(Manufacturing Engineering and Materials Processing 59) Suresh G. Advani, E. Murat Sozer-Process Modeling in Composites Manufacturing-Marcel Dekker (2003)
(Manufacturing Engineering and Materials Processing 59) Suresh G. Advani, E. Murat Sozer-Process Modeling in Composites Manufacturing-Marcel Dekker (2003)
(Manufacturing Engineering and Materials Processing 59) Suresh G. Advani, E. Murat Sozer-Process Modeling in Composites Manufacturing-Marcel Dekker (2003)
in Composites Manufacturing
Suresh G. Advani
University of Delaware
Newark, Delaware
E. Murat Sozer
Koc University
Istanbul, Turkey
Headquarters
Marcel Dekker, Inc.
270 Madison Avenue, New York, NY 10016
tel: 212-696-9000; fax: 212-685-4540
The publisher offers discounts on this book when ordered in bulk quantities. For more information,
write to Special Sales/Professional Marketing at the headquarters address above.
Neither this book nor any part may be reproduced or transmitted in any form or by any means, elec-
tronic or mechanical, including photocopying, microfilming, and recording, or by any information
storage and retrieval system, without permission in writing from the publisher.
EDITOR
loan Marinescu
University of Toledo
Toledo, Ohio
FOUNDING EDITOR
Geoffrey Boothroyd
Boothroyd Dewhurst, Inc.
Wakefleld, Rhode Island
Handbook of Induction Heating, Valery Rudnev, Don Loveless, and Ray Cook
Properties and performance of products made from fiber reinforced composites depend on
materials, design, and processing. This book is about polymer composites processing. Three
decades ago our understanding of mass, momentum, and energy transfer during composites
processing was nonexistent. As a result, almost all manufacturing was based on experience,
intuition and trial and error. We have come a long way since then. Many researchers did
delve into this difficult and poorly understood area to uncover the physics and chemistry of
processing and to develop the fundamental and constitutive laws to describe them.
There is currently a wealth of literature on modeling and simulation of polymer com-
posite manufacturing processes. However, we felt that there was a need to systematically
introduce how one would go about modeling a composite manufacturing process. Hence,
we focused on developing a textbook instead of a researcher's reference book to provide an
introduction to modeling of composite manufacturing processes for seniors and first-year
graduate students in material science and engineering, industrial, mechanical, and chemical
engineering. We have explained the basic principles, provided a primer in fluid mechanics
and heat transfer, and tried to create a self-contained text. Many example problems have
been solved to facilitate the use of back-of-the-envelope calculations to introduce a scientific
basis to manufacturing. The end of each chapter has questions and problems that reinforce
the content and help the instructor. "Fill in the Blanks" sections were created by Murat
Sozer to add to the qualitative knowledge of process modeling of composites manufacturing
that will develop the "experience base" of the manufacturing, materials, and design engineer
or scientist.
A project of this magnitude obviously cannot be realized without the help of others.
First, we thank Mr. Ali Gokce, graduate student at the University of Delaware, who created
many of the graphics in this book. Diane Kukich helped in technical editing. Of course we
thank all the graduate students in our research group who over the years have helped create
the research and the science base to develop models of composite manufacturing processes.
We would especially like to mention Petri Hepola, Steve Shuler, Terry Creasy, Krishna
Pillai, Sylvia Kueh, Simon Bickerton, Hubert Stadtfeld, Pavel Nedanov, Pavel Simacek,
Kuang-Ting Hsiao, Gonzalo Estrada, Jeffery Lawrence, and Roopesh Mathur. Some of the
examples and figures used in the book were first developed with their help.
The book contains eight chapters. The first two introduce the composite materials and
manufacturing processes. Chapters 3-5 provide the tools needed to model the processes,
and Chapters 6—8 apply these tools to some of the well known manufacturing processes.
Preface
1 Introduction
1.1 Motivation and Contents
1.2 Preliminaries
1.3 Polymer Matrices for Composites
1.3.1 Polymer Resins
1.3.2 Comparison Between Thermoplastic and Thermoset Polymers
1.3.3 Additives and Inert Fillers
1.4 Fibers
1.4.1 Fiber-Matrix Interface
1.5 Classification
1.5.1 Short F
1.5.2 Advanced Composites
1.6 General Approach to Modeli
1.7 Organization of the Book
1.8 Exercises
1.8.1 Qu
1.8.2 Fill in the Blanks
Introduction
1.2 Preliminaries
Composite materials generically consist of two different materials that are combined to-
gether. In engineering, the definition can be narrowed down to a combination of two or
more distinct materials into one with the intent of suppressing undesirable constituent
properties in favor of the desirable ones. Atomic level combinations such as metal alloys
and polymer blends are excluded from these definitions [9]. However, with the invention
of nanocomposites, one can probably group alloys and blends also under the umbrella of
composites.
In polymer composites, the individual constituents are polymer resin and fibers as shown
in Figure 1.1. The role of the polymer resin, which is also called the matrix phase of
the composite, is to primarily bind the fibers together, give the composite a nice surface
appearance in addition to environmental tolerance and provide overall durability. The fibers,
also known as the reinforcing phase, carry the structural load, reduce thermal stresses and
provide macroscopic stiffness and strength [8, 9]. The polymer matrix is either a thermoset
or a thermoplastic material. The fibers are made from glass, carbon or polymer. Some of
the fiber forms are shown in Figure 1.2.
From the processing and manufacturing viewpoint, the type of matrix plays an important
role. Thermoset materials are only 50 to 500 times more viscous than water and can
impregnate the empty spaces between the fibers readily. They do require an additional
processing step which involves chemical reaction of cross-linking the polymer chains known
as curing. This is schematically shown in Figure 1.3. On the other hand, thermoplastic
materials do not require this step but are highly viscous. Their viscosity can be as high
as a million times more than that of water. Hence, it is difficult to make them flow and
fill the tiny empty spaces between the reinforcing fibers. Figure 1.4 displays the important
differences between thermoplastics and thermosets. The constitutive equations that describe
the chemorheology of the matrix materials such as the influence of temperature, shear rate
and degree of cure on the viscosity will play an important role in the processing step during
•
M» JjnK
•""• X"» -
Fabric
\ Resin
Composite
FIBERS
Thermoplastics Thermosets
Recycles Cures
High Viscosity Low Viscosity
i: v
.-.:§V^:5
; '";£ vi* f$K£ V-i
V /* • ". •• • i *& ?'.' •* -..-"^^x
, • ** ».**•**. «"T »*«. ^t* '• I'.*'*""»•'• *."* •*- IT
Weave Fabrics:
Plain
(1 over, 1 under)
8 Harness 5 Harness
(7 over, 1 under) (4 over, 1 under)
I I I I 1
Relative
0) Processability
o
c
o Relative
o Performance
(Strength/Stiffness)
I
o
0)
0)
(0
LU
Figure 1.7: Schematic of role of fiber form on processing and performance [11].
(c) A
<•»'
o
Tm Temperature
10
CD
CL
10
10
Y(s 1 )
Figure 1.10: Change in viscosity as a function of temperature and applied shear for
polypropylene resin [9].
In Figure 1.8(c), one can notice the slightly cross-linked structure of an elastomer. Its
structure can be compared and modeled as a network of springs to capture their rubber
like character. Figure 1.8(d) shows the molecular structure of the family of thermosets.
Both thermosets and elastomers always cure to form an amorphous structure. Thermosets
initially consist of long chain molecules with weak bonds. Chemical reactions can initiate
covalent bonds that cross-link and cannot be melted. The high density of cross-linking
within the thermoset structure is responsible for the superior thermal stability and the
Table 1.1: Properties and typical applications for commonly used polymers [14].
Name / Abbreviation Material Application Important Application
Family Temperature Properties Examples
Range (C)
Polyethylene / PE Thermoplastic 100 Low strength, high Bottles, fuel tanks,
ductility, resists most sealing material, tub-
chemicals ing, plastic films
Polypropylene / PP Thermoplastic 110 In general, better Suitcases, tubing, en-
properties than PE closures, bottles
Polyvinylchloride / PVC Thermoplastic 60 Good chemical resis- Flooring material,
tance plastic films, tubing
Polytetrafluorethylen / PTFE Thermoplastic -200 to +270 Highest chemical Lab. Equipment,
resistance, strongest coatings (pans, etc.)
anti-adhesive
Poliamide / PA Thermoplastic -40 to +120 High strength and Ropes, bearings,
ductility gears, dowels
Polyethylentherephthalate / PET Thermoplastic 110 Low creeping ten- Most soda bottles,
dency, clear miniature parts,
parts with small
tolerances
Polycarbonate / PC Thermoplastic -100 to +130 High strength and Visors for helmets,
ductility, clear safety glasses, quality
flatware
Polyacrylate / PMMA Thermoplastic 70 Excellent optical Magnifying glasses,
properties (organic lenses of all kind,
glass), easy bonding showcases
Polystyrole / PS Thermoplastic 60 High strength, brit- Wrapping film, low
tle, glass clear, low quality cutlery and
chem. resistance plastic cups
Unsaturated Polyester / UP Thermoset 100 to 180 High tensile strength Structural parts for
(close to steel), good boats and cars, fish-
chemical resistance ing rods
Epoxy / EP Thermoset 80 to 200 Structural parts in Helicopter rotor
airplanes with high blades, fuselage,
demands for stiffness commonly used as
and strength adhesive
Phenolics / PF Thermoset 150 High strength and Enclosures, printed
stiffness, brittle circuit boards
Vinylesters / VE Thermoset 200 Room curing, high Applications in ma-
chem. resistance, rine industry, corro-
good strength and sion resistant tanks
ductility and pipes
Table 1.3: Summary of differences between thermoplastics and thermosets from processing
viewpoint.
Characteristic Thermoplastics Thermosets
Viscosity High Low
Initial state Usually solid Usually liquid
Post processing None Heat necessary
Reversibility Can be remelted and re- Once formed, vir-
formed gin state cannot be
recovered
Heat transfer requirement Heat needed to melt it Heat may be need to ini-
tiate cure
Processing temperature Usually high Can be at room temper-
ature
Usage Large volumes in injec- Mainly used in ad-
tion molding vanced composites
Solidification Cooling for change of Extraction of exother-
phase mic heat during curing
1.4 Fibers
Reinforcements carry structural loads and provide stiffness and strength to the composite.
They can be in the form of particles, whiskers or fibers. Particles or flakes have a low aspect
ratio of the diameter to the length, whiskers are usually 0.1 micron in diameter and made
from a single crystal. Particles do not provide a substantial change in mechanical properties
and whiskers are too expensive to manufacture.
Fibers are usually spun from a solution or a melt which orients the molecules of the
material. They are made from either glass, carbon or polymer. Their diameter is usually
less than 10 /um. Fibers come in various forms, shapes and materials and are primarily
used for reinforcements. Composites containing continuous fibers are known as advanced
composites. Composites containing discontinuous fibers are called short fiber or long fiber
reinforced plastics.
For advanced composites, the fibers are used in the form of rovings, yarns, strands and
tows. These yarns or tows can be combined in various forms to create a preform. Some of
these structures are shown in Figures 1.5 and 1.6.
Figure 1.11: A network of fiber tows containing 1000 to 2000 fibers in each tow stitched
together to form the fabric.
Many preforms are formed using fabric reinforcements. Fabrics are formed from a net-
work of continuous fibers. One large class of fabrics is manufactured by either weaving or
stitching together bundles ("tows") of fibers. These tows are generally elliptical in cross
section, and may contain from 100 to 48000 single fibers as can be seen in Figure 1.5. The
cross sectional width and thickness of tows are of the order of millimeters, as can be seen
in Figure 1.11. Another large class of preform fabrics include "chopped" and "continuous
strand" random mat also shown in Figure 1.5. These fabrics are typically formed from
low cost E-glass fibers, cheaper than woven and stitched fabrics, and used for low-strength
Figure 1.12: Fiber preform constructed from woven fabrics and placed in a mold that forms
a skeleton of the composite part.
1.5 Classification
Thermosets have been around much longer than thermoplastic materials; hence almost all
manufacturing techniques developed for thermoplastics today were originally derived from
the processes that used thermoset matrices, and the most important ones are listed in Table
1.4. The practice of choosing an appropriate manufacturing method is usually based on the
actual part size and geometry, the unit count, the precursor material (initial state of the
composite material), the selected components of the composite, i.e. the reinforcement and
the matrix, and the cost.
The first generation of composites used chopped or short fibers preimpregnated with
the thermoplastic polymer matrix in the form of pellets. The pellets are usually a few
centimeters in length and a few millimeters in diameter.
The composite types can be broadly divided into composites made from short fibers
(aspect ratio less than 100) and continuous fibers. The three most common mass produc-
tion processes for short fiber composites manufacturing are injection molding, compression
From the analysis, it was evident that composites with continuous fibers could enhance
the mechanical properties by one to two orders of magnitude as compared to short fiber
composites. These composites are referred to as advanced composites. There are several
primary steps that are common in manufacturing of advanced composites. First, all ad-
vanced composites require a skeleton structure of the fibers or fiber network that is tailored
for the particular part geometry and the property requirements. Second, this fiber structure
must be covered and impregnated by the liquid resin in some way. Finally, the part should
be supported by a rigid tool to allow the resin to solidify or cross-link, permanently freezing
the microstructure created by the fiber network.
Many fiber structures are available as seen from Figures 1.2, 1.5 and 1.6. Broadly, they
can be divided into two groups. The first group consists of continuous fibers in sheets, tapes
or tows aligned in one direction. These fibers may be prewet with the resin and then laid in
different directions by hand or by a machine to construct the desirable structure. Aligned
fibers allow for creation of very high fiber volume fraction and hence high specific in-plane
strengths and stiffnesses. The second group uses fiber interlacing to create two-dimensional
and three-dimensional interlocked textile structures as shown in Figure 1.6. This can allow
the composite to achieve higher stiffness and strength both in and out of plane directions,
and potentially allow the designer to tailor the mechanical or physical properties to the
desired application. However, as shown in Figure 1.14, the degree of complexity that can
be handled with short fiber composites cannot be duplicated with advanced composites.
Thus, the material manufacturers started to make resin impregnated prepregs with
continuous aligned fibers that could be bonded and fused together to make an advanced
composite. Textile preforms were being made where the fiber tows could be woven, stitched
or braided together to create the underlying microstructure to provide strength to the com-
posite. Various techniques were invented to induce the resin to wet the fibers and infiltrate
the empty spaces between the fibers to retain the integral structure of the composite. This
started the evolution of new composite manufacturing processes to make advanced com-
posites. The details of all the processes are explained very well in [9]. Here, we will briefly
introduce the general approach to model the processing step of composites manufacturing,
and in the next chapter we will briefly introduce some of the important manufacturing
processes and identify the underlying physics of transport of mass, momentum and energy
in these processes.
(a)
(b)
Figure 1.14: (a) Injection molded short fiber component with very high degree of geomet-
ric complexity [16]. (b) Resin transfer molded advanced composite with modest level of
geometric complexity [17].
The manufacturing process physics and modeling are greatly influenced by the type of
fibers being used: short or long, continuous or discontinuous, aligned or interlaced, etc. The
type of resins being used, thermoplastic or thermoset, also influences the process. Thus,
the fiber form and the matrix type play a key role in the selection of the manufacturing
process. The geometry of the part to be manufactured influences the decision and also if
the process is carried out in an open mold or a closed mold. These choices influence the
physics of mold filling.
The process modeling step in composite manufacturing is generally approached by re-
searchers on two scales. The macroscale is usually the order of the smallest dimension of
the composite being manufactured (millimeters). The microscale scale is more on the order
of a fiber or tow diameter (microns). At the macroscale, the modeler is generally inter-
ested in the overall relationship between the process parameters (such as pressure, flow rate
and temperature) and global deformation of the composite material that is being formed.
One can use a continuum mechanics approximation to describe this physics. However, as
composite materials are heterogenous materials by definition, macrolevel physics cannot
capture phenomena that occur on the scale of a fiber diameter (usually a few microns).
Hence one may need to model this physics separately and find an approach to couple it
with the macroscale physics.
1
Sizing is a chemical coating that is applied on the surface of the fibers, sometimes by grafting the
molecules on the surface in order to improve the adhesion between the resin and the fiber.
1.8 Exercises
1.8.1 Questions
1. What are the advantages of polymer composites over other materials?
2. List a few industries that use polymer composites.
3. When polymer composites were used a few decades ago, did the process engineers
rely on (i) experience and trial and error approaches, or (ii) accurate mathematical
modeling of process physics, in order to improve the manufacturability of a certain
prototype?
4. What are the two major ingredients of a composite material? How do they enhance
the properties of the composite?
5. What are the two types of polymer resins used in composites processing? What are
major differences between them?
Overview of Manufacturing
Processes
2.1 Background
Several industries have been using fiber reinforced composite materials for a few decades
now. Glass fibers were available commercially in the 1940s. Within a decade, composites
were being used by several industries; for example the automobile industry was producing
polyester panels with approximately 25% glass fibers [18].
Manufacturing with composite materials is very different from metals. This is because
when making a metal part, the properties of the virgin material and the finished part
are fundamentally unchanged. For composites, the manufacturing process plays a key role.
During composite processing, one makes not only the part of the desired shape, but also the
material itself with specific properties. In addition, the quality of the composite material
and the part fabricated depends on the manufacturing process, because it is during the
manufacturing process that the matrix material and the fiber reinforcement are combined
and consolidated to form the composite.
In early stages of development, the' cost of composite materials was very high and only
selected industries, for which the importance of the property of the material greatly out-
weighed the cost factor, were willing to use them. These industries were primarily aerospace
and the aeronautical industries. They valued the properties of the composites greatly and
could justify the higher costs because of the weight savings. Both industries took advan-
tage of the light weight and, in the case of defense oriented projects, the stealth properties.
The lack of automated and repeatable manufacturing processes drove the cost of composite
parts up and limited the number of potential users. Another industry that has been using
composites since the 1970s is the marine industry. It could deal with low production volume
and relatively high costs while taking advantage of the corrosion resistance properties of
composites. The majority of the manufacturing work done in these industries in the 1970s
was very labor intensive and not very cost effective as the manufacturing modus operandi
was "experience" and trial and error.
It was imperative that in order for composites to be widely used, especially by the con-
sumer goods industry, such as the automotive and sports industry, two major goals had to
be achieved. First of all, the cost of raw materials had to go south. Secondly, and most
importantly, manufacturing methods had to be developed to achieve high-volume produc-
tion that relied more on the fundamental understanding of the physics of the process rather
Force
Fluid Phase
Solid Phase
Mold
Part
zoom
Precursor Materials
The filled thermoplastic pellets usually contain a second, discontinuous, usually more rigid
phase blended into the polymer. When the aspect ratio (ratio of largest to smallest di-
mension) of the second component is around one, it is referred to as a filler. If the aspect
ratio is one to two orders of magnitude larger, then it is called a reinforcement. The most
commonly used reinforcements are particles, wiskers and short fibers usually less than one
inch in length. The parts obtained usually have a fiber volume fraction between 30% and
40%. Filled or reinforced materials provide much different properties than the base resin.
For example, reinforced polypropylene provides higher rigidity and lower warpage charac-
teristics than neat polypropylene. Also, the viscosity of the filled resin will be different in
magnitude and sometimes anisotropic as compared to the neat resin. In practice, fibrous
reinforcements used with glass fibers dominate the market although the carbon and aramid
fibers provide higher stiffness and strength but are seldom used due to the high cost of raw
materials.
The traditional injection molding process limits the fiber length that solidifies in the final
part since the high shear rates in the barrel and the passage of fibers through narrow gates
and openings in the mold cause significant fiber attrition. Usually, the fiber diameter is of
the order of a few microns, and the final length distribution, irrespective of the starting
fiber length, is of the order of 50 to 500 /mi. The starting length of these fibers in the
log-like pellets is usually of the order of 1 to 3 mm. As a result, new methods to produce
pellets containing longer fibers were developed in which the fibers were pultruded and stayed
bundled together and were not dispersed in the pellet by the action of compounding. These
pellets (see Figure 2.4) produced final parts that retained a higher percentage of longer
fibers and consequently showed a significant increase in modulus and impact toughness.
The thermoplastic matrix material selected also plays a role in the final physical and
optical properties. Most thermoplastic materials when they solidify do so as an amorphous
matrix or exhibit various degrees of crystalline behavior depending on the thermal history
the material undergoes during the injection molding process. Samples of crystalline and
amorphous materials are shown in Figure 2.5. More details on crystallization of thermo-
plastics will be discussed later.
Figure 2.5: Crystalline and amorphous materials, (a) Spherulite in malonamide containing
10% d-tartaric acid (low crystallinity). (b) Spherulite grown in mixture of isotactic and
atactic polypropylene (high crystallinity) (redrawn from [30]).
Transport Issues
The issues that relate to transport phenomena in this process are the flow of fiber suspen-
sions as they occupy the closed mold, the orientation of the fibers during flow, fiber length
distribution, fiber breakage and the heat transfer that changes the microstructure of the
resin. So, if we consider mass conservation, we have to account for the mass balance of
the suspension which can be treated as a continuum material at least for the short fiber
materials as the fiber length is of much smaller scale as compared to the domain dimen-
sions. One also needs to characterize and describe the orientation of the fibers in a flowing
suspension. The physical concept that one may have to invoke here is the conservation
of the orientation field, which simply states that if the orientation of the fibers disappears
in one direction, it should reappear in some other direction. One also needs to conserve
the momentum to describe the flow and the pressure field during the flow process. This
requires one to describe the constitutive equation between the stress applied and the strain
rate experienced by the material. For Newtonian fluids, this is usually constant and the
constant of proportionality is called viscosity. However, as the thermoplastic melts are shear
thinning, the viscosity is known to decrease with shear rate, and the addition of fibers can
change the stress strain rate behavior and even make it anisotropic. One needs a rheologi-
cal equation to describe this behavior. The energy conservation allows one to describe the
temperature history of the melt in the channel between the screw and the barrel, where it
gets its heat input from the heaters on the barrel and due to viscous dissipation caused by
the shearing of the suspension. It also allows one to keep track of the cooling history in
Figure 2.6: Schematic of fiber clustering for long fiber suspensions [31].
The coupling between the energy transport and the momentum creates a fountain flow
mechanism in injection molding. As the walls are cooler than the core, the suspension
viscosity is higher near the walls (polymer viscosity increases as temperature decreases) as
compared to the core. Hence, under the same pressure, the suspension in the core moves
ahead of the suspension near the walls, spreading from the center outwards like a fountain,
as shown in Figure 2.7.
Flow
direction
Figure 2.7: Schematic of fountain flow effect encountered during filling (redrawn from [32]).
The fibers align in the direction of shearing and also in the direction of stretching as
shown in Figure 2.8. The shear flow near the mold walls aligns the fibers in the direction of
I
s<n
''
Flow
direction / \/ \ / / \ \ / \ / / \ / \ / \/ \\ )
i i i \
core
} skin
Another phenomenon associated with flow and fibers at the micro level has been that
the fibers at the flow front tend to align along the flow boundary and not across it. This may
have to do with the surface tension phenomena between a solid, polymer and air. However,
because of this phenomenon, whenever two flow fronts meet, the boundary is called a weld
line and is usually the weak link for the mechanical properties because the fibers align along
the weld line and not across it, as shown in Figure 2.9. Thus, understanding of this issue
could help modelers address and develop flow management techniques to create strength
along the weld lines.
Applications
Nearly 20% of the goods manufactured nowadays use injection molding due to its versa-
tility and low cost. However as short fiber composites can improve the desired physical,
optical and mechanical properties, structural integrity and dimensional stability, injection
molding machines and the screw geometry were modified to handle fibers along with the
polymer. Many applications such as housing for electric tools, automotive parts under the
hood, plastic drawers, metal inserts and attachments, seats in airplanes, etc. are routinely
manufactured using injection molding with thermoplastic pellets containing discontinuous
fibers.
The fundamental advantage of injection molding is the ease of automating the process
and the short cycle times, which together allow for the possibility of high volume production.
In addition, molds can also be constructed to make more than one part at a time. The
major disadvantages are the high initial costs of the capital equipment and the molds and
the variation in part properties due to lack of control of fiber orientation and distribution.
,• Weld line
Injection (Fibers are aligned
gate —> parallel to weld line)
Figure 2.9: Schematic of flow front locations and weld line with fibers aligning along them
during injection molding.
2.3.2 Extrusion
Extrusion resembles injection molding because it contains a screw. The main difference
is that there is no closed mold in extrusion. Instead .a die is used to shape the polymer
suspension into specific cross sections. This process is used to plasticize and compound
polymer pellets containing short fibers and also for manufacturing continuous parts with
different cross sections. As in the case of injection molding, the screw melts the polymer and
acts as a piston to push the suspension into the die geometry. Inside the die, the suspension
takes the form of the die cross section and exits from the other side of the die and can be
continuously pulled to make long tubes, I-beams and reinforced pipes.
Solid pellets
Hopper
k) . .
Extr jder 1pie|
Sizer Cooler
to Cutter
Puller
Process
Figure 2.1C shows a simple sketch of an extrusion line. The process starts with a hopper
into which one pours solid pellets. The extruder melts the plastic (resin or polymer) and
may seriously cause fiber attrition. It pumps the fiber suspension through a die hole of the
desired shape. It then enters a sizing and cooling trough where the correct size and shape
are developed. Next, the formed product enters a puller which pulls it through the sizer.
Screw , Flight
Breaker plate
Transport Issues
The action of the screw as a pump is a complex process and involves drag flow and pressure
driven flow acting against each other. Hence the important transport issues are to under-
stand the interaction between the drag and pressure driven flow and their role in calculation
Dn,
Die swelling (see Figure 2.12) and melt fracture are significant processing issues for ex-
truded structures and with the addition of fibers can distort the final cross section. Trans-
port processes and their interaction with surface tension and normal stresses will allow one
to study and hopefully understand this relationship.
Applications
Extrusion is considered as the positive displacement pump for producing over six billion
polymer products each year. A partial list of extruded products includes films, pipes,
tubing, insulated wire, filaments for brush bristles, profiles for home siding, storm windows
and gaskets, etc. [35]. The process of making these products is termed extrusion. The
processing physics of flow and heat transfer of plastic melts in an extruder have been
studied in detail by many researchers [36, 37]. In the last few years, the process has been
slightly modified to allow extrusion of polymers containing reinforcements.
Continuous o
Strand Roving Chopper
Resin/Filler
10 Paste ^ Carrier Film
^ Take-up Roll
Figure 2.13: Schematic of sheet molding compound (SMC) production (redrawn from [38]).
To form the sheets of SMC material a specific procedure is used as shown in Figure 2.13.
A thin layer of resin is placed on a sheet of nonporous material, such as nylon. As the
nylon moves along the production line, fibers are added to it; these can be in random,
unidirectional, or other orientations. Next a layer of resin, placed on a cover sheet, is
applied onto the fibers so that the resin is in contact with the fibers. This sheet, enclosed
between the nylon sheet and the cover sheet, is then passed through several compaction
rollers. These serve two main purposes: they mix the resin and fibers together, and they
compact the sheets. The resin is now in a continuously changing state (i.e., it is slowly
curing); it is left to thicken for approximately 5 days, after which the SMC is ready. At
this point the SMC sheet must be used within a certain period of time, which can be
up to several weeks, and must be stored under certain environmental conditions, such as
low humidity. Several types of SMC are currently used in industry: SMC-R (reinforced
with fibers oriented randomly), SMC-C (reinforced with unidirectional continuous fibers),
SMC-C/R (reinforced with both randomly oriented and continuous unidirectional fibers),
SMC-D (reinforced with directional but discontinuous fibers). It is possible to use both
thermoplastics and thermosets in SMC, but the majority of SMC is done using thermosets
Process
The SMC or the thermoplastic material is stacked into the mold cavity. This is referred
to as the initial charge. The initial charge shape and its placement location in the mold
are crucial parameters as they influence the final properties of the product. Sometimes this
charge is preheated using dielectric sensors before closing the mold to initiate the flow. The
temperature field that results from this stage is part of the initial condition for the mold
filling stage. Mold filling begins when the polymer begins to flow and ends when the mold
cavity is filled. The heated top and bottom platen containing the two halves of the mold
cavity are brought together, generating heat and pressure to initiate the flow. Temperature
ranges between 135 and 160 degrees C, pressures between 3.5 and 15 MPa and cycle times
between 1 and 6 minutes. The amount of flow in compression molding is small but critical
to the properties and the quality of the part because the flow controls the orientation of
the short fibers and the final orientation pattern is what will determine the physical and
mechanical properties of the composite.
In-mold curing describes the next stage in compression molding of thermosets, where
the liquid resin starts to gel and cross-link and forms a solid part. The curing may initiate
during mold filling stage but the bulk of the curing takes place after the mold is filled. The
part is removed from the mold as soon as it is solid-like and may be placed in an oven for
post-cure to complete the curing process.
The mold is usually made of steel and it is hardened in key areas where the mold can
wear out more easily. This is important because the mold is subject to high pressures and
temperatures and also undergoes many cycles continuously. For these reasons mold design
is very important, and the overall cost of the molds is usually high. Several different resin
systems can be used in SMC. Vinyl ester and polyester are the most used in the automotive
industry, while epoxy resins are widely used in the aerospace industry [41]. There are several
variations and modifications that different industries have developed over time, in order to
improve the process and to tailor them to their own needs. The automotive industry, for
example, has a specific need for parts with excellent surface finish. For this, a technique
known as in-mold coating was developed.
In in-mold coating, after the part is partially cured inside the mold, the mold is opened
slightly and a resin, such as a urethane, is injected in the mold. Subsequently the mold is
closed again, causing the resin to coat the outside of the part, filling any voids on it. This
greatly improves the surface finish on the part and can save several stages in the painting
process.
Finished part
Performance of the manufactured part is tied to the flow, heat transfer and chemical
reactions which occur in the mold. For example, initial charge shape and location will change
the flow pattern, which in turn will influence the fiber orientation. The fiber orientation
will influence the physical and mechanical properties.
Platen
Exhaust
Figure 2.15: Composite sheet forming processing methods: (a) diaphragm forming, and (b)
matched die forming [42].
In diaphragm forming as shown in Figure 2.15(a), the blank is held between two dis-
posable, plastically deformable diaphragms of either super-plastic aluminum or polyimide
polymer. During the forming cycle, the diaphragm edges are clamped, heated along with
the blank and deformed through the use of air pressure to the tool surface. The diaphragms
serve to hold the blank in tension and prevent fiber buckling that can occur under com-
pressive stresses. When forming parts containing continuous fiber reinforcement, the di-
aphragms are clamped but the blank cannot be. This is due to the inextensibility of the
Figure 2.16: Linear beams can be stretch-formed into curved sections with favorable me-
chanical properties since the fibers follow the curvature of the beam [42].
The key to successful stretch forming is precise control over the final fiber placement.
This is achieved by clamping both ends of the unformed part, heating up the portion
between the clamps and then carefully forming the stock shape to the desired curvature
[42].
Process
The process involves dragging a combination of fiber and matrix materials from a supply
rack through a temperature controlled tool, which will determine the final part geometry.
Transport Issues
The transport phenomenon inside the heated die is of interest as most of the redistribu-
tion of the resin and the consolidation of the fibers and the resin takes place there. The
precursor material used will influence the transport phenomena modeling in this process.
Several different types of thermoplastic preforms can be used with this process. They are
either continuous fibers completely impregnated in the shape of thermoplastic tapes (e.g.
CF/PEEK) or glass fibers embedded in thermoplastic powder and enclosed by a thermo-
plastic tube of the same material or commingled polymer and reinforcing fibers. Figure
2.19 depicts available materials for usage in the thermoplastic pultrusion process.
'.U'VI | i r -
Figure 2.19: Schematic of different available precursor material for the pultrusion process
[13].
The die assembly for this process has two distinct sections. The first section is the
heated and the tapered entry region (Figure 2.20), which collects the preheated preform,
rearranges the fiber bundles to the desired shape and melts the polymer. The function of
the taper is to consolidate the preform, thus encouraging elimination of voids and complete
impregnation of the fibers with the polymer matrix. The second section of the die assembly
is known as the land region. It is of uniform cross section without a taper and may have
cooling lines attached to it. The role of this region is to solidify the matrix material forming
the final shape of the pultruded part. It is usually cooled to approximately 50-80 degrees C
(around the glass transition point (TG) of the thermoplastic matrix). The process velocity is
determined by the speed of the pulling system located immediately behind the die assembly.
The last step in this production process is to cut the final product into pieces of the desired
length.
Figure 2.20: Schematic of the tapered section of a standard pultrusion die [13].
There are two important parameters the modeling of this process should be able to
predict: (1) the pulling force required to run the operation at a reasonable speed to produce
parts that are free of voids and contain the desired fiber volume fraction, and (2) the desired
level of crystallinity in the matrix with minimal stress concentration in the heating and
cooling profile of the die. Hence the viscous flow physics and the heat transfer during the
process will play an important role in the determination of these key parameters such as
In this process, the tool, called the tow placement head, is designed to conform to the
geometry of the composite structure. Processing of thermoplastic composites is based on
melting and solidification of the matrix. The matrix requires energy input for melting and
energy extraction for solidification from the system. The method of energy transport can be
global where all of the thermoplastic matrix is melted as in compression molding, injection
molding, extrusion, pultrusion, etc., or can be localized as in filament winding and tape
lay-up where only a portion of the matrix is melted.
Process
In this process, 3-mm to 12-mm wide thermoplastic tape preimpregnated with continuous
fibers is placed on the tool surface (if it is the first layer) or on the substrate (previously
deposited material on the tool). The incoming tape or tow and the previously deposited
material on the head (the substrate) are preheated by laser, gas, or any other methods of
concentrated localized energy. Rollers are used to initiate intimate contact and consolidate
the incoming tape to the substrate below. The localized nature of heating demands the
consolidation process also be localized, and it is commonly referred to as in-situ consoli-
dation. Additional, local energy may be provided to heat all the layers underneath in the
thickness direction to further improve the overall degree of bonding, healing and intimate
contact. Void content within the tow and the substrate decreases under the pressure of the
consolidation rollers.
In most industrial applications the thermoplastic tape lay-up process is automated and
also known as automated tow placement (ATP) process. In this process, a relatively thin
tape is consolidated on a substrate under the application of heat and pressure (Figure 2.21).
In most cases, the feed tape, the heater (gas, induction, laser etc.) and the consolidation
rollers/shoes traverse the substrate at a predefined path and velocity.
The material used mostly in aerospace industry is carbon fiber preimpregnated with
PEEK or PEKK thermoplastic matrix. Typical applications are the fuselage and wing
structures of the aircraft. One of the important objectives and advantages of the ATP
process is to eliminate the use of a huge autoclave in order to make the process more cost
effective. Also, the ability to make the part out of the mold is attractive, in addition to
having the capability to create multi-axial laminates and moderately complex structures.
The downside is the investment required in automation and the cost of the tool head.
The demand to be cost effective forces the process to be conducted at the maximum
allowable speed. This requires optimization of the process parameters at desired processing
speed while maintaining the quality of the product. To achieve such goals, a fundamental
understanding of the process is necessary.
Transport Issues
Different aspects of the process that can be modeled are intimate contact, polymer heal-
ing and consolidation [47, 48, 49, 50]. To model these processes, one must quantify the
squeeze flow during application of the rollers and the transient heat transfer that governs
the temperature of the resin and affects its viscosity. The heat transfer during cooling in-
fluences the microstructure, entrapment of voids and the quality of the part. The quality
of the part also depends on melting, consolidation, solidification and through the thickness
temperature gradient.
Issues related to consolidation are intimate contact, void reduction and migration, gap
reduction between adjacent layers, adhesion and diffusion of matrix chains. The key issues
for modeling include tow-placement head configuration, consolidation, bonding and the heat
transfer between the incoming tape and the substrate interface. A good bond between the
substrate and the incoming tape requires the interface temperature to be greater than the
melting temperature of the thermoplastic. The temperature gradient through the thickness
is responsible for residual stress development in the composite. The critical issues in the
thermoplastic tape lay-up (or ATP) process are the heating of the tape above the melt
temperature for good bonding with the previous "substrate" layer, without overheating to
prevent degradation. So, the rate of heat input is a critical process parameter and will play
a role of selection on the type of heater used. The heater width is also a parameter to
provide an optimum heating zone for good consolidation.
The consolidation pressure is also a very important parameter. Consolidation pressure
is applied for void reduction and adhesion to the previous layer. Low pressure may create
pools of resin and poor bonding. Excessive force can squeeze the resin out creating a resin-
starved region, and at the same time can deform the fibers, which will reduce the local
strength. Residual stress development in the part is an outcome of the localized heating
and cooling process described earlier and is an important issue. Intimate contact, diffusion
An autoclave is a large pressure vessel with a heating facility, or one can think of an autoclave
as a large oven with an integral pressurizing facility. A schematic of an autoclave is shown
in Figure 2.22.
composite under
vacuum
vacuum line
blower to
circulate air
table or
mold
\
heat source
Fiber preforms
Finished part
Tool
(b) (d)
Figure 2.23: Schematic of four main stages of an autoclave: (a) placement of adhesive, (b)
placement of fiber preform, (c) autoclave cure, and (d) finished part.
Process
Autoclave process is the earliest method used to make advanced composites for aerospace
applications. The four stages involved are shown in Figure 2.23. The material and tool
preparation stage is initiated by first covering the tool surface with a release film that
allows one to detach the composite from the tool surface readily. The next stage involves
cutting the prepreg (continuous unidirectional fibers partially impregnated with the uncured
thermoset resin) layers and stacking them in a desired sequence on the tool surface to form
the composite lay-up. This is accomplished usually by hand lay up, although great advances
have been made in the use of automated tow placement and automated tape lay up for the
stacking sequence step. For example, the Boeing 777 aircraft tail assembly used a variety
of automated processes [51]. Despite these advances, many parts in the aerospace industry
still rely on hand lay-up for this step. The composite lay-up is covered by peel plies, release
fabric and bleeder material in that sequence. Peel plies provide surface texture and release
fabric allows resin to flow into the bleeders. On top of the bleeder material is the breather
material. The breather material distributes the vacuum over the surface area. A vacuum
bag envelopes the tool, the part and ancillary materials for vacuuming.
The third stage, as shown in Figure 2.23, involves transferring the part into the au-
toclave and initiating the curing step by exposing the assembly to elevated temperatures
and pressures for a predetermined length of time. The goal is to consolidate and solidify.
The elevated temperatures provide the heat to initiate the cure reaction, and the applied
pressure provides the force needed to drain the excess resin out of the composite, consoli-
date individual plies or prepreg layers and compress the voids. For thermoset composites
this step is irreversible. Hence, it is necessary to subject the composite part to the correct
processing window of temperatures and pressures to ensure a quality part.
The temperatures and pressures are of the order of 100-200 degrees C and 500-600 kPa
respectively. Also, if one wants to manufacture a large part, a large autoclave is necessary.
As the autoclave is a pressure vessel, it is usually made as a cylindrical or axisymmetric
tube with a door at one end. As the autoclave must be strong at high temperatures as
well, the autoclave is an expensive piece of equipment usually made out of welded steel. An
Transport Issues
To address these critical issues, one needs to understand the mass, momentum and heat
transfer that the composite undergoes during the curing cycle. The temperature and the
pressure of the autoclave influences the temperature of the composite, the degree of cure
of the resin, the resin viscosity, the resin flow, fiber volume fraction of the composite,
the change in the void sizes, residual stresses and strains in the composite, and the cure
time. We will discuss in a later chapter how to develop a model for autoclave processing.
However, due to the extensive requirements of material data and batch-to-batch variation
in properties, one can apply these models effectively only to simple geometries. For complex
geometries one must combine models with sensing and control to produce successful parts.
The advantages of autoclave processing are that it can produce composite structures
with very high fiber volume fraction. Also, a lot of empirical data is available on this
process which makes it attractive when reliability outweighs the cost. Also once the most
Process
Resin transfer molding consists of a mold cavity that is in the shape of the part to be
manufactured. The fiber preform is placed in the cavity. The mold is closed and clamped or
held under pressure in a press. The resin is injected into the compressed preform through
one or more gates from a pressurized container. Once the mold is full, the injection is
discontinued and the resin is allowed to cure. This cure may be initiated by either heating
the mold which heats the resin as it flows into it, or by addition of inhibitors that initiate
the cure after a time interval allowing the resin to first complete the impregnation of the
preform. The mold is opened once the cure is complete or the part is sufficiently hardened
to be demolded. These steps are depicted in Figure 2.25.
RTM offers the promise of producing low-cost composite parts with complex structures
and large near net shapes. Relatively fast cycle times with good surface definition and
appearance are achievable. The ability to consolidate parts allows considerable time saving
over conventional lay-up processes. Since RTM is not limited by the size of the autoclave or
by pressure, new tooling approaches can be utilized to fabricate large, complicated struc-
tures. However, the development of the RTM process has not fulfilled its full potential.
For example, the RTM process has yet to be automated in operations such as preforming,
reinforcement loading, demolding, and trimming. Therefore, RTM can be considered an
intermediate volume molding process [52, 53].
VARTM and SCRIMP are slight modification of the process where the top half of the
mold is replaced by a vacuum bag (as in an autoclave) and a permeable layer is introduced
at the top or the bottom to facilitate the distribution of the resin throughout the part
1) Preform
Manufacture
5) De-molding
\ /
3) Resin V
Injection A
2) Preform
Compression 4) Resin Cure
Mold
Vacuum pump
Stepl Step 2
Resin impregnates
Resin injection fibers and cures
StepS Step 4
This process has replaced RTM for some applications due to its simplicity, low initial
capital investment and the ability to manufacture large structures such as bridge sections
and rail carriages as shown in Figure 2.27.
Also, the cost is kept low due to low pressures used in the manufacturing process and
the reactions being carried out at room temperature. It only needs one tool surface, and
the top surface is bagged as in autoclave processing which also cuts down on the tooling
costs. The disadvantages of VARTM are poor surface finish on the bagging side, limitation
to nearly flat structures, time involved in material preparation, poor dimensional tolerances
and lack of automation. The co-injection process, as shown in Figure 2.28, can use RTM
or VARTM process where two different immiscible resins are injected and co-cured to form
a composite containing different resins.
/ Separation layer
Vinyl ester
Figure 2.28: Co-injection process which can be RTM or VARTM where two different im-
miscible resins are injected into a fiber preform and co-cured to form a composite [58].
Compaction
Nip-Point Heater
Mandrel
W V
J V
V Tensioner
V
Figure 2.30: Schematic of the filament winding process [59].
Issues
The process variables that can be selected and controlled independently are the winding
speed, fiber tension, and external temperature or heating rates. Hence, the process model
should provide information regarding the mandrel temperature and the temperatures inside
the composite, degree of cure, viscosity, fiber positions, stresses and strains, porosity, and
winding and curing times. The transport of the resin here through a moving fiber bed and
also the cure kinetics reaction that changes the viscosity of the resin requires one to address
the flow, energy and reaction kinetics.
2.6 Exercises
2.6.1 Questions
1. What is the main difference between metal and composite parts in terms of the prop-
erties of finished parts and the raw materials that are used to manufacture them?
2. Why were composite materials used mainly in some selected industries, such as the
aerospace industry, during the early stages of composite material development?
3. What are the two major classes of composite manufacturing processes in terms of
types of molds being used? Give examples.
4. Describe the injection molding process by using all of these words and terms: high-
volume production, thermoplastic resin, solid pellets, short fibers, fillers, feeding, hop-
per, barrel, screw, melting, mixing, functions as a piston, sprues, runners, solidifica-
tion, and ejection.
5. What is a pellet? What are the ingredients? What is the size of a typical pellet?
6. Why is the length of fibers that solidify in the final part limited in the traditional
injection molding process? What is an alternative approach to overcome this? What
is the advantage of using longer fibers in this new approach?
7. While modeling the transport phenomena in the injection molding process, what phys-
ical conservation laws are used? What are the independent and dependent variables
in them?
8. What is "fountain flow" in injection molding? What process and material parameters
determine its significance? What is the result of this flow?
9. What are "skin" and "core" layers in injection molding?
The philosophy behind a good model should be "prepare for the worst and hope for
the best." The assumptions made must still maintain the behavior and features of the
system we are trying to model intact. Otherwise, the model may not properly represent the
physical system.
The general approach to buildind a model is illustrated in Figure 3.1.
Physical Process
Objective
Assumptions and
Simplifications
Mathematical Model
Solution Method
Analytic or Numerical
Results
Adequate Solution NO
,rYES
NO Agree with
Process Physics?
,,YES
Useful Predictions/Designs
Here, the total time derivative term on the left hand side of Equation (3.1) is carried inside
the integral, by using the Leibniz rule,1 as a partial time derivative. The additional terms
from the Leibniz rule dropped out since the control volume is fixed in space.
The rate at which fluid mass enters the control volume V through its boundary S is the
flux integral and is given by
q = - f pn-UdA, (3.2)
Js
where n is the unit outward normal to S as shown in Figure 3.2.
Figure 3.2: Control volume for the derivation of conservation of mass equation.
1
The Leibniz rule states that
-
at, A^ A(t)
s(x,y,z,t)dV (3.3)
.'
mass per unit time.
The rate of increase of mass within V is equal to the rate at which mass enters V through
<S, minus the rate at which mass is lost:
F = 0. (3.7)
Since the considered control volume V is assumed to be of any arbitrary shape, and at
any arbitrary location of the flow region, not only the integral, but actually the integrand,
dp/dt + V • (pU) + s must be zero, too, then
^ + V - ( p U ) = -S. (3.8)
This is the conservation of mass (or, also known as continuity equation) of fluid mechanics.
In a Cartesian coordinate system, it takes the following form by expanding V • (/oU)
since the control volume and hence its sides Ax, Ay and Az are fixed, i.e., not changing
with time. The rate of fluid mass inflow into the control volume V through its face at x is
(pux)\xAy&z. The rate of fluid mass outflow from the control volume V through its face
at x + Ax is (pux)\x+A.x&yAz. Similar expressions can be written through the faces at y
and z for inflows, and at y + Ay and z + Az for outflows as shown in Figure 3.3. Hence,
the net rate of flow is
Rate of \ / Rate of \ . . , ,
. _ — n = AyAz r\(pux)N x — (pux)N
mass inflow / \ mass outflow I
[(pu,
\ A /Y> A -v \( ™,
~ (Puy)
+AxAy [(puz - (puz) IZ+AZ] • (3.13)
If there is a sink within V which absorbs fluid mass at a rate of s(x,y, z,t) mass per
unit volume per unit time, then the net rate of lost mass within V is equal to
Inflow flux through ^x x (jt + Ax, y + Ay, z + Az).*^ Outflow flux through
the face at x the face at x+Ax
\ «»
#*
/
_^P M ** + A*
Figure 3.3: Control volume for the pseudo derivation of conservation of mass equation in
Cartesian coordinate reference frame.
Dividing both sides of Equation (3.15) by the control volume AxAyAz leads to
y - ((My) (f)U Z) \Z ~ (f)U z
-s. (3.16)
dt Ax Ay Az
As Ax, Ay and Az —> 0, Equation (3.16) converges to
dp d , , d . , d (3.17)
This is identical to Equation (3.10) in Derivation 1. Or, using the differential vector oper-
ator, V, we can rewrite it as
which is the same as Equation (3.8). Here, dp/dt is the partial time derivative of p at a
fixed spatial location. Alternatively, the continuity equation can be written in terms of the
substantial time derivative Dp/Dt following a fluid particle. Since
5Qdz SQdy SQdz
Dt dt dx dt dy dt dz dt
3() d() d() d()
dt dx dy y dz
(3.19)
UL
then, dp/dt — Dp/Dt - U • V/9. Substituting this into Equation (3.18), and rewriting
V • (pU) as Vp • U + pV • U, we obtain the following
(3.20)
in xyz Cartesian, r9z cylindrical, and rO(f> spherical coordinate systems, respectively.
Since p\, =
d(eSp] _ _ 9(eSp) d(eSp) d(eSp) dux duy duz
The motion of an incompressible substance only determines its stress state to within an
arbitrary isotropic constant (a scalar multiple of 6ij), and that is the negative pressure,
— P. A positive pressure generates a negative normal stress. The viscous stress arises from
the fluid motion, and is related to the deformation of the fluid by constitutive equations
as we will study in the next section. If there is no fluid motion (hydrostatic case), then
all viscous stresses TIJ are zero since there is no fluid deformation. In this case, all shear
stresses cr^ are zero, and normal stresses an are equal to —P.
AyAz
Figure 3.5: Control volume for derivation of conservation of momentum equation for fluid
flow.
Let's consider the forces, in the x direction only, acting on the control volume shown
in Figure 3.5. Once we have the force balance in the x direction, the procedure can be
repeated in the y and z directions as well. The force balance on the control volume yields
x component of x component of
Dux
FT = m total stress forces I body forces, FB due to
on all six surfaces gravity, magnetism, etc.
Du^
— <7T AyAz
:+Aa
(3.30)
as Ax, Ay and Az —> 0 in the limit. Similarly, one can repeat the force balance in the y
and z directions as well, yielding
One can rewrite the momentum equation with substantial derivative terms in expanded
form as follows
u
x ^. 1 V i ' z y (3.35)
dt ' dx dy dz dy dx dy dz
duz duz duz duz dP drxz dryz drzz
Uy Uz ^~ - (3.36)
dt ' dx dy dz d^ dx dy ^dz
We can combine Equations (3.31)-(3.33) in vectorial form. The physical meaning of each
term is explained below the equation
This is called the conservation of momentum equation, or also known as the equation of
motion. These equations along with the continuity equation describe the physical laws for
the selected system. Considering gravitational force as the only body force (i.e., F = pg),
the equation of motion can be written in Cartesian, cylindrical and spherical coordinates
as listed in Tables 3.1, 3.2 and 3.3, respectively. However, one does need a relationship
between stresses and the deformation of the fluid before one can solve for either pressure or
velocities experienced by the fluid. The equations that describe these are discussed in the
next section.
(z-dir.:) p = _^
dx
+[ f^
dx
+dy^ + dz
P9x
dvt
(y-dir.:) p
(z-dir.:) p
H,
dt dr r 89 r dz
= _i^ + fl^2( r 2 T r ( ? ) + ±:
r 89 r dr r 89
(z-dir.:} SF
r aw
^
dz [r dr r 89 dz
(r-dir.:)
v
vr , Vr. <l> dvr
I dt T ur dr ~r r rsinl? 9c/>
c9P 1 1
r r rsm rsm9 r J
9-dir.:)
e
at dr "T" r <96> rsinS
IdP I 5 1 d r^ cot
r ~^"
dd ^^~
r3dr rsmOd9 rsm 36 — P9e
Mir.:)
I dP I d I d i cot 0
rsmt rsm
VU =
.dux
11 -.
dx dx dx
du. du7
dux duz
i (3.38)
~d7 ~fc
or, using tensorial notation,
dux 9uy Quz dui du2 du3 ~
dx dx dx dxi dxi 9xi
dux QUy Quz dui 9u2 9u3
VU = 9y 9y dy dx2 9X2 9X2
(3.39)
dux 9uy du,
. dx3 dx3 dx3 .
or, as (VU)jj = duj/dxi = Ujti in short. By adding and subtracting |(VU) T , velocity
gradient can be written as
VU =
\ JVU-(VU) T |
rate of strain tensor 7 vorticity tensor 10
dua dui
dxi dx3
(3.40)
The manipulation above was done to decompose the velocity gradient VU into a symmetric
part 5(VU+ (VU) T ), and an anti-symmetric part i(VU- (VU) Thus, one can express
the strain rate tensor as
in = ( u j i + uu).
> 1J \ J-)L ''•:}/
(3.41)
\ I
The rate of strain tensor, 7 describes the rate at which the material (fluid in our case)
is changing its shape irrespective of the translational or rotational motion. If one considers
two close points P and Q within the fluid domain, then the distance between them is
ds = \/<ix • <ix, where <ix is the position vector from one point to the other. One can show
that
2
d(ds}
_
, . , , . ,
— / T x • 'V • /7"V — fiT •'"V- • HT •
,„ ^
I i 4-/ 1
t-tA. I ti-iV \JjJui I'l'] UvJ-i 1 ' V *-* . T r i j J
at
Hence, the rate of change of the squared distance can be written in terms of 7 and ofoc only.
For that reason, 7 is also called the rate of deformation tensor. It is a symmetric tensor
since 7 — VU + (VU) , and A + AT is symmetric for any tensor A. The scalar magnitude
of 7 tensor, which is independent of the coordinate system is given by
(3.43)
Components of the rate of strain tensor are listed in Tables 3.4, 3.5 and 3.6 in Cartesian,
cylindrical and spherical coordinates, respectively.
In addition to deformation, the motion of a continuum may include solid body rotation.
Vorticity tensor, uj
Uij = Ujti - mj (3.44)
describes this solid body rotation. The fluid flow is called irrotational if the vorticity tensor
is a zero tensor, i.e., all the components are zero.
dx
7yy —
dy
dvy dvx
= 7ya
dx dy
dvz dvy
dy dz
dvx dvz
7xZ ~r\ "T~ ~^
dz dx
=2—
dr
dvr
lee = <* -
r 09 dr
Jzz =
dz
d 1 dvr
r~dJ
1 dvz dvg
7«0 — -r -57T
dO + ~x~~
dz
dvr dvz
~0 1 ^
dz dr
7rr =
dr
dvr
rd6 dr
1 dv^ vr vg cot i
rsin# dd> r r
d ve\ 1 dvr
= 70r = r — —
Or r r 89
sin 9 d f v^ \ 1 dvg
rsin^ dd>
1 dvr d fvd,
h r— ——
r sin 9 dd> dr V r
For this flow, the scalar magnitude of the strain rate tensor is equal to 7 = \/^(c 2 + c2) = c-
ux(x) = cx
y
Solution
The rate of strain tensor 7 and the vorticity tensor u are calculated to be
2c 0 0 0 0 0
7= 0 -2c 0 0 0 0 (3.46)
0 0 0 0 0 0
Since cj is zero, the flow is irrotational; hence there is no vorticity and because of this, the
simple elongational flow is sometimes called pure shear flow. The strain rate is equal to
7 = /i(4c 2 + 4c2) = 2c. •
ux(x) = cx
Solution
The rate of strain tensor 7 and the vorticity tensor LJ are calculated to be
2c 0 0 0 0 0
7= 0 -c 0 0 0 0 (3.47)
0 0 -c 0 0 0
The strain rate for this case is equal to 7 = -\A(4c2 + c2 + c 2 ) = \/3c. Again this is an
elongational flow and irrotational as u\ = 0. •
(3.49)
Hence, a normal viscous stress TH and a shear viscous stress r^ are given by
(3.50)
(3.51)
Although many fluids are characterized very well as Newtonian fluids, there are many
other fluids which behave differently. The relation between the viscous stress and the rate
of strain tensors is not a linear one as in the Newtonian fluids. Some typical non-Newtonian
fluids are "shear-thinning," "shear-thickening," and "Bingham plastic" fluids. These fluids
will be discussed in detail in the next chapter.
Let's return to Newtonian fluids. If the fluid density is assumed to be constant (incom-
pressible fluid), then according to the conservation of mass equation, dp/dt + V • (pU) =
0 + /oV • U = 0, which requires that V • U = 0. Then, the V • U term in the normal stresses
drops out. So, for constant density fluids, we have
Til = (3.52)
(3.53)
(3.54)
[ dux ^
r., 1
dt
u
x ^
dux
ox
|
1 u
y ^
dux
dy
1 Uz
dux
~d7
dP \82ux , d2ux , d2ux} , r
1
a 9 1 Q 9 ' a 9
i
OX
' r
arc c*y az J ' -^Sa:
(y-dir.:)
aw y l
x y
L dt dx dy
dP
n
dy ' A1 [ -~i o
ox^ ' ov
*"!y 9 ' oz*
0 9 '
J
By
dP
dz +^ [ 9x2
z
I
%2
z
1 z
dz2 J
I T^DBz
Table 3.8: The Navier-Stokes equations for incompressible resins in cylindrical coordinates
in £, 6 and z directions.
IdP d il d 2 <9ur
--5S-
r d0 "5"
dr \r-^-
dr T
r2 "aF
39
I du
--2z , 3w MO 9uz 9w
(z-dir.:) ^, - — + U r _ _z + —+u
dt dr r 69 dz
+ Ur +
(r-dir.:) P \ gt dr ~*~ r 56 rsm6 dd> r
dP \ dd ( 1 d . 2 ,\ 1 d ( sm. g. dur\ 1 9V
r
9r r 2 ^r(
o" V~? 9r "*-) / r29smOdd
• /i "57;
V ~o7T
96 / r 2o sin
• 2/i
9
2 d 2
T -^7( (u» sm 0) -
r 2 sin 6
,. , I » e e g
dir.:) p 1 -5T+ u '-~a~ H
r aa"
ad rsmv d<p r r J
1 92ua 2 9ur
r sin2
2
J
sin (
. ,, dtp
rsind a.i' ir~?"n~
2
dr V "5~
9r / ^J
r 2 ^S
96 V ~^~S
sin 6* ^7i ("*
Solution
As seen in Equation (3.34), the x-component of force due to viscous stresses is drxx/dx +
dTyX/dy + drzx/dz, and for constant density fluids, TXX = 2/j,dux/dx as derived from Equa-
tion (3.50). The x derivative on TXX results in drxx/dx = -^(2fj,dux/dx) = 2/j,d'2ux/dx^.
2
However, we have the resultant force as ^[d ux/dx^ + (etc.)] as seen in Equation (3.55). To
eliminate the coefficient 2 in the momentum equation, one needs to look at the entire terms
as
drxx ^ dryx . drzx _ ( d \ dux~l d \duy «9ux] d \duz
^~ ~^ r ~^ + TT"
dx dy dz ^ \ dx [ dx \ dy [ dx dy \ dz ["dx dz
jp
dy2 dxdy dxdz dz2
(3-59)
The three terms drop out since their summation is zero because Qjfg- + -^ + ^jf = 0 for
incompressible (constant density) fluids. Similar derivations can be repeated in the y and
z directions to show that V • r = /nV 2 U for incompressible Newtonian fluids. •
y. Fixed plate
Dynamic b.c.:
Kinematic b.c.:
«, =0, Kinematic b.c.:
ua=QRn K, =0
Figure 3.9: Four example situations for different boundary conditions, (a) Kinematic bound-
ary conditions at the upper and lower boundaries for fluid flow within a rectangular channel,
(b) The same as in (a), except that the pulling force F on the plate is specified instead
of the speed. So, this boundary condition on the upper boundary of (b) is dynamic, (c)
Kinematic boundary conditions on the outer and inner cylinder boundaries of fluid flow
within a concentric domain, (d) The same as in (c), except that the torque T applied to
the outer cylinder is specified instead of the angular speed. So, this boundary condition on
the outer boundary of (d) is dynamic.
The no-slip condition can be directly applied in (a). If the plates are impermeable (non-
porous), then there is no fluid flow in the normal directions on the plates, which can be
written as U • n = uy = 0 at y = 0 and y = h, where U is the fluid velocity vector, and n
Pulling speed
; Moving plate
Fluid B
Kinematic b.c,: u^ — 0,
/////////^^^^//////
«,*=o
//, Fixed plate
Below we will solve several fluid flow problems by following this solution procedure assuming
Newtonian constitutive law, and then in the following chapters, the procedure will be applied
to the solution of different composite manufacturing processes.
Solution
Step 1: Cartesian coordinate system is chosen and attached at a convenient location as
shown in Figure 3.9(a). The parameters are the speed of plate V, height of the channel h,
and viscosity of the fluid n.
Step 2: As we stated in the previous section, we assume that the problem is two-dimensional
in the xy plane, i.e., the z component of velocity uz is zero, and the variables do not vary
in the z direction (d(}/dz = 0). We also assume that the fluid is Newtonian.
Step 3: The governing differential equations are the conservation of mass and the conser-
vation of momentum equations, Equations (3.10) and (3.55). We will consider steady-state
flow (d(}/dt = 0 for any variable). Note that steady state doesn't mean D()/Dt — 0 for a
Step 4: The boundary conditions have already been shown in Figure 3.9(a) as ux = V,
uy = 0 at y = h, and ux = 0, uy — 0 at y = 0.
Step 5: For a steady state with no-sink effects, the conservation equation can be simplified
to
dx dy dz
where uz — 0 due to 2-D assumption, and dux/dx = 0 due to the fully developed velocity
assumption. Then, Equation (3.60) reduces to duy/dy = 0. Integration of duy/dy = 0
yields that uy = uy(x), i.e., uy is not a function of y, but can be a function of x. The
boundary conditions show that uy — 0 for all x values, and hence uy must be zero within
the entire domain as well since it is not a function of y either. The only non-zero velocity
component is then ux which is a function of y only. How do we find ux(y)l We use the
momentum equation in the x direction:
(dux dux dux dux\ dP \drxx dryx drzx\
P -^r + ux — h uyy — I- uz —— = - — + — h -r h -^— + rBx (3.61)
\ dt dx dy dz J dx [ dx dy dz J
If the fluid is a Newtonian fluid, needed components of the stress tensor are equal to
dUx
T
xx = o^W = n0 /-) co \
(3.6za)
dx
T
idu x duy\ dux
yx = M -«- + ^ = ^— (3.62b)
V dy dx J dy
rzx = ^(^ + ~} = 0 . (3.62c)
\ dz dx J
Hence,
d ,„, d (dun
^ ' dy \ dy
(3.63)
uy-
After eliminating the terms which are zero and considering that F# — — pgk, hence FBX = 0,
Equation (3.61) reduces to
The momentum equations in the y and z direction reveal that ^ and ^ are equal to zero
if one ignores gravitational effects.
As the fluid pressure is the same at the inlet and exit and there is no pressure build-up
inside the geometry, dP/dx = 0 as well, which reduces Equation (3.64) to
V
(3.66)
Figure 3.11: Velocity profile in a rectangular channel induced by a moving plate at constant
velocity V.
Step 6: The viscous shear stress Txy on fluid at any location is rxy = /j,[(dux/dy) +
(duy/dx)} = fJ,(V/h + 0) = [j,V/h which is a constant. The flow rate across a vertical
cross section of the channel is Q — /0 aux(y)dy — J0 a(Vy/h)dy — aVh/2 where a is the
dimension of the channel in the z direction. The flow rate is linearly proportional to V and
h. The shear stress rxy, which is equal to the required pulling force per unit plate area, is
linearly proportional to V, but inversely proportional to h. •
Solution
Step 1: Cartesian coordinate system is chosen as shown in Figure 3.12. The parameters
are V, h, L, /j,, Pi, Pe.
Step 3: The governing differential equations are Equations (3.10), (3.55), and (3.56) with
\Kinematic b.c.: ux = V ,
Fluid
Dynamic b.c.:/ Dynamic b.c.:
Kinematic b.c.: ux — 0,
Step 5: The solution is exactly the same as in Step 5 of Example 3.1 up to the momentum
equation in the x direction:
dP_
(3.67)
dx
In this case, dP/dx ^ 0 here as specified by the boundary conditions. The momentum
equation in the y direction, Equation (3.56) yields
0 = - (3.68)
dy
neglecting the body force. Thus P is a function of x only. ux was found to be a function
of y only from the continuity equation (see Example 3.1). The left hand side of Equation
(3.67) is a function of x only, and the right hand side is a function of y only. In order to hold
this equality for all x and y within the fluid domain, both sides must be constants. Hence,
dP/dx = c, which means that the pressure changes linearly in the x direction. Using the
dynamic boundary conditions at both ends of the channel, dP/dx = (Pe — Pi)/L. Equation
(3.67) reduces to
pe _ pi
2 (3.69)
dy p,L
Integrating Equation (3.69) twice with respect to y yields ux(y) = [(Pe —
ciy + c?. Boundary condition n x (0) = 0 yields c% — 0. c\ is found by using the last boundary
condition, ux(h) = V. Hence,
/ \
= G, i, / 9
" \ '
(3.70)
uCouette(y}+ Poiseuille (
ifPe<P0
Step 6: What is the flow rate in this flow? Flow rate can be calculated as Q = /0 ux(y) a dy,
where a is the width of the channel in z direction. For this flow, it will be
rh n V 1 , (Pi-Pe)ah3 ahV
(3.71)
Q = Jo I\- 2fj,L y
l^ \ ady
- 12ML +
^T'
Fluid
Pulling speed
Moving cylinder
Kinematic b.c.:'
u = 0, UQ = 0, M = 0 Kinematic b.c.:
ur = 0, UQ = 0, u z = V
Figure 3.14: Incompressible fluid flow between two concentric cylinders induced by the
motion of inner cylinder in the z direction [62].
Solution
Step 1: Cylindrical coordinate system is chosen as shown in Figure 3.14. The relevant
parameters are Ri, R0, V, and \JL.
Step 2: Steady state (d()/dt = 0), fully developed (no end effects, d\J/dz = 0), and
incompressible flow (Dp/Dt — 0, hence V • U = 0) is considered in this problem. Due
to symmetry, UQ = 0 and d()/d9 = 0. Also, no body force (including gravity) will be
considered here.
dur dur
I
Ug dur
1. 1_
Ua
V, I
~^ I ^v/i
~~dt or r ac/ r r dz
I d . d
- -^7.(rr8r) + ^-(
7T
dr r dff dz
and the conservation of momentum equations in the z direction:
duz duz UQ duz duz
~^r
ot
P\u HZ"
r 06 ~7S~
az
ur = 0, uz = V &t r = Ri (3.75)
ur = 0, uz = Q at r = fl0. (3.76)
Step 5: After eliminating the terms that are equal to zero, Equation (3.72) reduces to
A ( r « r ) = 0. .(3-77)
as rur is not a function of r, but could be a function of z. But, as flow is fully developed
flow (so dur/dz — 0), f ( z ) has to be equal to a constant; hence,
Applying boundary condition either (3.75) or (3.76), c\ = 0, hence ur = 0. The only non-
zero velocity component is then uz. For an incompressible Newtonian fluid, one can now
find the components of the stress tensor as follows:
f/1 1
Trr = 2n-—^ = 0 (3.80a)
or
O
<3-80b>
= 0 (3.80c)
d (u \ I du ]
r — —e + - —r-r =0 (3.80d)
or \ r J r oU \
dus 1 du
or a
The only non-zero stress component is rrz, and its derivative with respect to z is also zero.
Hence, one can simplify the momentum Equations 3.73 and 3.74 to
(3.82)
duz
— =0. (3.83)
or
duz
^~ (3.84)
or
and further integration results in
(3.85)
Boundary conditions (3.75) and (3.76) can be applied which allows us to find c\ and c^'.
ln(r/R0)
(3.1
Pulling speed
Moving cylinder
Fluid
y///////////////////
• Stationary cylinder^
Figure 3.15: Velocity profile due to inner cylinder moving with speed V in the z direction.
Q = I °uz(r)2Trrdr
JRi
27TV n (3.89)
l
°
dr [ In
(3.90)
The external force per unit z length, Fi, required to pull the cylinder to maintain its speed
at velocity V is calculated as
= AT,.
r=Ri
In (Ri/Ro) r=Ri
(3.91)
In
Note that F0 = (2vrRO)TZT\T=R O = 2Tr/j,VR%/ln (Ri/R0) acts on the stationary outer cylinder,
which is greater than Fj. •
U P (3.92)
Dt
Since (D/Dt)(A B) =B (DA/Dt) + A • (DB/Dt) for any two vectors A and B, one can
write (D/Dt)(U-U) = 2U-(DV/Dt). Hence, the first term in Equation (3.92) can be
rewritten as U • \p(DV/Dt)] = pD(±U -U)/Dt = pD(±\U\2)/Dt where i U - U = ^ U 2 is
the kinetic energy of fluid particles per unit mass. Hence, Equation (3.92) can be recast as
D /I
P U - -U-(VP) (V-r) (3.93)
Dt \2
( Rate of energy
. &y
increase due to
body forces FB J
\
+
\
/ „
^
,,
..
generation
'
\
I Rate of energy \
/
,_ ...
(3.94)
^gy \
Rate of energy ,
increase due . to ] = / n • (a • U)
U) dA (3.98)
JS
total stress a J
where a is the total stress tensor. The energy contribution due to the body force term will
be
Rate of energy
increase due to | = / U • FB dF (3.99)
body forces F^
Rate of energy
= / RdV (3.100)
generation Jv
where R is the rate of energy generation per unit volume due to internal energy sources,
(such as chemical and nuclear reactions). Equation (3.94) can be assembled in mathematical
form as follows:
= -f - fn-
Js Js
n-(a-\J)dA+ / U-FBdV + / RdV. (3.101)
s Jv Jv
Applying Gauss' divergence theorem (Equation (3.5)), the first three surface integrals on
the right hand side of Equation (3.101) are converted to volume integrals, and Equation
(3.101) becomes
+V • q + V • (<r • U) + U • FB + R\dV
(3.102)
Since the control volume V is arbitrary, not only the integral but actually the integrand
must be zero:
1
I (~) I / \ 0
). (3.103)
r V-(vU)~l3-FB-R = 0. (3.105)
Dt V 2' ') \ 2
Equation (3.105) is expanded by writing the total stress tensor as a = —PI + r
(3.106)
V • (r • U) that appears in Equation (3.106) can be expanded as T : (VU) + U • (V • T). The
double dot product between two tensors, say A and B, is denned as A : B = AijBji =
AnBu + AI^BII + Ais-Bai + 421-612 + ^22-622 + ^23-B32 + A3iB13 + A32B23 + ^33-633, which
is a scalar. Equation (3.106) then reduces to
-r : (VU) - U - (V • r) - U • FB - R = 0. (3.107)
The second term on the left-hand side of Equation (3.107) is rewritten by substituting the
expansion from Equation (3.93). Thus, Equation (3.107) expands to the form
DE_ U F 1
Dt
+ V • q + V • (PU) - T : (VU) - U • (V • r)
-U-F j B --R = 0. (3.108)
Using the fact that U • (VP) - V • (PU) = -PV • U, and simplifying Equation (3.108)
results in
DE
= - y - q - P V - U + r : (VU) + s (E +-\U\2} + R. (3.109)
L)t \ 2
This is the conservation of energy equation, but it is not very useful, since we cannot
evaluate the first term, DE/Dt directly.
The internal energy E will be written by using thermodynamic relations, and then
DE/Dt will be substituted into Equation (3.109) to obtain the final form of the conservation
of energy equation.
One can use the thermodynamics relation H = E + P/p in order to write the internal
energy E in terms of enthalpy H. Then DE/Dt is equal to
— - —
Dt ~ Tn
in the derivation of energy equation. Here v = l/p is the volume of fluid per unit mass,
and S is the entropy of fluid per unit mass.
Solution
As E is a function of v and T, dE can also be written as
dE dE
dE = dv dT. (3.119)
dv T
dT
DE dE Dv dE DT
+
~Dt dv T
~Dt ~dT V 7)t
( dS \ Dv dE DT
— 1 p _j_ jP
dv J 1 Dt _j_ (3.120)
V T ^ dT ~Dt'
Dv
Dt Dt
I Dp
(3.121)
where Dp/Dt was replaced by — pV • U — s using the conservation of mass equation. For
brevity, no sink case (s = 0) will be considered from now on, and the substitution of
Equation (3.121) into Equation (3.120) yields
v U-
OE DT
Dt T
^~dT V ~Dt
DT
V UH v (3.122)
P 9T n
~° Dt
DT
- V-\J + cv—-| = - V - q - P V - U + r : (VU) + P. (3.123)
LJ t
—PV • U on the left-hand side cancels —PV • U on the right-hand side, and expanding the
substantial derivative as DT/Dt = dT/dt + U • VT, Equation (3.123) reduces to
dT OP
— + pcv\J.VT = - V - q + r : ( V U ) - T — V-U+R. (3.124)
This is the conservation of energy equation, but it is seldom used due to the need for cal-
culation of the underlined term. For incompressible fluids, we do recover Equation (3.117).
For an isotropic thermal conductivity which is constant within the entire domain, k • VT
reduces to fcVT where k is the scalar thermal conductivity. Hence, the energy equation
further simplifies to
dT
pcp— + pcpU-VT = fcV2T + T : (VU) + R (3.127)
for materials with constant (uniform) isotropic thermal conductivity. However, composites
may have nonisotropic thermal conductivity especially when carbon fabric is used, in which
dT
V - ( k - V T ) = — (fc n —
d dT . dT , <9T
hK + k23 -Z-
9X2 8x2
d dT dT dT
-— (3.128)
0x3
Consider a composite of thickness h having thermal conductivities of k(, k'2 and k'3 along the
principal directions of the fibers and perpendicular to the fiber direction. This composite has
fibers at an angle 9 with respect to the large faces of the composite as shown in Figure 3.16.
The composite is held between two highly conductive materials maintained at temperatures
TU and TL as shown in Figure 3.16.
Figure 3.16: A composite section being heated by two isothermal surfaces at uniform tem-
peratures TU and TL. The fiber direction in the composite is at angle 0 with the vertical
direction x\ as shown.
How will you calculate the heat flux flowing normal to the large faces of the plate? Will
the heat flux vector be parallel to the temperature gradient VT as is the case for isotropic
materials?
- k — k — k — (3129)
dxi dx2 dx3
=
dT dT dT
— <?2 ™21 r K22 r K23 lo.loU)
5xi dx2 dx3
9T
— k —
dxi dx2 °~
Now as j^ = 0 and ^ = 0, Equations (3.129)-(3.131) reduce to
(3.132)
dT
= ^- (3.133)
dT
= k31— (3.134)
ox i
where qi, q%, and 53 are the components of the heat flux vector q = qii + q%j + gsk. Thus
q\ will be heat flux in the i direction. Also,
dT TU-TL
-- (3.135)
T = Tw\ is prescribed
Solution
Steps 1-5: Although the cylindrical coordinate system seems to be the logical choice here,
the Cartesian coordinate system is preferred because of its simplicity. The justification for
this choice is that since (R0 — Ri) <C Ri, the velocity component in the radial direction ur
is expected to be much smaller than the one in the tangential 0 direction ug. Hence the
flow is assumed to be only in the 0 direction. With a coordinate transformation such as
Moving plate
u =
Fluid
uy=Q
Figure 3.18: (a) Geometry for two concentric cylinders and boundary conditions. The outer
cylinder rotates while the inner one is stationary, (b) The geometry can be converted into
one-dimensional planar Couette flow if RQ — Ri <C RI-
Here -^- — 0 considering steady state, and R — 0 since there is no heat generation within
the resin. The only non-zero velocity component is ux as found above. Considering that
the resin should have the same properties at every 9 (hence x} value, the temperature T is
expected to be a function of y only. Hence, after dropping the zero terms, Equation (3.138)
reduces to
dT dT <9T
pcp — — = k + r : (VU)
dy dz dx2 ' dy* dz2
f dT d2T
pcp \ux(0) + (0) — + r : (VU)
The only non-zero terms in the viscous stress tensor are rxy and Tyx:
d d
^ = ryx = J^ + ^}=J^^
\ dy dx J + ^l}=^-.
\R0-Ri dxj R0-Ri (3.140)
M
^k) (0)
R0to
rio — rii J \ ILO — Jn
The constant A is the energy generated within the fluid due to the viscous dissipation.
By substituting T : (VU) from Equation (3.141) and replacing d2T/dy2 with d2T/dy2
as T is only a function of y, Equation (3.139) reduces to
f - £ -•
By using the nondimensionalized variables, T = T/(T0 — Ti) = T/AT and y = y/(R0 — Ri) =
y/h, Equation (3.142) takes the following form:
/AT\2
B
- (lF ) - D (3 143)
'
where D = f (:§^)2 = f(M') 2 - Integrating Equation (3.143) with respect to y twice yields
where the constants c\ and c% are to be determined by using the thermal boundary condi-
tions. Boundary condition T(0) = T;/AT results in c^ = Tj/AT, and the other boundary
condition T(l) = T0/AT results in c\ = 1 — D. Hence, the resin temperature distribution
is given by
(3.145)
For T0 > TJ case, a typical nondimensional temperature profile is shown in Figure 3.19
underneath the velocity profile. The nonlinearity is due to heat generation induced by
viscous dissipation. Step 6: Conductive heat flux in the x and z directions is zero since
the temperature doesn't change in those directions. The flux q (which is dimensional here)
through the top and bottom plates (where n — y and n — —y, respectively) are
OT _ dT _ fcAr dT
dn dy h dy
•> = yl(R0-Ri)
x
y- Fixed plate = To-Tt
- Ti
Iv +
AT
D = ^-
fcAT D)
(3.146)
h
-k—-k—- kAT 9T
y=0 dn dy h dy
kAT fcAT(l - D)
[20(0) + 1-D] = (3.147)
h
Note that, q
y=0 y=l
Although there is no energy source inside the fluid, it was assumed that the velocity and
temperature profiles do not change along the x direction, and the steady state was assumed,
the conductive heat flux at the two walls is not equal. This is because mechanical energy
is converted into heat due to viscous dissipation within the fluid. •
3.7 Exercises
3.7.1 Questions
1. Why does one formulate a process model?
2. What are the main ingredients needed to create a mathematical model?
3. The conservation of momentum equation can be succintly written as
ITU
P (3.148)
~Dt ~
3.7.2 Problems
1. The conservation of mass equation is
?T + V • (pU) + s = 0 (3.149)
or, alternatively
— + p V - U + s = 0. (3.150)
What is the difference between dp/dt and Dp/Dt? The second term in the first
equation is V • (pU) whereas it is pV • U in the second equation. Are they equivalent?
Explain.
2. For a two-dimensional, constant density and steady state fluid flow, the velocity com-
ponents are given as ux — ax and uy = by. What is the relation between a and 6?
Can a = 26? Explain.
3. Construct numerical examples of U in which 7 and ui (i) are both zero, (ii) are both
non-zero, and (iii) one of them is zero and the other one non-zero.
4. In a plane Couette flow configuration, the fluid is bounded by two parallel plates with
a depth of h. For a Newtonian fluid with viscosity p., what are the velocity, viscous
stress tensor, and vortex tensor if the upper plate is pulled with a constant speed of V
in the positive x direction, and the lower plate is pulled with velocity V in the negative
x direction? If both plates are pulled at the same speed in the same direction, what
is the fluid velocity profile? What is the viscous stress tensor, TJJ in this case?
5. What is the required force per unit plate area in order to pull a plate in plane Couette
flow? How does the force change with the viscosity and the density of the fluid?
6. In a combined Couette-Poiseuille flow (Example 3.6), if the flow rate is Q = 0, at
which y location does ux(y] = 0? Explain the importance of your result. In this case
(Q = 0), does it mean that there is no fluid flow? Explain.
4.1 Introduction
In the previous chapter, we familiarized ourselves with the physical laws needed to cre-
ate a mathematical model for any manufacturing process and tailored it for composites
processing. The important pieces needed to construct the model are
• identification of the system along with its boundaries,
• the governing equations for conservation of mass, momentum and energy,
• constitutive equations to describe the materials and their phenomenological behavior,
and
• boundary conditions to tailor the model to a specific composite manufacturing process.
We need constitutive laws because we cannot completely describe, from first principles,
some of the transport phenomena such as the nonlinear material behavior of resin, the resin
and fiber interactions and resin cure kinetics at the macro scale.
Constitutive equations are empirical relations between parameters of interest. They
endeavor to incorporate the physics observed from the experiments or studied and analyzed
at the micron and molecular level into the equations at the macro scale. They have to
be objective such that the relationship and the results do not change with the coordinate
frame. Almost all constitutive equations require the researchers to characterize constants
needed in the equation that are specific to the material and its state. For example, when
a fluid is subjected to a stress, if the stress is always directly proportional to the strain
rate the fluid undergoes, the constant of proportionality (which can be characterized by
experiments) is called viscosity. This constitutive relation is known as Newtonian law and
the fluids that exhibit this behavior are called Newtonian fluids. Constitutive equations
should not exhibit any singularity or instability that is an artifact in the processing regime
to be consistent with the physics.
Constitutive equations are necessary to describe the processing of composite materials
due to their heterogeneous nature, complex chemistry of the resin, its interaction with
fibers and fillers and simultaneous transport of mass, momentum and energy at the micro,
meso and macro levels. Also, issues that are analyzed at the micro scale, such as growth
Continuity Equation
Momentum Transfer
Figure 4.1: Flowchart illustrating the dependence of viscosity on temperature, cure, shear
rate, particle volume fraction and orientation. Velocity and pressure are dependent on
viscosity through the mass and momentum conservation.
Resin velocity and pressure are calculated by solving the equation of motion which
contains the viscosity. Hence the velocity and pressure are functions of viscosity.
Most thermoset resins can be treated as Newtonian, although this may not be strictly
true under high shear rates. However, their viscosity is affected by temperature and cure
kinetics. Thermoset molecules will cross-link and undergo exothermic reactions when initi-
Figure 4.2: Applied shear causing the molecules to align along the flow direction.
Thus, the resistance to flow can dramatically reduce due to the change in the struc-
ture under applied shear load reducing the viscosity by orders of magnitude. Such liquids
are known as shear thinning fluids, where the viscosity decreases with shear rate. Thus
thermoplastics belong to a class of non-Newtonian fluids. Thermoplastics may also display
viscoelastic behavior due to the long chained spring like molecules; however, in composites
processing, due to the presence of the fibers and low flow rates, one can ignore the viscoelas-
tic effects [65, 66]. Usually, one checks the Deborah number (De) to gauge the importance
of viscoelasticity. De is the ratio of elastic effects to viscous effects and if De <C 1, one can
safely ignore the elastic effects of the thermoplastic polymers. This is usually true for almost
all composites manufacturing processes. As there is no cross-linking or reactions that the
thermoplastics undergo in their melt state, cure kinetics characterization for thermoplastics
is not necessary.
The viscosity of the resin also changes with temperature for both thermoplastics and
thermosets. This effect is more significant for thermoplastics. The viscosity will usually
reduce by one to two orders of magnitude as the temperature of the material increases from
the melt temperature to a few degrees higher than the melt temperature. This is primarily
due to reduction in the resistance to flow as the molecules at high temperature can move
much more freely. Molecular dynamic simulations can be performed to monitor the motion
of each and every molecule under applied shear and temperature [67]. These simulations
confirm the macroscopically observed behavior. This information is not very useful from the
processing viewpoint, but can help in development of constitutive equations for temperature
dependence of viscosity.
Spriggs
10 100 1000
Shear Rate (1/s)
Figure 4.3: Graphical depiction of viscosity changes with the shear rate on a log-log scale
as described by various viscosity models.
Figure 4.3 graphically depicts how the viscosity changes with the shear rate (7). The
units of shear rate are 1/s. Here, 7 must be independent of the coordinate system to
maintain objectivity, chosen as the magnitude of the strain rate tensor, and given by
7 -r
9 t-~i (4.6)
Experiments are conducted in a cone and plate device or a capillary rheometer to mea-
sure the viscosity under a prescribed shear rate. Cone and plate, as shown in Figure 4.4, is
usually a suitable device for viscosity measurement at low shear rates. Capillary rheometer
as shown in Figure 4.5 is suitable to measure viscosity under high shear rates.
For Newtonian fluids, the value of the viscosity should be the same at different shear
rates. However, for thermoplastic resins, this is not true. At different shear rates, one
measures different viscosity values. To measure the change in the viscosity as a function of
the shear rate, one has to subject the material to different shear rates. Almost all of the
thermoplastics display a Newtonian plateau at low shear rates, then a logarithmic decrease
with increase in shear rate and again a plateau at high shear rates as shown in Figure 4.6.
Figure 4.3 shows how the constitutive Equations (4.2)-(4.5) try to capture the observed
behavior. The Carreau model correctly captures the behavior of shear thinning materials.
However, one needs to determine four parameters by fitting the form of Equation (4.5) to
the experimental data. One can assume that when the shear rate is infinity the viscosity
goes to zero, i.e., 77(00) = 0, and thus reduce the Carreau model to three parameters.
The Spriggs model needs only three parameters too, but does not predict the plateau at
high shear rates and does not have a smooth transition from the Newtonian plateau to the
power-law unlike the Carreau model. The power-law model requires only two parameters
and does not predict any Newtonian plateau. Thus, at low and high shear rates, the power-
law model will give physically incorrect results, whereas the truncated Spriggs model will
be inaccurate at very high shear rates. However, the power-law model is popular for its
simplicity and captures the essential behavior of shear thinning in the intermediate shear
rate region, which is usually the case when processing such materials.
2RP
03
0)1 Thermocouple
Melt
reservoir
"Gl
05
1
LUi
2R
600 - -
400 - -
Increasing temperature, T
1
o
en
>
g 200 +
CO
10 10
Shear Rate [1/s]
Figure 4.6: Schematic of viscosity versus shear rate at various temperatures for thermoplas-
tic resins.
3. What is the force necessary between regions 1 and 2 to drag the fiber with velocity
Solution
1. Attach the radial coordinate frame such that the z direction is along the center of the
rod as shown in Figure 4.7. Postulate that ty — vg = 0 and vz = vz(r) (satisfying the
continuity equation), which implies that
r = r(r). (4.7)
To find the velocity profile, one can use the z direction of motion
1
(4.8)
r dr
as all other terms are zero, if we ignore inertia and unsteady effects due to low Reynolds
number flow.
Note that now we need to subsitute the power-law constitutive model to relate stresses
to velocity gradients
>rz (4.9)
where
(4.10)
and
(4.11)
Since
dvz
= 7zr = (4.12)
dr '
Equation (4.11) reduces to
dvz dvz
7= (4.13)
dr dr
77 = m (4.14)
dr
Substitution of Equation (4.14) into Equation (4.9) gives
(4.15)
dr
Now if one substitutes Equation (4.15) into Equation (4.8), the following differential equa-
tion for vz is obtained:
d
' ' """ l ' = 0. (4.16)
dr [ \ dr
Integrating twice and using the boundary conditions
v
z — o a-t r = i (4.17)
vz = 0 at r = R0 (4.18)
results in
(4.19)
where re = Ri/R0.
3. The force is calculated as
(4.21)
(re - Kn)R0
where r/o is the viscosity at temperature TO and c\ and 0% are constants that depend on
the material. Temperature dependence of some of the common thermoplastics is shown in
Figures 4.6 and 4.8.
Thermosets also exhibit similar temperature dependence, but their temperature behav-
ior is coupled with the cure kinetics. Usually, the viscosity for thermosets reduces with
temperature, T and increases with degree of cure, a. Both phenomena can be described
well with exponential curves. Hence, the viscosity for a thermoset can be expressed in a
generic constitutive equation as
c
i (4.23)
??(T, a) = ?)o exp —- + c3a
where ci, c% and 03 depend on the thermoset material under consideration. There are many
other ways to characterize the viscosity dependence of reacting systems.
Two concurrent phenomena govern the rheological behavior of a reacting system: one
associated with the intensification of the mobility because of the increase in temperature,
responsible for decreasing the viscosity, and another one related to the growing size of
the molecules during cure, responsible for increasing the viscosity of the resin [73]. The
empirical Williams-Landel-Ferry (WLF) equation [74], valid near the glass transition and
based on the free-volume theory, is frequently used to represent the first phenomenon:
ry
In = (4.24)
In 77 — In r/oo + In Mu (4.25)
RT0
r/(T, a) =
exp ( —- + Ka (4.26)
\K1
where U is the activation energy of the viscous fluid, R is the gas constant and K is
a constant which accounts for the effect of the chemical reaction on the change in the
reacting mass viscosity and, consequently, on the dissipation intensity. It is assumed that
U is independent of the degree of cure. The values for these parameters for two systems are
listed in Table 4.1.
77 = r, exp (4.28)
\RT
where the fluid activation energy, En, and the frequency factor, A,,, are given by:
E^ — a + ba A^ — a0exp(-b0a), (4.29)
with a, 6, a0 and b0 as specific constants for each resin. For example, consider a system
composed of a partially cured unsaturated polyester resin, OC-E701 [91], 43 wt % of styrene
and t-butyl perbenzoate as an initiator. The constants for this system are: a=7.8 kcal/mole;
6=19.7 kcal/mole; a0=6.41xlO~5 N.s/m 2 ; 60=23.1. However, Lee and Han [91] pointed out
that for these values, at temperatures higher than 156.2°C, the viscosity calculated by
Equation (4.28) decreases with an increase in a, which is not physically correct. Hence,
they proposed two other equations, valid for the system used: one based on the Tajima and
Crozier [94] model:
26.8(41.63+ r-90.19a)
log 77(T, a) = (23.64 + 3.82a) - (4.30)
55.03 + T- 90.19a
-T- (4.31)
Hence, it should be clear to the reader that development of a universal formula for viscosity
as a function of temperature and cure does not exist. However, forms suggested in Equations
(4.26)-(4.28) can accomodate many thermoset resin systems.
Figure 4.9: The deformation response of a bundle of aligned fibers lubricated with viscous
resin. For transversely isotropic bundle or laminates, the two dominant shear modes are (i)
transverse shearing, and (ii) longitudinal shearing [6].
of these laminates. Unfortunately, the shear and extensional properties of long discontinu-
ous fiber reinforced composites in their melt state are difficult to obtain using traditional
rheological techniques, such as cone and plate and capillary rheometers, due to the presence
of continuous or long fibers.
One way to describe their shear behavior is to use squeeze flow to characterize the
bulk transverse shear viscosity of these highly filled viscous resin systems. Squeeze flow is
invoked by placing the material between two impermeable platens and applying a normal
force. One may characterize the material under constant applied load or under a constant
closure rate of the platens. Such industrial testing devices are referred to as the parallel
plate plastometer or squeeze plate viscometer [96, 97]. These devices measure radial flow
between two parallel disks and have been used to study the viscosity of polymer melts and
other viscous materials.
The advantages of this type of test include mechanical simplicity, inclusion of fibers
without the danger of attrition, achievement of very high shear rates, usage at high tem-
peratures and the ease with which high-viscosity materials can be tested [98]. Additionally,
this type of flow is of interest since it is encountered within lubrication systems and the
stamping of plastic sheets. Finally, it provides a technique for investigators to evaluate
rheological equations under transient conditions.
In the parallel plate plastometer, the test material, in the form of a cylinder, is placed
between parallel, circular flat plates. With the lower plate fixed, either a constant force
or a constant closure rate is applied to the upper plate. If a constant force is applied,
measurement of the resulting displacement of the upper plate is used to determine the
rheological properties of the specimen. If a constant closure rate is applied, measurement
of the resulting force on the upper plate is used to determine the viscosity dependence
on shear rate. Although these experiments involve unsteady shear flows, the flow rates
are usually modest so that the analyses are based on quasi-steady state solutions. Two
different sample arrangements may be used in the plastometer. Either the specimen sample
cross-sectional area can be equal to the cross-sectional area of the plates, in which case the
area under compression is constant, or the cross-sectional area of plates may be larger than
the specimen sample, in which case the volume of the specimen is constant.
Squeeze flow rheometers have been used to study both unfilled viscous and viscoelastic
liquids and have been modified to study fiber filled thermoplastic materials. Experimental
Air Jets
Hot Platen
Cartridge Heaters (*6)
Clamp
/l\K M Clamp
i Beam
W Thermocouples (*6) ^l//
Hot Platen
Figure 4.10: High temperature squeeze flow set up for thermoplastic composite laminates
in a servohydraulic test frame [42].
With this system, they tested three different combinations of aligned fiber materials:
Table 4.3: Power-law Viscosity Parameters for Unfilled and Fiber Filled Clay
Fiber Volume Percentage Platen Closure Rate [cm/min] m [Pa sn] n
0 0.0254 to 0.762 30,000 0.1
20% Nylon Fibers 0.0254 82,000 0.1
40% Nylon Fibers 0.0254 110,000 0.1
50% Nylon Fibers 0.0254 300,000 0.1
60% Nylon Fibers 0.0254 810,000 0.1
60% Glass Fibers 0.0254 1,650,000 0.1
They found that the unfilled clay and the filled samples behaved as shear thinning power-
law fluids. Table 4.3 summarizes the results that were plotted over a broad range of shear
rates. It reveals the decidedly nonlinear relationship between the composite's transverse
shear viscosity and the percentages of fibers. The clay filled with 60% nylon fibers resulted
in a transverse shear viscosity 27 times greater than the viscosity of the unfilled clay, while
60 volume percentage of the glass fibers resulted in a transverse shear viscosity 55 times
greater. From this information, it seems that adding the same volume percentage of smaller
diameter fibers has the effect of increasing the transverse shear viscosity even more than
the increase found using larger diameter fibers.
The reinforcing PEEK matrix polymer in the APC-2 laminates is known to behave as
a shear thinning Carreau-type fluid. The four parameter Carreau fluid can be reduced to
three parameters by assuming that the fluid viscosity goes to zero at infinite shear rate.
Thus, the three-parameter Carreau fluid model [71, 72] describes the viscosity as Newtonian
behavior at low shear rates followed by shear thinning power-law behavior at higher shear
rates, and is commonly expressed as:
(A7) (4.33)
where 770 is the zero shear rate viscosity — the viscosity of the material when deformed
infinitely slowly, n is the shear thinning exponent that adjusts the slope of the shear thinning
region and A is a time constant that adjusts the position of the "knee" in the r]vsj curve
where there is an onset of shear thinning behavior. The Newtonian model is recovered for
n = 1, and the power-law expression is approached for large values of A.
For the PEEK matrix, these viscous parameters were measured by [115] and are listed
in Table 4.4. Viewing the transverse shear flow of the fiber filled matrix as viscous flow
Shuler and Advani [2] collected the platen force data for the four closure rates and used
least-squares fit to the Carreau model to find the parameters for the bulk transverse shear
viscosity of the APC-2 laminates. The three Carreau parameters of Table 4.5 were found to
produce loading curves that closely matched the experimental platen force data collected.
Table 4.5: Carreau Model Parameters for the Transverse Shear Viscosity of APC-2 Material
at 370°C.
APC-2 Transverse Shear Viscosity Carreau Parameters at 370°C.
•to = 2.5 * 106Pa.s
A = 50.0sec
n = 0.65
The Carreau model curve for the transverse shear viscosity of the APC-2 material at
370°C based on the parameters listed in Table 4.5 are graphed in Figure 4.11. The Carreau
model viscosity of the neat PEEK matrix polymer at 370° C is graphed in Figure 4.11 as well.
It is noteworthy that the transverse shear viscosity of the composite is approximately 3600
times greater than viscosity of the PEEK matrix resin. Also, the onset of shear thinning
behavior occurs at a lower shear rate than for the neat resin. Since the Carreau model
captures both the Newtonian behavior at low shear rates and the shear thinning behavior
at higher shear rates, this three parameter model is able to describe the viscous behavior
of the unidirectional APC-2 laminates over a range of shear rates.
It is instructional to note that during squeeze flow the material experiences a range
of shear rates that vary both over time and throughout the volume of the material. The
transverse shear rates encountered in squeeze flow range from very low at the start of each
test and at the symmetry planes to a maximum which occurs at the largest closure rates
adjacent to the platen contact surfaces near the outer edges of the samples. Because squeeze
flow encompasses such a broad range of shear rates throughout the volume of the material,
a simple Newtonian or power-law fluid model cannot adequately describe the squeeze flow
behavior of the laminate that, depending on the shear rate, follows both Newtonian and
shear thinning behavior. Hence the use of the Carreau model is more appropriate for
composite laminates.
Figure 4.11: The Carreau model curve using the parameters listed that describes the trans-
verse shear viscosity of the APC-2 material and the resin PEEK at 370°C [42].
between two disks is proposed as shown in Figure 4.12, in which one can measure the force
and the displacement. Conduct the analysis to delineate the approach.
Solution
As h <C R, one can assume that shear forces will dominate. Also for GMT materials, the
viscosity is very high such that the Reynolds number is much less than unity. Hence, one
can ignore the inertia forces and assume a quasi-steady state process. A "quasi-steady
state" as we will see in Chapter 5, implies a "steady state" at a given instant. This allows
us to solve the steady-state problem at any height h, which will then change at the next
instant and the velocity profile will change accordingly and achieve steady state instantly.
Using mass conservation for a disk of any radius r < R, one can show that
f
-'fnrr = 2vrr / vr dz (4.34)
Jo
where VT is the radial velocity and h = dh/dt is the instantaneous disk velocity. See Figure
(4.13) for the derivation of Equation (4.34).
The mass that disappears in the ^-direction must appear in the radial direction as shown
, ^ dz.
pvrr — = p I 2iTrvr(z) , (4.35)
In the equation of motion, as only the shear stress component rrz is important, the r
component of the equation, ignoring inertia and gravity as well, reduces to
drrz dJP
(4.36)
dz dr '
Figure 4.12: Schematic of GMT material between two parallel disks of radius R with cylin-
drical coordinate system placed midway between the disks.
symmetry
Figure 4.13: Sketch of velocity profile of the material between the midplane and the upper
platen in compression molding of a radial charge.
Substitution of Equation (4.39) into Equation (4.40) and integration of vr with respect to
z results in
\ i+s"
(4 41)
1 + s V m dr s) '
where s = 1/n. Note that the velocity profile is identical to fully developed flow between
parallel plates separated by distance 2h. Now, one subsitutes Equation (4.41) into Equation
(4.34), and integrates with respect to z, to obtain the following differential equation for P(r):
(4 45)
-
where
One can find A and b from the plot of In F versus In ( ^ ) as shown in Figure 4.14. From b,
ln(F)
In (A)
In (1/h)
Figure 4.14: Approach to find the power law parameters from squeeze flow compression
experiment in which the material is squeezed at the constant rate of h and the required
force, F, is measured.
Next, we will discuss how the presence of short fibers influences the viscosity and how
one can model this physical effect.
- . (4.49)
When the fibers are long and slender (r <C 1) this requires a very small volume fraction. For
example, a suspension with fibers of aspect ratio equal to 100 must have a volume fraction
less than 0.01% to be dilute. If the fiber volume fraction is increased such that
1 1
or (4.50)
then the suspension is said to be semi-dilute. In this regime the fibers seldom touch,
but fibers experience frequent fluid-mechanical interactions from the perturbation of fluid
velocity caused by the nearby fibers. Some workers call this the semi-concentrated regime.
Increasing the volume fraction still more, to
1
<c or n> (4.51)
reaches the concentrated regime. Here the average interfiber spacing is of the order of a
fiber diameter, so that each fiber is always touching, or nearly touching, many other fibers.
Figure 4.15 maps out these regimes in terms of volume fraction and aspect ratio. As
shown in the figure, all commercial composites fall into the concentrated regime. No com-
mercial materials are truly dilute. Hence we will discuss here the phenomenological equation
to describe the viscosity of concentrated suspensions.
1000
£ 100
CO
o
0)
Q. 10
.001 .01 .1
Volume Fraction, c
Figure 4.15: Different regimes of concentrations for fiber suspensions [116].
The most important action of a fiber in a suspension is to resist stretching of the fluid
along the fiber axis. All fiber suspensions have high viscosities in the fiber direction. This
Figure 4.16: Definition of the orientation of a single fiber in a Cartesian coordinate frame
[117].
where p; (i=l,3) are Cartesian components of p. Orientation space is the set of all possible
directions of p and corresponds to the surface of a unit sphere. Sometimes, the fibers
are long and the parts are thin which constrains the fibers to lie in the plane. Thus the
orientation is called two-dimensional or planar orientation. In Figure 4.16, if one selects the
1-2 plane as the planar orientation plane, then 9 will be equal to vr/2 or p% equal to zero
for every fiber. The planar orientation can be described by either (f> or a two-dimensional
vector p. The orientation space for two dimensional orientation will be a unit circle.
Figure 4.17: Sample of fibers digitized from a radiograph of a plaque of sheet molding
compound [118].
The most basic description is the probability distribution function. If we consider a small
region from a composite or a suspension, we would expect to see many fibers with different
orientations of fibers. The probability distribution will give us a measure of the state of
fiber orientation at that location. Thus one can define if>(0, (f>) such that the probability of
Orientation Parameters
The probability distribution function is a complete description of orientation state, but
it is also cumbersome. A number of orientation parameters have been defined to provide
concise and more easily interpreted measures of orientation. The best known is the Hermans
orientation parameter. It concerns the special case of axisymmetric orientation, when the
fibers are uniformly distributed in <f> so that i/j is a function of 9 only. The extent of alignment
around the 9 = 0 axis is then measured by the parameter /, defined as
o
'• 1.0
m
5
0.5
-n/2 o
1 71/2
Orientation Angle,
Figure 4.18: Histogram and probability distribution function for the fiber orientation in
Figure 4.17 [118].
Orientation Tensors
A description which combines the generality of the probability distribution function and
the concise nature of the orientation parameters is a tensor description of orientation. A
second-rank orientation tensor can be defined as
dij = < p^ > (4.59)
and a fourth-rank tensor as
a>ijkl = < PiPjPkPl > (4.60)
That is, the components of the orientation tensors are formed by taking the orientation
average of the products of components of the vector p. These tensors are a generalization
of the orientation parameters. It is evident from their definition that a^ and a^i will
transform according to the rules of tensor transformation, so they are tensors. They are
also symmetric, i.e.
Oij = a-ji (4.61)
The normalization condition, Equation (4.54), implies that the trace of d^ is unity,
Oij = I (4.62)
is: •x-/ -i
0.5 Ol F0.7 Ol _ F 1 0
a
<t~[0 0.5] "V ~ [ 0 0.3J "*' [ 0 0
Figure 4.19: Planar fiber orientation states and the corresponding tensor representations.
There are a number of physical interpretations of the orientation tensors. They can
be thought of as a generalization of the orientation parameters, as the moments of the
probability distribution function or as the coefficients of a series expansion of the distribution
function. For a complete review see [119]. For the present we note that the orientation
tensors are free from assumptions about the shape or symmetry of the distribution function,
and can be readily expressed in any convenient coordinate system.
Solution
As the fibers are uniformly distributed, an = a^ = 033 and 012 = a23 = asi = 0. From
normalization, we know that an + 022 + 033 — 1. Thus,
i 0 0
0 i 0 (4.63)
0 0 I
It can also be proved that the nth-rank orientation tensor provides sufficient information
to predict the effect of orientation on any nth-rank property of the suspension or composite.
an =
where wk is the weighting factor for the fcth fiber. If the fibers are almost circular in cross
section, then wk can be taken as 1, which reduces Equation (4.64) to
where the coefficients 777 , Np and Ns depend on rjs, c, A, B and C, and a^ and a^i are the
second and fourth ranked orientation tensors. Now 777 contains all the isotropic contributions
to viscosity (from both the solvent and the particles), while anisotropic contributions by
the particles are represented by two dimensionless parameters, Ns and Np. Np is called the
particle number and N3 the shear number [121]. These parameters depend on the particle
aspect ratio and volume fraction, but (at least for dilute suspensions) not on the orientation
state of the fibers. When both of these parameters equal zero, Equation (4.68) reduces to
an isotropic Newtonian fluid.
The particle number Np is the key rheological parameter to describe the behavior of a
fiber suspension. Physically it represents the factor by which the suspension resists elonga-
tion parallel to the fiber direction, as compared to other deformations. Evans [126] was the
first to identify this parameter and its key role in suspension rheology. For a suspension
of slender particles Np can be quite large. Figure 4.15 shows how Np varies with volume
fraction and aspect ratio according to the theory of Dinh and Armstrong, using the aligned
assumption for inter-particle spacing. The other theories give similar predictions.
H O
The reaction can be tracked by monitoring the instantaneous concentrations of A and B,
denoted by CA and CB- They can be measured in moles per gram and will decrease with
time as more A's and B's react with one another to polymerize.
C* = (4.74)
C* is zero when there has been no reaction and equals unity when all of the A's have
been consumed and the reaction is complete. C* can also be called the degree of cure or
conversion (also sometimes the symbol a is used to denote degree of cure). Substitution of
this in Equation (4.71) with some algebraic manipulation gives
dC1*
— = k0 exp (-E/Rr)CAo(l - C*)(A - C*) (4.75)
with an initial condition of C* — 0. Here A is equal to CBO/CAO, and A equals one if A and
B are present in equal number of moles.
t (sec)
Figure 4.20: The degree of cure as a function of time for the parameters chosen in Example
4.4.
One can easily plot this equation for C* with respect to time. Selection of ko = 3 and
E/R = 300 results in the plots shown in Figure 4.20. •
For a reaction in which temperature varies with time, which is the case for all composite
processing methods, one needs to solve Equation (4.75) once the temperature history T(i)
is known.
If the polymerization mechanisms and the order of the reaction are known then it may be
straightforward to determine the molecular weight and the gel point explicitly as a function
of the extent of reaction at a single temperature. However, the chemistry is usually complex
and the reactions are taking place at different temperatures as the curing process by nature
is nonisothermal. Hence the order of the reaction is not known and one cannot usually
have a single type of constitutive equation to accommodate the chemical complexity of the
reaction.
Hence, kinetics equations due to complex reactions are strictly empirical. For example,
cure of unsaturated polyesters is fairly well represented by a kinetic equation such as
dC*
= k0exp(-E/RT)(C*)m(l-C*y (4.77)
dt
Since dC*/dt = 0 for C* = 0, one assumes that the resin starts with a small degree of cure,
say C* = 0.0001. One now needs to measure ko and m and n from experiments, where
(m + n) represents the reaction order. There are many other deviations from this type of
Here f ( C * ) is the function that describes the rate of reaction in terms of primary con-
stituents and empirical parameters, K is the rate constant denned by the Arrhenius type
of relationship and expressed as
where A is the frequency factor. Different forms are proposed for f ( C * } in the literature.
We will list two of the most common ones used. The first one corresponds to an nth degree
reaction which takes the form
f(C*) = (1 - C*)n. (4.80)
And the second one is related to the autocatalytic reaction as follows:
dC*
Q = HRM— (4.82)
at
where M is the mass of the sample. Kinetics are best characterized by reacting the sample at
a constant temperature and measuring Q with time. Several runs at different temperatures
are needed to find the temperature dependence.
This is by far the most widely utilized technique to obtain the degree and reaction rate of
cure as well as the specific heat of thermosetting resins. It is based on the measurement of the
differential voltage (converted into heat flow) necessary to obtain the thermal equilibrium
between a sample (resin) and an inert reference, both placed into a calorimeter [137, 138].
As a result, a thermogram, as shown in Figure 4.21 is obtained [139]. In this curve, the
area under the whole curve represents the total heat of reaction, HR, and the shadowed
area represents the enthalpy at a specific time.
W
"o
0
I
Temperature
Figure 4.21: Example of a Thermogram — rate of enthalpy change with temperature [129].
The DSC can operate under isothermal or nonisothermal conditions [140]. In the former
mode, two different methods can be used [141]:
i) A sample is placed into a previously heated calorimeter (faster equilibrium [142]) or
into an unheated calorimeter, whose temperature is raised as quickly as possible up to the
curing temperature.
ii) A sample is cured for various times until no additional curing can be detected; then,
the samples are scanned (heating rate ranging from 2 to 20 °C/min) in order to measure
the residual enthalpy, Hres. The degree of cure is calculated directly,
(4.83)
HR
but not the reaction rate, which is obtained by tangents to the curve of a versus time. For
reactions with very small exothermal heat, this method should be utilized instead of the
first method.
In the nonisothermal or dynamic mode, a sample is set into a calorimeter and the
temperature is raised at a certain constant heating rate up to the operational temperature.
The total heat of reaction is independent of the heating rate (recommended range is 2-
20°C/min [141, 134]). Three methods are possible:
i) All kinetic information (A and E in equation (4.79) can be obtained from only one
experiment. Although this method has been successfully applied for some first order re-
actions, it is not accurate for others; the heat of reaction is higher (around 10-30%) when
compared with the isothermal mode [141].
ii) A and E in equation (4.79) can be accurately measured from the values of the
peak exotherm temperature for various heating rates for all reactions [141]. This is the
recommended method.
iii) More general than the second method, it measures A and E from the values of
the temperature necessary to reach a constant conversion for various heating rates. This
technique needs considerable effort to be moderately successful.
Although many researchers use DSC for measuring enthalpies over the entire curing
process, it is not accurate after the gel point [143, 144, 145].
Several authors have pointed out that the parameters of the reaction rate expression for
isothermal and dynamic conditions may be different [143, 146, 147, 148, 149, 150, 151].
One possible explanation for different results between isothermal and dynamic analysis
was first given by [152], and it is related to the fact that the degree of conversion is a
where a = dT/dt. For isothermal experiments a=0, and there is no difference between
equations (4.78) and (4.85). But as the dynamic measurements are simpler and faster
than isothermal ones (for the first case, only one experiment is needed to obtain all kinetic
parameters), it is important to calculate the term (da/dT)t in order to acquire a relationship
between dynamic measurements and isothermal parameters.
Other thermal techniques to characterize cure are thermogravimetric analysis (TGA)
[76, 159], high pressure calorimeter (HPC) [141], thermomechanical analysis (TMA) [141,
160] and differential (or dynamic) thermal analysis (DTA) [161]. These are rarely used and
will not be discussed here.
However, we can briefly mention another technique used which is called torsional braid
analysis (TEA). This technique is a variation of torsion pendulum (TP) and was developed
by [162] in 1958. It consists of measuring the frequency and decay constants which charac-
terize each wave resultant of free torsional oscillations (at approximately 1 Hz) subjected
to a sample (a glass braid impregnated with a resin). These measurements are converted
into elastic (stored energy) and loss (loss energy) moduli, which are related to transition
(gelation and vitrification) times and temperatures.
TEA is a sensitive technique for determination of physical changes in the resin [163, 164,
165, 166, 167, 168], occurring at and after the gel point; it has a lower limit of detectability
[141]. TEA is not capable of measuring the degree of cure.
Other mechanical techniques are dynamic mechanical analysis (DMA) [169, 170] and
rheological dynamic spectrometer (RDS) [76, 171, 172, 173].
Due to the fact that there is no accurate technique valid for the entire curing reaction,
two techniques should be used to follow cure: one applied before gelation (DSC) and another
one applied after gelation (TEA, DMA, TMA, d.c. conductivity [174, 175, 176, 177]) .
•eo
Wavenumber (cm 1 )
The basic principle of all these methods is the comparison between the spectrum of refer-
ence substances and spectra of the reactants and products of a curing reaction subjected to
radiation. A qualitative and quantitative identification of the components is then possible.
However, microscopic characterization has not yet being linked to the phenomenological
parameters used in cure-kinetic equations but helps in understanding how the components
react.
Table 4.6: Kinetic Parameter Percentage Errors Between Composite and Neat Resin
It is clear that the surface of reinforcement can affect the kinetics. The disagreement
affects the extent. The type of resin and the temperature range are important considera-
tions. Generally at low temperatures, one must be cautious about applying the neat resin
phenomenological parameters to a system in which the resin will impregnate a network of
fibers.
4.6.1 Introduction
The term crystallization has its roots in the word crystal, which refers to an ordered arrange-
ment of molecules/units, which can be categorized in a specific way. Thus crystallization
refers to the phenomenon of turning into a crystal. First, it is important to distinguish
between the phenomenon of solidification and crystallization although one uses them as
synonyms. Solidification refers to the transition of a substance from the liquid state to the
solid state; it doesn't require it to solidify into an ordered structure. On the other hand,
crystallization specifically refers to the transition of a body from the liquid state to an or-
dered solid state or a crystalline state. While solidification refers to all substances in general,
crystallization usually refers to long chain structures, such as thermoplastics and proteins
for example. "Crystal" refers to solidified metal structures too, due to their regularity. An
important point to note is that although the term crystalline polymers can be theoretically
defined, it is not physically realizable, since there is always a separate noncrystalline phase.
Hence thermoplastics can be in either an amorphous state (no crystals) or a semicrystalline
state (partial crystallinity). Crystallization evolves either from a solution or from the melt.
These types of crystallization entail different concepts. We restrict ourselves here to only
crystallization from the melt as thermoplastic composites involve processing of melts.
When we consider the crystallization of thermoplastics, the complex nature of their
structure is influenced by the processing conditions. Depending on how the thermoplastic
is cooled, it will either crystallize (slow cooling) or solidify into a glassy state (fast cooling);
unlike a pure metal melt, which will always crystallize. This occurs due to the structure
of polymers, i.e., their chain-like molecules. The ordering of these chains does not occur as
easily as does the ordering of the atoms of a pure metal melt, where they organize them-
selves into a regular lattice (ordered structure/unit) when the thermodynamic conditions
are conducive. Instead, various effects come into play at the micro and the macro level of
the crystallizing molecules, which makes crystallization in the case of polymers not a truly
thermodynamic transition.
>,
Q.
LLJ
Temperature Temperature
o O
4-J~
0!
CD
I
O
tf=
O
CD
Temperature Temperature
fa) (b)
Figure 4.23: Variation of thermodynamic variables with temperature, (a) refers to a semi-
crystalline polymer and (b) refers to an amorphous polymer [200].
4.6.3 Background
The knowledge that a thermoplastic could crystallize was not realized until 1912, when
Von Laue showed that, because of the repetitive character of their lattices, crystals would
scatter a beam of x-rays into regular patterns which were analogous to optical diffraction
patterns [202]. Thus, subsequent work on polymeric materials demonstrated that they
exhibit crystallinity at the level of molecular packing. Thus the next step was to deter-
mine the development and nature of this crystallinity, which to this date is not completely
understood.
The chemical properties of a thermoplastic composite product are dictated by its matrix
The crystallization from the melt is also referred to as bulk crystallization, as the move-
ment of chains is hindered by other chains in their path and also by crystallized areas in the
bulk. Crystallization in the melt or bulk occurs by the formation of crystals in the radial
direction from a central nucleus, also known as spherulites. A schematic of a spherulite is
shown in Figure 4.25.
The formation of spherulites is described by the phenomenological theory first proposed
by [205]. They theorized that in order for growth of a spherulite to take place, two conditions
must be met: the medium must have a high viscosity, and noncrystallizable material (or
solute) must be present. Similar to solidification of metals, the solute is rejected from the
crystallizing solid and pushed ahead of the solidifying interface, and is assumed to promote
the instability of the solid/liquid interface. They suggested that branches of the spherulite
penetrate into the melt and grow preferentially in regions rich in crystallizable material.
Initially, the nucleation process must occur in order for the polymer chains to attach
themselves to the nuclei and facilitate growth of the crystal. Then nucleation and growth
occur together for a certain amount of time, which is followed by a free growth phase of
the spherulites. Free growth continues until the spherulites impinge on each other, and
the process of secondary crystallization takes place. Secondary crystallization refers to
the attachment of chains on a smooth surface of the crystal, where previously no growth
occurred. It is much slower compared to primary crystallization because the polymer chains
are not able to find nuclei to attach themselves to, and they are also hindered from diffusing
to other favorable growth sites. Finally, after crystallization is complete, the body cools
down to the ambient temperature by convection [206].
One can address the crystallization kinetics on two scales that are coupled. At the
microscopic scale, one needs to model the nucleation and growth of a spherulite and on a
macroscopic scale one needs to determine the overall crystallinity behavior of the polymer.
It is obvious that changes at the microscopic level will influence the macroscopic constitutive
equation.
Spherulitic growth is still not fully understood, but it has been shown that their radii often
grow linearly with time under isothermal conditions, which supports a kinetics controlled
growth theory [206, 207, 208].
The growth of the tips of the branches of a spherulite follows the kinetics of attachment
of polymer chains given by the Lauritzen and Hoffman theory [209, 210]. Thus, the growth
I
= nK(T)(l-X)ln (4.88)
dt
The models for nonisothermal crystallization more accurately describe experimental data
collected on the crystallization of various thermoplastics.
The constants don't have a physical meaning and cannot be easily connected to the
microscopic analysis. However, the constants can be easily determined through differential
scanning calorimetric (DSC) experiments, and no induction time (the time required for the
organization of molecules around a nucleus) factor is necessary. DSC is a technique used
to determine thermal transition points of various substances. The evolution of the absolute
crystallinity profile within a part during cooling can be predicted by coupling the model
with a heat transfer analysis [203]. This is done by representing the source term in the
energy balance equation by pcpHrdx/dt which tracks the heat evolved during solidification.
4.7 Permeability
Permeability characterizes the ease with which a fluid can flow through a porous medium.
In almost all composite processes that use a thermoset resin as the matrix material, the
resin has to be either infused into a fiber preform in processes such as liquid molding or
extracted out during consolidation as in autoclave processing. This motion of the resin is
usually modeled as flow through a porous medium in which the fiber network constitutes
the porous medium. The empirical Darcy's law [213] governs the flow of the resin as follows:
Q = ^ . (1.90)
Here Q is the flow rate across the cross section A, T] is the viscosity of the liquid, dP/dx is
the driving pressure gradient and K is the permeability of the porous medium.
This law allows the modelers to describe a macroscopic relationship between the flow
rate and the pressure gradient needed to drive the flow instead of using the momentum
conservation equations. If one wanted to use the momentum equations, one would have
to describe the geometry of every channel present in the fibrous network and solve the
momentum equations inside them to find the flow rate and pressure inside these channels.
There can be a network of over a million channels created by the arrangement of the fibers.
This task would be impossible to approach even with a sophisticated computer, and more
importantly we only need to know the overall relationship between the pressure drop and
the flow rate and not the details of the pressure at each and every location within every
channel while processing such materials. Thus, Darcy's law lumps the ease of flow within
these channels into a parameter called permeability which characterizes the mobility of the
resin through the fibrous porous media.
Permeability is a function of fiber network, fiber volume fraction and also the type of
sizing applied to the fibers. Also, as the porous medium formed by the fibrous network
will be anisotropic (it will offer different resistances in different directions) , the practice has
converged on generalization of Darcy's law to account for it:
u=-—-VP (4.91)
77
where u is the volume averaged "Darcy velocity" (see Figure 4.26), r\ is the viscosity of the
fluid, VP is the pressure gradient, and K is the permeability tensor of the preform. Here,
K is now a tensor instead of a scalar and can be represented as
and the three-dimensional form of Darcy's law is expressed in matrix form as follows:
The Darcy's law that now substitutes the momentum equations can be written down in
three directions as follows (by using Cartesian coordinates):
K^d_P_ _ K^d^ KxzdP
(4.94)
T) dx 77 dy r) dz '
KyxdP _ KyydP _ KyzdP
(4.95)
TI dx TI dy rj dz '
Kzx dP Kzy dP Kzz dP
(4.96)
T] dx rj dy TI dz'
Ku 0 0 \ / dP/dxi \
0 #22 0 dP/dx-2 . (4.97)
0 0 K33 ) \ dP/dx3 )
Here, the local principal coordinate system X i , x % , x 3 is along the principal axes of the
preform which coincides with the major and minor axes of the ellipsoid formed by the resin
volume injected as shown in Figure 4.27. Thus, by choosing the coordinate directions along
Figure 4.27: Local in-plane principal axes in the plane of a preform along the major and
minor axes of the ellipsoid formed by the injected resin volume.
the principal axes of the fiber preform, one can measure the principal permeability values of
the preform. Then by coordinate transformation, one can calculate the non-diagonal values
if one knows the flow direction with respect to the principal permeability direction.
Figure 4.28: Coordinate axes (x[, £ 2 ) and principal axes (x\, x2) of the preform. One
dimensional resin flow from left to right.
cavity wall. Resin fills this section first due to negligible resistance to the flow. The
significant flow through the preform starts after this air channel is filled. Note that no
significant racetracking channel should exist between the preform and horizontal mold walls
(see Figures 8.22, 8.23, and 8.24 for definition of racetracking). Consider the two coordinate
systems: (i) the principal axes x'y', and (ii) mold axes xy. The two-dimensional (in-plane)
permeability tensor is given by
Kn 0
(4.98)
0 K22
if x\ is along x\ and x'2 is along x2 . However, if x'x is at an angle 6 to the principal x\
axis, then the components of the permeability tensor in the x'y' frame can be calculated by
tensor transformation
L
n K(12 (4.99)
where
Solution
The one-dimensional experiment shown in Figure 4.28 will allow one to relate flow rate to
pressure drop as follows:
dP dP (4.103)
K(12
(4.105)
\dx(J
, dx'
(4.106)
dx' dx'
12
1
- (4.107)
= KU cos 9 (4.108)
dP\
+1
WJ tan2 9
where /3 — K^/Kn- u'i and dP/dx'^ can be measured accurately only if the ratio of
Kii/Kyz is small or if K\i is small. Three experiments should be carried out at angles 9,
9 + A#i and 9 + A$2 where A$i and A$2 are two distinct and known angles such as 15 and
30 degrees, respectively. This will change the right hand side and now we will have three
equations and three unknowns to solve for 6, KU and K^ for a given preform [214]. •
~ (4.109)
where Vf is the fiber volume fraction and K is the principal permeability component in the
flow direction. The constant A is supposed to be a function of the fiber network. However,
it was found that one needed a different value of A for the same network at different fiber
volume fractions. Also for flow across an aligned bed of fibers, once all the fibers touch each
other one would expect the permeability to be zero in that direction. However, the model
was not able to capture that physics either. More constitutive equations were developed by
[215, 216] to explain some of these shortcomings.
Bruschke and Advani [216] found that one could use lubrication theory and the cell model
concept to describe permeability across an array of fibers more accurately as a function of
fiber volume fraction
K
where L2 = 4V//7T, and r is the radius of a fiber. Note that no empirical parameters were
needed.
Gebart [215] used similar logic to develop principal permeability values as a function
of fiber volume fraction in both along the fiber direction and perpendicular to the fiber
direction:
K
\\ = |r (l ~v^ (4 m)
-
5/2
(4.112)
In all cases, one finds that permeability of the preform decreases as fiber volume fraction
is increased. This results in reduced empty space between the fibers; hence, the resin will
face more resistance to impregnating this preform. However, there is no universal formula
for the relation between the preform permeability and the fiber volume fraction. This
relation is mainly a function of fabric type (i.e., whether it is a glass or carbon) and fabric
structure (i.e., whether it is stitched, woven, random, etc.). A typical graph showing the
change in permeability with fiber volume fraction is shown in Figure 4.29 for a plain weave
glass fabric. Furthermore, the detailed structure of a preform is different depending on
whether it is draped over a tool surface with double curvature and/or whether it has some
nesting and racetracking channels. All these changes will affect the permeability.
1 E-10 - -
Figure 4.29: A typical graph showing the change in permeability with fiber volume fraction
for a plain weave.
from the soil mechanics literature, and by using an analogy of flow along and across an array
of cylindrical tubes. However, as none of these idealized geometric arrangements resemble
the preform geometries realistically, computational approaches to solve for the flow/pressure
drop relationship in a unit cell with periodic boundary conditions offer the possibility of
extending the calculations to find the permeability of complicated fiber arrangements as
shown in Figure 4.30 [217, 218, 219, 220, 221]. However, these approaches do not comment
on the unresolved issues of the influence of surface tension [222], fiber wetting, and void
entrapment on permeability predictions. Also nesting of fabrics can influence the perme-
ability, shown in Figure 4.31, as can the addition of binders and stitches that hold the
fabrics together, hence the need to express the permeability and its relationship to fiber
volume fraction and the preform type empirically.
Since it is impractical to solve for the detailed flow field in the interstices between
the medium particles, the basic assumption to formulate a predictive analytic model for
permeability is that the porous medium is homogeneous, and one expects to find a repetitive
arrangement of the idealized geometry of the pore structure. This repetitive arrangement
is also called a unit cell. All approaches assume that Darcy's law holds on a macroscopic
level, and by choosing a geometry that closely resembles the repetitive fiber arrangement
a relationship between the flow rate and the pressure drop in the selected geometry of the
unit cell may be calculated. By comparing it to Darcy's law, one is able to formulate
an expression for the permeability in terms of the parameters of the selected geometry.
This approach can improve one's understanding of the flow through small gaps and help
develop a more science-based constitutive relationship between the preform permeability
and compaction of the preform.
Figure 4.31: Modeling permeability of more than one fabric layer [223].
K= Q (4.113)
AP/L
The porous medium was considered isotropic so one needed just one value of permeability
for a given architecture of the porous media.
3.5m
Figure 4.32: A column of sand that was used by Darcy in 1856 in order to measure the
permeability of sand [213].
Fiber preforms offer different resistances in different directions due to their architecture.
Hence, one needs to find the complete permeability tensor for such preforms. Three different
methods have been developed to measure permeability of fiber preforms. They are one-
dimensional linear flow, two-dimensional radial flow and three-dimensional hemispherical
flow. In all the experiments, one needs a flow rate measurement method and a pressure
drop measurement device. The one-dimensional experiment is the easiest to comprehend.
By guiding the flow along one of the principal axes of the preform, one can determine the
permeability by monitoring the flow rate and the pressure drop, as shown in Figure 4.33(a).
The pitfalls to watch out are racetracking along the edges and bending of the mold plates
(which makes the fiber volume fraction of the preform uneven).
While one-dimensional flow is from a line source as seen in Figure 4.33(a), two-dimensional
flow is usually a radial flow from a point source. The shape of the ellipse and the direction
of its major and minor axes allows one to determine the in-plane permeabilities as shown in
Figure 4.33(b). However, one needs to conduct some analysis to calculate these values. De-
tails are given in the book chapter by [214]. Three-dimensional permeabilities can be found
in a similar way by using a point source as shown in Figure 4.34. The details are given in
[223, 224]. Two- and three-dimensional measurements usually require recording of how the
flow front moves with time. This could be accomplished by video taping the experiment in
plane using a transparent mold or digitizing the flow front or using a SMARTWeave sensor
system [225] to extract information about the flow front in three dimensions.
Side view
^/i **
'' i
Here <5jj is the Kronecker delta, where 8ij = 1 when i = j and zero otherwise. The stresses
carried by the fibers are represented by <Jij and are due to elastic deformation stresses. The
viscous stresses arise due to the viscous deformation of the material and can be represented
by the stress tensor in the fluid as shown in Chapter 3 (T^ = rfjij}. The elastic deformation
stresses can be further subdivided into an axial stress and a bulk stress. The axial stress
is along the fiber direction, and the bulk stress is usually the compressive transfer stress in
the 2-3 plane. The axial stress can be positive or negative, but the bulk stress is always
compressive (negative).
The axial behavior of a fiber bundle is important during processes such as making of a
prepreg, filament winding and pultrusion. When one applies an axial force or tension to a
fiber bundle, Gutowski proposed a nonlinear relationship between the force and the fiber
Figure 4.35: Examples of geometries in which fibers are compressed during processing [6].
A0V0
(4.115)
Vf
where modulus E represents the bending stiffness of the fibers, VQ is the volume fraction
under zero stress, Va is the maximum volume fraction that can be achieved, and j3 is
a constant that defines the state of the bundles. AQ is the initial cross-sectional area
across which the load is applied. This relationship is plotted in Figure 4.37. It shows how
tensioning of the fiber bundle increases the fiber volume fraction.
In many other composites manufacturing processes bulk compressive stresses are applied
to the fiber bundle. For example in autoclave processing, one can derive the relationship
between the bulk stress azz and the fiber volume fraction as shown in Equation (4.116). The
plot reveals the rapid load transfer from the resin to the fibers during processing. Lower
resin pressure can promote the growth of the initiated voids, which is detrimental to the
composite.
lvf
3TTE
a-,-, = (4.116)
Note that the above two equations are derived but are based on the assumption that
the fibers are aligned inside a tow and make multiple contacts and have a slight waviness
h\
Figure 4.38: Deformation of a slightly curved fiber in a cell (redrawn from [6]).
Recently, Chou and Chen have addressed the compaction of woven and stitched fabrics.
They have adopted a unit cell approach to develop the stress and compaction relationship
and also allow for nesting of the fabric [231, 232, 233]. This compaction of fabrics and
nesting effects also influence the flow of resin as the permeability of the preforms change.
See [234] for more details.
4.9 Exercises
4.9.1 Questions
1. Why do we need constitutive laws?
2. What is the viscosity of a fluid?
3. What are the dimensions of viscosity in terms of mass (M), length (L) and time (T)?
4. What are the three common constitutive models for viscosity of shear thinning ma-
terials, such as thermoplastics? Sketch typical viscosity versus shear rate profiles for
these three models as well as for a Newtonian material.
5. What is the mathematical expression for the magnitude of the strain rate tensor?
6. What are the two common devices used to measure the viscosity of fluids?
7. For shear thinning materials at low and high shear rates, the power-law model gives
physically incorrect values of viscosity whereas the Carreau model correctly captures
4.9.3 Problems
1. The following empirical equation was proposed by [88, 89, 90] for the viscosity of a
thermoset resin system:
NQ+6a
(E^f aa
= A exp
" « (Rr)(^t
where the constants of the equation are listed in Table 4.2. (i) For RIM2200 resin,
plot 77 as a function of T, 30 < T < 90 for three different values of degree of cure,
a — 0, 0.05 and 0.10. (ii) Repeat (i) except that the resin is commercial DSM resin,
and temperature domain is 30 < T < 60. (iii) Repeat (i) except that the resin is
DGEBA and TETA resin, and the temperature domain is in 70 < T < 90.
2. Consider the reaction between two chemical groups denoted by A and B which link
together two segments of polymer chain PI and P% as follows: PiA + PaB = PiABP2-
The kinetic equation is ^ = k0 exp (~E/RT)CAO(l - C*)(A - C*) where C* =
A
^~ A and A = CBO/GAO- Plot C*(t) if the rate constant fco = 4, the ratio of
activation energy for the reaction E to the universal gas constant R is E/R = 250,
the temperature of the polymer is held constant at T = 450K, A = 1, and the initial
concentration is CAO = 0.4, How long does it take for C* to reach values of 0.5 and
0.99? How do the results change if the temperature is increased to 500K?
**• x
K * 10-10m2.
(i) Calculate the principal permeability components K\\ and Kyi- (ii) Draw the x-
y and 1-2 coordinate systems, and indicate the angle between the two coordinate
systems, (iii) Draw a typical flow front indicating the aspect ratio of the elliptical
shape.
7. You are to conduct two experiments of mold filling in a rectangular mold. The flow will
be constrained to be one-dimensional as the injection will be through a line injection
gate as shown in Figure 4.39.
In the first experiment, you place in the mold a preform that occupies 25% of the
mold, and you are given that its permeability for a fiber volume fraction of 25% is
5.1 Introduction
A model is an idealized mathematical representation of a physical process or a system. The
intent of this book is to teach the fundamentals involved in building process models for
polymer composite manufacturing methods such as injection molding, pultrusion and resin
transfer molding.
Manufacturing of composites has relied on experience and trial-and-error methods to
design, develop and fabricate a product. However, this approach has proven to be expensive
in time and money to develop new prototype geometries, which usually translates into high
risk, and therefore has hindered the use of composite materials in many potential industrial
applications. Use of process models can accelerate the path from conception to prototype
development and make these materials and their processing operations competitive with
metal and other materials in terms of their cost.
Composites processing models are built on the foundation of physical laws and appro-
priate assumptions and boundary conditions based on the understanding of the physical
phenomena and constitutive laws derived from experimental data. Once the model is well
posed in mathematical equations, the next step is to examine the behavior of the model
in response to changes in the process and material variables. This information can prove
to be very useful to design a mold or a die for the manufacturing method or to alter the
manufacturing process to create a successful part. To investigate the behavior of a com-
posite manufacturing process in a routine manner and with minimum effort, the process
model can be incorporated into a computer simulation. A computer simulation or a virtual
processing scenario is a combination of an idealized process model expressed in mathemat-
ical equations, a numerical method to solve the equations, and the computer software to
carry out the solution and display the results graphically, mimicking the physical behavior
of the process. Thus, such virtual composite process scenarios can provide valuable and
detailed information about the process and improve the understanding.
Ventl
Ventl Ventl
Fill time = 51 s. Fill time = 118s. Fill time = 169s.
Figure 5.1: Different mold filling scenarios with resin injection from three different locations
in an L-shaped mold. The mold filling simulation depicts the location of the flow fronts and
can indicate the filling time to completely saturate the mold. The filling times are 51, 118
and 169 seconds for the three selected injection gates.
Solution
Different gate locations/conditions can be investigated to achieve the mold filling goals.
In the L-shaped mold shown in Figure 5.1, three different gate locations under the same
injection pressure are filled to find which location results in the minimum filling time. A
mold filling simulation based on flow through porous media may be used to simulate all
three cases. The results show that the filling times vary considerably: 51, 118 and 169
seconds, in (a), (b), and (c), respectively. Instead of conducting trial and error in actual
mold filling, investigating different simulations will save considerable design time and hence
filling time. It can also save the part from "dry regions without resin" by choosing the
gate location appropriately, and placing the vent(s) at the last points to be filled. In this
example, 3, 1 and 2 vents are needed in (a), (b) and (c), respectively, as shown in the same
figure. •
Problem Definition
Identify
Boundary Conditions
Important Parameters
Physical Laws
Empirical Laws
Build Model Assumptions
Analogies with
Other Systems
Simplifications
Experiments
(How well physics
Solution to Equations
is described?)
Check
Check Assumptions
Boundary Conditions
Model Revision
Check Check
Constitutive Equations Solution Method
=T
H
->^>^>^>^>^>i.
^* T=TL
Figure 5.3: Non-isothermal simple shear flow of a fiber suspension between two parallel
plates.
OT
• VT = fcV2T + r : (VU) + R. (5.3)
which are the conservation of mass, momentum and energy equations, respectively with the
following boundary and initial conditions:
ux — U0, uy = uz = 0 at y —h (5.4)
ux = uy — uz = 0 at y =0 (5.5)
at y=h (5.6)
T = TL at y =0 (5.7)
T = T0 at t =0 (5.8)
This full system will require investment of many hours and computer resources. One
could simplify the system by assuming that the flow and/or temperature profile is fully
developed. One could simplify it further by assuming that viscous dissipation is negligible
and solve for the temperature profile before solving for the momentum equation, or one
could assume that the dependence of viscosity on temperature has no effect on the solution.
If the aim of the model was to capture the effect of temperature boundary conditions on
the flow rate and if one chooses the last option, one has made an assumption that will not
result in capturing this effect.
Barrel
Screw T = TL
Figure 5.4: Flow of a polymer suspension between the extruder screw and the barrel.
Solution
If we assume a steady-state and fully developed velocity and temperature profile, it will
allow us to represent velocity and temperature profiles as
u = u(y), v = w =0 (5.9)
(5.10)
M = ^e-'P1-^ (5.14)
where /j,0 is the viscosity at a reference temperature TO, and a is a material constant.
Substitution of Equation (5.14) into Equation (5.13) leads to
The energy equation (see Chapter 3) in which we use steady-state and fully developed
assumptions and retain the influence on temperature due to viscous dissipation allow us to
simplify the equation as follows:
dy2 \dyj • -
Equations (5.15) and (5.16) are coupled as u depends on viscosity which is a function of
temperature and temperature depends on u. If one further assumes that viscous dissipation
was not important (a good assumption for low viscosity materials but questionable for high
viscosity materials; usually one could estimate the importance of viscous dissipation with
Brinkman's number as we will discuss in a later section) then
y= du (5.19)
J
For TH > TL, (i.e., 9\ is positive), the velocity profile u(y) is sketched in Figure 5.5 for
061=0.1 061=1
different aQ\ values. Notice that, the velocity profile becomes linear as aO\ —» 0. This
is not surprising, because as aQ\ —> 0, fj, = /j10e~a(T~T°^ —> \i,0\ hence the dependence
of viscosity on temperature is very weak, and the thermal boundary conditions will not
cause deviation on the linear velocity profile. On the other hand, as aQ\ increases, the
velocity profile becomes very nonlinear. For a better approximation of the temperature
profile with viscosity dissipation included, substitute the derivative of u from Equation
(5.23) into Equation (5.16) and integrate to find the nonlinear temperature profile due to
viscous dissipation. •
How we simplify the model, pose the correct problem and find its solution is an im-
portant aspect in process modeling. We will focus on some important tools that will be
useful in doing so. They are (a) dimensional analysis, (b) common assumptions in poly-
mer composites processing, (c) various simplifications to represent boundary conditions, (d)
simplified geometric models, and (e) mathematical manipulations to simplify the solutions.
where T is the dependent variable and t and x are independent variables. k\ to £4 are
coefficients in front of the governing equation. As explained above, to cast this equation
in nondimensional form, one must find characteristic values for each of the variables in the
above equation. The variables are T, t and x. Thus, one nondimensionalizes them using
characteristic values Tc, tc and xc
T* = , t* = - and x* = — . (5.25)
J. c Tc Xc
Here Tc, tc, and xc will be related to some physical situation of the problem. For example,
if x was the variable along the length of a mold geometry, L, then xc = L. If T was the
variable temperature and represented the temperature of the resin, then Tc could be the
temperature difference between the initial temperature of the resin Tj and the mold wall
temperature Tw. Thus Tc — Ti~Tw\ this will ensure that the T* is also of the order unity.
When possible, have all dimensionless variables range from zero to unity.
Thus, the nondimensional form of Equation (5.24) becomes
Tc dT* Tc dT* Tc
^ t c d r * ^ ^ d2T* £4 tc_ = , }
dt* fci xc dx** fci x2 d(x*)
* 22 (
' }
Note that, as the equation is dimensionless and as the coefficient in front of the first
term is unity and dimensionless, all other coefficients (|^ ^-, |^ 1%, and ^ |f ) will also be
dimensionless. Now, one can neglect the terms that have coefficients much less than one.
For example if ^-tc <C x2 and ^-tc <C xc, one could neglect the terms with dT* /dx* and
d2T*/d(x*)2 as the coefficients in front of them ^|% and f^, respectively, will be small.
Thus, the partial differential equation (5.27) reduces to the ordinary differential equation
given below
f + lt- <«•>
Also, if one did not have a characteristic value for tc, one could choose it to be tc = kiTc/k^
by making the coefficient equal to 1. This reduces the nondimensional equation to
dT*
— + 1 = 0. (5.29,
k =k =10W/m.C
xx yy
k = 1 W/m.C
Figure 5.6: Epoxy-carbon fiber prepregs that are stacked and placed inside a metal tool
with mold walls at temperature Tw = 200°C.
Solution
The energy equation reduces to
T -K — U (5.30)
dx2 dz2
T T
dy<
as there is no flow and no viscous dissipation. Here, kxx, kyy and kzz are thermal conduc-
tivities of the composite in the x, y and z directions, respectively, and R is the volumetric
rate of heat generated inside the composite due to the curing reaction. Also, steady-state
is assumed (usually not a valid assumption).
One can further simplify Equation (5.30) and find its solution as detailed using the
following procedure. First identify independent (x, y and z) and dependent (T) parameters.
Then, form dimensionless variables:
T — T-
x = y =—,
yc z=
and1 1rr* — * (5.31)
Physically, this implies that most of the heat is conducted through the thickness direction
as h <C L,W even if kzz is an order of magnitude less than kxx and kyy. Thus, Equation
(5.32) simplifies to
d2T* R h2 , N
(5.34)
dz* -*-i
(flT*
^T = -R" (5-35)
where
R
R* = T _T . (5.36)
JU
Thus, R* is a dimensionless number that compares the volumetric heat generation rate
with the heat being conducted through the composite. The solution to Equation (5.35) is
straightforward if R is assumed to be constant:
dT*
d
-L = -JZV + d (5.37)
dT*
T* = 1 at z* = -1 or -— = 0 at z* = 0. (5.40)
V ;
dz*
One can solve c\ and c<z by using these boundary conditions. Hence, Equation (5.38) is
rewritten as
2
T*
1 - -^Z* +— + I
~ 2 Z + 2 +1
= -~(z*2-l) + l. (5.41)
dT*
= -R*z*. (5.42)
dz
Thus, the temperature gradient is maximum near the edges, and zero at the center. The
larger the value of the gradient, the greater the chances of developing of thermal stresses.
Hence, it is important to have a low value of .R*. •
Formulate nondimensional conservation of mass and momentum equations for flow of viscous
Newtonian fiber suspensions in long and narrow channels.
(L » H)
(TXX)C
T??
T,, =
_d I £ —n ("5 45)
HL dx H dz ' ^ ' '
Thus, we can make the order of both the terms of the order unity as follows:
^ + ^ =0 (5.46)
dx dz
if (Txz)c is chosen as -j^-. As H <C L, then (rxz)c ^> (TZZ)C and (TXX)C. Thus, nondimension-
alization suggests that shear stress (rxz) is dominant as compared to normal stresses (TXX,
As viscous materials achieve steady-state and a fully developed profile very quickly,
steady-state flow is not an unreasonable assumption when considering conservation of mo-
mentum. Also, one can ignore body (gravity) forces in the equation as viscous forces
dominate. Thus, the 2-D equation of motion simplifies to
Equations (5.50) and (5.51) can be rewritten in terms of dimensionless variables as follows:
dx^ dzj ~ w L d x ] \ L ) 8x
H3
w +w^dw\ = Pc idP\ + (H\(dtxz + 3tzz\ (5 55)
^) -wU*) U)bi- -0rJ-
#2 \ /
'
One can now choose Pc = p,qL/H3, such that the coefficient in front of the pressure gradient
term in Equation (5.54) is unity. Note that Equations (5.54) and (5.55) are scaled properly
as the scaling term for pressure Pc — [iqL/H3 is the same in both Equations (5.54) and
(5.55). Also, as H <C L, one can ignore all terms that have H/L or their higher powers.
Thus, the complicated equation of motion for flow in narrow gaps simplifies to
(5.57)
Fo = ^. (5.58)
a
Lc is the characteristic length in the direction of heat removal or heat addition by conduc-
tion, a is the thermal diffusivity and is given by
a = 4r (5.59)
P^P
where p and Cp are the density and heat capacity of the composite respectively, k is
thermal conductivity of the composite in the direction of conduction, if the composite
exhibits anisotropy.
Most commonly used nondimensionless numbers in polymer and polymer composite
processing are listed in Table 5.1.
V2 inertia
Froude number Fr = — gravity
gL
Pearson number Pn =
heat required to alter viscosity
Br heat conduction due to AT
Consider a polymeric composite that one would like to cool from 200°C down to 70°C as
shown in Figure 5.8. The top and bottom surfaces of the composite are kept in perfect
thermal contact with an aluminum tool at Tw = 25°C. After how long should the engineer
remove the part from the tool?
V25C
The thermal conductivity of the polymer is k = 0.1 W/m.C and that of the S-glass used
is k = 10 W/m.C in both the longitudional and transverse directions. The density of S-2
glass is 2500 kg/m3 and Cp is 0.41kJ/kg.C. For the polypropylene matrix, Cp is 2.0kJ/kg.C
and density is 900kg/m 3 . The S-glass fraction in the composite is 50%.
Solution
Parameters:
geometry W,L,h
process AT(=Tw-Ti),
material k,p,Cp,Vf.
Governing equation:
dT
(5.60)
X y z
(5.62)
L' h' T
-L 111
—T-
-*- 7,
h2 86
^dl
here a = k/(pCp) where
k = (l- Vf)kies-m + libers (5.64)
and
PCP = (1 - Vf)(pCp)ies-m + Vf(pCp){ibers. (5.65)
As h <C L, h/L and h/W are much smaller than unity, one can safely ignore the conduction
in the x and y directions. Also one can choose tc — h?/a, such that the coefficient in front
of dO/dt will become unity. Thus, the nondimensional form becomes
86 (5 66)
'
The two boundary conditions and the initial condition are given as follows:
at i = 0 (5.67)
at z = I (5.68)
f)f)
— = 0 at z = 0. (5.69)
dz
Note that the temperature distribution is symmetric about z = 0 due to the geometry and
boundary conditions at z = ±1. Hence, one could solve this problem in — 1 < z < I domain
by using #(— 1) = 1 and 6(1) = 1. Or, alternatively, and as we did here, one can use the
symmetry condition and solve the problem in 0 < z < I domain by using dO/dz(0) = 0 and
6(1) = 1.
Equation (5.66) is a diffusion equation. One can apply the separation of variables
technique to solve it along with the boundary and initial conditions
(5.70)
Substitution of Equation (5.70) into Equation (5.66) yields
ZT' = Z"T. (5.71)
Division of Equation (5.71) by ZT results in
TV <7"
Since the left-hand side of Equation (5.73) is a function of t only, and the-right hand side
of Equation (5.73) is a function of z only, Equation (5.73) can be true only if
T' X" , ^
-= — — = constant = -K2Z . (5.73)
J A
(5.74)
(5.75)
_ J A cos KZ + BSW.KZ,
= \D + Ez,
and
0 = H + JcosKze-KZ*. (5.80)
Application of the other boundary condition, Equation (5.68), results in the following:
« =^ (5-82)
0 ( z , 0 ) = 0 = l+ £ Jncos^. (5.84)
n=l,3,5,...
By using Fourier series expansion on Equation (5.84), one can determine Jn:
Jn = - - l ] c o s d x . = - (_i)n-2 (5.85)
nvr
for n = 1, 3, 5 , . . . . Hence, the final solution is given by
oo ,
e(z,t) = l- ]T A(_l)(n-l)/2cos^±e-(nu/2ri_ ( 5 _g 6 )
n=l,3,5,...
Note that if we want to know how long it takes to cool the composite from 200°C (9 — 0)
to 70°C [9 = (70 - 200)/(25 - 200) = 0.7429] at the mid point (z = z = 0), one needs to
evaluate and plot 9(z,t) numerically from Equation (5.86) and find out at what t value,
0(0, t) = 0.7429. In order to numerically evaluate Equation (5.86), a simple program can
be written as shown below in FORTRAN, or any other programming languages such as C,
BASIC, or mathematical solvers such as MATHEMATICA, MATLAB or Maple.
Figure 5.9 was prepared, and i was found to be 0.6484. This corresponds to t — tct —
( h 2 / a ) f = (0.012/0.00000358)0.6484 = 18.1 seconds. •
O
Q.
0.8
0.7429
o
§ 0.6
2
&
Q.
E
0.4
"CD
c
g
'w
c 0.2
OJ
E
T3
C
O
0.5 1.5
S
o
Nondimensional time, t
Figure 5.9: Nondimensional temperature versus nondimensional time at the center of the
composite part. In order to cool the composite from 200°C (0 = 0) to 70°C (9 = (70 -
200)/(25 - 200) = 0.7429), the part should be kept in the mold for t = 0.6484 which
corresponds to 18.1 seconds.
A polymer suspension flows through a channel of thickness h and length L. The schematic
is shown in Figure 5.10. The temperature of the polymer is T; before it enters the mold.
The mold wall temperature is kept at Tw. Formulate the energy equation that depicts the
importance of viscous dissipation in heating of the suspension.
T = T:
'"• . W 1; .-&
Flow rate, q Az t
(Per unit width in 1 >.x Material parameters: k, p, Cn,(4, h
y direction) I
,,,,,,,,,,,,,,,,,,, T —T
w 'K- •;/•
« - . .. 1 ... k-
(L » h)
dT +u
dT\ = id2T
fe
d2T
W
ot ir}
ox J 4 1 + Tl +r:Vu + R. (5.87)
As the flow is only in the x direction, v and w components of the velocity vector u are zero.
There is no internal energy generation; hence R — 0. Also, as h <C L
= rxz— (5.90)
as v = 0, w = 0, and as h < L, therefore, | | > |^, |^, and rxz > r^^, rxy as seen from
the example on flow in thin gaps. One could have arrived at the same conclusion using
dimensionless analysis.
Using the following nondimensionalization of the variables:
u u
uc q/h
AT
X
T — —
L '
/ AT A 8G\
PCP (uc —u - J = -^- ^ J + -^ ^-J (5.91)
here AT = TW — Ti, and a Newtonian constitutive equation TXZ — /j,^ was assumed. We
can select the conduction term as the one whose coefficient should be unity. Hence, division
of Equation (5.91) by kt\T/h2 results in
a \LJ\ dx
89
(5.93)
\ /
A
Fully Developed Region
Developing Region
the flow profile is rearranging from plug flow to a parabolic profile for a viscous Newtonian
fluid. Usually, this region is proportional to the diameter of the tube and the Reynolds
number [238]
— = 0.054 Re. (5.94)
If the Reynolds number is less than one, the entrance effects last for less than the tube
diameter, and the assumption of fully developed flow will not influence the important physics
or the solution.
Similarly, if one wanted to assume a fully developed temperature profile along the flow
direction, the following criterion needs to be met [239]:
Tangential direction, x
Tangential direction, x
Let us review here how lubrication theory allows one to simplify the equations of motion
for a Newtonian viscous material. The conditions that should be met before applying this
analysis to a flow problem are
• The material should be incompressible.
• The flow should be isothermal.
• The Reynolds number should be less than one (that means the inertial forces should
be smaller than the viscous shear forces).
• The gap height should be very small compared to the in-plane dimensions.
• If the gap height is not a constant, it could vary very slowly with the in-plane dimen-
sions (for instance if the in-plane dimensions were x and y, dh/dx and dh/dy <C 1).
one can ignore the flow in the thickness direction where t and L are the thickness and in-
plane dimensions, and K± and KL are the permeability components along these directions,
respectively.
Solid surface
tangent, t
(Continuity of velocities)
(Continuity of shear stresses)
Fluid 1
Figure 5,15: Boundary conditions along mold walls, on a free surface and at an injection
point.
where n and t are the normal and tangential directions to the wall.
P = -Hnlet (5.102)
or
Q = Q inlet (5.103)
an can
as shown in Figure 5.15. Pjniet d Qmiet be constant, or may vary with time.
P=c+AP
= f(X)
Unit cell:
P=q(y)
dPldx = dP/dx = r
I Warpt
Cbc = T (5.105)
.
km
Here, km is the conductivity of the mold material, h is the heat transfer coefficient between
heating/cooling fluid and the pipes, and k is given as
(kf + kr) + Vf(kf - kr)
K (5.106)
(kf - Vf(kf - kr)
where Vf is the fiber volume fraction of the preform, and kf and kr are the conductivity
of the fiber preform and resin, respectively. When the mold is insulated well (not much
heating or cooling), then Cic approaches zero. The fixed temperature (perfect thermal
contact) boundary condition corresponds to having a large value of C^c. This formulation
allows for the change in the mold surface temperature as experienced in practice. Due
to the exothermic chemical cure reaction of polymeric resin, the temperature of the resin
and hence the mold surface might increase away from the injection ports. Hence, this
formulation allows the modeler flexibility in specifying the boundary condition. Usually the
radiative boundary condition is not employed unless the temperature of the mold or that
of the surroundings is very high.
• Ability to check the usefulness and validity of the phenomenological law or constitutive
equation that is used in the process model.
Figure 5.18: Mold filling simulations for complex composite structures. Different gray tones
indicate the location of resin front at different times during mold filling.
It is also possible to take simplified geometries and couple them to represent a complex
geometry. This may geometrically produce an approximate solution, but will be able to
address some of the complexities in the material modeling. Below, we will present results
for Newtonian and power-law fluids for selected simple cases that were studied earlier in
detail in Chapter 3.
Couette Flow
Couette flow is the flow in a rectangular channel induced by a moving upper boundary at
a constant speed V. The no-slip boundary condition is assumed at the lower and upper
solid walls. Hence, the resin velocity is zero and V at the lower and upper boundaries,
respectively. The velocity profile is found to be
ux(y} = — y . (5.107)
Figure 5.19: Velocity profile in a rectangular channel induced by a moving upper plate at
a constant velocity V (Couette flow).
Fluid
P = Pi
y\
//////////////
yy, Fixed plate
Figure 5.20: Velocity profile in a rectangular channel induced by a pressure drop in the
direction of flow (P; > Pe).
(5.109)
Circular tube
•r Fluid
x
= 2R
Figure 5.21: Velocity profile in a circular tube induced by a pressure drop in the direction
of flow (Pi > Pe).
linearly as follows,
P(t) =
Kxx(l-Vf]
«_v (5.110)
to the square root of the permeability components in those principal directions. If the fluid
is injected at a constant flow rate Q, then the injection pressure rises logarithmically
P(t] = ^Q
•^yy
where Kxx and Kyy are the permeability components in the major and minor axes, and h
is the mold cavity thickness.
Solution
From Figure 5.23, we can see that there are two simple geometry flows involved: flow
through a tube and radial flow through the preform. At any time, the pressure difference
between the resin container and the location of the flow front is constant and equal to
(Patm — Pvac}- However, the pressure Pi at the inlet to the mold will vary with time, as the
Flow rate, Q
Pgtm —
Qt — L
(5.112)
u-
R
JD
Qr = 2TTR—h(l - Vf) (5.113)
Although the flow rate, Qr may vary with time, at any given time it should be the same
for all r from r = R^ to r = R considering the conservation of mass (see Figure 5.25). Thus,
= ur(2irr)h. (5.114)
From Darcy's law
K_dP_
(5.115)
IJL dr
therefore
Qr = - (5.116)
Equating Equations (5.112) and (5.118), and assuming Pvac — 0, one can find Pi as a
function of R, the location of the flow front.
Pii =
l&KLh
which is derived by taking the volume of resin in the mold and differentiating it with time
to obtain the flow rate.
Equating Equation (5.120) to Equation (5.112) and substituting Equation (5.119) for
Pi gives us a partial differential equation for R with respect to t:
TTRf
1- Patm. (5.121)
IQKLh
1 +-
The liquid is injected from a point source (denoted by O), and the shape of the flow front
is that of the surface of an ellipsoid with major axes forming angles 0^- with the axes of
a general coordinate system x,y,z. The permeability tensor for this case is in the form
used in Equation (5.122). After introducing a second coordinate system with axes x',y',z'
coinciding with the major axes of the ellipsoid, the permeability tensor reduces to the form
fci 0 0
*-] = 0 k-2 0 (5.123)
0 0 k3
where ^1,^2,^3 are the principal permeabilities of the fibrous material along the axes
x',y',z', respectively. After the principal permeabilities ki,k-2,k% are determined, the com-
ponents of the permeability tensor [K] in a general coordinate system can be obtained by
the transformation:
[K] = CT[K']C (5.124)
where C is the matrix of direction cosines of general coordinate axes x, y, z with respect
to the principal axes x',y',z'. To convert the original anisotropic medium into an isotropic
medium with reference permeability k, one can perform another transformation of coordi-
nates to new independent variables X, Y, Z defined by the relations [64, 242]:
2
k\^ 1/2
z
1/2
X=(^-} x',
- (5.125)
or
The result of these transformations is that the general problem of point injection in an
anisotropic medium is reduced to point injection in an isotropic medium with permeability
k = (kik^kz)1^ . The shape of the flow front in this case is that of a spherical surface (see
the image on the right of Figure 5.26). •
5.9.2 Superposition
Whenever the equations are linear, one can superimpose solutions. For example, when one
has a pressure driven flow and drag flow occuring in a cylindrical geometry, one can add
the solutions for the velocity profile to find the combined solution which is helical for this
case.
Solution
The important physics to understand here is that we have a combination of drag driven flow
and pressure driven flow. The pressure gradient is created because of flow being constrained
in the z direction. If one neglects the end effects, the velocity profile, uzd due to drag force
will be
(5.12T)
IliK \
as shown in Figure 5.27. Here K = d/D and RQ = D/2. Also Qj, the flow rate due to drag
flow will be
R],_dP\ (5.129)
uzp = 1-
and
Q p = \ l - K (5.130)
dz
The uzp due to pressure driven flow is shown in Figure 5.28. Hence, the velocity profile in
the annulus is a superposition of drag flow and pressure driven flow as shown in Figure 5.28.
I inner cylinder |
Figure 5.28: Velocity profile due to drag and pressure driven flow.
Note that we do not know — §f. However, we do know that total Q = 0. Thus,
Q = Qd + QP = 0. This allows one to solve for — ^j and substitute in equation for uzp.
Hence, the final steady-state fully developed velocity profile will be
Mr/RJ 3 (_£)
r
(5.131)
RO 4u V dz /
To find the force, one can calculate the shear stress along the fiber surface (Figure 5.29):
U
o
Figure 5.29: Shear stress along the fiber as it is pulled with velocity U0 from a concentric
cylinder.
Thus one can use the superposition principle to solve difficult problems more elegantly.
Solution
There are four temperature scales here [63]:
• AT = (Tj — TU,), temperature difference between entering and wall temperature.
• The inverse of £, which has temperature units.
Entering
temperature, Tt
H
„/ V 1
• Temperature rise k ' due to the balance of dissipation and conduction across the
thickness, H. Here k is the thermal conductivity of the suspension.
• Temperature rise J^WH* ^ue ^° ^e balance of viscous dissipation and axial convection
along the flow direction. Here p and cp are the density and the heat capacity of the
suspension, respectively.
Hence the three dimensionless numbers listed in Table 5.1 are important:
5 135
u, =
?(-£)
4/j,\dzJ <- >
Copyright 2003 by Marcel Dekker, Inc. All Rights Reserved.
dP AP PR - R
(5.136)
L
dP
Q = (5.137)
(5.138)
h 8/i V dx
dP AP
(5.139)
L
9.
W
= (5.140)
(5.141)
4/j.rh
dP
(5.142)
dr TT
P(r) = Pi- (5.143)
P = Pi at r = Ri. Solution not valid near r = 0. (5.144)
1-
v, = 1- (r/Ro) (5.145)
ln(l//e) dz J
5PN
Q = 1-K4- 4
(5.146)
8/j, \ dz J
where
dP AP _ Pe-Pj
(5.147)
~T ~
K —
_ (5.148)
R0
(e) Tangential annular drag flow between two concentric cylinders where the outer cylin-
der is rotating with angular velocity fi:
2K2 In K - K2 + 1
Q = 2nR0(R0-Ri)u0 (5.151)
4(1 — K) In K
5.12 Validation
Almost all process models that are developed involve one or more assumptions, and many of
them use constitutive or phenomenological equations to represent material behavior. Thus,
it is important to do a reality check and see if the model predictions agree with experiments.
Hence, there is a need for validation.
If a simulation of a model has been developed, it can provide many details such as
the pressure and temperature distribution, fiber orientation state at every location, etc.
However, one can only measure finite quantities from an experiment. Hence, the choice of
which variables to measure and where to measure them become important considerations.
Figure 5.31: (a) Schematic of a simple porous medium exhibiting a single length scale, (b)
Schematic cross section of a woven fabric unit cell, demonstrating the dual length scale
porous medium.
5.13 Exercises
5.13.1 Questions
1. What are the advantages of using a mathematical model and its analytical or nu-
merical solution, compared to relying on a trial and error approach during composite
manufacturing?
2. What are the main ingredients of a mathematical model of a manufacturing process?
3. What are the major steps in developing a model? In order to rely on the results of a
model, what criteria should be satisfied?
4. Should models be accepted as representing the real process exactly?
5. Why do modelers usually make simplifying assumptions in their models instead of
retaining all the physical details? How can a skilled modeler decide on which simpli-
fications are acceptable?
6. Why do modelers need to verify and validate their models with controlled experi-
ments? What is the difference between validation and verification?
7. What are the important considerations when comparing results from a model with
the experiments?
8. Can a model be a good representation of a physical manufacturing process used in
one production plant but not in a different plant? If so, do you need to adjust your
model? How?
9. Is there a unique model for a particular process?
35. What are the mathematical tools to simplify governing differential equations and
boundary conditions? Explain briefly.
5.13.2 Problems
1. Consider a large polymeric composite plate with a thickness of 2cm. Suppose that one
would like to cool it from 300°C down to 70°C. The top and bottom surfaces of the
composite are kept in perfect thermal contact with an aluminum tool at Tw = 25°C.
Considering typical composite material data, how long should the part be kept in the
mold?
2. Considering flow through a circular tube, what dimensions and properties of the prob-
lem do you need in order to calculate the length of the entrance region? When is this
length significant?
3. For what values of Graetz number, Gz, are velocity and temperature profiles assumed
to be developed in the direction of flow in a tube? Explain.
6.1 Introduction
The polymer processing industry creates at least 30% of the objects that we see around
us. There is a good chance that these objects were fabricated using one of the following
three manufacturing processes: extrusion, injection molding or compression molding. They
are all high volume production processes. Extrusion is a continuous process, injection
molding is an automated process and compression molding is a semiautomated process but
can manufacture large and complex parts that cannot be easily accomplished by injection
molding or extrusion. Most of the components fabricated are for nonstructural applications
as the polymer structure can withstand only a limited amount of load and stress.
These methods were also being used to process food which had solid particles imbedded
in a viscous fluid. Hence it was natural to extend these manufacturing methods to process
fiber suspensions as well. Fiber suspensions refer to chopped fibers suspended in a polymer
melt. Compression molding involved placing the material containing fibers and resin in a
solid form called the charge inside the mold cavity. Usually, the heat conduction mode is
employed to heat the charge by direct contact with the mold platen to melt the polymer
and create a suspension of fibers in the polymer melt. This suspension is compressed by
the platen to occupy the mold and take the form of the mold cavity. After solidification,
the part is removed from the mold. Unlike compression molding, which uses a mold platen
to melt and displace the material, the extrusion and injection molding process use a screw
to melt and move the material. The function of the screw in both these processes is similar
to some extent during preparation of the suspension. Examples of screws used are shown
in Figure 6.1.
The screw by its rotation can generate shear flow which causes viscous dissipation that
melts the polymer and mixes the fibers with the polymer melt at the same time. The
screw also acts as a pump that can push the polymer melt suspension into a die during the
extrusion process, or into a mold during the injection molding process. However, modifica-
tions need to be made to the screw material and design to ensure that it can withstand the
abrasive environment created by the fibers. The channel depth of the screw is also usually
increased to avoid excessive fiber length attrition, as fiber length plays a crucial role in load
transfer.
The goal of the screw is to produce a well-mixed suspension suitable for forming op-
erations. In extrusion, the suspension is pumped through an opening into a die to form
a continuous product, whereas in injection molding the suspension enters a mold through
one or more inlet gates and takes the shape of the mold. In extrusion, one does not need
Figure 6.1: Examples of screws used in extrusion and injection molding for mixing and
pumping the fiber suspension into a die or a mold [102].
Assign suspension
properties
v V
Calculate Newtonian
flow field (or input
initial guess)
r v ^ ^
Output velocity and Calculate orientation
orientation values states based on
streamlines
j
\
r ^
Update Theological
properties of the
suspension
^ )
_—-— ^^\~~~~~~.
•-> Decoupled Solution
Output velocity and
-*• Coupled Solution orientation values
Figure 6.2: Decoupled and coupled solution approaches to describe the flow and fiber ori-
entation in a fiber suspension [244, 245].
Figure 6.3: Two typical parts of thickness b, which is much smaller than the in-plane
dimensions.
Figure 6.4: (a) A three-dimensional box, and (b) its mathematically and conceptually
"unfolded" version.
The next step is to mathematically describe the material behavior to formulate the
model. The simplest approximation is to assume the material to be Newtonian and isotropic.
However, if the physics and experiments suggest something different, one could refine the
material behavior by including nonlinearities. It is always useful to address the isothermal
case first, as one can understand the important parameters if the complexities are kept to
a minimum. For the moment, one can assume the viscosity to be a function of x, y and z.
With these assumptions, one can now delve into the modeling.
A sketch of a typical compression molding flow is shown in Figure 6.5 with the coordinate
system appropriately attached and some of the parameters such as the in-plane dimension
L, height h and the closing speed s identified. For thin parts, L 3> h. As shown in
Figure 6.5, the flow in the x-y directions is due to the squeezing motion of the upper half
of the mold, which moves downward with a prescribed speed of s which is the same as —h
or —dh/dt. The instantaneous thickness of the part is h(x,y] and the velocity components
of the material are denoted by u, v and w in the x, y and z directions, respectively.
Next, one can carry out an order of magnitude analysis to evaluate the magnitude of
velocities in the x, y and z directions. The mass conservation equation for incompressible
fluids may be stated as
du dv dw = 0 ,6 1 .
7r
ax + 7T
ay + ioz
r - C-)
Using this as the guide, and the characteristic dimensions in the in-plane directions (x and
y) and thickness direction (z) as L and h, respectively, we find that
du Uc dv Uc dw s
(6.2)
dx dy L' ~dz h
where "~" denotes "of the order." We assume that the characteristic velocity, Uc, in the x
and y directions is of the same order and the dimensions in the in-plane directions are of
the same magnitude. Thus, substituting this order in the mass conservation equation, we
These relations will hold everywhere except very close to the edge of the flow domain and
if L 3> h. Most of the flow will be planar and the velocity in the thickness direction u; can
be neglected. However, the velocities u and v are still functions of the z direction. Thus,
to eliminate the z direction dependence, one can integrate the continuity equation in the z
direction
x rh(x,y
^ du(x,y,z] dv(x,y,z) ftl(X,
h(x,y) dw(x,y,z}
dz dz+ / (6.4)
IQ dx JO dy Jo dz
which can be written as
d(uh] d(vh] _
=S (6.5)
dx dy 2=0
u(x,y) = — !
x, y} Jo
u(x,y,z)dz (6.6a)
I r
v(x,y)
,y} =
= 77—r / v(x,y,z)dz. (6.6b)
h(x,y) Jo
Equation (6.5) is the continuity equation for a compression molding situation in which the
lower mold platen is stationary and the top platen is moving down at the speed of s when
in contact with the charge. Thus, this equation must be satisfied everywhere in the charge
domain.
Figure 6.6: (a) A typical Hele-Shaw velocity profile in which the direction of the velocity
vector is not a function of z at a fixed ( x , y ) location, (b) A typical non-Hele-Shaw velocity
profile: velocity vector direction changes with z.
(6.9a)
dx
Uc
(6.9b)
dz
As h <C L, one can safely neglect Equation (6.9a) as compared to Equation (6.9b). This
reduces the equations of motion to
9P = drzx
(6.10a)
dx dz
OP dr..zy
(6.10b)
dy dz
dP
(6.10c)
Equation (6.10c) implies that pressure varies within the x-y plane only. For isothermal
Newtonian fluids, one can substitute the stress-strain rate relationship
(6.11)
d_P_ _
(6.12a)
dx
dP__
(6.12b)
~dy~
Since P is not a function of z, Equations (6.12a) and (6.12b) can be integrated twice to
obtain the velocity profile through the thickness
1 dP
—-— (6.13a)
2fj, dx
1 dP
— — (6.13b)
2M dy
The boundary conditions to find the constants are no-slip at the mold platen walls,
u = 0, f = 0 a t 2 : = 0 and at z = h. (6-14)
(6.15a)
2/j, dx
1 dp
v= ,22
h (6.15b)
2/z dy
-
v
This form states that the magnitude of the velocity vector will vary across the cavity
thickness but the direction will not. Thus, we have proved that this flow is a Hele-Shaw flow
as shown in Figure 6.6(a). Heiber and Shen [257] have shown that even in a nonisothermal
or non-Newtonian case, this type of flow has the characteristics of Hele-Shaw flow.
This simplifies the analysis to a great extent. Instead of calculating u and v at every
x, y and z location, we can instead find the average velocity at each ( x , y ) location. These
average velocities can be obtained by integrating Equations (6.6a) and (6.6b) to give [258,
259, 260, 109, 261, 2, 262]
0 = -^-^ (6.17a)
12/i dx ^ '
= -*-^. y(6.17b)
dy '
For nonisothermal and non-Newtonian cases in which viscosity varies with z, [257, 258, 259,
260] provide the general form
« = -h£ox
£ (0.18.)
, = -S-9^ { (6.18b)
h 8y '
where S is a measure of the ease with which the suspension flows locally and is given by
S= I ( * ~ A ) dz. (6.19)
Jo f]
Here T\ — TI(Z) is the viscosity and A is the value of z at which the shear stresses rzx and rzy
are zero. For an isothermal case, S will be equal to /i3/12/z.
To find the governing equation for the pressure distribution, one can substitute Equa-
tions (6.18a,b) into the integrated mass conservation equation (6.5). This results in
d fS <9P\\ dd_ fS
(S_dP_\
OP _
dx V h dx J dy \ h dy
This equation is a Poisson equation and can be solved for P(x,y), once the boundary
conditions are specified. The physical boundary conditions will be that P = 0 at the free or
moving boundary surface (assuming there is no back pressure in the mold) and dP/dn = 0
Z
-dz. (6.21b)
If A is not known a priori, one of the boundary conditions such as no-slip, i.e. u — 0, v = 0
at z — h can be used to find A.
The governing equation (6.20) can be solved to determine pressure and velocity dis-
tributions at any instant in time as we treat this problem as quasi-static in nature. The
time-dependent nature of the pressure and velocity field enters into the problem because
as the mold closes the redistribution of the charge takes place and the flow front position,
where a boundary condition is applied, keeps changing. The usual procedure to predict the
flow in compression molding can be outlined as follows [263]:
1. Specify initial conditions (location of the charge, boundaries of the charge, closing
speed, etc.).
2. Solve for the pressure distribution at t = 0 using Equation (6.20).
3. From the pressure distribution, find the average velocities and advance the material
to form the new boundaries of the charge.
4. Solve Equation (6.20) in the new domain with the new height h and the boundary
conditions.
5. Repeat steps 3 and 4 until the charge touches all the mold walls or until the maximum
force the compression molding machine can generate has been reached or until one
achieves the final thickness in the z direction. To calculate the force required at any
time t to move the platen at a speed s, one must integrate the pressure field at that
t over the charge domain.
The two main assumptions that are used in this model were negligible w and insignifi-
cant in-plane stresses. These assumptions are valid everywhere in the domain except near
the edges of the compressing material. Thus, one can gain tremendous simplification by
sacrificing a little bit of accuracy at the edges. The Hele-Shaw model is always valid as long
as the shear stress terms are larger than the in plane stresses. This implies that
drxz r\wuc drxx t)cuc
(6 22)
-^~-h^^^~^' '
In compression molding, usually a cold charge is molded in a hot mold so the viscosity of
the material near the mold wall, r)w, where most of the shearing dominates, is lower, as
The viscosity could differ at the most by an order of magnitude from the surface to the
center. However, usually if L is about an order of magnitude larger than h, which is
definitely true for a part such as an automobile hood, one can usually justify the use of the
Hele-Shaw model.
It further assumes that there is no shearing across the thickness. This leads to u and v
being only functions of x and y, and therefore u = u and v = v. The equation of motion
due to this assumption reduces to
dx dx dy
dP
ay = d^
ax + d^
ay (6.25b)
with the boundary conditions that velocity normal to the mold walls is zero and the total
stress normal to the free flow front is zero.
Lubricating squeeze flow solutions can often be trivial, providing bi-axial extensional
flow in the x and y directions. This implies that as you squeeze in the z direction, the
material elongates in the x and y directions equally, until it encounters a mold wall and
then it elongates only in the direction it can as shown in Figure 6.7. The pressure inside the
charge is usually constant. To find this pressure one can use the boundary condition that
the total stress normal to the boundary, ann is zero. If vn is the velocity component normal
to the free flow front boundary and n denotes the normal direction, it can be expressed as
(Jl I
<?nn = ^-^ ~P = 0 (6.26a)
f/1 1
or, 2ri—- = P. (6.26b)
CfiL
Here P is the pressure of the material inside the charge which is not equal to the pressure
outside.
Consider a polymer charge being squeezed between two parallel disks as shown in Figure
6.8. The disk surfaces are lubricated so that there is a slip boundary condition in the radial
direction. Assume that inertia and body forces are negligible. The thickness of the polymer
charge is h and the radius of the disks, and hence the charge is R. The disks are compressed
in a press with a closing speed of s — —dh/dt = —h.
Closing speed
s =-dh/dt
Figure 6.8: Schematic of a fiber reinforced polymer charge being squeezed between two
parallel disks in compression molding.
Find (i) velocity distribution, (ii) stresses and the pressure in the fluid, (hi) the total
force exerted on the upper disk by the fluid, and (iv) h(t) if /i(0) = h0 and the total weight
of the upper disk and the press is W.
1 9 , . dvz
vz = — s at z —h (6.28a)
Tzr = 0 at z =h (6.28b)
vz = 0 at z =0 (6.28c)
rzr = 0 at z =0 (6.28d)
vr = 0 at r =0 (6.28e)
arr = -Pa at r =# (6.28f)
where Pa is the atmospheric pressure. Let's assume that vr = vr(r) and vz — vz(z) consid-
ering a thin charge, R 3> h. The assumption of vz — vz(z) meets the boundary conditions
(6.28a) and (6.28c). With these assumptions, the partial derivatives in Equation (6.27)
reduce to ordinary derivatives, and hence Equation (6.27) can be rewritten as
Only the continuity equation (conservation of mass) was used to find the velocity field.
That means the equation of motion (conservation of momentum) was not used for the
(6.34b)
is a biaxial elongation tensor. If the constitutive law between the stress and strain rate
is known, one can also calculate the stress tensor. We will assume a Newtonian material
behavior to evaluate the viscous stress tensor for the material. The stress tensor is given by
(6.36)
What is the pressure distribution within the fluid (polymer charge)? One can use the
stress tensor in the equation of motion in the r and z directions to evaluate that. The
equation of motion in the r direction is as follows:
dP 1 d IT0e . 9rrz
OTrg
0 = (rr,. r ) 4
dr r dr de r f dz
r
dP I d /iS> I 8(0) (Ais)/h + d(0)
(r
dr ' r dr h )\ + r 89 r dz
0- +0 (6.37)
or rn rn
or, simply dP/dr = 0. The equation of motion in the z direction gives:
6P I d ZZ Idrg, OT
0 = ~-^- + --^-(rrrz)^ --- ^- + ^~
oz r or r oti oz
o = _
dz r dr r dO dz
OP
0 = -— + 0 + 0 + 0 (6.38)
dz
or, simply dP/dz == 0. Hence, P is a function of neither r nor z. So, it is a constant P = C.
In order to find C, we use the boundary condition (6.28f):
oyr = rrr - P
-Pa = ^-P. (6.39)
P = ~T- (6.40)
which is higher than the pressure outside Pa. As seen from Equation (6.40), the higher the
viscosity /j, and closing speed s, the higher the charge pressure P.
(ui) o~zz is the force per unit area exerted on the polymer charge. So, the force F on the
upper disk is the integration of — azz over the disk area:
F = - I azz dA — - 2?rr dr
R
(6.41)
= W + TtR (6.42)
force due to atmospheric pressure
force exerted by the polymer charge
Here, we neglect inertia forces. Figure 6.9 shows the force balance. Hence, W —
Pa (Atmospheric pressure)
mum m u
tmtrttmm. (Normal stress)
Figure 6.9: Force balance on the upper disk for Lubricated Squeeze flow Example 6.1.
If a constant closing speed s is desired, then W should be increased with time by increasing
the press force. One can see this mathematically due to the presence of h(t) in the denom-
inator. If W is kept constant, then the closing speed s will decrease with time. In this
problem, W is specified as a constant, and we will calculate h(t). The closing speed of the
disk s is —dh/dt, hence
dh Wh
(6.43a)
~dt
\nh = - (6.43c)
h(t)=c2e (6.43d)
In Equation (6.43b), integration of both sides with respect to t is carried out to obtain
Equation (6.43c). The constant c% (= eci) is found to be ho by using the initial condition
ft-(O) = ho. Hence,
Figure 6.10: The mold gap thickness as a function of time for Example 6.1.
Obviously, this result breaks down at smaller values of h, as the assumptions of slip BC and
no surface tension at the boundaries will no longer be valid. •
vi = frji (6.45)
at the mold walls. A schematic of the predicted shape of the velocity profile vr(z] is
presented at the bottom of Figure 6.11.
In the following example, the proposed alternate boundary condition for the Hele-Shaw
model will be investigated through an analytical case study.
Example 6.2: Hele-Shaw Model with Partial Slip Boundary Condition [264]
Investigate the flow induced by the compression molding of a cylindrical charge between
circular plates which can undergo partial slip with respect to the mold walls. The plates
are moving at a speed of h towards each other, and are of radius R as shown in Figure 6.12.
Solution
The position of the charge flow front is defined by R. The equations to be solved are as
follows:
fh
2-xr \ vrdz + Trr2h = 0 (6.46)
Jo
drrz dP (6 47)
-w - ~w -
Equation (6.46) was derived by integrating the continuity equation over the charge volume.
Equation (6.47) is the momentum equation with the assumption that shear stresses are
higher than normal stresses. The boundary conditions are assumed to be
Figure 6.11: Velocity profiles vr(z) for various flow models [264].
2h
dr J \ / \ dr
Then Equation (6.46) is used to find the pressure gradient:
dr o,, , , , , , - (6-5°)
If /3 is given a value of zero, the velocity profile reduces to that of the normal Hele-Shaw
solution. If - —> oo, the velocity profile reduces to that of the lubrication squeezing solution.
Several velocity profiles have been plotted to examine the influence of the arbitrary
coefficient /3 on the amount of slip occurring at the mold walls.
As the velocity profile varies linearly with h, and scales linearly with radius r, all profiles
have been calculated for a single h and radius. Calculated velocity profiles at different plate
separations are shown in Figure 6.13. As the separation of the two plates decreases during
the process, the average radial velocity through the thickness increases. Therefore each ve-
locity profile calculated is normalized to its center line value, vmax. These plots demonstrate
how the amount of slip at the mold boundaries develops as the process continues.
Three different sets of velocity profiles are plotted in Figure 6.13. In each plot, the
velocity profiles have been calculated at h values of 0.1, 0.05, 0.01, 0.005 and 0.001 m. Each
set of velocity profiles has been developed for a certain value of /? in m/(Pa.s).
For Newtonian fluids, the ratio of the slip velocity at the wall to the centerline velocity
increases as the separation of the mold wall decreases. The fluid is slipping more and more
as the mold closes.
The pressure profiles were found by integrating ^-, applying the boundary condition
P — Pa at r — R. For convenience, Pa has been set to zero at this point. For Newtonian
fluids,
The expression for pressure distribution is useful to calculate the force required for
compression of the charge.
The squeezing force required to move the circular disks towards each other will be
calculated. This force was found by integrating the pressure at the mold surface from r — 0
to r = R:
(R
P(r) r dr. (6.52)
Jo
The resultant force for a Newtonian fluid is calculated to be
Squeezing force will be calculated for a particular theoretical experiment. The initial
diameter of the charge is 0.3m, and it has a thickness of 8mm. The closing speed h is
h=0.1m
h=0.05in
h=0.01m
h=0.005m
h=0.001m
P =-0.005 m/Pa-s
decreasing h
0.4 0.6
Vr/Vma
h=0.1m—
h=0.05m--
h=0.01m ~
h=0.005m
h=0.001m
: ' ! decreasing -
0.4 0.6
VrA/ma
Figure 6.13: Normalized velocity profiles for compression molding flow of a Newtonian fluid
with partial slip boundary conditions along the mold walls [264]. Top: (3 = 0.0005, Middle:
(3 = 0.005, and Bottom: j3 = 0.05 in m/(Pa.s). In each plot, the profiles are shown for
different values of h in meters.
25000 -
20000 -
8 15000 -
10000 -
5000 -
Figure 6.14: Force required to compression mold a Newtonian material with partial slip at
the mold walls [264].
As f3 is increased from zero, the squeezing force approaches the lubricated squeeze flow
model result. This is to be expected, because the velocity profile approaches "plug" flow
as /3 increases. However, as j3 increases one would expect the Hele-Shaw model to break
down as the in-plane stresses will become as important as or more important than the shear
stresses. Thus, the results for the pressure distribution and the force are valid only at low
values of j3. •
Figure 6.15: Velocity profile and the migration paths of cold material entering a hot mold.
In (a), the no-slip velocity boundary condition leads to migration of the material from the
center to the mold walls. In (b), the slip boundary condition at the walls and temperature
dependence of viscosity cause the particles to migrate from the surface towards the center.
Solution
For lubricated squeeze flow, the velocity components u and v do not vary with z; therefore,
du dv
(6.54)
dz dz
Differentiating the equation of continuity with respect to z for an incompressible fluid,
d i du dv dw
(6.55)
x + oy
jr + To =°-
Interchanging the order of differentiation leads to
d i du\ d dv _j_ _ Q
(6.56)
dx\dzj ' dy \dz'
Substitution of Equation (6.54) into Equation (6.56) results in
= 0. (6.57)
Speed, s
then the fluid or the material particles will remain at the same position A throughout the
flow. However, the particles will move in the x-y plane.
Here dT/dt represents the time derivative at a fixed position A. Equation (6.61) is similar
to the conduction problem of a slab initially at temperature TO, whose walls are exposed
to Tw at t = 0 except the slab is shrinking across its thickness with time. The solution is
given by
= I + 2 £ - - exp -» cos (flnA) (6.62)
Tw — TO
an — TT I n -\— ] (6.63)
\ £)
and t*(t) is an integrated time
n dt
Here t' is the dummy variable and hf is the final thickness of the charge. The changing
thickness of the sheet warps the time scale, accelerating the conduction of heat as the sheet
becomes thinner, as illustrated in Figure 6.17. •
6.2.9 Cure
In modeling the cure and coupling it with heat transfer, one must usually account for the
dwell time, an initial period when no appreciable curing takes place due to the addition of
inhibitors. Curing will be slow initially but then will rapidly accelerate and eventually level
off. Isothermal curing behavior of unsaturated polyester is shown in Figure 6.18 by plotting
degree of cure c versus time t. The degree of cure is defined as
c =- (6.65)
It
where q is the cumulative amount of heat released from the beginning t = 0 to the current
time t as compared with the total heat of reaction qt- As listed in Chapter 4, there are
many models to describe the cure kinetics. For example,
(l-cT (6.66)
Figure 6.17: Nondimensional temperature at the center of the charge as a function of time
with an initial thickness hi = 10mm, final thickness hf = 3mm, closing speed s = Imm/s,
and thermal diffusivity a = Imm 2 /s.
9t
Time
where S = qtdc/dt. At the same time, dc/dt is dependent on cure and temperature as seen
from Chapter 4 and Equation (6.66). Hence these equations need to be solved simultane-
ously if the curing has initiated. The viscosity of the material also dramatically increases
with the cure rate, which makes the flow of the material difficult. Many process engineers
prefer to add inhibitors to the resin so that they can fill the part with low viscosity suspen-
sion, before the resin cure initiates. Therefore, the convective terms on the left-hand side of
Equation (6.68) vanish, and it becomes a transient heat conduction problem with a space
dependent energy generation term.
Solution
We will assume that the mold is saturated with resin when the resin cure initiates. As the
in-plane dimensions are much longer than the thickness h in the z direction, nondimen-
sionalization of the energy equation is carried out with conduction only in the z direction.
Other variables can be nondimensionalized as follows:
T
T* = - - (6.69)
Jo
(6.70)
c
*' = S I"-71'
where ATb = Tw — TQ (= difference between the wall and initial temperature) and tc =
pcph2 /k is the characteristic time for heat conduction. To make dc/dt dimensionless and
of the order unity, one must use the characteristic time of reaction, tr instead of tc. Thus
The value of tr will depend on the kinetic equation used. Also the adiabatic temperature
rise just due to the heat released from the reaction can be estimated as follows:
Dar
d(z*Y2 + *- (6-74)
Da = (6.75)
It represents the ratio of the rates of temperature rise due to reaction to conduction. Hence,
when Da is large, the reaction kinetics dominate the heat transfer, and if Da is small then
heat conduction dictates the heat transfer. •
Figure 6.19: The convection of fiber orientation during compression molding [117]. The
direction of the line shows the direction of most fibers and the length of the line shows the
strength of the difference between the fibers in that direction as compared to the direction
perpendicular to that direction. Thus a dot would represent equal orientation distribution
of fibers in all directions and a long line will denote that most fibers are oriented in that
direction [117].
6.3 Extrusion
Extrusion is the most important forming method in polymer processing. More polymers
and other materials are converted into useful objects by extrusion than any other method.
An extruder is essentially a pump to displace viscous materials. It is also capable of other
operations such as mixing along with pumping. The extruder can be fed with molten
material or with solid polymer chips, beads or pellets. In the case of composite materials,
the pellets have short fibers embedded in the polymer. A melting operation is achieved few
diameters downstream of the inlet. The operation is called plasticating, and the extruder is
known as a plasticating extruder. If molten material is fed to the extruder, it is known as
melt extruder. If two dissimilar polymers are to be blended, or if the pellets contain short
fibers, an additional operation of mixing is performed by the extruder.
The heart of the extruder is the screw. The details and the nomenclature for the screw
are shown in Figure 6.20. Table 6.1 shows the various geometric relationships. The screw
in the extruder is usually between 7.5 centimeters and 30 centimeters in diameter. The
speed is usually in the range from 20 to 500 revolutions per minute with a working pressure
between 50 and 1000 atmospheres. The flow rate could be as high as 2000 kg/hour. The
length to diameter ratio could be anywhere from 5 to 100.
The extruder can be divided into three zones, the solid conveying zone, the melting
zone and the metering (melt conveying zone). The extruder operations can be grouped as
follows: (1) melting, (2) mixing and devolatilization, and (3) pumping. The transport of
the material inside the barrel is similar to a nut held in a wrench with the screw rotating
in the nut. If a rotating screw is prevented from advancing, the nut will slide in the wrench
as shown in Figure 6.21. In a similar manner the material slides between the screw and the
barrel.
Nut (Material)
obstruction
preventing
IIS screw
;
advancement
Screw
Wrench (Barrel)
Figure 6.21: The principle of material advancement in an extruder can be compared to the
sliding of the nut if a rotating screw is prevented from advancing.
Barrel
Figure 6.22: Schematic of the unwrapped extruder channel and the coordinate system on
the screw.
Next, we need to make simplifying assumptions to pose our problem correctly and
uniquely. For our analysis, we can assume Newtonian fluid, although similar analysis has
been conducted for non-Newtonian fluids. We will assume isothermal conditions. This
allows us to solve the flow in the x and z directions separately and then add the solutions
linearly as the fluid is Newtonian. There is no coupling with the temperature field.
The relative motion between the screw and the barrel becomes equivalent to steady
motion of a plane at an angle 9 to the helical axis z, which is the helical angle of the screw
as shown in Figure 6.20. However, the drag flow is generated in the x and z directions.
With inertia neglected, the equations of motion for Newtonian fluids reduce to:
dP_ (6.76)
dx dx2 dy2 dz2
T
dx = M ^ +^ r l (6.78)
dP , ~ ^ . ~ u,z , (g 79)
The pressure gradient in the z direction is not a function of x or y, and ux and uz are not
a function of z as the flow is fully developed in the z direction. Thus, in Equation (6.79),
the pressure gradient has to be a constant along the z direction. Hence
' pji '•„,
dz 2r + 2r l' = C - (6.80)
function of z only not & function of
Thus
where PQ is the pressure near the exit of the screw and Pi is the pressure at the inlet. I is
the helical length of the screw and is given in Table 6.1. Now, one can solve
AP d2uz d2uz
T (6'82)
on all four walls. If W 3> H, one can ignore d2uz/dx2 as compared to d2uz/dy2 and the
solution will be valid everywhere except near the walls at x = 0 and x = W. The complete
solution for Equation (6.82) with boundary conditions (6.83) and (6.84) is given by an
infinite (Fourier) series. However, the most useful result is the volumetric flow rate
f ,- WH
Q= uzdxdy = —-AP(Fp}. (6.85)
JO JO 12/it
where Fp is the shape factor given by the infinite series and is a function of the aspect ratio
H/W [102]. The variation of this function with the aspect ratio is shown in Figure 6.23.
One can combine the drag driven flow in the z direction with the pressure driven flow
in the same direction since the fluid is Newtonian and one is dealing with the isothermal
case. Thus, the flow rate is
1 WH3 AP
(6.86)
where Vz — V cos 9 and I is the helical length as listed in Table 6.1. Note here the pressure
flow is "backflow" in the direction opposite to that of the drag flow. For most extruders
W/H is large; hence Fj and Fp are close to unity. The barrel velocity is V — irDN and
Vx = VsmO and Vz — VcosO and N is the screw speed in seconds.
There is also a transverse flow ux(x,y) due to drag flow in that direction induced by
Vx. For large W/H, ux will have weak dependence on x, except near the walls (the screw
flights). The momentum equation can be rewritten as
dP
0 = -TT-
dx +' ^ 8y2
(6.87)
H
}' —
dx
(6.90)
ux(H)=-Vx
Figure 6.24: The velocity component ux as a function of y inside the channel created by
the extruder screw and barrel.
LJ = F U
cW
xy UT + T, dxdz. (6.94)
OJO
H H
The -^ term in Equation (6.95) is zero since a uniform flow in the z direction is assumed,
and the -^j- term in Equation (6.96) is negligible compared to the other term ^ for large
W/H ratios. The velocity components are
6 98
<- '
Hence the power required is
2 2
(6 99)
-
For basic performance, one needs to calculate Q, <1> and AP. Equations (6.86) and (6.99)
are two of the three equations. We need a third equation to uniquely solve for Q, uj and
AP.
If there is no restriction at the end of the screw, AP = 0. That implies that there is no
mechanism to build up the pressure. Normally, there is a die downstream which establishes
the pressure rise down the extruder axis. Hence the additional equation comes from pressure
drop across the die associated with die output.
For Newtonian fluids
Kr
Q = —-AP (6.100)
M
where KG is related to die geometry and characteristics. To analyze this further, rewrite
Equation (6.86) as
AP
Q = AN-C - (6.101)
V
where
A = ]-WHFd cos 6-rrD (6.102)
and
WH3
C = ^rFp (6.103)
where the channel length is given by I = L/s'mO, and L is the screw length, and ./V is the
screw speed in seconds, then solving Equations (6.100) and (6.101) simultaneously results
in
(6 104)
11 A
-
(6.106)
The flow rate Q, which is normal to the helical axis should still be a constant; hence
(007)
and hence
(6.108)
As H(z) is a function of the screw dimensions, it is a known quantity. For example, consider
a linear approximation
H(z) = E0^
TT/ \ TT ,
(6.109)
dp
(6.110)
C = (6.113)
where
and
Q = AN--2L%-. (6.117)
fjL(z) az
Integrate with respect to z to obtain the pressure drop AP
rl
o
It is necessary to introduce the temperature rise into the equation
where the last term QdP in Equation (6.119) is neglected as it is small. Prom Equation
(6.99)
w = E»N2l + ANAP (6.120)
T. (6.123)
One can substitute Equation (6.123) into Equations (6.117) and (6.118), which results in
and
,025)
Q is now quadratic with N, and AT enters implicitly. Let us assume that the viscosity
dependence of temperature has the form
H = ae~bT. (6.126)
Recalling that
_
EN2
By multiplying both sides of Equation (6.128) by ebT and integrating with respect to z, the
following equation is obtained:
1 (6.129)
(6.130)
Q=K^P
H(Z) = Kc^P
V0
Q can also be written from the screw characteristic equation Equation (6.125):
. ,032)
AT
AT = ^ (6.133)
o
and from Equation (6.131)
By inserting Equations (6.133) and (6.134) into Equation (6.131), one can get
pCpKxmx
N1 (6.136)
where NI = C/KG and JV2 = (Apcp)/(NlV0bE). For any JVi and N2 one can obtain the
solution for x and hence find Q and AP. The results are plotted in Figure 6.26.
More on the modeling of extrusion can be found in textbook written by Tadmor and
Gogos on this topic [102].
Figure 6.26: Variation of screw parameters (Ni and A^) for Adiabatic case when the viscos-
ity is temperature dependent [3]. Find x from JVi and N%, and use it in Equations (6.133)-
(6.135) to calculate AT, Q and AP.
ejection of
the object
from the mold
6.4.2 Materials
Both thermoplastic and thermoset polymers can be injection molded. Most often used ther-
moplastics are PE (polyethelene), PPS (polyphenylenesulfi.de) and PA (polyamide), which
are usually glass-reinforced. High-performance thermoplastics such as PPS and PEEK
(polyetheretherketone) might be carbon-reinforced. Unsaturated polyesters and phenolics
are the most common thermosets used. The reinforcement is usually short glass fibers in
the range of 0.5-5 mm in length. The raw material supplied to the barrel of the injection
machine is usually in the form of pellets or powder.
6.4.3 Applications
One of the advantages of injection molding is that geometrically complex components with
highly accurate dimensions can be economically manufactured in high volumes. Also, as
the process is automated, the cost per part will decrease rapidly as the volume increases. It
is not uncommon to make several hundred thousand to a million parts with a single mold.
For example, most plastic household items are produced utilizing this process. Examples of
short fiber reinforced articles are components for light machinery, equipment housings and
sprockets, and a variety of automobile components including appearance parts such as tail-
gates, instrument clusters, lamp housings, oil pumps, water pumps, bumpers and air intake
manifolds. This wide range of applications is constantly increasing. Other possible appli-
cations include replacement of conventional structures consisting of numerous components
(manufactured separately and assembled) with a single, lightweight, net shape composite
part.
The number and variety of parts manufactured by injection molding is large. It includes
items as small as paper clips to as large as automobile bumpers. Nonetheless the applications
are limited to non-load carrying components because of the poor mechanical properties of
the injection molded parts caused by:
• Relatively low fiber volume fraction.
• Short length of reinforcement fibers.
• Not much control over the alignment of the fibers induced by the flow.
• Fiber degradation in the feed screw and injection gate.
Despite these limitations of the injection molding process it has numerous advantages such
as:
Unlimited geometrical complexity of the manufactured parts.
In order to obtain parts with desired mechanical properties, it is equally important to design
the mold and control the process in such a way as to get favorable fiber orientation states.
Thus, models that predict the evolution of the fiber orientation will also be useful.
2b
Figure 6.28: Typical dimensions of parts made using the injection molding process.
(6.137b)
ay \ ox2 ay
— = M ^-±2 (6.138a)
ox oz
=M—I- (6.138b)
Since P is not a function of z, each equation can be integrated twice to give the parabolic
velocity profiles
1 rlP
_i_±L^2 (6.139a)
u= + CiZ + C2
2fjL dx
v =- z2 + c3z + c4. (6.139b)
2^ dy
The boundary conditions are that the velocity gradients are zero at the center due to
symmetry and no-slip at the mold walls. Mathematically,
— = —=0 at z = 0 (6.140a)
oz oz
u =v=0 at z = b. (6.140b)
U
~ ~2u ~dx
',222
z -bL2) (6.141b)
dy ^ '
This solution looks similar to flow between infinite parallel plates subjected to a pressure
drop boundary condition. However, in this case, the pressure gradient terms, dP/dx and
dP/dy are functions of x and y and are still unknown.
Also, as in compression molding, one can take the ratio of u/v and find that it is not
a function of z, thus exhibiting the feature of Hele-Shaw flows. This allows one to average
the velocities in the z direction at every x and y and simplifies the problem. These average
velocities are
1 fb
u(x,y) = - u(x,y,z)dz (6.142a)
b Jo
i rb v(x,y,z)dz.
v(x,y) = - (6.142b)
o Jo
Once u and v are known functions of x and y, and as one knows the velocity profile in
the z direction from Equations (6.141a) and (6.141b), one can find the velocity as a function
,= - { (6.143b)
3M dy '
The incompressible continuity equation is
du dv dw , ,,
7T
ax + 7T «- = °'
ay + az °44
This contains the velocity component w which we have ignored so far in the analysis, so we
may be tempted to do so here as well. However, although we know that w will be much
smaller than u and v, this does not imply that dw/dz will be smaller than du/dx or dv/dy.
In fact, as the thickness is much smaller than the other two directions, one can safely argue
that the dw/dz term will not be smaller than the other two terms in the equation. Thus,
one cannot just drop dw/dz. To address this issue, the general approach adopted is to not
satisfy the continuity equation exactly at every location in the suspension. Instead, one can
integrate the continuity in the z direction across the cavity thickness,
,fe = 0. (8.145)
o ox o y o dz
This will ensure no material is lost. Rearranging the order of integration results in
a rb d fb rb
— \ udz+ — \ vdz+ / dw = 0. (6.146)
dx Jo dy Jo Jo
The first two integrals are bu and bv, while the third integral must equal zero (since w
equals zero at z = 0 and at z — b), which results in
Substitution of Equation (6.143) into Equation (6.147) results in the governing equation to
be solved for the pressure distribution
If b is not a function of x and y (if the part is of uniform thickness, 26) and if the viscosity
is constant, the equation can be further simplified to the Laplace equation
Qlp
^=a <6-149'
Hence, by using simplifications and physics of the process, we have been able to reduce the
problem from four variables (u, v, w, P) to one variable P. Once the pressure is found, one
can calculate the velocities. The main issue now reduces to solving Equation (6.149).
dn
Air
Polymer
dp_ = 0
dn
Figure 6.29: Boundary conditions on pressure for the domain containing the polymer sus-
pension.
To solve these problems for isothermal and Newtonian cases in any arbitrary two-
dimensional geometry, one has to resort to numerical methods. There are many commercial
and research software packages available that implement the model discussed above. The
most commonly used is called C-Mold [266]. However, we will present an example of injec-
tion molding of a center gated disk under isothermal and Newtonian conditions in which a
closed form solution is possible.
Example 6.5: Mold Filling Time for Center Gated Disk [3]
Consider injection molding of a fiber suspension into a radial mold to make disks of radius
R and thickness 2h. The gate is located in the center of the disk and the suspension is
injected under constant pressure (see Figure 6.30).
a) Find the time to fill the mold under the assumption that the suspension behaves as
a Newtonian viscous fluid.
b) What will be the fill time under a constant flow rate injection at Q? Under these
conditions, what will be the maximum pressure?
Solution
a) We need to make the following assumptions to simplify the solution to the problem:
• h«R.
• Entrance effects are negligible.
• Fiber orientation influence on viscosity is negligible.
• Quasi-steady state.
• Inertia effects are negligible.
• uz and UQ = 0.
Thus, mass conservation dictates
;!<"*> = » («M)
Therefore,
C(z,t)
3.151)
Equation of motion in the r direction for a Newtonian fluid, when h « R (only shear
stresses being important) reduces to
dP
(6.152)
Integrating C twice using the no slip boundary conditions at the mold walls, which translate
into
C = 0 at z = h (6.154)
dC/dz = 0 at z = 0 (6.155)
Substitution of Equations (6.151) and (6.156) into Equation (6.159) and its integration
results in
Thus,
A = -Rf-—l-(-2). (6.161)
One can integrate Rf with respect to t to find how the flow front moves with time to
fill the mold. When Rf = R, that value of time will give us the fill time.
If one defines
r* = ^ (6.163)
*9 i / *\ 1 / *9 ,. \ ^i-L £
^>*2 (6.166)
t/iu = ^ ^ (6.168)
Thus,
3D// Crt
(6.171)
As seen from Equation (6.171), the maximum pressure PQ will be when t = tfm.
See [267] to include the influence of fiber orientation on the flow velocity and the con-
ditions under which it is important. •
F
dz 1 + (ir
\dzjl + ?• (6-172)
cure kinetics
viscous dissipation
=0 (6.173)
ux \ UJL, / uy \ uy /
where
S= \ — dz. (6.174)
Jo rj
S is a measure of the fluidity, with b denoting the cavity half thickness in the z direction
and 77(7, T,P) denoting the shear viscosity of the molten polymer. For example, in terms
of the power-law model
r? = m(T, P) -y"-1 (6.175)
where
( r ) exp(bP) (6.176)
it follows that
S=I(—^y)^ (6.177)
where
A =
and
rh z ,
X = / z(-^-)ndz (6.179)
•A) g(T)
with
m(T, P) = m0g(T)ebp (6.180)
and
g(T) = exp(^ - p). (6.181)
Note that if the power law index was n — 1/4, the fluidity index as expressed by Equation
(6.177) will be proportional to the cube of the pressure gradient, which will make Equation
x 2h
Solution
As the viscosity, r\ is assumed to be not changing significantly with temperature, one can
first solve the velocity distribution by using the equation of motion alone. Secondly, the
calculated velocity distribution is used in the energy equation to solve the temperature
profile.
Together with the assumptions made above, we will assume no-slip boundary conditions
at the mold walls. These will lead to a Poiseuille flow. However, the flow rate Q is specified
instead of the usual pressure drop AP along the channel of a length L. The equation of
motion for this one-dimensional, steady-state flow is
dP du2
0 =- (6.182)
dx
For this fully developed flow, the pressure gradient in x direction is constant and can be
written as dP/dx — dP/dx — AP/L — (Pe — Pi)/L, where Pe and Pi are the exit and inlet
pressures and L is the length of the channel. One might object to this by saying that the
values of Pe, Pi and L were not specified in the problem. What we will do here is that we
will first solve for u(z) in terms of AP/L, and then calculate the corresponding flow rate
Q taking the integral of u(z) across the thickness per width W. Then, we will rewrite u in
Symmetry of the profile requires that du/dz = 0 at z — 0 hence c\ = 0. The no-slip bound-
ary condition is u = 0 at the walls z = ±h. This can be satisfied if c% = — APft- 2 /(2r/L).
Hence the velocity profile is given as
z2
~h2}' (6 185)
'
= 2W
o
(6.186)
(6 18T)
M J' '
'i. (6.188)
(6.189)
In order to solve the temperature profile, the energy equation will be used
dT 82T fdu\2
u7r = k--^ + r)( — ) . (6.190)
ox ozz \oz J
Assuming u(dT/dx) is negligible compared to the two other terms, Equation (6.190) can
be simplified to
(6.191)
dz2 k \dz '
d=J (6192)
~J^'
^dz
(6.194)
One can write A = BQ2 where B = r//(fcTiVF 2 /i 2 ) to easily see the effect of Q on 9 and
hence T while plotting the mid-plane temperature versus flow rate.
One can evaluate du*/dz* = —3z*/2 from Equation (6.189) and then substituted into
Equation (6.194) to have
6 195
^=-¥<-•>'• <- '
By integrating both sides of Equation (6.195) twice with respect to z*, one gets
The symmetry of the temperature profile about the mid-plane can be mathematically ex-
pressed as 39/dz* = 0 This requires 03 = 0. The other constant c4 is evaluated by using
the boundary condition 9 = (T\ — T\)JT\ — 0 at z* = ±1. Hence, the resulting profile is
given by
. . . O _/i
as a function of Q.
Equation (6.198) can be rewritten as follows:
I
CD T-i
0 Flow rate, Q
Figure 6.32: Mid-plane temperature, Tmjdpiane = T(z — 0), of thermoplastic material versus
flow rate Q. The material is injected at a flow rate of Q and the upper and lower mold
walls are kept at a constant temperature T\.
Solution
We will consider the isothermal case for simplicity. The suspension is injected through the
center gate at a known flow rate. It makes perfect sense to choose a cylindrical coordinate
system with its origin at the center of the disk, and z = ±b are the surfaces of the disk. As
the flow would be radial, under the assumption of creeping flow and fully developed flow,
the only nonzero component of the velocity will be vr. Except near the entrance and at the
flow fronts, the velocity can be written as
C
vr = — (6.200)
r
where C depends on the flow rate. The power-law index n is represented as l/s.
The rate of deformation tensor for this flow field has the form
2e0 0 70
0 -2e0 0 . (6.201)
70 0 0
Thus, the flow is a shearing/stretching flow. There is shearing of the flow along the walls at
±b due to the no-slip boundary condition along the walls of the mold. Along the midplane,
away from the walls (close to z = 0) the flow will stretch or elongate in the tangential (9)
direction due to the radial nature of the flow. If 7 is the shear rate, it can be easily found
by using equation 70 = dur/dz + duz/dr from Table 3.5,
7o = - (6.202)
r
If we label the stretch or elongation rate by e'o, one can use the definition of elongation rate
from Chapter 3 to find the stretch rate
— z 3+ 1 (6.203)
Stretching will align the fibers in the transverse direction, while shearing will align them in
the flow direction. The ratio
(6.204)
V
'
The objective is to determine the velocity field and fiber orientation of flow of fiber
suspensions not only through the thickness but also in the plane. There are two approaches
that can be followed depending on whether the flow and orientation fields are coupled or de-
coupled. If one considers a coupled approach, one cannot use the Hele-Shaw approximation
and hence Equation (6.173) cannot be used to find the pressure and derive the velocity field.
This is because in the coupled approach the shear and normal stresses are also functions of
fiber orientation.
The general form of the constitutive equation to describe the stress tensor as a function
of the strain rate tensor and the fiber orientation tensor is given according to [265] as:
Tij = r/s7ij + r]scjkiaijki + B py^a^- + %•%•] + C%- + 2FaijDr (6.205)
where 7^ = (duj/dxi) + (dui/dxj) is the rate of deformation tensor, r]s is the viscosity of
the solvent, c is the volume fraction of particles, A, B, C, F are material constants, Dr is the
rotary diffusivity due to Brownian motion, a^ and a^ki are second and fourth order tensors
describing the orientation state of the fibers and defined as: a^ = (piPj) = f PiPj'lP(p) dp,
and dijki = (piPjPkPl) — I PiPjPkPli)(p) dp, where ijj(p) is a probability distribution function
P(9i < 0 < 9i + dO, </>i < 4> <</>! + d(j)} sin Q\ dO d<j), p is a unit vector aligned with the axis
of symmetry of the fiber as shown in the Figure 6.34. More details are given in Section 4.4.
The rotary diffusivity term can be neglected [265] when the fibers are slender (which is
usually the case in injection molding) Equation (6.205) can be written in the form
T
ij = rii [jij + Np%iaijki + Ns {jikakj + aijjkj}] (6.206)
where Np is a dimensionless parameter, called the particle number, and ??/ contains isotropic
contributions to viscosity from the solvent and the fibers.
In addition, one needs a constitutive equation to predict and calculate fiber orientation
from flow. According to [119], the equation of change for the second order orientation tensor
a,ij can be written as:
where u>kj = (duj/dxk] - (duk/dxj) is the vorticity tensor, 7 = ^jij'jji is the scalar
magnitude of the rate of strain tensor. C/ is interaction coefficient, measuring the intensity
Figure 6.34: Definitions of 0, <j>, and p t o denote the direction of a short fiber.
of interactions between the fibers in the suspension. In order to solve Equation (6.207) a
closure approximation of the fourth order tensor aijki in terms of the second order tensor
dij is necessary. There are many such approximations available [270, 271]. Table 6.2 lists a
few.
Table 6.2: Closure approximations for a^ki in terms of the second order tensor
1
Linear closure
(Hand [272])
7
Quadratic closure dijkl — dijdkl
(Hindi and Leal [273])
1 - f
Hybrid closure avkl ^ V*< 1 • * , < 1 ^
(Advani and Tucker [274])
7
Under certain assumptions about the fibers and the flow, one can formulate the com-
plete set of governing equations to solve flow and fiber orientation simultaneously. We will
describe the procedure here for an isothermal case. For a nonisothermal case, one must also
include the energy equation such as Equation (6.172) in which the viscous dissipation term
is expressed as T : VU instead of just the velocity gradients multiplied by the viscosity.
Other assumptions that are reasonable are that fibers are considered rigid axisymmetric
particles that are smaller than any length scale of the flow field, the fibers are large enough
to neglect Brownian motion, and inertia and body forces are ignored due to the highly
viscous nature of the suspension. Thus using the continuity equation,
V-U = 0 (6.208)
and the equation of motion with inertia and body forces neglected
0 = -VP + V - r (6.209)
and combining the above two physical laws with the constitutive law for the stress and
fiber orientation, one has to solve the complete set of equations expressed by Equations
(6.206)-(6.209).
Suspension
Flow Field Flow Field
Rheology
1t
Mechanical Mechanical
Properties Properties
(a) (b)
Figure 6.35: The approach to solve (a) decoupled and (b) coupled orientation field.
What does one do with the fiber orientation information? This information is quite
useful in prediction of thermal or mechanical properties of a short fiber composite. For
example, one can write the stiffness tensor in terms of the orientation tensors. For details,
see [119].
6.5.3 Problems
1. The approximated governing equation for the extrusion process was found to be
AP
L^J. ff-ii
\J Uiy ff"ii z
\J Ujy
= £. _| -
2
/j,L dx dy2
Advanced Thermoplastic
Composite Manufacturing
Processes
7.1 Introduction
To predict how a material will behave under deformation applied in any manufacturing
operation will depend on the approach selected to describe its macroscopic behavior. For
most thermoplastic materials containing continuous fibers, fiber preforms, or collimated
fibers (long discontinuous fibers in an oriented fiber assembly), one can treat them as a
homogeneous material with anisotropic viscosity or transversely isotropic viscosity with
inextensibility in the fiber direction. However, at the microscopic level there are many phe-
nomena taking place that are conglomerated into constitutive constants at the macroscopic
level to allow for prediction of the deformation rate relationship with the applied force.
The microscopic phenomena that influence the constitutive equations at the macrolevel
are resin percolation, interply (in between the plies) and intraply (within a ply) slip and
shearing. Resin percolation and redistribution are functions of the fiber volume fraction and
the load applied, whereas the shearing within the ply or in between the plies is limited to the
resin rich areas. Experimental studies [275, 111, 276] suggest that this flow of the resin in
between the fibers is a truly viscous flow phenomenon and can be modeled in that fashion.
If the process time scales are large and viscosity is low as in the case of thermoset resins,
shaping operations can squeeze excess resin out. However, as the resin viscosity is high
and usually the process time scales are low for thermoplastics, if one does not understand
the deformation behavior the reinforcing fibers will not be able to adjust to the flow and
buckling and distortion may result. The resin flow is schematically illustrated in Figure 7.1
within a unidirectional fiber bed and fabric tows of a mat, hence the need to understand
the behavior of such materials.
Another change that the thermoplastic reinforced material undergoes is the level of
crystallinity. The rate at which the material is cooled to below the glass transition temper-
ature is the primary process variable, which governs the crystallization in the matrix. The
presence of fibers also provides nucleation sites and can cause secondary crystallization.
In addition to influencing the crystallinity of the matrix, cooling can also cause internal
stresses that can lead to microcracking. The internal stresses are caused due to four major
factors: the mismatch of the coefficient of thermal expansion between the resin and the
fibers, temperature gradients between the laminates, the orientation of the fibers and the
shape of the part to be manufactured. Usually as the composite is cooled from its processing
temperature below its glass transition temperature, the resin tries to shrink which is resisted
by the stiff inextensible fibers. This places the fibers in compression and the resin in tension.
Similar logic applies to the composite. The composite is cooled from outside to the center.
This would shrink and solidify the outside first, and then the interior is constrained during
its solidification. This puts the outside under compression and the interior under tension.
Hence controlled cooling is an important process parameter during manufacturing of such
materials.
In summary, the inextensibility of the fibers coupled with the high viscosity of the resin
limits the deformation processes that one can attempt for manufacturing operations to shape
these materials into useful end products. The challenges are to be able to model this material
and to ensure that stress build up during cooling is minimum. We will consider modeling
of three thermoplastic composites processing methods in this chapter that underpin the
technology of advanced thermoplastics composites manufacturing. They are sheet forming,
tape layup (also known as fiber tow placement) and thermoplastic pultrusion.
in a polymer sheet and stacked together to form the laminate that is then deformed to
contour to the shape of the tool or the mold. Long discontinuous fiber (LDF) material was
introduced to accommodate the stretching mechanism, as the fibers were long and aligned
but the discontinuous nature could create room for stretching.
There are four main composite forming processes: diaphragm forming, matched die
forming, stretch forming and roll forming. Many details of these processes are given in
chapters on sheet forming by O'Bradaigh in [277]. and Bhattacharrya [278]. Here we will
re-iterate the important features of these processes in terms of the deformation mechanisms.
Lower Diaphragm
Figure 7.3: Schematic of diaphragm forming process of thermoplastic composite sheets [42].
PLATEN
Mold
Heaters Clamping
Ring
Laminate
Insulation
Mold
mmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmmt
Heating Oven
t t t t t t
Figure 7.5: Consolidation of prepregs.
Roll forming uses a series of rotating rollers to deform a composite sheet into a curvilinear
shape as shown in Figure 7.6. This forming process is different from the others as the
deformation is completed in a series of steps using a series of rollers. Again one must
develop a model to describe the deformation of materials between two rollers to be able to
predict the deformation.
For more details on stretch and roll forming, refer to [9].
(D O
Figure 7.7: Schematic of flow of resin from resin-rich areas to resin-starved regions.
In processes such as matched die, there is also transverse flow in which the fibers and
resin move as a unit together when squeezed as shown in Figure 7.8. One can characterize
p
1111 p
O
0 0 0 5°°0°5go?>
O
11 1111
°Jo°uu
0°0°oO°°°'
Q o o oo %°°d- o° o uoob^u odo o oo o"
the rheology of squeezing flow of such materials as a single entity of a viscous material
in order to model their behavior. When forming single curvature parts, in addition to
transverse flow and resin redistribution, one also needs the individual plies to slip over
one another in the laminate to form the part as the stretching mode along the fibers is
unavailable due to the inextensibility of the fibers. Figures 7.9 and 7.10 show the phenomena
of interply slip. Here, one would like to know the force required to pull a ply out from a
stack of plies that have been consolidated. Rheological studies to characterize interply slip
are usually conducted by flexural testing of heated laminates or by ply pull out experiments
at elevated temperatures. Finally, if one is interested in forming double curvature parts, in
Figure 7.10: Interply deformation: (a) buckling of inner fibers and (b) interply slip [277].
Thus at the microscopic level within each sheet of the thermoplastic material with
unidirectional fibers, the fibers have to move past one another to accommodate a complex
curvature. The fibers can shear past one another. As each fiber is covered with a layer of
resin, one can estimate the force needed for this activity. In addition, the fibers can also
move in the transverse direction. These mechanisms are demonstrated in Figures 7.11(a)
and (b).
(7.1)
dt
Solution
(a) Integrating Equation (7.2) and using the fact that L — L0 at t = 0 gives
(7.5)
(7.6)
7
7.3 Pultrusion
Pultrusion is a continuous composites manufacturing process. The goal in pultrusion is
to produce a consolidated profile with a constant cross section by pulling impregnated
reinforced material through a heated forming die. The schematic for the process is shown
in Figure 7.13.
In this process, unidirectional preimpregnated tows, tapes or fiber rovings are guided
through a preheater where their temperature is raised to the melting temperature of the
polymer. Next, the heated polymer and fibers enter a die assembly. A die assembly consists
of a series of dies which consolidate, form and cool the material. Usually, a significant taper
is used in the entrance die to help consolidate the polymer with the fibers. Downstream,
the cooling dies cool the formed product and influences the properties such as crystallinity.
Also, faster cooling can allow one to locate the pulling mechanism closer to the die assembly.
The pulling speeds can vary from a few millimeters per minute to a few meters per minute.
Fiber
Resin
v
R(z)
Notice that so far we have not utilized the continuity equation. We know that the flow
rate of the resin has to be constant throughout the taper in order to satisfy the continuity
equation. The flow rate at the end of the taper and in the uniform section of the die, where
the velocity is uniform can be expressed as
(7.15)
where Re is the radius of the annulus at the end of the taper. One should expect that at
any axial position along z, the flow rate should be the same. Note that, v\ = irR^f/itR^,
and the radius of the annulus at the inlet, Ri is related to initial fiber volume fraction of
irR2,
V
° = ^R*-
At any cross section along z, the flow rate Q is the sum of plug flow induced by fiber
drag and backflow generated due to the opposing pressure gradient as shown in Figure 7.16.
As the resin is Newtonian, one can superimpose these flows
The drag flow is uniform due to the slip boundary condition at the end of the annulus. This
slip boundary condition is the same as symmetry or no flow out of the cell boundary. Thus
(7.17)
and hence
Qbackflow = Q ~ Qdrag = UlT -R2(z (7.18)
at any cross section along the axial direction, z.
One can also find the total flow rate at any cross section by integrating Equation (7.12)
to find the velocity profile, uz, and relating it to the flow rate as shown below:
27T fR(z)
Q= o JR
uzr dr d0. (7.19)
By equating Equation (7.16) and Equation (7.19), one can find details of the flow pattern
and pressure gradient, once we have a rheological form that describes the viscosity T] of the
thermoplastic material.
Let us assume the fluid is Newtonian and that the viscosity does not vary appreciably
with temperature. One can now integrate Equation (7.12) to obtain the velocity profile
which is
f
4// dz [ \ r )\
Note here that uz varies with r which is between R and Rf. Substituting Equation (7.20)
into Equation (7.19) and integrating it results in
drag flow
Qbackflow = V
TT_dP_
8/j, dz
Inverting Equation (7.22) gives the pressure gradient along the tapered section of the die
(7.23)
£
- 4 ( £ M ~ + (S r + 41n
One can integrate equation (7.23) with respect to z to find the pressure profile along
the z direction. This pressure profile will be a function of the viscosity of the material, the
pulling speed, the taper of the die and the final fiber volume fraction desired.
Solution:
Given,
V0 = 0.40
Vl = 0.50
Rf = 0.020 m
L = 0.2m
M = 1000 Pa.s
10 6 Pa
(7.24)
(7.25)
(7.26)
Substitute Equations (7.24), (7.25) and (7.26) into Equation (7.23), and integrate with
respect to z to obtain the pressure profile
- z,
P(L)-P(0) = Pmax
a
= L
L dz'. (7.28)
Substituting Pmax and L into Equation 7.28 and inverting the equation, U is found. One
can use a mathematical solver program such as Maple [282] or Matlab [283] to calculate
the integral and solve for U as shown in Figure 7.17. The critical value of U is found to be
vO «=• 0.40:
vl = 0.50:
Rf = 0.020:
:L ~ 0,2:
- 1000:
Pin ax •"• 1.0E6:
Rin
8w~. 03 162277660
> Rex = sqrt(Rf A 2/vl) ;
&j;t = .02828427125
> ratio = Rf/Rex;
miio =.7071067811
= ( (Rex-RIn) /I.,) *z+Rin;
U) =......01669252675Z r .03162277660
> u = Pmax /
lot < <8*mu*pi*(R<S3[:-"-2- < R < B ) } '"'2} ) /<p.i.
*(3-4*ratio A 2+ratio A 4+4*ln{ratio) )
z-0. . L ) ;
U~. 7480135211
r>
0.75 m/s. If one pulls the rovings with this critical speed, then the pressure at the exit of
the die will be 1 MPa.
The pressure profile along the z axis is plotted in Figure 7.18. •
= 7«7 (7.29)
ro
Q_
z[m]
Figure 7.18: Pressure profile along the axial z direction in the tapering die during the
pultrusion process. The rovings are pulled with 0.75 m/s.
where m and n are curve fitting parameters and 7 is the magnitude to the strain rate tensor.
One can recover the Newtonian fluid model by choosing m = /j, and n = 1. One can start
with Equation (7.10), and integrate with respect to r using the boundary condition that at
r — R, the shear stress is zero. This results in the shear stress profile
IdP
r— (7.30)
2~dz
For the power-law fluid model,
duz
—r- = —m (7.31)
dr dr V dr
The combination of Equations (7.30) and (7.31) results in
,1/n
duz _ R2
(7.32)
dr 2m dz
One must integrate uz with respect to r to find the velocity profile. Once the velocity profile
is known using the conservation of flow rate equation, one can obtain the expression for
dP/dz along the fiber direction.
Example 7.3: Derive the Pressure Profile for a Shear Thinning Thermoplastic
Composite Being Pultruded
Find the expression for dP/dz for a composite being pultruded through a die. The ther-
moplastic matrix in the composite exhibits shear thinning behavior which can be modeled
Solution (a):
Substituting n = 1/2 in Equation (7.32) and integrating with the boundary conditions listed
in Equations (7.13) and (7.14) gives
R L o RfL 1 RfL
uz = — U (7.33)
\2m dz J Rf+ R 3 R3
if X = R/Rf, then
/ 1 dP\*
= — -- (7.34)
dz
1 dP\' 8 J_ 3A __ A 3 A^ N
= -2?r (7.35)
2m dz J ~5 + 2A + T " Y + To
Thus
Solution (b):
For n = 1/3,
-(——V
\2m dz J 5ji-F+i*V l+l7 - ' (7.37)
Thus
Qbackflow -
Again, one must invert Equation (7.38) to be able to integrate the pressure as a function
of z. Figure 7.19 illustrates the effect of power-law exponent on the pressure profile using
dimensionless variables. •
^ ^RdJl(*nRd)
n=l
+ ——
4fc
/
'-
r6 JJQ\
n(\
A r}dr\
n')(J'l\
"vn> - ^z/u
Ae a (7
V-^460\
)
T2M D
Vf (7.47)
k
(pcp)c = (l-V})pmcm + V f p f c f (7.48)
where the subscripts m and / represent matrix and fiber materials, respectively. The
particular solution that will satisfy boundary condition in Equation (7.45b) and (7.45c) is
Ts(r) = TW + - L - Rl - r 2 . (7.49)
One needs to find the average heat generated per unit volume of the cell to be used in the
energy equation. This is done by using average shearing of uz in the radial direction
« , „ „ , , , „ , rdrdO
(IT /I average
*-"
where R refers to the local radius of the cell model. Thus, dissipation will not be constant
along the taper and will vary as R changes along z.
For Newtonian fluids
(7.53)
where A = Rf/R. Thus the viscous dissipation term varies with the square of the pressure
gradient and the square of R. If one substitutes Equation (7.23) for dP/dz in Equation
(7.53), one could plot how viscous dissipation changes with R(z).
This viscous dissipation can then be included in Equation (7.44) and solved to evaluate
the impact of the q term on the temperature profile. There is no closed form solution for
this, but one could easily develop a finite difference solution for this problem.
Solution
If we use the following nondimensional variables,
ctrrz . ,
(7.54a)
f= (7 54c)
JTd '
<fc = If (7.54d)
and choose
7? TJ
Pe = Peclet number = - (7.55a)
arr
riU2
Br — Brinkman number = - . (7.55b)
we get
Usually Pe « 102 to 103; hence one can easily discard |^ -p- ^^ term. The last term on
the right-hand side of Equation (7.56) can make a significant contribution even for small
values of Br number as there are many fibers and hence many cells inside the tapering die
directly adding to the heat generation. •
Prepreg spool
&
Tensioner
.«#*• Pre-heater
Consolidation roller
Tool
Figure 7.20: A thermoplastic fiber-reinforced tape being consolidated to the substrate layer
with heating mechanism and consolidation roller.
As shown in Figure 7.20, there are five main components in this process. They are
the preheater, tensioner, nip(contact)-point heater, consolidation roller and a tool for the
placement of the fiber tows to form the composite.
The tensioner is used to mechanically maintain constant tension within the fiber tow
from the moment it leaves the supply spool until it is wound on the tool or the previously
placed fiber tows on the tool.
The tool, which is known as a mandrel in filament winding and as a tape-laying head in
the tow placement manufacturing method respectively, is the surface on which the fiber tows
are placed and wound to create the geometric form of the composite structure. In the case
of filament winding, it is cylindrical in form, whereas in tow placement it can take various
complex forms based on the geometry of the part to be manufactured. Usually metals
such as stainless steel or aluminum are used to form the tool because of their durability
and resistance to oxidation. The tool is usually heated to avoid creation of large thermal
gradients between the heated tow and the mandrel. The large thermal mass of the tool
can act as a heat sink, lower the matrix temperature of incoming tow and limit the inter-
laminar flow of the molten polymer. Therefore, a heated tool is necessary to retain heat in
Find the heater power required to heat a tow from 30°C to 130°C. Assume that 20% of the
heater energy is lost to surroundings.
Solution
Qt = mcPt(Tp-Tr]
= PtVtwtttcPt(Tp - Tr) (7.57)
where
The requirements of the nip- or contact-point heater are the same as the preheater
except the nip-point heaters are focused only at the point of contact of the incoming tow
and the previously placed tows. The nip-point heater should induce rapid heating to bring
the contact area to above the melting point of the resin instantly as the contact time is
usually very short. A variety of heaters can be used for this purpose. Some examples are
open torch, laser, hot gas, infrared and heated rollers, or hot shoes.
Compaction Force
Machine Limit
Processing Speed
Required Porosity
Porosity
Maximum Speed
Figure 7.21: Obtaining the highest processing speed from a typical porosity-speed-force sur-
face, taking into account machine limits and quality requirements for on-line consolidation
force the machine is capable of providing may be used as the limiting condition. Section A
is taken at this level, and its projection on the processing speed-porosity plane is obtained.
The maximum acceptable porosity can now be used to determine the maximum processing
speed by following the line BCD. The line BC runs parallel to the processing speed axis
at the required porosity value until it intersects the projection of section A. The line CD
runs parallel to the porosity axis until it intersects the processing speed axis. The point
D represents the highest processing speed that can be used to obtain a part of acceptable
Compaction
Roller
Figure 7.22: Location of the coordinate axes for the consolidation process [284].
may be modeled as shown in Figure 7.23. A tow of a given height, hi, and width, Wi, enters
the region under the compaction roller. The coordinate system is defined such that x is
the width direction, y is the direction of tow movement and z is the thickness direction.
The tow speed, V, is specified. As the tow passes under the compaction roller, its height is
reduced to hf and its width changes to Wf. The consolidation force compresses the voids
and changes the height and width of the tow. Since the tow dimension in the y direction is
much larger than in the x and z dimensions, the y direction flow may be neglected. Due to
the high matrix resin viscosity, the flow may be treated as creeping motion and the inertial
effects can be neglected.
(a)
(c)
Figure 7.23: Schematic of the various views of the consolidation squeeze flow model with
the input variables [284].
o
dp d
-7^(pvz} = 0 (7.59a)
d~t
dP __d_
^~ (7.59b)
dx dx dx dz
dP d dvx dv d dvz (7.59c)
^~
dz "H~
dx ^~
dz ~5~
dz
The density and viscosity of the continuum are given by p and /i, the pressure in the
material by P, and the velocity by v. In actual tows, the thickness (0.006") is much smaller
than the width (0.25"). Therefore, the dominant velocity gradient term in the momentum
equations is the shear strain rate through the gap, dvx/dz and its derivative with respect
to the thickness direction, z. Therefore, the momentum equations, Equations (7.59b) and
(7.59c), may be simplified as,
dP_ d_
(7.60a)
dx ~dz dz
dP
(7.60b)
Equation (7.61) may be integrated in the thickness direction from 0 to h, the instantaneous
thickness of the domain.
h
d + / \ir(pVx) dz+ v
p( ^=o = °- (7-62)
The velocity vz at z = 0 is 0 and at z = h is h, the closing speed between the two
bounding surfaces. Substituting these values and interchanging the order of integration
and differentiation in Equation (7.62) results in
.„ , .„ . vxdz + ph — 0. (7.63)
at ox \ Jo )
Equation (7.60a) may be integrated with respect to z to yield,
„[£]= ,£ + Cl(l) . (7 . 64 )
Laz j ax
Integrating Equation (7.64) once more with respect to z,
,~x dP fz £ _ „ , . fz I ,.
(7.65)
ax Jo fi Jo ^
where £ is a dummy variable of integration. The value for vx from Equation (7.65) may be
substituted into Equation (7.63) to eliminate velocity as a variable and maintain pressure
as the only variable to be solved
The values of C\(x) and vx(0) may be obtained from two velocity boundary conditions.
Since we have a second order equation for pressure, two pressure boundary conditions are
also needed.
Let p* be the nondimensionalized density of the continuum defined as,
p* = ^ (7.67)
Pf
where p is the density of the continuum and pf is the density of the composite with no
voids. Equation (7.66) can be rewritten as,
1^7- ( * I fh
9 \P
W,+ -z- Z t
^> JA * /~i t —\
fZ
I
1
_7A I _7 _ 1 i _*7 n /y rjo\
at ax y Jo o fJ. Jo M ' '
0 0
° o&00ocb°o.0d00o
O
U O- OO o^\ Or\ r\°
O O oPOD °0 cO o
The initial radii of the void and the resin shell are R0 and S0, respectively. Pg and Pf are
the gas and resin pressure inside and outside the shell, respectively. The resin is assumed
to be incompressible; therefore
= Jr-3 n3
0 - ri0. (7.70)
(7.72)
Ro
R* = — (7.73)
RO
Equation (7.69) may now be rewritten using Equations (7.70) and (7.71) as,
dR* Pgo T
-1 =0 (7.74)
~dT T0
Mean void
fraction
O O
Axis of symmetry
Figure 7.25: Microscopic void compression model used to determine the bubble radius as a
function of the local pressure and temperature.
Solution
Initially, the composite material occupies a hollow spherical volume between r = R0 and
r = S0. The mass of the composite material excluding the void in it, m/ is given as
( 7 - 76 )
where pf is the density of the composite material excluding the void in it. At any time, the
mass of the cell (composite material including the void) is given as
m = m/ + mvoid = p — S3 (7.77)
o
where p is the density of the composite material including the void and S is the radius of
the cell at that time. If one neglects mvojd relative to m/, then
(7.78)
which reduces to
If one takes the time derivative of both sides of Equation (7.79), the following is obtained
dp* _ 8
~dt ~ ~di
= -3^ f. (7.80)
Time derivative of Equation (7.70) allows one to express dS/dt in terms of dR/dt
CL . _.o.. CL , ™Q T-.Q -r-.Q\
dt dt
2
dS _ R dR
~dt ~ 52 dr
= 0 0 dR
6
dt S dt
dR
dt
{
"°\wj '"'u,;
(S*3-1 + R*3}2 dt ' (
'
which is Equation (7.75). •
The ratio ETwau//i (j3 in Equation (6.45)) determines the type of boundary condition
that is imposed. If one expects the friction to be very high, i.e., Kwau//j, —> oo, it is clear
that wwaii must be zero such that the left hand side is finite. This is equivalent to a no-
slip boundary condition. On the other hand, if the friction is expected to be very low,
i.e., Kwaiif/j, —> 0, the normal derivative of the velocity goes to zero at the wall. This is
equivalent to a perfect slip boundary condition. Therefore, this friction factor boundary
condition may be used to impose all types of velocity boundary conditions ranging from
no-slip to perfect slip.
\ Ox / constrained edge = °
(TT)
While considering the tow that is being laid down, we apply the free surface bound-
ary condition because the unconsolidated tow width is smaller than the consolidated tow
width. However, for a tow that was placed previously, one applies the constrained boundary
condition because it has tows along both edges of the width.
Example 7.7: Characterization of the Viscosity of AS4 Fibers with PEKK Resin
A constant load with a heating unit attached is used to characterize the viscosity of AS4
fibers with PEKK resin. The specimen, which is a ten layered consolidated AS4/PEKK
laminate, was placed inside an oven using a dead weight as the source of the constant load.
The composite and the dead weight were held at the desired temperature for 20 minutes.
The dead weight was then placed on the specimen for a predetermined duration of time.
The dimensions of the specimen before and after the dead weight being placed on it were
measured to estimate the viscosity of the composite at different temperatures. Table 7.1
shows the results from these experiments. Find the temperature dependence of the viscosity
of this system.
Solution
The change in dimensions of the laminate can be related to its viscosity assuming that the
composite behaves as an isothermal incompressible Newtonian fluid. Thus,
5 Ft
(7.86)
8 LiW? tf -
Hence, at each temperature, one can find /j>.
It is clear that the average viscosity of the composite decreases with an increase in
temperature. The viscosity of AS4/PEKK can be approximated by the following functional
form (see Chapter 4 for details):
Pa.s (7.87)
where T is the temperature in K. By plotting viscosity values at 4 different temperatures, a
best-fit line was found. Note that viscosity is plotted on a log scale due to the exponential
nature of Equation 7.87. The values of A and B that best-fit the experimental data were
1.28E-4 and 16548.6. The best fit line and the experimental data are shown in Figure 7.26.
to
Q.
10
620 625 630 635 640 645 650 655 660 665
Temperature (K)
Isothermal Incompressible
Consider an isothermal incompressible Newtonian fluid with the following boundary condi-
tions. The friction factors at z = 0 and z — h are denoted by Kb and Kt in Equation 7.83.
The susbcripts b and t characterize the bottom and top, respectively. The pressure at x — w
is zero and the pressure gradient dP/dx = 0 at x = 0. Equation (7.65) can be used to solve
for the velocity profile.
u.oo
0.5 Kb= 1000 -
0.45
,,,<..,---•••••'"""""'"
c 0.4
o
ts
2 0.35 .,-•''".--•'''"
X
0.3 /('•••''' K t =1000
(D
±=
^C 0.25 /.••• 0.1
/•••' 0.001
0.2
8 /••'
<D
->
0 15 // _.— -.,
0.1 " / ^^^ "~~~~^^\"""""---
0.05
n
/^^ ^X;
0.2 0.4 0.6 0.8
Non-dimensional height through the gap
Figure 7.27: The velocity profile changes from a symmetric parabolic to an asymmetric
parabolic shape as the top surface friction factor goes to zero with the bottom surface surface
friction factor maintained constant at a high value (which models the no-slip boundary
condition).
for a given value of K£. It can be seen that the velocity profile changes from a symmetric
parabolic profile for the case of no-slip at both walls to an asymmetric profile when no-slip
is imposed on one wall and perfect slip is allowed at the other wall. Equation (7.68) can
now be used to determine the pressure distribution in the domain.
-i
p— (7.92)
7?
Figure (7.28) shows the pressure distribution in the domain for various friction factors. The
pressure distribution is seen to be parabolic in nature independent of the type of velocity
boundary condition that is applied. However, the pressure drop from the mid-plane to the
outer edge is the lowest when a perfect slip condition is imposed on the top surface and
the highest when a no-slip condition is imposed on the top surface, given that the bottom
surface has a no-slip condition prescribed.
The change in width or the movement of the free surface may be tracked as a through-
thickness average based on the velocity in the x direction at the free surface.
1.4 Kb=1000
1.2
E 0.8
73
C
O
0.6
0.4
0.2
0
0.2 0.4 0.6 0.8
Non-dimensional width
Figure 7.28: The pressure profiles maintain a parabolic shape. The pressure drop from the
middle of the tow to the outer edge decreases with decreasing top surface friction factor.
dw 1
(7.93)
dt ~ h
Substituting for vx from Equation (7.91), it can be shown that,
dw h
—— = — w — (7.94)
dt h
In other words,
w h = constant (7.95)
This result is not surprising because we are dealing with an incompressible fluid and expect
the total volume to be conserved. This result is, in fact, a verification that we conserve
volume in this formulation.
Solution
The force required for compaction, Fc, for this incompressible material may be obtained by
integrating the pressure distribution across the width and along the contact length.
Substituting for h and the pressure solution from Equations (7.90) and (7.92) respectively
and carrying out the integrations we obtain
(7.97)
wf
The compaction ratio hf/hi is governed by a single nondimensional parameter, Yin, given
by the right-hand side of Equation (7.97).
.. -—
n-M
1J _15 A *L
„ o
1 r (7.98)
+
2,
-^"1/5 (7.99)
11^ varies between zero and oo depending on the parameters of the problem. For example,
for high compaction forces, low tow speeds or low material viscosity, the magnitude of H^ is
large and thus hf/hi is a small number. Figure 7.29 shows the variation of the compaction
ratio with the nondimensional parameter HJJ. •
DC
to
Q.
o
O
dP
- - (7.100)
at 2 ) (I + K: + K?/K?} e
The velocity field is still given by Equation (7.91). Equation (7.100) needs to be solved
numerically due to the dependence of the density on both time and position, x.
The solution methodology is shown in Figure 7.30. The initial density in the domain is
Macroscopic Microscopic
assumed to be uniform. Equation (7.100) is solved with the pressure boundary conditions
- 0 (7.101)
' x=0
P(x = w(t)) = 0 (7.102)
using a central difference scheme for the spatial derivative of P. The velocity in the x
direction is calculated from Equation (7.91). Based on the average velocity of the material
in the x direction, the free surface and the other nodes are advanced to the new positions.
The microscopic void compression equation, Equation (7.74), is solved using a Runge-Kutta
algorithm at each node in the domain to obtain the updated void radii at all locations. The
densities at these locations are updated, and the Equation (7.100) is solved again. This
procedure is repeated at every time step during the solution.
The model solution is illustrated for the case of a no-slip condition at the bottom surface,
z = 0, and a perfect slip condition at the top surface, z = h, in order to limit the number
of results to be presented. The initial void fraction is maintained at 0.05 (5%), and a tow
speed of 0.05 m s"1 was used. The final void fraction distribution was calculated for the
compaction ratio, /i///i;, of 0.99 to 0.95. A constant viscosity of 105 Pa.s was used for
the medium. The initial pressure inside the voids was assumed to be atmospheric and the
initial void radius was assumed to be 10 /im. The final void fraction distribution for these
cases is shown in Figure 7.31. The final void fraction is seen to be highest at the outer
edge of the tow because the pressure is the lowest at that point. As the compaction ratio
decreases, i.e., the amount of compaction increases, the final void fraction decreases, which
is consistent with our intuition.
0.045
"§
CO 0.044
0.043
0.042
0.041
0.04
0.2 0.4 0.6 0.8 1 1.2 1.4
Non-dimensional position along the width
Figure 7.31: Final void fraction distribution in an isothermal compressible medium under-
going consolidation.
Nonisothermal Compressible
The most general case considers the continuum to be compressible and the domain to be
nonisothermal. The viscosity of a continuum is a function of the temperature and therefore
varies throughout the domain.
The values of ^(O) and C\(x] in Equation (7.65) may be evaluated from the velocity
boundary conditions and the equation may be rewritten as
dP
(7.103)
(7.104)
-o 0 o o o o— Region 1
o o o o o o \
\
o o o o o o \
'v2 \ Region II
0 0 o o o o \
regions I and II. Region I denotes the topmost tow that is being placed. The void fraction
in this tow reduces from the incoming value, <f>vi to an intermediate value, </)V2. Region II
denotes the layers beneath the topmost layer that are still at a temperature > Tg. The void
fraction in this region reduces from 4>V2 to the final void fraction ^3.
The pressure in region II is uniform and the continuity equation simplifies to
ph = constant. (7.106)
The combined solution of the change in height of regions I and II needs to be carried
out iteratively. The rates of change of the height of regions I and II are denoted by hi and
hn respectively. These are related to the total change in height, h by
h = hj + hn (7.107)
h is known from the geometric parameters and the tow speed. An initial guess is used for hj
and (j)v2 and the pressure distribution in region I is computed and averaged over the width
of the tow. This average pressure is used as the driving pressure for the void compaction
in region II. Based on the void compression model the change in density of region II and
the associated change in height may be computed. The sum of the changes in height is
compared to the known total change and the guess for hj modified appropriately. This
g
ra
'o
Figure 7.33: Variation in the void fraction in regions I and II as a function of the non-
dimensional tow position under the compaction roller for a tow speed of 0.005 m/s.
Summary
A model for consolidation and void reduction during thermoplastic composites processing
was discussed. This work is based on the assumption that the composite can be modeled as
a compressible continuum. Thus the problem is solved as the squeeze flow of a compressible
viscous fluid. After deriving a general formulation, three specific cases — isothermal in-
compressible, isothermal compressible and nonisothermal compressible — were solved. The
final one is the most applicable model for the fiber tow placement manufacturing process.
However, for an extended model of the consolidation process, the domain can be split into
two regions, the topmost tow and the other tows beneath it that are at a temperature
greater than Tg. Given the total thickness change in both regions, an iterative scheme can
be used to estimate the individual changes in thickness in the two regions. Simultaneously
the changes in their void fraction and the required compaction pressure to achieve the given
thickness change can be computed.
This analysis demonstrates how one can start with a simple model and refine the model
by relaxing the assumptions to move towards accurately describing the physics. Comparison
with experiments can usually help determine if these refinements are necessary. Revisions
and refinements are usually more time consuming in terms of model development and solu-
tion.
3. The composite is cooled from the outside to the center. This would shrink and so-
lidify the first, and then the is constrained during its solidifica-
tion. This puts the outside under and the interior under Hence
controlled cooling is an important process parameter during manufacturing of such
materials.
4. In the diaphragm forming process, a laminate of the composite material is sandwiched
and heated between two sheets of and a is used to form the
composite. The edges of the material are not as the inextensibility of the
material can cause fiber
5. In the matched die forming process, the composite laminate or sheet is held and
between two metal tools surfaces as is accomplished in sheet metal forming.
The process takes very time. Most match die forming has been practiced
on shapes due to nonuniform and distributions.
6. The stretch forming process can be used only with materials that can
stretch. fiber thermoplastics cannot do this but materials can.
7. The roll forming process uses a series of rotating to deform a composite
into a shape.
8. When forming single curvature parts, individual plies over one another in
the laminate. The stretching mode along the fibers is unavailable due to the
of the fibers.
11. In thermoset pultrusion, one uses law to model resin impregnation, whereas
in thermoplastic pultrusion, due to the high of the material, this law is not
the correct model.
12. The heat transfer model in thermoset pultrusion involves On the other
hand, the heat transfer model in thermoplastic pultrusion focuses on
13. In pultrusion, if one assumes that most of the flow is parallel to the , one
can divide the section of the die into self similar cells. A cell model is
formulated by considering a single tapered annulus of liquid surrounding
each This assumption allows one to simplify the problem to
problem in the direction of the fiber movement. Also, due to the cylindrical nature
of the fiber, one can assume that the annulus of the resin is also ; thus one
can choose the coordinate system to pose the problem.
14. In pultrusion, the mechanism of the fibers in the axial direction induces
a flow in that direction. However due to the geometry, the
material experiences a pressure build-up. This makes the resin pressure higher near
the of the die as compared to the Hence, a pressure gradient is
established in the direction to drag flow, which may give rise to
flow.
15. In pultrusion, at any cross section along the axial direction of a tapered die, the flow
rate Q is the sum of flow induced by the fiber , and
generated due to the opposing
16. In the tapered die of the pultrusion process, one would expect heat generation by
18. There are five main components in on-line consolidation process. They are the
, , , , and a for
the placement of the fiber tows to form the composite.
19. The is used to mechanically maintain constant tension within the fiber tow
from the moment it leaves the until it is wound on the tool or the previously
placed fiber tows on the tool.
8.1 Introduction
Thermoset resins are very conducive to processing, as their viscosity is reasonably low to
allow easy impregnation of the empty spaces between the fiber preforms. The earliest
composites manufacturing concentrated on processes such as wet hand layup, which is an
open mold process. This process employed a one-sided male or female mold, in which a
layer of resin called the gel coat is applied on the mold surface. The gel coat is usually
about 0.5 mm in thickness, has good environmental resistance and produces a smooth and
pigmented surface that covers the fiber preform. To manufacture a composite, a measured
amount of the thermoset resin is poured over the gel-coated surface. The fiber preform is
placed on top of the liquid resin and pressed by a hand held roller. The resin percolates
through the fibrous network due to the application of the pressure as shown in Figure 8.1.
Laminate
Dry reinforcement \ Gel coat
After one layer of the fiber preform is satisfactorily impregnated and compacted, another
layer is placed on top and the procedure is repeated until the desired lamination of the
composite is complete. Although this process was one of the first techniques to be used in
manufacturing composites, it is still widely used by the boat manufacturing industry due
to its versatility and low capital investment. However, as the resin that is widely used in
this open molding process contains styrene (which is emitted to the environment), there is
a push to replace it with closed mold processes due to environmental concerns.
All the processes we will discuss in this chapter evolved from this process in which the
[ Recirculating Air
VoJifc,. (Forced Convection)
D
in i n t
.'
i i
."
I
."
I /
/
/ '
i\ ; . •, . , •: „ o, _j
/ i
v i \ / y
\ /
Bagged Composite Parts
Autoclave processing is carried out under high pressure and temperature. The tempera-
ture and pressure requirements are set by the material systems used. For example, for high
temperature thermoset resins such as polyamides and PMR-15, the temperatures required
could be in the range of 300-400° C and pressure requirements may exceed 1 MPa. For such
cases, one may need special autoclaves. Thermoplastic composites can also be processed
in these autoclaves as most advanced thermoplastic resins such as PEKK and PEI melt at
high temperatures and require higher pressures for consolidation.
Autoclave consolidation has been the champion of the aerospace industry for years.
For aerospace applications, the parts can be very large, such as the case would be with
an airplane wing; consequently, the autoclaves need to be even larger. The autoclave
dimensions for some aerospace applications can be as large as 30 meters in diameter, and
over 50 meters in length. This is equivalent to a multi-floor building. The largest existing
autoclave in the U.S. is at the Boeing commercial aircraft plant in Seattle, WA — large
enough to fit an entire length of a commercial airplane wing. The autoclaves are usually
made of welded steel as they are subjected to high pressures and temperatures. Hence they
are expensive pieces of equipment [51]. This is the main reason for reluctance to use this
process in other industries besides aerospace, such as automotive and sporting.
Figure 8.3: Schematic of stacking order in autoclave processing (redrawn from [297].
Typically, thermosets have been used in autoclave processing. For most thermoset
resins, the heat of the autoclave is actually necessary to initiate the cure reaction. The
typical temperature range for autoclave processing of thermosets is around 100-200°C, with
pressures in the 600-700 kPa range. Thermoplastics can also be used, but as is typical with
thermoplastics, the temperature and pressure requirements are much higher than those for
thermosets. For example, thermoplastic matrix materials such as PEI or PEKK require
processing temperatures in the 300-400° C range with pressures of the order of 1 MPa.
The typical inputs to the autoclave process are the raw material quality, attributes of
part preparation and handling, and most importantly from the process modeling viewpoint,
the pressure and temperature cycle imposed on the autoclave for composite manufacturing.
The quality of the part manufactured is measured in terms of final thickness or dimensions,
degree of cure or cross-linking (for thermosets) and void or porosity content. The prepregs
used in the composite layup can be either impregnated (either partially or fully) tapes with
smooth surfaces, or fully impregnated tapes with rough surfaces. It was found that fully
impregnated tapes with smooth surfaces do not provide a path for the volatiles and air to
escape from the part and can lead to a lower quality part. The material preparation can
influence the part quality if the procedure of layup is not executed in a repeatable fashion.
However, as a process modeler, the most effective way to influence the part quality is by
controlling the pressure and the temperature cycle during the curing and consolidation of
the composite part.
An example of a pressure and cure cycle is shown in Figure 8.4. There are typically
three stages of autoclave processing. Stage I is the heating stage. At the same time, the
pressure is ramped up inside the autoclave during this stage. The first pressure ramp
and hold in stage I is for the viscosity to go down and for excess resin to flow vertically
into the bleeder material. The second ramp and hold is for polymerization of the resin to
initiate during which time the viscosity rises dramatically. Stage I ends when the selected
processing pressure and temperature are reached. Stage II is essentially a "hold" stage. The
autoclave is maintained at the processing pressure and temperature to allow consolidation
and curing to occur during this stage. As the cure reaction proceeds and generates heat
due to polymerization, in stage III, the temperature is lowered to allow the excess heat
from the reaction to diffuse through the part but the pressure is maintained to prevent
voids from growing. When the cure reaction is complete and the temperature of the part is
lowered, the pressure is released and the part is removed from the autoclave. The vacuum is
discontinued to allow for out-gassing of volatiles. Although the autoclave pressure is high,
Time
Figure 8.4: Typical autoclave temperature and pressure cycles during the three stages of
autoclave processing.
Solution
Consider the sketch shown in Figure 8.6 of a nucleated void of radius R. To relate, dR/dt,
the rate of growth of this void, to the resin pressure, one can balance the stress at the
interface between the resin outside the void and the air inside the void. One can write the
following equation that equates the normal stress on either side of the interface location
= R:
2a
— i-oaid = ~Pr (8.1)
~R
-=R
where Pvoid is the void pressure, Pr is the resin pressure, R is the radius of the void, a is
the surface tension of the resin, and rrr is the normal stress experienced by the resin. For a
Newtonian fluid, low Reynolds flow and considering spherical symmetry, the normal stress
UIUU
Fiber, f
Resin, r-
Figure 8.5: Redistribution of resin and fiber pressure as the network is compressed [61]. In
(a), the fiber network is carrying no load. In (b), as the network is compressed partially,
more and more load is being taken by the fibers, and in (c), all load is taken by the fibers.
Figure 8.6: Schematic of force balance during void growth in autoclave processing.
(8.2)
.n, -it
Prom Equation (8.2), one can easily conclude that, as Pr decreases, R increases. Hence,
it is important to have (Pvo;d ~~ -Pr) < 2a/R in order to suppress the void growth. •
Thus as resin pressure goes down, the rate of bubble growth could potentially increase.
It is crucial to understand and predict the resin pressure inside the part during consolidation
as it plays a dominant role in formation and suppression of voids. If resin pressure drops
to zero, voids will form freely inside the composite since there is no back pressure to resist
the void growth. Intuitively, one may want to increase the autoclave pressure to prevent
this. However, as seen from Figure 8.4 applying too much pressure can cause consolidation
to occur too quickly (in which case the pressure will be all borne by the fiber network), or
cause regions to lose too much resin. If this happens too quickly, then resin pressure will
be lost before the part has had a chance to fully cure initiating void growth. Also complete
wet-out of the fibers may not occur which is detrimental to its mechanical properties. Hence
it is important to control the resin pressure to allow slow enough consolidation to allow total
wet-out, while maintaining sufficient resin pressure until the part solidifies.
There are many combinations of pressure and temperature one could apply with time
to change the cycle shown in Figure 8.4 to produce the part. However one is limited by
how fast one can change the pressure and temperature in the autoclave. At a given power,
to find the time it takes to increase the temperature by 1°C, one would have to consider
the mass and heat capacity of the autoclave. In general the MCp term for the autoclave
and the tooling can be much larger than that of the part where M and Cp are the mass
and specific heat, respectively. Hence, large autoclaves are not well suited for rapid rates
in heating or cooling of the autoclave due to their large thermal inertia.
Solution
The schematic of the autoclave is shown in Figure 8.7. One can write the heat transfer
to the autoclave as Q = QAt = MCpAT. Here Q is the heater power, At is the time
to increase the temperature of the autoclave by AT = 1°C, Cp is the specific heat of the
autoclave material, and M is the mass of the autoclave which can be calculated as
M = Mtube + 2Mendplate
= P (Kube + 2T4ndplate)
f<V ~-~ ~~
'/ - - - ~ "".I I—"—"'"""
Taking Cp = 460 J/(kg °C) for steel, one can find the time to heat the autoclave by 1°C
At =
3 4 106
= - *: j
seconds .0 ^
(8.4)
Q
where Q is in Watts. Note that one Joule is equal to 1 Watt second. Figure 8.8 shows how
long it takes to heat the autoclave of these dimensions by 1°C at various power inputs. For
example, it takes 5 minutes and 40 seconds if the heater has a power of 10 kW, and only
1 minute and 8 seconds if the input power is 50 kW. The rate at which one can increase
the autoclave temperature is plotted as a function of the power input in Figure 8.9. This is
obtained by rewriting Equation 8.4 such that ^ is on the left-hand side:
^ °C/seconds (8.5)
At Mcp 3.4* 106
Thus, the heating rate is directly proportional to power input. •
For industrial autoclaves, which are generally heated by natural gas burners, it is not
uncommon to have heating rates of about 2 to 3°C per minute.
The most widely used method of pressurizing the autoclave is to use nitrogen gas. Air is
unsuitable at high temperatures due to the risk of auto-ignition. However as the approach to
increase the pressure is to use a compressible medium and as the volume to be compressed is
large, one can only expect sluggish rates of increase in pressure. However, there is sufficient
flexibility in application of the pressure to the part. Usually the parts are placed on one side
of the tool and wrapped in a plastic bagging material. Pressure is then applied to the part,
Figure 8.8: Time to heat the autoclave by 1°C at various power inputs.
Figure 8.9: Heating rate (temperature rise) AT/At at various power inputs.
- Prepreg stack
Release agent/film
Dam Mold
The next step involves heat, resin flow, consolidation and cross-linking of the composite
layup. In this step, the process parameters of pressure, temperature and time play an
important role in the quality of the part that will be made, as this process is irreversible for
thermosets. Insufficient pressurization can result in excess resin in the part. On the other
hand, high pressures can lead to resin starved regions. The spatial curing distribution and
the degree of cure is determined by the temperature cycle with time. Usually the focus is to
understand the flow and heat transfer dynamics in this step to effectively model this step of
autoclave processing. The final step of post processing usually will initiate with inspection
and based on the findings may require either rework including further heat treatment to
increase the degree of cure, trimming, machining or assembly.
Assumptions
• The fibers in the composite can be considered as nonlinear elastic porous media.
• The composite consists of incompressible resin and fibers (pr and pf are constant).
dx
Figure 8.11: Schematic showing the geometry and the deforming coordinate system of a
control volume.
(8.6)
Vf dz
Here V0 is the initial fiber volume fraction and Vf is the fiber volume fraction during the
consolidation stage. One can now use the control volume shown in Figure 8.11 and establish
V
v\
V}] °] n
- V}\ = °
Using Darcy's law instead of the momentum equations, one can introduce the velocity and
pressure gradient relationship through the permeability of the preform in the x, y, and z
direction
^x — ,-, i '-'y — o i ^2 — ,-,
H ox p, ay n oz
where Kxx, Kyy, Kzz are the permeabilities in the x, y, and z directions. It is assumed
that x, y, and z are the principal axes of the permeability tensor. Substitution of velocities
in terms of pressure gradients in the continuity equation and assuming that Vf does not
change with x or y direction results in
3.12)
(8.14)
Kzs l ( l - L ) 2 / 3 L t a n - g T . £2 . , , _2 4V
- -- —f (8.15)
Notice that all of these equations require an empirical constant such as pzz or kc except
for the Bruschke-Advani model, which is based on flow across solid cylinders for a certain
packing geometry and fiber volume fraction, r or r/ in these equations refers to the radius
of the fiber.
Substitution of any of the permeability relationship with the fiber volume fraction in
Equation (8.10) results in a partial differential equation for resin pressure in which Vf is
also unknown. However, using Equations (8.10), (8.11) and (8.12) one can solve for Vf
and Pr if the applied autoclave pressure is known. Another important quantity is the final
fiber volume fraction after the compaction since this will give an idea of the mechanical
properties of the panel, as well as the final consolidated thickness. These are related via
nAw
hh = —— fsi7\
(8.17)
where h is the consolidated thickness in [m] in SI units, AW is the weight of the fabric per
unit area in [N/m2], n is the number of layers, and p is the weight density of the fibers in
[N/m3].
Usually in autoclave processing, one strives to eliminate the flow in the in plane direction
(x and y direction) and focuses on bleeding the excess resin through the z direction into the
bleeder cloth. Thus one can easily show by dimensional analysis that if h <C a and 6, one
would expect the pressure gradient and the resin flow to be primarily in the z direction.
This allows one to reduce the Equation (8.10) to
V0 dV
Solution
To solve Equation (8.18) to obtain Vf as a function of time, first assume \i, to be constant
and also the autoclave pressure, PT to be constant. The approach is to define a new variable
e as follows:
I — Vf resin volume fraction
Vf fiber volume fraction '
Hence, Vf = 1/(1 + e) and dVf = —de/(l + e) 2 , and Equation (8.18) can be rewritten in
terms of e as follows:
where V0 =
One can solve the above equation for e, along with the equations for permeability and
fiber stress constitutive equations. The initial condition is that the fiber volume fraction is
V0 at t = 0 and two boundary conditions are necessary along the z direction. At z = h,
the boundary condition would be that as Pr = 0; therefore, azz will be equal to PT. Hence,
one can find the fiber volume fraction using Equation (8.12) at z=h, which will allow one
to calculate e at z = h. The second boundary condition is at the tool surface, that is at
z = 0, there is no flow of the resin out of the tool surface. Mathematically, this can be
expressed as dPr/dz = 0 at z — 0. One can take the derivative of Equation (8.11) at z — 0,
which gives dP^/dz = dPr/dz + do~zz/dz. This implies that dazz/dz = 0 at z = 0 as the
other two terms are zero as well. dazz/dz will be zero only if dVf/dz = 0 at z = 0. One
can express gradient of Vf in terms of de/dz. A typical solution of how the resin pressure
and the fiber volume fraction change with time due to known applied pressure gradient
are shown in Figures 8.12 and 8.13, respectively. Note the significant variation in the fiber
volume fraction as a function of thickness at early times, but as one approaches a steady
state, the Vf becomes more uniform. •
0.75
0.70
0.65
0.60
0.55
0.50
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
To simplify the analysis, one can assume that the resin bleeding is complete and can use
the energy conservation principle to write the governing equation to predict the temperature
field inside the composite as a function of time. This temperature field will be a function of
the thermophysical properties of the composite, the cure kinetics of the resin, the geometry
of the part and the temperature of the autoclave. One could control the temperature
history of the composite by changing the temperature of the autoclave. Mathematically,
this translates into boundary conditions on the composite surface that may be of the first,
second or third type as discussed in Chapter 3.
One can obtain the following governing equations for conservation of energy and resin
cure state with temperature as a primary variable for a composite that is undergoing curing
without any resin flow:
.dT d // £. &T\I J d /, dT\1 _| d (/ 1I. dT\I _l_ (.1
„ dC* ,,
T\fr\rt
d- 1
n f-.,
dt
—
— r-. I ^x ^ I ^ f-i
[ L
dx \ dx J dy \ dy JJ ~*~ dz
I ^V r-i o I ^2 r-i
\ dz JI "TV-*-
dt,. •
f7
VflPrflr
(8.23)
Here p is the density, Cp is the heat capacity and V/ is the fiber volume fraction. The
subscript r is for the resin and / is for the fibers. kx, ky and kz are conductivities of the
composite in the x, y and z directions, respectively. dC*/dt is the degree of cure expressed
rlC1*
^— = k0exp(-E/RT)(C*)m(l-C*)n. (8.24)
By selecting the cooordinate directions to coincide with the principal directions of the fiber
preforms, one can write the governing equation as cast in the form of Equation 8.23. The
composite is usually much longer in the in-plane directions (a; and y directions) as compared
to the thickness direction and hence one can usually eliminate the first and second terms
on the right hand side in the analysis (see Example 5.3 in Chapter 5 for details). The
constitutive form of the cure kinetics equation is usually a function of the temperature
and the current cure state (Equations (4.23) through (4.31)). One can solve Equations
(8.23) and (8.24) simultaneously to obtain the temperature and cure history as a function
of the imposed boundary and initial conditions. The initial condition for the composite
temperature is usually room temperature, and for cure it is zero cure. The boundary
conditions are related to the temperature of the autoclave. The simplest approximation is
that the temperature of the two surfaces at z = 0 and at z — h is equal to the autoclave
temperature Tw(t), and one can solve the coupled system of equations to obtain temperature
and cure simultaneously. See reference [300, 301, 302] for details.
Figure 8.15: Various mold sections to manufacture the composite part depicted on the
bottom right.
Under compressive clamping force, the mold parts completely enclose the fiber preform
that is placed inside. By conforming the preform to the desired dimensions of the final
part, a near-net-shaped part can be manufactured by this type of LCM process. However,
process and design engineers should be careful in selecting the mold material. If the thermal
expansion coefficients of the mold and composite materials are dissimiliar they will expand
or contract at different rates during the heating/cooling stages of cure. This will usually
lead to significant warpage or cracks in the composite part. Usually exact match of the
thermal coefficients is not achievable, and design engineers have to adjust the mold cavity
dimensions to compensate for the different expansion and contraction rates of the two
materials. However, the key advantage is that this type of process allows one to obtain very
tight tolerances which is rarely possible with one-sided molds.
RTM, the most widely used of the closed mold processes, was first introduced in the
1940s. A schematic of the RTM process is shown in Figure 8.16. One can describe the RTM
process in five steps: (i) manufacturing of reinforcing preform, (ii) draping of preform over
a tool surface, (iii) compaction of preform by closing and clamping of the mold, (iv) filling
the empty spaces between the fibers of preform with resin and cross-linking of the resin,
and (v) demolding.
1. Preform Manufacturing
5. Demolding and Final Processing
Resin
3. Mold Closure
t Vacuum
Tube
Figure 8.17: (a) "Vacuum bagged" VARTM and (b) SCRIMP [307, 58].
FASTRAC layer
graphite or
inner vacuum bag fiberglass
L
primary (inner)
vacuum line x>
secondary (FASTRAC )-
vacuum line
The major advantages of the single-sided LCM processes are (i) a very simple injection
system, (ii) a simple mold, (iii) visual control of filling stage on the surface through the
transparent plastic film (iv) capability to manufacture parts of the order of a few meters
such as train and bus compartment sides that can be upto 15 meters long, and (v) low
capital investment.
However, there are many disadvantages, including (i) long fill time due to low injection
pressure of 1 bar, (ii) low fiber volume fraction due to low compaction pressure of 1 bar,
(iii) low quality surface finish and low dimensional tolerance, (iv) requirement for skilled
labor during the bagging stage, (v) labor intensive as in Autoclave processing, (vi) material
wastage is much higher than other processes, and (vii) for thick parts, formation of voids
through the thickness, as one can visualize only the resin close to the top layer adjacent to
the plastic film.
In order to overcome the long fill time disadvatage of VARTM, The Seeman Corpo-
ration resin infusion molding process (SCRIMP) was developed and patented under their
company's name. A very porous fabric, known as the distribution medium, is placed on top
of the preform. Since it has very low resistance to resin flow, resin saturates this porous
layer quickly and then flows into the reinforcing preform in the thickness direction. Other
similar processes such as FASTRAC have been developed to improve the delivery approach
to get the resin from the container to the surface of the preform quickly. In such approaches,
channels are formed along one of the faces of the mold to distribute the resin along one
of the faces as shown in Figure 8.18 [308]. The quick distribution of the resin is shown in
Figure 8.19 [308, 309].
Resin System
Resin forms the matrix of the composite part after it polymerizes. A resin system may
consist of some or all of these ingredients: promoters, fillers, mold release agents, pigments
and catalysts. Some of the most commonly used resins in LCM are polyester, epoxy, vinyl
ester, urethane, and nylon. Fillers such as clay, aluminum trihydrate and calcium carbonate
are used to reduce costs. A good resin system should have the following properties:
• Low viscosity to ease injection.
• Effective wetting properties to wet the fabric.
• Uniform and consistent cure kinetics.
• Rapid cure after gelation.
• High transition temperature.
• High tensile strength to support the reinforcing fabric under load.
Low resin viscosity is desirable as it will require low injection pressure under constant
flow rate injection, or short filling time under constant pressure injection. If the resin
viscosity exceeds a typical upper bound of 500 centipoise (typically, one centipoise is the
viscosity of the water), it will not only require high injection pressure equipment, but might
also "wash-out" the fabric by forcing it to move from its originally placed configuration.
At the other extreme, injecting a resin with too low viscosity (typically lower than 50
centipoise) is much easier, but it may fail to "wet-out" the interior of each individual fiber
tow (fiber bundle) even if the resin impregnates the empty spaces between the fiber tows.
A thermosetting resin system can be injected either as a single component system or
a multiple component system. In multiple component systems, the resin is premixed with
the cataylst in a single container in the desired ratio and injected into the mold. This is
a simple and reliable system but one must use up the resin before it gels or have to clean
the container before the next injection. The other option, the single component system
requires separate containers for the resin and the catalyst or the accelerator, which are
mixed "on the fly" as they enter the mold. This allows the flexibility of changing the mix
ratio to accelerate or decelerate the cure. However this could introduce variability in the
part quality due to uneven mix or cure.
Table 8.2 shows a comparison of the most commonly used LCM thermosetting resins
from the viewpoint of performance, processing and cost [9].
Resin cures due to the chemical reactions to create a matrix around the reinforcing
fabric. Usually, the chemical reaction is a step- and/or chain-growth polymerization. The
length of the cure cycle depends on the type of resin system and temperature of the heating
system. During the cure cycle, the viscosity increases and then eventually gellation starts
at Tgei. The cure kinetics of resin systems, especially the length of the cure cycle, can be
adjusted by changing the amount of catalyst and accelerator that are added to the resin
system. Some resin systems are designed for room temperature cure.
The strength of the matrix starts to build up after Tgei. The part is demolded when the
matrix reaches a "green strength" such that the part will not deform or warp after being
taken out of the mold. Wedges, mallets or hydraulic ejection system can be used to demold
the part.
Demolding the part before it reaches the maximum strength has advantages and disad-
vantages. The main advantage is that the production cycle is shortened by enabling early
injection of the next part. Another advantage is that the damage due to exothermic resin
cure reaction on the mold surface is reduced. The disadvantage is that the demolded part
might warp or develop scratches on its surface. Hence, demolding should be done neither
too early nor too late. Some trimming might be needed to remove excess matrix along the
edges of the part.
Although most RTM applications use thermoset resin systems, research is underway to
use thermoplastic resin for RTM due to the added attraction of recyclibility. The research
is focused on reducing the viscosity of cyclic thermoplastic resins to the order of thermoset
resins to promote impregnation.
Mold (Tool)
Mold material must be selected considering the production volume, the temperature bounds
of the heating/cooling system, tolerances in part dimensions, and the quality of surface
finish required. Steel, aluminum and copper are the most commonly used metals. Steel
is preferred for high volume production because of its high resistance to wear and hence
can last longer. Aluminum and copper are easier to cut during tool making. For low
production volumes (less than several thousand parts), nonmetallic materials such as wood
and reinforced polyesters or epoxy composites are preferred because of easy and rapid tool
manufacturing and low cost.
Temperature fatigue, solvents, mold release agents and accidental scratches made on
the manufacturing floor cause wear of the mold surfaces. A nickel shell on the mold cavity
surface delays the wear of the surface and is usually applied to molds that need to produce
large number of parts.
Upper half
of mold
Composite Part
Lower half
of mold
Figure 8.20: A finite element model of a complex RTM mold with heating and cooling
channels and the composite part.
Runner systems (a network of resin-carrying tubes for the inlet gates) are sometimes
used to reduce the filling time or reduce injection pressure by having multiple resin carrying
tubes, T or multiple connectors, and valves between the injection machine and the mold,
instead of having a single tube and and a single gate. For large parts, runner systems are
required in order to inject the resin within a limited time (of the order of minutes), before
Injection System
Simple injection equipment can be used for single component resins, or batch mixed multi-
ple component resin systems. However, expensive injection equipment with the capability
of mixing the resin components is preferred for high production volumes of composite parts
with multiple component resin systems. Ideal injection equipment should allow the man-
ufacturing engineer (i) to change injection pressure or flow rate, (ii) to open and close
gates for sequential gating by opening/closing gates, and (iii) to monitor the temperature,
pressure, flow rate and cumulative amount of resin injected.
Most of the injection equipment is designed for either pressure controlled or flow rate
controlled injections. Flow rate controlled injection equipment should allow the user to set
an upper critical injection pressure to protect the mold from damage and bending. Once
the critical pressure value is reached, either the flow rate may be reduced, or the injection
may be continued at the critical pressure value.
Radius Engineering Inc. [314] and Liquid Control Corp. [315] are some injection equip-
ment manufacturers for LCM. Plastec and Liquid Control allow for single component sys-
tems, where one has separate containers for the resin and the catalyst and one can pre-select
the mix ratio before the injection. Figure 8.21 shows an injection machine with simple con-
trols from Radius Engineering Inc. [314].
Once the mold is closed, even in one-sided mold processes, one cannot know or see if the
resin has completely saturated the preform or if there are regions inside the composite with
only "dry" fibers (uncovered by the resin). Sometimes even very small regions without resin
can cause catastrophic failures of the composite part. Modeling of the flow will allow one
to investigate the resin impregnation process and strategically design gates and vents and
injection schemes to optimally fill the composite part without any dry spots. Almost all
of the issues that influence flow impregnation are related to the permeability in the mold.
Higher permeability along the walls and corners of the mold due to unintentional gaps
between the preform and the mold wall causes racetracking and changes the flow pattern
dramatically. Draping the fabric over the tool surface can make the permeability change
in magnitude and direction over doubly curved surfaces and introduce changes in the flow
pattern of the impregnating resin. If the permeability of the fiber tows is distinctly less
than the permeability between the fiber tows, it gives rise to a dual scale porous medium
flow in which the fiber tows may fill at a much slower rate than the flow in between the
tows. In addition, if the flow slows down dramatically, surface tension effects of the resin
can influence the microscopic flow pattern. These issues are schematically shown in Figure
8.22.
Thus, during the mold filling stage these flow issues can influence the final quality of
the manufactured part and the decision as to whether one can adopt the LCM process to
manufacture a selected net-shape part. The flow issues arise due to significant changes in
the flow pattern as a result of minor changes in the material or process parameters. Below,
we discuss the most crucial ones that could be modeled and incorporated into simulations
to understand and resolve such issues.
Racetracking
When the fiber preforms are placed in a mold, one may inadvertently create a small gap
between the mold wall and the preform. The flow always uses the path of least resistance
and will race along this edge. This phenomenon has been termed racetracking [54, 316, 317,
318, 319, 320, 321, 322]. Figure 8.23 shows experimental snapshots of racetracking along an
edge of a mold, and Figure 8.24 depicts the changes in the flow pattern due to racetracking
around edges and a corner in a simulated mold filling example.
This racetracking is manifested not only along the mold wall edges and corners but also
around ribs, T-joints and inserts which are usually present in complex net shaped structures.
Tackified preforms could provide a good solution to reduce the possibility of racetracking
channels or at least make such disturbances tractable. A predictable model of resin flow
with racetracking along the edge can allow the process engineer to take advantage of such
channels to impregnate the part. Hence modeling of such channels will be very useful.
Transverse Flow
Macro Void Formation
Unsaturated Effects
Figure 8.22: Flow issues in LCM filling process: (i) race-tracking, (ii) macrovoid formation,
(iii) transverse flow, (iv) preform deformation due to draping and compaction, (v) microvoid
formation, and (vi) unsaturation effects [54].
Figure 8.23: Experimental snapshots of racetracking along the left edge of a rectangular
mold cavity [316, 319]. Resin enters the mold cavity from the upper edge by creating a line
injection. Horizontal flow fronts would be developed if there was no racetracking.
zoom
zoom
zoom
I I
Flow fronts
Figure 8.24: Simulated racetracking along the edges of a mold: (a) mold gap with possible
regions for racetracking, (b) no low resistance areas; hence, simulation shows uniform filling
pattern, and (c) racetracking along the edges changes the flow pattern significantly and can
cause dry spots during manufacturing.
.V"
' * &•**'
•-•-•-
^ V
and
fh
Ul(z)dz = -^^-. (8.26)
One can also write the flow rate using Darcy's law:
AP
= Ui(wh) = ,^^^. (8.27)
The permeability of the racetracking channel, Krstce can be obtained by equating the two
flow rates, Qp0iseuille = QDarcy
^race = ~- (8.28)
Note that substitution of the values for h and w = O(l)mm gives an estimate for Krat
channel which is about two to three orders of magnitude larger than the .Kbulk = 1*10~10 m
Macrovoid Formation
Macrovoids are formed when the air being displaced by the resin gets entrapped or when
there is not sufficient resin pressure to overcome the resistance offered by the fiber preform
and/or the viscosity of the resin. Figure 8.26 shows snapshots of a mold filling situation
at different times in which the resin is impregnating the preform. The experiment clearly
shows the formation of the dry spot or the macrovoid.
These dry spots (voids) might significantly reduce the strength of the composite part
depending on the location and the size of them. The size and location can usually be
detected by nondestructive evaluation (NDE) methods such as ultrasound. Entrapment of
dry spots is a critical issue as it will cause a high rejection rate. Flow modeling can definitely
help in eliminating or reducing the size of the dryspot by identifying the last regions to fill,
once the injection gate locations are specified. This will allow the mold designer to place the
vent locations in the last regions to fill. The current trial and error methods allow the resin
to bleed out of the vent for sufficient time in the hope that the void will get transported
towards the vent and eliminated. This technique may be successful, if the void is within
the proximity of the vent. Morevover, flushing of the resin causes wastage and increases the
cycle time.
Transverse Flow
As most composite structures are thin, one can ignore the flow in the thickness direction
and assume an average value for it. This is a good assumption when the in-plane permeabil-
ities of the preform do not change by orders of magnitude through the thickness. However,
transverse flow can be significant if the transverse permeability component of a fabric pre-
form is much smaller than the in-plane components (see Example 5.8 in Chapter 5). Also,
for processes such as SCRIMP and FASTRAC, the main goal in the process is to speed up
the flow or reduce the pressure requirements by flooding the preform face with the resin
by introducing a very high in-plane permeability material. This will introduce flow across
the thickness of the composite which will be very nonuniform and may trap macrovoids
as shown in Figure 8.27. In VARTM, the transverse flow is significant if L\ ^> LI where
LI and 1/2 represent the distances between the injection gate and the flow fronts in the
distribution medium and bulk preform, respectively, as shown in Figure 8.28.
To address this issue, one must address the flow in the transverse direction. Either
analytic models [327, 328] can be developed or fully three dimensional simulations need to
be implemented [329, 330, 331, 332]. One would also need to characterize the transverse
permeability of the preform which is usually much more difficult to measure [333, 334, 335].
Trapped macrovoids
t=t
:f Mold wall
Distribution Media
A.
Inter-Bundle
Voids
1
Intra-Bundle
Voids
Figure 8.29: Microvoids are created because the permeability inside a fiber tow region is very
different than the permeability of the region between the fiber tows in woven and stitched
fabrics, (a) Interbundle voids are created as resin races along the tows and encircles itself
along a stitch, (b) Resin flow is faster in between the tows which traps voids in between
the fiber bundles [336].
However, if the flow is faster in between the fiber tows, the resin will encircle itself along
a stitch and entrap a microvoid within a fiber tow as shown in Figure 8.29. This issue is
usually addressed on the manufacturing floor by slowing the flow down, allowing sufficient
time for the regions within the tow to impregnate as in most instances, the fiber tow regions
are about two to three orders of magnitude less permeable than the regions in between the
fiber tows. Modeling can help understand the important parameters that influence this flow
in between and within the tows which should prove useful in eliminating the microvoids.
Hence, it may be important to include this phenomenon in flow modeling.
Solution
Under steady-state conditions, one can show that
where qi is the cooling rate of the thin composite of thickness h\ and kt is the transverse
conductivity of the composite. (Tw — Tc) is the temperature difference between the wall
and the center. Thus if the thickness is quadrupled, the cooling rate should be reduced to
one-fourth to continue to maintain the thermal design criteria. Hence the cooling rate for
the thicker composite should not exceed 2.5 Watts/sec. •
The above example was a very crude estimate as we assumed steady state. Polymers
and glass generally have a low thermal diffusivity and hence it takes them much longer to
reach steady state as compared to the processing time. To obtain more accurate predictions
it is useful to solve the governing equations with numerical methods to obtain more accurate
predictors. Example 5.5 shows how one could solve for the transient problem to obtain a
better prediction of the temperature field.
In the last two decades, numerical simulation codes have been developed that incorpo-
rate the process physics of mold filling and cure and thus allow composite process design
engineers to understand the physics and use the simulation as a design tool for prototype
development instead of using trial and error manufacturing approaches. The engineers can
change the process parameters in the simulation and evaluate the influence of the para-
meters on the simulation results in order to search for the optimal material and process
parameters such as locations of the injection gates and vents, injection pressure or flow
rate, cooling rate, resin system, fabric type and fiber volume fraction.
The focus of the next section is to model the impregnation of resin into the fabric
preform, which is usually called the filling stage of LCM, and also develop the non-isothermal
formulation with cure kinetics. The filling stage occurs in all LCM processes with slight
variations. However, in most cases the resin propagates through a stationary fabric preform
which is modeled as a porous medium. The filling model is based on Darcy's law which was
introduced in Chapter 4.
u = -— - V P (8.32)
V
where u is the volume averaged Darcy velocity, /J, is the viscosity of the fluid, VP is the
pressure gradient, and K is the permeability tensor of the preform. In expanded form,
Darcy's law can be rewritten as
where uxi uy and uz are the resin velocity components and Ktj are the permeability tensor
components. The details on measurement and prediction of the permeability components
were discussed in Chapters 4 and 5. Variations of RTM process such as SCRIMP involve
three-dimensional resin flow, as the high permeability distribution medium placed on top
of the preform creates a large transverse gradient flow through the thickness as shown in
Figure 8.28. However, resin flow in most LCM applications can be approximated by two-
dimensional flow because the transverse flow in the thickness direction is only of the order
of a few thicknesses as seen from the example in Chapter 5 and the in-plane dimensions
are usually two to three orders of magnitude larger than the thickness. Approximation
of three dimensional geometry with two dimensional flow brings about significant savings
in computational time and reduces the number of material parameters one must measure
or calculate. By ignoring the flow in the thickness direction, the two-dimensional form of
where ux and uy are Darcy velocities averaged through the preform thickness. This assump-
tion is a very good one as long as in-plane permeabilities of the preforms that are stacked in
the thickness direction are about the same order of magnitude. Note that in SCRIMP, the
permeability of the distribution medium may be two to three orders of magnitude larger
than the preform permeability in the inplane direction; therefore, the two-dimensional as-
sumption is not suitable. Thus, for this process one has to be ready to accept a larger error
in the calculations or resort to a fully three-dimensional flow model. The computational
time for the three-dimensional model would increase by an order of magnitude and one
would require data regarding the transverse permeability of the preform.
In most LCM processes there is insignificant transverse flow; hence, they lend themselves
to two dimensional flow in the plane of the preform. Hence we will restrict ourselves to
discussing the solution of the mathematical model that will be based on a two-dimensional
case for simplicity. The extension to three dimensions is straightforward and does not
involve any conceptual problem.
The equation of mass conservation is obtained by the divergence of velocity as
We also assume here that the tows impregnate at the same rate as the space in between the
tows. Substitution of the velocity terms from Equation (8.34) in terms of resin pressure P
yields the following second order partial differential equation for the resin flow:
d (Kxx dP\ d (Kxy dP\ d (Kyx dP\ d (Kyy dP\
=U
"~~
fj, ~3~~
ox / + «~ «~~ ) + a~
ox \ ^ ay ] ay \ fj, ~a~
ox)) + ay
a~
\ p, ~x~
ay) ' (8.35J
This governing equation is solved to calculate the resin pressure in the resin domain that
has a moving free surface. The solution domain changes continuously. However because
of the "quasi-steady state" assumption, one solves for the pressure in the resin occupied
domain at that instant in time. For open mold processes such as VARTM and SCRIMP,
there is a dynamic distribution of pressure between the preform and the resin which may
change the fiber volume fraction and the permeability due to the presence of the resin just
as depicted conceptually in Figure 8.5 for autoclave processing. More details on this can be
found in [339, 340, 341, 342].
With known material parameters of viscosity and permeability, which may vary with
position, this governing differential equation with boundary conditions along the boundary
of the resin domain has a unique solution. The boundary conditions are written for three
segments of the boundary: (i) at injection gate(s), (ii) on the free surface (flow front), and
(iii) along the mold wall. Usually, the injection system will inject the resin into the mold
under constant pressure or constant flow rate:
P = -Rnj (8.37a)
or
A(-— -VPV n = - Q i n j (8.37b)
PS = or, Pe = 0 (8.38)
where Pg and Pvent are the flow front pressure and vent pressure, respectively. Along the
in-plane mold wall boundary, the normal velocity component of velocity should be zero if
there is no resin leakage:
^ dP „ dP
un = — Knn— + Knt— = 0, (8.39)
an at
where n and t denote the directions normal and tangential to the mold wall, respectively.
These three types of boundary conditions are illustrated in Figure 8.30.
u.n=o
Gate
P-P
or _ I")
(u.n)A=-Q
inj
Vent
It is possible that due to the presence of inserts or multiple gates, multiple resin fronts
may result. When these fronts meet and are not linked to any vent, the air that they
displace gets entrapped in the regions encompassed by the resin flow fronts. For such cases,
the boundary condition given in Equation (8.38) can be modified to include the increase in
the flow front pressure due to the increase in the void pressure. The entrapped air may be
assumed to behave as an ideal gas satisfying
•'void ' void
= constant (8.40)
'void
where Pvoidi ^void and Vvoid are the pressure, temperature and volume of the air enclosed
[343]. However, this would require one to keep track of the void created by merging flow
fronts and calculate its volume at each instant of filling. In reality some of this air will
Figure 8.31: Fiber tows or bundles are impregnated much more slowly than the regions in
between the fiber tows.
In order to model this phenomenon, a sink term is added to the mass conservation
equation (see Equation (3.10) in Chapter 3): V - U = — s. Or in expanded form, Equation
(8.36) is replaced with the following:
_ KxxdP\ | d K^dp_\ K*-yx
VX8P\ . 8 (K _
= S. (8.41)
— )+ —
dx ^i dx J dx dy J dy dy)
Here, s is the rate at which the fluid is impregnating inside the fiber tows, and should .be
proportional to the permeability of the fiber tows. For typical LCM fabrics, individual tow
permeability is usually much lower than the overall fabric permeability. Besides the tow
permeability, the value of s is also dependent on pressure outside the tows, the geometry
of the tows and the density of these tows. For a circular tow, s can be derived as follows
[344, 345, 346, 347]:
s = (8.42)
-1 'la
where
dra Kit(P0 + Pc-
(8.43)
dt • O ln
Here, PQ, Pc, and Pvac are the free surface, capillary and vacuum pressures, respectively, €j
is the porosity of the inner region (intratow), r/a is the microfront radius inside the tow,
7*i0 and r-2a are the tow and the cell (the circular computational domain with the tow and
surrounding air channel) radii, Kit is the fiber tow permeability (see Figure 8.32).
Figure 8.32: Parameters to model the sink effect during impregnation with resin of fiber
tows of permeability, KH.
Thus one would have to solve Equation 8.36 and 8.41 simultaneously to solve for the
macro flow front and the saturation of each tow as the mold filling proceeds. This usually
will modify the pressure profile under constant injection conditions [348, 349].
Solution
(a) In Chapter 4, the relation between the permeability of a single scale preform and the
injection pressure was derived in detail under the conditions when resin is injected at con-
stant flow rate in one-dimensional flow. We will rewrite the final equations for completeness
of this problem:
t. (8.44)
Thus,
pM SQ\2
Kxx = d p (—) • (8.45)
dux
— —s. (8.46)
dx
Here s corresponds to the sink term due to delayed saturation of fiber bundles coinpared
to the empty spaces between the surrounding fiber tows. If s is a constant as stated in the
problem, one can integrate both sides of Equation (8.46) with respect to x, and apply the
boundary condition at the injection gate, ux(x = 0) = Q/(wh), which then results in the
following expression for the resin velocity as a function of x\
ux = -sx+~. (8.47)
One-dimensional Darcy's law relates the resin velocity also to the pressure gradient through
the resin viscosity and the preform permeability in the flow direction, Kxx
KT.xdP_
ux = -- (8.48)
dx
Equating the two ux's in Equations (8.47) and (8.48) yields dP/dx = -£—(sx — ^). One
can obtain P(x, t) by integrating both sides of this equation with respect to x, and applying
the boundary condition at the injection gate, P(x = 0,t) = Pinj(t)
v (8.49)
' ' Kxx | 2 wh
The boundary condition at the moving resin flow front, x = x/, yields the following:
A4 sxj Qxf
P(x = x f , t ) =0 = (8.50)
~2 wh
assuming that P(x — xj, i) = PVent = 0. One can rewrite Equation (8.50) in order to obtain
a relation between Kxx and other process parameters and variables
QXf(t)
(8.51)
Pinj(t) [ 2 wh
One needs to write x/(t) explicitly in order to use Equation (8.51). This is shown below
*/(*) =
Xf(t'}dt'. (8.52)
0
Xf(t) appears on both hand sides of Equation (8.52). In order to write Xf(i) explicitly, the
Laplace transform of Equation (8.52) will be taken as follows:
-Xf(p] = - (8.53)
where A = Q/(wh(/)) and B — s/<j) were used for the sake of simplicity. In Equation (8.53),
Xf(p) is the Laplace transform of x/(t), and p is the Laplace variable. Rearranging Equation
(8.53) by using partial functions, one can obtain the following:
(8 54)
'
One can now take the inverse Laplace transform of Equation (8.53) which results in the
following:
B
Q (8.55)
swh
Finally, one can substitute this expression into Equation (8.51)
Preform permeability Kxx along the flow direction is calculated by substituting injection
pressure at any time (except at t = 0) into Equation (8.56). At t = 0, that expression
is 0/0 and singular. As a good engineering practice, Kxx should be evaluated at several
time values, and an average should be taken due to minor variabilities in s, Kxx, Q, and
racetracking effects. Figure 8.33 illustrates some typical plots of Xf(t) and Pinj(t) with and
without the effect of dual scale flow. In reality, s is not a constant but is coupled to the
flow outside the tows because of the variation in resin pressure. The resulting equations are
nonlinear and require numerical methods to solve them. •
a contour levels
Figure 8.34: Schematic of resin cure during the mold filling stage, a is the degree of cure
[330].
For resins that need to be heated to initiate cure, usually warm or cold resin is injected
into a hot mold containing heated preforms. In such cases the heat is transfered from the
mold and the preforms to the resin as the resin flows into the mold. As the resin temperature
increases, the resin viscosity decreases. However, higher temperature of the resin will initiate
and accelerate its cure and as the degree of cure increases, the resin viscosity will increase.
Usually the viscosity due to the cure can increase by orders of magnitude whereas the drop
in viscosity due to increase in temperature for thermosets is only a few percent. To model
the change in the viscosity due to temperature and cure kinetics in the flow equation, one
must also characterize the viscosity dependence on these functions as discussed in Chapter
4.
The temperature and degree of cure will be nonuniform in the mold, as the heat dis-
tribution is nonuniform due to convection of the resin and also due to conduction of the
heat through the mold walls and the preform. Also, if one uses a multi-component injection
system then the initial concentration of the cataylst may be different for different regions in
the resin, initiating different rates of cure. The important processing step here is to design
heating systems to prevent early gelation of resin and hence incomplete filling stage. For
such situations, one would find the nonisothermal resin flow model along with the energy
Figure 8.35: Schematic of heat dispersion effect; the figure on the left (a) depicts the resin
impregnation directions across the fibers. The figure on the right (b) due to the assumption
of Darcy's law can account for convection only in the flow direction. The difference in heat
convection effect is expressed by a dispersion term.
A dimensional analysis shows that, for mold filling applications with a small Brinkman
number, viscous dissipation is negligible. Further, for thin shell-like domains, heat con-
dT
- + Cbc(T - =0 (8.58)
where n is the normal in the outward direction to the mold wall, and the boundary condition
constant C&c is written as follows:
l
3.59)
kt l/hh + l/hri
where hh and hm are the heat transfer coefficients between the heating fluid and the pipe,
and between the mold and resin, respectively, km is the thermal conductivity of mold
material, kt is the conductivity of the composite in the transverse direction, a is the distance
between the mold wall and the heating pipe, and T^ is the temperature of the heating fluid
as shown schematically in Figure 8.36.
heating
pipe
top mold a z
platen at
Resin o o OAP 0*0 o
flow ~~* o o o o o
Figure 8.36: Boundary condition at the mold wall across the thickness of a composite [54].
The other boundary condition is usually at the injection gate where the temperature of
the resin is usually prescribed. At the flow front, an energy balance allows one to account
for the convection into the unsaturated region as shown in Figure 8.37.
T0, unsaturated
preform
temperature
fArea^:!™"1"""""""1
m's IB m
^i^ mi»
^^
fluid filled preform
The s term in the energy equation models the heat released by the exothermic reaction
when the resin molecules are cross-linking as the resin cures. Hence one needs a constitutive
equation that can characterize cure and also provide the rate at which it generates energy.
Solution
The energy equation, Equation (8.57), for steady state reduces to
d(T)
= k, (8.60)
dx
with boundary conditions
Here (pcpjr is the heat capacity of the resin, (u) is the Darcy's velocity in the flow
direction (x direction) and (T) is the average temperature at a given location. Here kzz is
the conductivity of the medium in the thickness direction. Also as thickness is much less
than the length, one can ignore the conduction in the x direction.
When the filling speed is slow, the conduction across the thickness will dominate. At
steady state, the mold wall temperature will diffuse through the thickness, and resin and
preform will reach the mold wall temperature. The only exception will be the injection gate
region of width h, where the temperature changes from T{n (from inlet resin temperature
to T0) as sketched in Figure 8.38 [355, 358].
The dimensionless number that governs the ratio of the convection in the inflow direction
and conduction through the thickness direction is the Graetz number, Gz:
(u i "-ZZ
Gz = where azz = 3.63)
azzL
This can be obtained by nondimensionalization of Equation 8.60 by using the following
nondimensional variables:
(T) - Tin X
XT' - z'-Z (8.64)
T0-Tin ' ~L' h
Thus when Gz « 1, conduction dominates as shown in Figure 8.38. One can easily
solve Equation (8.66), by introducing a new variable
(8.68)
where z is the distance from the wall towards the center in the thickness direction. This
allows us to convert a partial differential equation into an ordinary differential equation by
similarity method and the solution can be easily obtained:
(T) - Ti
= l-erf(0 (8.69)
for Gz » 1. Here erf represents the Gaussian error funtion. Thus for Gz » 1 (large
filling speeds), the thermal boundary layer increases with distance from the entrance as
shown in Figure 8.39. The thickness of this boundary layer grows with the square root of
the distance from the mold entrance and is given by
(8.70)
s—
where
(8.73a)
(8.73b)
where R is the universal gas constant. The resin material constants m, n, AI, A%, E\ and
E-2 are measured experimentally, as discussed to some extent in Chapter 4.
The cure convects with the resin as it flows to occupy the mold; hence one can express
the conservation of cure in mathematical form as follows:
(8.75)
iwhere the activation energy E^ and the constant A^ are given as follows:
E^ = a + ba (8.76a)
(8.76b)
This set of governing equations can be solved at any instant during the filling process. The
zth component of [[<5e]][P] vector is the amount of mass generated per unit time at the
qij = n • K ds 3.81)
where Sij is the boundary between the two control volumes, h is the mold gap thickness at
the boundary, and n is the normal vector to the boundary lying in the plane of the preform.
Nodal fill factors are used to track the moving flow front. The fill factor for each node is
defined as the fraction of its control volume occupied by the fluid as shown in Figure 8.41
[337, 214, 330].
Fil I ed area of th §1
Consolidation
Composite Substrate
F-.ber Band
Fiber)
for customization of properties such as torsional stiffness, axial stiffness, tensile strength,
proportional elastic limit, hardness, abrasion and wear resistance.
During this process, the following physical phenomena occur:
• Thermochemical reaction and heat transfer: The resin cures due to the cross-linking
exothermic reactions of the thermoset matrix, and the heat generated is conducted to
the ambient.
• Resin Slippage: Since there is uncured resin between the fibers, the fibers slip from
their original position under tension.
• Void Formation: Voids are formed due to the air trapped between the bands of fibers.
• Thermal expansion effects and the development of stresses and strains: The mandrel
and the composite expand due to the changes in temperature which creates thermal
stresses and strains due to mismatch in thermal expansion. This adds to the stresses
developed due to the fiber tension.
Figure 8.45: Interrelationship of submodels for filament winding with thermosetting matrix
materials
In this section, we will consider only the thermochemical reaction and the fiber motion
or the resin seepage submodel. These two models are coupled via the resin viscosity, which
is temperature and cure dependent. It is assumed that the winding of the mandrel is in-
stantaneous. The modeling approach will be described with the use of a simple example.
Significant work in the modeling of the filament winding process for thermosetting matrix
composites has been undertaken by [368], and the models have been verified with experi-
ments. The following process submodels are adopted from their work. The goal here is to
introduce modeling and improve the understanding of the modeling process and not solve
comprehensively for all the details.
3T 1 d dT d dT
- 7T
•(8.82)
r <9r
where krr and kzz are the composite conductivities in the radial and axial directions, re-
spectively, p and C are the density and heat capacities of the composite. One could use
simple rule of mixtures to find these values knowing the density and the heat capacities for
the resin and the fibers and the fiber volume fraction. Hu is the heat of reaction for the
resin, da/dt is the rate of cure, where a is the degree of cure. One would need an equation
that can describe the cure-kinetic reaction. The cure-kinetic reaction is usually a function
of temperature. Hence the cure kinetics are coupled with the energy equation and one must
solve them simultaneously
— = rEV(a, HP +\
1 , t). fQ Q'3\
(a.od)
Hence our independent variables are t, r and z and our dependent variables are a and T.
The initial conditions for cure and temperature need to be specified (usually a is zero at
t = 0). As the temperature equation is a partial differential equation, boundary conditions
are required to solve for the temperature. Temperature, heat flux or a convective boundary
condition is usually specified on the composite outside diameter and on the mandrel inner
diameter. In Figure 8.46, constant temperature boundary conditions are applied in the
radial and axial directions. Here hm and h are the thickness of the mandrel and the
composite respectively.
R
mo R
mil
z=0 z=L
Darcy's law as applied to the fiber governs the movement of fiber through the resin. In liquid
molding processes, we considered the fiber network stationary and the resin moves relative
to the fiber network because of the pressure gradient created. In autoclave processing, the
resin was made to bleed across a network of fibers that was being compressed. In this
process, the fibers or the porous network formed by the fibers are forced through the resin
due to the applied tension as seen in Figure 8.47.
Fiber
Sheet
Resin
Figure 8.47: Schematic of fiber motion through the resin due to the applied tension.
Flow of the resin, expansion of the mandrel and expansion of the composite affect the
fiber position. However, in the fiber motion model we will consider fiber movement only
due to the flow of the resin. As seen in Figure 8.47, each composite layer is idealized as a
fiber sheet surrounded by thermoset resin. The motion of the fibers can be denoted by Xf.
Then, Darcy's law gives the rate at which the fibers move relative to the resin
dxs KrrdP
= (8 84)
-dT -—^' '
This model assumes that the resin flows transverse to the fibers in the radial direction.
Another critical assumption is that the resin carries the entire load and none is carried by
the fibers, unlike in autoclave processing. The consolidation pressure is provided by the
component of the tension in the filament over the distance r/, as recast below:
*^sin'*00.. (8.85)
dt [i rf
Here dxf/dt is the rate of radial movement of the fibers, <jf is the tension in the fiber,
Krr is the permeability of the fibrous network in the radial direction, since the fibers are
moving through and relative to the resin of viscosity, fj,. 3>0 is the angle at which the
fiber is wound onto the mandrel and r/ is the radius of the fiber layer. The fiber motion
submodel is coupled with the thermochemical submodel through the resin viscosity, fj,, which
is temperature dependent
). (8.86)
rf = r°f + xf (8.87)
where Ef is the elongational modulus of the fiber material. Hence we have a set of coupled
nonlinear equations for the fiber displacement and the fiber tension, which have to be
integrated to obtain the displacement for every layer. The initial conditions for the above
equations are
xf = 0 (8.89)
af = ^ (8.90)
rj = r°f (8.91)
at t = 0. Here F0 is the winding tension and r°, is the initial radius of the fiber layer. We
will conclude this section by demonstrating the approach to calculate the temperature, cure
kinetic and fiber motion by coupling the thermochemical and the fiber seepage submodels.
ktc l d prVrHu da
dt* pCpbtlfi I r* Or* \ dr*) \ L ) 0(z*}2 frp Tf,—r -^- (8-93)
-
T * -4OT!*- ,'
177 C ' . ' •;
:
= Cure C^cle
Figure 8.48: Sample composite part and cure cycle used for the model solution [365].
Using the documented values of the thermal capacity and heat conductivity of the composite
and aluminum, it was found that the time scale of heat conduction in the composite (45
s) was much slower than the conduction time scale in the mandrel. In addition, since the
composite layer is thin, compared to the length of the part, the terms for the conduction in
the z direction can be neglected.
2
£tot „, PCp&tlt __ PrVrHu
fc
T I ^^ ' t>c —
c
7, i P — n /rri rp \) c,Al ^ <"c,composite •
L J k
(8.94)
This leads to the following simplified thermochemical equation in nondimensional form:
The cure kinetics of the epoxy resin and the relation for resin viscosity can be recast as
follows:
^=(Kl + K2a1-03) (B - a1'22) (8.96)
= A W «P^*J (8.97)
-=tc~ (8.98)
dt* dt ^ '
Since the conduction time scale in the mandrel is much smaller than that of the composite,
the conduction in the mandrel can be assumed to be in quasi-steady state and the transients
there can be neglected. In addition, the chief mechanism of heat transfer from the autoclave
m . _ T"1 . ftrr
L
f~^
, . -i £ ., -'•air -*• composite, out j *~'-
Outside surface, composite: - ----'— = k—, at r — rout/'o -i n-i \
(8.101)
i | '"in, composite \
_ I _ \ nn.mandrel / _ 1
rl
•rt-convl — ~ Ti
i i -t"cona —
-ftcond n
9 ti ' conv2 — «
^-convz i •
in mandrel^ ^^"^niandrel ^^"^"out, composite'^
(8.102)
These boundary conditions can be nondimensionalized to give the following conditions:
The governing PDE may be discretized using finite differences to give the following
scheme:
Ai* At*
v =- FT, u = —r^- , V i = 1 . . . r steps and i — 0 . . . t steps v (8.107)
Ar* 2 r*[z]Ar* .
B .V^.
r~\ 1
S.
rpn+1 *
_i Q ^^
-^
^
Mconv. ,in~i~^cond.,mandrel
^ ,
/7^n+l *
J. yy = ^
H-\-Hco
^
d.,mandrel -^ -ttcoiiv.,out
(8.108)
where R — fc/(AttotAr*). The numerical scheme selected to solve this problem is fully
implicit since the explicit scheme has an inherent limitation in time-step size which can
increase the computational time substantially.
2
^ / - C - ^ (8.109)
Ftension
Initial conditions: x°f = 0, a°f = , r°f = Rf (8.112)
A
f
The discretized thermochemical model can be solved using an iterative solver for the
temperature and the cure rate. The viscosity can be determined from the calculated values
of the temperature and cure. These viscosity values can then be used to solve for the fiber
motion and tension. The fibers tend to move downwards through the resin pool, until the
resin gels and solidifies. As the fibers move down, they become slightly slack and the fiber
tension decreases. Hence fiber tension also decreases with time.
However, other phenomena — the development of stresses and strains, thermal expan-
sion of the mandrel and composite with temperature rise — have not been modeled here.
These phenomena will definitely have a significant effect on the fiber movement. Hence,
though the presented results are reflective of the process physics, a complete picture can be
obtained only by including all these effects.
8.6.3 Problems
1. Assume that (i) the resin used in an autoclave is Newtonian, (ii) the Reynolds number
is low, and (iii) the voids are spherical. If P is the void pressure, Pr is the resin
pressure, R is the radius of the void, and a is the surface tension, find an expression
for these process variables so that the void growth will be suppressed. If you can vary
the temperature and pressure within the autoclave, how will it affect the void growth?
Would you set the pressure to be maximum for autoclave control? Explain why.
2. Equation (8.2) relates the void growth to viscosity, void radius, surface tension, pres-
sures of void and the resin. Can you express R(t) in closed form? How would you
solve R(t) numerically, with an initial condition R(0) — R0, and what data will you
need?
3. Consider an autoclave that is made of welded steel of thickness 20 mm. The autoclave
is a cylindrical tube of diameter 0.5 m and length of 1 m. The tooling and part thermal
inertia can be considered negligible as compared to the autoclave inertia. How long
will it take to heat the autoclave from room temperature of 25°C to 350°C, if the
heater power supplied to the autoclave is 20 kW?
4. By referring to Figures 8.12 and 8.13, answer the following questions: (a) It takes ti
minutes for the mid-point of the composite to reach a fiber volume fraction of about
0.61 from the initial value of 0.53. How long will it take for the entire composite to
be above fiber volume fraction of about 0.70? (b) If it takes t% minutes for the resin
pressure at the top layer to reach its maximum value Pmax, at roughly what time(s)
the resin presure at this location reaches Pmax/27 Explain why there is more than
one unique time values to your answer.
5. Consider a resin flow racetracking along a rectangular channel of height of h = 5 mm
and a width of w = 2.5 mm. If all four walls of the channel are assumed to be
nonporous, calculate the permeability Krace of the channel to be used as a parameter
in one-dimensional Darcy's law.
6. Thermal design criteria require you to keep the temperature differential between the
center and the surface of the composite to less than 20° C. You know that a cooling
rate of 5 Watts/sec for a 6 mm thick composite was able to do so for an epoxy resin.
For a second part of 8 mm thickness, what is the maximum cooling rate you can
impose to meet the thermal design criteria?
[1] Sylvia Kueh. Investigation into the effects of applying pressure during compression
molding. Mechanical Engineering Department, University of Delaware, Report 99.1,
1999.
[2] M. R. Barone and D. A. Caulk. A model for the flow of a chopped fiber reinforced
polymer composite in compression molding. Journal of Applied Mechanics, 53(2) :361-
371, 1986.
[3] S. Middleman. Fundamentals of Polymer Processing. McGraw Hill, New York, 1977.
[4] Petri J. Hepola. Thermoplastic pultrusion with on-line dry powder impregnation of
fibers. PhD thesis, Mechanical Engineering Dept, University of Delaware, Newark,
DE, 1993.
[5] S. G. Advani, editor. Flow and Rheology in Polymer Composites Manufacturing.
Elsevier, Amsterdam; New York, 1994.
[6] T. G. Gutowski. Advanced Composites Manufacturing. Wiley, 1997.
[7] Dave and Loos, editors. Processing of Composites. Hanser Publishers, Munich, 2000.
[8] M. Schwartz. Composite Materials II: Processing, Fabrication and Applications. 1996.
[9] T. Astrom. Manufacturing of Polymer Composites. Chapman and Hall, London,
1997.
[10] T. S. Chou, R. L. McCullough, and R. B. Pipes. Composites. Scientific America,
254(18):193-203, 1986.
[11] S. Shuler and S. G. Advani. Transverse squeeze flow of concentrated aligned fibers in
viscous fluids. Journal of Non-Newtonian Fluid Mechanics, 65:47-74, 1996.
[12] H. Domininghaus. Die Kunststoffe und ihre Eigenschaften. VDI-Verlag, 4th Edition,
Dusseldorf, 1992.
[13] Hubert Stadtfeld. Thermoplastic pultrusion. Mechanical Engineering Department,
University of Delaware, Report 99.8, 1999.
[14] M. Pahl, W. Gleifile, and H.-M. Laun. Praktische Rheologie der Kunststoffe und
Elastomere. VDI-Verlag Dusseldorf, 4th Edition, 1995.
[15] D. W. Becker. International Encyclopedia of Composites, volume 3. Edited by Lee,
S. M., VCH Publishers, Inc., New York, 1990.
[52] Liquid moulding technologies: resin transfer moulding, structural reaction injection
moulding, and related processing techniques. SAE International and Woodhead Pub.
Ltd., Warrendale, PA; Cambridge, England, 1997.
[53] R. S. Parnas. Liquid composite molding. Hanser Publishers, Munich, Germany, 2000.
[55] Emanuele F. Gillio. Co-injection resin transfer molding of hybrid composites. Master's
thesis, University of Delaware, Newark, DE 19716.
[58] E. F. Gillio, S. G. Advani, B. K. Fink, and J.W. Gillespie Jr. Investigation of the
role of transverse flow in co-injection resin transfer molding. Polymer Composites,
19:738-749, 1998.
[61] R. Dave, J. L. Kardos, and M. P. Dudukovic. A model for resin flow during com-
posite processing: Part 1 - General Mathematical development. Polymer Composites,
8(l):29-38, 1987.
[64] M. N. Ozisik. Heat Conduction. John Wiley and Sons, Inc., New York, 1993.
[65] M. Kamal and M. E. Ryan. Fundamentals of Computer Modeling for Polymer Process-
ing. Edited by C. L. Tucker III, Hanser Publishers, New York, 1989.
[121] C. L. Tucker. Flow regimes for fiber suspensions in narrow gaps. J. Non-Newtonian
Fluid Mech, 39:239-268, 1991.
[122] H. Giesekus. Elasto-viskose flussigkeiten, fur die in stationaren schichtstromungen
samtliche normalspannungskomponenten verschieden gross siend. Rheol. Acta, 2:50-
62, 1962.
[123] G. G. Lipscomb, M. M. Denn, D. U. Hur, and D. V. Boger. The flow of fiber suspen-
sions in complex geometries. J. Non-Newtonian Fluid Mech., 26:297-325, 1988.
[131] W. Fisch, W. Hoffman, and R. Schmid. J. Appl. Polym. Sci., 13:295, 1969.
[137] Edward M. Barrall II and Julian F. Johnson. Differential Scanning Calorimetry The-
ory and Applications. In Jr. Philip E. Slade and Lloyd T. Jenkis, editors, Thermal
Characterization Techniques, volume 2, chapter 1, pages 1-39. Marcel Dekker, Inc.,
1970.
[174] M. A. Acitelli, R. B. Prime, and E. Sacher. Kinetics of epoxy cure: (1) The system
bisphenol-A Diglycidyl Ether/m-Phenylene Diamine. Polymer, 12:335-343, 1971.
[175] Yong Deng and George C. Martin. Modeling Diffusion during Thermoset Cure: An
Approach Based on Dielectric Analysis. Macromolecules, 27:5141-5146, 1994.
[178] Roger J. Morgan and Eleno T. Mones. The Cure Reactions, Network Structure, and
Mechanical Response of Diaminodiphenyl Sulphone-Cured Tetraglycidyl 4,4' Diamin-
odiphenyl Methane Epoxies. Journal of Applied Polymer Science, 33:999-1020, 1987.
[179] J. F. Stevenson. Free Radical Polymerization Models for Simulating Reactive Process-
ing. Polymer Engineering and Science, 26:746-759, 1986.
[180] V. Bellenger, J. Verdu, J. Francillette, P. Hoarau, and E. Morel. Infra-Red Study of
Hydrogen Bonding in Amine-Crosslinked Epoxies. Polymer, 28:1079-1086, 1987.
[181] A. Sabra, T. M. Lam, J. P. Pascault, M. F. Grenier-Loustalot, and P. Grenier. Charac-
terization and Behaviour of Epoxy-Based Diaminodiphenylsulphone Networks. Poly-
mer, 28:1030-1036, 1987.
[182] A. J. Attias and B. J. Bloch. Chemical Structure of Networks Resulting from Curing
of N,N-Diglycidylaniline - Type Resins with Aromatic Amines. IV. Characterization
of TGDDM/DDS and TGDDM/DDM Networks by High-Resolution Solid-State 13C-
NMR. Journal of Applied Polymer Science: Part A: Polymer Chemistry, 28:3445-
3466, 1990.
[183] M. G. Rogers. The Structure of Epoxy Resins using NMR and GPC Techniques.
Journal of Applied Polymer Science, 16:1953-1958, 1972.
[184] E. Mertzel and J. L. Koenig. Application of FT-IR and NMR to Epoxy Resins. In
K. Dusek, editor, Advances in Polymer Science - Epoxy Resins and Composites II,
volume 75, chapter 3, pages 73-112. Springer-Verlag, 1986.
[186] H. Batzer and S. A. Zahir. Studies in the Molecular Weight Distribution of Epoxide
Resins. II-Chain Branching in Epoxide Resins. Journal of Applied Polymer Science,
19:601-607, 1975.
[222] S. G. Advani and Dimitrovova. Capillary effects in fiber bundles. In Stanley Hartland,
editor, Surface Tension: Measurement, Theory and Applications. Marcel and Dekkar,
(in press), 2002.
[224] Pavel B. Nedanov, Suresh G. Advani, Shawn W. Walsh, and William O Ballata.
Determination of the permeability tensor of fibrous reinforcements for vartm. In
AS ME, 1999.
[225] S. Walsh. In-situ sensors method and device. U.S. Patent 5,210,499, May 1993.
[227] Alberto Belloli. Automated fiber preform permeability measurement station for vartm
process, diploma thesis, eth, Zurich. 2001.
[229] K.M. Pillai and S.G. Advani. Model for unsaturated flow in woven fiber pre-
forms during mold filling in resin transfer molding. Journal of Composite Materials,
32(19):1753-1783, 1998.
[230] Z. Cai and T. Gutowski. 3-d deformation behavior of a lubricated fiber bundle. Journal
of Composite Materials, 26(8):1207-1237, 1992.
[231] B. Chen, A.H.-D. Cheng, and T.-W. Chou. Nonlinear compaction model for fibrous
preforms. Composites - Part A: Applied Science and Manufacturing, 32(5):701-707,
2001.
[232] Baoxing Chen and Tsu-Wei Chou. Compaction of woven-fabric preforms: Nesting
and multi-layer deformation. Composites Science and Technology, 60(12-13) :2223-
2231, 2000.
[233] Chen, Baoxing, Chou, and Tsu-Wei. Composites Science and Technology,
59(10):1519-1526, 1999.
[338] Coulter John P. and Selcuk I. Guceri. Resin impregnation during the manufacturing
of composite materials subject to prescribed injection rate. Journal of Reinforced
Plastics and Composites, 7(3):200-219, 1988.