive:
𝑉 = (1 − 𝐹) ⋅ 𝑉𝑢 + 𝐹 ⋅ 𝑉𝑝 (23.10.28)
where 𝑉 is the total relative volume. Other mixture assumptions can and have been used
in different versions of DYNA2D/3D. The reaction rate law has the form:
∂𝐹
= GROW1(𝑝 + FREQ)𝑒𝑚 (𝐹 + FMXIG)𝑎𝑟1 (1 − 𝐹 + FMXIG)𝑒𝑠1
∂𝑡 (23.10.29)
+GROW2(𝑝 + FREQ)𝑒𝑛 (𝐹 + FMXIG)
If 𝐹 exceeds FMXGR, the GROW1 term is set equal to zero, and, if 𝐹 is less
thanFMNGR, the GROW2 term is zero. Thus, two separate (or overlapping) burn rates can
be used to describe the rate at which the propellant decomposes.
mixture states are used until the reaction is complete and then the reaction product
equation-of-state is used. The heat of reaction, ENQ, is assumed to be a constant and the
same at all values of 𝐹 but more complex energy release laws could be implemented.
24
Artificial Bulk Viscosity
Bulk viscosity is used to treat shock waves. Proposed in one spatial dimension by
von Neumann and Richtmyer [1950], the bulk viscosity method is now used in nearly all
wave propagation codes. A viscous term q is added to the pressure to smear the shock
discontinuities into rapidly varying but continuous transition regions. With this method
the solution is unperturbed away from a shock, the Hugoniot jump conditions remain valid
across the shock transition, and shocks are treated automatically. In our discussion of bulk
viscosity we draw heavily on works by Richtmyer and Morton [1967], Noh [1976], and
Wilkins [1980]. The following discussion of the bulk viscosity applies to solid elements
since strong shocks are not normally encountered in structures modeled with shell and
beam elements.
Consider a planar shock front moving through a material. Mass, momentum, and
energy are conserved across the front. Application of these conservation laws leads to the
well-known Rankine-Hugoniot jump conditions
𝜌(𝑢 − 𝑢0 )
𝑢𝑠 = , (24.1)
𝜌 − 𝜌0
𝜌 − 𝜌0 = 𝜌0 𝑢𝑠 (𝑢 − 𝑢0 ), (24.2)
𝑝 − 𝑝0 𝜌 − 𝜌0
𝑒 − 𝑒0 = , (24.3)
2 𝜌0 𝜌
σ
ε
x
Figure 24.1. If the sound speed increases as the stress increases the traveling
wave above will gradually steepen as it moves along the x-coordinate to form a
shock wave.
where Equation (24.3) is an expression of the energy jump condition using the results of
mass conservation, Equation (24.1), and momentum conservation, Equation (24.2). Here,𝑢𝑠
is the shock velocity, 𝑢 is the particle velocity, 𝜌 is the density, 𝑒 is the specific internal
energy, 𝑝 is the pressure, and the subscript, 0 , indicates the state ahead of the shock.
The energy equation relating the thermodynamic quantities density, pressure, and
energy must be satisfied for all shocks. The equation of state
𝑝 = 𝑝(𝜌, 𝑒), (24.4)
which defines all equilibrium states that can exist in a material and relating the same
quantities as the energy equation, must also be satisfied. We may use this equation to
eliminate energy from
Equation (24.3) and obtain a unique relationship between pressure and compression. This
relation, called the Hugoniot, determines all pressure-compression states achievable behind
the shock. Shocking takes place along the Rayleigh line and not the Hugoniot (Figure 24.1)
and because the Hugoniot curve closely approximates an isentrope, we may often assume
the unloading follows the Hugoniot. Combining Equations (24.1) and (24.2), we see that
the slope of the Rayleigh line is related to the shock speed:
1⁄
2
1 ⎡𝑝1 − 𝑝0 ⎤
𝑢𝑠 = ⎢ ⎥ . (24.5)
𝜌0 ⎢ 1 − 1 ⎥
⎣ 𝜌0 𝜌 ⎦
For the material of Figure 24.2, increasing pressure increases shock speed.
p1
p0
Figure 24.2. Shocking takes place along the Rayleigh line, and release closely
follows the Hugoniot. The cross-hatched area is the difference between the
internal energy behind the shock and the internal energy lost on release.
𝑝 = (𝛾 − 1)𝜌𝑒, (24.6)
where 𝛾 is the ratio of specific heats. Using the energy jump condition, we can eliminate e
and obtain the Hugoniot
𝑝 2𝑉 + (𝛾 − 1)(𝑉0 − 𝑉)
= 𝑝∗ = 0 , (24.7)
𝑝0 2𝑉 − (𝛾 − 1)(𝑉0 − 𝑉)
where 𝑉 is the relative volume. Figure 24.3 shows a plot of the Hugoniot and adiabat
where it is noted that for 𝑝∗ = 1, the slopes are equal. Thus for weak shocks, Hugoniot and
adiabat agree to the first order and can be ignored in numerical calculations. However,
special treatment is required for strong shocks, and in numerical calculations this special
treatment takes the form of bulk viscosity.
Limiting compression
Hugoniot
Adiabat
Figure 24.3. Hugoniot curve and adiabat for a g-law gas (from [Noh 1976]).
The viscosity proposed by von Neumann and Richtmyer [1950] in one spatial
dimension has the form
2
2 ∂𝑥̇ ∂𝑥̇
𝑞 = 𝐶0 𝜌(Δ𝑥) ( ) if <0
∂𝑥 ∂𝑥 (24.8)
∂𝑥̇
𝑞=0 if ≥0
∂𝑥
where 𝐶0 is a dimensionless constant and 𝑞 is added to the pressure in both the momentum
and energy equations. When 𝑞 is used, they proved the following for steady state shocks:
the hydrodynamic equations possess solutions without discontinuities; the shock thickness
is independent of shock strength and of the same order as the Δ𝑥 used in the calculations;
the q term is insignificant outside the shock layer; and the jump conditions are satisfied.
According to Noh, it is generally believed that these properties: “hold for all shocks, and
this has been borne out over the years by countless numerical experiments in which
excellent agreement has been obtained either with exact solutions or with hydrodynamical
experiments.”
In 1955, Landshoff [1955] suggested the addition of a linear term to the 𝑞 of von
Neumann and Richtmyer leading to a 𝑞 of the form
2
2 𝜕𝑥̇ 𝜕𝑥̇
𝑞 = 𝐶0 𝜌(Δ𝑥) ( ) if <0
𝜕𝑥 𝜕𝑥 (24.9)
2
2 ∂ẋ 𝜕𝑥̇
𝑞 = 𝐶0 𝜌(Δ𝑥) ( ) if ≥ 0,
∂x 𝜕𝑥
where 𝐶1 is a dimensionless constant and 𝑎 is the local sound speed. The linear term
rapidly damps numerical oscillations behind the shock front (Figure 24.3). A similar form
was proposed independently by Noh about the same time.
22-4 (Artificial Bulk Viscosity) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Artificial Bulk Viscosity
In converging two- and three-dimensional geometries, the strain rate 𝜀̇𝑘𝑘 is negative
and the 𝑞 term in Equation (24.15) is nonzero, even though no shocks may be generated.
This results in nonphysical 𝑞 heating. When the aspect ratios of the elements are poor (far
from unity), the use of a characteristic length based on √𝐴 or √ 3
𝑣 can also result in
nonphysical 𝑞 heating and even occasional numerical instabilities. Wilkins uses a bulk
viscosity that is based in part on earlier work by Richards [1965] that extends the von
LS-DYNA DEV 10/27/16 (r:8004) 22-5 (Artificial Bulk Viscosity)
Artificial Bulk Viscosity LS-DYNA Theory Manual
Neumann and Richtmyer formulations in a way that avoids these problems. This latter 𝑞
may be added in the future if the need arises.
Wilkins’ 𝑞 is defined as:
2
𝑑𝑠 𝑑𝑠
𝑞 = 𝐶0 𝜌𝑙 ( ) − 𝐶𝑙 𝜌𝑙𝑎∗
2
if 𝜀̇𝑘𝑘 < 0
𝑑𝑡 𝑑𝑡 (24.16)
𝑑𝑠
𝑞=0 if ≥0
𝑑𝑡
where 𝑙 and 𝑑𝑠
𝑑𝑡
are the thickness of the element and the strain rate in the direction of the
acceleration, respectively, and 𝑎∗ is the sound speed defined by (p/𝜌)1/2 if 𝑝 > 0. We use
the local sound speed in place of 𝑎∗ to reduce the noise at the low stress levels that are
typical of our applications.
Two disadvantages are associated with Equation (24.16). To compute the length
parameter and the strain rate, we need to know the direction of the acceleration through
the element. Since the nodal force vector becomes the acceleration vector in the explicit
integration scheme, we have to provide extra storage to save the direction. In three
dimensions our present storage capacity is marginal at best and sacrificing this storage for
storing the direction would make it even more so. Secondly, we need to compute l and 𝑑𝑠 𝑑𝑡
which results in a noticeable increase in computer cost even in two dimensions. For most
problems the additional refinement of Wilkins is not needed. However, users must be
aware of the pitfalls of Equation (24.15), i.e., when the element aspect ratios are poor or the
deformations are large, an anomalous 𝑞 may be generated.
25
Time Step Control
During the solution we loop through the elements to update the stresses and the
right hand side force vector. We also determine a new time step size by taking the
minimum value over all elements.
Δ𝑡𝑛+1 = 𝑎 ⋅ min{Δ𝑡1 , Δ𝑡2 , Δ𝑡3 , . . . , Δ𝑡𝑁 }, (25.1)
where 𝑁 is the number of elements. For stability reasons the scale factor 𝑎 is typically set to
a value of .90 (default) or some smaller value.
22.1
Time Step Control LS-DYNA Theory Manual
∂𝑝 ∂𝑝 ∂𝑝 ∂𝑈
) = ) + ) ), (25.5)
∂𝜌 𝑠 ∂𝜌 𝑈 ∂𝑈 𝜌 ∂𝜌 𝑠
and that along an isentrope the incremental energy, 𝑈, in the units of pressure is the
product of pressure, 𝑝, and the incremental relative volume, 𝑑𝑉:
𝑑𝑈 = −𝑝𝑑𝑉, (25.6)
we obtain
1⁄
⎡ 4𝐺 ∂𝑝 𝑝𝑉 2 ∂𝑝 ⎤ 2
𝑐=⎢ + ) + ) ⎥ . (25.7)
⎣3𝜌0 ∂𝜌 𝑈 𝜌0 ∂𝑈 𝜌 ⎦
For elastic materials with a constant bulk modulus the sound speed is given by:
𝐸(1 − 𝜐)
𝑐=√ (25.8)
(1 + 𝜐)(1 − 2𝜐)𝜌
where 𝐸 is Young’s modulus, and 𝜐 is Poisson’s ratio.
For the Belytschko beam the time step size given by the longitudinal sound speed is
used (Equation (25.9)), unless the bending-related time step size given by [Belytschko and
Tsay 1982]
0.5𝐿
Δ𝑡𝑒 =
3 (25.11)
𝑐√3𝐼 [ + 1 ]
12𝐼 + 𝐴𝐿2 𝐴𝐿2
is smaller, where 𝐼 and 𝐴 are the maximum value of the moment of inertia and area of the
cross section, respectively.
Comparison of critical time steps of the truss versus the elastic solid element shows
that it if Poisson's ratio, 𝜐 is nonzero the solid elements give a considerably smaller stable
time step size. If we define the ratio, 𝛼, as:
22.2
LS-DYNA Theory Manual Time Step Control
Table 22.1. Comparison of critical time step sizes for a truss versus a solid element.
𝐸
𝑐=√ . (25.14)
𝜌(1 − 𝜈 2 )
Three user options exists for choosing the characteristic length. In the default or first
option the characteristic length is given by:
(1 + 𝛽)𝐴𝑠
𝐿𝑠 = (25.15)
max(𝐿1 , 𝐿2 , 𝐿3 , (1 − 𝛽)𝐿4 )
where 𝛽 = 0 for quadrilateral and 1 for triangular shell elements, 𝐴𝑠 is the area, and𝐿𝑖 , (𝑖 =
1. . . .4) is the length of the sides defining the shell elements. In the second option a more
conservative value of 𝐿𝑠 is used:
(1 + 𝛽)𝐴𝑠
𝐿𝑠 = , (25.16)
max(𝐷1 , 𝐷2 )
where 𝐷𝑖 (𝑖 = 1,2) is the length of the diagonals. The third option provides the largest time
step size and is frequently used when triangular shell elements have very short altitudes.
The bar wave speed, Equation (21.10), is used to compute the time step size and 𝐿𝑠 is given
by
(1 + 𝛽)𝐴𝑠
𝐿𝑠 = max [ , min(𝐿1 , 𝐿2 , 𝐿3 , 𝐿4 + 𝛽1020 )]. (25.17)
max(𝐿1 , 𝐿2 , 𝐿3 , (1 − 𝛽)𝐿4 )
A comparison of critical time steps of truss versus shells is given in Table 22.2 with𝛽
defined as:
22.3
Time Step Control LS-DYNA Theory Manual
Δ𝑡2D-continuum 𝐶rod
𝛽= = = √1 − 𝜐 2 . (25.18)
Δ𝑡rod 𝐶
Table 22.2. Comparison of critical time step sizes for a truss versus a shell element.
The eigenvalue problem for the free vibration of spring with nodal masses 𝑚1 and
𝑚2 , and stiffness, 𝑘, is given by:
[𝑘 −𝑘 ] [𝑢1 ] − 𝜔2 [𝑚1 0 𝑢
] [𝑢1 ] = [0]. (25.20)
−𝑘 𝑘 𝑢2 0 𝑚2 2 0
Since the determinant of the characteristic equation must equal zero, we can solve for the
maximum eigenvalue:
𝑘 − 𝜔 2 𝑚1 −𝑘 2 𝑘(𝑚1 + 𝑚2 )
det [ ]= 0 → 𝜔max = , (25.21)
−𝑘 𝑘 − 𝜔 2 𝑚2 𝑚1 ⋅ 𝑚2
22.4
LS-DYNA Theory Manual Time Step Control
Therefore, in terms of the nodal mass we can write the critical time step size as:
2𝑀1 𝑀2
Δ𝑡𝑒 = 2√ . (25.24)
𝑘(𝑀1 + 𝑀2 )
The springs used in the contact interface are not checked for stability.
22.5
LS-DYNA Theory Manual Boundary and Loading Conditions
26
Boundary and Loading Conditions
∫ 𝑁 𝑡 𝑡𝑑𝑠, (26.1)
∂𝑏1
a Gaussian quadrature rule is used. To locate any point of the surface under consideration,
a position vector, 𝑟, is defined:
𝑟 = 𝑓1 (𝜉 , 𝜂)𝑖1 + 𝑓2 ((𝜉 , 𝜂)𝑖2 + 𝑓3 (𝜉 , 𝜂)𝑖3 , (26.2)
where
4
𝑗
𝑓𝑖 (𝜉 , 𝜂) = ∑ 𝜙𝑗 𝑥𝑖 , (26.3)
𝑗=1
Nodal quantities are interpolated over the four-node linear surface by the functions
1
𝜙𝑖 = (1 + 𝜉 𝜉𝑖 )(1 + 𝜂𝜂𝑖 ), (26.4)
4
so that the differential surface area 𝑑𝑠 may be written in terms of the curvilinear
coordinates as
𝑑𝑠 = |𝐽|𝑑𝜉𝑑𝜂, (26.5)
where |𝐽| is the surface Jacobian defined by
∂𝑟 ∂𝑟 1⁄
|𝐽| = ∣ × ∣ = (𝐸𝐺 − 𝐹2 ) 2 , (26.6)
∂𝜉 ∂𝜂
in which
3
2
x3
ξ
x2
4
i3 r(ξ,η) 1
i2
i1 x1
∂𝑟 ∂𝑟
𝐸= ⋅ ,
∂𝜉 ∂𝜉
∂𝑟 ∂𝑟
𝐹= ⋅ , (26.7)
∂𝜉 ∂𝜂
∂𝑟 ∂𝑟
𝐺= ⋅ .
∂𝜂 ∂𝜂
One such integral is computed for each surface segment on which a pressure loading acts.
Note that the Jacobians cancel when Equations (26.8) and (26.7) are put into Equation
(26.10). Equation (26.10) is evaluated with one-point integration analogous to that
employed in the volume integrals. The area of an element side is approximated by 4|𝐽|
where |𝐽| = |𝐽(0, 0)|.
𝐮𝑙
𝐱𝑙 = (26.13)
‖𝐮𝑙 ‖
𝐱𝑙 × 𝐯𝑙
𝐳𝑙 = (26.14)
‖𝐱𝑙 × 𝐯𝑙 ‖
𝐲𝑙 = 𝐳𝑙 × 𝐱𝑙 . (26.15)
𝛚̇ 𝐼𝑙 = 𝐪𝛚̇ 𝐼 , (26.18)
and the constrained components are zeroed. The modified vectors are then transformed
back to the global system:
𝐚𝐼 = 𝐪T 𝐚𝐼1 (26.19)
𝛚̇ 𝐼 = 𝐪T 𝛚̇ 𝐼𝑙 (26.20)
where 𝐚base is the base acceleration and 𝐦𝑒 is the element (lumped) mass matrix.
27
Time Integration
27.1 Background
Consider the single degree of freedom damped system in Figure 27.1.
p(t)
u(t) - displacements
Figure 27.1. Single degree of freedom damped system.
fI inertia force
elastic force
fs
m p(t) external forces
fD
damping forces
Figure 27.2. Forces acting on mass, m
𝑑2 𝑢
𝑓𝐼 = 𝑚𝑢̈; 𝑢̈ = acceleration
𝑑𝑡2
𝑑𝑢 (27.2)
𝑓𝐷 = 𝑐𝑢̇; 𝑢̇ = velocity
𝑑𝑡
𝑓int = 𝑘 ⋅ 𝑢; 𝑢 displacement
where 𝑐 is the damping coefficient, and k is the linear stiffness. For critical damping 𝑐 = ccr .
The equations of motion for linear behavior lead to a linear ordinary differential equation,
o.d.e.:
𝑚𝑢̈ + 𝑐𝑢̇ + 𝑘𝑢 = 𝑝(𝑡) (27.3)
but for the nonlinear case the internal force varies as a nonlinear function of the
displacement, leading to a nonlinear o.d.e.:
𝑚𝑢̈ + 𝑐𝑢̇ + 𝑓int (𝑢) = 𝑝(𝑡) (27.4)
𝑢0 = initial displacement
𝑢̇0 = initial velocity
𝑝0
= static displacement
𝑘
For nonlinear problems, only numerical solutions are possible. LS-DYNA uses the
explicit central difference scheme to integrate the equations of motion.
Substituting 𝑥̇𝑛 and 𝑥̈𝑛 into equation of motion at time 𝑡𝑛 leads to:
2 − 𝜔2 Δ𝑡2 1 − 2𝜉𝜔Δ𝑡 Δ𝑡2
𝑥𝑛+1 = 𝑥 − 𝑥 + 𝑌 , (27.17)
1 + 2𝜉𝜔Δ𝑡2 𝑛 1 + 2𝜉𝜔Δ𝑡 𝑛−1 1 + 2𝜉𝜔Δ𝑡2 𝑛
𝑥𝑛 = 𝑥𝑛 , (27.18)
which in matrix form leads to
Thus, damping reduces the critical time step size. The time step size is bounded by
the largest natural frequency of the structure which, in turn, is bounded by the highest
frequency of any individual element in the finite element mesh.
Figure 27.3. The right hand mesh is much more expensive to compute than the
left hand due to the presence of the thinner elements.
Hulbert and Hughes [1988] reviewed seven subcycling algorithms in which the
partitioning as either node or element based. The major differences within the two subcy-
cling methods lie in how elements along the interface between large and small elements are
handled, a subject which is beyond the scope of this theoretical manual. Nevertheless, they
concluded that three of the methods including the linear nodal interpolation method
chosen for LS-DYNA, provide both stable and accurate solutions for the problems they
studied. However, there was some concern about the lack of stability and accuracy proofs
for any of these methods when applied to problems in structural mechanics.
• Solid elements
• Beam elements
• Shell elements
• Brick shell elements
• Penalty based contact algorithms.
but intentionally excludes discrete elements since these elements generally contribute
insignificantly to the calculational costs. The interface stiffnesses used in the contact
algorithms are based on the minimum value of the slave node or master segment stiffness
and, consequently, the time step size determination for elements on either side of the
interface is assumed to be decoupled; thus, scaling penalty values to larger values when
subcycling is active can be a dangerous exercise indeed. Nodes that are included in
constraint equations, rigid bodies, or in contact with rigid walls are always assigned the
smallest time step sizes.
E2 = 4E1 F(t)
Material Group 1
A2 = A1
Material Group 2
ρ2 = ρ1
Figure 27.4. Subcycled beam problem from Hulbert and Hughes [1988].
minor cycle update. This linear variation of the major cycle nodal displacements during
the update of the element stresses improves accuracy and stability.
In the timing study results in Table 24.1, fifty solid elements were included in each
group for the beam depicted in Figure 27.4, and the ratio between 𝐸2 to 𝐸1 was varied from
1 to 128 giving a major time step size greater than 10 times the minor. Note that as the ratio
between the major and minor time step sizes approaches infinity the reduction in cost
should approach 50 percent for the subcycled case, and we see that this is indeed the case.
The effect of subcycling for the more expensive fully integrated elements is greater as may
be expected. The overhead of subcycling for the case where 𝐸1 = 𝐸2 is relatively large.
This provides some insight into why a decrease in speed is often observed when subcycling
is active. For subcycling to have a significant effect, the ratio of the major to minor time
step size should be large and the number of elements having the minor step size should be
small. In crashworthiness applications the typical mesh is very well planned and
generated to have uniform time step sizes; consequently, subcycling will probably give a
net increase in costs.
t
u2
v2
u1
v1
σ σ σ σ σ σ σ σ σ
t
u2
v2 σ σ σ σ σ σ σ
u1
v1
σ σ
u2 u2
v2
σ σ σ σ σ σ σ
v2 u1
v1
σ σ
t
σ σ σ σ σ σ σ σ σ u2
u2
v2
v2 u1
v1
𝐸2 = 𝐸1 180 22.09
180 22.75 (+3.0%)
𝐸2 = 4𝐸1 369 42.91
369 34.20 (-20.%)
𝐸2 = 16𝐸1 718 81.49
719 54.75 (-33.%)
𝐸2 = 64𝐸1 1424 159.2
1424 97.04 (-39.%)
𝐸2 = 128𝐸1 2034 226.8
2028 135.5 (-40.%)
Table 24.1. Timing study showing effects of the ratio of the major to minor time step
size.
The impact of the subcycling implementation in the software has a very significant
effect on the internal structure. The elements in LS-DYNA are now sorted three times
update velocities
update displacements
write databases
and new geometry
Sorting by Δ𝑡, interact with the second and third sorts and can result in the creation
of much smaller vector blocks and result in higher cost per element time step. During the
simulation elements can continuously change in time step size and resorting may be
required to maintain stability; consequently, we must check for this continuously. Sorting
cost, though not high when spread over the entire calculation, can become a factor that
results in higher overall cost if done too frequently especially if the factor, m, is relatively
small and the ratio of small to large elements is large.
28
Rigid Body Dynamics
𝑱𝜌 𝝎̇ + 𝝎 × 𝑱𝜌 𝝎 = 𝒇𝜌𝜔 , (28.2)
where 𝑀𝜌 is the physical mass, 𝑱𝜌 is the physical inertia tensor, 𝒙 is the location of the center
of mass, 𝝎 is the angular velocity of the body, and 𝒇𝜌𝑥 and 𝒇𝜌𝜔 are the forces and torques
applied to the rigid body through *LOAD_RIGID_BODY. These are equations that can be
found in a standard text book on rigid body mechanics. The physical properties of a rigid
body may come from three sources, these are
1.Integration of the mass density 𝜌 over a region 𝑉 occupied by the rigid body, for
which 𝑀𝜌 = ∫ 𝜌𝑑𝑉, and 𝑱𝜌 = ∫ 𝜌(𝒚 − 𝒙)⨂(𝒚 − 𝒙)𝑑𝑉. The initial rigid body coordi-
∫ 𝜌𝒚𝑑𝑉
nate 𝒙 is in this case determined from 𝒙 = 𝑀𝜌
. Here 𝒚 is the integrand variable.
2.Specifying properties using *PART_INERTIA, for which 𝑀𝜌 , 𝑱𝜌 and initial coordinate
𝒙 is simply specified in the keyword input deck.
3.For a nodal rigid body, *CONSTRAINED_NODAL_RIGID_BODY, the physical
properties vanish, i.e., 𝑀𝜌 = 0 and 𝑱𝜌 = 𝟎, and the position is arbitrary.
All rigid bodies possess slave nodes, which play a role when rigid bodies in LS-DYNA
interact with their surroundings. Slave nodes may come from the following.
1.The nodes in the finite element mesh for the part specified as rigid through
*MAT_RIGID.
2.Extra nodes definitions through *CONSTRAINED_EXTRA_NODES.
3.The set used for a nodal rigid body in *CONSTRAINED_NODAL_RIGID_BODY.
We use 𝑆 to denote the set of slave nodes to the rigid body, and these are constrained to the
rigid body through the equations
𝒙𝑖 = 𝒙 + 𝑸(𝒙𝑖0 − 𝒙 0 ), (28.3)
𝑸𝑖 = 𝑸, (28.4)
for all 𝑖 ∈ 𝑆. Here we have introduced the orientations 𝑸 and 𝑸𝑖 of the rigid body and
slave node 𝑖, respectively. Furthermore, 𝒙𝑖 is the coordinate of slave node 𝑖 and we use
superscript 0 to denote a quantity at time zero. The time evolution of 𝑸 is
𝑸̇ = 𝜴𝑸, (28.5)
where 𝛀𝒓 = 𝝎 × 𝒓 for an arbitrary vector 𝒓 and 𝑸 = 𝑰 (identity) at time zero. So equations
(28.3) and (28.4) can equivalently be put in rate form
𝒙̇𝑖 = 𝒙̇ + 𝝎 × 𝒓𝑖 , (28.6)
𝝎𝑖 = 𝝎, (28.7)
where 𝒓𝑖 = 𝒙𝑖 − 𝒙 and understandably 𝝎𝑖 rotational velocity of slave node 𝑖. This also
determines the space of admissible virtual displacement for the slave nodes in the context
of work principles, and for this reason we use a compact notation for this equation
𝒙̇ 𝑰 −𝑹𝑖 𝒙̇
[ 𝑖] = [ ] [ ]. (28.8)
𝝎𝑖 𝟎 𝑰 𝝎
where 𝑹𝑖 𝒓 = 𝒓𝑖 × 𝒓 for an arbitrary vector 𝒓.
Slave nodes may have masses 𝑚𝑖 , inertias 𝑱𝑖 and forces 𝒇𝑖𝑥 and 𝒇𝑖𝜔 associated with
them. The inertia properties may come from
1.Mass contributions from deformable elements connected to the rigid body, either
through *CONSTRAINED_EXTRA_NODES or
*CONSTRAINED_NODAL_RIGID_BODY or simply merged mesh.
2.Lumped masses through *ELEMENT_MASS or *ELEMENT_INERTIA.
Note that these quantities exclude any contributions from and on the rigid body itself, as
these are all collected in 𝑀𝜌 , 𝑱𝜌 , 𝒇𝜌𝑥 and 𝒇𝜌𝜔 . The motion of the slave nodes is governed by
their own equations of motion
𝑚𝑖 𝒙̈𝑖 = 𝒇𝑖𝑥 , (28.9)
𝑱𝑖 𝝎̇ 𝑖 + 𝝎𝑖 × 𝑱𝑖 𝝎𝑖 = 𝒇𝑖𝜔 , (28.10)
for 𝑖 ∈ 𝑆. We seek a set of equations for the rigid body that combines (28.1)-(28.2) and
(28.9)-(28.10) by condensing out the dependence of the slave nodes through (28.6)-(28.7).
Differentiating (28.6)-(28.7) with respect to time yields
𝒙̈𝑖 = 𝒙̈ + 𝝎̇ × 𝒓𝑖 + 𝝎 × 𝝎 × 𝒓𝑖 , (28.11)
𝝎̇ 𝑖 = 𝝎̇ , (28.12)
which can be inserted into (28.9)-(28.10) to yield
𝑚𝑖 [𝒙̈ + 𝝎̇ × 𝒓𝑖 ] = 𝒇𝑖𝑥 − 𝑚𝑖 𝝎 × 𝝎 × 𝒓𝑖 , (28.13)
𝑱𝑖 𝝎̇ = 𝒇𝑖𝜔 − 𝝎 × 𝑱𝑖 𝝎. (28.14)
This can be compactly written as
𝑚 −𝑚𝑖 𝑹𝑖 𝒙̈ 𝒇𝑖𝑥 − 𝑚𝑖 𝝎 × 𝝎 × 𝒓𝑖
[ 𝑖 ][ ] = [ 𝜔 ]. (28.15)
𝟎 𝑱𝑖 𝝎̇ 𝒇 𝑖 − 𝝎 × 𝑱𝑖 𝝎
It remains to use the principle of virtual work, employing (28.8), to reduce the number of
equations (rigid body and slave nodes) to the generalized rigid body equations. The result
of this endeavor is
𝑱𝝎̇ = 𝒇 𝜔 . (28.20)
where
𝑀 = 𝑀𝜌 + ∑ 𝑚𝑖 , (28.21)
𝑖∈𝑆
𝑱 = 𝑱𝜌 + ∑ 𝑱𝑖 + ∑ 𝑚𝑖 𝑹𝑖𝑇 𝑹𝑖 − 𝑀𝑹𝑧−𝑥
𝑇
𝑹𝑧−𝑥 , (28.23)
𝑖∈𝑆 𝑖∈𝑆
Equations (28.19)-(28.24) are the generalized rigid body equations to be solved for 𝒙
and 𝑸, cf. (28.5) and (28.18). The mass in (28.21) and inertia tensor in (28.23) are the
physical mass and physical inertia augmented by slave node properties; nodal masses 𝑚𝑖 ,
inertias 𝑱𝑖 and locations 𝒙𝑖 . We denote 𝑀 the algorithmic mass, which may not reflect what
the user intuitively expects when using *MAT_RIGID to make a part rigid. Similarly 𝑱 is
the algorithmic inertia tensor, and it is worth noting that a nodal rigid body must therefore
be connected to deformable elements or otherwise 𝑀 = 0 and 𝑱 = 𝟎 and its whereabouts
will be impossible to determine. For no mass scaling, all these properties are constant
(except for rotational updates of the inertia tensors) and can essentially be calculated at
time zero. If mass scaling is active, the slave nodal masses and inertias include the added
mass due to mass scaling and therefore change over time. This means that inertia
properties should be recomputed every time step to account for these changes, but the
default behavior is that this is done only for nodal rigid bodies and not for regular rigid
bodies. Presumably this is based on the assumption that the influence from slave nodes is
significant for nodal rigid bodies and not so much for regular rigid bodies, which is
probably true as long as the number of contiguous nodes is small compared to the total
number of nodes in the rigid body. Nevertheless, with RBSMS = 1 on *CONTROL_RIGID,
these extra masses are accounted for and equations (28.19)-(28.24) are solved as expressed
herein. This amounts to transforming 𝒙 to 𝒛 before the update, then update 𝒛, and
transform back to obtain the new 𝒙. As we now turn to the algorithmic update of the rigid
body location, we restrict ourselves to a special case for the sake of simplifying the
exposition;
From (28.19)-(28.20) can readily solve for the rigid body accelerations
𝒇𝑥
𝒙̈ = , (28.25)
𝑀
𝝎̇ = 𝑱 −1 𝒇 𝜔 . (28.26)
It turns out that the algorithmic mass 𝑀 can be calculated as
𝑀=∑ 𝑀𝑖 , (28.27)
𝑖∈𝑆
where 𝑀𝑖 is the mass of node 𝑖 as obtained from the mass of its associated elements
(integrating material density 𝜌 by shape functions 𝜑𝑖 over element domain). Furthermore𝒙
can be approximated from
𝑀𝒙 = ∑ 𝑀 𝑖 𝒙𝑖 . (28.28)
𝑖∈𝑆
Likewise, the inertia tensor is approximated by a nodal summation of the product of the
point masses with their moment arms
𝑱=∑ 𝑀𝑖 𝑹𝑖𝑇 𝑹𝑖 . (28.29)
𝑖∈𝑆
The initial velocities of the slave nodes are readily calculated for a rigid body from (28.6).
For arbitrary orientations of the body, the inertia tensor is transformed each time step
based on the incremental rotations using the standard rules of second-order tensors:
𝑱 𝑛+1 = 𝑨𝑱 𝑛 𝑨𝑇 (28.30)
where 𝑱 𝑛+1 is the updated inertia tensor components in the global frame. The transfor-
mation matrix 𝑨 is not stored since the formulation is incremental, but recomputed as
explained below. After calculating the rigid body accelerations from Equation (28.25) and
(28.26), the rigid body translational and rotational increment, ∆𝒙 and ∆𝜽, can be calculated
using the time step ∆𝑡 and the explicit time integration update. The translational
coordinate is then updated as
𝒙 𝑛+1 = 𝒙 𝑛 + ∆𝒙 (28.31)
and ∆𝜽 is used to calculate 𝑨 in (28.30) using the Hughes-Winget algorithm,
1 1
𝑨=𝑰+ 𝑇 ∆𝜽
(𝑰 + 𝛥𝑺) 𝛥𝑺, (28.32)
∆𝜽 2
1+
4
𝛥𝑺𝒓 = ∆𝜽 × 𝒓, ∀𝒓. (28.33)
The coordinates of the slave nodes are incrementally updated
𝒙𝑖𝑛+1 = 𝒙𝑖𝑛 + ∆𝒙 + (𝑨 − 𝑰)𝒓𝑖𝑛 , (28.34)
and the velocities of the nodes are calculated by differencing the coordinates
(𝒙𝑖𝑛+1 − 𝒙𝑖𝑛 )
𝒙̇𝑖 = . (28.35)
𝛥𝑡
A direct integration of the rigid body accelerations into velocity and displacements is not
used for two reasons: (1) calculating the rigid body accelerations of the nodes is more
expensive than the current algorithm, and (2) the second-order accuracy of the central
difference integration method would introduce distortion into the rigid bodies. Since the
accelerations are not needed within the program, they are calculated by a post-processor
using a difference scheme similar to the above.
∂𝐶(𝑥𝑖 , 𝑥𝑗 )
𝑓𝑗 = −𝑘𝐶(𝑥𝑖 , 𝑥𝑗 ) . (28.37)
∂𝑥𝑗
The forces acting at the nodes have to convert into forces acting on the rigid bodies.
Recall that velocities of a node i is related to the velocity of the center of mass of a rigid
body by Equation (28.6). By using Equation (28.6) and virtual power arguments, it may be
shown that the generalized forces are:
𝐹𝑖𝑥 = 𝑓𝑖 , (28.38)
The magnitude of the penalty stiffness 𝑘 is chosen so that it does not control the
stable time step size. For the central difference method, the stable time step Δ𝑡 is restricted
by the condition that,
2
Δ𝑡 = , (28.40)
𝛺
where 𝛺 is the highest frequency in the system. The six vibrational frequencies associated
with each rigid body are determined by solving their eigenvalue problems assuming 𝑘 = 1.
For a body with 𝑚 constraint equations, the linearized equations of the translational
degrees of freedom are
𝑀𝑋̈ + 𝑚𝑘𝑋 = 0, (28.41)
and the frequency is √𝑚𝑘/𝑀 where 𝑀 is the mass of the rigid body. The corresponding
rotational equations are
𝐉𝛉̈ + 𝐊𝛉 = 0, (28.42)
𝐉 is the inertia tensor and 𝐊 is the stiffness matrix for the moment contributions from the
penalty constraints. The stiffness matrix is derived by noting that the moment contribution
of a constraint may be approximated by
𝐅𝑥 = −k𝐫𝑖 × (𝛉 × 𝐫𝑖 ), (28.43)
𝐫𝑖 = 𝐱𝑖 − 𝐗cm , (28.44)
and noting the identity,
𝐀 × (𝐁 × 𝐂) = |𝐀 ⋅ 𝐂 − 𝐀 ⊗ 𝐂|𝐁, (28.45)
so that
𝑚
𝐾 = ∑ 𝑘[𝐫𝑖 ⋅ 𝐫𝑖 − 𝐫𝑖 ⊗ 𝐫𝑖 ]. (28.46)
𝑖=1
The rotational frequencies are the roots of the equation det∣𝐊 − 𝛺2 𝐉∣ = 0, which is
cubic in 𝛺2 . Defining the maximum frequency over all rigid bodies for 𝑘 = 1 as 𝛺max , and
introducing a time step scale factor TSSF, the equation for 𝑘 is
2TSSF 2
𝑘≤( ) , (28.47)
Δ𝑡Ωmax
The joint constraints are defined in terms of the displacements of individual nodes.
Regardless of whether the node belongs to a solid element or a structural element, only its
translational degrees of freedom are used in the constraint equations.
A spherical joint is defined for nodes 𝑖 and 𝑗 by the three constraint equations,
𝑥1𝑖 − 𝑥1𝑗 = 0 𝑥2𝑖 − 𝑥2𝑗 = 0 𝑥3𝑖 − 𝑥3𝑗 = 0, (28.48)
and a revolute joint, which requires five constraints, is defined by two spherical joints, for a
total of six constraint equations. Since a penalty formulation is used, the redundancy in the
joint constraint equations is unimportant. A cylindrical joint is defined by taking a revolute
joint and eliminating the penalty forces along the direction defined by the two spherical
joints. In a similar manner, a planar joint is defined by eliminating the penalty forces that
are perpendicular to the two spherical joints.
The translational joint is a cylindrical joint that permits sliding along its axis, but not
rotation. An additional pair of nodes is required off the axis to supply the additional
constraint. The only force active between the extra nodes acts in the direction normal to the
plane defined by the three pairs of nodes.
The universal joint is defined by four nodes. Let the nodes on one body be 𝑖 and 𝑘,
and the other body, 𝑗 and 𝑙. Two of them, 𝑖 and 𝑗, are used to define a spherical joint for the
first three constraint equations. The fourth constraint equation is,
𝐶(𝑥𝑖 , 𝑥𝑗 , 𝑥𝑘 , 𝑥𝑙 ) = (𝑥𝑘 − 𝑥𝑖 ) ⋅ (𝑥𝑖 − 𝑥𝑗 ) = 0, (28.49)
∂𝐶
and is differentiated to give the penalty forces 𝑓𝑛 = −𝑘𝐶 ∂𝑥 , where 𝑛 ranges over the four
𝑛
nodes numbers.
For rigid body switching to work properly the choice of the shell element
formulation is critical. The Hughes-Liu elements cannot currently be used for two reasons.
First, since these elements compute the strains from the rotations of the nodal fiber vectors
from one cycle to the next, the nodal fiber vectors should be updated with the rigid body
motions and this is not done. Secondly, the stresses are stored in the global system as
opposed to the co-rotational system. Therefore, the stresses would also need to be
transformed with the rigid body motions or zeroed out. The co-rotational elements of
Belytschko and co-workers do not reference nodal fibers for the strain computations and
the stresses are stored in the co-rotational coordinate system which eliminates the need for
the transformations; consequently, these elements can be safely used. The membrane
elements and airbag elements are closely related to the Belytschko shells and can be safely
used with the switching options.
The beam elements have nodal triads that are used to track the nodal rotations and
to calculate the deformation displacements from one cycle to the next. These nodal triads
are updated every cycle with the rigid body rotations to prevent non-physical behavior
when the rigid body is switched back to deformable. This applies to all beam element
formulations in LS-DYNA. The Belytschko beam formulations are preferred for the
switching options for like the shell elements, the Hughes-Liu beams keep the stresses in the
global system. Truss elements like the membrane elements are trivially treated and pose
no difficulties.
The brick elements store the stresses in the global system and upon switching the
rigid material to deformable the element stresses are zeroed to eliminate spurious behavior.
The implementation addresses many potential problems and has worked well in practice.
The current restrictions can be eliminated if the need arises and anyway should pose no
insurmountable problems. We will continue to improve this capability if we find that it is
becoming a popular option.
•Spot weld.
•Fillet weld
•Butt weld
•Cross fillet weld
•General weld
z z
node 2 node 3
node 1 node 2
y y
2 node spotweld 3 node spotweld
node 1
x x
z
node n
node n-1
n node spotweld
y
node 2
x
node 1
Figure 28.2. Nodal ordering and orientation of the local coordinate system is
important for determining spotweld failure.
Welds can fail three ways: by ductile failure which is based on the effective plastic
strain, by brittle failure which is based on the force resultants acting on the rigid body
weld, and by a failure time which is specified in the input. When effective plastic strain is
used the weld fails when the nodal plastic strain exceeds the input value. A least squares
algorithm is used to generate the nodal values of plastic strains at the nodes from the
element integration point values. The plastic strain is integrated through the element and
the average value is projected to the nodes with a least square fit. In the resultant based
brittle failure the resultant forces and moments on each node of the weld are computed.
These resultants are checked against a failure criterion which is expressed in terms of these
resultants. The forces may be averaged over a user specified number of time steps to
eliminate breakage due to spurious noise. After all nodes of a weld are released the rigid
body is removed from the calculation.
Spotwelds are shown in Figure 28.2. Spotweld failure due to plastic straining occurs
p
when the effective nodal plastic strain exceeds the input value, 𝜀fail . This option can model
the tearing out of a spotweld from the sheet metal since the plasticity is in the material that
surrounds the spotweld, not the spotweld itself. This option should only be used for the
material models related to metallic plasticity and can result is slightly increased run times.
Brittle failure of the spotwelds occurs when:
𝑛 𝑚
max(𝑓𝑛 , 0) ∣𝑓 ∣
( ) + ( 𝑠 ) ≥ 1, (28.50)
𝑆𝑛 𝑆𝑠
where 𝑓𝑛 and 𝑓𝑠 are the normal and shear interface force. Component 𝑓𝑛 contributes for
tensile values only. When the failure time, 𝑡f , is reached the nodal rigid body becomes
inactive and the constrained nodes may move freely. In Figure 28.2 the ordering of the
nodes is shown for the 2 and 3 noded spotwelds. This order is with respect to the local
coordinate system where the local 𝑧 axis determines the tensile direction. The nodes in the
spotweld may coincide but if they are offset the local system is not needed since the 𝑧-axis
is automatically oriented based on the locations of node 1, the origin, and node 2. The
failure of the 3 noded spotweld may occur gradually with first one node failing and later
the second node may fail. For 𝑛 noded spotwelds the failure is progressive starting with
the outer nodes (1 and 𝑛) and then moving inward to nodes 2 and 𝑛 − 1. Progressive
failure is necessary to preclude failures that would create new rigid bodies.
Ductile fillet weld failure, due to plastic straining, is treated identically to spotweld
failure. Brittle failure of the fillet welds occurs when:
z local coordinate z
system
α
2 2 node fillet weld
1 x y
a L
Figure 28.3. Nodal ordering and orientation of the local coordinate system is
shown for fillet weld failure.
where
𝜎𝑛 = normal stress
𝜏𝑛 = shear stress in direction of weld (local 𝑦)
𝜏𝑡 = shear stress normal to weld (local 𝑥)
𝜎𝑓 = failure stress
𝛽 = failure parameter
Component 𝜎𝑛 is nonzero for tensile values only. When the failure time, 𝑡𝑓 , is
reached the nodal rigid body becomes inactive and the constrained nodes may move freely.
In Figure 28.3 the ordering of the nodes is shown for the 2 node and 3 node fillet welds.
This order is with respect to the local coordinate system where the local z axis determines
the tensile direction. The nodes in the fillet weld may coincide. The failure of the 3 node
fillet weld may occur gradually with first one node failing and later the second node may
fail.
In Figure 28.4 the butt weld is shown. Ductile butt weld failure, due to plastic
straining, is treated identically to spotweld failure. Brittle failure of the butt welds occurs
when:
2 tied nodes
Lt
L y
4 tied nodes
Figure 28.4. Orientation of the local coordinate system and nodal ordering is
shown for butt weld failure.
where
26-12 (Rigid Body Dynamics) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Rigid Body Dynamics
2
3
1
z1 z2 z3
x2 x3
y1
x1 y2 2 y3
3 3
1
Figure 28.5. A simple cross fillet weld illustrates the required input. Here
NFW = 3 with nodal pairs (A = 2, B = 1), (A = 3, B = 1), and (A = 3, B = 2). The
local coordinate axes are shown. These axes are fixed in the rigid body and are
referenced to the local rigid body coordinate system which tracks the rigid body
rotation.
𝜎𝑛 = normal stress
𝜏𝑛 = shear stress in direction of weld (local y)
𝜏𝑡 = shear stress normal to weld (local z)
𝜎𝑓 = failure stress
𝛽 = failure parameter
Component 𝜎𝑛 is nonzero for tensile values only. When the failure time, 𝑡𝑓 , is reached the
nodal rigid body becomes inactive and the constrained nodes may move freely. The nodes
in the butt weld may coincide.
The cross fillet weld and general weld are shown in Figures 28.5 and 28.6,
respectively. The treatment of failure for these welds is based on the formulation for the
fillet and butt welds.
29
Contact-Impact Algorithm
29.1 Introduction
The treatment of sliding and impact along interfaces has always been an important
capability in DYNA3D and more recently in LS-DYNA. Three distinct methods for
handling this have been implemented, which we will refer to as the kinematic constraint
method, the penalty method, and the distributed parameter method. Of these, the first
approach is now used for tying interfaces. The relative merits of each approach are
discussed below.
Today, automatic contact definitions are commonly used. In this approach the slave
and master surfaces are generated internally within LS-DYNA from the part ID's given for
each surface. For automotive crash models it is quite common to include the entire vehicle
in one single surface contact definition where the all the nodes and elements within the
interface can interact.
Problems arise with this method when the master surface zoning is finer than the
slave surface zoning as shown in two dimensions in Figure 29.1. Here, certain master
nodes can penetrate through the slave surface without resistance and create a kink in the
slide line. Such kinks are relatively common with this formulation, and, when interface
pressures are high, these kinks occur whether one or more quadrature points are used in
the element integration. It may be argued, of course, that better zoning would minimize
such problems; but for many problems that are of interest, good zoning in the initial
configuration may be very poor zoning later. Such is the case, for example, when gaseous
products of a high explosive gas expand against the surface of a structural member.
slave surface
master surface
Quite in contrast to the nodal constraint method, the penalty method approach is
found to excite little if any mesh hourglassing. This lack of noise is undoubtedly
attributable to the symmetry of the approach. Momentum is exactly conserved without the
necessity of imposing impact and release conditions. Furthermore, no special treatment of
intersecting interfaces is required, greatly simplifying the implementation.
In the distributed parameter formulation, one-half the slave element mass of each
element in contact is distributed to the covered master surface area. Also, the internal
stress in each element determines a pressure distribution for the master surface area that
receives the mass. After completing this distribution of mass and pressure, we can update
∂B2
∂B 1
B20
B10
the acceleration of the master surface. Constraints are then imposed on slave node
accelerations and velocities to insure their movement along the master surface. Unlike the
finite difference hydro programs, we do not allow slave nodes to penetrate; therefore we
avoid “put back on” logic. In another simplification, our calculation of the slave element
relative volume ignores any intrusion of the master surfaces. The HEMP and TENSOR
codes consider the master surface in this calculation.
29.5 Preliminaries
S1
ns
S4
ms
S2
S3
X3
X2
X1
Figure 29.3. In this figure, four master segments can harbor slave node 𝑛𝑠 given
that 𝑚𝑠 is the nearest master node.
Consider a slave node, 𝑛𝑠 , sliding on a piecewise smooth master surface and assume
that a search of the master surface has located the master node, 𝑚𝑠 , lying nearest to 𝑛𝑠 .
Figure 29.3 depicts a portion of a master surface with nodes 𝑚𝑠 and 𝑛𝑠 labeled. If 𝑚𝑠 and 𝑛𝑠
do not coincide, 𝑛𝑠 can usually be shown to lie in a segment 𝑠1 via the following tests:
(𝐜𝑖 × 𝐬) ⋅ (𝐜𝑖 × 𝐜𝑖+1 ) > 0, (29.3)
Since the sliding constraints keep 𝑛𝑠 close but not necessarily on the master surface
and since 𝑛𝑠 may lie near or even on the intersection of two master segments, the
inequalities of Equation (29.3) may be inconclusive, i.e., they may fail to be satisfied or
more than one may give positive results. When this occurs 𝑛𝑠 is assumed to lie along the
intersection which yields the maximum value for the quantity
𝐠 ⋅ 𝐜𝑖
𝑖 = 1,2,3,4, .. (29.6)
|𝐜𝑖 |
When the contact surface is made up of badly shaped elements, the segment apparently
identified as containing the slave node actually may not, as shown in Figure 29.5.
Assume that a master segment has been located for slave node 𝑛𝑠 and that 𝑛𝑠 is not
identified as lying on the intersection of two master segments. Then the identification of
the contact point, defined as the point on the master segment which is nearest to 𝑛𝑠 ,
becomes nontrivial. For each master surface segment, 𝑠1 is given the parametric
representation of Equation (1.7), repeated here for clarity:
ns
s
g
ci+1 ms
X3
X2
X1
Figure 29.4. Projection of g onto master segment 𝑠1
1
2 8
7
6
3
5
4
Figure 29.5. When the nearest node fails to contain the segment that harbors the
slave node, segments numbered 1-8 are searched in the order shown.
Let t be a position vector drawn to slave node ns and assume that the master surface
segment s1 has been identified with ns . The contact point coordinates (𝜉c , 𝜂c ) on s1 must
satisfy
∂𝐫
(𝜉 , 𝜂 ) ⋅ [𝐭 − 𝐫(𝜉𝑐 , 𝜂𝑐 )] = 0, (29.10)
∂𝜉 𝑐 𝑐
∂𝐫
(𝜉 , 𝜂 ) ⋅ [𝐭 − 𝐫(𝜉𝑐 , 𝜂𝑐 )] = 0. (29.11)
∂𝜂 𝑐 𝑐
The physical problem is illustrated in Figure 29.6, which shows ns lying above the
master surface. Equations (29.10) and (29.11) are readily solved for 𝜉c and 𝜂c . One way to
accomplish this is to solve Equation (29.10) for 𝜉c in terms of 𝜂c , and substitute the results
3 2
ns
ξ
X3 t
4
1
r
X2
X1
Figure 29.6. Location of contact point when ns lies above master segment.
into Equation (29.11). This yields a cubic equation in 𝜂c which is presently solved
numerically in LS-DYNA. In the near future, we hope to implement a closed form solution
for the contact point.
The equations are solved numerically. When two nodes of a bilinear quadrilateral
are collapsed into a single node for a triangle, the Jacobian of the minimization problem is
singular at the collapsed node. Fortunately, there is an analytical solution for triangular
segments since three points define a plane. Newton-Raphson iteration is a natural choice
for solving these simple nonlinear equations. The method diverges with distorted elements
unless the initial guess is accurate. An exact contact point calculation is critical in post-
buckling calculations to prevent the solution from wandering away from the desired
buckling mode.
Three iterations with a least-squares projection are used to generate an initial guess:
𝜉0 = 0, 𝜂0 = 0,
𝐫,𝜉 Δ𝜉 𝐫,𝜉
[𝐫 ] [𝐫,𝜉 𝐫,𝜂 ] { } = [𝐫 ] {𝐫(𝜉𝑖, 𝜂𝑖 ) − 𝐭}, (29.12)
,𝜂 Δ𝜂 ,𝜂
In concave regions, a slave node may have isoparametric coordinates that lie outside
of the [−1, +1] range for all of the master segments, yet still have penetrated the surface. A
simple strategy is used for handling this case, but it can fail. The contact segment for each
node is saved every time step. If the slave node contact point defined in terms of the
isoparametric coordinates of the segment, is just outside of the segment, and the node
penetrated the isoparametric surface, and no other segment associated with the nearest
neighbor satisfies the inequality test, then the contact point is assumed to occur on the edge
of the segment. In effect, the definition of the master segments is extended so that they
overlap by a small amount. In the hydrocode literature, this approach is similar to the slide
line extensions used in two dimensions. This simple procedure works well for most cases,
but it can fail in situations involving sharp concave corners.
Penetration of the slave node ns through the master segment which contains its
contact point is indicated if
𝑙 = 𝐧𝑖 × [𝐭 − 𝐫(𝜉𝑐 , 𝜂𝑐 )] < 0, (29.14)
where
𝐧𝑖 = 𝐧𝑖 (𝜉𝑐 , 𝜂𝑐 ) (29.15)
is normal to the master segment at the contact point.
If slave node ns has penetrated through master segment 𝑠𝑖 , we add an interface force
vector 𝐟s :
𝐟𝑠 = −𝑙𝑘𝑖 𝐧𝑖 if 𝑙 < 0 (29.16)
to the degrees of freedom corresponding to ns and
𝑓𝑚𝑖 = 𝜙𝑖 (𝜉𝑐 , 𝜂𝑐 )𝑓𝑠 if 𝑙 < 0 (29.17)
to the four nodes (𝑖 = 1,2,3,4) that comprise master segment 𝑠𝑖 . The stiffness factor 𝑘𝑖 for
master segment 𝑠𝑖 is given in terms of the bulk modulus 𝐾𝑖 , the volume 𝑉𝑖 , and the face
area 𝐴𝑖 of the element that contains 𝑠𝑖 as
𝑓𝑠𝑖 𝐾𝑖 𝐴2𝑖
𝑘𝑖 = (29.18)
𝑉𝑖
for brick elements and
𝑓𝑠𝑖 𝐾𝑖 𝐴𝑖
𝑘𝑖 = (29.19)
max(shell diagonal)
for shell elements where 𝑓𝑠𝑖 is a scale factor for the interface stiffness and is normally
defaulted to .10. Larger values may cause instabilities unless the time step size is scaled
back in the time step calculation.
In LS-DYNA, a number of options are available for setting the penalty stiffness
value. This is often an issue since the materials in contact may have drastically different
bulk modulii. The calculational choices are:
The default may sometimes fail due to an excessively small stiffness. When this
occurs it is necessary to manually scale the interface stiffness. Care must be taken not to
induce an instability when such scaling is performed. If the soft material also has a low
density, it may be necessary to reduce the scale factor on the computed stable time step.
2
1
𝑘cs (𝑡) = 0.5 ⋅ SOFSCL ⋅ 𝑚∗ ⋅ ( ) , (29.20)
Δ𝑡𝑐 (𝑡)
where SOFSCL on Optional Card A of *CONTROL_CONTACT is the scale factor for the
Soft Constraint Penalty Formulation, 𝑚∗ is a function of the mass of the slave node and of
the master nodes. Δ𝑡cs is set to the initial solution timestep. If the solution time step grows,
Δ𝑡c is reset to the current time step to prevent unstable behavior.
A comparative check against the contact stiffness calculated with the traditional
penalty formulation, 𝑘soft=0 , and in general the maximum stiffness between the two is
taken,
𝑘soft=1 = max{𝑘cs , 𝑘soft=0 }. (29.21)
{SFS ⎫
⎧ 2
} 𝑚 𝑚 1
𝑘cs (𝑡) = 0.5 ⋅ SLSFAC ⋅ ⎨or ⎬ ( 1 2 ) ( ) . (29.22)
{
⎩SFM}⎭ 𝑚1 + 𝑚2 Δ𝑡𝑐 (𝑡)
Segment masses are used rather than nodal masses. Segment mass is equal to the
element mass for shell segments and half the element mass for solid element segments.
Like the Soft Constraint Penalty Formulation, 𝑑𝑡 is set to the initial solution time step which
is updated if the solution time step grows larger to prevent unstable behavior. However, it
differs from SOFT = 1 in how 𝑑𝑡 is updated. 𝑑𝑡 is updated only if the solution time step
grows by more than 5%. This allows 𝑑𝑡 to remain constant in most cases, even if the
solution time step slightly grows.
slave node
Figure 29.7. Failure to find the contact segment can be caused by poor aspect
ratios in the finite element mesh.
In metal forming applications, problems with the contact searching were found
when the rigid body stamping dies were meshed with elements having very poor aspect
ratios. The nearest node algorithm described above can break down since the nearest node
is not necessarily anywhere near the segment that harbors the slave node as is assumed in
Figure 29.5 (see Figure 29.7). Such distorted elements are commonly used in rigid bodies in
order to define the geometry accurately.
locate the possible segments that contain the slave node of interest. For each quadrilateral
27-12 (Contact-Impact Algorithm) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Contact-Impact Algorithm
Figure 29.9. Interior points are constructed in the segments for determining the
closest point to the slave node.
segment, four points are constructed at the centroids of the four triangles each defined by 3
nodes as shown in Figure 29.9 where the black point is the centroid of the quadrilateral.
These centroids are used to find the nearest point to the slave node and hence the nearest
segment. The nodes of the three nearest segments are then examined to identify the three
nearest nodes. Just one node from each segment is allowed to be a nearest node.
When the nearest segment fails to harbor the slave node, the adjacent segments are
checked. The old algorithm checks the segments labeled 1-3 (Figure 29.10), which do not
contain the slave node, and fails.
slave node
segment identified as containing slave node
Figure 29.10. In case the stored segment fails to contain the node, the adjacent
segments are checked.
metal. Unless thickness is considered in the contact, the effect of thinning on frictional
interface stresses due to membrane stretching will be difficult to treat. In the treatment of
thickness we project both the slave and master surfaces based on the mid-surface normal
projection vectors as shown in Figure 29.11. The surfaces, therefore, must be offset by an
amount equal to 1/2 their total thickness (Figure 29.12). This allows DYNA3D to check the
node numbering of the segments automatically to ensure that the shells are properly
oriented.
Thickness changes in the contact are accounted for if and only if the shell thickness
change option is flagged in the input. Each cycle, as the shell elements are processed, the
nodal thicknesses are stored for use in the contact algorithms. The interface stiffness may
change with thickness depending on the input options used.
Type 5 contact considers nodes interacting with a surface. This algorithm calls
exactly the same subroutines as surface-to-surface but not symmetrically: i.e., the
subroutines are called once, not twice. To account for the nodal thickness, the maximum
shell thickness of any shell connected to the node is taken as the nodal thickness and is
Figure 29.12. The slave and master surfaces must be offset in the input by one-
half the total shell thickness. This also allows the segments to be oriented
automatically.
updated every cycle. The projection of the node is done normal to the contact surface as
shown in Figure 29.13.
The need to offset contact surfaces to account for the thickness of the shell elements
contributes to initial contact interpenetrations. These interpenetrations can lead to severe
numerical problems when execution begins so they should be corrected if LS-DYNA is to
run successfully. Often an early growth of negative contact energy is one sign that initial
interpenetrations exist. Currently, warning messages are printed to the terminal, the
D3HSP file, and the MESSAG file to report interpenetrations of nodes through contact
segments and the modifications to the geometry made by LS-DYNA to eliminate the
interpenetrations. Sometimes such corrections simply move the problem elsewhere since it
is very possible that the physical location of the shell mid-surface and possibly the shell
thickness are incorrect. In the single surface contact algorithms any nodes still
interpenetrating on the second time step are removed from the contact with a warning
message.
generation phase.
LS-DYNA DEV 10/27/16 (r:8004) 27-15 (Contact-Impact Algorithm)
Contact-Impact Algorithm LS-DYNA Theory Manual
•Use consistently refined meshes on adjacent parts which have significant curva-
tures.
•Be very careful when defining thickness on shell and beam section definitions --
especially for rigid bodies.
•Scale back part thickness if necessary. Scaling a 1.5mm thickness to .75mm should
not cause problems but scaling to .075mm might. Alternatively, define a smaller
contact thickness by part ID. Warning: if the part is too thin contact failure will
probably occur
•Use spot welds instead of merged nodes to allow the shell mid surfaces to be
offset.
Brick
shell
where 𝑛𝑠𝑛 is the number of slave nodes, 𝑛𝑚𝑛 is the number of master nodes, Δ𝐹𝑖slave is the
interface force between the ith slave node and the contact segment Δ𝐹𝑖master is the interface
force between the ith master node and the contact segment, Δ𝑑𝑖𝑠𝑡𝑖slave is the incremental
distance the ith slave node has moved during the current time step, and Δ𝑑𝑖𝑠𝑡𝑖master is the
incremental distance the ith master node has moved during the current time step. In the
absence of friction the slave and master side energies should be close in magnitude but
opposite in sign. The sum, 𝐸contact , should equal the stored energy. Large negative contact
energy is usually caused by undetected penetrations. Contact energies are reported in the
SLEOUT file. In the presence of friction and damping discussed below the interface energy
can take on a substantial positive value especially if there is, in the case of friction,
substantial sliding.
The master mass 𝑚master is interpolated from the master nodes of the segment containing
the slave node using the basis functions evaluated at the contact point of the slave node.
Force oscillations often occur as curved surfaces undergo relative motion. In these
cases contact damping will eliminate the high frequency content in the contact reaction
forces but will be unable to damp the lower frequency oscillations caused by nodes moving
from segment to segment when there is a large angle change between the segments. This is
shown in the hemispherical punch deep drawing in Figure 29.16. The reaction forces with
and without contact damping in Figure 29.17 show only minor differences since the
oscillations are not due to the dynamic effects of explicit integration. However, refining the
mesh as shown in Figure 29.18 to include more elements around the die corner as in Figure
29.18 greatly reduces the oscillations as shown in Figure 29.19. This shows the importance
of using an adequate mesh density in applications where significant relative motion is
expected around sharp corners.
Figure 29.19. The oscillations are effectively eliminated by the mesh refinement.
Friction
Friction in LS-DYNA is based on a Coulomb formulation. Let 𝐟∗ be the trial force,𝐟𝑛
the normal force, 𝑘 the interface stiffness, 𝜇 the coefficient of friction, and 𝐟𝑛 the frictional
force at time 𝑛. The frictional algorithm, outlined below, uses the equivalent of an elastic
plastic spring. The steps are as follows:
𝑛+1
𝐹𝑦 𝐟∗
𝐟 = if ∣𝐟∗ ∣ > 𝐹𝑦 (29.29)
|𝐟∗ |
The interface shear stress that develops as a result of Coulomb friction can be very
large and in some cases may exceed the ability of the material to carry such a stress. We
therefore allow another limit to be placed on the value of the tangential force:
𝑛+1
𝑓 𝑛+1 = min(𝑓Coulomb , 𝜅𝐴master ), (29.32)
where 𝐴master is the area of the master segment and 𝜅 is the viscous coefficient. Since more
than one node may contribute to the shear stress of a segment, we recognize that the stress
may still in some cases exceed the limit 𝜅.
Typical values of friction, see Table 26.1, can be found in Marks Engineering
Handbook.
27-20 (Contact-Impact Algorithm) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Contact-Impact Algorithm
Tied interfaces include four interface options of which three are in the Sliding
Interface Definition Section in the LS-DYNA User’s Manual. These are:
• Type 7 for tying both translational and rotational degrees of freedom of nodes
The fourth option is in the “Tie-Breaking Shell Definitions” Section of the user’s
manual and is meant as a way of tying edges of adjacent shells together. Unlike Type 7 this
latter option does not require a surface definition, simply nodal lines, and includes a failure
model based on plastic strain which can be turned off by setting the plastic failure strain to
a high value. The first two options, which are equivalent in function but differ in the input
definition, can be properly applied to nodes of elements which lack rotational degrees of
freedom. The latter options must be used with element types that have rotational degrees
of freedom defined at their nodes such as the shell and beam elements. One important
application of Type 7 is that it allows edges of shells to be tied to shell surfaces. In such
transitions the shell thickness is not considered. Exceptions from these latter statements is
in case of invoking the implicit accuracy option, see *CONTROL_ACCURACY, for which a
node with rotational degrees of freedom can tie to any element with or without offset. In
this case moments are consistently transferred based on the kinematics of the chosen tied
interface, the theory for this is presented in Section.
Since the constraints are imposed only on the slave nodes, the more coarsely meshed
side of the interface is recommended as the master surface. Ideally, each master node
should coincide with a slave node to ensure complete displacement compatibility along the
interface, but in practice this is often difficult if not impossible to achieve. In other words,
master nodes that do not coincide with a slave node can interpenetrate through the slave
surface.
The interpolated contact point, (𝜉𝑐 , 𝜂𝑐 ), for each slave node is computed once, since
its relative position on the master segment is constant for the duration of the calculation. If
the closest point projection of the slave node to the master surface is non-orthogonal,
values of (𝜉𝑐 , 𝜂𝑐 ) greater than unity will be computed. To allow for slight errors in the
mesh definition, the slave node is left unconstrained if the magnitude of the contact point
27-22 (Contact-Impact Algorithm) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Contact-Impact Algorithm
exceeds 1.02. Great care should be exercised in setting up tied interfaces to ensure that the
slave nodes are covered by master segments.
Conflicting constraints must be avoided. Care should be taken not to include nodes
that are involved in a tied interfaces in another tied interface, in constraint sets such as
nodal constraint sets, in linear constraint equations, and in spot welds. Furthermore, tied
interfaces between rigid and deformable bodies are not permitted. LS-DYNA checks for
conflicting constraints on nodal points and if such conflicts are found, the calculation will
terminate with an error message identifying the conflict. Nodes in tied interfaces should
not be included as slave nodes in rigid wall definitions since interactions with stonewalls
will cause the constraints that were applied in the tied interface logic to be violated. We do
not currently check for this latter condition is LS-DYNA.
Tied interfaces require coincident surfaces and for shell element this means that the
mid-surfaces must be coincident. Consider Figure 29.21 where identical slave and master
surfaces are offset. In this case the tied constraints require that translational velocities of
tied nodes be identical, i.e.,
𝐯𝑠 = 𝐯𝑚 . (29.35)
Consequently, if the nodes are offset, rotations are not possible. The velocity of a tied slave
node in Figure 29.21 should account for the segment rotation:
𝐯𝑠 = 𝐯𝑚 − 𝑧̂ 𝐞3 × 𝛚, (29.36)
where 𝑧̂ is the distance to the slave node, 𝐞3 is the normal vector to the master surface at the
contact point, and 𝛚 is the angular velocity. Since this is not the case in the tied interfaces
logic, 𝑧̂ must be of zero length.
LS-DYNA projects tied slave nodes back to the master surface if possible and prints
warning messages for all projected offset nodes or nodes too far away to tie. This
projection eliminates the problems with rotational constraints but creates other difficulties:
• Geometry is modified
• Tied interfaces must be excluded from automatic generation since tied surfaces
cannot be mixed with automatic contact with thickness offsets.
Slave Surface
Master Surface
Figure 29.21. Offset tied interface.
An offset capability has been added to the tied interfaces which uses a penalty
approach. The penalty approach removes the major limitations of the constraint
formulation since with the offset option:
*CONTACT_TIED_NODES_TO_SURFACE_CONSTRAINED_OFFSET
*CONTACT_TIED_NODES_TO_SURFACE_OFFSET
*CONTACT_TIED_SHELL_EDGE_TO_SURFACE_CONSTRAINED_OFFSET
*CONTACT_TIED_SHELL_EDGE_TO_SURFACE_BEAM_OFFSET
The first of these two do not consider rotational degrees of freedom, whereas the other two
do. Furthermore, the first and third are constraint based and the other two are penalty
based, so all in all these four cover much of what a user expects from a tied interface. By
setting IACC to 1 on *CONTROL_ACCURACY any of the tied contact options mentioned
above (and the non-offset counterparts as a side effect, i.e.,
*CONTACT_TIED_NODES_TO_SURFACE and
*CONTACT_TIED_SHELL_EDGE_TO_SURFACE) are treated with this strongly objective
formulation. In addition to being strongly objective, i.e., forces and moments transform
correctly under superposed rigid body motions in a single implicit step, this formulation
applies rotational constraints consistently when and only when necessary. This means not
only that slave nodes without rotational degrees of freedom are not rotationally
constrained, but also that bending and torsional rotations are constrained to the master
segment’s rotational motion in a way that is physically justified. To be more specific, slave
node bending rotations (i.e., rotations in the plane of the master segment) are constrained
to the master segment rotational degrees of freedom if this happens to stem from a shell
element, otherwise they are constrained to the master segment rotation as determined from
its individual nodal translations. The slave node torsional rotations (i.e., rotations with
respect to the normal of the master segment) are always constrained according to this latter
philosophy, thus avoiding a torsional constraint to the relatively weak drilling mode of
shells. So this tied contact formulation properly treats bending and torsional rotations, here
slave node rotational degrees of freedom typically come from shell or beam elements. So in
effect, it is in a sense sufficient to only consider
*CONTACT_TIED_SHELL_EDGE_TO_SURFACE_CONSTRAINED_OFFSET
*CONTACT_TIED_SHELL_EDGE_TO_SURFACE_BEAM_OFFSET
for most situations (choosing between a constraint or penalty formulation) but the other
two
*CONTACT_TIED_NODES_TO_SURFACE_CONSTRAINED_OFFSET
*CONTACT_TIED_NODES_TO_SURFACE_OFFSET
𝑮 𝑬
𝑮0 𝑬0
𝑠 𝒅
𝛼𝜆 Tying
𝑚4
point
𝜆 𝑚3 𝑭
𝑭0
𝑀
𝑝
𝑡>0
𝑚1
𝑡=0
𝑚2
29.10.1 Kinematics
We consider a slave node 𝑠 tied to a master segment 𝑀 with an offset 𝜆, the slave node
projection onto the master segment is denoted 𝑝. Let 𝒙𝑠 denote the slave node coordinate
and
4 𝑖
𝒙𝑝 = ∑ ℎ𝑖 𝒙𝑚 (29.37)
𝑖=1
be the slave node projection on the master segment, where ℎ𝑖 are the constant isoparametric
weights. To each of 𝑠, 𝑝 and 𝑀 we associate orthonormal bases (coordinate systems)
represented by
𝑬 = {𝒆 1 𝒆 2 𝒆 3 } (29.38)
𝑭 = {𝒇1 𝒇2 𝒇3 } (29.39)
𝑮 = {𝒈1 𝒈2 𝒈3 }, (29.40)
respectively. The orthogonal matrix 𝑮 is the master segment coordinate system expressed
𝑖
as a function of the master coordinates 𝒙𝑚 with normal 𝒈3 . At 𝑡 = 0, we initialize𝑬0 = 𝑭0 =
𝑮0 but 𝑬 and 𝑭 then evolves independently based on the nodal rotational velocities 𝝎𝑠 and
4 𝑖
𝝎𝑝 = ∑ ℎ𝑖 𝝎 𝑚 . (29.41)
𝑖=1
In the numerical implementation, if ∆𝜽𝑠 = 𝝎𝑠 ∆𝑡 and ∆𝜽𝑝 = 𝝎𝑝 ∆𝑡 are the incremental
rotations of s and p at a given time step, then the coordinate systems are updated
𝑬𝑛+1 = 𝑹(∆𝜽𝑠 )𝑬𝑛 (29.42)
The formulae for 𝐾𝑓 and 𝐾𝑚 are found in earlier sections on tied interfaces while the
expressions for the matrices 𝑩𝑓 and 𝑩𝑚 are quite involved and omitted for the sake of
clarity, the nodal forces and moments are implemented by using (manual) algorithmic
differentiation.
The method consists of five steps. In the first step, the mass per unit area (mass/area) and
pressure are found at each node on the slave surface. Next, the contact point for each
master node is found, and the slave mass/area and slave pressure at each master node is
interpolated from the slave surface. In the third step, this pressure distribution is applied
to the master surface to update its acceleration. In the fourth step, the normal component
of the acceleration at each node on the master surface is scaled by its z-factor defined as the
27-28 (Contact-Impact Algorithm) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Contact-Impact Algorithm
tied interface
Figure 29.23. Incremental searching may fail on surfaces that are not simply
connected. The new contact algorithm in LS-DYNA avoids incremental searching
for nodal points that are not in contact and all these cases are considered.
mass/area of the master surface at the master node divided by the sum of the mass/area of
the slave surface at the master node. The last step consists of resetting the normal
acceleration and velocity components of all slave nodes to ensure compatibility.
The reasons for eliminating slave node tracking by incremental searching is illustrated in
Figure 29.23 where surfaces are shown which cause the incremental searches to fail. In
LS-DYNA tied interfaces are used extensively in many models creating what appears to the
contact algorithms to be topologically disjoint regions. For robustness, our new algorithms
account for such mesh transitions with only minor cost penalties. With bucket sorting
incremental searches may still be used but for reliability they are used after contact is
achieved. As contact is lost, the bucket sorting for the affected nodal points must resume.
In a direct search of a set of 𝑁 nodes to determine the nearest node, the number of distance
comparisons required is 𝑁 − 1. Since this comparison needs to be made for each node, the
LS-DYNA DEV 10/27/16 (r:8004) 27-29 (Contact-Impact Algorithm)
Contact-Impact Algorithm LS-DYNA Theory Manual
Bucket
1 8 7
X strips
2 6
3 4 5
Y Strips
Figure 29.24. One- and two-dimensional bucket sorting.
total number of comparisons is 𝑁(𝑁 − 1), with each of these comparisons requiring a
distance calculation
12 = (𝑥𝑖 − 𝑥𝑗 )2 + (𝑦𝑖 − 𝑦𝑗 )2 + (𝑧𝑖 − 𝑧𝑗 )2 , (29.56)
that uses eight mathematical operations. The cumulative effect of these mathematical
operations for 𝑁(𝑁 − 1) compares can dominate the solution cost at less than 100 elements.
The idea behind a bucket sort is to perform some grouping of the nodes so that the sort
operation need only calculate the distance of the nodes in the nearest groups. Consider the
partitioning of the one-dimensional domain shown in Figure 29.24. With this partitioning
the nearest node will either reside in the same bucket or in one of the two adjoining
buckets. The number of distance calculations is now given by
3𝑁
− 1, (29.57)
𝑎
where 𝑎 is the number of buckets. The total number of distance comparisons for the entire
one-dimensional surface is
3𝑁
𝑁( − 1). (29.58)
𝑎
Thus, if the number of buckets is greater than 3, then the bucket sort will require fewer
distance comparisons than a direct sort. It is easy to show that the corresponding number
of distance comparisons for two-dimensional and three-dimensional bucket sorts are given
by
9𝑁
𝑁( − 1) for 2D (29.59)
𝑎𝑏
27𝑁
𝑁( − 1) for 3D (29.60)
𝑎𝑏𝑐
where 𝑏 and 𝑐 are the number of partitions along the additional dimension.
The cost of the grouping operations, needed to form the buckets, is nearly linear with the
number of nodes 𝑁. For typical LS-DYNA applications, the bucket sort is 100 to 1000 times
faster than the corresponding direct sort. However, the sort is still an expensive part of the
contact algorithm, so that, to further minimize this cost, the sort is performed every ten or
fifteen cycles and the nearest three nodes are stored. Typically, three to five percent of the
calculational costs will be absorbed in the bucket sorting when most surface segments are
included in the contact definition.
considerably since the number of buckets will be reduced. In older versions of DYNA3D
this led to the error termination message “More than 1000 nodes in bucket.”
The formulas given by Belytschko and Lin [1985] are used to find the bucket containing a
node with coordinates (𝑥, 𝑦, 𝑧). The bucket pointers are given by
(𝑥 − 𝑥min )
𝑃𝑋 = 𝑁𝑋 ⋅ + 1, (29.65)
(xmax − xmin )
(𝑦 − 𝑦min )
PY = NY ⋅ + 1, (29.66)
(𝑦max − 𝑦min )
(𝑧 − 𝑧min )
PZ = NZ ⋅ + 1, (29.67)
(𝑧max − 𝑧min )
and are used to compute the bucket number given by
NB = PX + (PY − 1) ⋅ PX + (PZ − 1) ⋅ PX ⋅ PY. (29.68)
For each nodal point, 𝑘, in the contact surface we locate the three nearest neighboring nodes
by searching all nodes in buckets from
MAX(1, PX1), MIN(NX, PX + 1), (29.69)
A maximum of twenty-seven buckets are searched. Nodes that share a contact segment
with k are not considered in this nodal search. By storing the three nearest nodes and
rechecking these stored nodes every cycle to see if the nearest node has changed, we avoid
performing the bucket sorting every cycle. Typically, sorting every five to fifteen cycles is
adequate. Implicit in this approach is the assumption that a node will contact just one
surface. For this reason the single surface contact (TYPE 4 in LS-DYNA) is not applicable to
all problems. For example, in metal forming applications both surfaces of the workpiece
are often in contact.
The nearest contact segment to a given node, 𝑘, is defined to be the first segment
encountered when moving in a direction normal to the surface away from 𝑘. A major
deficiency with the nearest node search is depicted in Figure 29.25 where the nearest nodes
are not even members of the nearest contact segment. Obviously, this would not be a
problem for a more uniform mesh. To overcome this problem we have adopted segment
based searching in both surface to surface and single surface contact.
29.12.2 Bucket Sorting in Surface to Surface and TYPE 13 Single Surface Contact
The procedure is roughly the same as before except we no longer base the bucket size on
𝐿𝑀𝐴𝑋 which can result in as few as one bucket being generated. Rather, the product of the
number of buckets in each direction always approaches 𝑁𝑆𝑁 or 5000 whichever is smaller,
NX ⋅ NY ⋅ NZ ≤ MIN(NSN, 5000), (29.72)
where the coordinate pairs (𝑥min , 𝑥max ), (𝑦min , 𝑦max ), and (𝑧min , 𝑧max ) span the entire
contact surface. In the new procedure we loop over the segments rather than the nodal
points. For each segment we use a nested DO LOOP to loop through a subset of buckets
from IMIN to IMAX, JMIN to JMAX, and to KMAX where
IMIN = MIN(PX1, PX2, PX3, PX4), (29.73)
We check the orthogonal distance of all nodes in the bucket subset from the segment. As
each segment is processed, the minimum distance to a segment is determined for every
node in the surface and the two nearest segments are stored. Therefore the required
storage allocation is still deterministic. This would not be the case if we stored for each
segment a list of nodes that could possibly contact the segment.
We have now determined for each node, 𝑘, in the contact surface the two nearest segments
for contact. Having located these segments we permanently store the node on these
segments which is nearest to node 𝑘. When checking for interpenetrating nodes we check
the segments surrounding the node including the nearest segment since during the steps
1 2 3 4 5
Normal vector at
node 3
Figure 29.25. Nodes 2 and 4 share segments with node 3 and therefore the two
nearest nodes are1 and 5. The nearest contact segment is not considered since its
nodes are not members of the nearest node set.
Figure 29.26. The orthogonal distance of each slave node contained in the box
from the segment is determined. The box is subdivided into sixty buckets.
between bucket searches it is likely that the nearest segment may change. It is possible to
bypass nodes that are already in contact and save some computer time; however, if
multiple contacts per node are admissible then bypassing the search may lead to
unacceptable errors.
◦ Compute the normal segment vectors and accumulate an area weighted aver-
age at the nodal points to determine the normal vectors at the nodal points.
◦ Check all nearest nodes, stored from the bucket sort, and locate the node
which is nearest.
End of Loop
Of course, several obvious limitations of the above procedure exists. The normal vectors
that are used to project the contact surface are meaningless for nodes along an intersection
of two or more shell surfaces (Please see the sketch at the bottom of Figure 29.27). In this
case the normal vector will be arbitrarily skewed depending on the choice of the
numbering of the connectivities of the shells in the intersecting surfaces. Secondly, by
considering the possibility of just one contact segment per node, metal forming problems
cannot be handled within one contact definition. For example, if a workpiece is
constrained between a die and a blankholder then at least some nodal points in the
workpiece must necessarily be in contact with two segments-one in the die and the other in
the workpiece. These two important limitations have motivated the development of the
new bucket sorting procedure described above and the modified single surface contact
procedure, type 13.
A major change in type 13 contact from type 4 is the elimination of the normal nodal vector
projection by using the segment normal vector as shown in Figure 29.27.
Segment numbering within the contact surface is arbitrary when the segment normal is
used greatly simplifying the model input generation. However, additional complexity is
introduced since special handling of the nodal points is required at segment intersections
where nodes may approach undetected as depicted in Figure 29.28a.
To overcome this limitation an additional logic that put cylindrical cap at segment
intersections has been introduced in contact type 13 (and a3). See Figure 29.28b.
Assuming the segment based bucket sort has been completed and closest segments are
known for all slave nodes then the procedure for processing the type 13 contact simplifies
to:
Type 4
v
◦ If node is in contact, check to see if the contact segment has changed and if so,
then update the closest segment information and the orientation flag which
remembers the side in contact. Since no segment orientation information is
stored this flag may change as the node moves from segment to segment.
◦ Check the closest segment to see if the node is in contact if not then proceed to
the end of the loop. If the slave node or contact segment connectivity is a
member of a shell element, project both the node and the contact segment
along the segment normal vector to account for the shell thickness. A nodal
thickness is stored for each node and a segment thickness is stored for each
segment. A zero thickness is stored for solid elements. The thickness can be
optionally updated to account for membrane thinning.
◦ Check for interpenetrating nodes and if a node has penetrated apply a nodal
point force that is proportional to the penetration depth.
End of Loop
Note that type 13 contact does not require the calculation of nodal normal vectors.
a b
Figure 29.28.
In this constraint approach the accelerations, velocities, and displacements are first updated
to a trial configuration without accounting for interface interactions. After the update, a
penetration force is computed for the slave node as a function of the penetration distance
Δ𝐿:
𝑚𝑠 Δ𝐿
𝐟𝑝 =
𝐧, (29.78)
Δ𝑡2
where 𝐧 is the normal vector to the master surface.
We desire that the response of the normal component of the slave node acceleration vector,
𝐚s , of a slave node residing on master segment 𝑘 be consistent with the motion of the
master segment at its contact segment (𝑠c , 𝑡c ), i.e.,
as = 𝜙1 (𝑠c , 𝑡c )𝑎1𝑛𝑘 + 𝜙2 (𝑠c , 𝑡c )𝑎2𝑛𝑘 + 𝜙3 (𝑠c , 𝑡c )𝑎3𝑛𝑘 + 𝜙4 (𝑠c , 𝑡c )𝑎4𝑛𝑘 . (29.79)
For each slave node in contact with and penetrating through the master surface in its trial
configuration, its nodal mass and its penetration force given by Equation (29.72) is
accumulated to a global master surface mass and force vector:
where
𝑚𝑘𝑠 = 𝜙𝑘 𝑚𝑠 , (29.81)
𝐟𝑘𝑠 = 𝜙𝑘 𝐟𝑠 . (29.82)
After solving Equation (29.78) for the acceleration vector, 𝐚nk , we can obtain the
acceleration correction for the slave node as
𝐟p
𝐚ns = 𝐚s − . (29.83)
𝑚s
The above process is repeated after reversing the master and slave definitions. In the final
step the averaged final correction to the acceleration vector is found
1 1st pass 1 2nd pass
𝐚𝑛final = (1 − 𝛽)𝐚𝑛 + (1 + 𝛽)𝐚𝑛 , (29.84)
2 2
and used to compute the final acceleration at time 𝑛 + 1
𝐚𝑛+1 = 𝐚trial + 𝐚𝑛final , (29.85)
A trial tangential force is computed that will cancel the tangential velocity
𝑚𝜐
𝐟∗ = 𝑠 𝑡 , (29.88)
Δ𝑡
where υt is the magnitude of the tangential velocity vector
𝜐𝑡 = √𝐯𝑡 ⋅ 𝐯𝑡 . (29.89)
The magnitude of the tangential force is limited by the magnitude of the product of the
Coulomb friction constant with the normal force defined as
fn = ms 𝐚ns ⋅ 𝐧, (29.90)
𝐹𝑦 𝐟 ∗
𝐟𝑛+1 = |𝐟∗ |
if ∣𝐟∗ ∣ > 𝐹𝑦 . (29.93)
Therefore, using the above equations the modification to the tangential acceleration
component of the slave node is given by
∣𝐯s ∣
𝐚t = min (𝜇𝐚nt ⋅ 𝐧, ), (29.94)
Δ𝑡
which must act in the direction of the tangential vector defined as
𝐯
𝐧t = t . (29.95)
υt
The corrections to both the slave and master node acceleration components are:
ats = at 𝐧t , (29.96)
as ms
𝐚tk = −𝜙k 𝐧, (29.97)
mk t
The above process is again repeated after reversing the master and slave definitions. In the
final step the averaged final correction to the acceleration vector is found
1 1st pass 1 2nd pass
𝐚tfinal = (1 − 𝛽)𝐚t + (1 + 𝛽)𝐚t , (29.98)
2 2
and is used to compute the final acceleration at time 𝑛 + 1
𝐚𝑛+1 = 𝐚trial + 𝐚𝑛final + 𝐚tfinal . (29.99)
A significant disadvantage of the constraint method relative to the penalty method appears
if an interface node is subjected to additional constraints such as spot welds, constraint
equations, tied interfaces, and rigid bodies. Rigid bodies can often be used with this
contact algorithm if their motions are prescribed as is the case in metal forming. For the
more general cases involving rigid bodies, the above equations are not directly applicable
since the local nodal masses of rigid body nodes are usually meaningless. Subjecting the
two sides of a shell surface to this constraint algorithm will also lead to erroneous results
since an interface node cannot be constrained to move simultaneously on two mutually
independent surfaces. In the latter case the constraint technique could be used on one side
and the penalty method on the other.
The biggest advantage of the constraint algorithm is that interface nodes remain on or very
close to the surfaces they are in contact with. Furthermore, elastic vibrations that can occur
in penalty formulations are insignificant with the constraint technique. The problem
related to finding good penalty constants for the contact are totally avoided by the latter
approach. Having both methods available is possibly the best option of all.
n
l
n
V
n v
m
l
L
L
R
n
v
l
flat surface sphere
Figure 29.30. Vector n determines the orientation of the generalized stonewalls.
For the prescribed motion options the wall can be moved in the direction V as
shown.
velocity at the point of contact. Contact can occur with any of the surfaces which enclose
the volume. Currently four geometric shapes are available including the rectangular
prism, the cylinder, flat surface, and sphere. These are shown in Figure 29.30.
First, the user must specify which VDA/IGES surfaces, faces, and groups should be
attached to each material. This is done primarily through a special input file. Faces,
surfaces, and groups from several different VDA/IGES input files can be combined into
groups that later can be refered to by a user specified alias. For example, suppose a simple
sheetmetal forming problem is going to be run. The user might have an input file that
looks like this:
In this example, the user has specified that the punch will be made up of the group
"grp001" from the file "punch.vda". The VDA file is converted to a binary file "punch.bin".
If this simulation is ever rerun, the VDA input can be read directly from the binary file
thereby significantly reducing startup time. The die in this example is made up of several
surfaces and faces from 2 different VDA files. This format of input allows the user to
combine any number of faces, surfaces, and groups from any number of VDA files to
define a single part. This single part name is then referenced within the LS-DYNA input
file.
The contact algorithm works as follows. For the sake of simplicity, we will refer to one
point as being slaved to a single part. Again, this part will in general be made up of several
VDA surfaces and faces. First, the distance from the point to each VDA surface is
computed and stored. For that surface which is nearest the point, several other parameters
are stored such as the surface coordinates of the near point on the surface. Each time step
of the calculation this information is updated. For the nearest surface the new near point is
LS-DYNA DEV 10/27/16 (r:8004) 27-43 (Contact-Impact Algorithm)
Contact-Impact Algorithm LS-DYNA Theory Manual
slave point
P (x, y, z)
s
Figure 29.31. The geometry of the patch is a function of the parametric
coordinates 𝑠 and 𝑡.
calculated. For all other surfaces the distance the point moves is subtracted from the
distance to the surface. This continually gives a lower bound on the actual distance to each
VDA surface. When this lower bound drops below the thickness of the point being
tracked, the actual distance to the surface is recalculated. Actually, if the nearest surface is
further away from the point than some distance, the near point on the surface is not tracked
at all until the point comes close to some surface. These precautions result in the distance
from the point to a surface having to be totally recomputed every few hundred timesteps,
in exchange for not having to continually track the point on each surface.
To track the point on the nearest surface, a 2D form of Newton's method is used. The
vector function to be solved specifies that the displacement vector from the surface to the
point should be parallel to the surface normal vector. The surface tangent vectors are
computed with respect to each of the two surface patch parameters, and the dot product
taken with the displacement vector. See Figure 29.31 and Equation (29.104).
∂𝐪 ∂𝐪
(𝐩 − 𝐪) ⋅ = 0 and (𝐩 − 𝐪) ⋅ = 0. (29.105)
∂s ∂t
This vector equation is then solved using Newton's method as in Equation (1.106).
𝐪𝑖+1 = 𝐪𝑖 − (𝐅′ )−1 𝐅, (29.106)
where
∂𝐪
⎛(𝐩
⎜ − 𝐪) ⋅ ⎞
𝐅(𝑠, 𝑡) = ⎜
⎜ ∂𝑠 ⎟
⎟
⎟. (29.107)
⎜
⎜ ∂𝐪⎟⎟
⎝ (𝐩 − 𝐪) ⋅
∂𝑡 ⎠
27-44 (Contact-Impact Algorithm) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Contact-Impact Algorithm
previous location
new location
previous nearest
point
Figure 29.32. Newton iteration solves for the nearest point on the analytical
surface.
The convergence is damped in the sense that the surface point is not allowed to jump
completely outside of a surface patch in one iteration. If the iteration point tries to leave a
patch, it is placed in the neighboring patch, but on the adjoining boundary. This prevents
the point from moving merely continuous (i.e., when the surface has a crease in it).
Iteration continues until the maximum number of allowed iterations is reached, or a
convergence tolerance is met. The convergence tolerance (as measured in the surface patch
parameters) varies from patch to patch, and is based on the size and shape of the patch.
The convergence criterion is set for a patch to ensure that the actual surface point has
converged (in the spatial parameters x, y, and z) to some tolerance.
3 4
1
points 1, 2, 3, and 4 define drawbeads
Figure 29.33. The drawbead contact provides a simple way of including
drawbead behavior without the necessity of defining a finite element mesh for the
drawbeads. Since the draw bead is straight, each bead is defined by only two
nodes.
shell elements in the workpiece. A three-dimensional bucket search is used for the contact
searching to locate each point within a segment of the master surface.
The nodes defining the draw beads can be attached to rigid bodies by using the extra nodes
for rigid body input option. When defining draw beads, care should be taken to limit the
number of elements that are used in the master surface definition. If the entire blank is
specified the CPU cost increases significantly and the memory requirements can become
enormous. An automated draw bead box, which is defined by specifying the part ID for
the workpiece and the node set ID for the draw bead, is available. The automated box
option allows LS-DYNA determine the box dimensions. The size of this box is based on the
extent of the blank and the largest element in the workpiece as shown if Figure 29.34.
The input for the draw beads requires a load curve giving the force due to the bending and
unbending of the blank as it moves through the draw bead. The load curve may also
include the effect of friction. However, the coulomb friction coefficients must be set to zero
if the frictional component is included in the load curve. If the sign of the load curve ID is
positive the load curve gives the retaining force per unit draw bead length as a function of
displacement, δ. If the sign is negative the load curve defines the maximum retaining force
versus the normalized position along the draw bead. This position varies from 0 (at the
origin) to 1 (at the end) along the draw bead. See Figures 29.35 and 29.36.
Figure 29.34. The draw bead box option automatically size the box around the
draw bead. Any segments within the box are included as master segments in the
contact definition.
When friction is active the frictional force component normal to the bead in the plane of the
work piece is computed. Frictional forces tangent to the bead are not allowed. The second
load curve gives the normal force per unit draw bead length as a function of displacement,
δ. This force is due to bending the blank into the draw bead as the binder closes on the die
and represents a limiting value. The normal force begins to develop when the distance
between the die and binder is less than the draw bead depth. As the binder and die close
on the blank this force should diminish or reach a plateau. This load curve was originally
added to stabilize the calculation.
As the elements of the blank move under the draw bead, a plastic strain distribution
develops through the shell thickness due to membrane stretching and bending. To account
for this strain profile an optional load curve can be defined that gives the plastic strain
versus the parametric coordinate through the shell thickness where the parametric
coordinate is defined in the interval from –1 to 1. The value of the plastic strain at each
through thickness integration point is interpolated from this curve. If the plastic strain at
an integration point exceeds the value of the load curve at the time initialization occurs, the
plastic strain at the point will remain unchanged. A scale factor that multiplies the shell
thickness as the shell element moves under the draw bead can also be defined as a way of
accounting for any thinning that may occur.
Force
positive load curve ID
Penetration distande, δ
b
Force
0 1
Normalized draw bead length
Figure 29.35. Draw bead contact model defines a resisting force as a function of
draw bead displacement.
δ F = Ffriction +Fbending
Figure 29.36. Draw bead contact model defines a resisting force as a function of
draw bead displacement.
The basis of single edge contact is the proven single surface formulation and the input is
identical. The definition is by material ID. Edge determination is automatic. It is also
possible to use a manual definition by listing line segments. The single edge contact is type
22 in the structured input or *CONTACT_SINGLE_EDGE in the keyword input.
Figure 29.37. Contact between edges requires a special treatment since the nodes
do not make contact.
This contact only considers edge to edge contact of the type illustrated in Figure 29.38.
Here the tangent vectors to the plane of the shell and normal to the edge must point to each
other for contact to be considered.
tangent vectors in
plane of shell
Figure 29.38. Single edge contact considers contact between two edges whose
normals point towards each other.
of the area of the smallest rectangle that contains the cross section to avoid tracking the
27-50 (Contact-Impact Algorithm) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Contact-Impact Algorithm
orientation of the beam within the contact algorithm. Contact is found by finding the
intersection point between nearby beam elements and checking to see if their outer surfaces
overlap as seen in Figure 29.40. If the surfaces overlap the contact force is computed and is
applied to the nodal points of the interacting beam elements.
Contact surface
29.21.1 Kinematics
The Mortar contact is theoretically treated as a generalized finite element where each
ns
4
Slave segment 3
Ts
1 Π
X3 2
X2
O Master Segment
X1
Figure 29.41. Illustration of Mortar segment to segment contact
element in this context consists of a pair of contact segments. The friction model in the
Mortar contact is a standard Coulomb friction law. Each of the two segments has its iso-
27-52 (Contact-Impact Algorithm) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Contact-Impact Algorithm
parametric representation inherited from the underlying finite element formulation, so the
coordinates for the slave and master segments can be written
𝐱s = 𝑁s𝑖 (𝜉 , 𝜂)𝐱𝑖
𝑗
(29.21.108)
𝐱m = 𝑁m (𝜉 , 𝜂)𝐱𝑗 ,
where summation over repeated indices is implicitly understood, i.e., over the nodes. The
kinematics for the contact element can be written as the penetration
𝑑 = 𝐧sT (𝐱s − 𝐱̅m ), (29.21.109)
where 𝐧s is the slave segment normal and 𝐱̅m is the projected point on the master segment
along the slave segment normal. The element is only defined for the intersection between
the slave and master segment and for points where 𝑑 > 0, this domain is denoted Π and is
illustrated by gray in the Figure above. The sliding rate 𝐬̇ is similarly defined as
𝐬̇ = 𝐓sT (𝐱̇s − 𝐱̅m
̇ ), (29.21.110)
where 𝐓s are two co-rotational basis vectors pertaining to the slave segment.
and
⎧ 1 2 𝑑max
{ 𝑥 𝑥<
{ 4 2𝜀𝑑𝑐𝑠
𝑓 (𝑥) = ⎨ . (29.21.112)
{cubic function that depends on IGAP 𝑑max
{ ≤𝑥
⎩ 2𝜀𝑑𝑐𝑠
where 𝑑max is the maximum penetration to be given below. The Coulomb friction law is
expressed in terms of the tangential contact stress
𝐬 |𝐬|
𝛔t = 𝜇𝜎n 𝑔 ( ), (29.21.113)
|𝐬| 𝜇𝑑
where 𝜇 is the friction coefficient and
⎧ 𝑥 𝑥 ≤1−𝜀
{
{ 2
𝑔(𝑥) = ⎨1 − 𝜀 (1 + 𝜀 − 𝑥) 1 − 𝜀 < 𝑥 ≤ 1 + 𝜀. (29.21.114)
{
{ 4 𝜀
⎩ 1 1+𝜀<𝑥
The update of 𝐬 is done incrementally and is at the end of the step modified so that
|𝐬| ≤ 𝜇𝑑(1 + 𝜀) (29.21.115)
after the contact update.
and in this case the edge is smoothed by adding 4 segments between the two nodes
common to the two contact segments, while at the same time adjusting the size of the two
main contact segments. The effect is a rounded representation of the edge and a smoother
contact response, see Figure 27-42. The size of the smoothing is at most 5% of the size of
the smallest of the two contact segments, so the effect is reduced with mesh refinement.
Since the contact area of these edge segments may be small, the stiffness of the contact is for
those scaled by the factor 𝛽 = √3cot(𝜋−𝜃) where 𝜃 is the initial angle between the main
2
segment normals. This scale factor is applied for both the slave and master side, meaning
that if two right angled edges come into contact the stiffness is scaled by 𝛽𝑠 𝛽𝑚 =
𝜋−𝜋 𝜋−𝜋
√3 cot ( 2
) √3 cot ( 2
) = 9. Whether this is sufficient to handle most common
2 2
situations is currently unknown, so this design decision is subject for change in the future.
If a contact segment has two sharp edges with a common node, then a segment is created at
the location of that node to account for contact with the corner of the solid element
geometry. The stiffness scale factor for a corner node is the average of the scale factor for
the connected edges. The motivation behind the solid edge smoothing is two-fold; for one
thing it adds the feature of resisting penetration between a solid edge/corner and other
geometries and then it also aids establishing a physical contact state when solid elements
slide off sharp geometrical objects. More specifically it eliminates the ambiguity of which
contact segments are in contact with which, and presumably prevents sudden spikes in the
contact force, this is also illustrated in Figure 27-42.
It is important to stress that the automatic Mortar contact surfaces are always located on the
outer geometry for both slave and master sides, i.e., contact does not occur on the mid-
surface of shells. A common modelling problem is illustrated in Figure 27-43 that results in
extremely large contact stress unless the ignore flag is appropriately used. The forming
contact is treated differently not only in that shell edge or beam contact is not supported.
Here any shell master surface must be rigid with its segment normals oriented towards the
slave side of the contact. Furthermore the contact occurs here on the mid-surface in
contrast to the automatic option, while contact on the slave side still occurs on the outer
geometry. The segment orientation of the (deformable) slave side is on one hand arbitrary,
but if it consists of shell elements contact can only occur on one side for a given contact
definition. In a forming application for instance, the contact between the tool and blank
and the contact between the die and blank have to be defined using two different contact
interfaces since these contacts typically occur on different sides of the blank. Forming solid
master surfaces may be rigid or deformable, thus allowing for effects of tool deformation
and/or cooling effects.
(29.117)
where 𝑑max is the maximum penetration, 𝑑𝑐𝑠 is the characteristic length of the slave segment
and 𝑑𝑐𝑚 is the characteristic length of the master segment. This is illustrated in Figure 29.44
that shows the contact stress as function of the relative penetration. To minimize the risk of
releasing contacts one could increase the stiffness parameter SFS, but this may lead to
worse convergence in implicit analysis. To this end, the IGAP parameter can be used to
stiffen the contact for large penetrations without affecting moderate penetrations, this is
also illustrated in Figure 29.44 and can be used if the contact pressure is locally very high.
This also highlights the following important fact:
The Mortar contact has no “stick” option for improving implicit convergence.
4
IGAP=1
3.5 IGAP=2
IGAP=5
3 IGAP=10
Contact Stress
2.5
1.5
0.5
0
0 0.2 0.4 0.6 0.8 1
Penetration
This may on one hand lead to worse convergence characteristics but on the other rules out
sticky behavior and underreport of contact forces in the ascii database.
The characteristic length is for shells the shell thickness, whereas for solids it is the
smallest element size in the part that the segment belongs to. The latter may lead to
unrealistically high or low contact stiffness, or it may result in a too small maximum
penetration depth, all depending on the mesh. For a mesh independent behavior the user
may set SST to the characteristic length, which should correspond to some physical
member size in the model. A high value will provide a soft contact, and vice versa. It is
recommended to
Separate shells and solids into different contact definitions
Explicitly set SST to a characteristic length for the solid definitions
Element thickness T
On the other side of the spectrum, poor convergence could be due to a too stiff
contact. Since the stiffness for a Mortar contact segment pair only depends on the slave
segment it is recommended to
Put weak parts on the slave side
If a steel part is in contact with rubber for instance, the rubber part should be put as slave
side in the contact definition. One way of find contacts that cause poor convergence in
implicit is to turn on D3ITCTL on CONTROL_IMPLICIT_SOLUTION and RESPLT on
DATABASE_EXTENT_BINARY, which allows the user to isolate areas in the model where
convergence is poor. Not rarely this is due to contacts.
IGNORE < 0 See explanation for the corresponding positive value, the only
difference is that contact between segments belonging to the same part
is not treated
IGNORE = 0 Initial penetrations will give rise to initial contact stresses, i.e., the slave
contact surface is not modified
IGNORE = 1 Initial penetrations will be tracked, i.e., the slave contact surface is
translated to the level of the initial penetrations and subsequently
follow the master contact surface on separation until the unmodified
level is reached
IGNORE = 2 Initial penetrations will be ignored, i.e., the slave contact surface is
translated to the level of the initial penetrations, optionally with an
initial contact stress governed by MPAR1
IGNORE = 3 Initial penetrations will be removed over time, i.e., the slave contact
surface is translated to the level of the initial penetrations and pushed
back to its unmodified level over a time determined by MPAR1
IGNORE = 4 Same as IGNORE = 3 but it allows for large penetrations by also setting
MPAR2 to at least the maximum initial penetration
The use of IGNORE depends on the problem, if no initial penetrations are present
there is no need to use this parameter at all. If penetrations are relatively small in relation
to the maximum allowed penetration, then IGNORE = 1 or IGNORE = 2 seems to be the
appropriate choice. For IGNORE = 2 the user may specify an initial contact stress small
enough to not significantly affect the physics but large enough to eliminate rigid body
modes and thus singularities in the stiffness matrix. The intention with this is to constrain
loose parts that are initially close but not in contact by pushing out the contact surface
using SFST and applying the IGNORE = 2 option. Increasing SFST to a number larger than
unity will push the contact surface outside the geometry and contact will be detected
accordingly, see Figure. It is at least good for debugging problems with many singular
rigid body modes.
A drawback with IGNORE = 3 is that initial penetration must be smaller than half
the characteristic length of the contact or otherwise they will not be detected in the first
place. For this reason IGNORE = 4 was introduced where initial penetrations may be of
arbitrary size, but it requires that the user provides crude information on the level of
penetration of the contact interface. This is done in MPAR2 which must be larger than the
maximum penetration or otherwise an error termination will occur. IGNORE = 4 only
applies to solid elements at the moment.
𝑠1 𝑠2
𝑚2
𝑚1
𝑑(𝑡) 𝐴
𝑠1 vs 𝑚2 𝒙(𝑡) 𝒏
𝒏𝑠 𝒕
𝒙2
𝒙1 𝒚2
𝒚1
𝒏𝑚
𝒚(𝑡)
𝑠1 vs 𝑚1
𝑠2 vs 𝑚2
Kinematics
First a common tangential direction is determined, this is given as
𝒙 + 𝒚2 − 𝒙1 − 𝒚1
𝒕= 2
∥𝒙2 + 𝒚2 − 𝒙1 − 𝒚1 ∥
(29.118)
)
(29.119)
The overlapped interval is then further reduced to 𝑡 ∈ [𝑡1 , 𝑡2 ] for which 𝑑(𝑡) ≥ 0, note that
for the 𝑠2 vs 𝑚2 situation these two intervals are not the same. The interval [𝑡1 , 𝑡2 ] is
indicated by red in the illustration. This completes the kinematics in the normal direction.
In the tangential direction we need to define the kinematics for sliding and we do this by
associating a history variable to each slave segment. To this end, 𝑆 as a weighted measure
of the distance a slave segment has slid along the master surface and is defined as
𝑇
𝑆 = 𝑆𝑛−1 + ∑ ∫ {𝑠(𝑡) − 𝒕𝑛−1 (𝒙𝑛−1 (𝑡) − 𝒚𝑛−1 (𝑡))}𝑑𝐴𝑖
𝑖 𝐴𝑖
(29.120)
)
where 𝑠(𝑡) = 𝒕 𝑇 (𝒙(𝑡) − 𝒚(𝑡)) and the subscript 𝑛 − 1 refers to the corresponding value in
the previous step. The integral is taken over the domain 𝐴𝑖 of the intersected interval
between the slave segment and master segment 𝑖, accounting for plane strain or axial
symmetry. Note that the slave segment can be in contact with several master segments,
whence the sum. Further noting that by construction the first term in the integrand, 𝑠(𝑡) =
0, we can simplify this to
𝑇
𝑆 = 𝑆𝑛−1 − ∑ ∫ 𝒕𝑛−1 (𝒙𝑛−1 (𝑡) − 𝒚𝑛−1 (𝑡))𝑑𝐴𝑖 .
𝑖 𝐴𝑖
(29.121)
)
In the illustrated situation, slave segment 1 would get sliding contributions from master
segments 1 and 2, while slave segment 2 only from master segment 2. Likewise we define a
weighted penetration
𝐷 = ∑ ∫ 𝑑(𝑡)𝑑𝐴𝑖 .
𝑖 𝐴𝑖
(29.122)
)
Constitutive relations
For simplicity, we drop the explicit dependence on 𝑡 from now on and define the contact
stress (pressure) as
𝜎𝑛 = 𝐾 𝑑2
(29.123)
(29.124)
Furthermore, 𝛼 = PSF ∗ SLSFAC is a stiffness scale factor and 𝑇𝑠/𝑚 and 𝐾𝑠/𝑚 are
characteristic lengths and material stiffnesses of the slave and master segments,
respectively. The function 𝑓 is used to linearly reduce stiffness for segments that are not
parallel, note that 𝒏𝑠 and 𝒏𝑚 are the normals to the slave and master segments, and is given
as
⎧ 1
{ 0 − ≤𝑥
{ 2
{
{ 1 + 2𝑥 √ 3 1
{
𝑓 (𝑥) = ⎨ − ≤𝑥<− .
{1 − √3 2 2
{
{ √3
{
{ 1 𝑥 < − (29.125)
⎩ 2
(29.126)
with 𝜇 being the Coulomb friction coefficient and 𝑔 is a continuously differentiable function
defined as
⎧ 1, 𝑥 ≥ 1.03
{
{ 25
{
{ 1 − (𝑥 − 1.03)2 , 0.97 ≥ 𝑥 > 1.03
{ 3 (29.127)
𝑔(𝑥) = ⎨ 𝑥, −0.97 ≥ 𝑥 > 0.97
{ 25 )
{
{ −1 + (𝑥 + 1.03)2 , −1.03 ≥ 𝑥 > −0.97
{
{ 3
⎩ −1, −1.03 > 𝑥
The interpretation of this law is that the magnitude of friction stress 𝜎𝑡 is at most 𝜇𝜎𝑛 , and
the fraction thereof is determined by the appropriate relation between the accumulated
sliding and penetration. Upon convergence, to yield a proper friction behavior, 𝑆 is
updated for the next step according to 𝑆𝑛 = max(−1.03𝜇𝐷, min(1.03𝜇𝐷, 𝑆)).
Nodal forces
The nodal force contribution from a given segment pair is determined from the principle of
virtual work, i.e.,
𝛿𝑊 = ∫ 𝜎𝑛 𝛿𝑑 𝑑𝐴 + ∫ 𝜎𝑡 𝛿𝑠 𝑑𝐴.
𝐴 𝐴
(29.128)
Here 𝛿 is the variation operator, and the independent variables subject to this variation are
the nodal coordinates. Thus, replacing the left hand side with the expression involving
nodal forces and coordinates, and using (29.123) and (29.126) for the right hand side we get
∑ 𝒇𝐼𝑇 𝛿𝒙𝐼 = 𝐾 ∫ 𝑑2 𝛿𝑑 𝑑𝐴 + 𝜇𝐾𝑔 ∫ 𝑑2 𝛿𝑠 𝑑𝐴.
𝐼 𝐴 𝐴
(29.129)
)
where the exact expressions for the integrals in terms the nodal coordinates are, although
straightforward to derive, a bit lengthy and therefore omitted.
Stiffness matrix
The stiffness matrix is the second variation of the virtual work expression (29.128) where
we neglect the variation of the overlapped area and assume the variation can be taken on
the integrand directly
𝜕𝑔 1 𝑆 (29.131)
∆𝛿𝑊𝐼 = 2𝐾 ∫𝐴 𝑑 ∆𝑑 𝛿𝑑 𝑑𝐴 + 2𝜇𝐾𝑔 ∫𝐴 𝑑 ∆𝑑 𝛿𝑠 𝑑𝐴 + 𝜇 𝜕𝑥 {𝜇𝐷 ∆𝑆 − 𝜇𝐷 2
∆𝐷} ∫𝐴 𝜎𝑛 𝛿𝑠 𝑑𝐴
)
Here ∆ is again the variation operator, different notation to distinguish it from 𝛿 but
performing the exact same thing. At this point we emphasize that the (geometric) terms
involving ∆𝛿𝑑 and ∆𝛿𝑠 have been deliberately excluded as they, albeit being symmetric,
contribute to the indefiniteness of the tangent matrix. In (29.131), the first term on the right
hand side is the normal-normal interaction and is nicely symmetric, the remaining terms
come from friction (normal-tangent and tangent-tangent interaction) and needs
symmetrization and further simplification. To this end we neglect any terms involving
∆𝑑 𝛿𝑠 and ∆𝐷𝛿𝑠 which means that it remains to deal with the term involving ∆𝑆 𝛿𝑠. The
simplifications made are to approximate
1 (29.132)
𝜎𝑛 ≈ ∫ 𝜎𝑛 𝑑𝐴
𝐴 𝐴 )
in the last integral, and then neglect all terms in ∆𝑆 that do not pertain to the present slave
master segment pair. This results in
𝜕𝑔 1 1 (29.133)
∆𝒙𝐼𝑇 𝑲𝐼𝐽𝑇 𝛿𝒙𝐽 ≈ 2𝐾 ∫ 𝑑 ∆𝑑 𝛿𝑑 𝑑𝐴 + { ∫ 𝜎𝑛 𝑑𝐴} {∫ ∆𝑠 𝑑𝐴} {∫ 𝛿𝑠 𝑑𝐴}
𝐴 𝜕𝑥 𝐷 𝐴 𝐴 𝐴 𝐴 )
and the stiffness matrix can be identified as
𝜕𝑑 𝜕𝑑 𝜕𝑔 1 1 𝜕𝑠 𝜕𝑠 (29.134)
𝑲𝐼𝐽 = 2𝐾 ∫ 𝑑 𝑑𝐴 + { ∫ 𝜎𝑛 𝑑𝐴} {∫ 𝑑𝐴} {∫ 𝑑𝐴}.
𝐴 𝜕𝒙 𝐼 𝜕𝒙 𝐽 𝜕𝑥 𝐷 𝐴 𝐴 𝐴 𝜕𝒙 𝐼 𝐴 𝜕𝒙 𝐽 )
A characteristic feature of the Mortar contact, which can be deduced from this expression,
is that not only the nodal forces are continuous but also the stiffness matrix. This follows
from the 𝐶1 continuity of 𝑔, meaning that all involved functions above are continuous, and
that 𝑲𝐼𝐽 tends to zero as either 𝑑 or 𝐴 tends to zero. A mathematical treatment yields
𝜕𝑑(𝑡) 𝜕𝑑(𝑡) 𝜕𝑠(𝑡) 𝜕𝑠(𝑡) (29.135)
∥𝑲𝐼𝐽 ∥ ≤ 𝐾𝐴 max 𝑑(𝑡) {2 max ∥ ∥∥ ∥ + max ∥ ∥∥ ∥}
𝑡 𝑡 𝜕𝒙𝐼 𝜕𝒙𝐽 𝑡 𝜕𝒙𝐼 𝜕𝒙𝐽 )
which tends to zero as 𝑑 or 𝐴 tends to zero as the terms inside the right bracket are
bounded.
LS-DYNA DEV 10/27/16 (r:8004) 27-65 (Contact-Impact Algorithm)
Contact-Impact Algorithm LS-DYNA Theory Manual
the slave segment. Whence, for tied and tiebreak contacts SFST is not used to trans-
late the contact surface as it is for the one-sided Mortar contacts.
.
50000
-50000
-100000
-150000
0 10 20 30 40 50 60
Workpiece Displacement
Figure 27-47 A rubber compression example solved in implicit with Mortar contact
(Courtesy of Dellner Couplers AB). The graph shows the contact force between the rubber
parts and the moving workpiece and between the rubber parts and the two supports,
respectively.
30
Geometric Contact Entities
Contact algorithms in LS-DYNA currently can treat any arbitrarily shaped surface
by representing the surface with a faceted mesh. Occupant modeling can be treated this
way by using fine meshes to represent the head or knees. The generality of the faceted
mesh contact suffers drawbacks when modeling occupants, however, due to storage
requirements, computing costs, and mesh generation times. The geometric contact entities
were added as an alternate method to model cases of curved rigid bodies impacting
deformable surfaces. Much less storage is required and the computational cost decreases
dramatically when compared to the more general contact.
If the node is in contact with the solid, a restoring force must be applied to eliminate
further penetration. A number of methods are available to do this such as Lagrange
multipliers or momentum based methods. The penalty method was selected because it is
the simplest and most efficient method. Also, in our applications the impact velocities are
at a level where the penalty methods provide almost the identical answer as the exact
solution.
Using the penalty method, the restoring force is proportional to the penetration
distance into the solid and acts in the direction normal to the surface of the solid. Thus, the
G(x, y) < 0
R
G(x, y) > 0
G(x, y) = 0
penetration distance and the normal vector must be determined. The surface normal vector
is conveniently determined from the gradient of the scalar function.
∂𝐺 𝐢 + ∂𝐺 𝐣 + 𝜕𝐺 𝐤
∂𝑥 ∂𝑦 𝜕𝑧
𝐍(𝑥, 𝑦, 𝑧) = , (30.4)
2 2 2
√(∂𝐺) + (∂𝐺) + (∂𝐺)
∂𝑥 ∂𝑦 ∂𝑧
for all (𝑥, 𝑦, 𝑧) such that 𝐺(𝑥, 𝑦, 𝑧) = 0. The definition of 𝐺(𝑥, 𝑦, 𝑧) guarantees that this vector
faces in the outward direction. When penetration does occur, the function 𝐺(𝑥, 𝑦, 𝑧) will be
slightly less than zero. For curved surfaces this will result in some errors in calculating the
normal vector, because it is not evaluated exactly at the surface. In an implicit code, this
would be important, however, the explicit time integration scheme in DYNA3D uses such a
small time step that penetrations are negligible and the normal function can be evaluated
directly at the slave node ignoring any penetration.
𝐺(𝑥, 𝑦) = 𝑥2 + 𝑦2 − 𝑅2 , (30.5)
The penetrations distance is the last item to be calculated. In general, the penetration
distance, 𝑑, is determined by.
𝑑 = ∣𝐗𝑛 − 𝐗′𝑛 ∣, (30.6)
where 𝐗𝑛 is the location of node 𝑛 and 𝐗′𝑛 is the nearest point on the surface of the solid.
To determine 𝐗′𝑛 , a line function is defined which passes through 𝐗𝑛 and is normal
to the surface of the solid:
𝐋(𝑠) = 𝐗𝑛 + 𝑠𝐍 (𝐗𝑛 ). (30.7)
Substituting the line function into the definition of the Equation (30.2) surface of a solid
body gives:
𝐺(𝐗𝑛 + 𝑠𝐍(𝐗)) = 0. (30.8)
If Equation (30.8) has only one solution, this provides the parametric coordinates s which
locates 𝐗′𝑛 . If Equation (30.8) has more than one root, then the root which minimizes
Equation (30.6) locates the point 𝐗′𝑛 .
Inclusion of any structural elements into the occupant model will typically result in
very large stiffnesses due to the small time step and the (1/Δ𝑡)2 term. Thus the method is
highly effective even with impact velocities on the order of 1km/sec.
𝑎4 𝑏4 𝑐4
Substituting Equations (30.7) and (30.15) into Equation (27.2) gives:
𝑛 2 𝑛𝑦 2 𝑛 2 𝑛 𝑥 𝑛𝑦 𝑦𝑛 𝑛 𝑧
[( 𝑥 ) + ( ) + ( 𝑧 ) ] 𝑠 2 + 2 [ 𝑥2 𝑛 + 2 + 𝑧2 𝑛 ] 𝑠
𝑎 𝑏 𝑐 𝑎 𝑏 𝑐 (30.17)
𝑥𝑛 2 𝑦𝑛 2 𝑧𝑛 2
+ [( ) + ( ) + ( ) − 1] = 0.
𝑎 𝑏 𝑐
Solving this quadratic equation for 𝑠 provides the intercepts for the nearest point on
the ellipsoid and the opposite point of the ellipsoid where the normal vector, 𝐗𝑛 , also
points toward.
Currently, this method has been implemented for the case of an infinite plane, a
cylinder, a sphere, and an ellipsoid with appropriate simplifications. The ellipsoid is
intended to be used with rigid body dummy models. The methods are, however, quite
general so that many more shapes could be implemented. A direct coupling to solids
modeling packages should also be possible in the future.
31
Nodal Constraints
In this section nodal constraints and linear constraint equations are described.
Nodal constraint sets eliminate rigid body rotations in the body that contains the
node set and, therefore, must be applied very cautiously.
is initialized by linking to an implicit code to satisfy this equation at the beginning of the
calculation, the constant 𝐶0 is assumed to be zero. The first constrained degree of freedom
is eliminated from the equations of motion:
𝑛
𝐶𝑘
𝑢1 = 𝐶0 − ∑ 𝑢 . (31.3)
𝐶 𝑘
𝑘=2 1
32
Vectorization and Parallelization
32.1 Vectorization
In 1978, when the author first vectorized DYNA3D on the CRAY-1, a four-fold
increase in speed was attained. This increase was realized by recoding the solution phase
to process vectors in place of scalars. It was necessary to process elements in groups rather
than individually as had been done earlier on the CDC-7600 supercomputers.
Since vector registers are generally some multiple of 64 words, vector lengths of 64
or some multiple are appropriate. In LS-DYNA, groups of 128 elements or possibly some
larger integer multiple of 64 are utilized. Larger groups give a marginally faster code, but
can reduce computer time sharing efficiency because of increased core requirements. If
elements within the group reference more than one material model, subgroups are formed
for consecutive elements that reference the same model. LS-DYNA internally sorts
elements by material to maximize vector lengths.
E11=SGV1*PX1+SGV4*PY1+SGV6*PZ1
E21=SGV2*PY1+SGV4*PX1+SGV5*PZ1
E31=SGV3*PZ1+SGV6*PX1+SGV5*PY1
E12=SGV1*PX2+SGV4*PY2+SGV6*PZ2
E22=SGV2*PY2+SGV4*PX2+SGV5*PZ2
E32=SGV3*PZ2+SGV6*PX2+SGV5*PY2
E13=SGV1*PX3+SGV4*PY3+SGV6*PZ3
E23=SGV2*PY3+SGV4*PX3+SGV5*PZ3
E33=SGV3*PZ3+SGV6*PX3+SGV5*PY3
E14=SGV1*PX4+SGV4*PY4+SGV6*PZ4
E24=SGV2*PY4+SGV4*PX4+SGV5*PZ4
E34=SGV3*PZ4+SGV6*PX4+SGV5*PY4
DO 10 I = LFT, LLT
X1(I) = X(1,IX1(I))
Y1(I) = X(2,IX1(I))
Z1(I) = X(3,IX1(I))
VX1(I) = V(1,IX1(I))
VY1(I) = V(2,IX1(I))
VZ1(I) = V(3,IX1(I))
X2(I) = X(1,IX2(I))
Y2(I) = X(2,IX2(I))
Z2(I) = X(3,IX2(I))
VX2(I) = V(1,IX2(I))
VY2(I) = V(2,IX2(I))
VZ2(I) = V(3,IX2(I))
X3(I) = X(1,IX3(I))
X3(I) = X(2,IX3(I))
X3(I) = X(3,IX3(I))
X8(I) = X(1,IX8(I))
Y8(I) = X(2,IX8(I))
Z8(I) = X(3,IX8(I))
VX8(I) = V(1,IX8(I))
VY8(I) = V(2,IX8(I))
10 VZ8(I) = V(3,IX8(I))
initializes the nodal velocity and coordinate vector for each element in the subgroup LFT to
LLT. In the scatter operation, element nodal forces are added to the global force vector.
The force assembly does not vectorize unless special care is taken as described below.
DO 30 I = 1,NODFRC
DO 20 N = 1,NUMNOD
DO 10 L = LFT,LLT
RHS(I,IX(N,L))=RHS(I,IX(N,L))+FORCE(I,N,L)
10 CONTINUE
20 CONTINUE
30 CONTINUE
where NODFRC is the number of force components per node (3 for solid elements, 6 for
shells), LFT and LLT span the number of elements in the vector block, NUMNOD is the
number of nodes defining the element, FORCE contains the force components of the
individual elements, and RHS is the global force vector. This loop does not vectorize since
the possibility exists that more that one element may contribute force to the same node.
FORTRAN vector compilers recognize this and will vectorize only if directives are added
to the source code. If all elements in the loop bounded by the limits LFT and LLT are
disjoint, the compiler directives can be safely added. We therefore attempt to sort the
elements as shown in Figure 32.1 to guarantee disjointness.
Block 1 Block 2
Block 3 Block 3
The current implementation was strongly motivated by Benson [1989] and by work
performed at General Motors [Ginsberg and Johnson 1988, Ginsberg and Katnik 1989],
where it was shown that substantial improvements in execution speed could be realized by
blocking the elements in the force assembly. Katnik implemented element sorting in a
public domain version of DYNA3D for the Belytschko-Tsay shell element and added
compiler directives to force vectorization of the scatter operations associated with the
addition of element forces into the global force vector. The sorting was performed
immediately after the elements were read in so that subsequent references to the stored
element data were sequential. Benson performed the sorting in the element loops via
indirect addressing. In LS-DYNA the published GM approach is taken.
32.2 Parallelization
In parallelization, the biggest hurdle is overcoming Amdahl’s law for multitasking
[Cray Research Inc. 1990]
1
𝑆𝑚 = ,
𝑓𝑝 (32.1)
𝑓𝑠 +
𝑁
where
Table 29.1 shows that to obtain a speed factor of four on eight processors it is
necessary to have eighty-six percent of the job running in parallel. Obviously, to gain the
highest speed factors the entire code must run in parallel.
LS-DYNA has been substantially written to function on all shared memory parallel
machine architectures. Generally, shared memory parallel speed-ups of 5 on 8 processors
are possible but this is affected by the machine characteristics. We have observed speeds of
5.6 on full car crash models on a machine of one manufacturer only to see a speed-up of 3.5
on a different machine of another manufacturer.
Table 29.1.Maximum theoretical speedup Sm, on N CPUs with parallelism [Cray Research
Inc. 1990].
In the element loops element blocks with vector lengths of 64 or some multiple are
assembled and sent to separate processors. All elements are processed in parallel. On the
average a speed factor of 7.8 has been attained in each element class corresponding to
99.7% parallelization.
additional storage it is usually possible to eradicate the GUARDS from the coding. The
effort may not be worth the gains in execution speed.
The element blocks are defined at the highest level and each processor updates the
entire block including the right hand side force assembly. The user currently has two
options: GUARD compiler directives prevent simultaneous updates of the RHS vector
(recommended for single CPU processors or when running in a single CPU mode on a
multi-processor), or assemble the right hand side in parallel and let LS-DYNA prevent
conflicts between CPU’s. This usually provides the highest speed and is recommended,
i.e., no GUARDS.
When executing LS-DYNA in parallel, the order of operations will vary from run to
run. This variation will lead to slightly different numerical results due to round-off errors.
By the time the calculation reaches completion variations in nodal accelerations and
sometimes even velocities are observable. These variations are independent of the
precision and show up on both 32 and 64 bit machines. There is an option in LS-DYNA to
use an ordered summation of the global right hand side force vector to eliminate numerical
differences. To achieve this the element force vectors are stored. After leaving the element
loop, the global force vector is assembled in the same order that occurs on one processor.
The ordered summation option is slower and uses more memory than the default, but it
leads to nearly identical, if not identical, results run to run.
Parallelization in LS-DYNA was initially done with vector machines as the target
where the vector speed up is typically 10 times faster than scalar. On vector machines,
therefore, vectorization comes first. If the problem is large enough then parallelization is
automatic. If vector lengths are 128, for example, and if 256 beam elements are used only a
factor of 2 in speed can be anticipated while processing beam elements. Large contact
surfaces will effectively run in parallel, small surfaces having under 100 segments will not.
The speed up in the contact subroutines has only registered 7 on 8 processors due to the
presence of GUARD statements around the force assembly. Because real models often use
many special options that will not even vectorize efficiently it is unlikely that more than
95% of a given problem will run in parallel on a shared memory parallel machine.
33
Airbags
An alternative approach for calculating the airbag volume, that is both applicable
during the inflation phase and less computationally demanding, treats the airbag as a
control volume. The control volume is defined as the volume enclosed by a surface. In the
present case, the ‘control surface’ that defines the control volume is the surface modeled by
shell or membrane elements comprising the airbag fabric material.
Because the evolution of the control surface is known, i.e., the position, orientation,
and current surface area of the airbag fabric elements are computed and stored at each time
step, we can take advantage of these properties of the control surface elements to calculate
the control volume, i.e., the airbag volume. The area of the control surface can be related to
the control volume through Green’s Theorem
∂𝜓 ∂𝜙
∭𝜙 𝑑𝑥𝑑𝑦𝑑𝑧 = − ∭ 𝜓 𝑑𝑥𝑑𝑦𝑑𝑧 + ∮ 𝜙𝜓𝑛𝑥 𝑑𝛤 , (33.1)
∂𝑥 ∂𝑥
where the first two integrals are integrals over a closed volume, i.e., 𝑑𝑣 = 𝑑𝑥𝑑𝑦𝑑𝑧, the last
integral is an integral over the surface enclosing the volume, and 𝑛𝑥 is the direction cosine
between the surface normal and the 𝑥 direction (corresponding to the x-partial derivative);
similar forms can be written for the other two directions. The two arbitrary functions 𝜙
and 𝜓 need only be integrated over the volume and surface.
𝑉 = ∭ 𝑑𝑥𝑑𝑦𝑑𝑧. (33.2)
Comparing the first of the volume integrals in Equation (33.1) to Equation (33.2), we can
easily obtain the volume integral from Equation (33.1) by choosing for the two arbitrary
functions
𝜙 = 1, (33.3)
𝜓 = 𝑥𝑥 , (33.4)
leading to
The surface integral in Equation (33.5) can be approximated by a summation over all the
elements comprising the airbag, i.e.,
𝑁
∮ 𝑥𝑛𝑥 𝑑𝛤 ≈ ∑ 𝑥̅𝑖 𝑛𝑖𝑥 𝐴𝑖 , (33.6)
𝑖=1
where for each element i: 𝑥̅𝑖 is the average x coordinate, 𝑛𝑖𝑥 is the direction cosine between
the elements normal and the 𝑥 direction, and 𝐴𝑖 is the surface area of the element.
Although Equation (33.5) will provide the exact analytical volume for an arbitrary
direction, i.e., any 𝑛, the numerical implementation of Equation (33.5), and its approxima-
tion Equation (33.6), has been found to produce slightly different volumes, differing by a
few percent, depending on the choice of directions: if the integration direction is nearly
parallel to a surface element, i.e., the direction cosine is nearly zero, numerical precision
errors affect the volume calculation. The implementation uses as an integration direction, a
direction that is parallel to the maximum principle moment of inertia of the surface.
Numerical experiments have shown this choice of integration direction produces more
accurate volumes than the coordinate or other principle inertia directions.
Because airbag models may contain holes, e.g., holes for inflation and deflation, and
Green’s Theorem only applies to closed surfaces, a special treatment is needed for
calculating the volume of airbags with holes. This special treatment consists of the
following:
• The n-sized polygon defining the hole is identified automatically, using edge
locating algorithms in LS-DYNA.
• The n-sized polygon is projected onto a plane, i.e., it is assumed to be flat; this
is a good approximation for typical airbag hole geometries. Planar symmetry should work
with the control volume capability for one symmetry plane.
• The area of the flat n-sided polygon is calculated using Green’s Theorem in
two dimensions.
• The resulting holes are processed as another surface element in the airbag
control volume calculation.
The equation of state used for the airbag simulations is the usual ‘Gamma Law Gas
Equation of State’,
𝑝 = (𝑘 − 1)𝜌𝑒, (33.7)
where 𝑝 is the pressure, 𝑘 is a constant defined below, 𝜌 is the density, and 𝑒 is the specific
internal energy of the gas. The derivation of this equation of state is obtained from
thermodynamic considerations of the adiabatic expansion of an ideal gas. The incremental
change in internal energy, 𝑑𝑈, in 𝑛 moles of an ideal gas due to an incremental increase in
temperature, 𝑑𝑇, at constant volume is given by
𝑑𝑈 = 𝑛𝑐v 𝑑𝑇, (33.8)
where 𝑐v is the specific heat at constant volume. Using the ideal gas law we can relate the
change in temperature to a change in the pressure and total volume, 𝑣, as
𝑑(𝑝𝑣) = 𝑛𝑅𝑑𝑇, (33.9)
where 𝑅 is the universal gas constant. Solving the above for 𝑑𝑇 and substituting the result
into Equation (33.8) gives
𝑐v 𝑑(𝑝𝑣) 𝑑(𝑝𝑣)
𝑑𝑈 = = , (33.10)
𝑅 (𝑘 − 1)
where we have used the relationship
𝑅 = 𝑐p − 𝑐v , (33.11)
and the notation
𝑐p
𝑘= . (33.12)
𝑐v
The equation of state and the control volume calculation can only be used to
determine the pressure when the specific internal energy is also known. The evolution
equation for the internal energy is obtained by assuming the change in internal energy is
given by
𝑑𝑈 = −𝑝𝑑𝑣, (33.16)
where the minus sign is introduced to emphasize that the volume increment is negative
when the gas is being compressed. This expression can be written in terms of the specific
internal energy as
𝑑𝑈 𝑝𝑑𝑣
𝑑𝑒 = =− . (33.17)
𝜌 0 𝑣0 𝜌0 𝑣
Next, we divide the above by the equation of state, Equation (4.11.144), to obtain
𝑑𝑒 𝜌(𝑘 − 1)𝑑𝑣 (𝑘 − 1)𝑑𝑣
=− =− , (33.18)
𝑒 𝜌 0 𝑣0 𝑣
which may be integrated to yield
ln𝑒 = (1 − 𝑘)ln𝑉, (33.19)
or evaluating at two states and exponentiating both sides yields
𝑣 (1−𝑘)
𝑒2 = 𝑒1 ( 2 ) . (33.20)
𝑣1
The specific internal energy evolution equation, Equation (33.20), the equation of
state, Equation (4.11.144), and the control volume calculation completely define the
pressure-volume relation for an inflated airbag.
Wang and Nefske define the mass flow through the vents and leakage by
𝑝2 1⁄ 𝑘𝑅 𝑘−1⁄
𝑚̇ 23 = 𝐶23 𝐴23 𝑄 𝑘 √2𝑔 (33.24)
𝑐( ) (1 − 𝑄 𝑘 ),
𝑅√𝑇2 𝑘−1
and
𝑝2 1⁄ 𝑘𝑅 𝑘−1⁄
𝑚̇ ′23 = 𝐶′23 𝐴′23 𝑄 𝑘 √2𝑔
𝑐( ) (1 − 𝑄 𝑘 ), (33.25)
𝑅√𝑇2 𝑘−1
where 𝐶23 , 𝐴23 , 𝐶′23 , 𝐴′23 , 𝑅 and 𝑔𝑐 are the vent orifice coefficient, vent orifice area, the
orifice coefficient for leakage, the area for leakage, the gas constant, and the gravitational
conversion constant, respectively. The internal temperature of the airbag gas is denoted by
𝑇2 . We note that both 𝐴23 and 𝐴′23 can be defined as a function of pressure [Wang, 1992]
or if they are input as zero they are computed within LS-DYNA. This latter option requires
detailed modeling of the airbag with all holes included.
𝑝2 𝑉 = 𝑚2 𝑅𝑇2 . (33.26)
Solving for 𝑇2 :
𝑝2 𝑉
𝑇2 = , (33.27)
𝑚2 𝑅
and substituting Equation (33.27) into equations (33.25), we arrive at the mass transient
equation:
√ 2
√𝑘 (𝑄𝑘 − 𝑄𝑘+1⁄𝑘 )
√ (33.28)
𝑚̇ out = 𝑚̇ 23 + 𝑚̇ ′23 = 𝜇√2𝑝2 𝜌
⎷ 𝑘−1
where
𝜌 = density of airbag gas,
𝜇 = bag characterization parameter,
𝑚̇ out = total mass flow rate out of bag.
The mass flow rate and gas temperature are defined in load curves as a function of
time. Using the mass flow rate we can easily compute the increase in internal energy:
𝐸̇in = 𝑐p 𝑚̇ in 𝑇in , (33.30)
where 𝑇in is the temperature of the gas flowing into the airbag. Initializing the variables
pressure, 𝑝, density, 𝜌, and energy, 𝐸, to their values at time 𝑛, we can begin the iterations
loop to compute the new pressure, 𝑝𝑛 + 1 , at time 𝑛 + 1.
𝑛+1⁄2 𝑝𝑛 + 𝑝𝑛+1
𝑝 =
2
𝑛+1⁄2 𝜌𝑛 + 𝜌𝑛+1
𝜌 =
2
𝑛+1⁄2 𝐸𝑛 + 𝐸𝑛+1 (33.31)
𝐸 =
2
𝑛+1⁄2
⎛
⎜ 𝑝𝑒 ⎞
⎟
𝑄 = max ⎜
⎜ , 𝑄 ⎟.
⎜ 1 crit ⎟
⎟
𝑛+
2
⎝ 𝑝2 ⎠
The mass flow rate out of the bag, 𝑚̇ out can now be computed:
√ 2 𝑘+1⁄
√ ⎛ 𝑛+1⁄2 ⁄𝑘 𝑛+1⁄2 𝑘⎞
√𝑘 ⎜ 𝑄 −𝑄 ⎟ (33.32)
𝑛+1⁄2 𝑛+1⁄2 𝑛+1⁄ √ ⎝ ⎠
𝑚̇ out = 𝜇√2𝑝2 𝜌 2 ,
⎷ 𝑘−1
where
𝑛+1⁄2 𝑛+1⁄2
𝑝2 =𝑝 + 𝑝e , (33.33)
𝑛+1⁄2
where Δ𝑉 is the change in volume from time 𝑛 to 𝑛 + 1. The new pressure can now be
computed:
𝐸𝑛+1
𝑝𝑛+1 = (𝑘 − 1)
, (33.37)
𝑉 𝑛+1
which is the gamma-law (where 𝑘 = 𝛾) gas equation. This ends the iteration loop.
The rate of change of internal energy, the left hand side of Equation (33.38), can be
differentiated:
𝑑 𝑑𝑚 𝑑𝑢 𝑑𝑚 𝑑 𝑑𝑇
(𝑚𝑢) = 𝑢+𝑚 = 𝑢 + 𝑚 (𝑐v 𝑇) = 𝑚̇ 𝑢 + 𝑚𝑐v̇ 𝑇 + 𝑚𝑐v , (33.39)
𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡
where we have used the definition
𝑢 = 𝑐v 𝑇. (33.40)
Then, the energy equation can be re-written for the rate of change in temperature for
the airbag
𝑑𝑇cv ∑ 𝑚̇ 𝑖 ℎ𝑖 − ∑ 𝑚̇ 𝑜 ℎ𝑜 − 𝑊̇ cv − 𝑄̇cv − (𝑚̇ 𝑢)cv − (𝑚𝑐v̇ 𝑇)cv
= (33.41)
𝑑𝑡 (𝑚𝑐v )cv
Temperature dependent heat capacities are used. The constant pressure molar heat
capacity is taken as:
𝑐p̅ = 𝑎 ̅ + 𝑏̅𝑇, (33.42)
and the constant volume molar heat capacity as:
𝑐v̅ = 𝑎 ̅ + 𝑏̅𝑇 − 𝑟 ̅, (33.43)
where
𝑟 ̅ = gas constant = 8.314 J/gm-mole K
𝑎 ̅ = constant [J/gm-mole K]
𝑏̅ = constant [J/gm-mole K2]
Mass based values are obtained by dividing the molar quantities by the molecular
weight, 𝑀, of the gas
𝑎̅ 𝑏̅ 𝑟̅
𝑎= , 𝑏= , 𝑟= . (33.44)
𝑀 𝑀 𝑀
The constant pressure and volume specific heats are then given by
𝑐p = 𝑎 + 𝑏𝑇 (33.45)
𝑐v = 𝑎 + 𝑏𝑇 − 𝑟. (33.46)
The specific enthalpy and internal energy becomes:
𝑇
𝑏𝑇 2
ℎ = ∫ 𝑐𝑝 𝑑𝑇 = 𝑎𝑇 + (33.47)
2
0
𝑇
𝑏𝑇 2
𝑢 = ∫ 𝑐𝑣 𝑑𝑇 = 𝑎𝑇 + − 𝑟𝑇. (33.48)
2
0
𝑐v = ∑ 𝑓𝑖 𝑐v(𝑖) , (33.51)
where
𝑓𝑖 = mass fraction of gas 𝑖
𝑀𝑖 = molecular weight of gas 𝑖
𝑐p(𝑖) = constant pressure specific heat of gas 𝑖
𝑐v(𝑖) = constant volume specific heat of gas 𝑖.
The specific enthalpy and internal energy for an ideal gas mixture with temperature
dependent heat capacity are
𝑇
𝑏𝑖 𝑇 2
ℎ = ∫ ∑ 𝑓𝑖 𝑐p(𝑖) 𝑑𝑇 = ∑ 𝑓𝑖 (𝑎𝑖 𝑇 + ) (33.52)
2
0
𝑇
𝑏𝑖 𝑇 2
𝑢 = ∫ ∑ 𝑓𝑖 𝑐v(𝑖) 𝑑𝑇 = ∑ 𝑓𝑖 (𝑎𝑖 𝑇 + − 𝑟𝑖 𝑇). (33.53)
2
0
2
𝑏𝑖 𝑇cv
∑ 𝑚̇ 𝑜 ℎ𝑜 = ∑ 𝑚̇ 𝑜 [ ∑ 𝑓𝑖 (𝑎𝑖 𝑇𝑐𝑣 + )]. (33.56)
gases 2
The gas leaves the airbag at the control volume temperature 𝑇𝑐𝑣 . The mass flow rate
out through vents and fabric leakage is calculated by the one dimensional isentropic flow
equations per Wang and Nefske. The work done by the airbag expansion is given by:
𝑊̇ cv = ∫ 𝑃𝑑𝑉̇ , (33.57)
𝑃 is calculated by the equation of state for a perfect gas, 𝑝 = 𝜌𝑅𝑇 and 𝑉̇ is calculated by
LS-DYNA
For the energy balance, we must compute the energy terms (𝑚̇ 𝑢)cv and (𝑚𝑐v )cv .
Conservation of mass leads to:
𝑚̇ cv = 𝑚̇ 𝑖 − 𝑚̇ 𝑜
(33.58)
𝑚cv = ∫ 𝑚̇ cv 𝑑𝑡.
34
Dynamic Relaxation and System
Damping
In LS-DYNA we have two methods of damping the solution. The first named
“dynamic relaxation” is used in the beginning of the solution phase to obtain the initial
stress and displacement field prior to beginning the analysis. The second is system
damping which can be applied anytime during the solution phase either globally or on a
material basis.
𝐐𝑛 (𝐝) = 𝐅𝑛 − 𝐏𝑛 − 𝐇𝑛 , (34.2)
LS-DYNA DEV 10/27/16 (r:8004) 32-1 (Dynamic Relaxation and System Damping)
Dynamic Relaxation and System Damping LS-DYNA Theory Manual
where we recall that 𝐌 is the mass matrix, 𝐂 is the damping matrix, 𝑛 indicates the nth
time step, 𝐚𝑛 is the acceleration, 𝐯𝑛 the velocity, and 𝐝 is the displacement vector. With Δ𝑡
as the fixed time increment we get for the central difference scheme:
𝑛+1⁄2 𝑛−1⁄2
(𝐝𝑛+1 − 𝐝𝑛 ) (𝐯 −𝐯 )
𝑛+1⁄2 (34.3)
𝐯 = ; 𝐚𝑛 = .
Δ𝑡 Δ𝑡
For 𝐯𝑛 we can assume an averaged value
1 𝑛+1⁄ 1
𝐯𝑛 = (𝐯 2 + 𝐯𝑛− ⁄2 ), (34.4)
2
and obtain
−1
𝑛+1⁄2 1 1 1 1 𝑛−1⁄2
𝐯 =( 𝐌 + 𝐂) [( 𝐌 − 𝐂) 𝐯 − 𝐐𝑛 ], (34.5)
Δ𝑡 2 Δ𝑡 2
𝑛+1⁄2 (34.6)
𝐝𝑛+1 = 𝐝𝑛 + Δ𝑡𝐯 .
In order to preserve the explicit form of the central difference integrator, 𝐌 and 𝐂
must be diagonal. For the dynamic relaxation scheme 𝐂 has the form
𝐂 = 𝑐 ⋅ 𝐌. (34.7)
If Equation (34.7) is substituted into (34.5) the following form is achieved
𝑛+1⁄2 2 − 𝑐Δ𝑡 𝑛−1⁄ 2Δ𝑡
𝐯 = 𝐯 2+ ⋅ 𝐌−1 ⋅ 𝐐𝑛 . (34.8)
2 + 𝑐Δ𝑡 2 + 𝑐Δ𝑡
Since 𝐌 is diagonal, each solution vector component may be computed individually from
𝑛+1⁄2 2 − 𝑐Δ𝑡 𝑛−1⁄2 2Δ𝑡 𝐐𝑛𝑖
𝐯𝑖 = 𝐯𝑖 + . (34.9)
2 + 𝑐Δ𝑡 2 + 𝑐Δ𝑡 𝑚𝑖
As a starting procedure it is suggested by Underwood
𝐯0 = 0
(34.10)
𝐝0 = 0.
Since the average value is used for 𝐯𝑛 , which must be zero at the beginning for a quasi-
static solution
−1⁄2 1⁄
(34.11)
𝐯 = −𝐯 2,
32-2 (Dynamic Relaxation and System Damping) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Dynamic Relaxation and System Damping
Papadrakakis [1981]. Then dynamic relaxation is nothing else but a critically damped
dynamic system
𝐶 = 𝐶cr = 2𝜔min 𝑚, (34.13)
with 𝑚 as modal mass. The problem is finding the dominant eigenvalue in the structure
related to the “pseudo-dynamic” behavior of the structure. As the exact estimate would be
rather costly and not fit into the explicit algorithm, an estimate must be used.
Papadrakakis suggests
∥𝐝𝑛+1 − 𝐝𝑛 ∥
𝜆𝐷 = . (34.14)
∥ 𝐝𝑛 − 𝐝𝑛−1 ∥
When this quantity has converged to an almost constant value, the minimum eigenvalue of
the structure can be estimated:
2
(𝜆2𝐷 − 𝜆𝐷 ⋅ 𝛽 + 𝛼)
𝜔min =− , (34.15)
𝜆𝐷 ⋅ 𝛾
where
2 − 𝑐Δ𝑡
𝛼=
2 + 𝑐Δ𝑡
𝛽=𝛼+1 (34.16)
2Δ𝑡2
𝛾= .
2 + 𝑐Δ𝑡
The maximum eigenvalue determines the time step and is already known from the
model
2 4.0
𝜔max = . (34.17)
(Δ𝑡)2
Now the automatic adjustment of the damping parameter closely follows the paper of
Papadrakakis, checking the current convergence rate compared to the optimal convergence
rate. If the ratio is reasonably close, then an update of the iteration parameters is
performed.
2 2
4.0 √𝜔min ⋅ 𝜔max
(34.18)
𝑐= 2
.
Δ𝑡 (𝜔min 2
+ 𝜔max )
As is clearly visible from Equation (34.18) the value of highest frequency has always a
rather high influence on the damping ratio. This results in a non-optimal damping ratio, if
the solution is dominated by the response in a very low frequency compared to the highest
frequency of the structure. This is typically the case in shell structures, when bending
dominates the solution. It was our observation that the automatic choice following
Papadrakakis results in very slow convergence for such structures, and this is also
mentioned by Underwood for similar problems. The damping ratio should then be fully
adjusted to the lowest frequency by hand by simply choosing a rather high damping ratio.
LS-DYNA DEV 10/27/16 (r:8004) 32-3 (Dynamic Relaxation and System Damping)
Dynamic Relaxation and System Damping LS-DYNA Theory Manual
An automatic adjustment for such cases is under preparation. For structures with
dominant frequencies rather close to the highest frequency, convergence is really improved
with the automatically adjusted parameter.
where 𝜂is an input damping factor (defaulted to .995). The factor, 𝜂, is equivalent to the
corresponding factor in Equations (31.7- 31.8).
The relaxation process continues until a convergence criterion based on the global
kinetic energy is met, i.e., convergence is assumed if
𝐸ke < CVTOL ⋅ 𝐸max
𝑘𝑒 , (34.20)
where CVTOL is the convergence tolerance (defaulted to .001). The kinetic energy excludes
any rigid body component. Initial velocities assigned in the input are stored during the
relaxation. Once convergence is attained the velocity field is initialized to the input values.
A termination time for the dynamic relaxation phase may be included in the input and is
recommended since if convergence fails, LS-DYNA will continue to execute indefinitely.
As seen from Figure 34.1 and as discussed above the best damping constant for the
system is usually the critical damping constant: Therefore,
𝐷𝑠 = 2𝜔min (34.23)
is recommended.
32-4 (Dynamic Relaxation and System Damping) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Dynamic Relaxation and System Damping
The number of cycles required to reduce the amplitude of the dynamic response by a factor
of 10 can be approximated by [see Stone, Krieg, and Beisinger 1985]
𝜔
ncycle = 1.15 max . (34.24)
𝜔min
Structural problems which involve shell and beam elements can have a very large ratio and
consequently very slow convergence.
LS-DYNA DEV 10/27/16 (r:8004) 32-5 (Dynamic Relaxation and System Damping)
LS-DYNA Theory Manual Heat Transfer
35
Heat Transfer
LS-DYNA can be used to solve for the steady state or transient temperature field on
three-dimensional geometries. Material properties may be temperature dependent and
either isotropic or orthotropic. A variety of time and temperature dependent boundary
conditions can be specified including temperature, flux, convection, and radiation. The
implementation of heat conduction into LS-DYNA is based on the work of Shapiro [1985].
subject to the boundary conditions, 𝜃 = 𝜃𝑠 on Γ1, 𝑘𝑖𝑗 𝜃,𝑗 𝑛𝑖 + 𝛽𝜃 = 𝛾 on Γ2, and initial
conditions at 𝑡0 :
𝜃Γ = 𝜃0 (𝑥𝑖 ) at 𝑡 = 𝑡0 . (35.2)
where
𝜃 = 𝜃(𝑥𝑖 , 𝑡) temperature
𝑥𝑖 = 𝑥𝑖 (𝑡) coordinates as a function of time
𝜌 = 𝜌(𝑥𝑖 ) density
𝑐𝑝 = 𝑐𝑝 (𝑥𝑖 , 𝜃) specific heat
𝑘𝑖𝑗 = 𝑘𝑖𝑗 (𝑥𝑖 , 𝜃) specific heat
𝑄 = 𝑄(𝑥𝑖 , 𝜃) internal heat generation rate per unit volume Ω
𝜃Γ = prescribed temperature on Γ1
𝑛𝑖 = normal vector to Γ2
Equations (35.1)-(35.2) represent the strong form of a boundary value problem to be solved
for the temperature field within the solid.
32.1
Heat Transfer LS-DYNA Theory Manual
DYNA3D employs essentially the same theory as TOPAZ [Shapiro 1985] in solving
Equation (35.1) by the finite element method. Those interested in a more detailed
description of the theory are referred to the TOPAZ User’s Manual. Brick elements are
integrated with a 2 × 2 × 2 Gauss quadrature rule, with temperature dependence of the
properties accounted for at the Gauss points. Time integration is performed using a
generalized trapezoidal method shown by Hughes to be unconditionally stable for
nonlinear problems. Newton’s method is used to satisfy equilibrium in nonlinear
problems.
The finite element method provides the following equations for the numerical
solution of Equations (35.1)-(35.2)
𝐶𝑛+𝛼
[ + 𝛼𝐻𝑛+𝛼 ] {𝜃𝑛+1 − 𝜃𝑛 } = {𝐹𝑛+𝛼 − 𝐻𝑛+𝛼 𝜃𝑛 } (35.3)
Δ𝑡
where
[𝐻] = ∑[𝐻𝑖𝑗𝑒 ] = ∑ ⎡
⎢ ∫ ∇ 𝑁𝑖 𝐾∇𝑁𝑗 𝑑Ω + ∫ 𝑁𝑖 𝛽𝑁𝑗 𝑑Γ⎤
𝑇
⎥ (35.5)
𝑒 𝑒 ⎣Ω𝑒 Γ𝑒 ⎦
[𝐹] = ∑[𝐹𝑖𝑒 ] = ∑ ⎡
⎢ ∫ 𝑁𝑖 𝑞𝑔 𝑑Ω + ∫ 𝑁𝑖 𝛾𝑑Γ⎤
⎥ (35.6)
𝑒 𝑒 ⎣Ω𝑒 Γ 𝑒 ⎦
32.2
LS-DYNA Theory Manual Heat Transfer
By convention, heat flow is positive in the direction of the surface outward normal vector.
Surface definition is in accordance with the right hand rule. The outward normal vector
points to the right as one progresses from node N1 to N2 to N3 and finally to N4. See
Figure 35.1.
1
𝑇 = (𝑇surf + 𝑇∞ ) (35.8)
2
N4
N1
N3
N2
Figure 35.1. Definition of the outward normal vector
32.3
Heat Transfer LS-DYNA Theory Manual
In a steady state nonlinear problem, an initial guess should be made of the final
temperature distribution and included in the input file as an initial condition. If your guess
is good, a considerable savings in computation time is achieved.
35.8 Units
Any consistent set of units with the governing equation may be used. Examples are:
Quantity Units
temperature K C F
space m cm ft
time s s hr
density kg/m3 g/cm3 Lbm/ft3
heat capacity J/kg k cal/g c Btu/LbmF
thermal conductivity W/m K cal/s cm C Btu/hr ft F
thermal generation W/M3 cal/s cm3 Btu/hr ft3
heat flux W/m2 cal/s cm2 Btu/hr ft2
32.5
LS-DYNA Theory Manual Adaptivity
36
Adaptivity
LS-DYNA uses an h-adaptive process, where parts of the mesh are selectively
refined during the course of the solution procedure. The methodology used is based on
Belytschko, Wong, and Plaskacz [1989]. In the former, elements were also fused or
combined when it was felt that they were no longer needed. It was found that the
implementation of fusing procedures for general meshes, such as occur in typical
applications of commercial programs, is too complex, so only fission is included.
Adaptivity in LS-DYNA can be restricted to specific groups of shell elements. Elements
that fall in this group are said to be in the active adaptivity domain.
Figure 36.1. One level adaptive calculation on a square cross section beam.
Figure 36.2. Aluminum blank with 400 shells in blank and four rigid tools.
In the h-adaptive process, elements are subdivided into smaller elements where
more accuracy is needed; this process is called fission. The elements involved in the fission
process are subdivided into elements with sides ℎ/2, where ℎ is the characteristic size of the
original elements. This is illustrated in Figure 36.5 for a quadrilateral element. In fission,
each quadrilateral is subdivided into four quadrilaterals (as indicated in Figure 36.2) by
using the mid-points of the sides and the centroid of the element to generate four new
quadrilaterals.
Figure 36.4. Final shape of formed part with 4315 shell elements per quarter.
The fission process for a triangular element is shown in Figure 36.6 where the
element is subdivided into four triangles by using the mid-points of the three sides. The
adaptive process can consist of several levels of fission. Figure 36.5 shows one subdivision,
which is called the second refinement level. In subsequent steps, the fissioned elements can
again be fissioned in a third refinement level, and these elements can again, in turn, be
fissioned in a fourth level, as shown in Figure 36.7. The levels of adaptivity that occur in a
mesh are restricted by three rules:
• The number of levels is restricted by the maximum level of adaptivity that is
allowed in the mesh, which is generally set at 3 or 4. At the fourth level up to 64
elements will be generated for each element in the initial mesh.
• The levels of adaptivity implemented in a mesh must be such that the levels of
adaptivity implemented in adjacent elements differ by, at most, one level.
• The total number of elements can be restricted by available memory. Once the
specified memory usage is reached, adaptivity ceases.
The second rule is used to enforce a 2-to-1 rule given by Oden, Devloo and
Strouboulis [1986], which restricts the number of elements along the side of any element in
the mesh to two. The enforcement of this rule is necessary to accommodate limitations in
the data structure.
The original mesh provided by the user is known as the parent mesh, the elements
of this mesh are called the parent elements, and the nodes are called parent nodes. Any
elements that are generated by the adaptive process are called descendant elements, and
34-4 (Adaptivity) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Adaptivity
any nodes that are generated by the adaptive process are called descendant nodes.
Elements generated by the second level of adaptivity are called first-generation elements,
those generated by third level of adaptivity are called second-generation elements, etc.
The coordinates of the descendant nodes are generated by using linear interpolation.
Thus, the coordinates of any node generated during fission of an element are given by
1
𝑥𝑁 = (𝑥𝐼 + 𝑥𝐽 ), (36.1)
2
where 𝑥𝑁 is the position of the generated node and 𝑥𝐼 and 𝑥𝐽 are the nodes along the side
on which 𝑥𝑁 was generated for a typical element as shown in Fig. 33.1. The coordinate of
the mid-point node, which is generated by fission of a quadrilateral element, is given by
1
𝑥𝑀 = (𝑥𝐼 + 𝑥𝐽 + 𝑥𝐾 + 𝑥𝐿 ), (36.2)
4
where 𝑥𝑀 is the new midpoint node of the fissioned quadrilateral and 𝑥𝐼 , 𝑥𝐽 , 𝑥𝐾 and 𝑥𝐿 are
the nodes of the original quadrilateral. The velocities of the nodes are also given by linear
interpolation. The velocities of edge nodes are given by
1
𝑣𝑁 = (𝑣𝐼 + 𝑣𝐽 ), (36.3)
2
and the angular velocities are given by
1
𝜔𝑁 = (𝜔𝐼 + 𝜔𝐽 ). (36.4)
2
The velocities of a mid-point node of a fissioned quadrilateral element are given by
1
𝑣𝑀 = (𝑣𝐼 + 𝑣𝐽 + 𝑣𝐾 + 𝑣𝐿 ), (36.5)
4
1
𝜔𝑀 = (𝜔𝐼 + 𝜔𝐽 + 𝜔𝐾 + 𝜔𝐿 ). (36.6)
4
The stresses in the descendant element are obtained from the parent element by
setting the stresses in the descendant elements equal to the stresses in the parent element at
the corresponding through-the-thickness quadrature points.
out-of-plane
undeformed deformed
Figure 36.8. Refinement indicator based on angle change.
In subsequent steps, nodes which are not corner nodes of an all attached elements
are treated as slave nodes. They are handled by the simple constraint equation.
Refinement indicators are used to decide the locations of mesh refinement. One
deformation based approach checks for a change in angles between contiguous elements as
shown in Figure 36.8. If 𝜁 > 𝜁tol then refinement is indicated, where 𝜁tol is user defined.
Figure 36.9. The input parameter, ADPASS, controls whether LS-DYNA backs
up and repeats the calculation after adaptive refinement.
After the mesh refinement is determined, we can refine the mesh and continue the
calculation or back up to an earlier time and repeat part of the calculation with the new
mesh. For accuracy and stability reasons the latter method is generally preferred; however,
the former method is preferred for speed. Whether LS-DYNA backs up and repeats the
calculation or continues after remeshing is determined by an input parameter, ADPASS.
This is illustrated in Figure 36.9.
37
Implicit
37.1 Introduction
Implicit solvers are properly applied to static, quasi-static, and dynamic problems with a
low frequency content. Such applications include but are not limited to
• Static and quasi-static structural design and analysis
• Metal forming, especially, the binderwrap and springback
• Gravitational loading of automotive structures
• Linear buckling and vibration analysis
An advantage of the implicit solver on explicit integration is that the number of load or
time steps is typically 100 to 10000 times fewer. The major disadvantage is that the cost per
step is unknown since the speed depends mostly on the convergence behavior of the
equilibrium iterations which can vary widely from problem to problem.
37.2 Equations
37.2.1 Discretization
Neglecting constraints, discretization formally leads to the matrix equations of motion
𝑹 = 𝑴𝒙̈ + 𝑭𝑖 − 𝑭𝑒 = 𝟎 (37.1)
where
𝒙̈ = acceleration vector of length 𝑛
𝑴 = 𝑛 × 𝑛 mass matrix
𝑭𝑒 = body force and external load vector of length 𝑛
𝑭𝑖 = internal force vector of length 𝑛.
6 This is true prior to time discretization, vectorsand depend on and on only through the exact time
differentiation. Once time discretization is done according to some scheme, the dependence is on and not
necessarily on the discretized but rather ∆ /∆ , see Section 1.4.1.
35-2 (Implicit) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Implicit
where 𝜌 is material density and 𝑉 denotes the body over which integration occurs.
The primary nonlinearities, which are due to geometric effects and inelastic material
behavior, are accounted for in 𝑭𝑖 ,
𝑭𝑖 = ∫ 𝑩𝑇 𝝈𝑑𝑉 , (37.3)
𝑉
Explicit integration trivially satisfies (37.1) since the calculation of the acceleration
guarantees equilibrium, i.e., from time step 𝑗 to 𝑗 + 1 we use
𝑗 𝑗
𝒙̈𝑗 = 𝑴 −1 [𝑭𝑒 − 𝑭𝑖 ]. (37.4)
The explicit update of the velocities and coordinates is given by
𝒙̇ 𝑗+1/2 = 𝒙̇ 𝑗−1/2 + 𝛥𝑡𝑗 𝒙̈𝑗 , (37.5)
37.2.2 Constraints
Constraints is obviously an important ingredient in nonlinear finite elements, and may
include for instance simple point or motion constraints, slave nodes constrained to rigid
bodies, joints and tied contacts. The principle behind (37.1) is that of virtual work, stating
that
𝛿𝒙 𝑇 𝑹 = 0 (37.7)
for any admissible virtual displacement field 𝛿𝒙. If there are no constraints, any
displacement field is admissible and hence 𝑹 = 𝟎. The presence of constraints put
restrictions on 𝛿𝒙 and the following is an attempt to derive the proper nonlinear equations
in this context.
The dependent part of the solution vector is in general analogous to (37.20) of the
indendent part
𝑘+1 𝑘 𝑘
𝒙𝐷 = 𝒙𝐷 + 𝑠∆𝒙𝐷 (37.27)
except for those constraint equations where there is an explicit expression 𝒙𝐷 = 𝒙𝐷 (𝒙𝐼 ),
then of course
𝑘+1
𝒙𝐷 = 𝒙𝐷 (𝒙𝐼𝑘+1 ). (37.28)
The jacobian matrix 𝑲 is the assembly of the tangent moduli of materials, external loads,
contacts, etc., and is in practice only an approximation due to the complexity of taking all
dependencies into account. In particular it does not include the geometric stiffness
contribution from internal forces by default, see IGS on CONTROL_IMPLICIT_GENERAL.
A speculative reason is that this has a smoothing effect and eliminates negative eigenvalues
due to compressive stresses. If the deformation mode is known to be mainly in tension or
if the material is hyper-elastic, including the geometric stiffness could improve
convergence however. Often the stiffness matrix is assumed symmetric and positive
definite, but is not limited to those characteristics. Note also that (37.25) indicates that both
𝑸 and 𝑯 must vanish to render a zero 𝑭, thus a zero displacement increment ∆𝒙 in the
iterative scheme.
Furthermore, 𝑷 is constant due to trivial constraints and the stiffness matrix 𝑲 in (37.26)
thus evaluates from the linearization of internal forces
𝑲 = 𝑷 ∫ 𝑩𝑇 𝑬𝑩𝑑𝑉 𝑷 𝑇 , (37.29)
𝑉
where we used 𝑬 = 𝜕𝝈 𝜕𝜺
to denote the constitutive matrix. So in more conventional terms,
the linear equation is written
𝑲𝒖 = 𝑭. (37.30)
Once solved, the stress
𝝈 = 𝑬𝑩𝒖 (37.31)
can be evaluated from the constitutive law for the resulting deformation. The stiffness
matrix 𝑲 in (37.30) is symmetric and positive definite if 𝑬 is and 𝑷 eliminates all rigid body
modes, whence the solution 𝒖 is unique. Furthermore, linearity implies that substituting
the right hand side for 𝑭𝜆 = 𝜆𝑭, the solution changes to 𝒖𝜆 = 𝜆𝒖 and the resulting stress
changes to
𝝈𝜆 = 𝜆𝝈. (37.32)
A linear solution is obtained by putting NSOLVR = 1 on CON-
TROL_IMPLICIT_SOLUTION.
Now go back to the nonlinear static equation and scale a constant external load 𝑭𝑒 with 𝜆
𝑷[𝑭𝑖 − 𝜆𝑭𝑒 ] = 𝟎 (37.33)
and assume an updated configuration 𝒙𝜆 with resulting stress 𝝈𝜆 to be the solution. To see
how the solution changes with 𝜆 we can perturb (37.33) to obtain
𝜕𝑭
𝑷 [ 𝑖 𝑷 𝑇 𝛿𝒙𝜆 − 𝛿𝜆𝑭𝑒 ] = 𝟎 (37.34)
𝜕𝒙
where we applied (37.14). This is conveniently rewritten as
𝑲𝜆 𝛿𝒙𝜆 = 𝛿𝜆𝑭, (37.35)
where the presence of stress will change the evaluation of the stiffness matrix from that in
(37.29) to
𝑲𝜆 = 𝑲 + 𝑷 ∫ 𝓑 𝑇 𝝈𝜆 𝓑𝑑𝑉 𝑷 𝑇 . (37.36)
𝑉
Here the second term is the geometric contribution to the tangent stiffness matrix where𝓑
is a matrix consisting of shape function derivatives for the finite element. We are interested
in the question of uniqueness of 𝛿𝒙𝜆 as the solution to (37.35), and this uniqueness is lost
when
det(𝑲𝜆 ) = 0. (37.37)
If one assumes that deformations are small enough to keep 𝑉, 𝑩, 𝑬 and 𝓑 independent of 𝜆,
and that (37.32) holds with (37.31) and (37.30), then (37.37) can be stated as the following
eigenvalue problem
𝑲𝛿𝒖 = −𝜆𝑲𝜎 𝛿𝒖 (37.38)
with the geometric (nonlinear) stiffness matrix given by
𝑲𝜎 = 𝑷 ∫ 𝓑 𝑇 𝝈𝓑𝑑𝑉 𝑷 𝑇 (37.39)
𝑉
and 𝑲 is the material (linear) stiffness matrix again given by (37.29). This is the theory
behind linear buckling, see CONTROL_IMPLICIT_BUCKLE, with 𝜆 being the buckling
load parameter and 𝛿𝒖 the associated buckling mode. Usually solutions with 𝜆 > 0 are of
interest, and from (37.38) and (37.39) it then follows that the principal stresses cannot be
positive throughout the domain of integration. In other words, the model must somehow
be in a compressed state. The full procedure is
1.Assemble 𝑲 by (37.29).
2.Solve (37.30) for a constant reference load 𝑭.
3.Evaluate resulting stress 𝝈 by (37.31).
4.Assemble 𝑲𝜎 by (37.39).
5.Solve (37.38) for 𝜆 and 𝛿𝒖.
𝐹0 𝐹0
𝐹 𝐹
𝐹1 𝐹1
𝐹2
𝐹2
𝑥 𝑥
𝑥2 𝑥1 𝑥0 𝑥2 𝑥1 𝑥0
Figure 3735-2 Full Newton (left) compared to quasi-Newton secant updates (right),
the linear approximation to obtain 𝑥2 is for full Newton the exact tangent in
(𝑥1 , 𝐹1 ) while it is the linear extension between the points (𝑥0 , 𝐹0 ) and (𝑥1 , 𝐹1 ) for
independent system of variables so the residual vector 𝑭 is given by (37.25). We here just
substitute the sub-index 𝐼 in the incremental displacement ∆𝒙 for the iteration counter 𝑘 to
simplify the notation in the following. The coordinate vector is updated
𝒙𝑘+1 = 𝒙𝑘 + sΔ𝒙𝑘 , (37.41)
where 𝑠 is a parameter between 0 and 1 found from a line search.
If the tangent stiffness matrix is calculated as (37.26) for each 𝑘 then this is termed a full
Newton method, but it may be beneficial to use so called quasi-Newton updates of 𝑲 −1 to
avoid the cost of solving a linear system of equations in each iteration. Four such methods
for updating the stiffness matrix are available
• Broyden’s first method
• Davidon
• DFP
• BFGS
and these involve rank 1 or rank 2 stiffness updates. Quasi-Newton methods are less
expensive than the full Newton method but often still result in robust program. In one
dimension, quasi-Newton corresponds to secant iterations and this method compared to
full Newton is illustrated in Figure 3735-2. Henceforth we assume that 𝑲0 represents the
last assembled matrix according to (37.26) and then 𝑲𝑘 , 𝑘 = 1,2,3, … are quasi-Newton
updates to be given in the following.
𝑭𝑘 𝒚 𝑇
𝑲𝑘 = 𝑲𝑘−1 − , (37.48)
𝒚𝑇 𝛥𝒙𝑘−1
resulting generally in non-symmetric secant matrices. The inverse forms are found by the
Sherman-Morrison formula
−1 𝑨−1 𝒂 𝒃𝑇 𝑨−1
(𝑨 + 𝒂𝒃𝑇 ) = 𝑨−1 − . (37.49)
1 + 𝒃𝑇 𝑨−1 𝒂
where 𝑨 is a nonsingular matrix and 𝒂 and 𝒃 are arbitrary vectors such that 1 + 𝒃𝑇 𝑨−1 𝒂 ≠
0. The inverse form for (37.48) can be found by letting
−𝑭𝑘
𝑨 = 𝑲𝑘−1 , 𝒂= , 𝒃 = 𝒚, (37.50)
𝒚𝑇 𝛥𝒙 𝑘−1
Again, recalling the quasi-Newton equation (37.42), and substituting (37.45) gives
𝑲𝑘−1 𝛥𝒙𝑘−1 + 𝛼𝒛𝒚𝑇 𝛥𝒙𝑘−1 + 𝛽𝒗𝒖𝑇 𝛥𝒙𝑘−1 = 𝛥𝑭𝑘 , (37.54)
1 1
and set 𝛼 = 𝒚𝑇 𝛥𝒙 , 𝒛 = −𝑭𝑘−1 , 𝛽 = 𝒖𝑇 𝛥𝒙 and 𝒗 = 𝛥𝑭𝑘 . Here 𝒚 and 𝒖 are arbitrary vectors
𝑘−1 𝑘−1
that are non-orthogonal to 𝛥𝒙𝑘−1 , i.e.,
𝒚𝑇 𝛥𝒙𝑘−1 ≠ 0, (37.55)
and
𝒖𝑇 𝛥𝒙𝑘−1 ≠ 0. (37.56)
In the BFGS method
𝒚 = 𝑭𝑘−1 𝒖 = 𝛥𝑭𝑘 , (37.57)
which leads to the following update formula
𝛥𝑭𝑘 𝛥𝑭𝑘𝑇 𝑇
𝑭𝑘−1 𝑭𝑘−1
𝑲𝑘 = 𝑲𝑘−1 + 𝑇
− 𝑇
, (37.58)
𝛥𝒙𝑘−1 𝛥𝑭𝑘 𝛥𝒙𝑘−1 𝑭𝑘−1
that preserves symmetry of the secant matrix. A double application of the Sherman-
Morrison formula then leads to the inverse form.
Special product forms have been derived for the DFP and BFGS updates and exploited by
Matthies and Strang [1979],
𝑲𝑘−1 = [𝑰 + 𝒒𝑘 𝒑𝑇𝑘 ] 𝑲𝑘−1
−1
[𝑰 + 𝒑𝑘 𝒒𝑘𝑇 ]. (37.59)
The primary advantage of the product form is that the determinant of 𝑲𝑘 and therefore, the
change in condition number can be easily computed to control updates. The updates
vectors are defined as
𝑇
𝛥𝒙𝑘−1 𝛥𝑭𝑘
𝒑𝑘 = −𝛥𝑭𝑘 − 𝑭𝑘−1 √ , (37.60)
𝑇
𝛥𝒙𝑘−1 𝑭𝑘−1
𝛥𝒙𝑘−1
𝒒𝑘 = 𝑇
. (37.61)
𝛥𝒙𝑘−1 𝛥𝑭𝑘
Noting that the determinant of 𝑲𝑘 is given by
2
det(𝑲𝑘 ) = det(𝑲𝑘−1 )[1 + 𝒒𝑘𝑇 𝒑𝑘 ] , (37.62)
it can be shown that the change in condition number, 𝑐, is
4
[√𝒑𝑇𝑘 𝒑𝑘 𝒒𝑘𝑇 𝒒𝑘 + √𝒑𝑇𝑘 𝒑𝑘 𝒒𝑘𝑇 𝒒𝑘 + 4 (1 + 𝒑𝑇𝑘 𝒒𝑘 )]
𝑐= . (37.63)
2
[4 (1 + 𝒑𝑇𝑘 𝒒𝑘 )]
Following the approach of Matthies and Strang [1979] this condition number is used to
decide whether or not to do a given update. The quasi-Newton condition (37.42) is easily
verified using (37.59) for a real non-singular tangent matrix 𝑲𝑘−1 and it follows that BFGS
preserves not only symmetry but also positive definiteness. Interestingly enough the
converse is not true; if 𝑲𝑘−1 is indefinite then 𝑲𝑘 may still be positive definite.
There is also the option LCPACK = 3, for which LS-DYNA uses a non-symmetric matrix
assembly when solving the linear system of equations. This adds to the computational cost
for each iteration but my improve convergence because of better tangents of certain
features. The non-symmetric solver is illustrated in Figure 3735-3 for simulating a clamped
cantilever beam subject to a follower load. In general the stiffness matrix is symmetric
when the force is derived from an energy potential, or in other words is conservative. This is
for instance the case for a hyperelastic material response or a physically admissible
pressure load, see Schweizerhof and Ramm [1984] for a discussion. Non-symmetry arise
e.g. from non-conservative forces, such as frictional contact, or physically inadmissible
design dependent loads. The example in Figure 3735-3 serves as one of the latter since the
load is not coming from a physical source, such as a water pressure, and it should be seen
as an academic example to prove a point.
Follower load 𝐹
Figure 3735-3 Nonlinear implicit solution of an elastic cantilever beam with a follower
force, von Mises stress is fringed..
The first criterion is simply a user defined upper limit of 𝑘, governed by ILIMIT on
CONTROL_IMPLICIT_SOLUTION. Using ILIMIT = 1 means that no quasi-Newton
updates are performed and should be used for highly nonlinear problems, larger values of
ILIMIT can be used if the nonlinearities are considered less severe. ILIMIT = 11 is the
default and is a reasonable value to use for starters.
A second criterion is that of increased residual norm, i.e., if any of the quantities
𝑇𝑘 = √𝑭𝑘𝑇 𝑱𝑡 𝑭𝑘 (37.65)
or
𝑅𝑘 = √𝑭𝑘𝑇 𝑱𝑟 𝑭𝑘 (37.66)
for some 𝑘 is their respective largest attained since start iterating, the stiffness matrix is
reformed. Here 𝑱𝑡 and 𝑱𝑟 are diagonal matrices with ones or zeros on the diagonal that
extract the translational and rotational degrees of freedom, respectively, and one speaks of
translational or rotational divergence.
criteria available is that the norm of residual 𝑭 must somehow decrease, LSMTD = 2, but
this will inevitably lead to ridiculously small steps and is not recommended or presented
further here. A more useful criterion is instead that an iterate 𝑘 + 1 is accepted if
∣∆𝒙𝑘𝑇 𝑭𝑘+1 ∣ ≤ 𝜀𝑠 ∆𝒙𝑘𝑇 𝑭𝑘 (37.70)
where 𝜀𝑠 is the line search convergence tolerance LSTOL. See Figure 3735-4 for an
illustration. This criterion is derived from a hypothetical assumption of the existence of an
energy potential 𝑊(𝒙𝑘 + 𝑠∆𝒙𝑘 ) for the residual force 𝑭 and that the potential is minimized
along the search direction
𝜕𝑊 𝜕𝑊 𝜕𝒙
= = 𝑭 𝑇 ∆𝒙𝑘 = 0. (37.71)
𝜕𝑠 𝜕𝒙 𝜕𝑠
𝑭 𝑇 ∆𝒙𝑘 (1.60) not solvable
Acceptable interval
according to (37.70)
𝑭𝑘𝑇 ∆𝒙𝑘
𝜀𝑠 𝑭𝑘𝑇 ∆𝒙𝑘
𝑠=1
Negative starting values
indicate negative
eigenvalues in tangent
A closer examination (and numerical experiments) reveals that 𝑭𝑘+1 is not necessarily
bounded by (37.70). Therefore this criterion can be complemented with
𝑇
√𝑭𝑘+1 𝑭𝑘+1 ≤ [1 + 𝜀𝑠 ]√𝑭𝑘𝑇 𝑭𝑘 . (37.72)
This option is invoked with LSMTD = 5 and means that both (37.70) and (37.72) are to be
satisfied for accepting an iterate 𝑘 + 1. While this is a more robust approach, it also
requires more residual force evaluations and is not recommended as default. It has proved
to work well for implicit rubber simulations and complicated contact problems.
Another criterion that must be met for convergence is that simple prescribed motion
constraints on nodes and rigid bodies, if they exist, must be “almost” satisfied. The
background is that only partial line searches (𝑠 < 1) will not satisfy simple motion
constraints, and at least one full line search step (𝑠 = 1) must be accomplished during the
iterations. For difficult problems, this sometimes never happens and therefore convergence
is prevented due to unfulfilled boundary conditions until all prescribed motion is satisfied
to within 1%.
With the automatic time stepper turned on, LS-DYNA will not only cut the time step for
convergence failure but also adjust the time step when converging. The adjustment is
based on a user defined iteration window, i.e., a range of iteration numbers that are
deemed acceptable for convergence, and if the number of iterations for convergence falls
outside this window the next time step will be adjusted as follows. If convergence is
attained for more iterations than acceptable then the time step for the next implicit step is
given by (37.77). If instead convergence is attained for fewer iterations then the next time
step is given by
∆𝑡new = min(∆𝑡max , 100.2 ∆𝑡old ) (37.78)
where ∆𝑡max is a maximum defined step. In this way LS-DYNA narrows in on an optimum
time step for which the number of iterations to converge falls within the specified window.
There is much more information regarding this option in the keyword manual and the user
is referred thereto for practical issues.
𝒙 = 𝒙 𝑗 + 𝛥𝒙. (37.82)
Here, Δ𝑡 is the time step size, 𝛽 and 𝛾 are the free parameters of integration and we have
used ∆𝒙 to denote the total displacement from step 𝑗 to step 𝑗 + 1. For 𝛾 = 1 ⁄ 2 and 𝛽 = 1 ⁄
4 the method becomes the trapezoidal rule and is energy conserving. If
1
𝛾> , (37.83)
2
2
1 1
𝛽 > ( + 𝛾) , (37.84)
4 2
numerical damping is induced into the solution leading to a loss of energy and momentum.
By inserting (37.80) and (37.81) into (37.79) and using (37.82) to eliminate ∆𝒙, we are back to
an equation in 𝒙 only and can apply the algorithm starting with (37.18) and everything
thereafter holds.
procedure. In fact, the eigenvalues of the resulting tangent can be made arbitrarily large by
decreasing ∆𝑡 at the cost of requiring more steps to obtain the solution.
Not only robustness but also the accuracy in dynamic implicit depends on the size of the
time step ∆𝑡, roughly speaking only frequencies up to ~∆𝑡−1 can be resolved. The method
is therefore not perfectly suitable for contact-impact and restitution problems, for contact a
crude rule of thumb is that the time a node or body is in contact should span at least a few
time steps to appropriately resolve the resulting impulse. Having said this, these problems
are still difficult to solve reasonably well.
Consider undamped free vibration in (37.86), i.e, 𝑫 = 𝟎 and 𝑭 = 𝟎, and transform this
equation from time to frequency plane assuming constant 𝑴 and 𝑲. This results in an
eigenvalue problem
𝑲𝒖 = 𝜔2 𝑴𝒖 (37.88)
that can be solved in LS-DYNA for the angular frequency 𝜔 and mode shape 𝒖. The stress
𝝈 in (37.87) affects the frequency 𝜔 in the following way. If the model is in a tensile state,
i.e., the principal stresses are positive, then the eigenvalues of 𝑲 increase compared to a
stress free state and from (37.88) the frequencies will increase. Conversely, the frequencies
will decrease if the principal stresses decrease, this is the effect of tuning a guitar string by
increasing or decreasing its tension. If linearizing with respect to a contact state, the last
term in (37.87) will in effect constrain relative motion between parts or nodes in contact.
More specifically, the relative normal displacement in 𝒖 will be zero, and the relative
tangential motion will be governed by the present stick/slip condition. If in stick mode the
relative tangential displacement in 𝒖 will be zero, and if in slip mode it is unconstrained.
See CONTROL_IMPLICIT_EIGENVALUE for the available options to solve (37.88).
𝑓 = 15 𝐻𝑧
𝑓 40 𝐻 𝑓 43 𝐻
Figure 3735-5 Intermittent eigenvalue analysis of tire. Model shown in top
left, followed by lowest frequency modes for unpressurized tire (top right),
inflated tire (bottom left) and inflated tire and frictional contact (bottom
right). Resultant mode displacements are fringed.
noise and erratic behavior. Noticable is that if 𝑫 = 𝟎 and 𝜇 is an eigenvalue then – 𝜇 is also
an eigenvalue, so then complex eigenvalues come in clusters of four, {𝜇, 𝜇̅̅̅, −𝜇, −𝜇̅̅̅}. Then it
suffice that eigenvalues have nonzero real parts 𝑟𝑗 ≠ 0 for a system to be unstable, which
corresponds to negative eigenvalues in (37.88).
To solve the system (37.91), it is transformed to a first order eigenvalue problem using𝚽 ̃=
𝜇𝚽, resulting in
[ 𝟎 𝑰 ][𝚽] = 𝜇[ 𝑰 𝟎 ][𝚽], (37.94)
−𝑲 −𝑫 𝚽 ̃ 𝟎 𝑴 𝚽 ̃
which can be solved by well-established eigenvalue algorithms. In the output files, LS-
DYNA reports eigenvalues with positive imaginary parts only, i.e., 𝑠𝑗 > 0, and the real and
imaginary parts (when non-zero) of the associated eigenvector 𝚽.
0.02
0.015
0.01
0.005
Damp ratio
-0.005
-0.01
-0.015
-0.02
0 2 4 6 8 10 12 14 16
Mode
Figure 3735-6 Brake squeal application. Pressure pads are applied to a rotating
disc and a nonsymmetric eigenvalue solution reveals friction instabilities. The
damp ratio 𝜗 is plotted at the top as function of mode number, see (37.93). Mode
#6 is the unstable mode and is depicted bottom right with displacements fringed.
34.22
LS-DYNA Theory Manual Arc-length
38
Arc-length
Arc-length methods are available in LS-DYNA for NSOLVR specified between 6 and 9, this
and all other parameters in this Section are located on the *CONTROL_IMPLICIT_SOLU-
TION keyword. These solvers use the Riks/Crisfield methods but unfortunately go under
the old LSDIR.EQ.1 option which makes them somewhat limited in terms of applicability.
For LSDIR.EQ.2 the arc-length method described in Ritto-Corrêa and Camotim [2008] is
implemented for the combination of NSOLVR.EQ.12 and ARCMTH.EQ.3. For this method
the parameters ARCPSI (𝜓), ARCALF (𝛼) and ARCTIM apply, out of which the last simply
tells at what time arc-length is initiated and the first two are to be described in more detail
below. We begin by an explanatory overview of the arc-length method in general for
which we will constantly be referring to Figure 38.7 below. After that the mathematical
details are revealed.
38.1 Overview
An implicit static problem is driven by a parameter 𝑡 referred to as the time, and assuming
that a solution is obtained at 𝑡 = 𝑡𝑛 the objective is to determine the solution given the
constraint 𝑡 = 𝑡𝑛+1 , where 𝑡𝑛+1 is given. We assume that the problem can be associated
with representative force and displacement parameters 𝑓 and 𝑑 and we distinguish between
load- and displacement-driven problems. An example of a load-driven problem is when
𝑓 = 𝑓 (𝑡) is an external load and 𝑑 is the resulting displacement of the associated nodes, and
likewise a displacement-driven problem is when 𝑑 = 𝑑(𝑡) is a prescribed displacement and
𝑓 is the corresponding reaction force. A solution is likely obtained if the problem is stable,
i.e., the force-displacement curve is monotonically increasing, but this method is not
designed to handle limit or turning points. A limit point is illustrated in figure (a) and a
turning point in figure (b), the solution in the next step may not be the one desired or not
even exist, in both cases we want to find the solution that continuously follow the path
corresponding to the force-displacement curve.
The solution for this is to set the stepping parameter 𝑡 free, i.e., replace the constraint 𝑡 =
𝑡𝑛+1 with an arc-length constraint 𝑔(𝑑, 𝑓 ) = 0. In words, this basically means that a
multidimensional sphere (arc) is put around the last converged solution and the next
solution is to be found on that given sphere, the stepping parameter 𝑡 is now a solution
variable. This makes the problem well-posed but unfortunately there are multiple
solutions to the problem, and it may turn out that the wrong solution is found. In figures
(c) and (d), the effect of the arc-length constraints is illustrated and there are two possible
solutions, one feasible that allows us to continue in the right direction and one infeasible
that takes us in the wrong direction. The latter phenomenon is termed doubling back, and
is in practice not easily avoided. Two additional parameters are available that have shown
to improve the robustness in this respect.
a) f b) f
d d
c) f d) f
d d
e) f f) f
d d
g) f h) f
d d
In figure (d), two of the infeasible solutions can in practice be avoided by including the
stepping parameter in the arc-length constraint, thus converting a cylinder to a sphere in
space-time. This is adjusted by the paremeter 0 ≤ 𝜓 < 1, and the constraint reads
(1 − 𝜓)𝑔(𝑑, 𝑓 ) + 𝜓(𝑡 − 𝑡𝑛 )2 = 0, the effect of this constraint is illustrated in figures (e) and
(f), and should be compared with figures (c) and (d). Note that two infeasible solutions are
avoided when comparing figures (d) and (f), it may sometimes be worth using a non-zero
value for 𝜓, e.g., 𝜓 = 0.1.
Another problem is that the feasible and infeasible solution may be too close to the last
converged solution, making the result from the simulation very unpredictable. For this a
parameter 𝛼 is introduced that translates the center of the spatial sphere in the direction of
the linear prediction (i.e., the first Newton iterate of the implicit solution procedure).
Assuming that this prediction is in the direction we want, using 𝛼 < 0 will move the center,
and consequently the infeasible solution, away from where the iterates are taking place. In
addition, the radius of the arc will increase making it less probable to find the incorrect
solution. This option has shown effective in solving snap-through problems when using
small steps to resolve maximum load values, and is illustrated in figures (g) and (h). For
snap-back problems, using 𝛼 = 1 could be an interesting choice since this centers the arc
right between the previously converged point and the first predictor in the arc length
method, thus encouraging the next solution to be found in the reversed direction. An
example of a snap-back problem is shown in Figure 35-9.
that we assume can be divided into a set of independent and dependent variables.
Furthermore we have the time parameter 𝑡 which may serve as the actual time (for
dynamic problems) or just a stepping parameter (for quasi-static problems). The division
into independent and dependent variables is motivated by the constraint equation that
must be fulfilled, i.e.,
𝒉(𝒙, 𝑡) = 𝒉(𝒙𝐼 , 𝒙𝐷 , 𝑡) = 𝟎. (38.96)
From the constraint, the constraint matrix is evaluated as
𝜕𝒉 𝜕𝒉 𝜕𝒉
=[ ] = [𝑪𝐷𝐼 𝑪𝐷𝐷 ], (38.97)
𝜕𝒙 𝜕𝒙𝐼 𝜕𝒙𝐷
which in turn determines the space of trial functions used to establish the nonlinear finite
element equation,
−1 𝑇 𝒓𝐼
[𝑰𝐼𝐼 −(𝑪𝐷𝐷 𝑪𝐷𝐼 ) ][𝒓𝐷 ] = 𝟎, (38.98)
where
𝒓
𝒓(𝒙, 𝑡) = 𝒓(𝒙𝐼 , 𝒙𝐷 , 𝑡) = [𝒓 𝐼 ] (38.99)
𝐷
is the full residual divided into the set of independent and dependent variables. See
previous chapter for further details leading up to (38.98).
−1 𝑇 𝒓𝐼 𝑲𝐼𝐷 −1
𝒓𝐼̂ = [𝑰𝐼𝐼 −(𝑪𝐷𝐷 𝑪𝐷𝐼 ) ] {[𝒓𝐷 ] − [𝑲 ] 𝑪𝐷𝐷 𝒉} ,
𝐷𝐷 (38.105)
⎧ 𝜕𝒓𝐼 ⎫
𝜕𝒓𝐼̂ {
{ ⎡ ⎤ 𝜕𝒉 }
}
= [𝑰𝐼𝐼 −1 𝑇
] ⎢ 𝜕𝑡 ⎥ − [ 𝑲𝐼𝐷 ] 𝑪 −1 ,
−(𝑪𝐷𝐷 𝑪𝐷𝐼 ) ⎨⎢ ⎥
𝜕𝑡 { ⎢𝜕𝒓𝐷 ⎥ 𝑲𝐷𝐷 𝐷𝐷
𝜕𝑡 ⎬
}
{ }
⎩⎣ 𝜕𝑡 ⎦ ⎭
and the independent search direction is given by
𝜕𝒙𝐼𝑠
𝛿𝒙𝐼 = 𝑠𝛿𝒙𝐼𝑠 + 𝛿𝑡 . (38.106)
𝜕𝑡
Here 𝑠 is the line search parameter and
𝛿𝒙𝐼𝑠 = −𝑲̂ 𝐼𝐼−1 𝒓𝐼̂ ,
𝜕𝒙𝐼𝑠 𝜕𝒓 ̂ (38.107)
= −𝑲̂ 𝐼𝐼−1 𝐼 .
𝜕𝑡 𝜕𝑡
The full search direction is completed by computing the dependent part as
𝑠
𝑠 𝜕𝒙𝐷
𝛿𝒙𝐷 = 𝑠𝛿𝒙𝐷 + 𝛿𝑡 , (38.108)
𝜕𝑡
where
𝑠
𝛿𝒙𝐷 −1
= −𝑪𝐷𝐷 𝑪𝐷𝐼 𝛿𝒙𝐼𝑠 − 𝑪𝐷𝐷
−1
𝒉,
𝜕𝒙𝐷𝑠
𝜕𝒙 𝑠 (38.109)
−1 𝜕𝒉
= −𝑪𝐷𝐷−1
𝑪𝐷𝐼 𝐼 − 𝑪𝐷𝐷 .
𝜕𝑡 𝜕𝑡 𝜕𝑡
Finally the new configuration is updated by means of
∆𝒙𝐼(𝑛,𝑖+1) = ∆𝒙𝐼(𝑛,𝑖) + 𝛿𝒙𝐼 ,
(𝑛,𝑖+1) (𝑛,𝑖) (38.110)
∆𝒙𝐷 = ∆𝒙𝐷 + 𝛿𝒙𝐷 ,
∆𝑡(𝑛,𝑖+1) = ∆𝑡(𝑛,𝑖) + 𝛿𝑡.
Upon convergence we set
∆𝒙𝐼(𝑛) = ∆𝒙𝐼(𝑛,𝑖+1) ,
(𝑛) (𝑛,𝑖+1) (38.111)
∆𝒙𝐷 = ∆𝒙𝐷 ,
∆𝑡(𝑛) = ∆𝑡(𝑛,𝑖+1) ,
and hence
𝒙𝐼(𝑛+1) = 𝒙𝐼(𝑛) + ∆𝒙𝐼(𝑛) ,
(𝑛+1) (𝑛) (𝑛) (38.112)
𝒙𝐷 = 𝒙𝐷 + ∆𝒙𝐷 ,
𝑡(𝑛+1) = 𝑡(𝑛) + ∆𝑡(𝑛) .
where
1−𝜓
𝛼𝑥 = 2 𝑇
(1 − 𝛼) (∆𝒙𝐼(0,1) ) ∆𝒙𝐼(0,1)
2 (38.120)
𝜓
𝛼𝑡 = (0,1) (0,1) .
∆𝑡 ∆𝑡
This can be written in terms of a polynomial in 𝑠 and 𝛿𝑡 as
𝑎𝑠𝑠 𝑠2 + 𝑎𝑡𝑡 𝛿𝑡2 + 2𝑎𝑠𝑡 𝑠𝛿𝑡 + 2𝑎𝑠 𝑠 + 2𝑎𝑡 𝛿𝑡 (38.121)
where
𝑎𝑠𝑠 = 𝛼𝑥 (𝛿𝒙𝐼𝑠 )𝑇 𝛿𝒙𝐼𝑠
𝑇
𝜕𝒙 𝑠 𝜕𝒙𝐼𝑠
𝑎𝑡𝑡 = 𝛼𝑥 ( 𝐼 ) + 𝛼𝑡
𝜕𝑡 𝜕𝑡
𝜕𝒙𝐼𝑠
𝑎𝑠𝑡 = 𝛼𝑥 (𝛿𝒙𝐼𝑠 )𝑇
𝜕𝑡
(38.122)
𝜕𝒙 𝑠
𝑎𝑠𝑡 = 𝛼𝑥 (𝛿𝒙𝐼𝑠 )𝑇 𝐼
𝜕𝑡
𝛼 𝑇
𝑎𝑠 = 𝛼𝑥 (∆𝒙𝐼(𝑛,𝑖) − ∆𝒙𝐼(𝑛,1) ) 𝛿𝒙𝐼𝑠
2
𝛼 𝑇 𝜕𝒙 𝑠
𝐼
𝑎𝑡 = 𝛼𝑥 (∆𝒙𝐼(𝑛,𝑖) − ∆𝒙𝐼(𝑛,1) ) + 𝛼𝑡 ∆𝑡(𝑛,𝑖) .
2 𝜕𝑡
For a given line search parameter value, the time increment can have two possible values
Panel Response
1
0.8
0.6
Force
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Displacement
39
Sparse Direct Linear Equation Solvers
LS-DYNA has 5 options for direct solution of the sparse systems of linear equations
that arise in LS-DYNA. All 5 options are based on the multifrontal algorithm [Duff and
Reid, 1983]. Multifrontal is a member of the current generation of sparsity preserving
factorization algorithms that also have very fast computational rates. That is multifrontal
works with a sparsity preserving ordering to reduce the overall size of the direct
factorization and the amount of work it takes to compute that factorization.
LS-DYNA DEV 10/27/16 (r:8004) 36-1 (Sparse Direct Linear Equation Solvers)
Sparse Direct Linear Equation Solvers LS-DYNA Theory Manual
The multifrontal algorithm instead follows a tree of computations where the tree
structure is established by the sparsity preserving orderings, See Figure 39.1. It is this tree
structure that greatly reduces the work required to compute a factorization and the size of
the resulting factorization. At the bottom of the tree, a frontal matrix is assembled with the
original matrix data and those columns that are fully assembled are eliminated. The
remainder of the frontal matrix is updated from the factored columns and passed up the
tree to the parent front in what is called an update matrix. As the computation works its
way up the tree, a frontal matrix is formed by assembling the original matrix data and the
update matrices from its children in the tree. The fully assembled columns are factored
and the remaining columns updated and passed up the tree. At the root (top) of the tree,
the remaining columns are factored.
36-2 (Sparse Direct Linear Equation Solvers) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Sparse Direct Linear Equation Solvers
To parent front
Solver Method
LS-DYNA DEV 10/27/16 (r:8004) 36-3 (Sparse Direct Linear Equation Solvers)
Sparse Direct Linear Equation Solvers LS-DYNA Theory Manual
No.
1 Older implementaion of Solver No. 4. Uses Real*4 arithmetic, has out of
memory capabilities as well as distributed memory parallelism. Only uses
MMD ordering. Was former default method. Retained for backward
compatibility. We recommend switching to Solver No. 4 for improved
performance.
3 Same as 1 except uses Real*8 arithmetic. We recommend switching to Solver
No. 5 or 6 for improved performance.
4 Real*4 implementation of multifrontal which includes automatic out-of-memory
capabilities as well as distributed memory parallelism. Can use either MMD or
METIS orderings. Default method.
5 Real*8 implementation of Solver No. 4.
6 Multifrontal solver from BCSLIB-EXT [Boeing Company, 1999]. Uses Real*8
arithmetic with extensive capabilities for large problems and some Shared
Memory Parallelism.
Can use either MMD or METIS orderings. If the other solvers cannot factor the
problem in the allocated memory, try using this solver.
We strongly recommend using Solvers 4 through 6. Solvers 1 and 3 are included for
backward compatibility with older versions of LS-DYNA but are slower the Solvers 4
through 6. Solvers 4 and 5 are 2 to 6 times faster than the older versions, respectively.
Solver 6 on a single processor computer should be comparable to Solver 5 but has more
extensive capabilities for solving very large problems with limited memory. Solvers 4 and
5 should be used for distributed memory parallel implementations of LS-DYNA. Solver 6
can be used in shared memory parallel.
In an installation of LS-DYNA where both integer and real numbers are stored in 8
byte quantities, then Solvers 1 and 3 are equivalent and Solvers 4 and 5 are equivalent.
The first way LS-DYNA has for preventing such matrix singularities is to add a
small amount of stiffness in the normal direction at each node of every shell element that
has the drilling rotation problem. This “drilling” stiffness matrix is orthogonal to rigid
36-4 (Sparse Direct Linear Equation Solvers) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Sparse Direct Linear Equation Solvers
body motions. The user can control whether this approach is used and how much stiffness
is added via the *CONTROL_IMPLICIT_SOLUTION keyword card. DRLMTH and DRL-
PARM are set in fields 5 and 6. If DRLMTH = 1 then this approach is used. The amount of
stiffness added is controlled via DRLPARM. The default for DRLPARM is 1.0 for linear
problems and 100.0 for nonlinear problems. DRLPARM ∗ .0001 is added in the normal
direction at each node to the diagonal terms associated with the rotational degrees of
freedom for certain types of elemental matrices. For eigenvalue problems the amount of
stiffness added is 1.E-12.
Adding stiffness to handle the drilling rotation problem has been used extensively.
While a robust and reliable approach, its drawback is that the added stiffness may affect
the quality of the computed results. The user can also select not to use this approach and
depend solely on AUTOSPC, the other method for preventing matrix singularities.
LS-DYNA DEV 10/27/16 (r:8004) 36-5 (Sparse Direct Linear Equation Solvers)
LS-DYNA Theory Manual Sparse Eigensolver
40
Sparse Eigensolver
LS-DYNA now includes the Block Shift and Invert Lanczos eigensolver from
BCSLIB-EXT. This eigensolver is used in LS-DYNA to compute the normal modes and
mode shapes for the vibration analysis problem
𝐊𝚽 = 𝐌𝚽𝚲, (40.1)
where 𝐊 and 𝐌are the assembled stiffness and mass matrices, 𝚽 are the eigenvectors
(normal mode shapes) and 𝚲 are the eigenvalues (normal modes).
The Lanczos algorithm iteratively computes a better and better approximation to the
extreme eigenvalues and the corresponding eigenvectors of the ordinary eigenvalue
problem 𝐀𝚽 = 𝚽𝚲where 𝐀 is a real symmetric matrix using only matrix-vector multiplies.
To use Lanczos on the vibration analysis problem it must be changed to
(𝐊 − 𝛔𝐌)−1 𝐌𝚽 = 𝚽𝚯, (40.2)
where each shifted and inverted eigenvalue 𝜃𝑖 = 1/(𝜆𝑖 − 𝜎). This change to an ordinary
eigenvalue problem makes the eigenvalues of the original problem near become the
extreme eigenvalues of the ordinary eigenvalue problem. This helps the Lanczos algorithm
compute those eigenvalues quickly.
function of the speed of the I/O subsystem on the computer as the CPU time. Parallelism
can only speed up the CPU time and does nothing to speed-up the I/O time.
The user can request how many and which eigenvalues to compute using the
keyword *CONTROL_IMPLICIT_EIGENVALUE. Via the parameters on this keyword, the
user can request any of the following problems:
• Compute the lowest 50 modes (that is nearest to zero)
• Compute the 20 modes nearest to 30 Hz.
• Compute the lowest 20 modes between 10 Hz and 50 Hz.
• Compute all of the modes between 10 Hz and 50 Hz.
• Compute all of the modes below 50 Hz.
• Compute the 30 modes nearest to 30 Hz between 10 Hz and 50 Hz.
For a system with a constant angular velocity 𝛚 = {𝜔𝑥 , 𝜔𝑦 , 𝜔𝑧 }T , the body force
added to the applied load is
𝐅B = −𝐌{2𝛚 × 𝐮̇ + 𝛚 × (𝛚 × (𝐫 + 𝐮))}. (40.3)
The damping and stiffness terms are easily derived in matrix form once the cross
product is expressed in matrix form.
0 −𝜔𝑧 𝜔𝑦
⎡ ⎤ 𝑥
⎢ 𝜔𝑧
𝛚 × 𝐫 = 𝛀𝐫 = ⎢ 0 −𝜔𝑥 ⎥
⎥ {𝑦}. (40.4)
⎣−𝜔𝑦 𝜔𝑥 0 ⎦ 𝑧
The linearized equation for vibration is
𝐌𝐮̈ + 𝐂𝐮̇ + [𝐊 + 𝐊σ ]𝐮 = −𝐌{𝛀𝐮̇ + 𝛀2 𝐮}. (40.5)
Rewriting this equation into the traditional form for eigenvalue analysis produces:
𝐌𝐮̈ + 𝐂R 𝐮̇ + 𝐊R 𝐮 = 0
𝐂R = 𝐂 + 𝐌𝛀 (40.6)
𝐌𝐮̈ + 𝐂R 𝐮̇ + 𝐊R 𝐮 = 0.
The inertial contribution to the damping matrix is not symmetric, nor does it fulfill
the requirements for Rayleigh damping, and therefore the resulting eigenvectors and
eigenvalues are complex. The inertial term to the stiffness matrix is, however, symmetric
and it softens the structure, thereby reducing its natural frequencies.
If the damping term is omitted, the matrices are real and symmetric, and the
resulting eigenvalue problem may be solved with the standard eigenvalue methods. The
natural frequencies won’t be correct, but they are typically close enough to the complex
solution that they can be used for initial design calculations.
41
Boundary Element Method
LS-DYNA can be used to solve for the steady state or transient fluid flow about a
body using a boundary element method. The method is based on the work of Maskew
[1987], with the extension to unsteady flow with arbitrary body motion following the work
of Katz and Maskew [1988]. The theory which underlies the method is restricted to
inviscid, incompressible, attached fluid flow. The method should not be used to analyze
flows where shocks or cavitation are present.
In practice the method can be successfully applied to a wider class of fluid flow
problems than the assumption of inviscid, incompressible, attached flow would imply.
Many flows of practical engineering significance have large Reynolds numbers (above 1
million). For these flows the effects of fluid viscosity are small if the flow remains attached,
and the assumption of zero viscosity may not be a significant limitation. Flow separation
does not necessarily invalidate the analysis. If well-defined separation lines exist on the
body, then wakes can be attached to these separation lines and reasonable results can be
obtained. The Prandtl-Glauert rule can be used to correct for non-zero Mach numbers in
air, so the effects of aerodynamic compressibility can be correctly modeled (as long as no
shocks are present).
the assumption of incompressibility implies an infinite sound speed; any disturbance is felt
everywhere in the fluid instantaneously. Although this is not true for real fluids, it is a
valid approximation for a wide class of low-speed flow problems.
Equation (41.1) is solved by discretizing the surface of the body with a set of
quadrilateral or triangular surface segments (boundary elements). Each segment has an
associated source and doublet strength. The source strengths are computed from the free-
stream velocity, and the doublet strengths are determined from the boundary condition.
By requiring that the normal component of the fluid velocity be zero at the center of each
surface segment, a linear system of equations is formed with the number of equations equal
to the number of unknown doublet strengths. When this system is solved, the doublet
strengths are known. The source and doublet distributions on the surface of the body then
completely determine the flow everywhere in the fluid.
The linear system for the unknown doublet strengths is shown in Equation (1.2).
[mic]{𝜇} = {rhs}. (41.2)
In this equation are the doublet strengths, [mic is the matrix of influence
coefficients which relate the doublet strength of a given segment to the normal velocity at
another segment’s mid-point, and rhs is a right-hand-side vector computed from the
known source strengths. Note that mic is a fully-populated matrix. Thus, the cost to
compute and store the matrix increases with the square of the number of segments used to
discretize the surface of the body, while the cost to factor this matrix increases with the
cube of the number of segments. Users should keep these relations in mind when defining
the surface segments. A surface of 1000 segments can be easily handled on most any
computer, but a 10,000 segment representation would not be feasible on any but the most
powerful supercomputers.
The nodes used to define the corners of the boundary element segments must be
ordered to provide a normal vector which points into the fluid (Figure 41.1).
normal
node 4 node 3
node 1 node 2
Figure 41.1. Counter-clockwise ordering of nodes when viewed from fluid
looking towards solid provides unit normal vector pointing into the fluid.
Triangular segments are specified by using the same node for the 3rd and 4th corner
of the segment (the same convention used for shell elements in LS-DYNA). Very large
segments can be used with no loss of accuracy in regions of the flow where the velocity
gradients are small. The size of the elements should be reduced in areas where large
velocity gradients are present. Finite-precision arithmetic on the computer will cause
problems if the segment aspect ratios are extremely large (greater than 1000). The most
accurate results will be obtained if the segments are rectangular, and triangular segments
should be avoided except for cases where they are absolutely required.
41.3 The Neighbor Array The fluid velocities (and, therefore, the fluid
pressures) are determined by the gradient of the velocity potential. On the surface of the
body, this can be most easily computed by taking derivatives of the doublet distribution on
the surface. These derivatives are computed using the doublet strengths on the boundary
element segments. The “Neighbor Array” is used to specify how the gradient is computed
for each boundary element segment. Thus, accurate results will not be obtained unless the
neighbor array is correctly specified by the user.
Each boundary element segment has 4 sides (see Figure 41.2). Side 1 connects the 1st
and 2nd nodes, side 2 connects the 2nd and 3rd nodes, etc. The 4th side is null for
triangular segments.
node 4 node 3
side 3
side 4
side 2
side 1
node 1 node 2
Figure 41.2. Each segment has 4 sides.
For most segments the specification of neighbors is straightforward. For the typical
case a rectangular segment is surrounded by 4 other segments, and the neighbor array is as
shown in Figure 41.3. A biquadratic curve fit is computed, and the gradient is computed as
the analytical derivative of this biquadratic curve fit evaluated at the center of segment j.
There are several situations which call for a different specification of the neighbor
array. For example, boundary element wakes result in discontinuous doublet distributions,
and the biquadratic curve fit should not be computed across a wake. Figure 41.4 illustrates
a situation where a wake is attached to side 2 of segment 𝑗. For this situation two options
exist. If neighbor (2, 𝑗) is set to zero, then a linear computation of the gradient in the side 2
to side 4 direction will be made using the difference between the doublet strengths on
segment 𝑗 and segment neighbor (4, 𝑗). By specifying neighbor (2, 𝑗) as a negative number
the biquadratic curve fit will be retained. The curve fit will use segment 𝑗, segment
neighbor (4, 𝑗), and segment –neighbor (2, 𝑗); which is located on the opposite side of
segment neighbor (4, 𝑗) as segment 𝑗. The derivative in the side 2 to side 4 direction is then
analytically evaluated at the center of segement j using the quadratic curve fit of the
doublet strengths on the three segments shown.
neighbor (3, j)
neighbor(4, j) side 3
segment j neighbor (2, j)
side 4 side 2
side 1
neighbor (1, j)
A final possibility is that no neighbors at all are available in the side 2 to side 4
direction. In this case both neighbor (2, 𝑗) and neighbor (4, 𝑗) can be set to zero, and the
gradient in that direction will be assumed to be zero. This option should be used with
caution, as the resulting fluid pressures will not be accurate for three-dimensional flows.
However, this option is occaisionally useful where quasi-two dimensional results are
desired. All of the above options apply to the side 1 to side 3 direction in the obvious ways.
For triangular boundary element segments side 4 is null. Gradients in the side 2 to
side 4 direction can be computed as described above by setting neighbor (4, 𝑗) to zero (for a
linear derivative computation) or to a negative number (to use the segment on the other
side of neighbor (2, 𝑗) and a quadratic curve fit). There may also be another triangular
segment which can be used as neighbor (4, 𝑗) (see Figure 41.5).
41.4 Wakes
Wakes should be attached to the boundary element segments at the trailing edge of
a lifting surface (such as a wing, propeller blade, rudder, or diving plane). Wakes should
also be attached to known separation lines (such as the sharp leading edge of a delata wing
neighbor(4, j)
segment j
side 2
at high angles of attack). Wakes are required for the correct computation of surface
LS-DYNA DEV 10/27/16 (r:8004) 38-5 (Boundary Element Method)
Boundary Element Method LS-DYNA Theory Manual
pressures for these situations. As described above, two segments on opposite sides of a
wake should never be used as neighbors. Correct specification of the wakes is required for
accurate results.
Wakes convect with the free-stream velocity. The number of segments in the wake
is controlled by the user, and should be set to provide a total wake length equal to 5-10
times the characteristic streamwise dimension of the surface to which the wake is attached.
For example, if the wake is attached to the trailing edge of a wing whose chord is 1, then
the total length of the wake should at least 5, and there is little point in making it longer
than 10. Note that each wake segment has a streamwise length equal to the magnitude of
the free stream velocity times the time increment between calls to the Boundary Element
Method routine. This time increment is the maximum of the LS-DYNA time step and
DTBEM specified on Card 1 of the BEM input. The influence coefficients for the wake
segments must be recomputed for each call to the Boundary Element Method, but these
influence coefficients do not enter into the matrix of influence coefficients which must be
factored.
The parameter DTBEM on Card 1 of the BEM input is used to control the time
increment between calls to the boundary element method. The fluid pressures computed
during the last call to the BEM will continue to be used for subsequent LS-DYNA iterations
until DTBEM has elapsed.
A further reduction in execution time may be obtained for some applications using
the input parameter IUPBEM. This parameter controls the number of calls to the BEM
routine between computation (and factorization) of the matrix of influence coefficients
(these are time-consuming procedures). If the motion of the body is entirely rigid body
motion there is no need to recompute and factor the matrix of influence coefficients, and
the execution time of the BEM can be significantly reduced by setting IUPBEM to a very
large number. For situations where the motion of the body is largely rigid body motion
with some structural deformation an intermediate value (e.g. 10) for IUPBEM can be used.
It is the user’s responsibility to verify the accuracy of calculations obtained with IUPBEM
greater than 1.
The final parameter for controlling the execution time of the boundary element
method is FARBEM. The routine which calculates the influence coefficients switches
between an expensive near-field and an inexpensive far-field calculation depending on the
distance from the boundary element segment to the point of interest. FARBEM is a
nondimensional parameter which determines where the far-field boundary lies. Values of
FARBEM of 5 and greater will provide the most accurate results, while values as low as 2
will provide slightly reduced accuracy with a 50% reduction in the time required to
compute the matrix of influence coefficients.
42
SPH
𝑊(𝐱, ℎ) should be a centrally peaked function. The most common smoothing kernel
used by the SPH community is the cubic B-spline which is defined by choosing 𝜃 as:
⎧ 3 3
{
{ 1 − 𝑢2 + 𝑢3 for |𝑢| ≤ 1
{ 2 4
𝜃(𝑢) = 𝐶 × ⎨1 3 (42.3)
{ (2 − 𝑢) for 1 ≤ |𝑢| ≤ 2
{
{4
⎩0 for 2 < |𝑢|
where C is a constant of normalization that depends on the number of space dimensions.
The SPH method is based on a quadrature formula for moving particles ((𝐱𝑖 (𝑡)) 𝑖 ∈
{1. . 𝑁}, where 𝐱𝑖 (𝑡) is the location of particle 𝑖, which moves along the velocity field v.
The SPH formalism implies a derivative operator. A particle approximation for the
derivative operator must be defined. Before giving the definition of this approximation,
we define the gradient of a function as:
∇𝑓 (𝑥) = ∇𝑓 (𝑥) − 𝑓 (𝑥)∇1(𝑥), (42.5)
where 1 is the unit function.
Starting from this relation, we can define the particle approximation to the gradient
of a function:
𝑁 𝑚𝑗
Πℎ ∇𝑓 (𝐱𝑖 ) = ∑ [𝑓 (𝐱𝑗 )𝐴𝑖𝑗 − 𝑓 (𝐱𝑖 )𝐴𝑖𝑗 ], (42.6)
𝑗=1
𝜌𝑗
1 ||𝐱𝑖 −𝐱𝑗 ||
where 𝐴𝑖𝑗 = 𝜃′( ℎ ).
ℎ𝑑+1
We can also define the particle approximation of the partial derivative ∂𝑥∂𝛼:
𝑁
∂𝑓
Π ( 𝛼 )(𝐱𝑖 ) = ∑ 𝑤𝑗 𝑓 (𝐱𝑗 𝐴𝛼 (𝐱𝑖 , 𝐱𝑗 ),
ℎ
(42.7)
∂𝑥 𝑗=1
A discrete adjoint operator for the partial derivative is also necessary, and is taken to
be the 𝛼 − 𝑡ℎ component of the operator:
𝑁
𝐷∗𝛼 𝜙(𝐱𝑖 ) = ∑ 𝑤𝑗 𝜙(𝐱𝑗 )𝐴𝛼 (𝐱𝑖 , 𝐱𝑗 ) − 𝑤𝑗 𝜙(𝐱𝑖 )𝐴𝛼 (𝐱𝑗 , 𝐱𝑖 ) (42.15)
𝑗=1
These definitions are leading to a conservative method. Hence, all the conservative
equations encountered in the SPH method will be solved using the weak form.
𝑑𝐯𝛼 𝑁
⎛ 𝜎 𝛼,𝛽 (𝐱𝑖 ) 𝜎 𝛼,𝛽 (𝐱𝑗 ) ⎞
(𝐱 ) = ∑ 𝑚𝑗 ⎜ 𝐴𝑖𝑗 − 𝐴𝑗𝑖 ⎟. (42.17)
𝑑𝑡 𝑖 𝑗=1 ⎝ 𝜌𝑖
2 𝜌2𝑗 ⎠
For example, if we choose the smoothing function to be symmetric, this can lead to
the following equation:
𝛼,𝛽
𝑑𝐯𝛼 𝑁
⎛ 𝜎 𝛼,𝛽 (𝐱𝑖 ) 𝜎 (𝐱𝑗 )⎞
(𝐱 ) = ∑ 𝑚𝑗 ⎜ + ⎟ 𝐴𝑖𝑗 . (42.20)
𝑑𝑡 𝑖 𝑗=1 ⎝ 𝜌𝑖
2 𝜌2𝑗 ⎠
This is what we call the “symmetric formulation”, which is chosen in the *CONTROL_SPH
card (IFORM = 2).
𝛼,𝛽
𝑑𝐯𝛼 𝑁
⎛ 𝜎 𝛼,𝛽 (𝐱𝑖 ) 𝜎 (𝐱𝑗 ) ⎞
(𝐱𝑖 ) = ∑ 𝑚𝑗 ⎜ 𝐴𝑖𝑗 − 𝐴𝑗𝑖 ⎟. (42.21)
𝑑𝑡 𝑗=1 ⎝ 𝜌𝑖 𝜌𝑗 𝜌𝑖 𝜌𝑗 ⎠
This is the “fluid formulation” invoked with IFORM = 5 which gives better results
than other SPH formulations when fluid material are present, or when material with very
different stiffness are used.
42.2 Sorting
In the SPH method, the location of neighboring particles is important. The sorting
consists of finding which particles interact with which others at a given time. A bucket sort
is used that consists of partitioning the domain into boxes where the sort is performed.
With this partitioning the closest neighbors will reside in the same box or in the closest
boxes. This method reduces the number of distance calculations and therefore the CPU
time.
The notation 𝑋 ̅̅̅ ̅𝑖𝑗 = 1 (𝑋𝑖 + 𝑋𝑗 ) has been used for median between 𝑋𝑖 and 𝑋𝑗 , 𝑐 is the
2
adiabatic sound speed, and
⎧ 𝑣𝑖𝑗 𝑟𝑖𝑗
{ℎ̅𝑖𝑗 2 if 𝑣𝑖𝑗 𝑟𝑖𝑗 < 0
𝜇𝑖𝑗 = ⎨ 𝑟𝑖𝑗 + 𝜂2 (42.23)
{
⎩ 0 otherwise
Here, 𝑣𝑖𝑗 = (𝑣𝑖 − 𝑣𝑗 ), and 𝜂 = 0.01ℎ̅𝑖𝑗 which prevents the denominator from vanishing.
2 2
Physical Properties:
The mass, density, constitutive laws are defined in the ELEMENT_SPH and the
PART cards.
Geometrical Properties:
39-6 (SPH) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual SPH
The geometrical properties of the model concern the way particles are initially
placed. Two different parameters are to be fixed: Δ𝑥𝑖 lengths and the CSLH coefficient.
These parameters are defined in the SECTION_SPH card.
A proper SPH mesh must satisfy the following conditions: it must be as regular as
possible and must not contain too large variations.
For instance, if we consider a cylinder SPH mesh, we have at least two possibilities:
43
Element-Free Galerkin
where 𝑛 is the order of completeness in this approximation, the monomial 𝐻𝑖 (𝐱) are basis
functions, and 𝑏𝑖 (𝐱) are the coefficients of the approximation.
The coefficients 𝑏𝑖 (𝐱) at any point 𝐱 are depending on the sampling points 𝐱𝐼 that are
collected by a weighting function 𝑤𝑎 (𝐱 − 𝐱𝐼 ). This weighting function is defined to have a
compact support measured by ‘a’, i.e., the sub-domain over which it is nonzero is small
relative to the rest of the domain. Each sub-domain ΔΩ𝐼 is associated with a node 𝐼. The
most commonly used sub-domains are disks or balls. A typical numerical model is shown
in Figure 43.1.
LS-DYNA DEV 10/27/16 (r:8004) 40-1 (Element-Free Galerkin)
Element-Free Galerkin LS-DYNA Theory Manual
ΩI
In this development, we employ the cubic B-spline kernel function as the weighting
function:
⎧2 2 3
{ − 4 (‖𝐱 − 𝐱𝐼 ‖) + 4 (‖𝐱 − 𝐱𝐼 ‖) ‖𝐱 − 𝐱𝐼 ‖ 1 ⎫
{3 for 0 ≤ ≤ }
{ 𝑎 𝑎 𝑎 2}}
{ }
{4 ‖𝐱 − 𝐱𝐼 ‖ ‖𝐱 − 𝐱𝐼 ‖ 2 4 ‖𝐱 − 𝐱𝐼 ‖ 2 1 ‖𝐱 − 𝐱𝐼 ‖ }
𝑤𝑎 (𝐱 − 𝐱𝐼 ) = ⎨ − 4 ( ) + 4( ) − ( ) for < ≤ 1⎬ (43.2)
{3 𝑎 𝑎 3 𝑎 2 𝑎 }
{ }
{ }
{0 otherwise }
{ }
⎩ ⎭
where NP is the number of nodes within the support of 𝐱 for which the weighting function
𝑤𝑎 (𝐱 − 𝐱𝐼 ) ≠ 0.
{𝐇(𝐱1 )}T
⎡ ⎤
𝐇 = ⎢⋯ ⎥, (43.6)
T
⎣{𝐇(𝐱NP )} ⎦
free shape functions in the Galerkin approximation, however, does not guarantee a linear
exactness in the solution of the Galerkin method. It has been shown by Chen et al. [2001]
that two integration constraints are required for the linear exactness solution in the
Galerkin approximation.
NIT
∑ ∇ΨI (x̂L )AL = 0 for {𝐼: supp(Ψ𝐼 ) ∩ Γ = 0}, (43.16)
𝐿=1
NIT NITh
∑ ∇ΨI (x̂L )AL = ∑ nΨ𝐼 (x̃𝐿 )𝑠𝐿 for {𝐼: supp(Ψ𝐼 ) ∩ Γℎ ≠ 0}. (43.17)
𝐿=1 𝐿=1
where Γℎ is the natural boundary, Γ is the total boundary, 𝐧 is the surface normal on Γℎ , x̂𝐿
and 𝐴𝐿 are the spatial co-ordinate and weight of the domain integration point, respectively,
x̃𝐿 and 𝑠𝐿 are the spatial co-ordinate and weight of the domain of natural boundary
integration point, respectively, NIT is the number of integration points for domain
integration and NITh is the number of integration points for natural boundary integration.
It can be shown that the smoothed EFG shape function gradient ∇ ̃ Ψ𝐼 (xL ) meets the
integration constraints in Equations (43.16) and (43.17) regardless of the numerical
integration employed.
To introduce the Lagrangian EFG shape function into the approximation of a path-
dependent problem, the strain increment Δ𝑢𝑖,𝑗 is computed by
∂Δ𝑢𝑖 ∂Δ𝑢𝑖 −1 −1
Δ𝑢𝑖,𝑗 = = 𝐹 = Δ𝐹𝑖𝑘 𝐹𝑘𝑗 . (43.21)
∂𝑥𝑗 ∂𝑋𝑘 𝑘𝑗
Following the derivation for explicit time integration, the equations to be solved
have the form
δ𝐮T 𝐌𝐮̈ = δ𝐮T 𝐑, (43.28)
where
̈ 𝑑2𝐼̈ , 𝑑3𝐼̈ ]𝑇
𝐮̈𝐼 = [𝑑1𝐼,
where B̃𝑇𝐼 (x) is the smoothed gradient matrix obtained from Equation (43.22), 𝑑𝑖𝐼 is the
coefficient of the approximation or the “generalized” displacement.
This is because, in general, the mesh-free shape functions are not interpolation
functions. As a result, a special treatment is required to enforce essential boundary
conditions. There are many techniques for mesh-free methods to impose the essential
boundary condition. Here, we adopt the transformation method as originally proposed for
the RKPM method by Chen et al. [1996].
and
̂ = 𝐀−T 𝐌𝐀−1 ; 𝐅̂ int = 𝐀−T 𝐅int .
𝐌 (43.34)
The nodes are partitioned into three groups: a boundary group 𝐺𝐵1 which contains
all the nodes subjected to kinematic constraints; group 𝐺𝐵2 which contains all the nodes
whose kernel supports cover nodes in group 𝐺𝐵1 ; and internal group 𝐺𝐼 which contains the
rest of nodes. Nodes numbers are re-arranged in the following order in the generalized
displacement vector:
u𝐵1
𝐮 = ⎢u𝐵2 ⎤
⎡
⎥ (43.35)
⎣u𝐼 ⎦
where 𝐮𝐵1 , 𝐮𝐵2 and 𝐮𝐼 are the generalized displacement vectors associated with groups 𝐺𝐵1 ,
𝐺𝐵2 and 𝐺𝐼 respectively. The transformation in Equation (43.32) is also re-arranged as
𝐵 BB
𝐮̂ = [𝐮̂𝐼 ] [Λ IB Λ BI ] [𝐮𝐵 ] ≡ 𝚲
̂ 𝐮, (43.36)
𝐮̂ Λ Λ II 𝐮𝐼
where
𝐵1 𝐵1 B1 B1
𝐮̂B = [û 𝐵 ] ; 𝐮B = [u𝐵 ] ; 𝚲BB = [Λ B 𝐵 Λ B1B2 ] ; 𝚲BI = [0 ] ; 𝚲IB
û 2 u 2 Λ 2 1 Λ B2B2 Λ𝐵2𝐼 (43.37)
= [Λ𝐼𝐵1 Λ𝐼𝐵2 ].
Using the mixed coordinates in Equation (43.38), the transformed discrete Equation
(43.31) becomes
T T
δ𝐮∗ 𝐌∗ 𝐮̈∗ = δ𝐮∗ 𝐑∗ , (43.40)
where
−T −1 −T
𝐌∗ = 𝐀∗ 𝐌𝐀∗ ; 𝐑 ∗ = 𝐀∗ 𝐑. (43.41)
The computation in Equations (43.41) is much less intensive than that in Equation
(43.31), especially when the number of boundary and contact nodes is much smaller than
the number of interior nodes.
Projection
ξ
The mesh-free shape functions are then defined with those locally projected
coordinates of the nodes
Ψ𝐼 (𝐗) = Ψ𝐼 (𝑥̂, 𝑦̂). (43.44)
However, the shape functions obtained directly above are non-conforming, i.e.
^
zi
z¯ I
y¯ I ^i
y
i M-plane
M-plane
I x¯ I ^xi
J J K
When the shell structure degenerates to a plate, the constant stress condition cannot
be recovered. To remedy this problem, an area-weighed smoothing across different
projected planes is used to obtain the conforming shape functions that are given by
NIE
∑𝑖=1 Ψ𝐼 (𝑥̂𝑖 , 𝑦̂𝑖 )𝐴𝑖
̃𝐼 (𝐗) = Ψ
Ψ ̃𝐼 (𝑥̂, 𝑦̂) = . (43.46)
𝑁𝐼𝐸
∑𝑖=1 𝐴𝑖
where NIE is the number of surrounding projected planes that can be evaluated at point X,
𝑨𝒊 is the area of the element 𝑖, and (𝑥̂𝑖 , 𝑦̂𝑖 ) is the local coordinates of point X in the projected
plane 𝑖.
With this smoothing technique, we can prove that the modified shape functions satisfy at
least the partition of unity property in the general shell problems. This property is
important for the shell formulation to preserve the rigid-body translation.
When the shell degenerates to a plate, we can also prove that the shape functions
obtained from this smoothing technique will meet the n-th order completeness condition as
NP
𝑖 𝑘𝑗 𝑗
̃𝐼 (𝐗)𝑋1𝐼
∑Ψ 𝑋2𝐼 𝑋3𝐼 = 𝑋1𝑖 𝑋2 𝑋3𝑘 , 𝑖 + 𝑗 + 𝑘 = 𝑛. (43.47)
𝐼=1
This is a necessary condition for the plate to pass the constant bending patch test.
V3 ζ
^
z
y^
X x^
z
x̄
y
x
Figure 43.4. Geometry of a shell.
surface is the mid-surface of the shell, the global coordinates and displacements at an
arbitrary point within the shell body are given by
ℎ
𝐱 = 𝐱̅ + ζ 𝐕3 , (43.48)
2
ℎ
𝐮=𝐮 ̅̅̅̅ + ζ 𝐔. (43.49)
2
where 𝐱̅ and 𝐮 ̅̅̅̅ are the position vector and displacement of the reference surface,
respectively. 𝐕3 is the fiber director and 𝐔 is the displacement resulting from the fiber
rotation (see Figures 43.4 and 43.5). ℎ is the length of the fiber.
With the mesh-free approximation, the motion and displacements are given by
𝑁𝑃 𝑁𝑃
𝜁 ℎ𝐼
̃𝐼 (𝜉 , 𝜂)𝐱𝐼 + ∑ Ψ
𝐱(𝜉 , 𝜂, 𝜁 ) = 𝐱̅(𝜉 , 𝜂) + 𝐕(𝜉 , 𝜂, 𝜁 ) ≈ ∑ Ψ ̃𝐼 (𝜉 , 𝜂) 𝐕 , (43.50)
𝐼=1 𝐼=1
2 3𝐼
Initial configuration
u U
V03
X
x V3
X̄ ū
x̄
Deformed Configuration
Figure 43.5. Deformation of a shell.
𝑁𝑃 𝑁𝑃 𝛼
𝜁 ℎ𝐼
𝐮(𝜉 , 𝜂, 𝜁 ) = 𝐮 ̃𝐼 (𝜉 , 𝜂)𝐮𝐼 + ∑ Ψ
̅̅̅̅(𝜉 , 𝜂) + 𝐔(𝜉 , 𝜂, 𝜁 ) ≈ ∑ Ψ ̃𝐼 (𝜉 , 𝜂) [−𝐕2𝐼 𝐕1𝐼 ] {𝛽𝐼 },(43.51)
𝐼=1 𝐼=1
2 𝐼
where 𝐱𝐼 and 𝐮𝐼 are the global coordinates and displacements at mesh-free node 𝐼,
respectively. 𝐕3𝐼 is the unit vector of the fiber director and 𝐕1𝐼 , 𝐕2𝐼 are the base vectors of
the nodal coordinate system at node 𝐼. 𝛼𝐼 and 𝛽𝐼 are the rotations of the director vector 𝐕3𝐼
about the 𝐕1𝐼 and 𝐕2𝐼 axes. ℎ𝐼 is the thickness. The variables with a superscripted bar refer
̃𝐼 is the 2D mesh-free shape functions constructed based on one
to the shell mid-surface. Ψ
of the two mesh-free surface representations described in the previous section, with (𝜉 , 𝜂)
either the parametric coordinates or local coordinates of the evaluated point.
The local co-rotational coordinate system (𝑥̂, 𝑦̂, 𝑧̂) is defined at each integration point
on the shell reference surface, with 𝑥̂ and 𝑦̂ tangent to the reference surface and 𝑧̂ in the
thickness direction (see Figure 43.6). The base vectors are given as
𝐱,ξ 𝐱,ξ × 𝐱,η
𝐞̂1 = , 𝐞̂3 = , 𝐞̂2 = 𝐞̂3 × 𝐞̂1 . (43.52)
∥𝐱,ξ ∥ ∥𝐱,ξ × 𝐱,η ∥
The rotation of the fiber director is then obtained from the global rotations:
𝛼 𝐕T
{𝛽} = [ 1T ] Δθ, Δθ = [Δ𝜃1 Δ𝜃2 Δ𝜃3 ]T . (43.54)
𝐕 2
z^ z^s y^s
V3 y^
V2
x^ s
V1
x^
z
y
x
Figure 43.6. Local co-rotational and nodal coordinate systems.
In the local co-rotational coordinate system, the motion and displacements are
approximated by the mesh-free shape functions
NP NP
hI
̃I x̂iI + ζ ∑ Ψ
x̂i = ∑ Ψ ̃I 𝑉̂ , (43.55)
I=1 I=1
2 3𝑖𝐼
NP NP
hI α
û i = ∑ Ψ ̃I
̃I ûiI + ζ ∑ Ψ ̂ 2iI
[−V ̂ 1iI ] { I }.
V (43.56)
2 β I
I=1 I=1
where the smoothed strain operators are calculated by averaging the consistent strain
operators over an area around the evaluated point
1 1 1
̃m
𝐁 𝐼 (𝐱𝑙 ) = ∫ 𝐁 ̂m ̃b
𝐼 𝑑𝐴 , 𝐁𝐼 (𝐱𝑙 ) = ∫ 𝐁 ̂ b𝐼 𝑑𝐴 , 𝐁
̃ s𝐼 (𝐱𝐿 ) = ∫ 𝐁 ̂ s𝐼 𝑑𝐴, (43.58)
𝐴𝑙 Ω𝑙 𝐴𝑙 Ω𝑙 𝐴𝐿 Ω𝐿
with
̂m
𝐁 𝐼
−1 ̃ ℎ𝐼 −1 ̃ ℎ𝐼
⎡Ψ̃𝐼,𝑥 0 0 −𝐽13 Ψ𝐼 𝑉̂ 𝐽13 Ψ𝐼 𝑉̂ ⎤
⎢ 2 2𝑥𝐼 2 1𝑥𝐼 ⎥
⎢ ⎥ (43.59)
=⎢ ̃𝐼,𝑦 −1 ̃ ℎ𝐼 ̂ −1 ̃ ℎ𝐼 ̂ ⎥,
⎢0 Ψ 0 −𝐽23 Ψ𝐼 𝑉2𝑦𝐼
2
𝐽23 Ψ𝐼 𝑉1𝑦𝐼
2 ⎥
⎢ ⎥
⎢ −1 ̃ ℎ𝐼 ̂ −1 ̃ ℎ𝐼 ̂ −1 ̃ ℎ𝐼 ̂ −1 ̃ ℎ𝐼 ̂ ⎥
̃ ̃𝐼,𝑥
⎣Ψ𝐼,𝑦 Ψ 0 −𝐽23 Ψ𝐼 𝑉2𝑥𝐼 − 𝐽13
2
Ψ𝐼 𝑉2𝑦𝐼
2
𝐽23 Ψ𝐼 𝑉1𝑥𝐼 + 𝐽13 Ψ𝐼 𝑉1𝑦𝐼 ⎦
2 2
Γl
Ωl ζ
ξ
xl η
ΓL
xL
ℎ𝐼 ℎ𝐼
⎡0 0 0 ̃𝐼,𝑥
−Ψ 𝑉̂ ̃𝐼,𝑥
Ψ 𝑉̂ ⎤
⎢ 2 2𝑥𝐼 2 1𝑥𝐼 ⎥
b
⎢ ℎ ℎ𝐼 ⎥
̂ 𝐼 = ⎢0
𝐁 0 0 ̃𝐼,𝑦 𝐼 𝑉̂2𝑦𝐼
−Ψ ̃𝐼,𝑦 𝑉̂1𝑦𝐼
Ψ ⎥, (43.60)
⎢ 2 2 ⎥
⎢ ⎥
⎢ ℎ𝐼 ℎ ℎ𝐼 ℎ𝐼 ⎥
̃𝐼,𝑥 𝐼 𝑉̂2𝑦𝐼
̃𝐼,𝑦 𝑉̂2𝑥𝐼 − Ψ ̃ ̂ ̃ ̂
⎣0 0 0 −Ψ
2 2
Ψ𝐼,𝑦 𝑉1𝑥𝐼 + Ψ𝐼,𝑥 𝑉1𝑦𝐼 ⎦
2 2
−1 ̃ ℎ𝐼 −1 ̃ ℎ𝐼 ̂ −1 ̃ ℎ𝐼 −1 ̃ ℎ𝐼 ̂
⎡0 0 ̃𝐼,𝑦
Ψ −𝐽33 Ψ𝐼 𝑉̂2𝑦𝐼 − 𝐽23 Ψ𝐼 𝑉2𝑧𝐼 𝐽33 Ψ𝐼 𝑉̂1𝑦𝐼 + 𝐽23 Ψ𝐼 𝑉1𝑧𝐼 ⎤
̂ s𝐼 = ⎢
𝐁 2 2 2 2 ⎥, (43.61)
⎢
⎢ ⎥
⎥
̃𝐼,𝑥 −1 ̃ ℎ𝐼 ̂ −1 ̃ ℎ𝐼 ̂ −1 ̃ ℎ𝐼 ̂ −1 ̃ ℎ𝐼 ̂
⎣0 0 Ψ −𝐽33 Ψ𝐼 𝑉2𝑥𝐼 − 𝐽13 Ψ𝐼 𝑉2𝑧𝐼
2 2
𝐽33 Ψ𝐼 𝑉1𝑥𝐼 + 𝐽13 Ψ𝐼 𝑉1𝑧𝐼 ⎦
2 2
and 𝐉−1 is the inverse of the Jacobian matrix at the integration point. The local degrees-of-
freedom are
𝐝̂ 𝐼 = [𝑢̂𝑥𝐼 𝑢̂𝑦𝐼 𝑢̂𝑧𝐼 𝛼𝐼 𝛽𝐼 ]T . (43.62)
The internal nodal force vector is
T T
𝐅̂Iint = ∫ 𝐁
̃m ̃b ̃sT
I σ̂ dΩ + ∫ ζ𝐁I σ̂ dΩ + ∫ 𝐁I σ̂ dΩ. (43.63)
Ω Ω Ω
The above integrals are calculated with the local boundary integration method.
Each background finite element is divided into four integration zones, shown as Ω𝑙 in
Figure 43.7. In order to avoid shear locking in the analysis of thin shells, the shear term
(third term in Eq. (43.63)), should be under-integrated by using one integration zone in
each background element (Ω𝐿 in Figure 43.7). Accordingly, the co-rotational coordinate
systems are defined separately at the center of each integration zone, as shown in Figure
43.6.
The use of the updated Lagrangian formulation implies that the reference coordinate
system is defined by the co-rotational system in the configuration at time t. Therefore, the
40-14 (Element-Free Galerkin) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Element-Free Galerkin
local nodal force and displacement vectors referred to this coordinate system must be
transformed to the global coordinate system prior to assemblage.
44
Linear shells
Plate
Membrane
where the complete details of the element are provided. A condensed overview is given
here. The shell theory makes the following assumptions:
The fiber remains straight and inextensible
The local x and y rotations of the shell are interpolated from the equations:
4 8
𝜃𝑥 (𝑟, 𝑠) = ∑ 𝑁𝑖 (𝑟, 𝑠)𝜃𝑥𝑖 + ∑ 𝑁𝑖 (𝑟, 𝑠)Δ𝜃𝑥𝑖
𝑖=1 𝑖=5
(44.1)
4 8
𝜃𝑦 (𝑟, 𝑠) = ∑ 𝑁𝑖 (𝑟, 𝑠)𝜃𝑦𝑖 + ∑ 𝑁𝑖 (𝑟, 𝑠)Δ𝜃𝑦𝑖 ,
𝑖=1 𝑖=5
where nodes 5-8 are at the mid side of the element. The interpolation functions are given
by
1 1
𝑁1 = (1 − 𝑟)(1 − 𝑠) 𝑁5 = (1 − 𝑟2 )(1 − 𝑠)
4 2
1 1
𝑁2 = (1 + 𝑟)(1 − 𝑠) 𝑁6 = (1 + 𝑟)(1 − 𝑠2 )
4 2
(44.2)
1 1
𝑁3 = (1 + 𝑟)(1 + 𝑠) 𝑁7 = (1 − 𝑟2 )(1 + 𝑠)
4 2
1 1
𝑁4 = (1 − 𝑟)(1 + 𝑠) 𝑁8 = (1 − 𝑟)(1 − 𝑠2 ).
4 2
In his formulation, Wilson resolves the rotation of the mid side node into tangential and
normal components relative to the shell edges. The tangential component is set to zero
leaving the normal component as the unknown, which reduces the rotational degrees-of-
freedom from 16 to 12, see Figure 44.2.
Δ𝜃𝑥 = sin𝛼𝑖𝑗 Δ𝜃𝑖𝑗
(44.3)
Δ𝜃𝑦 = −cos𝛼𝑖𝑗 Δ𝜃𝑖𝑗
Θj
ΔΘy
ΔΘx
i = 1, 2, 3, 4
ΔΘij j = 2, 3, 4, 1
αij m =5, 6, 7, 8
Θi
Figure 44.2. Element edge [Wilson, 2000]
4 8
𝜃𝑥 (𝑟, 𝑠) = ∑ 𝑁𝑖 (𝑟, 𝑠)𝜃𝑥𝑖 + ∑ 𝑀𝑥𝑖 (𝑟, 𝑠)Δ𝜃𝑖
𝑖=1 𝑖=5
(44.4)
4 8
𝜃𝑦 (𝑟, 𝑠) = ∑ 𝑁𝑖 (𝑟, 𝑠)𝜃𝑦𝑖 + ∑ 𝑀𝑦𝑖 (𝑟, 𝑠)Δ𝜃𝑖 .
𝑖=1 𝑖=5
Ultimately, the 4 mid side rotations are eliminated by using static condensation, a
procedure that makes this shell very costly if used in explicit calculations.
The local 𝑥 and 𝑦 displacements relative to the mid surface are functions of the 𝑧-
coordinate and rotations:
𝑢𝑥 (𝑟, 𝑠) = 𝑧𝜃𝑦 (𝑟, 𝑠)
(44.5)
𝑢𝑦 (𝑟, 𝑠) = −𝑧𝜃𝑥 (𝑟, 𝑠).
Wilson shows, where it is assumed that the normal displacement along each side is
cubic, that the transverse shear strain along each side is given by,
1 1 2
𝛾𝑖𝑗 = (𝑢𝑧𝑗 − 𝑢𝑧𝑖 ) − (𝜃𝑖 + 𝜃𝑗 ) − Δ𝜃𝑖𝑗 , (44.6)
𝐿 2 3
which can be rewritten, referring to Figure 44.3 as:
1 sin𝛼𝑖𝑗 cos𝛼𝑖𝑗 2
𝛾𝑖𝑗 = (𝑢𝑧𝑗 − 𝑢𝑧𝑖 ) − (𝜃𝑥𝑖 + 𝜃𝑥𝑗 ) + (𝜃𝑦𝑖 + 𝜃𝑦𝑗 ) − Δ𝜃𝑖𝑗 , (44.7)
𝐿 2 2 3
The nodal shears are then written in terms of the side shears as
𝛾 cos𝛼𝑖𝑗 sin𝛼𝑖𝑗 𝛾𝑥𝑧
[𝛾𝑖𝑗 ] = [ ] [ ], (44.8)
𝑘𝑖 cos𝛼𝑘𝑖 sin𝛼𝑘𝑖 𝛾𝑦𝑧
which can be inverted to obtain the nodal shears:
y
αki Θy
Θx
k
j
γki γij
ΔΘki
i = 1, 2, 3, 4
ΔΘij j = 2, 3, 4, 1
γxz αij m =5, 6, 7, 8
i γyz x
The standard bilinear basis functions are used to interpolate the nodal shears to the
integration points.
The inplane displacement field for the 8 node membrane is interpolated, using the
serendipity shape functions with the mid-side relative displacements, from:
4 8
𝑢𝑥 (𝑟, 𝑠) = ∑ 𝑁𝑖 (𝑟, 𝑠)𝑢𝑥𝑖 + ∑ 𝑁𝑖 (𝑟, 𝑠)Δ𝑢𝑥𝑖
𝑖=1 𝑖=5
(44.10)
4 8
𝑢𝑦 (𝑟, 𝑠) = ∑ 𝑁𝑖 (𝑟, 𝑠)𝑢𝑦𝑖 + ∑ 𝑁𝑖 (𝑟, 𝑠)Δ𝑢𝑦𝑖 .
𝑖=1 𝑖=5
It is desired to replace the mid side relative displacment by drilling rotations at the
corner nodes. Consider Figure 44.5: the mid-side normal displacements along the edge are
parabolic, i.e.,
𝐿𝑖𝑗
Δ𝑢𝑖𝑗 = (Δ𝜃𝑗 − Δ𝜃𝑖 ), (44.11)
8
while the mid-side tangential displacements are interpolated linearly from the end node
displacements, thus,
𝐿𝑖𝑗
Δ𝑢𝑥 (𝑟, 𝑠) = cos𝛼𝑖𝑗 Δ𝑢𝑖𝑗 = cos𝛼𝑖𝑗 (Δ𝜃𝑗 − Δ𝜃𝑖 )
8
(44.12)
𝐿𝑖𝑗
Δ𝑢𝑦 (𝑟, 𝑠) = −sin𝛼𝑖𝑗 Δ𝑢𝑖𝑗 = −sin𝛼𝑖𝑗 (Δ𝜃𝑗 − Δ𝜃𝑖 ),
8
4 8
𝑢𝑥 (𝑟, 𝑠) = ∑ 𝑁𝑖 (𝑟, 𝑠)𝑢𝑥𝑖 + ∑ 𝑀𝑥𝑖 (𝑟, 𝑠)Δ𝜃𝑖
𝑖=1 𝑖=5
(44.13)
4 8
𝑢𝑦 (𝑟, 𝑠) = ∑ 𝑁𝑖 (𝑟, 𝑠)𝑢𝑦𝑖 + ∑ 𝑀𝑦𝑖 (𝑟, 𝑠)Δ𝜃𝑖 .
𝑖=1 𝑖=5
This element has one singularity in the drilling mode of equal corner rotations, see
Figure 44.6.
s
6 3
4
7 r
5
1 8 2
s
4 3
r
1 2
Figure 44.4. Eight node membrane element.
Δuy i = 1, 2, 3, 4
ΔΘj j = 2, 3, 4, 1
Lij m = 5, 6, 7, 8
Δux
Δuij
αij
ΔΘi
Figure 44.5. Corner node drilling rotations and mid side edge normal
displacement [Wilson, 2000].
at the center of the element. The element performance is highly insensitive to the chosen
value of the penalty factor and some fraction of the elastic modulii, G or E, is frequently
used.
• 5, 8 or 9 point quadrature can be applied. The 5 and 8 point schemes induce a ‘soft’
first deformational mode, whereas the 9 point Gaussian quadrature results in a
stiffer mode.
• A membrane locking correction (Taylor) is applied to (i) alleviate a membrane-
bending interaction associated with the drilling degrees of freedom and (ii) allow
the standard application of the consistent nodal load at the edge. The correction has
a slight stiffening effect (see e.g. Cook Cantilever).
• A warping correction is applied using the rigid link correction (see Figure 44.7).
Flat Element
45
Random Geometrical Imperfections
45.2 Methodology
45.2.1 Generation of random fields using Karhunen-Loève expansion
The Karhunen-Loève expansion (e.g., Ghanem and Spanos [2003]) provides an
attractive way of representing a random (stochastic) field (process) through a spectral
decomposition, , as a function of x (e.g., two spatial variables):
∞
𝜛(x, 𝜃) = 𝜛(̅ x) + ∑ √𝜆i 𝜉𝑖 (𝜃)𝑓𝑖 (x), (45.1)
𝑖=1
where the 𝜉𝑖 are uncorrelated zero-mean random variables with unit variance, and 𝜛(̅ x)is
the average random field or mean of the process. The functions 𝑓𝑖 are the eigenfunctions of
the covariance kernel, C, with 𝜆𝑖 the associated eigenvalues, obtained from the spectral
decomposition of the covariance function via the solution on a domain 𝐷 of the Fredholm
integral equation of the second kind,
If 𝜛 is Gaussian, then 𝜉𝑖 are also Gaussian. For non-Gaussian processes (with arbitrary but
specified marginal distributions), 𝜉𝑖 are unknown. Phoon et al ([2002a], [2005]) give an
iterative procedure for obtaining 𝜉𝑖 given a target marginal distribution.
45.2.2 Solution of Fredholm integral of the second kind for analytical covariance
functions
By defining a set of basis functions: 𝜑1 (𝑥), 𝜑2 (𝑥),…,𝜑𝑁 (𝑥), each eigenfunction 𝑓𝑖 (𝑥)
can be approximated by the linear combination:
𝑁
𝑓𝑖 (𝑥) = ∑ 𝑑𝑖𝑘 𝜑𝑘 (𝑥), (45.5)
𝑘=1
where the 𝑑𝑖𝑘 are constant coefficients. By substituting (45.5) into Fredholm equation (45.2)
(using a scalar 𝑥 (one-dimensional random process) as an example) and writing as an error:
𝑁 𝑁
∫ 𝐶(x1 , x2 ) ∑ 𝑑𝑖𝑘 𝜑𝑘 (𝑥1 )𝑑x1 − 𝜆𝑖 ∑ 𝑑𝑖𝑘 𝜑𝑘 (𝑥1 ) = 0. (45.6)
𝐷 𝑘=1 𝑘=1
or the eigensystem
𝐀𝐃 = 𝚲𝐁𝐃 (45.9)
1
with Λ𝑖𝑗 = 𝛿𝑖𝑗 𝜆𝑗 . Orthogonal wavelets 𝜓(𝑥) are used (∫0 𝜓𝑗 (x)𝜓𝑘 (x)𝑑x = ℎ𝑗 𝛿𝑗𝑘 ) as basis
functions,
𝑁
𝐟𝑖 (𝑥) = ∑ 𝑑𝑖𝑘 𝜓𝑘 (𝑥) = 𝛙T (𝑥)𝐃(𝑖) , (45.10)
𝑘=1
−|𝑥1 − 𝑥2 |
𝐶(x1 , x2 ) = exp ( ) (45.16)
𝐿𝑐
|𝑥1 − 𝑥2 |
𝐶(x1 , x2 ) = 1 − (45.17)
𝐿𝑐
sin𝐿𝑐 (𝑥1 − 𝑥2 )
𝐶(x1 , x2 ) = (45.18)
𝐿𝑐 (𝑥1 − 𝑥2 )
−|𝑥1 − 𝑥2 |2
𝐶(x1 , x2 ) = exp ( ) (45.19)
𝐿𝑐
−|𝑥1 − 𝑥2 |
𝐶(x1 , x2 ) = exp(−(𝑥1 − 𝑥2 ))exp (45.21)
𝐿𝑐
with 𝐿𝑐 the correlation length in the respective direction.
which implies that the 𝑀 – 1 eigenvectors, 𝐀𝐯𝑖 , are also eigenvectors of 𝐀𝐀T or 𝐂. The
The 𝑀 – 1 eigenvalues of 𝐋 and 𝐂 are identical, i.e., 𝜆𝑖 = 𝜇𝑖 . Finally 𝐟𝑖 and 𝜆𝑖 are used in
46
Frequency Domain
where 𝜙𝑛 is the n-th mode shape and 𝑞𝑛 (𝑡) is the n-th modal coordinates.
With the substitution of Equation (46.1.2) into Equation (46.1.30), the governing
equation can be rewritten as
𝐦𝚽𝐪̈ + 𝐜𝚽𝐪̇ + 𝐤𝚽𝐪 = 𝐩(𝑡). (46.1.3)
Pre-multiplying by ΦT gives
𝐌𝐪̈ + 𝐂𝐪̇ + 𝐊𝐪 = 𝐩(𝑡), (46.1.4)
The orthogonality of natural modes implies that the following square matrices are
diagonal:
𝑃𝑛 (𝑡)
𝑞𝑛̈ + 2𝜁𝑛 𝜔𝑛 𝑞𝑛̇ + 𝜔𝑛2 𝑞𝑛 = (46.1.11)
𝑀𝑛
where the modal damping coefficient n is defined as
𝐶𝑛
𝜁𝑛 = (46.1.12)
2𝑀𝑛 𝜔𝑛
Applying Fourier transform to both sides of Equation (46.1.11), one obtains
𝑃𝑛 (𝜔)
(−𝜔2 + 2𝑖𝜁𝑛 𝜔𝑛 𝜔 + 𝜔𝑛2 )𝑞𝑛 (𝜔) = (46.1.13)
𝑀𝑛
The structural displacement response in frequency domain can be represented as
𝑁
𝜙𝑛 𝑃𝑛 (𝜔)
𝐮(𝜔) = ∑ (46.1.14)
𝑛=1
(−𝜔2 2
+ 2𝑖𝜁𝑛 𝜔𝑛 𝜔 + 𝜔𝑛 ) 𝑀𝑛
𝑁
𝜙𝑛 (𝑥𝑘 ) 𝑃̃ 𝑛 (𝑥𝑗 )
𝐅𝐑𝐅𝒖 (𝑥𝑗 , 𝑥𝑘 , 𝜔) = ∑ (46.1.15)
𝑛=1
(−𝜔2 + 2𝑖𝜁𝑛 𝜔𝑛 𝜔 + 𝜔𝑛2 ) 𝑀𝑛
The velocity frequency response function (Mobility) can be expressed as
𝑁
𝜙𝑛 (𝑥𝑘 ) 𝑃̃ 𝑛 (𝑥𝑗 )
𝐅𝐑𝐅𝒗 (𝑥𝑗 , 𝑥𝑘 , 𝜔) = 𝜔𝑖 ∑ (46.1.16)
𝑛=1
(−𝜔2 + 2𝑖𝜁𝑛 𝜔𝑛 𝜔 + 𝜔𝑛2 ) 𝑀𝑛
The acceleration frequency response function (Accelerance) can be expressed as
𝑁
𝜙𝑛 (𝑥𝑘 ) 𝑃̃ 𝑛 (𝑥𝑗 )
𝐅𝐑𝐅𝒂 (𝑥𝑗 , 𝑥𝑘 , 𝜔) = −𝜔2 ∑ (46.1.17)
𝑛=1
(−𝜔2 + 2𝑖𝜁𝑛 𝜔𝑛 𝜔 + 𝜔𝑛2 ) 𝑀𝑛
~
where Pn ( x j ) is obtained as
𝐜 = 𝛼𝐦 + 𝛽𝐤 (46.1.19)
This type of damping is normally referred to as Rayleigh damping. For classically
damped system,
𝜕𝑝
− ∫ ∇𝑝∇𝑁𝑖 𝑑𝑉 + 𝑘 2 ∫ 𝑝𝑁𝑖 𝑑𝑉 = − ∫ 𝑁 𝑑Γ (46.2.4)
𝜕𝑛 𝑖
𝑉 𝑉 Γ
With the substitution of the boundary condition (46.2.2) into Equation (46.2.4), and
taking the nodal pressure as the unknown variables, a linear equation system can be
established and solved in frequency domain. Since there is only one variable on each node,
this method is very fast.
47
Rotor Dynamics
47.1 Introduction
Rotor dynamics is a specialized branch of engineering science concerned with the behavior
and diagnosis of rotating structures. It is a study of vibration of rotating parts found in a
wide range of equipment including engine, turbine, aircraft, hard disk drive and more.
The analysis of the rotator dynamics involves two coordinate systems: rotating and fixed
coordinate systems. The equations of motion in the two coordinate systems are both
introduced.
We assume that the origin of the rotating coordinate system is fixed, so that:
𝑽 =𝑨=𝟎. (47.32)
By substituting (1.4) to (1.3):
𝑑𝜴
𝒂̅ = 𝒂 + 𝟐𝜴 × 𝒗 + 𝜴 × (𝜴 × 𝒓) + ×𝒓 . (47.33)
𝑑𝑡
𝐝𝛀
𝑭 = 𝑚𝒂 = 𝑚𝒂̅ − 2𝑚𝜴 × 𝒗 − 𝑚𝜴 × (𝜴 × 𝒓) − 𝑚 × 𝐫. (47.34)
𝐝t
The first term 𝑚𝒂̅ is the force in the fixed coordinate system. All other terms on the right
hand side are inertia forces arising in the rotating system. The Coriolis force is the
following quantity:
𝑭𝑪 = −2𝑚𝜴 × 𝒗. (47.35)
The third term produces the familiar centrifugal force:
𝑭𝑪𝒇 = −𝑚𝜴 × (𝜴 × 𝒓). (47.36)
The last term introduces the Euler fore when there is a nonzero rate of change in the
magnitude of the rotation vector:
𝒅𝜴
𝑭𝑬 = −𝑚 × 𝒓. (47.37)
𝒅𝑡
{𝑥̅⎫
⎧ } 𝑐𝑜𝑠Ωt −𝑠𝑖𝑛Ωt 0 𝑥 𝑥
𝐫 ̅ = ⎨𝑦̅⎬ = ⎡
⎢𝑠𝑖𝑛Ωt 𝑐𝑜𝑠Ωt ⎤
0⎥ {𝑦} = 𝑯 {𝑦}. (47.40)
{
⎩ 𝑧̅}
⎭ ⎣ 0 0 1⎦ 𝑧 𝑧
𝑯 is the transformation matrix. It is easy to get that 𝑯 𝑇 𝑯 = 𝑰, where 𝑰 is he identity
matrix. Other matrices that may be used later are also given here:
−𝑠𝑖𝑛Ωt −𝑐𝑜𝑠Ωt 0
𝑯̇ = Ω ⎡
⎢ 𝑐𝑜𝑠Ωt −𝑠𝑖𝑛Ωt 0⎤ ̅̅̅̅̅̅,
⎥ = Ω𝐇 (47.41)
⎣ 0 0 0⎦
−𝑐𝑜𝑠Ωt sinΩt 0
𝑯̈ = Ω2 ⎡
⎢−𝑠𝑖𝑛Ωt −𝑐𝑜𝑠Ωt 0⎤
2 ̿̿̿̿̿
⎥ = Ω ̿𝐇 , (47.42)
⎣ 0 0 0⎦
0 1 0
̅̅̅ ̅ 𝑇 𝑯 = ⎡
𝑯 ⎢−1 0 0⎤⎥ = 𝑷, (47.43)
⎣0 0 0⎦
0 −1 0
̅̅̅̅̅ = ⎡
𝑯𝑇𝑯 ⎢1 0 0⎤
T
⎥ = 𝐏 = −𝑷, (47.44)
⎣0 0 0⎦
1 0 0
̅̅̅̅̅ 𝑇 = ⎡
̅̅̅̅̅ 𝑇 𝑯
𝑯 ⎢0 1 0⎤⎥ = 𝐉. (47.45)
⎣0 0 0 ⎦
Figure 1.4 Node with six masses located at offset x’, y’ and z’ from the node center.
With this, the location vector in the fixed coordinate is of the form if only considers
rotational displacement:
𝒓 ̅ = 𝑯(𝒓 + 𝒖) = 𝑯(𝒓 + 𝑨𝛉), (47.50)
where
0 𝑧 −𝑦
⎡
𝑨 = ⎢−𝑧 𝑥 ⎤
0 ⎥. (47.51)
⎣𝑦 −𝑥 0 ⎦
Then we can get the kinetic energy due to rotational displacement as:
𝑚Ω2 𝑇 𝑚
𝑇= (𝒓 𝑱𝒓 + 𝟐𝒓 𝑇 𝑱𝑨𝛉 + 𝛉𝑻 𝑨𝑇 𝑱𝑨𝛉) + (2Ω𝒓 𝑇 𝑷𝑨𝛉̇ + 2Ω𝛉𝑇 𝑨𝑇 𝑷𝑨𝛉̇ + 𝛉̇𝑇 𝑨𝑇 𝑨𝛉̇). (47.52)
2 2
42-14 (Rotor Dynamics) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual Rotor Dynamics
Applying Lagrange’s equation, then we can obtain the final equation of motion as follows:
𝑴𝒖 𝟎 𝑪 𝟎 𝒁 𝟎 𝒖 𝑭
[ ] {𝒖̈̈} + 2Ω [ 𝒖 ] {𝒖̇̇ } − Ω2 [ 𝒖 ] { } = { 𝒄𝒖 }, (47.53)
𝟎 𝑴𝛉 𝛉 𝟎 𝑪𝛉 𝛉 𝟎 𝒁𝛉 𝛉 𝑭𝒄𝛉
where 𝑴𝒖 and 𝑴𝛉 are the mass and inertia matrices; 𝑪𝒖 and 𝑪𝛉 are the gyroscopic
matrices; 𝒁𝒖 and 𝒁𝛉 are the centrifugal softening matrices; 𝑭𝒄𝒖 and 𝑭𝒄𝛉 are centrifugal force.
Note that we don’t consider the other system damping and external force terms here, but
they can be added to (1.25) accordingly.
𝒖 = 𝑻𝒖. (47.58)
where, 𝒓 and 𝒖 are the location vector and displacement in the rotating coordinate system
Oxyz.
The equation of motion in (1.25) and (1.27) can be simplified written as:
𝑴𝟎 𝒂 + 𝑪𝟎 𝒗 + 𝑲𝟎 𝒖 = 𝑭𝟎 , (47.59)
And the force term can be written as
𝑭𝟎 = 𝒇𝟎 𝒓, (47.60)
By substituting (1.28), (1.29) and (1.30) to the equation of motion, we can get new mass,
damping, stiffness matrices and force vector as follows:
𝑴 = 𝑻 𝑻 𝑴𝟎 𝑻 , (47.61)
𝑪 = 𝑻 𝑻 𝑪𝟎 𝑻, (47.62)
𝑲 = 𝑻 𝑻 𝑲𝟎 𝑻, (47.63)
𝑭 = 𝑻 𝑻 𝒇𝟎 𝑻𝒓. (47.64)
So the equation of motion becomes:
𝑴𝒂 + 𝑪𝒗 + 𝑲𝒖 = 𝑭 . (47.65)
After the equation of motion is obtained, it can then be solved using the implicit solver.
Especially, the damping and stiffness matrices are related to the rotational velocity, so the
eigen-frequencies might change with the change of rotational velocity. A diagram to
represent this relationship is called Campbell diagram. An example is given in Figure 1.6.
Figure 1.6 A disk is spinning with the center axis, the mode frequencies change with the
increase of rotating speed.
48
References
Addessio, F.L., D.E. Carroll, J.K. Dukowicz, F.H. Harlow, J.N. Johnson, B.A. Kashiwa,
M.E. Maltrud, H.M. Ruppel, “CAVEAT: A Computer Code for Fluid Dynamics
Problems with Large Distortion and Internal Slip,” Report LA-10613-MS, UC-32, Los
Alamos National Laboratory (1986).
Ahmad, S., Irons, B.M. and Zienkiewicz, O.C., “Analysis of Thick and Thin Shell
Structures by Curved Finite Elements,” Int. J. Numer. Meths. Eng., 2 (1970).
Amdsden, A., A., Ruppel, H. M., and Hirt, C. W., “SALE: A Simplified ALE Computer
Program for Fluid Flow at All Speeds,” Los Alamos Scientific Laboratory (1980).
Argyris, J.H., Kelsey, S., and Kamel, H., “Matrix Methods of Structural Analysis: A
Precis of recent Developments,” Matrix Methods of Structural Analysis, Pergamon
Press (1964).
Arruda, E., and M. Boyce, "A Three-Dimensional Constitutive Model for the Large
Stretch Behavior of Rubber Elastic Materials," published in the Journal of the Mechanics
and Physics of Solids, Vol. 41, No. 2, 389-412 (1993).
Auricchio, F., Taylor, R.L. and Lubliner J., “Shape-memory alloys: macromodelling and
numerical simulations of the superelastic behavior”, Computer Methods in Applied
Mechanics and Engineering 146, 281-312 (1997).
Back, S.Y., and Will, K.M., “A Shear-flexible Element with Warping for Thin-Walled
Open Sections,” International Journal for Numerical Methods in Engineering, 43, 1173-1191
(1998).
Bahler AS: The series elastic element of mammalian skeletal muscle. Am J Physiol
213:1560-1564, 1967.
Bammann, D. J., and E.C. Aifantis, “A Model for Finite-Deformation Plasticity,” Acta
Mechanica, 70, 1-13 (1987).
Bammann, D. J., “Modeling the Temperature and Strain Rate Dependent Large
Deformation of Metals,” Proceedings of the 11th US National Congress of Applied
Mechanics, Tucson, AZ (1989).
Bammann, D. J., and Johnson, G., “On the Kinematics of Finite-Deformation Plasticity,”
Acta Mechanica 69, 97-117 (1987).
Bammann, D.J., Chiesa, M.L., McDonald, A., Kawahara, W.A., Dike, J.J. and Revelli,
V.D., “Predictions of Ductile Failure in Metal Structures,” in AMD-Vol. 107, Failure
Criteria and Analysis in Dynamic Response, edited by. H.E. Lindberg, 7-12 (1990).
Barlat and Lian, J., “Plastic Behavior and Stretchability of Sheet Metals, Part I: A Yield
Function for Orthotropic Sheets Under Plane Stress Conditions,” International Journal of
Plasticity, 5, 51-66 (1989).
Barlat, F., Lege, D.J., and Brem, J.C., “A Six-Component Yield Function for Anisotropic
Materials,” International Journal of Plasticity, 7, 693-712 (1991).
Bathe, K. J., and Wilson, E.L., Numerical Methods in Finite Element Analysis, Prentice-
Hall (1976).
Bathe, K.J., and Dvorkin, E.N., “A Continuum Mechanics Based Four Node Shell
Element for General Nonlinear Analysis,” Int. J. Computer-Aided Eng. and Software,
Vol. 1, 77-88 (1984).
Bathe, K.-J. and Dvorkin, E.N. A four node plate bending element based on Mindlin-
Reissner plate theory and a mixed interpolation, Int. J. Num. Meth. Eng., 21, 367-383,
1985.
Batoz, J.L. and Ben Tahar, M. Evaluation of a new quadrilateral thin plate bending
element, Int. J. Num. Meth. Eng., 18, 1644-1677, 1982.
Battini, J., and Pacoste C., “Co-rotational Beam Elements with Warping Effects in
Instability Problems,” Computational Methods in Applied Mechanical Engineering, 191,
1755-1789, (2002).
Bazeley, G.P., Cheung, W.K., Irons, B.M. and Zienkiewicz, O.C., “Triangular Elements
in Plate Bending—Conforming and Nonconforming Solutions in Matrix Methods and
Structural Mechanics,” Proc. Conf. on Matrix Methods in Structural Analysis, Rept.
AFFDL-R-66-80, Wright Patterson AFB, 547-576 (1965).
Belytschko, T., and Bindeman, L. P., "Assumed Strain Stabilization of the Eight Node
Hexahedral Element," Comp. Meth. Appl. Mech. Eng. 105, 225-260 (1993).
Belytschko, T., and Hsieh, B.J., ”Nonlinear Transient Finite Element Analysis with
Convected Coordinates,” Int. J. Num. Meths. Engrg., 7, 255–271 (1973).
Belytschko, T., and Leviathan, I., “Projection schemes for one-point quadrature shell
elements,” Comp. Meth. Appl. Mech. Eng., 115, 277-286 (1994).
Belytschko, T., and Lin, J., “A New Interaction Algorithm with Erosion for EPIC-3”,
Contract Report BRL-CR-540, U.S. Army Ballistic Research Laboratory, Aberdeen
Proving Ground, Maryland (1985).
Belytschko, T., Lin, J., and Tsay, C.S., “Explicit Algorithms for Nonlinear Dynamics of
Shells,” Comp. Meth. Appl. Mech. Eng. 42, 225-251 (1984) [a].
Belytschko, T., Ong, J. S.-J., Liu, W.K. and Kennedy, J.M., “Hourglass Control in Linear
and Nonlinear Problems”, Comput. Meths. Appl. Mech. Engrg., 43, 251–276 (1984b).
Belytschko, T., Stolarski, H., and Carpenter, N., “A C0 Triangular Plate Element with
One-Point Quadrature,” International Journal for Numerical Methods in Engineering,
20, 787-802 (1984) [b].
Belytschko, T., Schwer, L., and Klein, M. J., “Large Displacement Transient Analysis of
Space Frames,” International Journal for Numerical and Analytical Methods in
Engineering, 11, 65-84 (1977).
Belytschko, T., and Tsay, C.S., “Explicit Algorithms for Nonlinear Dynamics of Shells,”
AMD, 48, ASME, 209-231 (1981).
Belytschko, T., and Tsay, C. S., “A Stabilization Procedure for the Quadrilateral Plate
Element with One-Point Quadrature,” Int. J. Num. Method. Eng. 19, 405-419 (1983).
Belytschko, T., Yen, H. R., and Mullen R., “Mixed Methods for Time Integration,”
Computer Methods in Applied Mechanics and Engineering, 17, 259-175 (1979).
Belytschko, T., “Partitioned and Adaptive Algorithms for Explicit Time Integration,” in
Nonlinear Finite Element Analysis in Structural Mechanics, ed. by Wunderlich, W.
Stein, E, and Bathe, J. J., 572-584 (1980).
Belytschko, T., Wong, B.L., and Chiang, H.Y., “Improvements in Low-Order Shell
Elements for Explicit Transient Analysis,” Analytical and Computational Models of
Shells, A.K. Noor, T. Belytschko, and J. Simo, editors, ASME, CED, 3, 383-398 (1989).
Belytschko, T., Wong, B.L., and Chiang, H.Y., “Advances in One-Point Quadrature Shell
Elements,” Comp. Meths. Appl. Mech. Eng., 96, 93-107 (1992).
Belytschko, T., Wong, B. L., Plaskacz, E. J., "Fission - Fusion Adaptivity in Finite
Elements for Nonlinear Dynamics of Shells," Computers and Structures, Vol. 33, 1307-
1323 (1989).
Belytschko, T., Lu, Y.Y. and Gu, L., “Element-free Galerkin Methods,” Int. J. Numer.
Methods Engrg. 37, 229-256 (1994).
Benson, D.J., “Vectorizing the Right-Hand Side Assembly in an Explicit Finite Element
Program,” Comp. Meths. Appl. Mech. Eng., 73, 147-152 (1989).
Benson, D. J., and Hallquist, J.O., “A Simple Rigid Body Algorithm for Structural
Dynamics Program,” Int. J. Numer. Meth. Eng., 22 (1986).
Benson, D.J., and Hallquist J.O., “A Single Surface Contact Algorithm for the
Postbuckling Analysis of Shell Structures,” Comp. Meths. Appl. Mech. Eng., 78, 141-
163 (1990).
43-4 (References) LS-DYNA DEV 10/27/16 (r:8004)
LS-DYNA Theory Manual References
Bischoff M. and Ramm E., “Shear deformable shell elements for large strains and
rotations,”
. Int. J. Numer. Methods, 40, 4427-4449 (1997).
Blatz, P.J., and Ko, W.L., “Application of Finite Element Theory to the Deformation of
Rubbery Materials,” Trans. Soc. of Rheology, 6, 223-251 (1962).
Bodig, Jozsef and Benjamin A. Jayne, Mechanics of Wood and Wood Composites,
Krieger Publishing Company, Malabar, FL (1993).
Boeing, Boeing Extreme Mathematical Library BCSLIB-EXT User's Guide, The Boeing
Company, Document Number 20462-0520-R4 (2000).
Borrvall, T., Revision of the implementation of material 36 for shell elements in LS-
DYNA, ERAB Report E0307, Engineering Research Nordic AB, Linköping (2003).
Björklund O., “Ductile Failure in High Strength Steel Sheets,” Linköping Studies in
Science and Technology. Dissertations No. 1579, ISSN 0345-7524 (2014).
Burton, D.E., et. al., “Physics and Numerics of the TENSOR Code,” Lawrence
Livermore National Laboratory, Internal Document UCID-19428 (July 1982).
Cardoso, R.P.R. and Yoon J-W., “One point quadrature shell element with through-
thickness stretch, Computer Methods in Applied Mechanics and Engineering,” 194(9),
1161-1199 (2005).
Chang, F.K., and Chang, K.Y., “Post-Failure Analysis of Bolted Composite Joints in
Tension or Shear-Out Mode Failure,” J. of Composite Materials, 21, 809-833 (1987).[a]
Chang, F.K., and Chang, K.Y., “A Progressive Damage Model for Laminated
Composites Containing Stress Concentration,” J. of Composite Materials, 21, 834-855
(1987).[b]
Chang, F.S., J.O. Hallquist, D. X. Lu, B. K. Shahidi, C. M. Kudelko, and J.P. Tekelly,
“Finite Element Analysis of Low Density High-Hystersis Foam Materials and the
Application in the Automotive Industry,” SAE Technical Paper 940908, in Saftey
Technology (SP-1041), International Congress and Exposition, Detroit, Michigan (1994).
Chen, J.S., Pan, C., Wu, C.T. and Liu, W.K., “Reproducing Kernel Particle Methods for
Large Deformation Analysis of Nonlinear Structures,” Comput. Methods Appl. Mech.
Engrg. 139, 195-227 (1996).
Chen, J.S. and Wu, C.T., “Generalized Nonlocal Meshfree Method in Strain
Localization,” Proceeding of International Conference on Computational Engineering Science,
Atlanta, Georiga, 6-9 October (1998).
Chen, J.S. and Wang, H.P. “New Boundary Condition Treatmens in Meshfree
Computation of Contact Problems,” Computer Methods in Applied Mechanics and
Engineering, 187, 441-468, (2001a).
Chen, J.S., Wu, C.T., Yoon, S. and You, Y. “A Stabilized Conforming Nodal Integration
for Galerkin Meshfree Methods,” Int. J. Numer. Methods Engrg. 50, 435-466 (2001b).
Chen, W.F., and Baladi, G.Y., Soil Plasticity: Theory and Implementation, Elesvier, New
York, (1985).
Chung, K., and K. Shah, “Finite Element Simulation of Sheet Metal Forming for Planar
Anisotropic Metals,” Int. J. of Plasticity, 8, 453-476 (1992).
Cochran, S.G., and Chan, J., “Shock Initiation and Detonation Models in One and Two
Dimensions,” University of California, Lawrence Livermore National Laboratory, Rept.
UCID-18024 (1979).
Cohen, M., and Jennings, P.C., “Silent Boundary Methods for Transient Analysis”, in
Computational Methods for Transient Analysis, T. Belytschko and T.J.R. Hughes,
editors, North-Holland, New York, 301-360 (1983).
Cook, R. D., Concepts and Applications of Finite Element Analysis, John Wiley and
Sons, Inc. (1974).
Couch, R., Albright, E. and Alexander, “The JOY Computer Code,” University of
California, Lawrence Livermore National Laboratory, Rept. UCID-19688 (1983).
Cray Research Inc., “CF77 Compiling System, Volume 4: Parallel Processing Guide,”
SG-3074 4.0, Mendota Heights, MN (1990).
Crisfield M.A., Non-linear Finite Element Analysis of Solids and Structures, Volume 2,
Advanced Topics, John Wiley, New York (1997).
Deshpande, V.S. and N.A. Fleck, “Isotropic Models for Metallic Foams,” Journal of the
Mechanics and Physics of Solids, Vol. 48, 1253-1283, (2000).
Duff, I. S., and Reid, J. K., "The Multifrontal Solution of Indefinite Sparse Symmetric
Linear Equations," ACM Transactions of Mathematical Software, 9, 302-325 (1983).
Dvorkin, E.N. and Bathe, K.J. “A continuum mechanics based four-node shell element
for general nonlinear analysis,” International Journal for Computer-Aided Engineering
and Software, 1, 77-88 (1984).
Englemann, B.E. and Whirley, R.G., “A New Explicit Shell Element Formulation for
Impact Analysis,” In Computational Aspects of Contact Impact and Penetration, Kulak,
R.F., and Schwer, L.E., Editors, Elmepress International, Lausanne, Switzerland, 51-90
(1991).
Englemann, B.E., Whirley, R.G., and Goudreau, G.L., “A Simple Shell Element
Formulation for Large-Scale Elastoplastic Analysis,” In Analytical and Computational
Models of Shells, Noor, A.K., Belytschko, T., and Simo, J.C., Eds., CED-Vol. 3, ASME,
New York, New York (1989).
Farhoomand, I., and Wilson, E.L., “A Nonlinear Finite Element Code for Analyzing the
Blast Response of Underground Structures,” U.S. Army Waterways Experiment Station,
Contract Rept. N-70-1 (1970).
Flanagan, D.P. and Belytschko, T., “A Uniform Strain Hexahedron and Quadrilateral
and Orthogonal Hourglass Control,” Int. J. Numer. Meths. Eng. 17, 679-706 (1981)
Fleck, J.T., “Validation of the Crash Victim Simulator,” I - IV, Report No. DOT-HS-806
279 (1981).
Fleischer M., Borrvall T., and Bletzinger K-U., “Experience from using recently
implemented enhancements for Material 36 in LS-DYNA 971 performing a virtual
tensile test”, 6th European LS-DYNA Users Conference, Gothenburg, Sweden, 2007.
Galbraith, P.C., M.J. Finn, S.R. MacEwen, et.al., "Evaluation of an LS-DYNA3D Model
for Deep-Drawing of Aluminum Sheet", FE-Simulation of 3-D Sheet Metal Forming
Processes in Automotive Industry, VDI Berichte, 894, 441-466 (1991).
Ghanem RG, Spanos PD. Stochastic Finite Elements – A Spectral Approach, Springer-
Verlag, 1991, Revised Edition, Dover, 2003.
Gingold, R.A., and Monaghan, J.J., “Smoothed Particle Hydrodynamics: Theory and
Application to Non-Spherical Stars,” Mon. Not. R. Astron. Soc. 181, 375-389 (1977).
Govindjee, S., Kay G.J., and Simo, J.C., “Anisotropic Modeling and Numerical
Simulation of Brittle Damage in Concrete,” Report Number UCB/SEM M-94/18,
University of California at Berkeley, Department of Civil Engineering (1994).
Govindjee, S., Kay G.J., and Simo, J.C., “Anisotropic Modeling and Numerical
Simulation of Brittle Damage in Concrete,” International Journal for Numerical Methods in
Engineering, Volume 38, pages 3611-3633 (1995).
Groenwold, A.A. and Stander, N. An efficient 4-node 24 d.o.f. thick shell finite
element with 5-point quadrature. Eng. Comput. 12(8), 723-747, 1995.
Grimes, Roger, Lewis, John G., and Simon, Horst D., "A Shifted Block Lanczos
Algorithm for Solving Sparse Symmetric Generalized Eigenproblems," SIAM Journal of
Matrix Analyis and Applications, 15, 228-272 (1994).
Hallquist, J.O., “Preliminary User’s Manuals for DYNA3D and DYNAP (Nonlinear
Dynamic Analysis of Solids in Three Dimension),” University of California, Lawrence
Livermore National Laboratory, Rept. UCID-17268 (1976) and Rev. 1 (1979).[a]
Hallquist, J.O., “A Numerical Treatment of Sliding Interfaces and Impact,” in: K.C. Park
and D.K. Gartling (eds.) Computational Techniques for Interface Problems, AMD, 30,
ASME, New York (1978).
Hallquist, J.O., “NIKE2D: An Implicit, Finite-Element Code for Analyzing the Static and
Dynamic Response of Two-Dimensional Solids,” University of California, Lawrence
Livermore National Laboratory, Rept. UCRL-52678 (1979).[b]
Hallquist, J.O., “User's Manual for DYNA3D and DYNAP (Nonlinear Dynamic
Analysis of Solids in Three Dimensions),” University of California, Lawrence Livermore
National Laboratory, Rept. UCID-19156 (1981).[a]
Hallquist, J.O., LS-DYNA Keyword User’s Manual, version 970, Livermore Software
Technology Corporation (April, 2003).
Structures, T.J.R. Hughes and E. Hinton, Editors, 394-431, Pineridge Press Int., Swanea,
U.K. (1986).
Hallquist, J.O., and Benson, D.J., “DYNA3D User’s Manual (Nonlinear Dynamic
Analysis of Solids in Three Dimensions),” University of California, Lawrence Livermore
National Laboratory, Rept. UCID-19156 (1986, Rev. 2).
Hallquist, J.O. and Benson, D.J., “DYNA3D User’s Manual (Nonlinear Dynamic
Analysis of Solids in Three Dimensions),” University of California, Lawrence Livermore
National Laboratory, Rept. UCID-19156 (1987, Rev. 3).
Hallquist, J.O., Stillman, D.W., Hughes, T.J.R., and Tarver, C.,”Modeling of Airbags
Using MVMA/DYNA3D,” LSTC Report (1990).
Harten, A., “ENO Schemes with Subcell Resolution,” J. of Computational Physics, 83,
148-184, (1989).
Herrmann, L.R. and Peterson, F.E., “A Numerical Procedure for Viscoelastic Stress
Analysis,” Seventh Meeting of ICRPG Mechanical Behavior Working Group, Orlando,
FL, CPIA Publication No. 177 (1968).
Hill A.V., "The heat of shortening and the dynamic constants of muscle," Proc Roy Soc
B126:136-195, (1938).
Hill, R., “A Theory of the Yielding and Plastic Flow of Anisotropic Metals,” Proceedings
of the Royal Society of London, Series A., 193, 281 (1948).
Holmquist, T.J., G.R. Johnson, and W.H. Cook, "A Computational Constitutive Model
for Concrete Subjected to Large Strains, High Strain Rates, and High Pressures, pp. 591-
600, (1993).
Hopperstad, O.S. and Remseth, S., “A Return Mapping Algorithm for a Class of Cyclic
Plasticity Models,” International Journal for Numerical Methods in Engineering, 38, 549-564
(1995).
Hughes, T.J.R., The Finite Element Method, Linear Static and Dynamic Finite Element
Analysis, Prentice-Hall Inc., Englewood Cliffs, NJ (1987).
Hughes, T.J.R. and Carnoy, E., “Nonlinear Finite Element Shell Formulation
Accounting for Large Membrane Strains,” AMD-Vol.48, ASME, 193-208 (1981).
Hughes, T.J.R. and Liu, W.K., “Nonlinear Finite Element Analysis of Shells: Part I.
Two-Dimensional Shells.” Comp. Meths. Appl. Mechs. 27, 167-181 (1981).
Hughes, T.J.R. and Liu, W.K., “Nonlinear Finite Element Analysis of Shells: Part II.
Three-Dimensional Shells.” Comp. Meths. Appl. Mechs. 27, 331-362 (1981).
Hughes, T.J.R., Liu,W.K., and Levit, I., “Nonlinear Dynamics Finite Element Analysis of
Shells.” Nonlinear Finite Element Analysis in Struct. Mech., Eds. W. Wunderlich, E.
Stein, and K.J. Bathe, Springer-Verlag, Berlin, 151–168 (1981).
Hughes, T.J.R., Taylor, R. L., Sackman, J. L., Curnier, A.C., and Kanoknukulchai, W.,
“A Finite Element Method for a Class of Contact-Impact Problems,” J. Comp. Meths.
Appl. Mechs. Eng. 8, 249-276 (1976).
Hughes, T.J.R., and Winget, J., “Finite Rotation Effects in Numerical Integration of Rate
Constitutive Equations Arising in Large-Deformation Analysis,” Int. J. Numer. Meths.
Eng., 15, 1862-1867, (1980).
Isenberg, J., D.K. Vaughn, and I. Sandler, "Nonlinear Soil-Structure Interaction," EPRI
Report MP-945, Weidlinger Associates, December (1978).
Jabareen, M., and Rubin, M.B., A Generalized Cosserat Point Element (CPE) for Isotropic
Nonlinear Elastic Materials including Irregular 3-D Brick and Thin Structures, J. Mech.
Mat. And Struct., Vol 3-8, 1465-1498 (2008).
Jabareen, M., Hanukah, E. and Rubin, M.B., A Ten Node Tetrahedral Cosserat Point
Element (CPE) for Nonlinear Isotropic Elastic Materials, J. Comput. Mech. 52, 257-285
(2013).
Johnson, G.C. and Bammann D.J., “A Discussion of Stress Rates in Finite Deformation
Problems" International Journal of Solids and Structures, 20, 725-737 (1984).
Johnson, G.R. and Cook,W. H., “A Constitutive Model and Data for Metals Subjected
to Large Strains, High Strain Rates and High Temperatures,” presented at the Seventh
International Symposium on Ballistics, The Hague, The Netherlands, April (1983).
Johnson, G.R. and T.J. Holmquist, "An Improved Computational Model for Brittle
Materials" in High-Pressure Science and Technology - 1993 American Institute of
Physics Conference Proceedings 309 (c 1994) pp.981-984 ISBN 1-56396-219-5.
Johnson, C., Navert, U., and Pitkaranta, J., “Finite Element Methods for Linear
Hyperbolic Problems,” Computer Methods in Applied Mechanics and Engineering, 45,
285-312, (1984).
Jones, R.M., Mechanics of Composite Materials, Hemisphere Publishing Co., New York
(1975).
Karypis, G., and Kumar V., "METIS: A Software Package for Partitioning Unstructured
Graphs, Partitioning Meshes, and Computing Fill-Reducing Orderings of Sparse
Matrices," Department of Computer Science, University of Minnesota (1998).
Katz, J., and Maskew, B., “Unsteady Low-Speed Aerodynamic Model for Complete
Aircraft Configurations,” Journal of Aircraft 25, 4 (1988).
Kennedy, J. M., Belytschko,T., and Lin, J. I., “Recent Developments in Explicit Finite
Element Techniques and their Applications to Reactor Structures,” Nuclear Engineering
and Design 97, 1-24 (1986).
Kim, N., Seo, K., and Kim, M., “Free Vibration and Spatial Stability of Non-symmetric
Thin-Walled Curved Beams with Variable Curvatures,” International Journal of Solids and
Structures, 40, 3107-3128, (2003).
Klisinski M., “Degradation and Plastic Deformation of Concrete”, Ph.D. thesis, Polish
Academy of Sciences, 1985, IFTR report 38.
Kreyszig, E., Advanced Engineering Mathematics, John Wiley and Sons, New York,
New York (1972).
Krieg, R.D.,”A Simple Constitutive Description for Cellular Concrete,” Sandia National
Laboratories, Albuquerque, NM, Rept. SC-DR-72-0883 (1972).
Krieg, R.D. and Key, S.W., Implementation of a Time Dependent Plasticity Theory into
Structural Computer Programs, Vol. 20 of Constitutive Equations in Viscoplasticity:
Computational and Engineering Aspects (American Society of Mechanical Engineers,
New York, N.Y., 1976), 125-137.
Kumar, P., Nukala, V. V., and White, D. W., “A Mixed Finite Element for Three-
dimensional Nonlinear Analysis of Steel Frames,” Computational Methods in Applied
Mechanical Engineering, 193, 2507-2545, (2004).
Landshoff, R., “A Numerical Method for Treating Fluid Flow in the Presence of
Shocks,” Los Alamos Scientific Laboratory, Rept. LA-1930 (1955).
Lemmen, P.P.M and Meijer, G.J., “Failure Prediction Tool Theory and User Manual,”
TNO Building and Construction Research, The Netherlands, Rept. 2000-CMC-R0018,
(2001).
Lewis B.A., “Developing and Implementing a Road Side Safety Soil Model into LS-
DYNA”, FHWA Research and Development Turner-Fairbank Highway Research
Center. Oct, 1999.
Librescu, L. Qin, Z., and Ambur D.R., “Implications of Warping Restraint on Statics
and Dynamics of Elastically Tailored Thin-Walled Composite Beams,” International
Journal of Mechanical Sciences, 45, 1247-1267 (2003).
Liu, W.K. and Belytschko, T., “Efficient Linear and Nonlinear Heat Conduction with a
Quadrilateral Element,” Int. J. Num. Meths. Engrg., 20, 931–948 (1984).
Liu, W.K., Belytschko, T., Ong, J.S.J. and Law, E., “Use of Stabilization Matrices in
Nonlinear Finite Element Analysis,” Engineering Computations, 2, 47–55 (1985).
Liu, W.K., Guo, Y., Tang, S. and Belytschko T., “A Multiple-Quadrature Eight-node
Hexahedral Finite Element for Large Deformation Elastoplastic Analysis,” Comput.
Meths. Appl. Mech. Engrg., 154, 69–132 (1998).
Liu, W.K., Hu, Y.K., and Belytschko, T., “Multiple Quadrature Underintegrated Finite
Elements,” Int. J. Num. Meths. Engrg., 37, 3263–3289 (1994).
Liu, W. K., Chang, H., and Belytschko, T., “Arbitrary Lagrangian-Eulerian Petrov-
Galerkin Finite Elements for Nonlinear Continua,” Computer Methods in Applied
Mechanics and Engineering, to be published.
Liu, W.K., Jun, S. and Zhang, Y.F., “Reproducing Kernel Particle Method,” Int. J.
Numer. Methods Fluids 20, 1081-1106 (1995).
Lucy, L.B., “Numerical Approach to Testing the Fission Hyphothesis,” Astron. J. 82,
1013-1024 (1977).
Lysmer, J. and Kuhlemeyer, R.L., “Finite Dynamic Model for Infinite Media”, J. Eng.
Mech. Div. ASCE, 859-877 (1969).
MacNeal R.H. and Harder R.L., “A Proposed Standard Set of Problems to Test Finite
Element Accuracy,” Finite Elements Anal. Des., (1985, 3–20 (1985).
Maenchen, G. and Sack, S., “The Tensor Code,” Meth. Comp. Phys. 3, (Academic
Press), 181-263 (1964).
Marchertas, A. H., and Belytschko,T. B., “Nonlinear Finite Element Formulation for
Transient Analysis of Three Dimensional Thin Structures,” Report ANL-8104, LMFBR
Safety, Argonne National Laboratory, Argonne, IL (1974).
Matthies, H., and G. Strang, The Solution of Nonlinear Finite Element Equations, Int.,
Journal for Numerical Methods in Engineering, 14, No. 11, 1613-1626.
Mauldin, P.J., R.F. Davidson, and R.J. Henninger, “Implementation and Assessment of
the Mechanical-Threshold-Stress Model Using the EPIC2 and PINON Computer
Codes,” Report LA-11895-MS, Los Alamos National Laboratory (1990).
Mindlin, R.D., “Influence of Rotary Inertia and Shear on Flexural Motions of Isotropic,
Elastic Plates,” J. Appl. Mech. 18, 31-38 (1951).
Mizukami, A., and Hughes, T. J. R., “A Petrov-Galerkin Finite Element Method for
Convection Dominated Flows: An Accurate Upwinding Technique for Satisfying the
Maximum Principle,” Computer Methods in Applied Mechanics and Engineering, Vol.
50, pp. 181-193 (1985).
Monaghan, J.J., and Gingold, R.A., “Shock Simulation by the Particle Method of SPH,”
Journal of Computational Physics, 52, 374-381 (1983).
Murray Y.D., “Modeling Rate Effects in Rock and Concrete”, Proceedings of the 8th
International Symposium on Interaction of the Effects of Munitions with Structures,
Defense Special Weapons Agency, McLean VA, USA, (1997).
Murray, Y. D., “Modeling Rate Effects in Rock and Concrete,” Proceedings of the 8th
International Symposium on Interaction of the Effects of Munitions with Structures,
Defense Special Weapons Agency, McLean, VA, USA, (1997).
Murray, Y.D., Users Manual for Transversely Isotropic Wood Model, aptek technical
report to fhwa (to be published), (2002).
Murray, Y.D., and J. Reid, Evaluation of Wood Model for Roadside Safety Applications,
aptek technical report to fhwa (to be puslished), (2002).
Nagtegaal, J.C., Parks, D.M., and Rice J.R., “On Numerically Accurate Finite Element
Solution in the Fully Plastic Range”, Computer Methods in Applied Mechanics and
Engineering, 4, 153 (1974).
Neilsen, M.K., Morgan, H.S., and Krieg, R.D., “A Phenomenological Constitutive Model
for Low Density Polyurethane Foams,” Rept. SAND86-2927, Sandia National
Laboratories, Albuquerque, NM (1987).
Oden, J.T, P. Devloo, and T. Strouboulis, "Adaptive Finite Element Methods for the
Analysis of Inviscid Compressible Flow: Part I. Fast Refinement/Unrefinement and
LS-DYNA DEV 10/27/16 (r:8004) 43-17 (References)
References LS-DYNA Theory Manual
Ogden, R.W., Non-Linear Elastic Deformations, Ellis Horwood Ltd., Chichester, Great
Britian (1984).
Okuda T., Yamamae Y. and Yasuki T., Request for MAT126 Modification, Microsoft
Power Point presentation, Toyota Communication Systems and Toyota Motor
Corporation, 2003.
Papadrakakis, M., “A Method for the Automated Evaluation of the Dynamic Relaxation
Parameters,” Comp. Meth. Appl. Mech. Eng. 25, 35-48 (1981).
Phoon KK, Huang SP, Quek ST. Implementation of Karhunen–Loeve expansion for
simulation using a wavelet-Galerkin scheme. Probabilist Eng Mech 2002;17(3):293–303.
(2002a)
Phoon KK, Huang SP, Quek ST. Simulation of second-order processes using
Karhunen–Loeve expansion. Comput Struct 2002; 80(12):1049–60. (2002b)
Phoon KK, Huang HW, Quek ST. Simulation of strongly non-Gaussian processes using
Karhunen–Loeve expansion, Probabilistic Engineering Mechanics 20 (2005) 188–198.
Pian, T.H.H., and Sumihara, K., “Rational Approach for Assumed Stress Elements,” Int.
J. Num. Meth. Eng., 20, 1685-1695 (1985).
Prokic, A., “New Warping Function for Thin-Walled Beams. I: Theory,” Journal of
Structural Engineering, 122, 1437-1442, (1994).
Reid, S.R. and C. Peng, “Dynamic Uniaxial Crushing of Wood,” Int. J. Impact
Engineering, Vol. 19, No. 5-6, pp. 531-570, (1997).
Richtmyer, R.D., and Morton, K.W., Difference Equations for Initial-Value Problems,
Interscience Publishers, New York (1967).
Sandler, I.S., and D. Rubin, “An Algorithm and a modular subroutine for the cap
model,” Int. J. Numer. Analy. Meth. Geomech., 3, 173-186 (1979).
Schwer, L.E., Cheva, W., and Hallquist, J.O., “A Simple Viscoelastic Model for Energy
Absorbers Used in Vehicle-Barrier Impacts,” In Computational Aspects of Contact
Impact and Penetration, Kulak, R.F., and Schwer, L.E., Editors, Elmepress International,
Lausanne, Switzerland, 99-117 (1991).
Shapiro, A.B., “TOPAZ3D - A Three Dimensional Finite Element Heat Transfer Code,”
University of California, Lawrence Livermore National Laboratory, Report UCID-20481
(1985).
Shapiro, A.B., “REMAP: a Computer Code That Transfers Node Information Between
Dissimilar Grids,” Lawrence Livermore National Laboratory, UCRL-ID-104090, (1990).
Simo, J.C. and Govindjee, S., ``Exact Closed-Form Solution of the Return Mapping
Algorithm in Plane Stress Elasto-Viscoplasticity," Engineering Computations, 5, 254-258
(1988).
Simo, J.C. and Hughes, T.J.R. “On the variational foundations of assumed strain
methods,” Journal of Applied Mechanics, 53, 1685-1695 (1986).
Simo, J.C., Ju, J.W., “Stress and Strain Based Continuum Damage Models”, Parts I & II,
Int. J. of Solids and Structures, Vol 23, No 7, (1987).
Simo, J.C., Ju, J.W., Pister, K.S., and Taylor, R.L. “An Assessment of the Cap Model:
Consistent Return Algorithms and Rate-Dependent Extension,” J. Eng. Mech., 114,
No. 2, 191-218 (1988a).
Simo, J.C., Ju, J.W., Pister, K.S., and Taylor, R.L. “Softening Response, Completeness
Condition, and Numerical Algorithms for the Cap Model,” Int. J. Numer. Analy.
Meth. Eng. (1990).
LS-DYNA DEV 10/27/16 (r:8004) 43-19 (References)
References LS-DYNA Theory Manual
Simo, J.C. and Taylor, R.L., “A Return Mapping Algorithm for Plane Stress
Elastoplasticity,” International Journal for Numerical Methods in Engineering, 22, 649-670
(1986).
Steinberg, D.J. and Guinan, M.W., “A High-Strain-Rate Constitutive Model for Metals,”
University of California, Lawrence Livermore National Laboratory, Rept. UCRL-80465
(1978).
Stolarski, H., and Belytschko, T., “Shear and Membrane Locking in Curved Elements,”
Comput. Meths. Appl. Mech. Engrg., 41, 279–296 (1983).
Stone, C.M., Krieg, R.D., and Beisinger, Z.E., “Sancho, A Finite Element Computer
Program for the Quasistatic, Large Deformation, Inelastic Response of Two-
Dimensional Solids,” Sandia Report, SAND 84-2618, UC-32, Albuquerque, NM (1985).
Sturt, R.M.V., and B.D. Walker, J.C. Miles, A. Giles, and N. Grew, “Modelling the
Occupant in a Vehicle Context-An Integrated Approach,” 13th International ESV
Conference, Paris, November 4-7 (1991).
Tarver, C.M., and Hallquist, J.O., “Modeling of Two-Dimensional Shock Initiation and
Detonation Wave Phenomena in PBX 9404 and LX-17,” University of California,
Lawrence Livermore National Laboratory, Rept. UCID84990 (1981).
Taylor, R.L. and Simo, J.C. Bending and membrane elements for the analysis of thick
and thin shells, Proc. of NUMETA Conference, Swansea, 1985.
Taylor, R.L., Finite element anlysis of linear shell problems, in Whiteman, J.R. (ed.),
Proc. of the Mathematics in Finite Elements and Applications, Academic Press, New York,
191-203 (1987).
Thompson, R., L., and Maffeo, R. L., “A Computer Analysis Program for Interfacing
Thermal and Structural Codes,” NASA Lewis Research Center, Report NASA-TM-
87021 (1985).
Trefethen, L.N., “Group Velocity in Finite Difference Schemes,” SIAM Review, 24, No.
2 (1982).
Tsai, S.W., and Wu, E.M., “A General Theory of Strength for Anisotropic Materials,”
Journal of Composite Materials, 58-80 (1971).
Tuler, F.R. and B.M. Butcher, "A Criterion for the Time Dependence of Dynamic
Fracture," The International Journal of Fracture Mechanics, 4, No. 4 (1968).
Turkalj, G. Brnic, J., and Prpic-Orsic J., “Large Rotation Analysis of Elastic Thin-Walled
Beam-Type Structures Using ESA Approach,” Computers & Structures, 81, 1851-1864,
(2003).
Van Leer, B., “Towards the Ultimate Conservative Difference Scheme. IV. A New
Approach to Numerical Convection,” Journal of Computational Physics, 23, 276-299
(1977).
Vawter, D., "A Finite Element Model for Macroscopic Deformation of the Lung,"
published in the Journal of Biomechanical Engineering, 102, 1-7, (1980).
von Neumann, J., and Richtmyer, R.D., “A Method for the Numerical Calculation of
Hydrodynamical Shocks,” J. Appl. Phys., 21, 232 (1950).
Walker, B.D., and P.R.B. Dallard, “An integrated Approach to Vehicle Crashworthiness
and Occupant Protection Systems”, SAE International Congress and Exposition, Detroit,
Michigan, (910148), February 25-March 1 (1991).
Wang, J. T., and O. J. Nefske, “A New CAL3D Airbag Inflation Model,” SAE paper
880654, (1988).
Wang, J.T., "An Analytical Model for an Airbag with a Hybrid Inflator", Publication
R&D 8332, General Motors Development Center, Warren, MI (1995).
Wang, J.T., "An Analytical Model for an Airbag with a Hybrid Inflator", AMD-Vol. 210,
BED 30, ASME, 467-497 (1995).
Whirley, R. G., Hallquist, J. O., and Goudreau, G. L., “An Assessment of Numerical
Algorithms for Plane Stress and Shell Elastoplasticity on Supercomputers,” Engineering
Computations, 6, 116-126 (June, 1989).
Wilkins, M.L., “Calculations of Elastic Plastic Flow,” Meth. Comp. Phys., 3, (Academic
Press), 211-263 (1964).
Wilkins, M.L., Blum, R.E., Cronshagen, E., and Grantham, P., “A Method for Computer
Simulation of Problems in Solid Mechanics and Gas Dynamics in Three Dimensions and
Time,” University of California, Lawrence Livermore National Laboratory, Rept.
UCRL-51574 (1974).
Winters J.M. and Stark L., "Estimated mechanical properties of synergistic muscles
involved in movements of a variety of human joints,": J Biomechanics 21:1027-1042,
(1988).
Yunus, S.M., Pawlak, T.P., and Cook, R.D., “ Solid Elements with Rotational Degrees of
Freedom, Part I-Hexahedron Elements,” To be published, (1989).
Zajac F.E., "Muscle and tendon: Properties, models, scaling, and application to
biomechanics and motor control, "CRC Critical Reviews in Biomedical Engineering
17(4):359-411, (1989).