Download as pdf or txt
Download as pdf or txt
You are on page 1of 720

Injection and

Compression Molding
Fundcunentals
PLASTICS ENGINEERING

Series Editor
Donald E. Hudgin
Princeton Polymer Laboratories
Plainsboro, New Jersey

1. Plastics Waste: Recovery o f Economic Value, Jacob Leidner


2. Polyester Molding Compounds, R ob ert Burns
3. Carbon Black-Polymer Composites: The Physics o f Electrically Conducting
Composites, Edited by Enid K eil Sichel
4. The Strength and Stiffness o f Polymers, Edited by Anagnostis E.
Zachariades and R oger S. Porter
5. Selecting Thermoplastics for Engineering Applications, Charles P. M acDerm ott
6. Engineering with Rigid PVC: Processability and Applications, Edited by
/. Luis Gomez
1. Computer-Aided Design o f Polymers and Composites, D. H. Kaelble
8. Engineering Thermoplastics: Properties and Applications, Edited by James
M. Margolis
9. Structural Foam: A Purchasing and Design Guide, Bruce C. Wendle
10. Plastics in Architecture: A Guide to Acrylic and Polycarbonate, Ralph
Montella
11. Metal-Filled Polymers: Properties and Applications, Edited by Swapan K.
Bhattacharya
12. Plastics Technology Handbook, Manas Chanda and Salil K. R oy
13. Reaction Injection Molding Machinery and Processes, F. Melvin Sweeney
14. Practical Thermo forming: Principles and Applications, John Florian
15. Injection and Compression Molding Fundamentals, Edited by Avraam I.
Isayev

Other Volumes in Preparation


Injection and
Compression Molding
Fundamentals
Edited by
Avraam I. Isayev
THE UNIVERSITY OF AKRON
AKRON, OHIO

MARCEL DEKKER, IN C N ew York and Basel


L ib ra ry o f C ongress C a ta lo g in g -in -P u b lic a tio n Data

Injection and compression molding fundamentals.

(Plastics engineering ; 15)


Includes bibliographies and index.
1. Injection molding of plastics. 2. Plastics--Mold­
ing. I. Isayev, Avraam I. II. Series; Plastics

1. Injection molding of plastics. 2. Plastics--


Molding, I, Isayev, Avraam I. II. Series; Plastics
engineering (Marcel Dekker, In c.) ; 15.
TP1150.I535 1987 668.4»12 87-13489
ISBN 0-8247-7670-4

C o p y rig h t © 1987 by MARCEL D E K K E R , IN C . A ll R ights R eserved

Neither this book nor any part may be reproduced or transmitted in


any form or by any means, electronic or mechanical, including photo­
copying, microfilming, and recording, or by any information storage
and retrieval system, without permission in writing from the publisher.

MARCEL DEKKER,INC.
270 Madison Avenue, New York, New York 10016

Current printing (last digit) ;


10 9 8 7 6 5 4 3 2 1

PRINTED IN THE UNITED STATES OF AMERICA


Preface

During the past ten years, injection and compression molding of


polymers has undergone considerable development, laying a solid foun­
dation for a science-based molding technology. At present, extensive
use of the computer in mold design and manufacturing is gradually re­
placing skilled mold designers whose expertise in solving molding prob­
lems has been based on experience and intuition. Till now, informa­
tion concerning developments in the field of polymer technology was
dispersed among numerous scientific and technical journals and re­
ports, and there has been no single source or reference book in which
the present state of the art of molding technology has been updated.
To a certain degree, this has complicated proper utilization of these
recent developments by those engaged in the molding industry.
Accordingly, this book is an attempt at presenting a detañed, up-
to-date account of theoretical and experimental investigations con­
ducted in the molding area by different research groups. Various as­
pects of the molding operation are considered, including simulation
and experimentation of flow dynamics in the cavity and delivery sys­
tem, rheology and processing-property relationships in molded parts,
and computer-aided design and manufacturing of molds. Specific fea­
tures as well as problems related to the molding of thermoplastics,
rubbers, and thermosets are discussed. This volume is not simply a
collection of various papers but a unique volume that combines vari­
ous aspects of the molding operation.

Avraam Í . Isayev

III
Introduction

Generally speaking, this book can be divided into three succes­


sive sections. The first section, which includes the first three chap­
ters, is primarily concerned with various aspects of molding of ther­
moplastic polymers. Chapter 1 is devoted to computer simulation of
cavity filling and flow in the delivery system, based on an inelastic
formulation of the rheological behavior of thermoplastic polymeric
fluids, coupled with heat conduction, convection, and viscous-heat­
ing effects in the energy equation. Both one- and two-dimensional
approximations of flow dynamics are given for simulating the filling
of single- and multi-gated thin cavities with and without obstacles.
Simplified methods for cavity filling simulation of complex cavities are
described. Some typical representative comparisons of simulation with
experimental runs are presented.
On the other hand, Chapter 2 is relevant mainly to flow problems
in the juncture regions of injection-molding systems. Typical visco­
elastic effects encountered in the flow region involving abrupt or grad­
ual contractions or expansions of injection-molding systems are pre­
sented, along with various attempts at simulating entry-flow problems
by using inelastic and viscoelastic formulations. In particular, vari­
ous integral and differential-type viscoelastic models are used. Some
predicted results are compared with experimental data and certain im­
plications for juncture-type flow in mold design are drawn.
The following chapter is devoted to consideration of theoretical and
experimental aspects of the orientation, residual stresses, density,
and shrinkage development in injection-molded thermoplastic products
as affected by molding conditions, thermal, rheological and relaxation
Introduction

properties of the polymeric materials and equation of state of polymers.


Experimental techniques for their characterization are described, to­
gether with physical models developed in the modern literature.
The second section, which comprises Chapters 4 to 7, is concerned
with the molding process accompanied by chemical reaction of poly­
merization or crosslinking. In particular. Chapter 4 deals with funda­
mental aspects of thermoset molding including resin characterization
and process simulation. Models of isothermal and nonisothermal ki­
netics, rheology, and thermal properties of unfilled and filled ther­
moset resins are considered. Experimental techniques suitable for
characterization of kinetic and rheological behavior of thermosets are
described, together with thermoset molding equipment. Modeling and
computer simulation of various stages of the thermoset molding pro­
cess are considered and results of model predictions compared with
experimental data.
Chapters 5 and 6 are devoted to the molding technology of elasto­
mers. In particular. Chapter 5 is primarily concerned with the rhe­
ology and relaxation in raw elastomeric materials and the chemorheol-
ogy of elastomers during curing and describes how this information
can be utilized in computer-aided mold design and in determining ap­
propriate molding conditions. In addition, some molding aspects of
rubber-laminated composites are considered. Chapter 6, on the other
hand, gives a comprehensive review of the literature devoted to in­
jection molding of rubber compounds, including injection-mol din g ma­
chines, rheology, scorching, and moldability of rubber compounds,
various modeling aspects of the process, and characterization of ru b ­
ber moldings. Suggestions for future work are outlined in order to
develop a scientific basis for injection molding of rubber compounds
in which the process is considered as one integrated system.
Chapter 7 combines a comprehensive review of the state of the art
together with an extensive investigation devoted entirely to the com­
pression molding of unfilled and filled crosslinking and thermoplastic
polymers. Starting with process definitions and a description of com­
pression molding, this chapter deals with various aspects of experi­
mentation and with the simulation of squeezing type flows in the cavi­
ties of arbitrary shape using finite-difference and finite-element meth­
ods. Strongly nonisothermal flow and various rheological effects as
well as curing behavior and heat transfer on different stages of com­
pression molding are elucidated. Fiber orientation due to flow in com­
pression-molded short-fiber composites are experimentally determined
and tested against well-known and recently proposed theoretical mod­
els.
The third section of this book, consisting of Chapters 8 and 9,
deals with computer-aided design, engineering, and manufacturing
of molds. In particular. Chapter 8 presents principles of optimal
Introduction

mold-cooling-system design based on an extensive use of computers


which models the thermal balance in terms of a composite of two-di­
mensional cyclically steady heat transfer. Significant information
about cooling-system arrangements in the mold, properties of cool­
ing reagents, and thermal properties of plastics and mold materials
is given. An emphasis has been placed on the methodology of laying
out a cooling system in order to optimize the mold design and minimize
the cycle time. The succeeding chapter summarizes the state of the
art in computer-aided mold design and manufacturing. Concepts of
geometric-modeling languages are given which are commonly used for
mold design and manufacturing. Procedures for interactive mesh gen­
eration in cavity filling simulation, mold-assembly design via interac­
tive graphics, and principles of code generation for numerical control
of machining the mold components are discussed. In addition, mold-
cooling-system design by means of the 3-D boundary element method
is described.
This book is a first attempt at combining available information from
various engineering disciplines in order to formulate the fundamentals
of science-based technology for injection and compression molding of
polymers. Readers will decide whether this attempt has been success­
ful in providing a basis for industrial applications.
Contributors

Richard J. Ambrose Corporate Research Center, Lord Corporation,


Cary, North Carolina
Cornelius A. Hieber Sibley School of Mechanical and Aerospace En­
gineering, Cornell University, Ithaca, New York
Henry S .-Y . Hsich Corporate Research Center, Lord Corporation,
Cary, North Carolina
Musa R. Kamal Department of Chemical Engineering, McGill Univer­
sity, Montreal, Canada
Avraam I. Isayev Polymer Engineering Center, The University of
Akron, Akron, Ohio
Michael E. Ryan Department of Chemical Engineering, State Univer­
sity of New York at Buffalo, Buffalo, New York
Kamar J. Singh* Corporate Research and Development, General
Electric Company, Schenectady, New York
Charles L. Tucker III Department of Mechanical and Industrial En­
gineering, University of Illinois at Urbana-Champaign, Urbana,
Illinois

*Current affiliation: Aircraft Engine Business Group, General Elec­


tric Company, Cincinnati, Ohio.

VI I
vtii Contributors

Ram K. Upadhyay Corporate Research and Development, General


Electric Company, Schenectady, New York
K. K. Wang Sibley School of Mechanical and Aerospace Engineering,
Cornell University, Ithaca, New York
V. W. Wang* Sibley School of Mechanical and Aerospace Engineering,
Cornell University, Ithaca, New York

*Current affiliation: Advanced CAE Technology, Inc. Ithaca, New


York.
Contents

Preface 111
Introduction iv
Contributors vii

M e lt-V is c o s ity C h a ra c te riza tio n and Its


A pp lication to Injection Molding
Cornelius A . Hieber
I. Introduction 1
II. Melt-Viscosity Characterization 2
A. Viscosity Fits Using Cross-Arrhenius Model 2
B . Master Plots 8
C. Incorporating Pressure Dependence 15
III. Simulation of Cavity Filling 23
A. Simple Geometries 23
B . Composites of Simple Geometries 36
C. Thin Cavities of Arbitrary Planar Geometry 44
IV . Post-Filling Stage 61
Appendices
A. Dependence of n upon Mw/Mn 78
B . Dependence of x * upon Mw /Mn 79
C. Dependence of on Temperature Level 80
D. Dependence of no upon Mw 85
E. Relationship between x* and Jg® 86
F. Juncture-Pressure Correlation 96
Contents

G. Melt-Fracture Correlation 101


H. Comparison of Cross and Carreau Models 104
I. Viscous-Heating Effects in Capillary Rheometry 107
J. Fitting Capillary-Rheometer Data Accounting for
Pressure Dependence and Viscous Heating 111
K. ’’Fountain” Effect 124
References 129

2. Flow o f Polymeric Melts in J u n tu re Regions o f


Injection Molding 137
Avraam I. Isayev and Ram K, Upadhyay

I. Introduction 137
II. Experimental Results for Two-Dimensional
Flow of Polymers 138
A. Flow-Through Channels with Contraction
or Expansion 138
B. Entry and Exit Flow Effects 152
III. Modeling Two-Dimensional Flow 176
A . General 176
B . Generalized Newtonian Fluid Models 179
C. Viscoelastic Models 186
IV. Predicted Results Compared with Experimental
Results 206
Implications for Juncture-Type Flow in Mold
Design 218
References 221

3. O rie n ta tio n , Residual S tre s s e s , and Volum etric


E ffects in Injection Molding 227
Avraam L Isayev

I. Introduction 227
II. Development of Orientation in Injection Molding 228
A. Experimental Techniques 228
B . Modeling Orientation in Terms of Birefringence 233
C. Thermal- and Flow-Induced Birefringence 242
D. Orientation in Molded Parts of Amorphous
Polymers 246
III. Development of Residual Stress in Injection '
Molding 257
A. Experimental Techniques for Residual-Stress
Measurements 257
B. Thermal- and Flow-Induced Stresses 264
C. Modeling of Residual Stresses 273
Contents XI

D. Effects of Processing Conditions on Residual


Stresses 285
E. Comparison of Flow , Thermal, and Residual
Stresses in Molded Parts 294
IV. Development of Density and Shrinkage in
Injection Molding 296
V. Concluding Remarks 319
References 322

4. Therm oset In jectio n Molding 329


Musa R. Kamal and Michael E, Ryan

I. Introduction 329
II. Thermoset Resin Characterization 330
A. Kinetic Characterization 330
B . Rheological Characterization 344
C. Thermal Properties 350
D. Effect of Fillers on Kinetics and Rheological
and Thermal Properties 357
III. Thermoset Injection-Molding Simulation 359
A. Injection-Molding Systems: Equipment
and Resins 359
B . Mathematical Modeling and Computer Simulation 363
IV. Conclusions 372
References 372

5. Rheological B ehavior and Molding Technology


o f Elastomers 377
Henry S .-Y . Hsich and Richard J, Ambrose

I. Introduction 377
II. Relaxation Spectra of Elastomers 379
A. The Rubbery State 381
B . A Hybrid Model for Mechanical Spectra of
Filled and Unfilled Elastomers 381
C. Composite Mechanics of Filled Elastomers 389
III. Flow Rheology of Elastomers 392
A. Viscosity Theory of Polymers 393
B . Viscosity Studies of Elastomers and Shear-Flow-
Induced Structural Changes of Molecules 395
C. Effects of Filler on the Viscosity of Elastomers 398
IV . Chemorheology of Elastomers 402
A. Chemical Reactions of Cure 403
B . Kinetic Model of Cure as an Aid to Property
and Processing Control 409
XII Contents

Molding Technology of Elastomers 421


A. Elastomers as Engineering Materials 421
B . Elastomer Molding 422
C. The Use of Elastomer Rheological Properties in
Molding Laminated Elastomer Composites 424
D. Summary 430
References 431

6. Injection Molding o f R u b b e r Compounds 435


Avraam I, Isayev

I. Introduction 435
II. Progress in Injection Molding of Rubber 436
A. Introductory Remarks 436
B . Injection Machines 437
C . Rheology of Rubber Compounds 439
D. Scorching and Moldability 444
E. Modeling of the Injection-Molding Process 446
F. Characterization of Rubber Molding 459
III. Suggestions for Future Work 464
A . General 464
B . Characterization of Rubber Compounds 464
C. Experimental Investigation and Modeling
of the Molding Process 466
IV. Concluding Remarks 473
References 473

7. Compression Molding o f Polymers and Composites 481


Charles L, Tucker III

I . Introduction 481
A. Technological Significance of Compression
Molding 481
B. Critical Issues 482
C. Scope of This Chapter 483
D. Stages in the Compression Molding Process 483
II. Flow Models for Thin Cavities 485
A. Approximations for Thin Parts 485
B . Generalized Hele-Shaw Flow Model 488
C. Lubricated Squeezing Flow Model 497
D. Controlling Flow in the Mold 504
III. Heat Transfer and Curing 504
A. Heat Transfer During Mold Filling 506
B. Kinetics of the Curing Reaction 510
C . Heat Transfer and Curing 516
Contents

D . Thermal Design of Molds 520


E. Residual Stresses 523
IV. More Complex Rheological Effects 528
A. Effect on Mold Filling Pattern 528
B. Effect on Mold Closing Force 533
V. Fiber Orientation 539
A. Phenomenology 539
B . Models for Concentrated Suspensions 542
C. Predicting Fiber Orientation 547
D . Measuring and Characterizing Fiber Orientation 554
E. Fiber Damage During Flow 559
VI. Summary and Conclusions 560
References 562

8. Design o f Mold Cooling System 567


Kamar J. Singh

I. Introduction 567
II. Heat Transfer in Molds 568
A . Conduction 569
B . Convection 570
C . Radiation 574
III. Cooling of Plastic 575
IV. Heat Flow Through Molds 580
V. Cooling System 581
A. Cooling Lines 581
B . Heat Pipes 581
C. Bubblers 583
D. Baffles 585
VI. Coolants 586
VII. Design of Cooling System 587
VIII. Optimum Parameters 572
IX. Defects in Plastic Parts 593
A . Warpage 594
B . Residual Stress 595
Computer-Aided Design of Cooling Systems 597
A. Computer Software 598
B . Example Problems 598
References 605

9. C o m p u ter-A id ed Mold Design and M a n u fa c tu rin g 607

K. K. Wang and V . W. Wang

I. Introduction 607
A. Typical Mold Configurations 607
XIV Contents

B. Conventional Mold Design and Manufacturing


Methods 608
C. Computer-Based Mold Design and Manufacture 609
II. Overview of InteractiveComputer Graphics and
Geometric Modeling 610
A. Display Devices 611
B . Input Devices and Techniques 613
C. Geometric Modeling 615
III. CAD/CAE/CAM Software for Injection Molding 620
A. Runner System Design 621
B. Mold-Filling Simulation Programs 634
C. Cooling-System Design Program with Three-
Dimensional Boundary Element Method 649
D. Mold Assembly Design 657
E. Computer-Aided Mold Manufacturing 659
F. Integrated System for Mold Design and
Manufacture 662
IV. Concluding Remarks 667
References 668

Author Index 671


Subject Index 687
1
Melt-Viscosity Characterization and
Its Application to Injection Molding

C orneliu s A . H ieb er
Sibley School of Mechanical and Aerospace Engineering
Cornell University
Ithaca, New York

I. IN T R O D U C T IO N

The scientific investigation of the injection-molding process can be


traced back at least to the pioneering work of Gilmore and Spencer
[1,2] at Dow, during the early 1950s, and to Ballman and co-work­
ers [3,4] at Monsanto, during the late 1950s. A finite-difference
formulation for numerical simiilation was presented by Pearson [5] in
the mid-1960s but not implemented until the early 1970s in the well-
known works by Kamal and Kenig [6] and by Gogos and co-workers
[7 ,8 ]. Since then, the application of computer simulations to the in­
jection-molding process has clearly intensified, and its commercializa­
tion, pioneered by Austin [9] , has resulted in viable computer-aided
engineering (CAE) tools for mold design.
One main objective in this chapter is to present a detailed descrip­
tion of the basic elements involved in simulating the cavity-filling stage
of the injection-molding process. This is done in Section III, with the
formulation being given in terms of a fluid whose shear viscosity can
have an arbitrary dependence upon shear rate, temperature, and pres­
sure . Accordingly, a second objective is to present detailed results
concerning the characterization of such viscosity. This is done in
Section II. Thirdly, although a general viscosity model may not al­
ways be needed for the cavity-filling stage. Section IV shows that
such a model is essential when extending the simulation to the post­
filling stage.
2 Hieber

II. M E L T -V IS C O S IT Y C H A R A C T E R IZ A T IO N

A. V isco sity Fits Using C ro s s -A rrh e n iu s Model

In modeling the flow dynamics in injection molding, the single most


important polymer property is the shear viscosity n- Accordingly,
this section is concerned with the characterization of ri> making ex­
tensive use of a viscosity data bank which has recently been devel­
oped by the Cornell Injection Molding Program (GIM P). Currently,
this data bank consists of approximately 300 sets of viscosity mea­
surements with corresponding fits. Many of these sets are from the
open literature, such as the well-known compendium of Westover [10]
and the recent book Rheologie [11]. Arranged according to generic
material, viscosity measurements from the following sources have been
fitted: ABS [1 1 -1 4 ], HOPE [10,11,15-34], LDPE [10,11,15-19,21-
23,26,27,31,35-40], NYLON [10,11,41-45], PC [41,46-48], PMMA
[10,11,49-51], PP [11,17,21,23,25,26,28,31,44,45,52-55], PS [10,11,
14,17,21-26,31,38,41,54,56-69], PVC [10,11,70-74], SAN [10,11,
14,31,75,76].
Each data set has been fit by using a simplex method [77] and the
following four-constant (n, t*, B, T]^) model, in which the depen­
dence upon shear rate (Y ) is of a modified Cross [78—81] form:

no(T)
n=
.l-n
( 1)
1+

with rio(T) based upon an Arrhenius-type dependence:

no(T) = B exp (2)

That is, rio denotes the zero-shear-rate viscosity, with being a


measure of the temperature sensitivity of no characterizing
the shear-stress level at which n is in transition between the New­
tonian limit no^^^ power-law asymptote corresponding to large Y:

n , ^ vl-n -n-1 _ .
n ^ Tio ( t^*) y = A exp ( ^ ) (3)
t” - ‘

where

T =nT,, A = (4)
a b

In particular, note that the temperature sensitivity is reduced by the


factor n in the power-law region.
Melt-Viscosity Characterization

o
0
CO
*
E
o
E
CD

10 10 10 10
7 a (sec "' )

Fig. 1 Sample viscosity data set from GIMP data bank for PS (Styron
678U/DOW) by Chung [82] on Sieglaff-McKelvey capillary rheometer
(L/D = 38, D = 0.050 cm) at 180°C ( o ) , 200^C ( □ ) , and 230°C ( a ) .
Curve fits correspond to (8, 2, 9).

Figure 1 shows a typical plot from the CIMP data bank. In this
case the results are in terms of apparent shear rate and apparent
shear viscosity,

Y = 8 ^ ^ (5)
\ ®D IT
and

(6)
4 Hieber

where U, D, and Q denote the bulk velocity, capillary diameter, and


volumetric flow rate, respectively, with being the wall shear stress
as determined from the overall pressure drop for a tube of length L :

Ap (7)
w 4L

In particular, the curves in Fig, 1 correspond to the same functional


form as (1) but in terms of (Y^, coordinates, namely.

t1o(T )
(8)
1 + (n

with rio(T) given by (2) and

(n , T^*, B , = ^0.312, 2.16 x iQ®


cm2~ >

2.26 X 10’ ^ ---- , 12,500 k ) (9)


cm*sec /

which fits the data with an rms (root-mean-square) deviation of 7.3%.


As a simple means of transforming from (Y^, t\q) to (Y, n) coordi-
n ates~i.e., from (8, 2) to (1, 2)—the customary Weissenberg-
Rabinowitsch correction [83] can be fairly accurately effected by
merely setting

n/(l-n)
T* = f , ( n ) x x a ^ ( 10)

and maintaining n, B , unchanged. That is, the corrected flow


curves in the present case are then given by (1, 2) with

dyne
(n , T * , B, Tjj) = ^0.312, 1.77 x lo^
cm^

2.26 X 10’ ^ —
cm •sec
12,500 k )
/ ( 11)

This is illustrated in Fig. 2(a), in which the solid curve corresponds


to the 180°C curve from Fig. 1, given by (8, 2, 9) at T = 453 K, and
the dashed curve is the resulting transformed curve in ( t , n) coor­
dinates given by (1, 2, 11) at T = 453 K. For comparison, the sym­
bols indicate the transformation of individual points from (ta> co­
ordinates (solid) to (Y j n) coordinates (open) via the standard Weis-
senberg-Rabinowitsch correction [83] :
Melt-Viscosity Characterization

or r (sec“ ')
Id' 10" 10 10*

10”

10^

3x10*

F ig . 2 (a ) Illustration of shift from (Ya, na) fo (Y, n) coordinates


due to Weissenberg-Rabinowitsch correction; see text for definition
of symbols and curves, (b ) Plot of f^Cn) = (curve 1) based
on (10) and f 2 (n ) = A/Aa (curve 2) based on (14).

9 In n
______ a
9 In Y
( 12)

and

T
W
(13)
n=T
with (j) evaluated by basing na(Ya» upon (8, 2, 9). Note that the
transformed points are very well described by the dashed curve. Fur­
ther , although individual data points may undergo a significant shift,
the shift in the flow curve is still small, being at most about 13% in the
power-law region. This is seen from Fig. 2 (b ), in which curve 2 cor­
responds to the shift in the power-law region where the multiplicative
Hieber

F ig . 3 Values of n from viscosity fits in CIMP data bank according to


generic material: HDPE ( o ) , PP ( a ) , l DPE ( a ) , NYLON (▼ ), PC ( □ ) ,
SAN ( ♦ ) , PS ( 0 ) , ABS ( • ) , PVC ( v ) , and PMMA ( ■ ) . PS subsets
include NMWD (<►), HIPS (❖ ) and glass-fiber filled ( ❖ ) ; when known,
rigid ( v ) or plasticized ( v ) PVC has been denoted. Points lying off
graph: two O (n < 0) , one ▼ (n > 0. 5) and five o (n > 0. 5).

constant in (4) is related to the corresponding constant in the limit­


ing behavior of ria(^a» T ) by

A = f2(n) X Ag (14)

Figures 3—5 show resulting values for n, x*, and T]r) for some 300
viscosity data sets contained in the CIMP data bank. In particular,
these results are shown according to generic material, ranging from
the least temperature-sensitive high-density polyethylene (HDPE) to
the most highly temperature-sensitive polyvinyl chloride (PVC ) and
Melt-Viscosity Characterization

- 1— I I I I M I--------r--- \— I—f r i I 11-------- 1---- 1— r r -r I T--- 1


— I—I I I I I I
W
T V ▼ ^ V WV V v9 v

• • •

♦♦

no am o a

▼ ▼▼▼ ▼

AA A
1 -^

O O O CO 00 O Oo8:P 00 &^Oo£oOOO^ 00 o
I I_i I I I I ____I___I__1 I 1
.1I II______I___I__I_I !.. Mil_____ I___l__I_I III I 1
2x10^ 10^ 10® 10® 10^ 2x10^
T* (dyne / cm^ )

Fig. 4 Same as Fig. 3 but for t*; data sets lying essentially entirely
in power-law region have been omitted.

polymethylmethacrylate (PMMA), as can be seen from the values for Tb


in Fig. 5. On the other hand. Fig. 6 presents corresponding results
for no = no(T) based upon (2 ), with T denoting a typical processing
temperature for each generic material. Corresponding results in terms
of a characteristic relaxation time, 0 e tìo/ t *, are then shown in Fig. 7.
For those data sets lying essentially entirely in the power-law re­
gion, results for t * and B will be ill defined [although the product
will be well defined, corresponding to the multiplicative
constant in the power-law limit (3, 4 )]; accordingly, such data sets
have not been included in Figs. 4, 6, and 7. On the other hand,
these same data sets can be used to generate values for Tb = T^/n,
since both and n will be defined. Of course, if the data set is
only at one temperature, then no information on the temperature sen­
sitivity in terms of Tb (or Ta) can be obtained.
An analysis concerning the spread exhibited by some of these fig­
ures is presented in the appendixes in terms of the dependence of
Hieber

F ig . 5 Same as Fig. 3 but for T]^; data sets at only one temperature
have been omitted.

n upon Mw/Mn (Appendix A ) , t * upon M^/Mn (Appendix B ) , T]^ upon


temperature level (Appendix C ), and fjo upon (Appendix D) . As
shown in Appendix E, 1/t * is related to the steady-state compliance,
Je° 5 such that is related to the recoverable strain and might
therefore be used to correlate such effects as juncture pressure
losses (Appendix F) and melt fracture (Appendix G ) . In addition,
T* is an important quantity in developing master plots for the shear
viscosity, as shown in the next section.

B. M aster Plots

One of the more popular forms for collapsing viscosity flow curves in­
to master plots is that generally attributed to Vinogradov and Malkin
[84—86] in terms of n/no versus rio^* particular, it was found in
Melt-Viscosity Characterization

-i------1
---1—r-n~iHI------ 1
--- 1
—i i ri i ii------ 1--- 1—r -r-r i i 1 — I— I I I 1 1 II

• . . I m m§ m

9 V V V w ^ v V vv

• ••• • t

cx> ^

cPog

▼▼ ▼ ▼? ▼ w ▼▼

A AA^ jkXVsA /SiA A A A

A a U aa a

O OOOD & 0 ^ O O O O ^ 8 8 COO o%

_l__ I.. .........I_____1_ _ l____ I I 11 ■ I l l


2xIQ2 IQ3 10'* 10® 10^ 10^

Vcm sec/

Fig. 6 Same as Fig. 3 but for fio = B exp (Tfo/T) where T = 190^C
(463 K) for HDPE, LDPE, and P V C ; 210°C (483 K) for PP, SAN, PS,
ABS, and PMMA; 270°C (543 K) for NYLON and PC; data sets lying
essentially entirely in power-law region have been omitted, as well as
those corresponding to only one temperature (unless it happens to co­
incide with T ) .

[84] that the viscosity of many different polymers [polyethylene (PE) ,


polypropylene (P P ), polystyrene (P S ), polyvinyl butyral (P V B ), and
poly isobutylene (P IB )] as well as natural and synthetic rubber could
be correlated to within a factor of 2 above and below (see Fig. 17 of
[87]) by means of the following master curve:

[1 + C i(n o Y )“ " + C ain oY )“ " ] ’ '- (15)


no
10 Hieber

— I------ 1— r - r | ------ 1-------1— r-TT------ 1


■ i • i 1 1 I 1 I 1 t 1 1 I 1 1 1 1 11 1 111
■ ■ ■ ■ a t ■ a f aaa 1 . a

V VV 17 ^ V % V T

• i t • •• •

« □ □ n o□
□ □ □ □

f ▼w ▼ \ ▼

A AA aA A A A

A * ^ # 1 6 4 A f ^

0 8 00 0 0 ot?8ooo8oo 0 0 0 0 0 0 0 0 00
0
_____ 1_____ 1___ L..I 1___ J____1__1 iJ ___ j-------1—1 j-J____1------- 1__t i l ____1____ 1__i.. 1 1____ 1____1__« « 1 « I I I
10“ 10“^ 10 “ I0“L I 10 10* lO’*
^(sec)

Fig. 7 Same as Fig. 6 but for 0 = fjo/T*.

where

mi = .355, m2 = 2mi, Ci = .00612 (cm^/dyne)^^

€ 2 = .000235 (cm^/dyne)”^^ (16)

In later work by Kalinchev et al. [41,88], it was found that materials


such as nylon and polycarbonate (P C ) are not well correlated by (15,
16). In particular, viscosity measurements for 10 different commercial
grades of PC were found [88] to correlate within ±15% of (15) if the
constants are taken as

mi = .343, m2 = 2mi, C 3 = .00274 (cm ^/dyne)”^^.

C^ = .0000274 (emVdyne)"*^ (17)


Melt-Viscosity Characterization 11

Fig. 8 Master plot of cumulative data for PS-BMWD; solid curve based
on (1) with n = 0.25 and t * = 2.7 x 10^ dyne/cm^; dashed curve cor­
responds to (15, 16).

For this present work, (1) expresses n/no a function of not


given n and t *. Accordingly, for each data set in Section A (except
those that lie essentially entirely in the power-law region) we can use
the corresponding four-constant (n, t *, B, Tfo) fit to transform each
data point into (not» il/no) coordinates by calculating no based on (2 ).
In turn, by combining all such transformed data points for a given
generic material, we can then generate corresponding best values for
n and t *. This has been done for the various materials presented in
Section A, with illustrative plots being shown in Figs. 8 and 9 for PS-
BMWD and PC, respectively. Table 1 summarizes the resulting best
12 Hieber

Vo ' y (dyne/cm**2)

Fig. 9 Master plot of cumulative data for PC; solid curve based on
(1) with n = 0.17 and t * = 6.4 > 10^ dyne/cm^; lower and upper
dashed curves correspond to (15, 16) and (15, 17), respectively.

values of n and x* for each generic material, with Fig. 10 showing


the corresponding master curves.
In particular, the PS-BMWD results in Fig. 8 tend to lie system­
atically above (15, 16), although generally within the scatter noted
in [84] (or see Fig. 17 of [8 7 ]). On the other hand, the PC results
in Fig. 9 agree fairly well with (15, 17), although lying low at the
larger values of rioY* In this regard, it is not explicitly indicated
whether the results in [88] were Weissenberg-Rabinowitsch corrected.
If not, then the dashed curve based on (15, 17) should be shifted
down by about 13% [according to curve 2 in Fig. 2 (b )] in the large
Melt-Viscosity Characterization 13

T a b le 1 Master-Plot Constants Based on (1) for Various


Generic Materials^

#Data Rms dev


Material sets n, T* (dyne/cm (%)^

ABS 9 (0.27, 5.6 X 10^) 21


HOPE 47 (0.36, 2.2 X 10^) 49

HIPS 11 (0.26, 3.5 X 10^) 17

LDPE 47 (0.34, 1.6 X 10^) 24

NYLON 16 (0.27, 2.2 X 10^) 15

PC 9 (0.17, 6.4 X 10^) 14

PMMA 18 ( 0 . 20 , 1.02 X 10®) 21


PP 52 (0.31, 1.6 X 10®) 20
PS-BMWD 55 (0.25, 2.7 X 10®) 17

PS-NMWD 19 ( 0. 12 , 1.05 X 10®) 12


PVC-PLAS 6 (0.24, 3.8 X 10®) 22
PVC-RIGD 12 ( 0. 21 , 1.7 X 10®) 30

SAN 10 (0.26, 6.6 X 10®) 15

'^All results have been Weissenberg-Rabinowitsch cor­


rected. Data sets lying essentially entirely in power-
law region have been omitted.
^»Relative to average of experimental and fitted values.

rioY range of Fig. 9. A further complication in this large not i^ange


is a definite possibility of significant viscous-heating effects (due to
large shear-stress levels combined with a fairly temperature-sensitive
viscosity). Although the capillary diameter employed in [88] is not
explicitly given, smaller diameters would result in smaller viscous-
heating effects (see Appendix I for further details) and hence in a
higher (more correct) viscosity curve.
For more comparisons with the present results, note that Robinson
[89] (as reported in [90], pp. 155 and 156) has fitted the PS-NMWD
data from [58] and [59] with the same shear-rate dependence as (1)
u Hieber

5^
\
P-

10 10

'Ho * y (dyne/cm**2)

Fig. 10 Master plots for various amorphous (a ) and semicrystalline


(b ) generic materials based upon (1) with (n , t * ) as given in Table
1; lower and upper dashed curves correspond to (15, 16) and (15,
17), respectively.

and obtained a value of n = 0.11, which agrees well with the corres­
ponding result in Table 1. For PMMA, Panova et al. [91] (as reported
in [86], p. 176) have also used the same shear-rate dependence as
(1 ), with the resulting fit corresponding to n = 0.20, in excellent
agreement with Table 1, and

-1,69
X* = 7.48 X (18)
cm^ \ M
\ n
Melt-Viscosity Characterization 15

10

(b )
.......... .

’ -V
10
^ S a%
••• %N.
*Sa!9 o.
* aa’ 2o ’ %

••., A“ X T T_
% \

10 o HOPE
A LDPE \* V ° ° o ■
▼ NYLON \ *X$A V
A PP
▼ P V C -P L A S
V P V C -R IG D ■■■• V :

........... 1
________ ^
____ 1_ jj _ ii_
10
5
10 10 10 10 10 10

Vo ' y (dyne/cm**2)

Although molecular-weight information was not available for the 18


PMMA data sets used in Table 1, the corresponding value for x* of
1.02 X 10® dyne/cm ^ would translate on the basis of (18) to M^/Mn-
3.3, which seems to be a reasonable representative value for such
commercial grade PMMA samples.

C. In co rp o ratin g P ressu re D ependence

One advantage in using a viscosity model in which n/no is expressed


as a function of rioY is that the pressure dependence of n can then
be incorporated solely through the pressure dependence of no- This
has been shown experimentally for PS in [92] and, subsequently, for
PMMA in [50] and PC in [47]. That is, we can write
T6 Hieber

n(Y, T, p) = rio(T, p) X f ( 0 , 5 = noY (19)

where f (^ ) should become unity for small ^ and be proportional to


^n-1 power-law limit at sufficiently large 5*
In dealing with n(t> T, p) , we introduce for convenience the fol­
lowing quantities relating to its temperature and pressure sensitiv­
ities :

( 20)
^ /n ' ' Y>P 'T, p

and

, I d In rio\ g, , / 3 In n\ g„ ^ /3 In n\
( 21)
“• - ( 3P ® - ( 9P ^ - { 3p ) ' t, T

where x = riY is the shear stress. As applied to (19), it follows by


chain-rule differentiation that

d In f d In f
a’ = ao {, 8mT / ( 22)
dS ’ d?
Y. P

and

d In f , +/1£\ din f (23)


dC ’
'y. T

However,

T = riY = noY f ( C ) = g ( ? ) (24)

That is, T= t( 5) such that

(25)

Accordingly, within the framework of (19), it follows that

a” = ao and 3” = 3o (26)

That is, the temperature and pressure sensitivities of n(Y> T, p) at


fixed shear stress are the same as those of no(T, p) . On the other
hand,
Melt-Viscosity Characterization 17

(27)
y> p

and

(28)
(H ),, = ^- «
Hence,

d In f
= ao|l + (29)
d In 5

(cf. [93], [40], or [50], e . g . ) . In particular, if we base f(5 ) upon


( 1 ), then

1 -n
din f ^ (g/T*)
(30)
d In g 1 -n
1 + ( g/ x* )

such that

1 -nj
¡l+n(g/x*) U +n(5/x*)^
= ao (31)
1 -n ( ’ ' ‘’i n r
( l+ (g / T * )

Accordingly,

X. ao, 3’ 3o g/x* 0 (32)


and
^ nao, 3’ n3o as g/x* ->• (33)

Such limiting behaviors will be satisfied by any reasonable f ( g ) —i.e. ,


any such function that is essentially unity for sufficiently small g and
which is of order g^“^ for sufficiently large g. Accordingly, we see
that, whereas and 3 ” are independent of shear-stress level, and
3’ continually decrease with increasing shear rate. In fact, the behav­
ior of aJ in (33) is directly related to (4 ), namely that = nTi^.
Within the framework of (19) with f (g ) based on (1 ), it then be­
comes necessary to specify rio(T, p ). One evident possibility is to
extend ( 2 ) as follows:

rio(T, p) = B exp exp(gp) (34)


Ì8 Hieber

which therefore results in a five-constant (n, t *, B , T^, 3) model


for which

-Z t (35)
^0 - iji
rp2 2>

Alternativelyj one might employ a Williams, Landel, Ferry (WLF)-type


[94] relation:

no(T, p) = D i e x p { - (36)

where T * might be taken as the pressure-dependent glass-transition


temperature, as in [ 95] and [ 96] :

T * = T g ( p ) = D2 + D3P (37)

This gives a seven-constant (n , t *, D^, D 2 , D 3 , A^, A 2 ) model for


which

A1 A 2D3
Oio =
(38)
[A 2 + (T -T * )]^ [A 2 + (T - T * )]'

In particular, then,

(39)

with representative experimental values for T g (p ) being cited in [96]


(e . g . , 0.03°C/bar, 0.043°C/bar, and 0.023°C/bar for PS, PC, and
PMMA, respectively). In turn, these results compare fairly well with
Van Krevelen ([9 7 ], p. 348), who proposes that the less readily avail­
able 3o might be calculated from ao based upon 3o/cto “ 0.04®C/bar.
One shortcoming with this model, however, is that it indicates that
3 o is both temperature and pressure dependent, whereas there exists

fairly strong experimental evidence in the literature that 3 o is essen­


tially independent of pressure (up to 1500 bar for PS in [92], e . g . ,
and up to 1000 bar for PMMA in [5 0 ]). That is, at a fixed tempera­
ture no should have a purely exponential dependence upon pressure.
Such a deficiency in (36, 37) was noted in early work by Ferry and
Stratton [98] and O^Reilly [99]. Most simply, it can be remedied by
letting A 2 also be pressure dependent, namely

A2 — 2 3P (40)
Melt-Viscosity Characterization 19

such that the denominator in (36) becomes independent of pressure.


This extended seven-constant (n , x*, Di, D 2 , D 3 , A^, A 2 ) model then
gives

- Ai(Á2+D3P) A1 D3
Oi-O —
• 9 3o (41)
[Á2+(T-D2)]2 A2+CT-D2)

Concerning experimental results for viscosity pressure dependence,


one of the most extensive sources for polymer melts is [ 1 0 0 ] , which in­
cludes rio measurements for HDPE, LDPE (low-density polyethylene),
PP, PS, PMMA, and PC at various temperatures and pressure levels.
These results were obtained with a pressurized rotating viscometer,
as described in [101]. Such an instrument has also been used to gen­
erate results for PS [92] and various grades of PMMA [50] and HDPE
[102]. On the other hand, results for PS have been obtained in [103]
with translating pressurized coaxial cylinders. In addition, measure­
ments have been made with the pressurized double-piston capillary
rheometer of Maxwell and Westover [104,105], including results for
PP, PS, LDPE, and HDPE in [105], PC in [47], and ABS (acryloni­
trile-butadiene-styrene), PS, HDPE, PP, and SAN (styrene-acrylo­
nitrile) in [106]. Also related to viscosity pressure dependence are
numerous instances of nonlinear Bagley plots, such as for PS [95,107].
LDPE [35,40], HDPE [108], PC [46], PMMA [51], and PVC [74].
Comparisons between the foregoing data and fits based upon (38)
and (41) are shown in Figs. 11, 12, and 13 for PS, PC, and PMMA.
Experimental results for other materials are listed in Table 2. In par­
ticular, curves 1 and 2 in Figs. 11—13 are based upon (38) at pres­
sures of 1 bar and 500 bar, respectively, with (A^, A 2 , D 2 ) based
upon the best local-T^^ fits from Appendix C and D 3 based upon the
values for T g (p ) available from the literature. In all three cases,
curves 1 and 2 describe both the level of 3 o a^^d its temperature sen­
sitivity quite well. On the other hand, we repeat that the difference
between curves 1 and 2 represents a shortcoming of the model be­
cause experimental results generally indicate that 3 o is independent
of pressure. Curve 3, based upon (41), is independent of pressure,
but it tends to underestimate the temperature sensitivity of 3 o-
addition, the chosen values for D 3 are decidedly less than the cor­
responding values for T g ( p ) , such that the physical meaning of D 3
becomes unclear in this case.
The results in Figs. 11—13 indicate a definite decrease in 3o with
increasing temperature, but experimental results (Table 2) for semi­
crystalline materials such as HDPE and PP indicate no such evident
trend. In turn, this is directly analogous to the local-Tb results in
Appendix C, where Tt> (= T^ao) shows a definite decrease with in­
creasing temperature for PS, PC, and PMMA but not for PP. As
20 Hieber

Fig. 11 Variation of 3 q temperature for PS. Experimental re ­


sults from Ref. 92: □ , a , h , e ; from Ref. 95: ▼, from Ref. 100:
A ; from Ref. 103: from Ref. 105: ◄; from Ref. 106: ; from
Ref. 107: • . Curves 1 and 2 correspond to pressures of 1 bar and
500 bar based upon (36, 37) with (A^ , A 2 , D 2 ) = (27.2, 60.0°C,
100°C), corresponding to best fit in Fig. 46 of Appendix C, and
D 3 = 0.034°C/bar, based upon experimental results for T g (p ) in the
literature (e . g . , 0.036°C/bar [109] and 0.031°C/bar [110]); curve 3
based upon (36, 37, 40) with (A i, A 2 , D 3 ) = (27.2, 60.0°C, 100°C)
and D 3 chosen empirically as 0.023^C/bar. Results > ► and
• obtained via method described in Appendix J.

indicated in Appendix C, however, both the relatively low level of Tb


and its weak variation with temperature for PP are compatible with a
WLF representation in which T -T g is large. Although the WLF model
is typically considered mainly applicable to T -T g between 0 and 100°C,
useful correlations can still be obtained above this range, as has been
vividly shown by Menges and co-workers [114,115] in terms of ao v e r­
sus T -T g up to 350°C.
Due to the temperature dependence of 3o exhibited in Figs. 11—13,
it follows that a 0 will be pressure dependent. That is, from the def­
inition of ao and 3 o in ( 2 0 ) and ( 2 1 ) it follows that
Melt-Viscosity Characterization 21

Fig. 12 Variation of temperature for PC. Experimental re­


sults from Ref. 46: ▼; from Ref. 47: ♦ ; from Ref. 100: Curves
1 and 2 same as in Fig. 11 but with ( A i , A 2 , D 2 ) = (25.9, 61.2°C,
144°C), corresponding to best fit in Fig. 47 of Appendix C, and D 3 =
0.043°C/bar (cf. T g (p ) = 0.044°C/bar [99] and 0.041°C/bar [111]);
curve 3 same as in Fig. 11 but with (A ^ , A 2 ? D 2 ) = (25.9, 61.2°C,
1 4 4 0 c ) and D 3 chosen empirically as 0.019°C/bar. Results ▼and ♦
obtained via method described in Appendix J.

/ 3«o\ (42)
(» )/■ I ' p /t

since both quantities correspond to 9^(ln rio)/9T9p. Accordingly,


since Figs. 11—13 show that the left side of (42) is negative, it then
follows that ao should increase with pressure. This effect has not
been considered in the local-Tb fits in Appendix C. The implication,
however, is that data sets corresponding to higher pressure levels
should exhibit larger values for ag (or Tt>), thus resulting in a lar­
ger apparent experimental scatter if not accounted for.
To summarize these results, following Van Krevelen [97] we show
the results in Fig. 14 in terms of representative values of 3o ^0
for various generic materials at a typical processing temperature (T ,
as chosen in Fig. 6 ) for each. In particular, the values of 3o are
based on results in Figs. 11—13 for PS, PC, and PMMA and on
22 Hìeber

a>
c

■o
CSI
E
o
0>
'o
oa®

T{®C)

F ig . 13 Variation of 3o temperature for PMMA. Experimental re ­


sults from Ref. 50: a, b , b , h , h ; from Ref. 51: v , ▼; from Ref.
100: A. Curves 1 and 2 same as in Fig. 11 but with (A i, A 2 , D 2 ) =
( 38.2, 101. 3°C, 104°C), corresponding to best fit in Fig. 48 of A p ­
pendix C, and’ Ds = 0.024°C/bar (cf. T g (p ) = 0.023^C/bar [112] and
0.024°C/bar [113]); curve 3 same as in Fig. 11 but with (A i, À 2 , D 2 ) =
(38.2, 101. 3°C, 104°C) and D 3 chosen empirically as 0.017°C/bar. Re­
sult ▼ obtained via method described in Appendix J.

representative results from Table 2 for the remaining materials. On


the other hand, the values for ao = are based on the best fits
for PS, PC, and PMMA in Appendix C and on representative Tb re ­
sults in Fig. 5 for the remaining materials. With the exception of
PMMA, all results in Fig. 14 are well described by the dashed line,
which corresponds to 3o/oio - 0.056°C/bar.
In closing this section, note that since (34) corresponds to a con­
stant 3 o and fairly constant ao (= Tb/T^) , it follows that this model
will not be appropriate when the temperature dependence of ao or 3 o
may be significant. As applied to injection molding, this model proves
to be adequate for the filling stage, in which the bulk of the flow oc­
curs in a hot core region of essentially uniform temperature, but to­
tally inadequate for predicting pressure variations in the cavity dur­
ing the post-filling stage in Which significant cooling occurs through­
out, thus requiring a more refined representation such as one of the
seven-constant models described above. This is documented in the
following sections.
Melt-Viscosity Characterization 23

T a b le 2 Experimental Results for 3o of Polymer Melts


other Than PS, PC and PMMA^

Material T (^C) 3o (10 ^ cm^ dyne Source

ABS 190 (2.9) [106]

ACETAL 190 1.0 [100]


ACETAL 190 (1.4) [106]

HOPE 150-190 0.6 [100]


HDPE 160-200 0.9 [102]
HDPE 250 (0.9) [105]
HDPE 190 (2.7) [106]
HDPE 190 2.4 (1.7) [108]

LDPE 180 1.3 (1.0) [35]


LDPE 150 1.0-6.0 (5.9) [40]
LDPE 150-270 1.5 [100]
LDPE 250 (1.3) [105]

PP 190-230 1.5 [100]


PP 250 (1.6) [105]
PP 190 (2.2) [106]

PPO 260-300 5.4 [100]


PSF 257-277 4.5 [100]
PVC 170-200 3.8 (5.0) [74]
SAN 190 (3.9) [106]

^Values within ( ) obtained via method of Appendix J.

III. S IM U L A T IO N OF C A V I T Y F IL L IN G

A. Simple Geometries

In simulating the filling of simple geometries, the substantial related


literature, such as [5—8,47,116—119] , can be fairly well summarized
in terms of the following set of equations, shown here for the case of
a strip mold of width w and half-gap thickness b:

b
Q = w / u dy (43)
-b
24 Hieber

Fig. 14 Representative values for ao and 3o af T = T (see Fig. 6) for


ABS ( • ) , HDPE ( o ) , LDPE ( a ) , PC ( □ ) , PMMA ( ■ ) , PP ( a ) , PS ( o ) ,
PVC ( V ) , and SAN ( ♦ ) . Dashed line corresponds to 3o/oio 0.056°C/
bar.

8 / 3u\
(44)
3y ay)' 9x

/3T 8T 8^T
pc +u =k ny (45)
P\ai 8x th 8y"

where Q denotes the volumetric flow rate (typically taken as a speci­


fied constant), x the streamwise direction, y the gapwise (transverse)
direction, and t the time. Further, u denotes the velocity in the x di­
rection, n the shear viscosity, y = l8u/3y | the shear rate, p the pres­
sure, T the temperature, with p, Cp and kth denoting the melt density,
specific heat and thermal conductivity, all assumed to have constant val­
ues representative of the melt state.
In particular, Eq. (43) represents a balance of mass (or volume,
since density is assumed constant), whereas (44) gives the force bal­
ance in the streamwise direction between the viscous shear stress
Melt-Viscosity Characterization 25

and the pressure gradient. On the other hand, (45) indicates that
the change in temperature as one follows a fluid particle is due to
the net effect arising from the gap wise thermal conduction and the
viscous heating on the right side. Inertial effects have been omitted
from the left side of (44) due to the large viscosities typical of poly­
mer melts. For example, in terms of the characteristic Reynolds num­
ber. Re E pUb/n, we might typically have p ~ 1 g/cm^, U ~ 10 cm/
sec, b « 10“^ cm, and n ~ 10^ g/(cm*sec), giving Re « 10“^. On the
other hand, the thermal diffusivity, a = kth/pCp, of such polymer
melts is typically ~ 10“^ cm^/s. Hence, the characteristic Prandtl
number, Pr e v/a, where v e n/p « 10^ cm^/s is the kinematic vis­
cosity, will typically be about 10®. Accordingly, the Peclet number,
Pe E Ub/a = Re Pr, will be about 10®. That is, in terms of classical
fluid mechanics, the cavity-filling process corresponds to creeping
flow (Re « 1) but with the flow coupled to a temperature field char­
acterized by thin cold thermal boundary layers (due to Pe » 1) sur­
rounding a hot core region.
Finally, the governing equations (4 3 )-(4 5 ) must be supplemented
with a viscosity modeling. Typically, results in the literature have
been based upon a temperature-dependent power-law model such as

n = A exp E m(T) (46)

Here, however, we will also consider more general ri(y, T, p ) models


based upon the results in Section II.
Concerning associated boundary conditions on u (x , y , t) , T (x , y, t) ,
and p (x , t ) , it would typically be assumed that, at the inlet (x = 0),

u(0, y , t) = Uo(y, t) (47a)

and

T(0, y, t) = To (47b)

where Uo(y, t) is the corresponding fully-developed velocity profile


satisfying (43) and (44) for the particular n(y, T, p ) model evaluated
at T = T q, the injection melt temperature. The complication that Uq
may be time dependent is due to the possible pressure dependence of
ri and the increasing inlet pressure level as the melt advances further
into the cavity. If n is independent of pressure, as in (46), then
Uo = Uo(y).
Concerning conditions in the gap wise direction, it would typically
be required that at the cavity wall.
26 Hieber

u (x , b, t) = 0 (48a)

and

T (x , b, t) = T w (48b)

where is most simply taken as the coolant temperature. On the


other hand, assuming symmetry about y = 0, it then follows that

3u
(x , 0, t) = 0 (49a)
9y

and

AT (x , 0, t) = 0 (49b)
9y

Finally, along the advancing melt front, we neglect possible curva­


ture and associated transverse flow in this region (for more details of
this fountain-flow region, see Appendix K ), taking the front to be
flat and to advance uniformly based upon the bulk-velocity condition.
In addition, the temperature along this front, x = X m f(t), is taken to
be uniform and equal to the calculated centerline temperature at the
streamwise location immediately upstream of the front. Further de­
tails will be given in the finite-difference description to come. The
point here is that this ad hoc procedure is intended to model the fact
that the polymer melt along the front is supplied by the hot core re ­
gion .
Within the framework of the foregoing model equations, it is pos­
sible to manipulate the results into a simpler form. For example, in­
troducing A E -8p/9x and integrating (44) in the gapwise direction,
we get

9u ,
n 5 ^ = -A y ( S O)

where we have used (49a). Hence, since Y = •3u/9y for y >0 , it fol-
lows that

y = Ay
(51)
n

On the other hand, integrating (43) by parts and using (48a), we ob­
tain
Melt-Viscosity Characterization 27

Q = 2w i yY dy (52)
0

which, when combined with (51), gives

2
Q = 2wA J — dy (53)
0

Hence, introducing

_2
./ L a , (54)
0

which is a measure of the fluidity of the melt, we have

(55)
^ 2wS

In particular, if the viscosity is represented by the power-law


model, (46), the foregoing results can be simplified further. For ex­
ample, (51) can then be expressed as
1/n

(if) (56)

such that (52) then becomes


y , . 1/n ^b , . 1/n
(57)

where m(T) e mog(T) in (46) with g(To) e 1, such that

/T T \
g (T ) = exp (58)

Hence, introducing
K 1/n
y \ (59)
(ifc )
28 Hleber

we obtain

(60)
- ( 4 )”
An important point, which will be exploited in the subsequent numer­
ics, is that the dependence of the flow field on the temperature dis­
tribution factors out for the power-law model, being solely contained
in the factor $. Further, since 3 evidently also depends on the gap
thickness, it is appropriate to scale out this latter dependence by in­
troducing

(61)

where = (2n+l)/n, such that is nondimensional and solely de­


pends on the temperature field, being e 1 when T e T q.
For a center-gated disk cavity or a circular runner, analogous re ­
sults can be obtained. In particular. Table 3 summarizes such results
for these three simple geometries, both for a power-law fluid and for
an arbitrary ri(Y> T, p ) dependence.
Numerically, the foregoing problem for such simple geometries can
be handled quite adequately in terms of a finite-difference approach.
Introducing a variable-mesh grid (xi, yj) such that, e .g . , Ti^ j, k =
T (x i, yj, tj^), we get a standard finite-difference representation of
(45):

/T . - 1 T. . T.
, / i,j,k + l
pC . I
i,],k Ax. .
P\ 1-1

r 2T. 2T
1, i.i,k+l
"" N h ' Ayj-iAy^.

2T
i,3+l,k+l
+ <i>. (62)
Ayj(Ay._^+Ay^ i,3,k
?1
where $ e ny^ and Xj e + Axi_i, yj e y j - i + Ayj_x and tj^+i e tj^
+ Atj^. In particular, an upwind difference has been used here for
stability reasons in the stream wise convection term, whereas an im­
plicit representation in time has been used for the temperature in the
conduction term, thereby allowing relatively large values for At;j^.
fD
Table 3 Governing Relations in Various One-Dimensional Flows

Power-law fluid w
o
n =m (T )t>^-l, m (T) = mog(T) Arbitrary n ( Y , T, p) o
(A

Tube flow
n
3-
0}
-5
Q}
Q = ^/yYdy Y = , A = 2 m ,(^ ) , $ ^ f dy O
^ 2n ’ ttS ’ n ^ O
-s

o
Disk flow 3
(sector angle 6)

.b .. a/n / n b 2
Ay
Q = 29x/yYdy Y = (-^ ) > A = m o (^ ) , 5 H/ y (| ) dy Y= ^ A = -^
n 20xS

Strip flow
(width w)

r’" •
Q = 2w J yY dy
0
= (^ r. ^ t
x: Streamwise coordinate, A = -9p/8x.
y : Transverse coordinate (between centerline and w all).
30 Hieber

Further, the flow quantities u and ^ are based upon their known val­
ues from the start of the time step.
In particular, suppose there are N2 increments across the half-gap
such that Y j - i = 0 and y n 2+1 = ^ (with Ayj typically chosen to be
smaller near the wall, to resolve the thin thermal boundary layer) and
the time step has been chosen to be uniform with the strip mold of
length L filling in N1 time steps (i.e . , At = L/(N 1*U ), where U = Q/
2bw). Accordingly, at time t^+l = k At, the melt front will lie at
Xk+i = k Ax, where Ax = L/Nl. Supposing then that all quantities
have been determined through time t^, we would first update the
temperature field at time t^+l. In particular, based upon (47b), we
start at i=l with

for 1 < i < N2+1 ( 63 )

and then successively march forward to i = 2 ,3 ,... ,k, making use of


(62) to determine at each new i. In this regard, (62) ap­
plies only at the interior nodes, 2 < j < N2; at the centerline, j = 1,
the finite-difference representation of the conduction term gets ap­
propriately modified to

^th(Ayi^)('^i,l,k+l ’^ i , 2 , k + l )

whereas at the wall, j = N2 + 1, the equation gets replaced by the


boundary condition (48b), namely

( 64 )
^i,N 2+l,k+l

Accordingly, at each new i, the problem is tridiagonal in terms of


Ti,j,k+1 (1 1 j 1 N2+1) and can be solved directly by standard tri­
diagonal elimination. Having thus determined (1 < ] < N2+1)
for i = 2 , 3 , . . . , k , we then treat the temperature along the melt front,
^k+1 = ^ as discussed earlier, such that

for 1 < i < N2+1 ( 65 )


\+ l,i,k + l '^k,l,k+l

With the temperature field updated at time tk+i, it is then neces­


sary to update the various flow quantities. If we are dealing with a
general ti(Y> T, p) model, this latter update will require an iterative
procedure. For example, we can first evaluate S in (54) by numerical
(trapezoidal) quadrature, basing n upon the old shear rate and pres­
sure fields but upon the updated temperature field. (For simplicity,
the Y distribution along the new melt front, might be taken
Melt-Viscosity Characterization 31

initially equal to the old Y at xj^.) With such tentative values for
Si,k+1 (1 1 i 1 k + l), we can then calculate corresponding values
for Ai,k+1 f^om (55). Accordingly, the shear-rate distribution can
then be calculated from (51), and the pressure field determined by
numerical quadrature of Ai,k+1’ starting with a zero (gage-pressure)
value at the melt front and integrating back towards the inlet. This
cycle can then be repeated, with S being recalculated based upon a
reevaluated ri, etc., and the iterative process can be continued until
a prescribed convergence criterion is satisfied between successive
cycles. Following convergence, u is then calculated by numerical
quadrature of Y, starting with u = 0 at the wall and integrating in to
the centerline.
On the other hand, if the viscosity is based upon a power-law
model, such as (46), then there is no need for an iterative proce­
dure. That is, once the temperature field has been updated, one can
then evaluate $ from (59), based upon numerical quadrature, and then
directly update A and Y, based upon (60) and (56). Subsequently,
p (u ) can be generated from A (Y) by numerical quadrature.
During the first time step, when the melt front is at X2 = Ax, the
distributions for T, u, Y, e tc ., at i = 2 are merely taken to be the
same as at the inlet. Thereafter, we march forward in time as de­
scribed above, continuing the calculation until the cavity fills.
Figures 15 and 16 show the resulting calculated temperature and
velocity distributions at time of fill for a strip of dimensions (L , 2b,
w) = (20 cm, 0.2 cm, 2.5 cm) with Q = 10 cmVsec (tfiii = 1 sec), T q =
200°C, and T ^ = 27°C. The calculation has been based on (46) with
1-n
(n. A , T ) = 0.312, 34.3 g 3910 K (66)
a -( cm •sec \sec /

which corresponds to the power-law limit (3, 4) of the four-constant


fit (1, 2, 11) for the PS material [Styron 678U (D ow )] shown in Fig. 1
of Section II. In addition, the thermal properties have been taken as

gm
(p , C , k ) 0.94 . 2 1 X 10’^
p th cm^ ’ gm -°C ’
e rg \
1.5 X 10*^ (67)
sec-cm -°C j

which are representative of PS melts. The results in Figs. 15 and 16


are based upon N1 = 20 ( i . e . , Ax = L/20) and N2 = 9 with

= 0, 0.15, 0.30, 0.45, 0.60, 0.70, 0.80, 0.90, 0.95, 1.00 (68)
32 Hieber

X (cm)

F ig . 15 Predicted isotherms (K ) at t = tfni = 1 sec; power-law ap­


proximation for PS (Styron 678U/Dow); Q = 10 cm^/sec, T q = 473 K,
Tw = 300 K.

In particular, the calculated isotherms at tfin in Fig. 15 indicate


that the hot core is slightly above the injection temperature due to
viscous heating, whereas the cold layer next to the wall is thickest
at some intermediate streamwise location, which is typically closer to
the entrance than to the end of the cavity. Such a structure for the
cold boundary layer is compatible with an order-of-magnitude analy­
sis as presented in Appendix K. The corresponding contours of con­
stant u shown in Fig. 16 indicate that the flow stagnates in the cold
wall layer, thus requiring increased velocities in the hot core to main­
tain the specified volumetric flow rate (corresponding to a bulk ve­
locity of 20 cm/sec, in this case).
How the temperature field can affect the required injection pres­
sure is shown in Fig. 17 by plots at tfju of the calculated streamwise
variation of ^*, as defined in (61), for various values of Q. In p ar­
ticular, the curve for Q = 10 cm^/sec corresponds to the results in
Figs. 15 and 16, from which we see that the smallest value of 5* oc­
curs at X 6 cm. That is, the melt is coolest, in terms of its flow
resistance, at this streamwise location. On the other hand, at the
high Q of 100 cm^/sec (tfiu = 0.1 sec), the flow is basically less
resistive than at the inlet due to a dominant viscous-heating effect.
Melt-VIscosity Characterization 33

Fig. 16 Predicted contours of constant u (cm/sec) at t = tfni = 1 sec;


same conditions as in Fig. 15.

(The initial dip in this curve indicates that conduction cooling domi­
nates near the entry, but viscous heating eventually dominates fu r­
ther downstream; such a behavior is also found in simulating steady
flow in cold-walled runners.) At the other extreme, for a Q of 1
cm^/sec (tfin = 10 sec), the cooling effect has extended even into the
core, as evidenced by the lowered value of 5* at x = L [correspond­
ing to the melt front at tfjn, hence reflecting the centerline tempera­
ture at L - Ax, based upon (6 5 )]. In assessing the effect of 5* upon
the pressure field, we note from (60, 61) that Ap will be proportional
to ( 5 * ) ”^, where ( ) denotes a stream wise-averaged value. Hence,
e .g . , from the curve for Q = 3 cm^/sec in Fig. 17, we see that «
10"^, indicating that Ap should be higher than Ap^goth ^ factor of
about (100)^ ~ 4 in this case. This is borne out by Fig. 18, which
shows Ap versus Q at tfin based upon various models. In particular,
curve 1 corresponds to the foregoing power-law calculation with 1’
being the associated isothermal result for which Ap is proportional to
Q^. The predicted Ap according to curve 1 is indeed about four
times higher than the corresponding isothermal result at Q = 3 cm^/
sec. Also indicated in Fig. 18 are results based upon the full four-
constant (n , T*, B , T^)) model (1, 2, 11), namely curves 2 and 2’ .
In particular, curve 2^ lies below 1’ at lower Q, which is what would
34 Hleber

X (cm)

Fig. 17 Predicted distribution at t = same conditions as in


Fig. 15 but for Q = 1, 3, 10, 30, 100 cm^/sec.

be expected since the power-law limit overpredicts the viscosity when


extended to low shear rates. On the other hand, curves 1 and 2 are
essentially coincident. This might be interpreted on the basis that,
at the same shear rate, the four-constant model is more temperature
sensitive than the limiting power-law model [as is seen from (31—33),
a’ of the former model will lie above nao and only approach the latter
Melt-Viscosity Characterization 35

(/)
Q.

Q.
<

Fig. 18 Predicted pressure drop at t = tfin based upon various models


for same conditions as in Fig. 15 but for different values of Q. Curve
1: power-law limit (46, 6 6 ) ; 2 : four-constant model ( 1 , 2 , 1 1 ) ; 3:
five-constant model (1, 34, 140); 4; seven-constant model (1, 36, 37,
40, 141). Primed curves: corresponding isothermal case [obtained
by setting Tw = and omitting f^o>^ (62)] .

at larger Q] such that the cooling effect at lower Q will be larger in


the case of the four-constant model (as reflected by a larger shift be­
tween curves 2 and 2^ compared with that between curves 1 and V ) .
Alternatively, and more simply, the essential coincidence of curves 1
and 2 in Fig. 18 might be explained on the basis of the associated wall
shear-stress levels, in particular, a pressure drop of 6.9 x io® dyne/
cm^ ( 1 0 ^* psi) over a length of 2 0 cin with a half-gap thickness of 0 . 1
cm translatés to an average wall shear stress of « 3.5 x lo® dyne/cm^.
Hehce, compared with a t * of 1.77 x 1 0 ^ dyne/cm^ in the present case
(11), it foliotATS that the power-law model should indeed be a very good
approximation in the present nonisothermal case.
Also indicated in Fig. 18 are results based upon the five-constant
(n, T * , B, T]^, 3) model, curve 3, and the seven-constant (n , x*,
D i, D 2 , D 3 , A i, Â 2 ) rtiodel, curve 4. In these cases the model
36 Hieber

constants have been obtained by holding 3 or D 2 , D 3 , A i , and A 2


fixed, based upon reasonable values for PS from Section II.C and
Appendix C, and by applying the numerical program described in
Appendix J to the capillary-rheometer data for Styron 678U in Fig. 1.
The resulting values are given in Eqs. (140) and (141). As expected,
the incorporation of pressure dependence in terms of a nonzero 3 re ­
sults in higher pressure-drop predictions, which we see by comparing
curve 3 with curve 2. On the other hand, curve 4 is seen to cross
curve 3 at an intermediate Q , lying below for high Q and above for
low Q. This relative behavior at low Q is expected due to the in­
creased temperature and pressure sensitivities of the WLF-based
model at low temperatures, as exhibited by the corresponding curves
for Ti3 in Fig. 46 of Appendix C and for 3o Fig* H of Section II.C.
On the other hand, the fact that curve 5 lies below curve 4 at high Q
can also be attributed to a higher temperature sensitivity and its re ­
sponse to an elevated temperature due to viscous heating. In partic­
ular, the D 2 , A i, and A 2 upon which the present seven-constant fit,
(141), is based corresponds to curve 5 in Fig. 46. Accordingly, at
atmospheric pressure this model corresponds to Ti^ « 14,300 K at the
injection temperature of 2 0 0 ° C , compared with a constant T^ of 12,500
K in the present five-constant model, (140). In addition, ao of the
former model will actually be larger at the elevated pressures due to
the D3 P term in the expression for ag in (41).
Although the lower values of Q in Fig. 18 are not typical for the
filling stage. Section IV shows that a similar flow regime does arise
during the post-filling stage, where significant cooling occurs and
the difference between the five-constant and seven-constant models
becomes crucial. Before proceeding to such considerations, however,
we direct our attention to extending the cavity-filling simulation to
more general geometries.

B. Composites o f Simple Geometries

As a step toward handling more complicated geometries, one natural


approach is to represent the cavity and delivery system as a compos­
ite of one-dimensional-flow segments. Such an approach has been
used in [118,120,121] and, on a commercial basis, by MOLDFLOW [9 ].
Here an analogous technique is described, which is based on work at
Cornell University, namely the ’’coupled-flow-pathmethod [122].
As indicated schematically in Fig. 19, the delivery/cavity system
is represented as a combination of flow paths, each consisting of a
series of one-dimensional-flow segments. For Fig. 19 the system con­
sists of five flow paths comprising nine segments. Each segment can
be a circular tube, a center-gated-disk sector, or an end-gated strip,
for each of which the governing dynamic equations are listed in Table 3.
Melt-Viscosity Characterization 37

Fig. 19 Schematic of a system represented in terms of five flow paths


consisting of nine segments.

For noncircular runners the procedure is to use an equivalent cir­


cular runner based on the hydraulic radius (twice the cross-sectional
area divided by the perimeter) but with the effective volumetric flow
rate scaled down to maintain the actual bulk velocity. This approach
was found [123] to work well for polymer melts under isothermal con­
ditions, with similar results being found independently in [22] and
[124]. For cold-wall, noncircular runners such a procedure still
works fairly well [125].
The basic idea then in the coupled-flow-path approach is that the
volumetric flow rate along each flow path or arm, QARM (I), should
be adjusted each new time step such that the corresponding pressure
drop along each arm, DPARM (I), be the same for all I. During each
time step the procedure is to first advance the melt front in each par­
tially filled segment based on the known volumetric flow rates at the
start of that time step. Next, we update the temperature field in each
segment in which the melt is flowing, using a finite-difference repre­
sentation of the energy equation as described in Section A, with the
flow quantities again being based on their known values from the start
of that time step. At this point the flow resistance along each flow
path will have changed due to the advancement of the melt front and
a change in the temperature field. It is then necessary to adjust the
QARM(I) such that the DP ARM (I ) will all be equal again. We do this
via a Newton-Raphson procedure. In particular, for each arm I

NNARM
8PARM(1)
DPARM(I) + X X 6QARMCJ) = DPT (69)
8QARM(J)
J=1

where DPARM(I) denotes the calculated pressure drop along arm I


based on the newly advanced melt-front position and updated tem­
perature field but the old volumetric flow rates, whereas 6QARM(J)
denotes the required change in QARM(J) Such that the resulting
pressure drop along each arm will equal DPT (the injection pressure
38 Hieber

required to maintain the total prescribed Q at new time). Lastly,


8PARM(I)/3QARM(J) represents the sensitivity of DPARM(I) to
changes in QARM (J). This quantity is evaluated by successively
changing each QARM(J) by 1% and calculating the resulting change
in DPARM (I). In general, the off-diagonal terms (I J) will be non­
zero if arms I and J share a common segment ( e . g . , all five arms in
Fig. 19 share the sprue segment 1, whereas four of the five arms
share segment 3). NNARM denotes the total number of flow paths or
arms such that (69) constitutes NNARM equations but involves NNARM
+1 unknowns, namely 6QARM(J), 1 < J < NNARM, and DPT. The fi­
nal required equation is obtained from the global flow condition,
namely

NNARM
J 5 Q A R M (J )= 0 (70)
J=1

which follows from the fact that the unperturbed QARM(J) from the
start of the time step will have already added up to Q .
In evaluating DPARM(I) and 8PARM(I)/9QARM(J) in (69), we note
that

DPARM(I) 2 DPSEG(K) (71)


K

where the summation is over all segments lying along arm I. Accord­
ingly, in general, as we successively perturb each QARM(J) by 1% to
evaluate the sensitivity factor in (69), we must evaluate the pressure
drop along each segment, making use of the relations in Section A,
where the corresponding volumetric flow rate in segment I is

QSEG(I) = X QARM(K) (72)


K

Where the summation is over all arms passing through segment I.


In the special case of a power-law fluid, the preceding calculation
is made much easier because Ap That is, in this case it is ap­
propriate to introduce the quantity C S E G (I), where

DPSEG(I) E CSEG(I) *Q S E G (I)**E N (73)

and EN denotes the power-law index from (46). In particular, CSEG(I)


can be determined by calculating DPSEG(I) based on the old QSEG(I)
Melt-Viscosity Characterization 39

but the updated temperature and melt-front position, and then divid­
ing by the QSEG(I)**EN factor based on the old QSEG(I). Note that
CSEG(I) merely reflects the geometry of segment I together with the
effect of the current temperature distribution upon the viscosity field.
For example, for an end-gated strip segment as considered in Section
A, it would follow from (60) and (73) that CSEG(I) would correspond
to the quantity

mn
dx (74)
(2w)
-» /■ (if
0

with the upper limit x* being equal to either the current melt-front po­
sition, if the segment is partially filled, or to the length of segment I,
if it is filled. Similar relations would apply for the circular runner
and center-gated-disk segments, based on the analogous power-law re­
lations indicated in Table 3. In any case, CSEG(I) is unaffected by the
flow rate such that it remains constant during the Newton-Raphson pro­
cedure, thereby greatly reducing the execution time required to eval­
uate the sensitivity factor in (69). Once DPARM(I) and 9PARM(I)/
9QARM(J) have been evaluated, the resulting system of NNARM + 1
equations, corresponding to (69) and (70), is used to determine the
required changes in the various volumetric flow rates, 6QARM(I),
1 < I < NNARM, together with the corresponding injection pressure,
DPT, required to sustain the specified total volumetric flow rate. The
system of NNARM + 1 equations can be solved by standard Gaussian
elimination.
For reasons of numerical stability it is appropriate to ’’underrelax
the foregoing procedure and not allow any of the QARM(I) to change
by more than (typically) 10% during any such iteration. The need
for such a constraint typically arises when the melt front enters a
segment having a drastically different flow area or when an arm fills
such that the flow rate in that arm must get distributed among the
remaining arms that are still flowing. In such cases, the Newton-
Raphson procedure will be repeated until all the 6QARM(I) lie below
the 10% criterion, at which point the full values for the 6QARM(I) will
be used and the iteration ended.
A further complication concerns the situation where two (or more)
arms have not yet split. For example, referring to the schematic in
Fig. 19, we see that if segment 5 is currently partially filled, it means
that arms 3, 4, and 5, passing through segment 5, are indistinguish­
able at this point. Hence, to avoid redundant equations in (69), we
must replace Eq. (69) for arms 4 and 5 by the following relations:

6QARM(3) - 5QARM(4) = 0, 6QARM(3) - 6QARM(5) = 0 (75)


40 Hieber

That is, the changes in QARM(4) and QARM(5) must be the same as
the change in QARM(3). In addition, the total number of equations
depends on whether any of the flow paths has filled. For example,
if segment 9 in Fig. 19 has already filled and segments 4 and 5 are
currently partially filled, then the system will consist of five equa­
tions—namely Eq. (69) for arms 2 and 3, Eq. (75) for arms 4 and 5,
andE q. (70).
One further special consideration relates to when a segment fills
which does not lie at the end of a flow path. For example, when seg­
ment 3 in Fig. 19 fills, it becomes necessary to distribute QSEG(3)
among segments 4 and 5 since values for QSEG(4) and QSEG(5) will
be needed to advance the melt front in these respective segments dur­
ing the next time step. In general, the distribution of the upstream
QSEG among the downstream segments should be done on the basis
that the resulting pressure drop along each downstream segment be
the same for small subsequent times. For example, if the downstream
segments are all circular tubes, then for each downstream segment.

AP. A. X U.t (76)


3 3

i.e ., the product of the pressure gradient and the distance advanced
in time t measured from when the upstream segment fills. In particu­
lar, it is reasonable to assume that the temperature distribution in
each such downstream runner segment will initially be the same, such
that thermal effects will scale out. Hence, since

w.
A. — i- (77)
] a.

and, if we assume a power-law fluid.

n
T ^ m (T) Y. (78)
w. 3

whereas

(79)

and


Y. 3
—3 (80)
Melt-Viscosity Characterization 41

it then follows that

(81)
3 \ 3 / 3

which will be the same for all j provided Qj aj^. That is, if the
downstream segments are all circular tubes, then each Qj should be
proportional to the cube of its radius, with the constant of propor­
tionality chosen such that the Qj sum to the QSEG coming from the
upstream segment. In addition, Yj will then be the same in each down­
stream segment, indicating that the foregoing distribution for the Qj
will also apply even if the shear-rate dependence of the viscosity is
other than the simple power-law relation. By similar means it can be
shown that the downstream Qj should be proportional to Wjbj^ if all
are end-gated strips, to 6jbj^ if all are center-gated-disk sectors of
zero inner radius, and to Ojbj^j if all are center-gated-disk sectors
of nonzero inner radius (r j) .
Our formulation neglects juncture pressure losses between succes­
sive segments, although some progress has recently been made in this
direction [126] , particularly in the gate region, by extending the
semianalytical results developed in [127]. Also see Appendix F and
Chapter 2. On the other hand, concerning the handling of the tem­
perature field between successive segments, we distinguish between
whether the upstream and downstream segments are both of the same
^’type^’ ( i . e . , both axisymmetric or both plan ar). If they are, then
the calculated temperature profile at the exit of the upstream segment
is taken as the inlet temperature profile for the downstream segment ;
if not, then the calculated bulk temperature ( ”mixing-cup” tempera­
ture) at the exit of the upstream segment is taken as the uniform tem­
perature at the entrance to the downstream segment. Although some­
what arbitrary, some such procedure is needed since there is evi­
dently no simple way to relate streamlines between axisymmetric and
planar configurations.
As an illustration. Fig. 20 shows an 18-segment representation of
a double-gated rectangular plate with intentionally unbalanced runners,
as presented in [122]. The system has been represented in terms of
four flow paths, with the weldline location being based on a more de­
tailed calculation [128] using a finite-element—finite-difference proce­
dure described in Section C. Figure 21 shows resulting predictions
for the pressure versus time at the end of segment 1 (the sprue) for
a polypropylene with T q = 260°C, T^^^ = 27°C, a constant Q of 110 cm^/
sec and a cavity thickness (2b) of 0.318 cm. The viscosity has been
modeled in terms of a power-law fluid, (46), with
1-n
n = 0.323, A = 5400 (g/cm*sec)(l/sec) T« = 1470 K (82)
42 Hieber

Fig. 20 Representation of a sprue, runner, and double-gated rectan­


gular cavity in terms of four flow paths comprising a total of 18 seg­
ments.

Fig. 21 Predictions for pressure vs. time at end of segment 1 in Fig.


20 based upon coupled-flow-path method (open symbols) and FEM/
FDM simulation (solid symbols) described in Section C; PP (Profax
6523/Hercules) with Q = llOcm^/sec, T q = 260°C, and Tw = 27^C;
2b = 0.318 cm.
Melt-Viscosity Characterization 43

and the thermal properties taken as

p = 0 .7 7 -^ , C =3.1x10^.^?^,
p gm K

erg
1.5 X 10^ (83)
th sec*cm K

which are representative of molten P P . In particular, the predictions


in Fig. 21 based on the coupled-flow-path calculation (open symbols)
agree well with the results from the more accurate finite-element—fi­
nite-difference formulation (solid symbols). In turn, the latter pre­
dictions have been found [128] to agree well with corresponding ex­
perimental pressures.
From Fig. 21 note that the runners fill at t = 0.63 sec, at which
time there is an abrupt rise in pressure as the melt enters the rela­
tively more resistive thin cavity. Subsequently, there is another
abrupt rise at t « 2.3 sec, corresponding to when the left end of the
cavity fills, and then a final dramatic rise in pressure at t « 2.6 sec,
when the weldline forms across the entire cavity, thus tending to
block the flow from the left gate. This is illustrated even more clearly
in Fig. 22, which shows predictions for the volumetric flow rate out of

F ig . 22 Predictions for Ql /Q versus time; symbols and conditions


are the same as in Fig. 21.
44 Hieber

the left gate, ’■QL’^ divided by the constant total Q of 110 cmVsec.
In particular, the rise in Ql at t - 0.63 sec indicates that segment
3 apparently fills slightly earlier than 2 such that drops as the
melt first enters the more resistive cavity through the right gate.
However, once segment 2 fills and the melt enters through the left
gate, the flow resistance in the cavity due to the initially large value
of Ql will rapidly build up in the domain emanating from the left
gate, resulting in a decreasing Ql - Further, Fig. 22 clearly indi­
cates abrupt drops in Ql at t 2.3 sec, corresponding to when the
left end of the cavity fills, and at t =^2.6 sec, when the weldline
forms across the entire cavity. The coupled-flow-path results in­
dicate that Ql = 0 after the weldline forms, whereas the more accu­
rate calculation, based on the finite-element method—finite-difference
method (FEM—FDM) formulation described in the next section, allows
the weldline to subsequently move, resulting in nonzero but small
values of Ql /Q-

C. T h in Cavities of A r b i t r a r y Planar Geometry

As a further step in complexity, consider the filling of thin cavities


of arbitrary planar geometry; then a major added complication con­
cerns predicting the advancing melt front as a function of time in
planar coordinates. Notable contributions and/or applications in this
area have been made in [63,128—131]. In particular, the treatment
given here follows that in [128] , which uses a hybrid FEM—FDM for­
mulation together with a two-step, predictor-corrector procedure for
tracking the advancing melt front. An extension of this work, which
combines the FEM—FDM formulation of [128] with a variant of the pro­
cedure developed in [129] for advancing the melt front, is described
and illustrated in Chapter 9. This latter program can handle thin
three-dimensional parts, treated as composites of thin two-dimensional
components, and can be viewed as the present state-of-the-art in this
area, being comparable in versatility to the proprietary FEM package
ofMOLDFLOW [9 ].
In terms of a thin cavity of arbitrary planar geometry, the appro­
priate extension of the governing equations for one-dimensional-flow
segments considered in Section A can be written as

9u\
0 (84)
9x 9z \ dzi

3v ^
(85)
9y 9z\" 9z 1

n /3T ^ 8T
4-
~ xVr
+ nY" ( 86)
^P\9t 9x 9y / N h 9z^
Melt-Viscosity Characterization 45

¿ (b u ) + ^ (b v ) = 0 (87)

where (u , v ) are the velocity components in the (x , y ) planar coordi­


nates, whereas z is the gapwise coordinate and the ( “) indicates a
gap wise-averaged value. That is, the force balance in the x and y di­
rections is again between the pressure gradient and viscous-stress
terms, with inertial effects being omitted from the left sides of (84) and
(85). As in Section A, the transverse convection term, pCp(w9T/9z)
in the present context, is again omitted from the energy equation (see
Appendix K for some further discussion concerning the structure of
the temperature field and related references). An underlying assump­
tion throughout is that the gap thickness, 2b (which can be a func­
tion of X and y, as illustrated specifically in [131]), is much smaller
than any characteristic length L in the (x , y ) plane. Accordingly,
3/Bx, 3/3y « 3/3z, such that corresponding viscous-diffusion terms
involving x and y derivatives have been omitted from (84) and (85) as
well as such thermal-diffusion terms from (86). Further, the shear
rate is then given by

Y= ( 88)

Finally, the viscosity must be specified as a function of y, T, and


possibly p.
Concerning boundary conditions in the gap wise direction, we fol­
low the same approach as in Section A, thus requiring at the wall that

u (x , y, b, t) = 0 = v (x , y, b, t) (89a)

and

T (x , y, b, t) = (89b)

whereas along the centerline

g ( x , y. 0, t) = 0 = |J (X, y, 0, t) (90a)

and

(x , y, 0, t) = 0 (90b)
3z
46 Hieber

Accordingly, by introducing

(91)

and integrating (84) and (85) in the gapwise direction, making use of
( 90a), we obtain

3u A 9v
n TT- = ■ A ^ z , n ^ = - A z (92)
8z X 3z y

which, combined with (88), gives

• - Az (93)
^ ~ n

where

A 5 7a ^+ a ^ = Ivp (94)
'' X y ' ^

Equation (93) is a direct extension of (51) in Section A. By integra­


ting (92) again, making use of (89a), we get

b ~ b ~
u = A f -dz, v = A f ~dz (95)
y n

which indicates that (u , v ) °c -(9p/9x, 9p/9y). That is, the velocity


is in the direction of -Vp, as in classical Hele-Shaw flow [132], namely
viscous-dominated flow between closely spaced plates, which is here
being extended to a non-Newtonian viscosity under nonisothermal con­
ditions. Finally, performing another integration in the gapwise direc­
tion, using integration by parts, we obtain from (95) that

- _ S - S ,
u Ax> (96)
b

where

dz (97)
0

Substituting (96) into (87) then gives the following governing equation
for the pressure field:
Melt-Viscosity Characterization

(98)
8x

which is a nonlinear (since S depends on the pressure field, in gen­


eral) , elliptic partial differential equation.
For the special case of a power-law fluid, (46), the foregoing sim­
plifies just as in Section A. That is, as in (56), we now have from
(93) and (46) that

(99)
m " "

such that

(n -l)/ n
Az ( 100)
-(f)'
and

( 101)

where

(I) - ( 102)

In particular, the dependence of the pressure field on the temperature


field and cavity thickness is completely embedded in $, For the clas­
sical Hele-Shaw case [132], the fluid is Newtonian (n = l) and the cav­
ity thickness and temperature are uniform, such that g = 1 and both
^ and S are constant, with (98) thus reducing to Laplace^s equation.
For injection-molding applications, however, there is substantial shear
thinning with n « 1/3 or 1/4, such that, according to (101), S ^
or A^, respectively. That is, the governing equation for pressure,
(98), is highly nonlinear. In addition, as indicated by the sample
problem in Section A, the factor ^ will tend to be highly non uniform
due to nonisothermal effects, even for uniform cavity thickness.
Concerning boundary conditions in the planar coordinates, we re ­
fer to the schematic in Fig. 23. In general, since the governing force-
balance equations, (84) and (85), neglect viscous diffusion in the x
and y directions, the fluid seems to be inviscid when viewed in the
x -y plane. Indeed, this basic property was exploited by Hele-Shaw
48 Hieber

[133] in his original experimental work in which the observed stream­


lines around arbitrary bodies (aerofoils, in later work) correspond to
inviscid, potential flow. In this context it follows that along the outer
boundaries of the cavity, C q , or along any insert, Cj, one can only
require that the normal velocity be zero (impermeability condition).
Hence, when expressed in terms of pressure, this results in the nor­
mal derivative being zero, i.e. ,

IP 0 on C and C . (103)
3n o 1

It can be shown ( e . g . , [134]) that to impose a zero tangential veloc­


ity (no-slip condition), one would have to retain viscous diffusion in
the x -y plane within a region of 0 ( b ) , the cavity thickness, from such
boundaries. However, since the underlying assumption is that the
cavity thickness is much smaller than any characteristic length in the
x -y plane, it follows that such regions will be thin and can be neg­
lected in leading-order approximation.
For the same reason, when any weldline forms due to the impinge­
ment of two melt fronts, a situation which arises when the cavity is
fed by more than one gate or when there exist some inserts in the
cavity, one can only require that the pressure and normal velocity
be the same on either side of the weldline. That is.
Melt-Vrscosity Characterization 49

+
P = and on C (104)
wl

where we have used (96), noting will be continuous across such


boundaries, and the minus sign in front of (S 9p/9n)‘ is due to the
present convention that n"^ and n" are in opposite directions, both
being directed outward from their respective domains on either side
of the weldline. Again, due to the neglect of viscous diffusion in the
x -y plane , the present formulation must allow a discontinuity in the
tangential velocity (slip) across the weldline (which could also be
remedied by introducing viscous diffusion in the x -y plane within a
thin layer of 0 ( b ) in width on either side of the weldline).
Along the melt front, assuming the cavity to be properly vented,
we require that

0 on C (105)
mf

where p is taken as gauge pressure. Finally, along the gate we typi­


cally assume a uniform but time-dependent P q ( 1) whose value is to be
determined such that a specified total volumetric flow rate Q is main­
tained. In turn, this requires evaluating a line contour such that

Q / (2 b )v ds (106)
'' n
C
where C denotes any line contour which encloses the gate. Typically,
satisfying (106) by properly choosing P q ( 1) requires an iterative pro­
cedure, which will be described later.
Since thermal diffusion in the x -y plane is not included in (86), we
cannot impose any thermal conditions along Cq or Cj in Fig. 23. In
assessing the possible inaccuracy of this approximation, we note that
a cold mold temperature along Cq or Ci would, at most, penetrate a
distance of order /it into the melt during filling (convection would
actually reduce this purely conductive distance) , where a, the ther­
mal diffusivity (kth/pCp), is typically about 10'^ cm^/sec for polymer
melts. Hence, with fill times of about 1 sec or less, it follows that the
actual width of cold layers along Cq or would be about 0.03 cm,
which will be negligible compared with the characteristic length in
the x -y plane.
Similarly, along any weldline we must allow for a possible discon­
tinuity in temperature (due to different thermal histories on either
side of C ^ i ) . Again , this discontinuity could be smoothed out by in­
cluding thermal-diffusion layers of 0 (/ a t) in width on either side of
Cwi‘
On the other hand, along the melt front we adopt the same proce­
dure as in Section A . We can write this as
50 Hieber

on C (107)
mf

where denotes the centerline temperature of the nearest nodal


point lying upstream of the given mesh point on Cmf. Such a proce­
dure is intended to mimic the fountain-effect phenomenon (discussed
further in Appendix K ) , typically resulting in a thermal boundary
layer on the upper and lower plates which is thickest somewhere in­
termediate between the gate and the melt front, as illustrated in Figs.
15 and 17 of Section A for a strip cavity.
Finally, concerning the inlet thermal boundary condition, we typi­
cally require that

f = Tn on (108)

where Cg corresponds to a point in Fig. 23 but, more generally,


could be a line contour, whereas T q might be taken as the air-shot
temperature coming out of the nozzle of the injection-molding machine
or, preferably, it might be taken as the resulting calculated bulk
temperature coming out of the runner/gate system based on a simu­
lation in the manner of Section A.
In solving the preceding problem for given geometry and flow
conditions (Q , To» T ^ ) as well as specified thermal properties (p ,
Cp, k^}^) and viscosity n(Y, T, p ) , a FEM—FDM formulation has been
developed [128] in which triangular finite elements are used to de­
scribe the geometry in the x -y plane and a finite-difference grid is
used to represent the gapwise variation of such quantities as T , u,
V, Y, and p. In particular, the pressure field on element 1 is rep­
resented as

6
P
(1),(x , y, t) (109)
S
k=l

in which Pk^^^(t) is the pressure at the kth node of element 1 (see


Fig. 24), which is to be determined as a function of time, and Qk(^)
(x , y) denotes the standard quadratic shape function [135,136] for
a six-noded triangle. On the other hand, the temperature field on
element 1 is represented as

y, z, t) = 2 L^^Y x , y ) t) ( 110 )
k=l

where y ) denotes the standard linear shape function [135,


136] on a triangle and T j^(l)(z, t) denotes the temperature profile
Melt-Vìscoslty Characterization 51

k=3

k=2

Fig. 24 Nodal convention within an element.

at the kth vertex node of element 1, which is described in terms of a


finite-difference grid across the half-gap.
That is, the energy equation, (86), is solved at each vertex node
by using finite-difference representations for 8T/9t and 3^T/9z^, as
described in Section A, but with the linear shape functions in the
(x , y ) plane being used to evaluate 9T/9x and 9T/9y in the thermal
convection term (for stability reasons, contributions are only included
from those elements lying upwind of the given vertex node, as de­
scribed in [128]). With the flow quantities u, v, and nY^ in (86)
being based on their known values at the start of each time step, in
the same manner as in Section A, it follows that the temperature field
can be updated directly (see [128] for further details).
With the temperature field updated, the corresponding pressure
field is then determined by a Galerkin procedure, which is solved by
successive underrelaxation. In particular, for each node N we intro­
duce the weighting function p n ( x , y ) , which is constructed from the
quadratic shape functions in (109), such that pj^ = 1 at node N but
vanishes at each of the remaining five nodes of each element contain­
ing node N, as well as throughout any elements not containing N. Ac­
cordingly, multiplying (98) by pj^, integrating over the entire melt do­
main , and using the divergence theorem, we get

0 / p V -(S V p ) dü
a ^

= I - I N
(SVp) dQ ( 111)

But if node N is an interior node, then PN = 0 along the outer contour


C such that the line integral in (111) vanishes; on the other hand.
52 Hieber

if N lies along impermeable boundaries C© or Ci in Fig. 23, then


9p/9n = 0 and the line integral again vanishes. Hence, for any in­
terior node or node lying along an impermeable boundary, it follows
that

/ s(Vp^-vp) = 0 ( 112 )

On the other hand, if N lies along a weldline, then from the sec^
ond condition in (104) it follows that the line integral in (111) will
have equal but opposite values for nodes on either side of C^i. Hence,
if we treat these nodes as pairs ( i . e . , having the same (x , y ) coordi­
nates but possibly different temperatures and tangential velocities, as
discussed earlier), noting from the first condition in (104) that the
pressures at these nodes will be the same, it follows that we can add
Eq. ( I l l ) for these two nodes such that the line integral terms will
cancel and the resulting governing equation will again be given by
( 112 ) .
Finally, if N lies along the melt fron t, then we require pj^ = 0, as
in (105), whereas if the node lies along the gate, then we require that
Pn “ where P Q (t) is to be determined such that (106) is satis­
fied. In particular, (106) can be written in terms of the pressure
field as

Q = -2 / s |É d .
8n
(113)

where contour C encloses the gate.


In the integral of (112), Vp]^ and Vp are linear functions in (x , y ) ;
hence, if we treat S as a constant on each element (evaluated at its
centroid, for example), then the integral in (113) can be evaluated
in closed form. We get

6
2 (114)
1]
1' i-1

where N = NELNOD(l', i) and N' = NELNOD(l', j) (i .e ., N and N' are


the ith and jth nodes of element 1' in the convention of F ig. 24), with
the summation in 1' being over all elements containing node N . Since
B ij(l’) in (114) depends on pj^ through the S factor, it follows that
(114) is a nonlinear problem, which therefore requires an iterative-
type solution. In particular, we have solved the system of equations
by successive underrelaxation, typically taking 60% new and 40% old
as we scan the various nodes. The underrelaxation is necessary for
Melt-Viscosity Characterization 53

stability reasons and is due to the shear-thinning effect on S (in par­


ticular, this fluidity factor increases as |Vpl increases). Otherwise,
overrelaxation would be possible, as in the case of Laplace’s equation.
Typically, all nodes get scanned four times before updating the value
of S at the centroid of each element (hence the value of in (114))
and updating the value of p Q (t) based on evaluating the integral in
(113). Updating PQÍt) also requires underrelaxation. This proce­
dure is repeated until the pressure field satisfies the specified con­
vergence criterion between successive scannings as well as the global
volumetric flow condition.
Once the temperature and pressure fields have been updated, we
must advance the melt front for the next time step. We do this by
evaluating v^ = -(S / b )O p / 9 n ) at each vertex node along Cmf and de­
termining the advanced position of these fluid particles for a given
time increment. Further detaüs are given in [128] , where it is shown
that a two-step, predictor-corrector scheme is appropriate to give
more stable melt-front contours. It is the melt-front advancement
which greatly complicates the problem, particularly from the stand­
point of automating the numerical procedure. For example, when the
melt front impinges on an impermeable boundary, the explicit formula­
tion in [128] requires direct intervention by the user to adjust the
new melt front on the basis that it remain within the cavity and sat­
isfy conservation of mass as well as be perpendicular to the imperme­
able boundary (which follows from the observation that the melt front
is an isobar such that the velocity is normal to the front but must
also be tangential to the impermeable boundary). An associated com­
plication is the addition of new elements to the system as the melt
front advances. Fortunately, such complications have been eliminated
in the formulation described by Wang and Wang in Chapter 9. That
is, following the work in [129] , they use a fixed finite-element mesh
and associate a control volume with each vertex node together with a
scalar function f, which defines the fraction of the control volume
presently filled with melt, so that the melt-front advancement be­
comes amenable to an automated algorithm, although with some loss
in resolution.
Thus, it is seen that, in addition to making it possible to handle
arbitrary planar geometry, the finite-element formulation in the x -y
plane allows nodes along impermeable boundaries and weldlines to be
processed just like interior nodes ( ’’natural” boundary conditions).
This is an advantage over the finite-difference scheme, where such
boundaries would require special treatment with explicit representa­
tions for Sp/8n. Also, solving the resulting nonlinear system by re ­
laxation has the advantage of avoiding any concern with the order in
which the nodes are numbered (unlike direct solvers, where ’’band­
width” considerations are of paramount importance to execution time).
m Hieber

Fig. 25 Calculated instantaneous contours of constant normalized


pressure (aolid) and stream function (dashed) for n = 1 under iso­
thermal conditions. Attachment stream function ( — ) : ~ 58%.
Absolute length scale is arbitrary.

Furthermore, storage requirements are minimized, because no sub­


scripted variable varies more than linearly with the size ( i . e . , total
number of nodes or elements) of the system.
Some characteristics of the governing equation for the pressure
field, (98), are illustrated in Fig. 25, which shows resulting calcu­
lated isobars (solid curves) at a certain instant for a hypothetical
flow situation involving a rectangular cavity with an insert and as­
sumed melt front. In particular, a uniform normalized pressure of
unity has been imposed along the circular arc (C g ) in the lower left
corner with a pressure of zero along the melt front at the right. The
resulting intermediate contours have been obtained for a Newtonian
fluid under isothermal conditions such that (98) and its resulting nu­
merical form, (114), are then equivalent to Laplace^s equation. Based
on the calculated pressure field, the associated stream function has
then been constructed by making use of (96), from which it follows
that

^2 (2bv ) ds ds (115)
n ((= i)
Melt-Viscosity Characterization 55

In particular/the resulting dashed contours in Fig. 25 correspond to


normalized values of in increments of 0.2 of the total flow rate, with
about 58% of the flow going under the insert.
As an indication of the shear-thinning effect of the viscosity upon
the flow field, Fig. 26 shows corresponding calculated results for the
same situation as in Fig. 25, but with n = 0.3 rather than 1.0. In
particular, note that the shear-thinning effect increases the flow rate
going under the insert from about 58% to 70% (or, in terms of the ratio
of the flow rate below to that above, from about 1*4 to 2.3). Indeed,
this is what would be expected, since the shear rate in the Newtonian
case is already higher in the region below the insert, such that any
shear thinning would therefore have a larger effect in reducing the
flow resistance in that region. Similarly, since the overall highest
shear rates occur in the vicinity of the ’’gate” it follows that the
shear thinning should cause the greatest reduction in flow resistance
in this latter region, which is reflected by much lower pressure gra­
dients in the lower left region for n = 0.3. On the other hand, since
the overall normalized pressure drop is the same in both cases, it
follows that the pressure gradient increases in the slow-moving re ­
gion at the right, where the shear thinning has its smallest effect.
In general, then, we can conclude from Figs. 25 and 26 that the
shear-thinning effect of the melt viscosity can significantly affect the
pressure and velocity distributions. We emphasize this point because
several published works assume that variations of S in (98) can be
56 Hieber

F ig . 27 Plan view and side views for cavities (a ) and ( b ) , including


location of three flush-mounted pressure transducers; dimensions
given in units of mm. (From Ref. 131.)

neglected, thus giving Laplace’s equation for pressure. Although


such an approximation greatly reduces the numerical problem, the
resulting inaccuracies can be substantial, which is reflected by the
differences between the distributions in Figs. 25 and 26. Further,
in actual applications, nonisothermal effects arise which contribute
to additional variations in S . This has been shown in Section A and
will be further illustrated in the following example, in which varia­
tions in S also arise from a variable cavity thickness.
Figure 27 shows the mold configuration considered in [131], which
consists of two variable-gap-thickness cavities, (a ) and ( b ) , with the
latter also containing a circular insert, as indicated. In particular,
results based on the power-law model, (46), are presented in [131]
for PP [Profax 6523 (Hercules)] and PS [Styron 678U (D o w )]. Fig­
ure 28 shows the predicted advancing melt front and finite-element
mesh generated for cavity (b ) being filled with PP under the flow
conditions indicated. The simulation assumes that the rectangular
reservoir (at left of Fig. 27) gets completely filled before the melt
enters the thin cavity. Further, pressure variations in the large
cross-sectioned reservoir are neglected, such that the pressure
along the entrance to the cavity (x = 0, in Fig. 28) is assumed to
be uniform. The actual calculation in this case was begun at t =
0.093 s, with the melt front being linear and the melt domain con­
sisting of 15 elements. This initial melt front directly reflects the
Melt-Viscosity Characterization 57

3.81

6.99
X (cm)

Fig. 28 Predicted melt-front positions (heavier lines) and underly­


ing finite-element grid for PP filling cavity (b ) with T q = 197°C,
Tw = 27°C and Q = 7.1 cm^/sec. Dots denote locations of transdu­
cers 2 and 3; predicted weldline denoted by +hh- . (From Ref. 131.)

variation of the gap wise thickness (shown cross-hatched at right of


Fig. 27); in particular, under isothermal conditions and assuming a
uniform pressure along x = 0 with no flow in the y direction, it fol­
lows that the speed of the melt-front advancement will be linearly
proportional to the gap thickness. (This is completely analogous to
results relating to the coupled-flow-path approach in Section B; spe­
cifically, when the melt front reaches a branch point, the initial dis­
tribution of flow among downstream segments is proportional to wjbj^
if all are rectangular strips, such that the bulk velocity Uj = Qj/wjbj
is linearly proportional to bj, the gap thickness.) As the calculation
proceeds, however, we see that there is substantial flow in the y di­
rection, which is enhanced by significant cooling in the thin-sectioned
portion of the cavity. The melt along y = 0 substantially stagnates
due to cooling, thus requiring cross flow from the thicker sections to
fill the domain along y = 0. Due to this phenomenon, in this calcula­
tion we have relaxed the usual constraint that the melt front be per­
pendicular to impermeable boundaries. This is also borne out by the
experimental short-shot contours in Fig. 29, where the solid curves
meet the lower boundary at other than a 90° angle. In general, the
58 Hi'eber

Fig. 29 Comparison between predicted melt fronts (dashed) and ex­


perimental short shots (solid) for same conditions as in Fig, 28. Ex­
perimental weldline denoted by 4H4-. (From Ref. 131.)

predicted shape of the advancing melt front in Fig. 29 is in reason­


able agreement with the experimental short shots, particularly with
regard to the position of the weldline. As a more quantitative com­
parison, Fig. 30 shows the experimental pressure traces at transdu­
cers 1—3 (locations indicated in Fig. 27) together with corresponding
predictions. In general, the agreement concerning pressure level Is
quite good. One systematic discrepancy, namely the underprediction
of the pressure at transducer 1, might be attributed mainly to junc­
ture losses between the reservoir and cavity which would be reflected
in the experimental value, but have not been included in the simula­
tion. In addition, the change in slope of experimental trace 1 at t
0.39 sec apparently corresponds to when the reservoir fills. Lastly,
the discrepancy between the predicted and experimental times at
which transducers 1 and 2 lift off might be attributed mainly to a
volumetric flow rate which actually decreased with time. In this re ­
gard, note that Ap in the power-law region; hence, with n «
1/3 in the present case, a 15% change in Q will cause only a 5% change
in pressure. That is, uncertainties in Q could cause the discrepancies
concerning time of lift off of various transducers without having much
effect on the corresponding pressure levels.
As a more detailed description of the foregoing predictions. Fig.
31 shows the simulated isobars at time of fill, when the required gate
Melt-Viscosity Characterization 59

Fig. 30 Comparison between predicted pressures at gate (A ) and


transducers 2 (o) and 3 (+) with experimental traces (curves) for
same conditions as in Fig. 28. (From Ref. 131.)

pressure along x = 0 is 1470 psi. Also shown are some correspond­


ing predictions under isothermal conditions (dashed contours for 500
and 1000 psi) for which the required gate pressure is 1110 psi. In
assessing nonisothermal effects for such variable-gap-thickness cavi­
ties, we find it convenient, when employing the power-law approxima­
tion, to scale out the gap-thickness dependence of $ by introducing
as in (61). In particular, predicted contours of constant values
of at tfiii are shown in Fig. 32, where the small values in the lower
left indicate a highly resistive, low-temperature region. When com­
paring Figs. 31 and 32, recall [from (60)] that the pressure gradient,
hence the pressure drop (for fixed spatial domain), is proportional to
(^ * )" ^ . Hence, since the nonuniform field in Fig. 32 has an asso­
ciated overall Ap of 1470 psi, as compared with a value of 1110 psi
when = 1, it follows that the effective average value of is
60 Hieber

Fig, 31 Predicted isobars (6.9 x 10® dyne/cm^ = 10^ psi) at tfiil


0.69 sec with dashed curves corresponding to 5 and 10 x lo^ psi in
isothermal case; same flow conditions as in Fig. 28.

J* ~ (1110/1470)^^°*®^® - 0.42, which is compatible with the dis­


tribution in Fig. 32, particularly since the bulk of the flow passes
above the insert.
Finally, Fig. 33 plots predicted contours of constant wall-shear
stress, T^, corresponding to the nonisothermal pressure field in
Fig. 31. In this regard, it follows from (92) that = Ab, where

Fig. 32 Predicted contours of constant at tfj^i 0.69 sec; same flow


conditions as in Fig. 28.
Melt-Viscosity Characterization 61

*Fig. 33 Predicted contours of constant wall shear stress (10^ dyne/


cm^) at tfiii = 0.69 sec; same flow conditions as in Fig. 28.

A E |Vp| can be calculated from the pressure field and 2b(y) corres­
ponds to the cross-hatched cavity thickness shown at the right of
Fig. 27. From the resulting values in Fig. 33 we see that the bulk
of the material is subjected to a wall shear stress in excess of 1 x 1 0 ®
dyne/cm^. On the other hand, the present PP material, in terms of
the four-constant (n , t *, B, model, corresponds to a i * of
1.14 X 10^ dyne/cm^. Hence, (T^lavg/^^* ^ 8 , such that we would
expect that the power-law approximation should be quite accurate in
the present case. We have verified this by repeating the calculation
for the pressure field at tfm, using the same mesh and temperature
distribution as before, but now replacing the power-law asymptote
(46, 82) by the underlying four-constant model with

( 1 I 3 T * , B, T ^ ) = 1^0.323, 1.14 X 10= 2 , 9.0 , 4550 K


cm cm •sec

The resulting predicted overall pressure is within 2% of the above


1470 psi, confirming that the power-law approximation is indeed ac­
curate in the present case.

IV . P O S T -F IL L IN G STAGE

Since the pioneering work of Gilmore and Spencer [1 ,2 ], investiga­


tors have considered the development of the cavity pressure distri­
bution during the post-filling stage. Notable examples include [ 6 ,
137—141]. In particular, the most recent investigations have been
62 Hìeber

concerned with simulating the brief post-filling ( ’’packing”) stage, in


which the nonuniform pressure field developed during the filling stage
quickly approaches a uniform packing-pressure level by a diffusion-
like process. Due to the shortness of this stage, these simulations
have typically neglected nonisothermal effects; further, the viscos­
ity has usually been taken to be Newtonian, on the assumption that
the associated shear rates are small.
As an illustration, we might consider the case recently treated in
[141], namely that of a Newtonian fluid under isothermal conditions
with the equation of state corresponding to that of Spencer and Gil­
more [142]:

(P (116)

where p, p, R are constants. Any fluid motion which arises after


tfiii is due solely to variations in density, which, in turn, can be
due to pressure or temperature variations. In particular, if we con­
sider the case of a thin cavity, as in [141], then the local conserva­
tion of mass is given by

(117)

where w is here the velocity in the gap wise z direction. On the other
hand, if we neglect inertial and elastic effects, the force balance in
the X and y directions is the same as in (84) and (85) , such that u
and V are again given by (96), and S by (97). However, with T e
constant, and neglecting any pressure dependence in the viscosity,
we see that S = b^/3rio> where no is the uniform viscosity. Hence, in
this case.

-C ^ .c | E (118)
3y

where C e b^/3no« Further, since T is uniform and p is independent


of z (in the leading approximation, since the force balance in the z
direction indicates that pressure variations in the gapwise direction
are 0(b/L)^ smaller than in the x -y plane, where L denotes the char­
acteristic length in the latter p lan e), it follows that p is independent
of z so that integrating (117) across the gap thickness gives

+ 3-|(pu) + (p v ) + pw 1^ = 0 (119)
3t 3y
Melt-Viscosity Characterization 63

But w vanishes both at z = 0 (by symmetry) and at z = b (no pene­


tration), so the surface terms in (119) drop out. Hence, substitu­
ting (118) into (119) gives

( 120)

On the other hand, with T = constant, it follows from (116) that

dp\_
0 ( 121)

That is,

dp = r dp ( 122)

where

p +p
r E (123)
p ( i - p/p)

Hence, from (120) and (122) it follows that

(124)

Note, however, that

pV^p p ^

whereas

^ ^ ip ^ ^ ^
■ m 3x 9x L L

Hence, since Ap/p « 1, it follows that the last terms in (124) can
be neglected (analogous to the underlying approximation in acous­
tics) such that, again making use of (122), we see that (124) then
reduces to

^ ^ cnrrv ' ^ p (125)


9t
64 Hieber

where

p +p (126)
pr
1 - P/P

That is, the pressure field is governed by a diffusion equation, In


particular, the characteristic diffusion time is such that

At

that is.

At 3rio (127)
cr ir ff)© '
Accordingly, this characteristic time increases with increasing viscos­
ity or decreasing cavity thickness; that is, an increased flow resis­
tance means that the flow cannot as readily respond to the mass-flow-
rate demands associated with the varying density. On the other hand.
At decreases with increasing p, reflecting a decreasing density de­
pendence on pressure, such that mass-flow requirements associated
with a given change in pressure are decreased.
For P S , for example, representative values for the constants in
(116) are given by [142] :

(p, p, R) = (l.86 X 10® 1 . 2 2 - ^ , 8.00 x 105^^1:) (128)


^ \ cm® g-Kf

such that, typically, p + p ~ p and p/p « 5/6. Further, a typical


value for rio might be taken (see Fig. 6 of Section II.A ) as 3 x lo^
g/(cm*sec). Hence, with L 10 cm and b 0.1cm,

1/6
At ^ 10^ 10 ^ sec (129)
cm* sec 2 x lo^ dyne/cm'

That is, the transition from the nonuniform pressure field at tfip to a
uniform packing-pressure level should take about 0.1 s.
This result is illustrated by the experimental pressure traces in
Fig. 34, corresponding to the instrumented strip mold in Fig. 35,
which indicates the locations of the three flush-mounted pressure
transducers [PT 467E (D ynisco)]. In particular, after rising with a
fairly constant slope during the filling process, the pressure traces
in Fig. 34 rapidly increase and quickly tend to coalesce after tf^^.
Melt-Viscosity Characterization 65

Q.
to
O

Fig. 3 4 Pressure traces at transducers 1, 2, and 3 of Fig. 35 for PS


(Styron 678U/Dow). Cavity thickness = 0.254 cm (0.100 in .), To =
200°C, Tw = 32°C, velocity setting = 10%, hydraulic hold pressure =
6 . 9 X 10^ dyne/cm^ (1000 p si), hold time = 15 sec.

All three traces do not reach the same packing-pressure leve, however,
since the cavity pressures start to decay. This is shown more clearly
in Fig. 36.
In particular, we might suspect that such a decay in the cavity
pressure traces could be due to solidification of the melt over the
sensing element, with little direct correlation to the actual pressures
experienced within the bulk polymer. However, traces such as in
Fig. 36 are highly reproducible and show systematic trends with cav­
ity thickness. This is illustrated by the reduced experimental re­
sults in Fig. 37 for Api- 3 versus t-tfm for the preceding PS, as well
as a PP, with two different cavity thicknesses. For both polymers
the pressure decays sooner in the thinner cavity. In turn, this is
compatible with a smaller characteristic cooling time, which, for
transient one-dimensional conduction in a slab, is b^/a, where a is
the thermal diffusivity. Typically, a ~ 10“^ cm^ s"^ for polymers
such that, for the present cavity thicknesses, the characteristic
66 Hieber

Fig. 35 Sprue, runner, reservoir, and cavity of test mold with three
flush-mounted pressure transducers. Cavity thickness = 0.216 cm
(0.085 in .) or 0.254 cm (0.100 in .).

cooling time should be about 10 sec with that of the thinner cavity in
Fig. 37 is smaller by about 30%. It is noted that experiments con­
ducted with a gap thickness of 0.508 cm (0.200 in .) indicated essen­
tially no pressure decay while the packing pressure was held for 15
sec. In addition, note that the delayed sharp rise in the PP curves
in Fig. 37 is presumably associated with crystallinity effects.
The general picture, then, is that the post-filling phase is char­
acterized by a short packing stage (about 10"^ sec or even less, as
will be shown presently) and a relatively long ( « 10 sec) cooling
stage. Although most published work has been focused on modeling
the former stage, we mention the notable recent experimental and
numerical work of Deterre [143], which considers the latter stage
for both amorphous and semicrystalline polymers. In addition,
Tayler [144] has presented an analysis which delineates the leading-
order effects and associated model equations for all three stages.
Melt-Viscosity Characterization 67

- 6

Q.
ro
4 2

- 2

Fig. 36 Same as Fig. 34 but with longer time scale.

namely the filling, packing, and cooling phases, basically in terms of


a Newtonian fluid.
As one means of modeling some of these experimental results, we
might extend the one-dimensional-filling simulation of a strip cavity,
described in Section III.A , by incorporating compressibility effects.
In particular, if we still neglect inertial and elastic effects and treat
the fluid as being purely viscous, then the force and energy bal­
ances are still given by (44) and (45), respectively. However^ due
to a variable density, the volumetric flow rate is no longer uniform in
X, so (43) must be replaced by an appropriate mass balance. In par­
ticular, the local conservation of mass balance might be written as

3u
+ u (pv) = 0 (130)
8t ^ 3x 9x 9y

where the last term will drop out once we integrate across the gap.
On the other hand, at least during the post-filling stage.

9u U tt
L’
68 Hieber

Q.
<

Fig. 37 Experimental results for Api_ 3 vs. time for PS (Styron 678U/
Dow: open symbols) and PP (Profax 6323/Hercules: solid symbols)
for cavity thickness of 0.216 cm (triangles) and 0.254 cm (squares).
Contours are faired curve fits of the reduced data points.

such that the latter term is negligible, since Ap/p « 1 (in the same
way that the last terms in (124) could be neglected). Hence, omit­
ting the third term in (130) and integrating from y = 0 to y = b,
we get

du
dy = 0 (131)
3t 3x

However, since Ap/p « 1, we can simplify the second term in (131)


by treating p as some representative constant value (po); therefore,

allowing for the possibility that b = b ( x ) . Or, to the same accuracy,

8 In p _ _ 1 J. (132)
(b u )
8 t b 3x
Melt-Viscosity Characterization 69

On the other hand, by integrating (44) repeatedly with respect to y,


we get

- _ S ap
(133)
^ ' b 3x

as in (96) , with S again being given by (54). Hence,

9 In p i A / s 1e \ (134)
at b ax \ 3x/

In particular, if density variations are neglected, as in Section III,


then (134) reduces to the one-dimensional case of (98). However, in
this case we can integrate once with respect to x such that

-S ^ = constant = bu = b
3x 2bw

which corresponds to (55), with w denoting the width of strip.


However, if we do include variable-density effects, which are es­
sential for the post-filling phase, and if we base p(p, t) upon the
Spencer-Gilmore [142] equation of state, as given by (116), then

which can be rewritten as

d In p = p/p dp - d In T
p +p (‘ i )
Hence, (134) can then be expressed as

3p 1 ±/s ^ \ =
G (x , t) ■F(x, t) (135)
at b 3x\ dx)
where

(136)

and

In T
F (x , t) E - e dy (137)
3t
70 Hìeber

where p, but not T, is independent of y. Based on (116) and (128),


we observe that the density of PS will typically lie within the range of
0.90—1.00 g/cm^ for temperatures and pressures of interest. Hence,
to the same approximation involved in deriving (134), we could re ­
place the 1 - p/p factor in (136) and (137) by a constant value of ap­
proximately 1 - 0.95/1.22 = 0.22, thereby eliminating entirely the in­
tegral in (136).
Equations (135) —(137) can be used as the basis for a unified for­
mulation for the filling, packing, and cooling stages. For example,
during the filling stage we expect that both the G and F terms will
be negligible, with (135) being dominated by the diffusion term and
the flow being driven by a specified flow condition at the inlet,
namely

Qo (138)
-( 2wS
'x=0

where Qo denotes the constant volumetric flow rate supplied to the


cavity (note: since the pressure (hence, density) level at the cav­
ity entrance will be increasing as the melt front advances further in­
to the cavity, Qo will actually correspond to a slightly increasing mass
flow rate—a higher-order effect in our formulation) , and S, defined
by (54), must be determined by iteration based on the particular vis­
cosity model, n(Ÿ> T, p ), evaluated at the inlet melt temperature. To*
During the packing stage the G term becomes important, with F
still having a negligible effect. In particular, starting with the non-
uniform pressure field at tfm, the pressure now adjusts as the zero-
gage-pressure condition along the melt front gets replaced by an im­
permeability (no-flow) condition at x = L, namely

(^ \ = 0 (139)
T
^ ' x=L

and typically an abruptly increased packing pressure gets imposed at


X = 0. Throughout this phase, p will be increasing with T essentially
unchanged, such that the density will be increasing.
During the final cooling stage both F and G become important, with
F being the driving force. Both T and p will be decreasing during
this phase and have opposite effects upon the density.
Numerically, (135) is amenable to a tridiagonal form for pressure,
using an implicit formulation with finite-difference representations
such as in Section III.A . However, since S (and G, to a lesser de­
gree) depends on p, an iterative procedure is used. In particular.
Melt-Viscosity Characterization 71

for numerical stability it is necessary to underrelax the values of S


due to the shear thinning of the viscosity, as was the case for the
FEM—FDM formulation described in Section III.C . Further, T gets
updated in the same manner as for the strip cavity described in Sec­
tion III. A.
As quantitative illustrations, we now consider the strip mold in
Fig. 35 for PS [Styron 678U (D o w )]. In particular, two different
representations of the viscosity will be considered, namely models I
and III of Appendix J . For the present material we thus get

dyne
(n, T*, B , g) = ^0.297, 1.83 x 10®

2.80 X lo"^ 12,400 K, 3.4 x 1 0 ‘®


cm •sec

and

i 5 dyne
(n, T*, Di, Da, D 3 , A i, Aa) 0.284, 1.78 x 10
cm^ ’

1.75 X 10^ 100°C,


cm •sec

-e c m ^ * ° C
2.3 X 10 27.2, 60.0°C (141)
dyne ’

Both sets of values were obtained by using the simplex—finite-differ­


ence program described in Appendix J as applied to the cap illary-
rheometer data in Fig. 1 of Section II.A . That is, end losses have
been neglected, but pressure-dependence and viscous-heating ef­
fects in the data have been accounted for by the finite-difference
simulation. For (140), the value of 3 has been held fixed at a repre­
sentative value based on the results in Fig. 11 of Section II.C . On
the other hand, the values of D 2 , A^, and A 2 in (141) have been
fixed based on the results in Fig. 46 of Appendix C, with the value
of D 3 specified on the basis of Fig. 11. Figure 38 shows the plots
based on (140) and (141). As we would expect, both fits agree quite
well over the experimental temperature and shear-rate range. How­
ever , as we extrapolate to lower temperatures, model III gives a sys­
tematically higher viscosity. In turn, this reflects an increased tem­
perature sensitivity at low temperatures, as shown by the upturn in
curve 5 in Fig. 46, as well as an increased pressure sensitivity, as
shown by curve 3 in Fig. 11. Further, the solid curve 4 in Fig. 38
72 Hieber

Fig. 38 Viscosity curves based upon models I (dashed) and III (solid)
for PS (Styron 678U/Dow) at values of (T , p ) as follows: 1 (230°C, 1
b a r), 2 (180°C, 1 b a r), 3 (150°C, 1 b a r ) , 4(150°C, 500 b a r).

indicates a power-law behavior essentially all the way down to a shear


rate of 10"^ sec"^. turn, this apparently reflects a shear-stress
level that is still significant relative to t *.
Focusing our attention first on the packing stage, we see in Figs.
39 and 40 the resulting predictions for the pressure at transducer 3
of Fig. 35 for two different packing pressures, as indicated. In both
cases the results correspond to the tinner cavity [b = 0.108 cm/2b =
0.085 in .] under isothermal conditions. That is, nonisothermal effects
have been omitted by setting Tw = 200°C = T q and omitting the viscous-
heating term from (62), which gives T = 200®C throughout. During
the filling stage both F and G in (135) have been omitted, with the
pressure field determined succesively as the melt front advances, and
Qo specified as 10 cm^/sec in (138). Based upon model I with the con­
stants in (140), this results in predicted pressures at tfj^j = 0.628 sec
Melt-Viscosity Characterization 73

o.
M
g
lO
o.

Fig. 39 Predicted pressure trace at transducer 3 in Fig. 35 during


isothermal packing with pack pressure = 6.31 x lo ’' dyne/cm^ (914
p si), corresponding to pressure drop at tfui for case in which Qo =
10 cm^/sec, T = 200°C, b = 0.108 cm (2b = 0.085 in .) . Results are
for PS (Styron 678U/Dow) based upon model I (curve 1) and its New­
tonian (curve 0) and power-law (dashed) limits, with primed curves
denoting omission of pressure dependence (3 = 0).

of 914 psi for transducer 1 in Fig. 35 (corresponding to x = 0 in sim­


ulation; i . e . , any pressure loss in reservoir or between reservoir and
cavity is neglected) and 600 (184) psi for transducer 2 (3 ). Subse­
quently, the peak pressure (at x = 0) is either maintained at 914 psi
(Fig. 39) or else instantaneously raised and held at 7250 psi (Fig.
40), corresponding to the experimental levels attained in Figs. 34
and 36.
The calculation for t > tf^p has then proceeded by including G in
(135) and employing various forms of model I, as indicated in Fig. 39.
In particular, an exponentially increasing temporal mesh size has been
used ( f E t - tfj2 i> increasing by a factor of 10 typically every 10 time
n Hieber

Fig. 40 Same as Fig. 39 but with pack pressure = 5 x lo® dyne/cm ^


(7250 p si).

steps, starting with At = 10~^ sec) together with 18 fairly uniform


mesh increments in the streamwise direction and nine nonuniform in­
crements in the transverse direction (defined by (68) of Section
I II.A ).
Observing the results in Figs. 39 and 40, we note that including
the pressure dependence leads to a slower diffusion process in all
cases, due to an increased viscosity [cf. (127)], with the effect evi­
dently being larger in Fig. 40 due to a higher pressure level. Fur­
ther, it is noted that the 0^ curve, which corresponds to a constant
viscosity throughout (namely, n = B exp (T]3/473 K) =6.80 x 10^
gm/(cm « s e c )), has essentially the same characteristic time in both
figures, with the process being fairly well completed by t ^ 10”^ sec,
in agreement with the estimate in (129).
For the power-law approximation, the dashed curves agree quite
well (markedly better in Fig. 40, due to a larger pressure drop,
hence larger shear-stress level) with the full model initially and only
break down toward the end as a uniform pressure is approached. In
particular, when the overall pressure drop in the cavity has dropped
to 1.5 X 10^ dyne/cm^ ( ~ 200 p si), A ^ Ap/L ~ 2 x 10® dyne/cm ^ and
Melt-Viscosity Characterization 75

= Ab ^ 2 X 10^ dyne/cm ^ « t *. Hence, we would expect the


power-law approximation to become inaccurate as the pressure field
approaches within 200 psi of the asymptotic packing level. On the
other hand, note from Figs. 34 and 36 that the experimental pressure
traces indicate that Ap does not get below about 400 psi before cool­
ing effects cause the pressure drop to increase. Hence, we would
expect the power-law approximation to be sufficiently accurate in
the actual case, which has also been confirmed by Deterre [143].
Curve 1 indicates a markedly quicker response for Fig. 40, which
is evidently due'to shear-thinning effects. In particular, in this lat­
ter case the entrance pressure gets abruptly raised at tfjii from
6.31 X 10^ dyne/cm ^ (914 psi) to 5 x 10® dyne/cm^ (7250 psi) such
that, in effect, the characteristic pressure gradient, Ap/L, is in­
creased by a factor of about 8. Since the full model is essentially
in the power-law region in this case (see Fig. 40), it follows that
the characteristic velocity increases by a factor of 8^/^, which is
approximately 10^ for n = 0.297. Hence, since U ~ 10 cm/sec dur­
ing filling, U ^ 10^ cm/sec during the brief packing.
Alternatively, if we generalize (127) to the shear-thinning case,
using (135), it follows that

bL^ 1 - p/p
At (142)
s p + p

In particular, for the power-law limit under isothermal conditions, it


follows from Section III.A that

.1/n
S = /— \ ^ /Ab\ (143)
\2n+l/ A Vm /

where, in terms of the limiting behavior of model I,

m(T, p ) = b ” exp|— exp( n3p) ( T*) ^ ^ (144)

Hence, from (142) and (143) it follows that

\ l / n , . /T X 2

At ^ (145)

In particular, if n = 0.297, then At is proportional to indicating


a strong shear-thinning effect. Indeed, it is this factor which accounts
for the much quicker response of curve 1 in Fig. 40. Specifically,
76 Hieber

4 *^0
cT

Fig. 41 Comparison between experimental traces (curves) for trans­


ducers 1—3 of Fig. 35 and corresponding predictions (symbols) based
upon model I (open) and III (solid) for PS (Styron 678U/Dow); b =
0.108 cm (2b = 0.085 in .), T q = 200°C, = 32°C, Q q = 10 cmVsec;
trace 1 used as imposed pack pressure for t > tfiu.

with A typically larger by a factor of 8 in Fig. 40, it follows that the


diffusion time should be smaller by a factor of approximately
« 10”^, which is roughly borne out by the results for curve 1 in Figs.
39 and 40. Further, if we evaluate (145) for the conditions in Fig.
40, taking A - 5 x 10® (dyne/cm^)/7.62 cm - 7 x 10^ dyne/cm® and
p 2.5 X 10® dyne/cm^ as representative values, together with p -
0.95 g/cm®, it then follows from (140), (144), and (145) that At 10”^
sec with a representative shear rate of (Ab/m )l/^ « 2.5 x 10^ sec'^
(cf. (56) of Section III.A ) and characteristic velocity on the order of
bY/3 - 10** cm/sec, in agreement with the previous estimate.
Such large velocities arising within such a short time indicate that
inertial effects, particularly p(9u/9t) in the x-force balance, may be­
come important. Such effects have been considered in the extensive
analysis by Kuo and Kamal [138]. Here, however, concern with the
behavior at the smaller times in Fig. 40 is not really of practical in­
terest, since it requires that the rise time for the imposed packing
Melt-Viscosity Characterization 77

5 i
Q.
ro
4 9

Fig. 42 Same as Fig, 41 but for b = 0.127 cm (2b = 0.100 in .),

pressure be even smaller. In the detailed simulations to be presented,


the specified pack pressure is based on experimental trace 1, for
which the rise time is about 10”^ sec, as shown in Fig. 34. Hence,
although the experimental packing-pressure level corresponds to
that in Fig. 40, the actual response time of the system may be closer
to that in Fig. 39, for which the characteristic diffusion time is itself
about 10"^ sec. That is, if the rise in the packing pressure occurs
on a time scale of 10"^ sec, the system should be able to respond al­
most instantaneously, such that the maximum pressure drop (hence,
pressure gradient) during the packing phase will be on the same or­
der as during the filling. This is consistent with the almost coinci­
dental rise of traces 1—3 in Fig. 34.
Finally, detailed comparisons between the preceding formulation
and the current data are shown in Figs. 41 and 42 for PS. In both
cases, F and G have been omitted during the filling stage but re­
tained for t > tfiii, with the temperature field being updated through­
out, based on the procedure described in Section III.A . The calcu­
lations have been continued until the centerline temperature drops
below 100®C ( T g ) , corresponding to t - tfin “ 7 sec (10 sec) in Fig.
41 (42), beyond which one could not expect the present viscosity
78 Hieber

modeling to apply. Indeed, the comparisons are very encouraging, in­


dicating that the nonuniform pressure field during the cooling phase
can be simulated quite well, provided we use a model such as III, which
should be fairly realistic for amorphous polymers down to Tg. Below
that, evidently memory effects would have to be incorporated to han­
dle the final residual cavity pressures evidenced in Figs. 36 and 41.

A p p en d ix A . DEPENDENCE OF n UPON M ^ /M n

In assessing a possible dependence of the power-law index upon molec’-


ular-weight distribution, we show in Fig. 43 values for 1-n versus

Fig. 43 Plot of (1 -n) vs Mw/Mn for (a ) PS, (b ) PP, and (c ) HDPE.


Best-fit curves correspond to (146) with = 0.87 (a ), 0.97 ( b ) , and
0.95 ( c ) . Two blends (<>) at right of (a ) and four lowest points (all
from same source) as well as point at right of (c ) have been omitted
in obtaining best fits.
Melt-Viscosity Characterization 79

Mw /Mr for those PS, PP, and HDPE data sets of Section II.A for which
Mw/Mn is known. Despite evident scatter, the results are compatible
with a functional dependence of the form

0.2

■"=M i) (146)

with Cl 0.87, 0.97, and 0.95 for PS, PP, and HDPE, respectively.
Such a negative one-fifth power was obtained by Cross [80] but with
a viscosity model somewhat different than ( 1 ), namely

no - Hop
n = r io (147)
1 -n
1 + (noY/T^*)

and with the result that Ci = 1.00. For model (1 ), a zero value of n,
corresponding to Ci = 1 and Mw/Mn = 1 , is not acceptable since for
sufficiently large shear rate it would result in the shear stress being
independent of shear rate, which corresponds to a physically unstable
(indeterminate) situation.

A p p en d ix B . DEPENDENCE OF x * UPON M w /M p

Figures 44 and 45 show results in terms of x* versus M^/Mn for those


PS and PP data sets of Section II .A which lie not entirely in the power-
law region (such that x* should be well defined) and for which M^/Mr
is known. In both cases the results can be approximated by the func­
tional form

/M
(148)

with values as given. The same functional form was found by Panova
et al. [91] to apply to commercial-grade PMMA with (C 2 , C 3 ) -(7 .4 8
X 10^ dyne cm"^, 1.69), as given in (18) of Section II.B .
In Fig. 44, some of the PS blends are exceptional (as in Fig. 43 also,
although to a lesser extent). In this regard, extensive experimental
results of Van Krevelen et al. [145] indicate that higher-order molec­
ular-weight ratios can be correlated in terms of M^/Mn for commercial
grade polymers but usually not for polymer blends. Therefore, where­
as higher-order molecular-weight ratios have been used (e . g . , [55,
146,147]) to correlate Je° (hence, x*, based on Appendix E ), the re ­
sults from [145] then indicate that such results should also correlate
80 Hieber

Fig. 44 Plot of T * vs. MyqiMn for PS data sets (including HIPS: ❖ ).


Best-fit curve corresponds to (148) with ( C2, C3) = (1.02 x 1 0 ®
dyne/cm^, 1.52), for which the two blends (i>) at right have been
omitted.

with Mw/Mn, at least for commercial grade polymers. The lack of


correlation for HDPE (not plotted) could therefore be due largely to
experimental-grade polymers and blends.

A p p en d ix C . DEPENDENCE OF T b ON
TEM PER ATU RE LEVEL

To assess the dependence of Tb on the temperature level, we plot


results in Fig. 46 for those PS data sets from Section II.A which in­
volve more than one melt temperature. Specifically, the points plotted
in Fig. 46 correspond to local-Tb fits involving only two successive
temperature levels within any given data set. For example, for the
Melt-Viscosity Characterization 81

Fig. 45 Plot o f T* versus Mw/Mn for PP data sets. Best-fit curve


corresponds to (148) with (C 2 , C 3 ) = (1.48 x 10® dyne/cm 1.29).

PS data set in Fig. 1, four-constant fits have now been obtained sep­
arately for the (180°C, 200°C) data ànd for the (200°C, 230°C) data,
resulting in respective values for of 16,900 K and 10,100 K, which
have been plotted in Fig. 46 at T = 190°C and 215^C, respectively.
In particular, the smaller at the higher temperature is compatible
with the general trend evidenced by Fig. 46. In turn, this downward-
sloping trend indicates the inadequacy of the Arrhenius-type model,
( 2 ) , when applied over too large a temperature range.
As a more accurate representation of the viscosity temperature sen­
sitivity, the curves shown in Fig. 46 are all based on the WLF-type
[94] model, as given by (36) of Section II.C , but with no pressure
82 Hieber

o
E

o
u

Fig. 46 Viscosity temperature sensitivity of PS: comparison between


various WLF curves and local four-constant fits of PS data sets from
Section II.A , including HIPS ( ❖ ) , GFF ( ❖ ) , and NMWD (<>). Curves
are based upon (149) with (A i, A 2 , T *) = (40.1, 51.6°C, 100°C) : 1
(classical WLF from Ref. 94 with T * = Tg = lOO'^C from Ref. 97); (31.5,
50.0°G, 100°C): 2 (A^ and A 2 specialized for PS from Ref. 96); (20.4,
101.6 °C, 140°C): 3 (»’Tg" version from Ref. 94, with T * = Tg = 140°C
from Ref. 97); (26.7, 57.3°C, 100®C) : 4 (best fit of all points); (27.2,
60.0^C, 100°C): 5 (best fit without HIPS and GFF); (24.3, 56.0®C,
100°C) : 6 (best fit of HIPS alone) .

dependence included. By comparing the expressions for ag in (35)


and (38), we see that the local value of based on the WLF func­
tional form corresponds to
Melt-Viscosity Characterization 83

Fig. 47 Same as Fig. 46 but for PC with curves corresponding to


(A i, A 2 , T *) = (40.1, 51.6°C, 144°C): 1 (classical WLF from Ref.
94 with T * = Tg = 144°C from Ref. 96); (34.5, 72.0®C, 144°C) : 2
(A i and A 2 specialized for PC from Refs. I l l , 96); (37.5, 23.3°C,
I 4 4 0 C ); 3 (A i and A 2 specialized for PC from Refs. 46, 96); (25.9,
61.2°C, 144^0 : 4 (best fit of present results).

AiAaT^
T^ao = (149)
[A 2 + (T - T * )]'

Therefore, curves 1—6 in Fig. 46 are all based on (149) with various
sets of values for (A^, A 2 , T * ), as indicated. The suggested best
fit is curve 5 for PS ( i . e . , BMWD or NMWD) and curve 6 for HIPS.
Analogous results for PC and PMMA are shown in Figs. 47 and 48,
where the best fit is given by curve 4 in both instances. On the
other hand, corresponding results for semicrystalline PP, shown in
84 Hieber

Fig. 48 Same as Fig. 46 but for PMMA with curves corresponding to


(A i, A 2 , T *) = (40.1, 51.6°C, 104°C): 1 (classical WLF from Ref.
94 with T * = Tg = 104^C from Ref. 96) ; (32.4, 90.7°C, 104°C) : 2
(A i and A 2 specialized for PMMA from Refs. 51, 96); (20.4, 101.6 °C ,
160°C): 3 ( ”T s” version from Ref. 94 with T * = Tg = 160^C [9 7 ]);
(38.2, 101.3°C, 104°C): 4 (best fit of present results).

Fig. 49, indicate that is essentially independent of temperature


level and well approximated by a value of 5500 K, in agreement with
Van der Vegt [52]. On the other hand, the classical WLF represen­
tation does predict the level of Tb fairly well, as indicated by the
curve in Fig. 49, together with a fairly weak variation with tempera­
ture (due to T - Tg being so large). Although the WLF equation is
usually considered to be mainly applicable to T - between 0® and
Melt-Viscosity Characterization 85

14 -
o
12 f
E
o*
10 o

UJ

Fig. 49 Same as Fig. 46 but for PP with curves corresponding to


(A i, A 2 , T *) = (40.1, 51.6^C, -10°C), i.e ., the classical WLF from
Ref. 94 with T * = To* = -10®C from Refs. 114, 115.

° C , it can still be useful above that range, as demonstrated,


1 0 0

by Menges [114,115], and Shenoy [31,148], and their co­


workers .

A p p en d ix D . DEPENDENCE OF no UPON

Probably the most well-known relationship concerning melt-viscosity


characterization is that between no namely

no = c,M ^3--* (150)

which typically applies when is sufficiently large (such as for


commercial grade polymers). In particular, this 3.4-power relation­
ship can be traced back at least to the classical work of Fox and Flory
[149,150].
86 Hieber

Fig. 50 Dependence of fio upon for PS, including NMWD (<►),


HIPS ( ❖ ) , blends (i>) and reduced data ( ♦ ) from Fox and Flory
[149, 150] , Best fit (omitting O points) corresponds to (150) with
Cij = 3.0 X 1 0 ~^^ g/(cm*sec).

Figures 50 and 51 are plots for no = no(T) versus for PS (T =


210®C) and HDPE (T = 190°C) based on those corresponding data sets
in Section II,A which lie not entirely in the power-law region and for
which Mw is known. Despite evident scatter, (150) describes the re ­
sults in Figs. 50 and 51 fairly well. For PS representative points
from Fox and Flory [149,150] have been included, after reevaluating
their molecular-weight values based on the more recent intrinsic-vis­
cosity correlation in [151] and after using curve 5 in Fig. 46 to trans­
form from 217 to 210°C. Curve 5 has also been used to convert other
points in Fig. 50 when the data set corresponds to only one tempera­
ture (such that the temperature sensitivity cannot be deduced directly
from the data set). On the other hand, results for HDPE in Fig. 51
have been transformed to 190®C, based on a representative value for
Tjj of 3200 K (see Fig. 5), when the data set corresponds to only one
temperature.

A pp en dix E . R E L A T IO N S H IP BETWEEN
T * and Je°

Prest, Porter, and O’Reilly [154] have developed an empirical correla­


tion for polymer melts which indicates that
Melt-Viscosity Characterization 87

Fig. 51 Dependence of Hq upon for HDPE, including additional


data points from Miltz and Ram from Ref. 36: ©; 'iuhg from Ref.
152: Schreiber from Ref. 153: ® . Best fit (of all points) cor­
responds to (150) with Cn = 8 . 1 X 1 0 "^^ gy(cm*sec).

- at T]Qb0 = 1 (151)
MO ^

where n* is the dynamic shear viscosity at frequency o), with Jg® the
steady-state compliance (see p. 60 df Graessldy [147] for a very gdod
summary of ways to evaluate Je°) • By using the Cox-Merz [155] re­
lation, namely

at CO = Ÿ (152)

we get

^ =I at noY Je° = 1 (153)


no

By comparison, according to the viscosity model in (1 ),

noY (154)
— = I at
no 2

which immediately indicates that Je° and 1/t * are related. More spe­
cifically, if we introduce
88 Hieber

Fig. 52 Relationship between c = n/i1o ^ = t / t * based upon (1 ),


as rewritten in (156), for various values of n.

_ n _ noY ^ _ ny -
C , Oq = and o = — = (155)
no y X * T * T *

then Qq = a / c and ( 1 ) can be rewritten as

(156)

which is graphed in Fig. 52 for various values of n. In particular, it


then follows from (155) and (156) that

n / (l-n )
at ' (157)
no ^ ‘ H D
On the other hand, (153) can be expressed as

^ “ I at t = I (158)
rio "5 e 3
Melt-Viscosity Characterization 89

Fig. 53 Je° versus Mw/M^ for PS and PP based upon (146, 148, 159)
and respective values for (C i, C2, C3) as given in Appendixes A
and B .

Hence, from (157) and (158) it follows that

(159)

To illustrate, if we use the correlations in Appendixes A and B for


n and x* as functions of M^/Mn> we can then use (159) to generate
results for Je° versus This has been done in Fig. 53 for PS
and PP. In addition, if we use the values for (n , x*) from the mas­
ter plots in Table 1 of Section II.B , it then follows from (159) that
Je° = 9 . 3 X 10"^ (PS-BMWD), 2.1 x 10"^ (PS-NMWD), and 1 . 7 x 10"^
(PP) cm^/dyne. From the curves in Fig. 53 it then follows that
90 Hieber

Fig. 54 Experimental results for versus for PS-NMWD based


upon shear creep (A: from Ref. 156), stress relaxation ( o : from
Ref. 157, > : from Ref. 158), dynamic data ( o : from Ref. 146, □ :
from Ref. 154, +: from Ref. 159, ♦ : from Ref. 160), and steady-
shear data (< : from Ref. 64, v: from Ref. 161). Solid curve is
graphical fit whereas dashed line is based on (159) with (n , * =
t

(0.12, 1.05 X 10® dyne/cm^) , corresponding to master viscosity fit


for PS-NMWD in Table 1 of Section II.B .

representative values for M^/Mn in these cases are - 2.4 (PS-BM W D),
1.0 (PS-NMWD), and 5.6 (P P ).
As a direct check on (159), we consider PS-NMWD, for which ex­
tensive data exist in the literature for Je°. This is shown in Fig. 54,
which includes results based on various experimental methods, as in­
dicated. In particular, the dynamic-data results are based on the
relation

T 0 _ lim (160)
e 03-^0 (G ’’)^

where G’ and G^’ are the storage and loss moduli, respectively; the
steady-shear results are based on
Melt-Viscosity Characterization 91

Fig. 55 Results for * versus


t based on four-constant (n, * , B,
t

T ^) fits (1, 2) of the 19 PS-NMWD data sets reported in Table 1 of


Section II.B . Curve 1 corresponds to 1/Je° and curve 2 to ,
where n 0 . 1 2 , and Je°(M^) is given graphically by solid curve in

Fig. 54.

0 . lim
— (161)
~ T->-0 2t ^

where = Th - T2 2 is the first normal-stress difference. As noted


by Graessley [147] , Wales [64], and Janeschitz-Kriegl [162], Je° in
this case tends to approach a constant asymptotic value of approxi­
mately 1 . 8 X 10"® cm^/dyne for sufficiently large In particular,
the solid curve in Fig. 54 is a graphical fit of the experimental re­
sults which takes account of this limiting behavior. On the other
hand, by using the values for (n , x*) from the master viscosity plot
for PS-NMWD in Table 1, we obtain from (159) that Jq° - 2 . 1 x 10"®
cm^/dyne. This is shown dashed in Fig. 54 for 8 x 10** < < 4 .6 x
10®, which corresponds to the range of the 19 data sets used to gener­
ate the master plot for PS-NMWD in Section II.B . Alternatively, re ­
sults from fitting these 19 PS-NMWD viscosity data sets with (1, 2)
are shown in Fig. 55 in terms of x* versus M^. For comparison,
92 Hieber

Fig. 56 Steady-shear results for N i /2t^ versus based on PS-BMWD


t

data from Ref. 21: +; from Ref. 23: □ ; from Ref. 24: ■ ; from Ref.
25: < ; from Ref. 54: V; from Ref. 65: O; from Ref. 6 6 : A; from
Ref. 67: > ; from Ref. 163: ^ ; from Ref. 164: o. Dashed-dotted
line corresponds to onset of melt instability according to Gleissle [39]^
namely N i /t = 4.6, whereas dashed line is based on (159) with (n , x*)
= (0.25, 2 . 7 X 1 0 ^ dyne/cm ^), corresponding to master viscosity fit for
PS-BMWD in Table 1 of Section I I . B .

curve 1 corresponds to 1 /Je°> where Je°(Mw) is given by the solid


curve in Fig. 54; similarly, curve 2 corresponds to 21/(1"^)/Jg° =
2.20/Je°, ^ = 0.12. Hence, the results in Fig. 55 indicate reason­
able agreement with (159), although replacing 21/(1"^) with a value
of about 1.5 would be more appropriate in this case.
As a second check on (159), we consider PS-BMWD. Figure 56
shows plots of steady-shear experimental results for N i/ 2 x^ which,
according to (161) , should approach Je° for sufficiently small shear
stress. In general, however, the bulk of the results in Fig. 56 show
no sign of leveling off. Further, any eventual leveling off will clearly
be much higher than the level of the dashed line, even though we ex­
pect that the latter should represent the present PS-BMWD case.
Melt-Viscosity Characterization 93

Fig. 57 Values for Je° based upon various methods as applied to PS


data of Refs. 6 6 (Mw/Mn = 2.76) and 163 (M^/Mn = 4.1) as well as
PP data (Mw/Mn = 3.5 and 17) from Ref. 165. Solid curves for PS
and PP are same as in Fig. 53.

To investigate this dilemma further, we consider some of the more


extensive data sets which include both dynamic and steady-shear
measurements. In particular, Fig. 57 plots results for Jg° based on
various methods as applied to the PS-BMWD data from [ 6 6 ] and [163]
as well as the data for two PP melts from [165]. For the data from
[163] , all four methods, based on (151) (Note: 0 0 2 / 3 Fig* 57
94 Hieber

Fig. 58 Storage modulus G’ versus G* = / (GO^+(G")^ based upon PS-


BMWD data from Refs. 6 6 : A (T = 170°C, Mw/Mn = 2.76), Ref. 163:
A (170°C, 4.1), and Ref. 166: • (160°C, 2.9).

denotes the frequency at which n^/no = 2/3), (159), (160), and (161),
agree reasonably well with each other as well as with the PS curve
from Fig. 53. Concerning the data from [ 6 6 ] , however, we note that,
whereas the results based on (151) and (159) agree well with each
other as well as with the PS curve, the corresponding result based on
the limiting behavior of the dynamic data is high by about a factor of
4. To address this discrepancy in more detail, in Figs. 58 and 59 we
Melt-Viscosity Characterization 95

log,Q 6 * (dyne/cm )

Fig. 59 Same as Fig, 58 but for loss modulus G^’ versus G*.

plot the dynamic data from [ 6 6 ] and [163] together with correspond­
ing results from [166] for a similar PS-BWMD. [G * has been used as
the abscissa rather than oo so that the temperature effect scales out
(the data in [166] being at 160*^0, as compared with 170*^C in [ 6 6 ]
and [1 6 3 ]).] The dashed line at large G* in Fig. 58 corresponds to
G^ ^ G*, since G^^ « G’ in this range. On the other hand, G^ « G’^
for sufficiently small G*, such that G” ^ G* in this range, as Fig. 59
shows. Of prime interest here, however, is the behavior of G’ at
small G*. In particular, both dashed lines in the lower left of Fig.
58 correspond to a slope of 2, indicating that both the a and a data
extend into the quadratic range for G^(o)). However, the multiplica­
tive factor, corresponding to Je°, is larger by more than a factor of
2 for the results from [ 6 6 ] , even though M^/Mn is smaller. Hence,
these results from [ 6 6 ] and [163] are incompatible and, on the basis
of the corresponding results in Fig. 57, it appears that the low-G* a
points in Fig. 58 are the suspicious ones. That is, on the basis of the
A points in Fig. 58, and because the a data correspond to a smaller
96 Híeber

Mw/Mn» one would expect that, with decreasing G*, the latter points
would reach a quadratic behavior by G* 10** dyne cm”^ rather than
the 10^ dyne cm"^ range exhibited by the a results in Fig. 58.
In Fig. 57, all four methods agree fairly well when applied to the
PP sample from [165] for which M^/Mn = 3.5. On the other hand,
for the PP with a much broader molecular-weight distribution, neither
the dynamic nor the steady-shear data extend into the quadratic range
for G*((jo) or N i(y ) such that only results based on (151) and (159)
are given. In particular, the significant difference between the cor­
responding A and ■ results in this case is not surprising, given that
the associated limiting behaviors of n*(oi3) and ti( y ) give values for rio
which differ by more than a factor of 2, as shown by Fig. 2 of [165].
In particular, by comparing the a and ■ results at Mw/M^ = 17 with
the PP curve in Fig. 57, we see that the ti* ( íjü) results are more re ­
liable in this case; i.e ., the ri(y) curve levels off and approaches
its Newtonian limit at too large a shear rate, given the large value of
As a more detailed plot of the results from [165], Fig. 60
shows results for the Mw/Mn = 3.5 sample in terms of both dynamic
and steady-shear measurements. In particular, the dynamic and
steady-shear birefringence measurements agree fairly well, whereas
the cone-and-plate data show no tendency to level off with decreas­
ing T. In turn, such results seem to draw into question the reliabil­
ity of the corresponding steady-shear mechanical measurements at low
T in Fig. 56.
Figure 56 also includes a contour corresponding to the onset of
melt instability or melt fracture, according to the criterion docu­
mented by Gleissle [39]. In terms of this result, the large- t points
from [23] based on an indirect method of deriving from measured
exit pressures, seem to underestimate since they diverge from the
dashed contour, implying that the flow would always remain stable.
Other than those points, the bulk of the large- t points in Fig. 56
suggest crossing the dashed contour at t - 8 x lO^ dyne/cm^, which
seems to be a reasonable critical stress value in this case; this is
discussed further in Appendix G.

A p p en d ix F . JU N C TU R E -P R ES S U R E
C O R R E LA TIO N

In capillary rheometry, the extra (or juncture) pressure loss is usually


investigated by a Bagley [167] plot, i.e ., Ap versus L/D for fixed
y^ and T, with Ape corresponding to the intercept at L/D = 0. A l­
ternatively, such results are obtained (e . g . , [11,168]) by using
pressure transducers in the capillary or slit along the length of the
die and then deducing Ape by extrapolation. Typically, (e . g . , [22,
Melt-Viscosity Characterization 97

Fig. 60 Results for PP (Mw/Mn = 3.5) at 210®C from Ref. 165 based
on dynamic data ( • ) and steady-shear birefringence (▼) and cone-
plate ( ♦ ) measurements. Solid curves are curve fits of experimental
points. Upper and lower dashed lines correspond to 2 l/ (l" n ) /x* and
l/no^2 / 3 based upon respective fits of riiy) and n *(w ).

26]) results for Ape at different temperatures can be collapsed by


plotting versus t^ . Such a procedure is used in Figs. 61—64 for
representative Ape data from the literature for HDPE, LDPE, PP, and
PS, respectively. In each case the best fit has been generated based
on the functional form

^Pe “ ^5 w (162)

with resulting values for C 5 and Ce as indicated. Further, in an at­


tempt to possibly reduce the scatter, the experimental results have
been replotted in Figs. 65—68 in terms of Ape/i* versus with
the corresponding best fit based on the function al form

- (^) (163)
98 Hieber

Fig. 61 Ejttra pressure loss based on HDPE data fpom Ref. 11: a ;
from Ref. 15: ■ ; from Ref. 16: < ; from Ref. 22: V; from Refs.
23,168: > ; from Ref. 26: O; from Ref. 29: ▼; from Ref. 30:
from Ref. 32: +; from Ref. 33: ♦ ; from Ref. 34: o ; from Ref.
169: A . Best-fif curve corresponds to (162) with C 5 = 0.0271
(cm^/dyne) -^^^ and Cg = 1*399.

In general, though j (163) does little td improve the correlation. The


scatter is actually increased in the PS case^ for which the results in
Fig. 6 8 separiite into two groups with respective fits as indicated
(dashed).
Melt-Viscosity Characterization 99

Fig, 62 ilxtra pressure loss based on LDPE data from Ref. 11: a *
from Ref. 15: P ; from Ref. 16: < ; from Ref. 22: v ; from Ref. 26:
o ; from Ref. 34: <>; from Ref. 169: a ; from Ref. 170: ° . Best-fit
qU3?ve corresponds to (162) with C 5 - 0.161 (cm^/dyne)*^^^ and Ce =
1.321.

Nevertheless, the fits in Figs. 61” 64 can still be useful. For ex­
ample, if we introduce an effective extra capillary length, Le, based
on (7) with Ap replaced by Apg, then

= 1 ^ = £ 5 . C e-l
(164)
D 4 4

which is plotted in Fig. 69 based on the fits in Figs. 61-^64. For


examplé, for the PS data set in Fig. 1, the maximum value of
100 Hieber

Fig. 63 Extra pressure loss based on PP data from Ref. 11: a ; from
Refs. 23,168: > ; from Ref. 26: O; from Ref. 32: +; from Ref. 53:
□ ; from Ref. 69: o; from Ref. 171: < . Best-fit curve corresponds
to (162) with Cs = 2.87 x 10"^ (c m ^ / d y n e )^ *a n d Cg = 2.098.

is approximately 600 sec"^ x 3 x 1 0 ^ g/(cm*sec) - 2 x lO® dyne/cm^,


which, from the PS curve in Fig. 69, translates to a maximum L q /B
of about 6 . Hence, since L/D = 38 in Fig. 1, it follows that correct­
ing for Apg would lower the viscosity by at most 15%, and typically
Melt-Viscosity Characterization 101

F ig . 64 Extra pressure loss based on PS data from Ref. 11; a ;


from Ref. 22: v ; from Ref. 26: o; from Ref. 32: +; from Ref. 61:
□ ; from Ref, 69: O; from Ref. 171: < ; from Ref. 172: > . Best-
fit curve (omitting □ points) corresponds to (162) with Cg = 2 . 5 7 x
1 0 ”® (cm^/dyne)^*^°® and Cg = 2.108.

by much less at the lower stress levels. If this effect were signifi­
cant, however, the fits for Apg could be used to obtain a corrected
by solving

Ap = Cs + 4 ^ Ti
Lvy (165)

iteratively for given Ap and L /D.

A pp en dix e . M E L T -F R A C T U R E C O R R E LA TIO N

It is well known that as polymer melts flow through a die, the extru-
date becomes irregular above a certain flow rate [173], often with
102 Hleber

T^/T«

Fig. 65 Results for HDPE, as in Fig. 56, but normalized with respect
to T * . Points from Ref. 169 not plotted since * unknown. Best-fit
t

curve corresponds to (163) with Cy = 4.39 and Ce = 1.257.

associated pressure oscillations [174]. One possible criterion for the


onset of such irregular flow was put forth by Bartos [17], in which
the occurrence is associated with when nw/no - 0.025, where nw
notes the shear viscosity at the waU. More recently, Gleissle [39]
associated the occurrence with when (N i /t )^ - 4.6, where Ni denotes
the first normal-stress difference.
From our work we see from Fig. 52 in Appendix E that n/no =
fn(T/T*; n ) . In particular, for the range of n of interest (see Table
1) we see from Fig. 52 that the value of / * corresponding to n / n o =
t t

0.025 varies monotonically from about 2.5 at n = 0.2 to 7.5 at n = 0.35.


Hence, from our work and the correlation of Bartos [17], we would
expect the critical value of corresponding to the onset of melt
Melt-Viscosity Characterization 103

Fig. 6 6 Results for LDPE, as in Fig. 62, but normalized with respect
to T * . Points from Refs. 34, 169, and 170 not plotted since t * un­
known. Best-fit curve corresponds to (163) with Cy = 7.36 and C q =
1.307.

fracture to be about 5t * , with the factor being somewhat smaller at


the lower values of n and somewhat larger at the higher values of
n (such as for HDPE and LDPE). This tends to be borne out by Fig.
70, which plots various available results from the literature.
To relate to the criterion of Gleissle [39], we refer to Fig. 56 of
Appendix E and consider the data from [163]. Specifically, if the a
points at large t are extrapolated, they cross the dashed curve from
Gleissle [39] at t = Tcritical “ 8 x 1 0 ^ dyne/cm^. On the other hand,
fits of the viscosity data from [163] give n = 0.323 and t * = 1.24 x
10^ dyne/cm^, such that it follows from Fig. 52 that n/no = 0.025 at
t /t * - 6 . Hence, according to the criterion of Bartos [17], Tcritical
- 6 X 1.24 X 1 0 ^ dyne/cm^. That is, as applied to the data of
Gortemaker et al. [163], both the criterion of Bartos [17] and that
of Gleissle [39] predict that the onset of melt fracture would occur
when - 8 X lo^ dyne/cm
104 Hleber

Fig. 67 Results for PP, as in Fig. 63, but normalized with respect to
T * . Best-fit curve corresponds to (163) with C 7 = 2.89 and G q = 1.795.

A p p en d ix H . COMPARISON OF CROSS
AND C AR REAU MODELS

As an alternative to the modified Cross model, (1 ), one could represent


the shear-rate dependence by the equally popular Carreau [177] model,
which can be expressed as

no (166)
2 -,(l-n ) / 2
[1 + (tIoY/t^*) 1
Melt-Viscosity Characterization 105

Fig. 6 8 Results for PS, as in Fig. 64, but normalized with respect to
T*. Points from Refs. 32 and 172 not plotted since x* unknown. Best-
fit curve (omitting □ points) corresponds to (163) with C 7 = 6.17 and
Ca = 1.543. Upper dashed curve is best fit of results from Ref. 11:
A , from R ef. 22: v , from R ef. 26: o with C 7 = 4.60 and Ca = 2. 201
whereas lower dashed curve is best fit of Ref. 69: o , Ref. 173: <1
with C 7 = 1.31 and Ca = 2.146.

Although (1) and (166) have the same asymptotic limits for small and
large Y, the transition between these limits is much more rapid for
(166), where the behavior of ( tìoY/t^*)^» rather than ( tìoY/'^*) >
as for ( 1 ) , relative to 1 is of prime importance.
To illustrate these differences, we apply the two models to the vis­
cosity flow curves in Fig. 71, which are based on results from [62],
for PS-NMWD and PS-BMWD materials. In particular. Fig. 72 shows
the corresponding best fits for the NMWD case, with the solid curve
corresponding to ( 1 ) with

dyne
6
(n , T*, no) = ( 0 . 079, 1.32 X 10
cm^ ’
2.20 X 10^ cm_S_ (167)
*sec
106 Hieber

Fig. 69 Results for Lg/D based upon (164) and the fits in Figs.
61“ 64.

which fits the indicated points with an rms deviation of 5%, whereas
the dashed curve corresponds to (166) with

(n , T * , no) = (o .2 3 1 , 6 .8 4 x
^ \
10=
cm'^
1.99 x 10 = — ----
cm* sec/
\ (168)

and an rms deviation of 2%. Similarly, Fig. 73 shows analogous re ­


sults for the BMWD case, with the solid curve based on (1) and

(n , T * , no) = (o .2 8 7 , 2 .3 9 x io= 3 .1 8 x 10= — S— \ (1 6 9 )


^ \ cm cm* sec/

and an rms deviation of 2 %, whereas the dashed curve is based on


(166) with

(n , T * , no) = ( 0. 371 , 1.43 X 10= 2 .3 9 x 10= — S— \ ( 170 )


\ cm"^ cm-sec/

and an rms deviation of 8 %.


Melt-Viscosity Characterization 107

Fig. 70 Results for critical wall shear stress corresponding to onset


of melt fracture or oscillatory pressure traces for HDPE from Refs.
16, 17, 19, 21, 34: O; l DPE from Refs. 16, 17, 19, 21, 39, 174, 175,
176: A; n y l o n from Ref. 175: ▼; PMMA from Refs. 17, 175: ■ ;
PMMA/PS from Ref. 17: a ; PP from Refs. 17, 21: a ; PS from Refs.
17, 21, 22, 6 6 , 174: o. For points from Ref. 175, in which no flow
curves are given, the values of x* have been taken from correspond­
ing master fits in Table 1 of Section II.B . The curves correspond to
•■w/ * = 5 and 20.
t

As might be expected, the Carreau model does better in the NMWD


case, which is characterized by a more rapid transition between the
Newtonian and power-law asymptotes, whereas the Cross model does
better in the BMWD case. Since the BMWD case is more representa­
tive of commercial grade polymers, it follows that the Cross model is
more suitable for present purposes.

A p p en d ix 1. V IS C O U S -H E A T IN G EFFECTS
IN C A P IL L A R Y RHEOMETER

The effects of viscous heating upon viscosity measurements, p ar­


ticularly in capillary rheometers at high shear rates, have received
108 Hieber

Fig. 71 Viscosity flow curves at 180®C for two PS melts of NMWD and
BMWD, based upon Fig. 2 of Graessley et al. [62].

considerable attention in the literature (e . g . , [178--191]). This sec­


tion uses the theoretical perturbation expansion from [187] together
with representative viscosity fits for various generic melts to gener­
ate curves which can be used to determine when viscous heating is
important.
Specifically, Fig. 74 shows plots for nylon, PC, PMMA, PP, PS-
BMWD, and PVC-RIGD, based on representative property values
listed in Table 4. The curves in these figures correspond to the
flow conditions under which viscous heating accounts for a 1 0 %re ­
duction in the overall pressure drop. For a given L/D we can at­
tain a higher shear rate before viscous heating becomes important
by using a smaller diameter or a higher melt temprature. On the
other hand, when expressed in terms of noYa» we can attain higher
values by using a smaller-diameter tube or a lower melt temperature.
This is illustrated by Fig. 75 for the PC case. In turn, the results
in Fig. 75 strongly suggest that viscous heating is at least partly
responsible for the scatter exhibited at the higher values of rioY
the PC master plot in Fig. 9 of Section I I . B .
The symbols in Fig. 74 are corresponding predictions based on
the finite-difference formulation described in Appendix J, and they
Melt-Viscosity Characterization 109

Ü
0)
*eo
E
Ü
E
cn

y (sec’’ )

Fig. 72 Viscosity data points from NMWD curve in Fig. 71 with solid
and dashed curves being best fits based upon Cross and Carreau
models, respectively, as given by (1, 167) and (166, 168).

serve as a check on the perturbation theory [187] and its extension


to a more general viscosity model. That is, whereas the perturbation
expansion in [187] is for a power-law fluid, it has now been applied
to the more general viscosity model, ( 1 ) , by introducing a local
power-law index,

1 -n
1 + nK noYa (171)
1 -n
1+ K
110 Hieber

o
0)
*(0
g
o
£
O)

y (sec ’ )

Fig. 73 Same as Fig. 72 but for BMWD case with best fits given by
(1, 169) and (166, 170).

which corresponds to

In r\\
n E 1 + (172)
3 In Y ^

evaluated at y = Ya* The results in Fig. 74 do not include the pres­


sure dependence, such that no in (171) has been evaluated at T q (uni­
form inlet melt temperature) and atmospheric pressure. In addition,
the temperature-sensitivity factor used in [187] actually corresponds
Melt-Viscosity Characterization 111

to a ’ , in the notation of Section II.C ; in turn, it follows from (31)


that actuals nao when evaluated at Ya* Accordingly, the pertur­
bation expansion in [187] indicates that the leading effect of viscous
heating is to lower the overall pressure drop by the factor

1 - ef(Ç ^; n) (173)

where, based on the preceding extension.

n
a
£ = haoAT, AT = (174)

and

11 a - (175)

based on the particular model used in Table 4. That is, e represents


the product of the temperature sensitivity of the viscosity and a char­
acteristic temperature rise due to viscous heating. Oh the other hand,
the function f(? L ; n) has been determined [187] for n = . 1, .2, . . . , 1 . 0
by a series solution in terms of

Ua _ Nh
—, a = ^-77- (176)
^ ¡k - '■'= ^ a pC

where, in heat transfer terminology, 5l inverse Graetz number


and Pe is the Peclet number. Hence, the curves in Fig. 74 correspond
to where ef = 1 /1 0 .

A p p en d ix J . F IT T IN G C A P IL L A R Y -R H E O M E T E R
D A T A A C C O U N T IN G FOR PRES­
SURE DEPENDENCE AND
V IS C O U S H EA T IN G

To account for the possible effects of viscous heating or pressure de­


pendence on capillary-rheometer measurements, we developed a com­
puter program which combines the simplex method of Nelder and Mead
[77] with a finite-difference modeling of the nonisothermal steady flow
of a generalized viscous fluid in a circular tube. That is, for each
data point the program simulates the pressure drop and, on the basis
112 Hleber

^’ 10% viscous-heating” contours for various generic materials;


F ig . 74
symbols denote finite-difference results for ( T q, D) = (higher T q,
0.025 cm) : > and (lower T q, 0. 200 cm) : < .

of the comparison between the predicted and experimental Ap , uses


the simplex method to locate the best set of parameters in the viscos­
ity model such that the rms deviation for the total set of data points
is a minimum. In addition, the program allows one to fix certain pa­
rameters in the viscosity model when appropriate.
Melt-Viscosity Characterization 113

Based on Section II, the program presently considers three vis­


cosity models, all based on ( 1 ) but with various representations of
rio(T, p ), namely the five-constant (n , t *, B, Tb, 3) model I given
by (1,34), the seven-constant (n, t *, D x , D 2 , D 3 , Ax, A 2 ) model II
given by (1, 36, 37), and the seven-constant (n , x*, Dx, D 2 , D 3 , Ax,
A 2 ) model III given by (1, 36, 37, 40). The finite-difference repre­
sentation of the energy equation follows along the lines of Section
III.A , (62), but for the steady-axisymmetric case, namely
H ieb er

nu

AX.
2T. . . ^'^i+l.i+1 - 1+
f J ^i+i.i-1 ^ ^ ■" Ä y lÄ y T T ^ n
k.

k
1_— 4rr^ ■' “ÂÿTTÂÿT
Fr\Ay5-i^^yri^3^ ’ ’
(177)
AV j-i'^i+l,i'''l\ + $. .
Melt-Viscosity Characterization 115

except along the centerline, j=l, where the equation reduces to

4k (178)
p Ax.

and at the wall, where

T = T (179)
i+l,N 2+l w

where Tw is taken to be T q, the uniform melt temperature at inlet. In


particular, one can solve this system directly for T i + i j (1 < ] < N2+1)
by standard tridiagonal solvers, with the corresponding flow quantités
116 Hieber

at the new streamwise location being updated by iteration for the given
ri(Yj T, p ) , making use of appropriate relations for tube flow such as
given in Table 3 of Section III.A . Hence, for a given set of flow con­
ditions, the calculation starts with a uniform inlet melt temperature
and corresponding fully-developed velocity profile at x = xi=i = 0 and
successively marches forward to x = = L, giving the overall
pressure drop. In particular, the following variable streamwise and
radial mesh gives sufficiently accurate results:

L
= 0, (1 0 )V ® x . (i= 2 ,3 ........ 9) (180)
^^ = 1 0 ’
Melt-Viscosity Characterization 117

and

-¿ = 0, 0.2, 0.4, 0.55, 0.70, 0.80, 0.90, 0.95, 1.00 (181)


a

where a denotes the tube radius.


Figure 76 shows resulting fits for the PS and PP data generated
by Westover [105] with the pressurized double-piston, capillary-
rheometer device. The corresponding curves for PS and PP are
given, respectively, by
118 Hieber

Table 4 Representative Property Values Used in Viscous-Heating


Calculations Without Pressure Dependence^

(n , 1 *, B, T b ) or
Material (n , x*, D i, D 2 , A^, A 2 ) ( P > C p , k-tb)

NYLON (0.27, 2.2 X 10®, 2. X 1 0 (0.99, 4.4 X l0 ^ 2.5 X 1 0 “)


9 X 1 0 ®)
PC (0.17, 6.4 X 1 0 ®, 4. X l 0 ®•^ ( 1 .0 6 , 1 . 9 X l0 ^ 2 . 4 X 1 0 “)
144, 25.9, 61.2)
PMMA (0.20, 1.0 X 1 0 ®, 9. X 1 0 “ , (1.04, 2.3 X l0 ^ 2.1 X lO")
104, 38.2, 101.3)
PP (0.31, 1.6 X 10®, 7. X lO’ S (0.77, 3.1 X l 0 ^ 1.5 X 1 0 “)
5.5 X 1 0 ®)

PS-BMWD (0.25, 2.7 x 1 0 ®, 3. x 10^®, (0.94, 2.1 X 10^ 1 . 5 X 10“*)


1 0 0 , 27.2, 60.0)

PVC-RIGD (0.21, 1.7 X 10®, 2. x lo'®-“, (1.32, 1.8 X 1 0 ’ , 1.8 X 1 0 “*)


20. X 10®)

^Values given in following units: x* (dyne/cm®), B [g / (c m -se c )], Tb


(K ), Di [g/(cm -sec)], D 2 (° C ) , A 2 (° C ) , p (g/cm®), Cp (e r g / g .°C ),
kth [e rg / (se c 'c m -°C )].

( n , X * , B , Tb, 3) = (o.256, 2.24 x 1 0 ®


cm

5.25 X 10'® ---- S— , 12,600 K, 2.75 x 10'® ^ 2 ^


cm •sec dyne /
1 (182)

and

dyne
( n , T * , B, Tb, 6 ) ^0.253, 2.06 X 1 0 ®
cm

.613 ---- i — , 6230 K, 1.63 x lo'^ (183)


cm
— •sec""" dyne /

If the calculations are repeated with viscous heating omitted and


the constants in (182) and (183) held fixed, the resulting curves in
Fig. 76 shift up by no more than 7% and 2% for the PS and PP, re ­
spectively, indicating that viscous-heating effects are not significant
Melt-Viscosity Characterization 119

for these data sets. Further, if representative wall-shear-stress lev­


els are calculated based on (7) of Section II.A , it follows that var­
ies between 4 x lo^ and 2 x lo^ dyne/cm ^ for PS and between 1 x lo^
and 1 . 3 X lo® dyne/cm^ for PP. Compared with the resulting values
for T*, these shear-stress levels indicate that these data sets do not
lie entirely in the power-law region, so the resulting values for *t

and rio should be significant. In particular, the values for x* in


(182) and (183) do agree reasonably well with the values from the
corresponding master plots in Table 1 of Section II.B , as do the val­
ues for n. Further, the values for 3 (= 3o) are also reasonable, as
120 Hieber

y (sec )
10 10 ^ 10 "

O
4
C
>
>%
T3

Q.
<

Fig. 76 End-corrected data from pressurized capillary rheometer of


Westover [105] for PS and PP with T q = 250°C, D = 0.228 cm, L/D =
2 0 , and reservoir pressure (in units of 6 . 9 x lO^ dyne/cm^ = 1 0 ^ psi)
= 2 (A ), 5 ( o ) , 10 ( □ ) , 20 ( 0 ) , and 25 ( v ) . Curves are correspond­
ing best fits based upon model I as given by (182) for PS and (183)
for PP with rms deviations of 9% and 3%, respectively.

plotted in Fig. 11 of Section II.C for PS and tabulated in Table 2 for


PP.
Figure 77 shows results from fitting data of Nakajima and Collins
^ [74] for an unplasticized PVC. The resulting best fit based on model
I is

dyne
(n, T * , B, Tb, 3) |o.300, 5. 08 X 10®
cm^ ’

cm"
6.20 X 10-1 0 __ g , 15,700 K, 4 . 6 X 10” (184)
cm •sec dyne /

For comparison, if we repeat the simulation with the same model con­
stants as in (184) but with viscous heating omitted from (178), we get
the dashed results in Fig. 77. Accordingly, the viscous-heating effect
Melt-Viscosity Characterization 121

Fig. 77 End-corrected cap illary-rheometer data of Nakajima and Col­


lins [74] for unplasticized PVC with T q = 190°C, D = 0.127 cm, and
L/D = 5.1 ( A ) , 10.1 ( o ) , 19.8 ( □ ) , and 39.4 ( o ). Solid curves cor­
respond to best fit based upon model I as given by (184), with rms
deviation of 4%. Dashed curves are corresponding predictions when
viscous heating is omitted, but constants are kept the same as in
(184).

lowers Ap by as much as 25% for the largest L/D and highest Ya* Cor­
responding fits of data from [74] at 170°, 180°, and 200°C give val­
ues for 3 of 5.4 X 10“^, 4 . 7 x 10"^, and 5 . 0 x 1 0 "^ cm^/dyne, respec­
tively, with corresponding rms deviations of 10%, 2%, and 6 %. For
comparison, Nakajima and Collins [74] deduced a representative value
from their measurements of 3 . 8 x 10“^ cm^/dyne. This somewhat small­
er value is compatible with the fact that viscous-heating effects were
not included in the data reduction in [74].
For some cases the data is essentially entirely in the power-law re ­
gion, and viscous-heating effects are negligible, such that the fore­
going finite-difference simulation can be replaced by closed-form re ­
sults. In particular, if we consider the power-law limit of model I,
we have

n- 1 bp
n(Y, T, p) = m(T, p)Y m(T, p) = m(T)e (185)
122 Hieber

where

m(T) = ^ exp(^-Y^j, b=n (186)

Under isothermal conditions, T = T j such that fh (T ) is a constant, m,,,


and it follows that the axial force balance is amenable to separation of
variables, resulting in

-bp dp
C = - (187)
dx H M t i]

By successive radial integration of the right half of (187) , we obtain

^ ^2 m
0 n T3n+1
| 3n+l ». 1 ^
(188)
a L 4n ^aj

where Ya = 4U/a = 4 Q/ 7ra^. On the other hand, axial integration of


the left half of (187), with p = po at x = 0 and p = pe at x = L, gives

exp(-bP g) - exp(-bp „)
) (189)

Hence, combining (188) and (189), we have

L r Sn+l .
4m, e xp(-bp ^) - exp(-bP|j)j. (190)
D L 4n

In particular, if the pres sure-sensitivity effect is small, such that


both bpe and bpo « 1, then the right-hand side of (190) reduces to
Po " Pe = ^P> recovers the usual result for fully-developed
tube flow of a power-law fluid.
Therefore, for a given data set at a single temperature, (190) can
be combined with the simplex method to determine the best values of
(n, mo> b ) . On the other hand, if the data set corresponds to more
than one temperature (but each run is still at a uniform temperature) ,
then results for (n . A, Ta, b ) can be obtained where A = B ^ ( t* ) 1'^
and Ta = nT^ from (186). In either case the parameter b corresponds
to 3^ in the notation of Section II.C , such that, from (33), 3o = b/n.
For example, the foregoing procedure was applied to the nonlinear
Bagley plot for PMMA at 180°C from Casale et al. [51]. The residts are

(n , nlo, b ) = ^0.250, 1.38 x lo® (sec)*^, 0.86 x 10"® ^ i )


Melt-Viscosity Characterization 123

Q.
<

Fig. 78 End-corrected Bagley plot of PMMA data from Ref. 51 at


Ya = 1 ( ^ ) » 5 ( o ) , 10 ( □ ) , 20 ( 0 ) , 30 (+ ), and 40 ( v ) s e c '^ T q =
180°C; 0.0508 cm < D < 0.1524 cm. Solid curves correspond to best
fit based on model I, as given by (192), whereas dashed curves cor­
respond to best fit based on (185) with T e T q, as given by (191).

which corresponds to the dashed curves in Fig. 78 and fits the data
with an rms deviation of 6 %. Figure 78 also shows results based on
the best fit generated by the finite-difference—simplex program using
model I , namely
124 H ieber

(n , T * , B , T, e) = (o.246, 1.63 x 10=


D ^ cm

1.42 X 1 0
■11 __ £_ -, 20,600 K, 4.36 x lO*^ cm (192)
cm* sec' dyne /

which corresponds to the solid curves and fits the data with an rms
deviation of 3%. In particular, from (186) it follows that the power-
law limit of (192) corresponds to

(n , nio, b ) = |o.246, 1.32 x 1 0 = (sec )”^, 1.07 x 1 0 '= (193)

which agrees reasonably well with (191).


The procedure developed by Penwell et al. [192] for deducing val­
ues for b is based upon a variant of (190). Specifically, if one con­
siders two different flow rates and corresponding pressure drops for
the same capillary tube and temperature, it follows from (190) that
/ * A n
- 1l~exp (-bA p i)
(194)
n l-e x p (-b A p 2 )

noting that bpe « 1 if the exit pressure is atmospheric. For example,


if we consider the data points in Fig. 78 and assume n = 0.25, based
on (191) and (193), then it follows that (194) can be used to gener­
ate a value for b from each pair of data points at a given L/D. This
has been done with resulting values for b plotted in Fig. 79. Despite
evident scatter, particularly at the smaller L/D where the pressure
differences are rather small, the results do center about a value of
b 1.0 X 1 0 "^ cm^/dyne, which is compatible with the results in (191)
and (193). On the other hand, with n = 1/4, this then corresponds
to a value for 3 o of about 4 . 0 x lO"^ cm^/dyne.

A p p en d ix K. "F O U N T A IN " EFFECT

The "fountain" region, a term coined by Rose [193], refers to the re ­


gion in the vicinity of an advancing free surface, such as in a capil­
lary tube. Within the context of dynamic contact angles, this phe­
nomenon has received considerable attention in the literature, as sum­
marized by Dussan [194] and as exemplified by more recent works
such as [195—198]. As applied to injection molding, the earliest anal­
ysis of this phenomenon seems to have been that of Tadmor [199].
Melt-Viscosity Characterization 125

Fig. 79 Results for b based upon (194) with n = 0.25 as applied to


data points in Fig. 78; additional points lie off the graph for L/D =
12. 2 .

Subsequently, the relevance of this region to the developing tempera­


ture field during the filling stage has been addressed from various
perspectives in [119,200—203]. An interesting experimental study re ­
lating to the kinematics in the fountain region is found in [204].
Schematically, the injection-molding situation is as shown in Fig.
80 for a strip cavity. In particular, the advancing melt front is
curved, extending over a distance of 0 (b ) in the streamwise direc­
tion. Further, as noted in Section III.A , since the Peclet number,
Pe E Ub/a ( a e kth/pCp), is large, the thermal diffusion from the
cold walls is restricted to thin thermal boundary layers which sur­
round an essentially isothermal core region in which T T q, the in­
jection temperature. Although the direct effect of the fountain region
upon the overall pressure drop will be small if b « L (the underlying
thin-cavity assumption) , its contribution to the developing temperature
126 Hieber

Fig. 80 Schematic of advancing front in strip cavity together with


cold layer (dashed) adjacent to wall and associated effect on veloc­
ity profile.

field, and possible orientation in the surface of the final solidified


part [199], is more crucial. Specifically, the cold layer that de­
velops behind the advancing melt front arises from material that has
been transported from the hot core region to the wall layer by pass­
ing through the fountain region.
As indicated in Fig. 80, and studied more extensively in [119,200—
203], the temperature field is characterized by two thermal boundary
layers, namely a steady-state layer in the vicinity of the entrance, of
order (axb/U)^/^ in thickness (^’Leveque” type [205]), and a transi­
ent layer developing from behind the advancing front, with thickness
of order (at)^^^, where t = (xm f(t)-x )/ U = t - x/U in the constant U
case. In turn, there will be an intermediate overlap region where
these two regions merge, characterized by where both boundary lay­
ers are of the same thickness. Specifically, at tfip, when x ^ f = L,
it follows then that the overlap region will occur where x/L = 0 ( 5 ^^^^) ,
in which Çl = L/t>Pe is typically a small quantity, with the thickness
being (a L ) 1 / 2 That is, the thermal boundary layer at tfUl will be
thickest in a region near the entrance, with the transient structure
extending over most of the length of the cavity. Such a streamwise
variation in the boundary-layer thickness is exhibited by the formu­
lation in Section III, as shown by Figs. 15 and 17, For a more de­
tailed treatment of this thermal structure, including the effects of a
temperature-dependent viscosity but still without the effects of a
curved melt front, see [119,200—203].
For the flow field in the vicinity of the curved melt front, the sim­
plest case is evidently that of a Newtonian fluid under isothermal con­
ditions. In this regard, we obtained a solution some years ago [206] ,
Melt-Viscosity Characterization 127

F ig . 81 Predicted free-interface shape at r = 0 , 1, and 3 together


with corresponding finite-element configuration; from Ref. 206. Co­
ordinates are shown normalized with respect to half-gap thickness.

via a finite-element formlation for creeping Newtonian flow in the two-


dimensional case. Specifically, we used quadratic (linear) shape func­
tions for the velocity components (pressure) together with triangular
elements, the latter having one curved side (cubic) along the melt
front. In addition, we used special elements at the attachment point
and used the local singular behavior. Figure 81 shows the predic­
tions for the curved melt front and underlying finite-element grid for
various values of the nondimensional parameter T e a/rioU> where a
denotes the surface tension. In particular, an approach to a more
circular interface with increasing T is noted, as would be expected.
Corresponding results for the velocity field are indicated in Fig. 82
for r = 1. The velocities are shown relative to the front, such that
the upstream velocity at the left corresponds to a superposition of
Poiseuille flow to the right and uniform flow to the left.
The preceding solution has a basic flaw : it cannot describe the
kinematics of the fluid lying in front of the interface. In particular,
although the present formulation neglects the viscosity of the latter
fluid (a very good approximation in the present context, since the
dynamic viscosity of air is smaller than that of polymer melts by a
factor of 1 0 “^ or 1 0 “®) , such that the shear stress vanishes along
the interface in the present solution (a natural boundary condition,
128 Hieber

Fig. 82 Calculated velocities (scaled relative to U) corresponding to


r = 1 in F ig . 81; in frame fixed to front.

in finite-element formulation), we still require the tangential velocity


to be continuous across the interface. However, if we were to impose
the interfacial tangential velocity shown in Fig. 82 upon the fluid in
front of the interface, the material would move toward the attachment
point with nowhere to go from there. Indeed, it is such a kinematic
dilemma which has led to some controversy , including the introduction
of slip at the wall, in treating dynamic contact angles in general (cf.
[194-198]).
Even if this limitation were overlooked and the formulation were ex­
tended to the shear-thinning case, the remaining problem would be to
incorporate this local fountain-flow solution within the framework of an
overall, unsteady, nonisothermal situation with temperature-dependent
viscosity. Work in this direction awaits further development.

ACKNOW LEDGMENTS

This work has been supported by the Cornell Injection Molding Pro­
gram (K. K. Wang, Director), which is financed by the National Sci­
ence Foundation (Grant MEA-8200743) and an industrial consortium
consisting of approximately 20 companies. Special thanks are due to
other co-workers over the years, particularly A. I. Isayev, S. F.
Shen, V. W. Wang, J. F. Stevenson, and Simon Chung.
Melt-Viscosity Characterization 129

REFERENCES

1. G. D. Gilmore and R. S. Spencer, Modern Plastics, 27(8): 143


(1950) .
2 . R. S. Spencer and G. D. Gilmore, J. Colloid Sci,, 6: 118
(1951) .
3. R. L. Ballman, L. Shusman, and H. L. Toor, Modern Plastics,
37(1): 105 (1959).
4. H. L. Toor, R. L. Ballman and L. Cooper, Modern Plastics,
38(1): 117 (1960).
5. J. R. A. Pearson, Mechanical Principles of Polymer Process­
ing, Pergamon Press, Oxford, 1966.
6. M. R. Kamal and S. Kenig, Polym. Eng, Sci,, 12: 294, 302
(1972) .
7. J. L. Berger and C. G. Gogos, Polym, Eng, Sci, > 13: 102
(1973) .
8. P. C. Wu, C. F. Huang and C. G. Gogos, Polym, Eng, Sci.,
14: 223 (1974).
9. C. Austin, in Computer Aided Engineering for Injection Mold­
ing, Chapter 9, E. C. Bernhardt (e d .), Hanser, Munich, 1983.
10 . R. F. Westover, in Processing of Thermoplastic Materials, Sec­
tion III, E. C. Bernhardt (e d .). Van Nostrand Reinhold, Prince­
ton, N .J ., 1959.
11. Rheologie, Verband Deutscher Maschinen und Anlagenbau (e d .),
Carl Hanser Verlag, Munich, 1982.
12 . R. L. Bergen and H. L. Morris, Proc. Fifth Intern. Congr.
Rheol., 4: 433 (1970).
13. H. W. Cox and C. W. Macosko, A .I.C h .E .J ., 20: 785 (1974).
14. M. Kasajima, A. Suganuma, D. Kunii, and K. Ito, Proc. Int.
Conf. Polym. Processing, N. P. Suh and N. H. Sung (e d s .),
M .I.T . , 1977, p. 473.
15. H. Schott and W. S. Kaghan, J. Appi. Polym. Sei., 5: 175
(1961).
16. T. Arai and H. Aoyama, Trans. Soc. Rheol., 7: 333 (1963).
17. O. Bartos, J. Appi. P h ys., 35: 2767 (1964).
18. R. A. Mendelson, Trans. Soc. Rheol., 9: 53 (1965).
19. T. Kataoka and S. Ueda, J. Appi. Polym. Sei., 12: 939 (1968).
20. R. A. Mendelson, W. A. Bowles, and F. L. Finger, J. Polym.
Sei. A-2, 8: 127 (1970).
21. T. F. Ballenger, I.-J . Chen, J. W. Crowder, G. E. Hagler,
D. C. Bogue, and J. L. White, Trans. Soc. Rheol,, 15: 195
(1971).
22 . R. Ramsteiner, Kunststoffe, 61: 943 (1971).
23. C. D. Han, K. U. Kim, N. Siskovic, and C. R. Huang, J. Appi.
Polym. Sei. , 17: 95 (1973).
130 Hieber

24. L. A. Utracki, Z. Bakerdjian, and M. R. Kamal, J. AppL


Polym. Sei., 19: 481 (1975).
25. J. L. White and J. F. Roman, J. AppL Polym. Sei., 20: 1005
(1976).
26. H. Münstedt, Kunststoffe, 68: 92 (1978).
27. M. S. Jacovic, D. Pollock, and R. S. Porter, J. Appl. Polym.
Sei., 22: 517 (1979).
28. N. Alle and J. Lyngaae-Jorgensen, Rheol. Aeta, 19: 94 (1980).
29. E. Boudreaux and J. A . Cuculo, J. Appl. Polym. Sei., 27: 301
(1982).
30. F. P. La Mantia, A. Valenza, and D. Acierno, Rheol. Aeta, 22:
299(1983).
31. A . V. Shenoy, S. Chattopadhyay, and V. M. Nadkarni, Rheol.
Aeta, 22: 90, 209 (1983).
32. T. Nishimura, Rheol. Aeta, 23: 617 (1984).
33. L. A. Utracki and M. M. Dumoulin, Polym. -Plast. Teehnol.
Eng., 23: 193 (1984).
34. L. A. Utracki and R. Gendron, J. Rheol., 28: 601 (1984).
35. I. J. Duvdevani and I. Klein, SPE J. , 23(12): 41 (1967).
36. J. Miltz and A. Ram, Polym. Eng. Sei., 13: 273 (1973).
37. G. Ehrmann, G. Robins and M. H. Wagner, Kunststoffe, 64: 463
(1974).
38. J. L. White and H. B. Dee, Polym. Eng. Sei., 14: 212 (1974).
39. W. Gleissle, RheoL Acia, 21: 484 (1982).
40. H. M. Laun, Rheol. Aeta, 22: 171 (1983).
41. E. L. Kalinchev and M. B . Sakovtseva, Int. Polym. Sei. Teeh.,
4(8): T 167 (1911)
42. V. G. Bankar, J. E. Spruiell, and J. L. White, J. Appl. Polym.
Sei. , 21: 2125 (1977).
43. H. M. Laun, Rheol. Aeta, 18: 478 (1979).
44. Z. Tadmor and C. G. Gogos, Prineiples of Polymer Proeessing,
Wiley, New York, 1979.
45. H. W. Cox, C. C. Mentzer, and R. C. Custer, Polym. Eng. Sei.,
24: 501 (1984).
46. M. Yamada and R. S. Porter, J. Appl. Polym. Sei., 18: 1711
(1974).
47. H. A. Lord, Polym. Eng. Sei., 19: 469 (1979).
48. B. A. Knutsson and J. L. White, SPE Teeh. Papers, 30: 786
(1984).
49. J. M. Lupton, Chem. Eng. Progr. Symp. Series No. 49, 60: 11
(1964).
50. H. D. Herrmann and W. Knappe, Rheol. Aeta, 8: 384 (1969).
51. A. Casale, R. C. Penwell, and R. S. Porter, Rheol. Aeta, 10:
412(1971).
Melt-Viscosity Characterization 131

52. A. K. Van der Vegt, Trans.J. Plastics Inst., 32; 165 (1964).
53. G. V. Vinogradov and N. V. Prozorovskaya, Rheol. Acta, 2: 156
(1964).
54. D. C. Huang and J. L. White, Polym. Eng. Sei., 19: 609 (1979).
55. W. Minoshima, J. L. White, and J. E. Spruiell, Polym. Eng. Sei.,
20: 1166 (1980).
56. R. S. Spencer and R. E. Dillon, J. Colloid Sei., 4: 241 (1949).
57. J. F. Rudd, J. Polym. Sei., 44: 459 (1960); J. Polym. Sei., 60:
S7 (1962).
58. R. L. Ballman and R. H. M. Simon, J. Polym. Sei. A, 2: 3557
(1964).
59. R. A . Stratton, J. Colloid Interface Sei., 22: 517 (1966).
60. K. S. Hyun and H. J. Karem, Trans. Soc. Rheol., 13: 335 (1969).
61. D. P. Thomas and R. S. Hagan, Polym. Eng. Sei., 9: 164 (1969).
62. W. W. Graessley, S. D. Glasscock, and R. L .. Crawley, Trans.
Soc. Rheol., 14: 519 (1970).
63. Y. Kuo and M. R. Kamal, A .I.C h .E .J ., 22: 661 (1976).
64. J. L. S. Wales, The Application of Flow Birefringence to Rheo­
logical Studies of Polymer Melts, Delit University Press, Delft,
1976.
65. G. H. Pearson and L. J. Garfield, Polym. Eng. Sei., 18: 583
(1978).
6 6 . H. M. Laun, M. H. Wagner, and H. Janeschitz-Kriegl, Rheol.
Acta, 18: 615 (1979).
67. R. Racin and D. C. Bogue, J. Rheol., 23: 263 (1979).
6 8 . M. D. Dalai and R. C. Armstrong, Polym. Eng. Sei., 22: 684
(1982).
69. A. I. Isayev and B . Chung, Polym. Eng. Sei., 25: 264 (1985).
70. E. A. Collins and C. A. Krier, Trans. Soc. Rheol., 11: 225
(1967).
71. E. A . Collins and A. P. Metzger, Polym. Eng. Sei., 10: 57
(1970).
72. L. A. Utracki, J. Polym. Sei. (P h y s .), 12: 563 (1974).
73. E. A. Collins, C. A . Daniels, and C. E. Wilkes, in Polymer Hand­
book, 2nd e d.. Chapter V, J. Brandrup and E. H. Immergut
(e d s .), Wiley, New York, 1975.
74. N. Nakajima and E. A. Collins, J. Appl. Polym. Sei., 22: 2435
(1978).
75. R. A. Mendelson, Polym. Eng. Sei., 16: 690 (1976).
76. H. Münstedt, Polym. Eng. Sei., 21: 259 (1981).
77. J. A. Neider and R. Mead, Computer J., 5: 308 (1965).
78. M. M. Cross, J. Colloid S et, 20: 417 (1965).
79. M. M. Cross, Europ. Polym. J., 2: 299 (1966).
80. M. M. Cross, J. Appl. Polym. Sei., 13: 765 (1969).
132 Hieber

81. M. M. Cross, Rheol. Acta, 18: 609 (1979).


82. B. Chung, private communication, Chem. Eng. Dept., Cornell
University, 1983.
83. B. Rabinowitsch, Z. Phys, Chem,, A145: 1 (1929).
84. G. V. Vinogradov, A . Y. Malkin, N. V. Prozorovskaya, and
V. A. Kargin, Doklady Akad, Nauk SSSR, 154: 890 (1964).
85. G. V. Vinogradov and A. Y. Malkin, J. Polym, Sci, A-2, 4:
135 (1966).
8 6 . G. V. Vinogradov and A. Y. Malkin, Rheology of Polymers,
Mir, Moscow, 1980.
87. V. Semjonow, Adv. Polym, Sci,, 5: 387 (1968).
8 8. E. L. Kalinchev, M. B. Sakovtseva, M. G. Evdokimova, and
A. Y. Dogadushkina, Int, Polym, Sci, Tech,, 4(10): T/lOO
(1977).
89. D. Robinson, M.S. Thesis, University of Rochester, 1967.
90. S. Middleman, The Flow of High Polymers, Interscience,
New York, 1968.
91. G . D . Panova, L . I . Myasnikova, D . N . Emelyanov, and
A. V. Ryabov, Vyokomol, Soedin, , 18B: 273 (1976).
92. K. H. Hellwege, W. Knappe, F. Paul, and V. Semjonow,
Rheol, Acta, 6: 165 (1967).
93. V. Semjonow, RheoL Acta, 3: 98 (1963).
94. M. L. Williams, R. F. Landel, and J. D. Ferry, J, Am, Chem,
Soc,, 77: 3701 (1955).
95. R. C. Penwell and R. S. Porter, J, Polym, Sci, A-2, 9: 463
(1971) .
96. P. H. Goldblatt and R. S. Porter, J, Appl, Polym, Sci,, 20:
1199 (1976).
97. D. W. Van Krevelen, Properties of Polymers, 2nd e d ., Else­
vier, Amsterdam, 1976.
98. J. D. Ferry and R. A. Stratton, Kolloid Z . , 171: 107 (1960).
99. J. M. O'Reilly, J, Polym, Sci,, 57: 429 (1962).
lOO* F. N. Cogswell and J. C. McGowan, Brit, Polym, J,, 4:183
(1972) .
101. V. Semjonow, Rheol, Acta, 2: 138 (1962).
102. L. Christmann and W. Knappe, Colloid Polym, Sci,, 252: 705
(1974).
103. F. Ramsteiner, Rheol, Acta, 9: 374 (1970).
104. B, Maxwell and A. Jung, Modern Plastics, 35(3): 174 (1957).
105. R. F. Westover, SPE Trans,, 1: 14 (1961).
106. K. Ito, Rept, Progr, Polym, Phys, Japan, 12: 131 (1969).
107. R. L. Ballman,Nature, 202: 288 (1964).
108. S. Y, Choi and N. Nakajima, Fifth Int, Congr, Rheol,, 4: 287
(1970).
109. S. Matsuoka and B . Maxwell, J, Polym, Sci,, 32: 131 (1958).
Melt-Viscosity Characterization 133

110. K. H. Kellwege, W. Knappe, and P. Lehmann, Kolloid Z . , 183:


110(1962).
111. Y. Ishida and S. Matsuoka, Am. Chem. Soc, Polym, Preprints,
6: 795 (1965).
112. N. Shishkin, Soviet Phys. Solid State, 2: 322 (1960).
113. A. Zosel, Kolloid Z., 199: 113 (1964).
114. G. Menges, P. Thienel, and G. Wubken, Kunststoffe, 66: 42
(1976).
115. G. Menges, J. Wortberg, and W. Michaeli, Kunststoffe, 68: 47
(1978).
116. G. Williams and H. A. Lord, Polym. Eng. Sei., 15: 553, 569
(1975).
117. J. F. Stevenson, A. Galskoy, K. K. Wang, I. Chen, and D. H.
Reber, Polym. Eng. Sei., 17: 706 (1977).
118. S. M. Richardson, H. J. Pearson, and J. R. A . Pearson, Plas­
ties and Rubber: Proeessing, 5: 55(1980).
119. H. Van Wijngaarden, J. F. Dijksman, and P. Wesseling, J.
Non-Newt. Fluid Meeh., 11: 175 (1982).
120. Y. Kuo, SPE Teeh. Papers, 24: 135 (1978).
121. J. F. Stevenson and W. Chuck, Polym. Eng. Sei., 19: 849 (1979),
122. C. A. Hieber, SPE Teeh. Papers, 28: 356 (1982).
123. A. I. Isayev, K. D. Vachagin, and A. M. Naberezhnov, J. Eng.
Phys., 27: 998 (1974).
124. C. Miller, Ind. Eng. Chem. Fundam., 11: 524 (1972)
125. C. A. Hieber, R. K. Upadhyay, and A . I. Isayev, SPE Teeh.
Papers, 29: 698 (1983).
126. T. H. Kwon, D. Chu, and K. K. Wang, SPE Teeh. Papers, 31:
809(1985).
127. F. N. Cogswell, Polym. Eng. Sei., 12: 64 (1972).
128. C. A . Hieber and S. F. Shen, J. Non-Newt. Fluid Meeh., 7: 1
(1980).
129. E. Broyer, C. Gutfinger, and Z. Tadmor, Trans. Soe. Rheol. ,
19: 423 (1975).
130. W. L. Krueger and Z. Tadmor, Polym. Eng. Sei., 20: 426
(1980).
131. C. A. Hieber, L. S. Socha, S. F. Shen, K. K. Wang, and A. I.
Isayev, Polym. Eng. Sei., 23: 20 (1983).
132. H. Schlichting, Boundary Layer Theory, 6 th e d ., McGraw-Hill,
New York, 1968.
133. H. S. Hele-Shaw, Proe. Royal Inst., 16: 49 (1899).
134. B. W. Thompson, J. Fluid Meeh., 31: 379 (1968).
135. O, C. Zienkiewicz, Finite Element Method, McGraw-Hill, New
York, 1977.
136. K. H. Huebner and E. A. Thornton, The Finite Element Method
for Engineers, 2nd e d ., Wiley, New York, 1982.
134 Hieber

137. D. C. Paulson, SPE Tech, Papers, 13: 1009 (1967).


138. Y. Kuo and M. R. Kamal, Proc, Int. Conf, Polym. Processing,
N. P. Suh and N. H. Sung (e d s .), M .I.T ., 1977, p. 329.
139. T. S. Chung and M. E. Ryan, Polym. Eng. Sei., 21: 271(1981).
140. T. S. Chung and Y. Ide, J. Appl. Polym. Sei., 28:2999(1983).
141. T. S. Chung, Polym. Eng. Sei., 25: 111 (1985).
142. R. S. Spencer and G. D. Gilmore, J. Appl. P h y s .,20: 502
(1949).
143. R. Deterre, Ph.D. Thesis, Louis Pasteur University, Strasbourg,
France, 1984.
144. A. B. Tayler, Tech. Rept. No. 25, Cornell Injection Molding
Program, Cornell University, 1978.
145. D. W. Van Krevelen, D. J. Goedhart, and P. J. Hoftijzer, Poly­
mer, 18: 750 (1977).
146. N. J. Müls and A. Nevin, J. Polym. Sei. A-2, 9: 267 (1971).
147. W. W. Graessley, Adv. Polym. Sei., Vol. 16, Springer-Verlag,
Berlin, 1974.
148. A. V. Shenoy and D. R. Saini, Rheol. Acta, 23: 368 (1984).
149. T. G. Fox and P. J. Flory, J. Phys. Chem. , 55: 221 (1951).
150. T. G. Fox and P. J. Flory, J. Polym. Sei., 14: 315 (1954).
151. T. Kotaka and N. Donkai, J. Polym. Sei., A-2, 6: 1457 (1968).
152. L. H. Tung, J. Polym. S e t , 46: 409 (1960).
153. H. P. Schreiber, J. Appl. Polym. S e t , 9: 2101 (1965).
154. W. M. Prest, R. S. Porter, and J. M. O^Reilly, J. Appl. Polym.
Sei., 14: 2697 (1970).
155. W. P. Cox and E. H. Merz, J. Polym. Sei., 28: 619 (1958).
156. D. J. Plazek and V. M. O^Rourke, J. Polym. Sei. A-2, 9: 209
(1971).
157. A. V. Tobolsky, J. J. Aklonis, and G. Akovali, J. Chem. Phys.,
42: 123 (1965).
158. G. Akovali, J. Polym. Sei. A-2, 5: 875 (1967).
159. S. Onogi, T. Masuda, and K . Kitagaua, Macromolecules, 3: 109
(1970).
160. G. Marin and W. W. Graessley, Rheol. Acta, 16: 527 (1977).
161. H. J. M. A. Mieras and C. F. H. van Rijn, Nature, 218: 865
(1968).
162. H. Janeschitz-Kriegl, Polymer Melt Rheology and Flow Birefrin­
gence, Springer-Verlag, Berlin, 1983.
163. F. H. Gortemaker, H. Janeschitz-Kriegl, and K. te Nijenhuis,
Rheol. Acta, 15: 487 (1976).
164. J. L. White and A. Kondo, J. Non-Newt. Fluid Mech., 3: 41
(1977).
165. J. W. C. Adamse, H. Janeschitz-Kriegl, J. L. den Otter, and
J. L. S. Wales, J. Polym. Sei. A-2, 6: 871 (1968).
MeIt-V¡scos¡ty Characterization 135

166. J. P. Montfort, G. Marin, J. Arman, and Ph. Monge, RheoL


Acta, 18: 623 (1979).
167. E. B . Bagley, J. Appi. P h ys., 28: 624 (1957).
168. C. D. Han and M. Charles, Trans. Soc. Rheol., 15: 371 (1971).
169. J. L. S. Wales, J. L. den Otter, and H. Janeschitz-Kriegl,
Rheol. Acta, 4: 146 (1965).
170. C. D. Han and C. A. Villamizar, J. Appi. Polym. Sci., 22: 1677
(1978).
171. H. C. Tseng, G. E. Grant, C. A. Hieber, K. K. Wang, and
H. H. Chiang, SPE Tech. Papers, 21: 716 (1985).
172. A . Santamaria and G. M. Guzman, Polym. Eng. Sci., 22: 365
(1982).
173. J. P. Tordella, in Rheology, Voi. 5, Chapter 2, F. E. Eirich
(e d .). Academic Press, New York, 1969.
174. C. D. Han and R. R. Lamont, Polym. Eng. Sci., 11: 385 (1971).
175. J. P. Tordella, J. Appi. Phys., 27: 454 (1956).
176. J. Meissner, Pure Appi. Chem. , 42: 553 (1975).
177. P. J. Carreau, Ph.D. Thesis, University of Wisconsin, 1968.
178. H. C. Brinkman, Appi. Sci. Res., A2: 120 (1951).
179. R. B . Bird, SPE J., 11: 35 (1955).
180. H. L. Toor, Trans. Soc. Rheol., 1: 177(1957).
181. J. E. Gerrard, F. E. Steidler, and J. K. Appledoorn, I.E .C .
Fun dam., 4: 332 (1965); I.E .C . Fun dam. , 5: 260 (1966).
182. R. Daryanani, H. Janeschitz-Kriegl, R. van Donselaar, and
J. van Dam, Rheol. Acta, 12: 19 (1973).
183. M. R. Kamal and H. Nyun, Trans. Soc. Rheol., 17: 271 (1973);
Polym. Eng. Sci., 20: 109 (1980).
184. P. C. Sukanek and R. L. Laurence, A .I.C h .E . J., 20: 474
(1974).
185. J. Jacobsen and W. O. Winer, J. Lubr. Tech., 97: 472 (1975).
186. N. Galili, R. Takserman-Krozer, and Z. Rigbi, Rheol. Acta,
14: 550, 816 (1975).
187. C. A. Hieber, Rheol. Acta, 16: 553 (1977).
188. H. H. Winter, in Advances in Heat Transfer, Voi. 13, J. P.
Hartnett and T. F. Irvine, Jr. (e d s .). Academic Press, New
York, 1977.
189. J. R. A. Pearson, Polym. Eng. Sci., 18: 222 (1978).
190. H. Ockendon, J. Fluid Mech., 93: 737 (1979) .
191. G. D. Galvin, J. F. Hutton, and B . Jones, J. Non-Newt. Fluid
Mech., 8: 11 (1981).
192. R. C. Penwell, R. S. Porter, and S. Middleman, J. Polym. Sci.
A-2, 9: 731 (1971).
193. W. Rose, Nature, 191: 242 (1961).
194. E. B. Dussan V. , Ann. Rev. Fluid Mech., 11: 371 (1979).
136 Hieber

195. J. Loundes, J. Fluid Mech., 101: 631 (1980).


196. L. M. Hocking, Q. J. Mech. Appl. Math,, 34: 37(1981).
197. C. G. Ngan and E. B . Dussan V ., J. Fluid Mech,, 118: 27
(1982).
198. P. Bach and O. Hassager, J, Fluid Mech,, 152: 173 (1985).
199. Z. Tadmor, J, Appl, Polym, Sci,, 18: 1753 (1974).
200. H. Janeschitz-Kriegl, Rheol. Acta, 16: 327 (1977); Rheol, Acta,
18: 693 (1979).
201. W. Dietz and J. L. White, Rheol, Acta, 17: 676 (1978).
202. A. B. Tayler and M. O. Nicholas, IMA J. Appl, Math,, 28: 75
(1982).
203. S. M. Richardson, Rheol, Acta, 22: 223 (1983).
204. L. R. Schmidt, Polym, Eng, Sci,, 14: 797 (1974).
205. R. B. Bird, W. E. Stewart, and E. N. Lightfoot,Transport
Phenomena, Wiley, New York, 1960.
206. C. A. Hieber and S. F. Shen, Progress Report No. 6 , Section
4.3, Cornell Injection Molding Program, 1979.
Flow of Polymeric Melts in Juncture
Regions of Injection Molding

Avraam I. Isayev Ram K . U padhyay


Polymer Engineering Center Corporate Research and Development
The University of Akron General Electric Company
Akron, Ohio Schenectady, New York

1. IN T R O D U C T IO N

During various processing operations, polymeric melts frequently flow


through juncture regions—areas with a sudden or gradual change in
cross section (contraction or expansion). In injection molding such
regions are usually encountered at the entrance to runners, gates,
and cavities, and within the cavities themselves due to possible ribs,
bosses, obstacles, or dimensional nonuniformities. Although this sit­
uation in many respects resembles flow through fittings, considered
in fluid mechanics, the problem of polymer melt flow through junc­
tures is much more complicated, especially when evaluating stresses
and pressure losses. The major complication arises due to the visco­
elastic nature of polymeric melts exhibiting highly non-Newtonian flow
behavior accompanied by significant elasticity. Moreover, the non-
isothermal character of the flow in most polymer processing adds an­
other dimension of complexity. For Newtonian flows in classical fluid
mechanics, various experimental charts and tables have been compiled
to evaluate the pressure required to push the fluid through various
fittings [1] . Such a generalized approach to polymeric fluids is ex­
tremely difficult and does not seem to be amenable at the present time,
although these difficult problems are receiving extensive attention both
theoretically and experimentally. In particular, the theoretical work
includes numerical simulation of polymeric flow in juncture regions in
terms of various constitutive models. Despite considerable effort
by investigators, the progress has been slow, and more time will be

137
138 Isayev and Upadhyay

needed before the results can be used in practical applications. On


the other hand, the trend in experimental work includes establishing
some semiempirical correlations (master curves) for different geome­
tries useful for industrial applications . Due to limitations of both the
numerical simulations and the experimental work, however, there has
been little common ground between these two approaches. Obviously,
to verify the predictive capability of existing constitutive models for
describing real experiments, we need extensive comparisons between
simulated results and experimental data. Thus, the main scope of
this chapter is to give an up-to-date account of the developments in
the area of juncture-type flow problems along with some implications
for mold design.

II. E X P E R IM E N TA L R ESU LTS FOR TW O -


D IM E N SIO N A L FLOW OF POLYMERS

Since the actual flow in juncture regions of injection molding systems


may be nonisothermal and three-dimensional in nature, an experimental
investigation of such flow would be very complex, especially when the
effect of viscoelasticity is involved. Hence, the approach in this area
has been limited mainly to isothermal two-dimensional flow situations.
In particular, a planar and axisymmetric flow in converging and di­
verging channels with various contraction or expansion ratios and
angles will be considered in this section.

A. F lo w -T h ro u g h Channels w ith C on tractio n o r Expansion

The simple case of channel flow with contraction or expansion is en­


countered when polymer flows from a large/small to a small/large slit.
It will be referred to here as planar two-dimensional flow. The axi­
symmetric counterpart of this is when polymer flows from a large/small
to a small/large tube. Transition from a large to a small cross-sec­
tional area may be abrupt or gradual. The case of 90° entry or exit
angle will be called abrupt contraction or abrupt expansion, respec­
tively, and when this angle is less than 90°, it will be referred to as
the converging or diverging case.
In such planar or axisymmetric two-dimensional flows, changes in
the cross-sectional flow area can cause significant adjustments in the
kinematic, stress, and pressure fields, which depend not only on the
geometric configuration but also on the rheological properties of the
fluid [2 ]. In particular, the upstream section will be characterized
by a fully developed velocity profile which undergoes a transition
within the converging or diverging area and in the entrance region
to the downstream section, which becomes fully developed again if
Flow of Polymeric Melts 139

this section is sufficiently long. Similar changes occur in the shear-


and normal-stress fields. In general, the pressure drops during tran­
sition are affected by inertial, viscous, and elastic effects. For most
polymer melts, because of their high viscosity, the inertial effects are
negligible. However, a substantial contribution to the pressure drop
comes from the interaction of viscous and elastic properties of the ma­
terial .
Extensive information about the flow dynamics can be obtained by
the rheo-optical method [3] , which uses the ability of polymers to ex­
hibit birefringence during flow together with stress-optical laws re ­
lating optical and mechanical quantities for viscoelastic media. The
applicability of stress-optical rules to polymers was proposed by Lodge
[4] and experimentally verified by Philippoff [5] and Janeschitz-Kriegl
[ 6 ] for one-dimensional flows. In particular, the stress-optical rule
states that the birefringence is linearly proportional to the principal
stress difference Aa, namely An = C Ao, and the orientation of the op­
tical axis, Xo> 9 .nd the principal stress axis, Xg» coincide; i.e ., Xq =
Xg= X. In particular, C is the stress-optical coefficient, and X is the
extinction angle.
The rheo-optical technique has been applied by Bogue et al. [7—9]
to study planar two-dimensional flow of polymer solutions in converg­
ing and diverging channels. The following mathematical relations [ 8 ]
in cartesian coordinates can be written as

( 1)

a - a = Aa cos 2X^ = ^ cos 2x ( 2)


XX yy s c o

where Oxy is the shear stress and Oxx ~ Oyy is the first-normal-stress
difference. The transformation of stresses in polar coordinates can be
performed as follows:
Converging

= -Ö )sin 2 0 - a cos 2 0 ( 3)
re 2 XX yy xy

a - = (a - a )cos 2 0 - 2 a sin 26 (4)


rr 06 XX yy xy

Diverging

a . = -^ (a -a ) sin 2 0 + 0 cos 2 0 (5)


re 2 XX yy xy
a - a._ = (a - a )cos 2 0 + 2 a sin 2 0 (6)
rr 00 XX yy xy
140 Isayev and Upadhyay

Thus, the shear stress and the first-normal-stress difference can be


calculated from the measurements of An and X at a given point.
The typical results [ 8 ] from such measurements for a constant flow
rate are shown in Fig. 1 for a 12% solution of polystyrene in Aroclor
1242, with the geometry of the working cell being given on the right
of Fig. 1. These figures show the distribution of shear stress and
the first-normal-stress difference in the thickness direction at dif­
ferent locations along the flow direction measured from the entry edge
for converging and diverging flows. In both cases the shear stress
increases linearly in the thickness direction from zero at the center-
line to a maximum at the wall. The first-normal-stress difference is
nonzero everywhere with its value increasing or decreasing from the
centerline towards the wall, depending on the axial location in the
converging case, but always increasing in the diverging case. How­
ever, these measurements have been reported [ 8 ] to be close to the
viscoelastic behavior of the second-order fluid. In particular, the
second-order fluid describes the viscoelastic behavior of polymer
melts only in low-strain-rate range.
Significant contribution to our understanding of the viscoelastic
behavior of polymer flow through converging dies has been made by
several authors using the rheo-optical technique. In particular, Han
et al. [ 2 , 1 0 , 1 1 ] have studied the stress distribution in converging
dies for different polymeric melts. The first-normal-stress differ­
ence along the centerline - cree for a polystyrene melt is shown
in Fig. 2. It can be seen that Orr ~ <^60 first increases, goes through
a maximum before the melt reaches the exit of the die, and then de­
creases.
Further studies of converging flow have been made by Brizitsky
et al. [ 1 2 ] for polybutadienes and polyisoprenes of various molecu­
lar-weight distributions. Figure 3 shows a typical distribution of
^ x x " ^yy along the centerline in the flow direction at various flow
rates and contraction angles a in converging flows of polybutadiene.
The position of the maximum is independent of the contraction angle.
At the same time the distance upstream from the duct entrance where
the first-normal-stress difference starts to build up increases with
decreasing convergence angle. It seems to follow from these results
that the distance between the entrance and the cross section of the
small channel at which the first-normal-stress difference starts to
build up cannot be determined. This is apparently due to an insuf­
ficiently long preentrance channel, indicating that the stress pro­
file was not fully developed in the latter section. However, the
lengths of the small channel were sufficiently long, resulting in full
relaxation of the normal stresses. The nondimensional distance over
which the first-normal-stress difference drops to zero, X q/Yi (h is the
[a) (a)
- x /h = 3 .7 o
- x /h = 3 .7
THuy
Developed

“D
O,
■<
Depths.89cm 3
(D
n’

r+
in

°rr-°68'P^
Fig. 1 Gapwise distribution of shear stress, a^e (a ) and first-normal-stress difference,
(b ) during flow of 12% wt. polystyrene (Styron 6 6 6 /Dow) in Aroclor 1242 through converging and di­
verging channel at various cross sections and temperature of 25°C. (Results from Ref. 8 .) Flow
rate is 10.4 gm/sec.
142 Isayev and Upadhyay

Fig. 2 First difference of normal stresses a^r " ci0 0 along centerline
of a channel with converging angle of 60° for polystyrene (Styron
678/Dow) at flow rate of 1.5 cm^/min and temperature of 200°C. (From
Ref. 11.)

duct thickness) has been found [ 1 2 ] to be related to the value of the


molecular mass of the polymer chain between physical entanglement,
Me, and the molecular-weight distribution. In particular, for vari­
ous narrow- and broad-distribution polybutadienes and polyisoprenes,
a unique dependence of (xo/h)/Me^ on Yw^o exists and is shown in
Fig. 4. Here, Yw is the shear rate at the wall of the small channel,
and 0 Q is the zero-shear-rate relaxation time determined as a prod­
uct pf the zero-shear-rate viscosity and elastic compliance of the
polymer. Since 0q increases with brotóening of molecular-weight
distribution [13], it follows from the master curve in Fig. 4 that
stress relaxation in polymers with wider molecular-weight distribu­
tion is slower than in those with narrow distribution.
Flow of Polymeric Melts U3

Fig. 3 Curves of distribution of first-normal-stress difference along


flow axis in reservoir and inside ducts for polybutadiene (Mw = 1 . 8 x
10®, Mw/Mn = 1.1) at 25°C. Shear stress on duct wall 0i2,w ^’^ units
of 10® N/m^; (a ) curve 1-1.51; 2-1.82; 3-2.22; 4-2.40; (b ) curve
1- 0.58; 2-1.16; 3-1.70; 4-1.99; 5-2.29; 6 -2 .4 ; (c ) curve 1-0.65;
2- 1.06; 3-1.41; 4-1.70; 5-2.10; (d ) curve 1-0.43; 2-0.83; 3 -
1.08; 4-1.33; 5-1.67; 6-2.08; 7-2.22. (From Ref. 12.)
144 Isayev and Upadhyay

Fig. 4 Master curve for reduced length over which first-normal-stress


difference relaxes to zero vs. reduced deformation rate for various
poly butadienes and polyisoprenes at 25®C. Polybutadienes: a — =
1.35 X 10\ Mw/Mn = 1.1; ■ - Mw = 1.8 X l0 ^ Mw/Mn = 1 . 1 ; • - Mw
3.2 X l0 ^ Mw/Mn = 1.1; ▼ - Mw = 2.4 X l 0 ^ Mw/Mn = 3.0; polyiso-
prenes: a —Mw = 3.75 x 1 0 ^, Mw/Mn = 1.14; n-M w = 5.75 x 10^ Mw/
= 1 . 0 2 ; o —M ^ = 8.3 x 1 0 ^, M^v /Mn = 1 . 1 2 ; v - Mw = 6 . 0 x l 0 ^
Mw/Mn = 2.4. (From Ref. 12.)

Some interesting viscoelastic effects in polymer flows can be ob­


served by reversing the flow direction in a given juncture geometry
and by measuring velocity and stress fields. It is well known that
the reversal of flow direction has no effect in generalized Newtonian
fluids; however, there exist substantial differences when the flow di­
rection of viscoelastic fluids is reversed, especially at high flow rates
[14—17], namely beyond the second-order-fluid range. In particular,
the pressure loss and birefringence profiles differ for contraction and
expansion flows for the same geometry and flow rates. Extensive ex­
perimental data will be presented here for the planar working cells of
geometry depicted in Fig. 5 and will show various aspects of this ir ­
reversibility phenomena.
Figure 6 shows the experimental birefringence at different cross
sections located at dimensionless length x/h (distance along flow axis
Flow of Polymeric Melts U5

DEPTH = 2.0 cm

Cell H, cm 2b, cm L, cm a degrees

0.795 0.108 0.95 14.3

0.803 0.148 0.80 29.6

0.806 0.110 0.80 42.6

0.800 0.400 0.95 90.0

0.822 0.121 0.95 90.0

2b

Fig. 5 Schematic sketch of the testing cells (all dimensions in centi­


meters). Points 1 and 2 indicate the locations of pressure transdu­
cers .

in terms of gap thickness of small channel) from the entrance to the


small channel for working cells A , D, and E with (a ) denoting that
the flow is directed from the large channel to the small, and (b ) de­
noting that the flow is from the small channel to the large. In most
of these cases the birefringence measurements have been made only
in the large channel and in the tapered part, due to limitations of
the optical system. There are substantial differences in the gapwise
variation of birefringence among the different cells. In the converg­
ing channel the birefringence at any cross section increases mono-
tonically toward the wall, indicating the absence of any secondary
flow. On the other hand, in the abrupt-contraction cases, there is
a maximum with subsequent decay toward the wall (cells D and E ) ,
indicating a low-intensity deformation region. Further, the Dosi-
tion of the gapwise maximum shifts towards the centerline as one
CELL A CELL D CELL E
(a)
.pr
Oi

W
Q)
a>
<
0)
D
a
c
■D
0)
a
DT
0}
Fig. 6 Gapwise birefringence distribution at different cross sections in converging (a ) and diverging (b )
cases for a constant volumetric flow rate. Data for working cells A, D, and E in Fig. 5. Arrows indicate
the location of wall.
Flow of Polymeric Melts U7

Fig. 7 Position of the gap wise-birefringence maximum, Ymax’ vari­


ous upstream cross sections vs. distance, -x , from the entrance to the
small channel at two volumetric flow rates, Q. Data for working cell D
in Fig. 5: Q = 3.08 x lO“^ ( ■ ) and 1.08 x 10"^ m^/sec ( □ ) . Data for
cell E in Fig. 5: Q = 7.16 x 10‘ ® ( • ) and 1.32 x 10"^ m^/sec ( o ) . A r­
rows indicate the transverse locations of upstream and downstream
walls. (From Ref. 17.)

approaches the entrance to the small channel. Results for cells D and
E also show that the extent of the low-deformation-intensity region
increases as one moves towards the entrance and the main flow goes
through a converging funnel formed naturally by the polymer stream
in the core region where the birefringence behavior is similar to that
of the converging flow in cell A ( i . e . , the birefringence increases in
the gap wise direction). The angle of this funnel, a, can be deter­
mined from Fig. 7, which shows the gap wise position of the maximum,
Ymax» different upstream cross sections versus distance (- x ) from
the entrance to the small channel for two flow rates and both test
cells with abrupt contraction. Within the investigated range of the
flow rates, this angle is about 33°—35° and independent of contrac­
tion ratio and flow rate.
U8 Isayev and Upadhyay

CELL A
0 KlO^,m^/$ec

Fig. 8 Birefringence along centerline for converging (a ) and diverg­


ing (b ) flows at different volumetric flow rates. Data for cells A, G,
D, and E. Arrows on x/h axis indicate start or end of converging or
diverging section.

Similar data are presented for diverging and abrupt-expansion flows


in Fig. 6 (b ). For diverging flow the qualitative behavior is similar to
that of converging flow. However, a comparison of the two abrupt-ex­
pansion cases (cells D and E) shows a qualitative difference. For
H/h = 2 (cell D) the birefringence values at first increase and then de­
crease from the centerline toward the wall, with a local maximum near
the wall, but for H/h = 6.79 (cell E) there is a monotonie decrease.
It seems that the gapwise location of this maximum shifts towards the
centerline with increasing contraction ratio, indicating a more exten­
sive low-intensity deformation. By implication, the maximum may ap­
proach the centerline at a very high contraction ratio.
Figure 8 shows the development of birefringence along the center-
line for various cells A~E for contraction and expansion flows. Each
Flow of Polymeric Melts U9

CELL C

X/ h

curve in these figures corresponds to a specific flow rate, Q. The


distance is measured from the start of the small channel, with posi­
tive values denoting positions along it. In contraction flow the bi­
refringence starts to build up in the large channel [Fig. 8 ( a ) ] , con­
tinues to increase in the preentrance region, and then passes through
a maximum, close to the entrance to the small channel. Thereafter,
the relaxation occurs in the small channel. Depending on the driving
pressure or flow rate, complete relaxation may or may not occur with­
in the length of the channel. In diverging flow [Fig. 8 (b )] the b i­
refringence seems to start building up just before the exit from the
small channel, passes through a maximum after the exit, and then
150 Isayev and Upadhyay

CELL D
(b)

relaxes. Complete relaxation occurs further downstream in the large


channel. The position of the maximum shifts slightly downstream with
increasing flow rate. At the two highest flow rates for the abrupt-ex­
pansion flow with H/h = 2 (cell D ) , a fully developed flow has appar­
ently not been established in the small channel, as indicated by an
extension of curves 3 and 4. This nonzero birefringence in the small
channel is probably due to a corresponding small length to thickness
ratio L/h (8.75 in this case).
Flow of Polymeric Melts 151

Figure 9 presents a comparison of the centerline birefringence max­


imum, An„jax» versus pressure drop, Ap (measured by transducers 1
and 2, shown in Fig. 5), and flow rate, Q, for contraction and expan­
sion flows. This comparison indicates that An^ax is larger in contrac­
tion than in expansion, and the difference between the two increases
with increasing Ap or Q. Further, to compare the results among the
different cells (corresponding to different gap thicknesses) in order
to understand the effect of contraction ratio and contraction angle,
we see that Figs. 10(a) and (b ) show An^ax a function of wall
shear rate, for contraction and expansion flows in cells D and E,
and A, C, and E, respectively. It is clear from Fig. 10(a) that the
contraction ratio significantly influences An^ax —namely, higher b i­
refringence for higher contraction ratio at the same value of wall
shear rate. Fig. 10(b) shows that the higher contraction angle re ­
sults in higher birefringence. The effect of contraction ratio is es­
sentially the same in Fig. 10(b) because the test cells have essen­
tially the same values (7.36, 8.06, and 6.79) of contraction ratios.
Physically, the value of total elongational strain is the same in all
three cases (contraction ratios are essentially the same), but the
length of channel for build-up decreases with large contraction
angle, which causes higher elongational strain rates and hence
higher birefringence. The effect of contraction angle a is clearly
seen in Fig. 11, which shows An^ax a function of a for = 5.0
and 31.6 sec"^. In particular, the birefringence maximum in con­
traction and expansion flows increases monotonically with a, with
the slope of the curves being larger for small a. In the low-shear-
rate region, where the flow is approaching the Newtonian or sec­
ond-order-fluid limit, the difference between the contraction- and
expansion-flow results becomes negligible, as shown in Fig. 10.
Similar observations for contraction flow have previously been
made [12], and it seems clear that the stresses in polymeric melts
increase with increasing contraction ratio and angle. However, the
behavior at very large contraction ratios is not clear. Besides con­
traction ratio, the value of L/h can also influence the stresses. In
particular, this effect has been observed [12] for L/h < 7, with the
stresses being found to be independent of L/h for L/h > 7.
Figure 12 shows the volumetric flow rate Q as a function of Ap.
It also shows the nondimensional axial position Xo/h at which the cen­
terline birefringence starts to build up in contraction flow or com­
pletely relaxes in expansion flow. The effect of the reversal of flow
direction is apparent in all the test cells, and this effect is stronger
at higher contraction ratio and contraction angle and also at higher
flow rates for each individual cell. In general, it can be concluded
that polymeric melts require a smaller pressure drop in a contraction-
flow situation than in expansion at the same flow rate. Further, Fig.
12(b) shows that the contraction flows have lower Xo/h values which,
in all cases, increase with Q.
152 Isayev and Upadhyay

CELL A

Ap xio'®. Po

F ig . 9 Maximum of birefringence along centerline vs. volumetric flow


rate (a ) and pressure drop (b ) for various cells in Fig. 5 for contrac­
tion (1) and expansion ( 2).

Finally, Fig. 13 shows the ratio of expansion to contraction pres­


sure drop as a function of a for = 1.0 and 31.6 sec~^ in test cells
with approximately the same contraction ratios. This ratio is always
greater than unity and increases with both contraction angle and flow
rate.

B. E n try and E x it Flow E ffects

Results in the last section have shown that the dependence of pres­
sure drop on the volumetric flow rate is quite different in contrac­
tion and expansion flows. Evidently these differences come from the
entry and exit flow effects. Additional energy is required in deform­
ing the melt at the entrance to contraction and expansion areas and
for the development of the velocity and stress fields in the transition
region. Capillary rheometry is one general area where the entry and
exit effects have been carefully studied and quantified to properly
determine the viscosity of polymeric melts. Some of the procedures
Flow of Polymeric Melts 153

CELL B

CELL C

log Ap, Pa log Q, m^/sec

used in capillary rheometry will be briefly discussed here. There


have also been major efforts to understand the flow patterns in the
die entry region. In this regard, review papers by White and Kondo
[18] and Boger [19] are of particular interest.
154 Isayev and Upadhyay

CELL D

9 3
M
K
o
E
c
< 2

CELL E
Fig. 10 Maximum of birefringence along centerline vs. wall shear rate in small channel, (a ) For abrupt-con-
traction (curves 1 and 3) and abrupt-expansion (curves 2 and 4) flows. Curves 1 and 2 correspond to cell E,
and curves 3 and 4 to cell D. (b ) For abrupt-contraction (curve 1), abrupt-expansion (curve 2), converging
(curves 3 and 5), and diverging (curves 4 and 6) flows. Curves 1 and 2 correspond to cell E, 3 and 4 to C ,
and 5 and 6 to A .
156 Isayev and Upadhyay

y = 31.6 s e c
' w

V =5 s e c ’
'w

F ig . 11 Dependence of birefringence maximum along centerline vs.


contraction angle (curves 2 and 4 for converging flow, and 1 and 3
for diverging flow) at two wall shear rates in small channel for cells
A, E, and C having approximately same contraction ratio.

CELL A

Fig. 12 (a ) Pressure drop Ap vs. Q for contraction (curve 1) and


expansion (curve 2) flow in various cells, (b ) Channel length, Xo/h,
at which centerline birefringence starts to build in contraction (curve
1) or completely relaxes in expansion (curve 2) flow in various cells.
Flow of Polymeric Melts 157

CELL B

(b )

-7

lO
■e
o -8

_L_
10

CELL C

-7

-8 -

J
10

Bagley [20,21] proposed a correction factor ne expressed in terms


of the capillary radius R as follows:

(7)
R 2a
12,w

where ai2,w is the wall shear stress, and Le and Apg are an effec­
tive extra length and pressure drop to be determined by a linear
158 Isayev and Upadhyay

CELL D

E
d

CELL E

extrapolation of pressure drop versus capillary length to zero length.


The terms ne and Ape include both the entrance and exit effects. The
separation of these effects requires detailed measurements of the pres­
sure development along the channel. At present it seems to be well
accepted to neglect the exit effects in the total correction factor.
Flow of Polymeric Melts 159

= 31.6 s e c " ’
Isec"'

Fig. 13 Dependence of pressure drop ratio for expansion and con­


traction flow vs. angle at two wall shear rates for cells A , C, and E
having approximately same contraction ratio.

In general, ng and Apg have been found to be much higher for


polymer melts than for generalized Newtonian fluids, especially at high
shear rates. More than two decades ago, Philippoff and Gaskins [22]
proposed to separate ng into two components corresponding to vis­
cous dissipation and stored elastic energy, namely

N
l,w
n« = n„ + s^, ( 8)
2a
1 2,w

where Uy is the Couette correction for a purely viscous fluid, s^ is


elastic recoil, and N i,w is the first normal-stress difference at the
capillary wall. The viscous dissipation term comes from the energy
loss due to a reorganization of the velocity fields at the entrance and
exit. For the value of ny, there is no single accepted opinion in the
literature. Barr [23] has suggested 0.9, Schurz [24] 0.5—1.0, and
Sulzhenko and Kuvshinski [25] 1.146. Philippoff [22], and, later,
others [26—29] have noted that ny is small in comparison with Sj* and
does not depend on the volumetric flow rate. On the other hand,
Philipp and Wulf [30] found a strong dependence of ny on the strain
rate. At high strain rates ny is quite large. These contradictions
seem to be due to the fact that ny has to be measured on inelastic
fluids, whereas polymeric melts are highly viscoelastic.
160 Isayev and Upadhyay

Further, for the extra pressure loss, Ape> in flow from an infinite
reservoir to a hole, many different relations have been proposed. For
Newtonian fluids, Weissberg [31] used a creeping flow analysis to de­
termine

3nQ (9)
2R®

where r\ is the viscosity. Later, La Nieve and Bogue [32] showed


that Eq. (9) agrees well with experimental data. Chong et al. [33]
then generalized Weissberg^s equation by using the Helmholtz theorem
of minimum energy dissipation in the following form:

cnQ
( 10 )

with c being a hole constant. However, Volkov et al. [34] experimen­


tally found that c is not always a constant, but it depends on the vol­
umetric flow rate and fluid viscosity. They proposed the following
empirical relation, valid for their Newtonian fluids (polymer solutions);

Ap^ = 6 . 0 3 Q ° - % 0 - 2 4 + 0. 2 1ogQ ( 11 )

Holmes [35] proposed the following general equation on the basis of


dimensional analysis combined with boundary-layer and creeping-flow
limiting behavior:

AP. ( 12)

where ki and k£ are Hagenbach and Couette coefficients, respectively,


and Re is the Reynolds number . The first term in this equation comes
from boundary-layer analysis [36—39], and the second term corres­
ponds to the creeping-flow solution [31]. Sylvester and Rosen [40]
experimentally verified Eq. (12) for Newtonian fluids and found k^ and
k£ to be an increasing function of contraction ratio. For viscoelastic
fluids Sylvester and Rosen [41] found much higher extra pressure
losses than predicted by Eq. (12) for inelastic fluids with comparable
viscous properties. By assuming Hooke’s law to be valid for visco­
elastic fluids, they derived the following relation for the extra pres­
sure loss:

3n + 1 12,w
Ap (13)
e 4(5n + 1)
Flow of Polymeric Melts 161

where n is the power-law index of the shear viscosity and G is the


Hookean shear modulus. This equation is valid for various polymer
solutions at low shear stresses but overpredicts at high shear stress­
es, which, according to the authors, is due to a breakdown of Hookers
law.
Using the power-law model. Boles et al. [42] have derived the ex­
pressions for entrance pressure drops in contraction flow treated as
tapered cone. Their results are in good agreement with experimen­
tal data at low shear rates, but they underpredict at higher shear
rates. This discrepancy seemingly arises from the significant con­
tribution of fluid elasticity to the pressure drop. Also, it has been
observed that the measured pressure drop is essentially independent
of cone angles at high shear rates due to the channeling effect.
Based on the assumption that elongational flow at the entrance re­
gion is predominant, Cogswell [43] has proposed the following equa­
tions for the extra pressure loss for flow of a power-law fluid from a
large reservoir into a slit or a circular die, respectively:
slit die

(n + l)/2
2 I2 n + 1 ^ ^ ^ ./2 /JQ V
(4noK) (14)
l2Wh/
® /3( n . 1 ) ^ >
circular die

(n + l)/2
4 /2
Ape = 3(n (15)

where no is the initial Newtonian viscosity, K is a preexponent in the


power-law equation, and W is the width of the slit. The derivation of
Eq. (14) assumes that elongational viscosity is constant and given by
X = 4no foi* a two-dimensional, wedge-type extension at the slit en­
trance. Similarly, Eq. (15) is based on a constant elongational vis­
cosity assumption, X = 3rio> uniaxial extension at a circular die en­
trance. Generally, it has been found that Cogswell’s equations tend
to overpredict experimental entrance pressure losses, particularly at
high flow rates relevant to injection molding [44,45]. Kwon et al.
[44,45] have modified the Cogswell equation by introducing power-
law dependence of elongational viscosity and an additional term due
to first normal-stress difference.
Recent analyses of entrance and exit effects have been presented
by Boger and Denn [46], using a macroscopic momentum balance, and
by Choplin and Carreau [47], employing a microscopic mechanical
energy balance. Although the results of both papers seem to be
162 Isayev and Upadhyay

equivalent, the former has mainly reviewed the capillary and slit meth­
ods for normal-stress measurements, whereas the latter has consid­
ered the problem of extra pressure loss, which is more relevant to
this chapter. The following relations for the extra length, which are
similar to those of PhUippoff and Gaskins [22] [Eq. ( 8 ) ] , have been
derived [47] for entrance and exit flows:
entry flow

n - n + n, - n. ^ (16)
e V k N,

exit flow

n =n (17)
e V k N,

where n^ is the kinetic energy correction, which is negligible for high-


viscosity fluids, and nN¿ is the correction due to normal stresses
(stored elastic energy). The stored elastic energy term is subtracted
from the viscous dissipation term, thereby reducing the entry loss,
whereas most experiments indicate otherwise. Therefore the authors
[47] have suggested that coupling elastic properties with viscous
properties may increase n v For exit flow, on the other hand, the
elastic term is added to the viscous term. The sum of the entry and
exit corrections is then the total correction for a slit or capillary,
which is the same as the Bagley correction factor. The kinetic and
elastic energy corrections do not contribute to the Bagley correction.
Based on this, the authors [47] conclude that the large extra pres­
sure loss in viscoelastic fluids is due to a coupling between elastic
and shear-thinning properties, which seems to be substantiated by
experimental data with highly elastic and inelastic fluids with shear
thinning and with elastic fluids having little shear thinning.
Laun [48] has measured the pressure gradient along a rectangular
channel in order to separate the entrance and exit pressure losses in
a low-density polyethylene (L D P E ). The exit losses were found to be
much smaller than the entrance losses reported earlier for polymer so­
lutions. In addition, Laun [48] has determined the Bagley correction
for a circular die flow and has found that the sum of the entrance and
exit pressure losses in the slit are the same as the Bagley correction
for the circular dies. This indicates that the difference between the
planar and axisymmetric geometries has minor influence on the total
extra pressure loss.
Further, La Mantia et al. [49] have carried out a comprehensive
experimental study of entrance and exit effects in capillary flow for
various high-density polyethylenes (HDPE) and HDPE—LDPE blends.
Flow of Polymeric Melts 163

7.77

Fig. 14 Schematic representation of die. Point 1 indicates location of


flush-mounted pressure transducer—thermocouple. All dimensions are
in millimeters.

with particular attention given to the influence of molecular weight,


Mw, molecular-weight distribution, and temperature. Generalized
curves for end correction ne versus ya^ have been obtained, which
show a temperature independence but a dependence on molecular
weight and index of poly disper sity , Mw/M. In particular, val-
ues of Uq have been shown to increase with and In addition,
the correction factors for the blends were found to lie between the
ne values of the components when the two homopolymers had very
different values. However, they were found to be much larger when
two were close. For the investigated blends this latter behavior was
noted when the molecular weight of HDPE was larger than LDPE.
Recent measurements, using a screw extruder, of the extra pres­
sure loss in circular dies [50] have been made for various glass-fiber-
filled and unfilled polymer melts in the shear-rate range of 10^—10^
sec"^. The die consists of a reservoir and an interchangeable capil­
lary (see Fig. 14, which shows the cross-sectional geometry of the
reservoir and the location of the pressure transducer/thermocouple)
and is attached to the extruder for continuous extrusion. The cross
section of the reservoir has a shallow slot along the flow direction
which is suitable for mounting the transducer flush to the internal
surface. The hydraulic radius of the reservoir cross section is 4.654
mm, and six interchangeable capillary tubes of 0.75 mm radius with
lengths varying from 0.674 mm to 20 mm have been used. The poly­
mers investigated were polystyrene (P S ), polypropylene (P P ), and
unfilled polycarbonate (P C ). The unfilled polystyrenes were Styron
678U (D ow ), with = 2. 51 x 10^ and Mw/Mn = 2.04, and Mobü llOR,
with Mw = 2.71 X 10^ and M^/Mn = 1.78. The latter was also investi­
gated with 20% (by weight) of glass fibers. The PP was also studied
with 20% and 40% glass fibers, whereas the PC sample was Lexan 141-
112 (G . E .). In addition to extrusion experiments, a Sieglaff-McKel-
vey capillary rheometer was used to obtain flow curves.
164 Isayev and Upadhyay

log Q, cm^/sec log Q , cm^ / sec


Fig. 15 Ap versus Q in capillaries of various length, L, at different
temperatures for PS Styron 678U (Dow). L(10“^ m) = 0.647 (X ), 1.155
( • ) , 2.03 ( □ ) , 5.00 ( V ) , 10.0 ( A ) , 20.0 ( o ) . (From Ref. 50.)

Figure 15 shows the pressure drop as a function of volumetric flow


rate in PS 678U for various capillary lengths and different tempera­
tures. The slopes of these curves decrease with an increase in cap­
illary length (results in large deformations and longer residence times) ,
indicating stronger non-Newtonian behavior at larger lengths. The
true flow curves are shown in Fig. 16 after making Bagley and Rabino-
witsch corrections to the data in Fig. 15 along with the data from a
capillary rheometer. The rheometer and extruder data are in good
agreement, indicating that no destructive changes are introduced in
the viscosity of the PS melt. The 172°C curve shows a tendency to­
ward Newtonian behavior at high shear rates, which may be attributed
Flow of Polymeric Melts 165

Fig. 16 Flow curves for PS Styron 678U (Dow) at different tempera­


tures. Open and filled symbols are from extruder and capillary rhe­
ometer. (From Ref. 50.)

to a forced glass transition induced by high-shear rate and low tem­


perature. Similarly, Ap-versus-Q curves for PC melt are shown in
Fig. 17.
From Bagley plots (pressure drop versus capillary length) shown
in Fig. 18, the extra pressure loss (corresponding to zero length of
capillary) Ape, which is the sum of entry and exit losses, is easily de­
termined. The nondimensional extra length L q /R can then be calcu­
lated from Eq. (7 ). In particular, resulting plots of Ape and Lg/R
versus wall shear rate shown in Fig. 19. These quantities in­
crease with shear rate and decrease with temperature. At high shear
rates Apg is as large as 10 MPa, which can make a significant contri­
bution to the overall injection pressure. The extra length of the PS
log Q , cm ^/sec

Fig. 17 Ap versus Q for PC Lexan 141-112 (G . E .) at 280®C for


L(10'^) = 1.155 ( • ) , 2.03 ( □ ) , 5.00 ( v ) , 10.0 ( ) .
a

(a) (b)

Fig. 18 Ap vs. L for different Q (cm^/sec) for PS (a ) and PC (b )


melts.
Flow of Polymeric Melt^ 167

(a)

Fig. 19 Dependence of extra pressure loss, Ape, (a ), and nondimen-


sional extra length, Le/R ( b ) , on shear rate, Yw> for PS Styron at
various temperatures and for PC Lexan at 280°C. (From Ref. 50.)

melt is as large as 15 radii, but it is only about 1.2 for PC melt (in
agreement with Yamada and Porter [5 1 ]). However, the actual mag­
nitude of Ape for PC is comparable to that of PS (because of a rela­
tively high viscosity for the PC melt) and may be significant. In ad­
dition, the Le/R curve for PS at 172°C shows a maximum, reflecting
an approach toward a second Newtonian behavior (see Fig. 16).
Viscosity-versus-shear-rate curves obtained from the capillary
rheometer and extruder are shown in Fig. 20 for pure and glass-fi­
ber-filled PS and PP at different temperatures. No extruder data
could be obtained at 180®C for the PP melt or at 200°C for the PS
melt because of excessive power requirements. The rheometer and
extruder data are seen to be in good agreement for pure PS and PP,
but extruder systematically measures lower viscosities in glass-fiber-
filled cases (a, b ). However, the rheometer data of the melt collected
from the die exit agrees with the extruder data ( c ) , which implies
that the rheological properties of the filled polymer are affected by
intense shearing in the extruder. This is apparently due to fiber
breakage and structural changes in the filled polymers [52,53]. Vis­
cosity reduction in glass-fiber-filled polymers has also been observed
168 Isayev and Upadhyay

Fig. 20 (a ) Shear viscosity, n> vs. shear rate, y, at different tem­


peratures for PS melt (Mobil llO R ); open and filled symbols represent
data from extruder and rheometer, respectively, (b ) n vs. Y for PP
melt at different temperatures and for 40% glass-fiber-filled PP (♦ , o)
at 200°C; open and filled symbols represent data from extruder and
rheometer, (c ) processing effect on the viscosity curves of 20% filled
PS melt. Data from extruder or rheometer after various processing
conditions at 200°C ( ) , 230°C ( ) and 260°C ( □ ) are original/rheom­
a o

eter (fu ll), injection-molded—rheometer (open), extruded/rheometer


(upper f u ll), extruded/extruder (lower full). (From Ref. 50.)

in injection molding (Fig. 20c). Hence, such a change in rheological


properties should be taken into account in the design of molding or
extrusion processes for such materials.
Figure 21 shows Ape~versus-Yw curves for pure PS (a ), 20% glass-
fiber-filled PS ( b ) , and both pure PP and 40% filled PP (c ) at various
temperatures, respectively. The extra pressure loss Ape for the pure
polymers is significantly lower than for the filled polymers and de­
creases with increasing temperature. The corresponding Lq /R-v e r­
sus-Yw curves are shown in Fig. 22, where we see that the extra
length can reach as high as 30 radii for the filled polymers. That is,
the normalized extra pressure loss Le/R is larger for the filled melts
even though the apparent viscosity (which, in effect, occurs in the
denominator of Lg/R) is also larger for the filled material.
Flow of Polymeric Melts 169
170 Isayev and Upadhyay

Fig. 21 Ape vs. Yw ^or (a ) pure PS (Mobil llOR) , (b ) 20% glass-fi­


ber-filled PS and (c ) pure PP melt (open symbols) at different tem­
peratures and 40% filled PP melt (solid) at 200®C.

From an applications point of view, it would be useful to find some


empirical correlations between the extra pressure loss and the struc­
tural, rheological, and processing characteristics of polymers. It has
been found that Apg data can be correlated by the parameter Y tIo/3,
where 3, the index of polydispersity, represents the elasticity of the
Flow of Polymeric Melts 171

material. Figure 23 shows this correlation for the previous two in­
vestigated polystyrenes at different temperatures and the data ob­
tained by La Mantia et al. [49] for six HDPEs with various molecu­
lar weights and 3. In both cases the data collapse onto a single
curve. Further work is needed to obtain such a correlation for other
classes of polymers.
The pressure loss data in this section have included both entry
and exit losses. To separate the exit losses, Han [54] and Boger and
Denn [46] have proposed the following differential equations:
for a capillary

\T a.
Ni + 2N2=p^ + — — (18)
W

for a slit
3p
s
Ni = p + T (19)
s w 9t
w

where Ni and N 2 are the first- and second-normal-stress differences,


whereas pc and ps denote the exit pressure losses in a capillary and
slit. There has been considerable discussion in the literature con­
cerning the validity of Eqs. (18) and (19), because their derivation
172 Isayev and Upadhyay

Fig. 22 L q / R v s . for (a ) PS (Mobil llOR) melt, (b ) 20% filled PS


melt and (c ) pure PP melt at different temperatures and 40% filled PP
melt at 200°C.

disregards the velocity rearrangement near the die exit [46,55]. Ex­
perimental checks of the validity of these equations is difficult due to
many sources of error involved in measuring the exit pressure, in­
cluding an extrapolation of the pressure distribution along the die
wall to the die exit, viscous heating, and pressure-dependent viscos­
ity [48,55,57]. In particular, the values of Ni obtained from Pg by
Eq. (19) are consistently higher than those obtained from a cone-plate
device [56]. Thus, the main usefulness of these equations is to give
Flow of Polymeric Melts 173

an upper bound for the exit pressure and its contribution to the over­
all Ape. From Eqs. (18) or (19) the entry and exit pressure losses
can be separated from the total extra pressure loss. To integrate
Eqs. (18) and (19) , we need the dependence of Ni and N 2 on the
shear rate Y. In particular, N i(y ) and N 2 (y ) can be obtained by
using the Leonov constitutive equation [58], and the model parame­
ters can be determined by fitting the experimental steady-shear-vis­
cosity curves to the corresponding theoretical function (see Section
3.2 for details of the Leonov model and its viscometric functions). In
a multimode Leonov model the parameters are rik represent­
ing a viscosity and relaxation time of the kth relaxation mode, and s
lies between 0 and 1. Numerical values of these parameters for two
different polymers using a three-mode representation are given and
the corresponding fits with the experimental data are shown in Fig.
24.

polystyrene (Mobil llOR ) at 230^C


s = 1.00 X 10'^

Tik = 1.37 X l0 ^ 3.72 X 10^, 7.69 x 10^ Pa-sec

0j^ = 1.13 X loS 9.66 X lo■^ 6.00 x l o '- sec


174 Isayev and Upadhyay

Fig. 23 Master curves for Ape versus Yno/3 for (a ) two PS melts at
different temperatures and (b ) six HDPE melts at 190°C [48]. (From
Ref. 50.)
Flow of Polymeric Melts 175

Fig. 24 n vs. Y for PS (Mobil llOR) at 230°C (curve 1) and LDPE at


150®C (curve 2). Curves: theoretical fits; symbols: experimental
data. Data for LDPE from Laun [48]. (From Ref. 50.)

LDPE at ISO^C

s = 4.69 X 10"^

rik= 3.42 X 10\ 5.17 x l0 ^ 3.73 x 10^ Pa*sec

Qy. = 5.66 X 10°, 1.32 x 10"\ 2.12 x 10"^ sec

Figure 25 shows the experimental Ape and the calculated exit pres­
sure loss as a function of Y^ for PS. The exit loss accounts for only
15—20% of the total extra loss.
For the LDPE melt Laun’s [48] experimental data of exit pressure
loss as a function of wall shear stress t w , and his corresponding the­
oretical results are presented in Fig. 26. At low shear stresses, the
calculated results are in the middle of experimental scatter but signi­
ficantly higher at higher stresses. This discrepancy may be due to
flow instability at high shear stresses which can influence the exit
pressure measurements.
Further, the effect of geometry (capillary-versus-slit flow) is seen
in Fig. 27, which shows a consistently higher exit loss for the capil­
lary. The observed differences may vary from 100% to 40%, going from
low- to high-shear-stress range.
176 Isayev and Upadhyay

o
Q.
(D
O

log y , sec '


w

Fig. 25 Experimental Ape and calculated exit pressure loss, Ap^^^j^,


vs. shear rate for PS (Mobil llOR) at 230®C. (From Ref. 50.)

III. MODELING T W O -D IM E N S IO N A L FLOW

A. General
A realistic simulation of two-dimensional flow requires a constitutive
equation which can represent the fluid behavior in complex flow sit­
uations. Many constitutive equations are available, and a particular
choice depends largely on intended application and the degree of nu­
merical complexities associated with it. This section examines differ­
ent fluid models, ranging from the simple Newtonian to general visco­
elastic, and the computational difficulties associated with them. The
focus is on the abrupt-contraction-flow problem with contraction ratio,
3, in both planar and axisymmetric configurations [see Fig. 2 8(a)].
This problem has become a bench-mark in two-dimensional flow simu­
lation.
Since polymer solutions and polymer melts are usually quite vis­
cous, inertia terms have little or no effect upon the flow field and so
are usually neglected in computations. The governing equations for
steady, inertia-free, incompressible flow with no body force, which
are to be solved with a given constitutive equation, are as follows;
continuity

v -V = 0 ( 20 )
Flow of Polymeric Melts 177

Fig. 26 Experimental [48] (symbols) and predicted (curve) exit pres­


sure loss vs. shear stress for LDPE melt at 150°C. (From Ref. 50.)

Fig. 27 Calculated exit pressure loss vs. shear stress in capillary


and slit rheometer for LDPE melt at 150°C.
178 Isayev and Upadhyay

(a)
t
3 b
i
; _______________________~ ____________________ b

(b)

Fig. 28 (a) Schematic of the abrupt-contraction geometry and (b )


normalized extra length as a function of contraction ratio for New­
tonian fluids in planar and axisymmetric geometries.

momentum

0 = -Vp + A* t ( 21)

constitutive

F ( t , V , VV) = 0 (22)
Flow of Polymeric Melts 179

where V is the velocity vector, p the isotropic part of the stress, x


the extra-stress tensor , and F a functional describing the constitu­
tive equation. The total stress tensor a is

a = - p6 + T (23)

with 6 being the Kronecker delta. The dependent variables to be


solved from these equations, together with appropriate boundary con­
ditions, are V, p, and x. The level of computational difficulty enters
through the complexity of the constitutive equation. Here, we will
classify different fluid models in two broad categories: generalized
Newtonian and viscoelastic.

B. G eneralized Newtonian Fluid Models

The generalized Newtonian class contains all the fluids for which the
deviatonic stress can be expressed as

X = n(Y) e (24)

where e is the rate-of-deformation tensor, n is the viscosity function,


and Y is related to the second invariant of e:

(e : e) (25)

These fluid models do not describe the so-called elastic effects, i.e. ,
the time-dependent growth and relaxation of stresses after imposition
and cessation of strain or strain rate. In other words, the stresses
and strains in the fluid adjust instantaneously, and the fluid does not
exhibit any memory effects. Further, a reversal of the flow direction
does not have any effect on the velocity or stress field.
Numerical methods for calculating the two-dimensional flow of gen­
eralized Newtonian fluids have been well investigated and in most
cases are a simple adaptation of methods developed for Newtonian
flows by incorporating the variable viscosity and using iterative tech­
niques; i . e . , the viscosity is updated from the previous velocity field.
Early contributions to this area include papers by Duda and Vrentas
[59] and Nickell et al. [60]. Rheometrical applications based on finite-
difference calculations may be found in papers by Paddon and Walters
[61] and Williams [62]; Halmos and Boger [63] obtained a solution for
contraction flow that is useful for interpreting experimental observa­
tions. The finite-element technique, which is very convenient for
handling arbitrary geometries, is increasingly being used as a basis
for flow simulations.
18 0 Isayev and Upadhyay

We now discuss some of the viscosity functions which have been


used to solve the entrance-flow problem.

N e w to n ia n M o d e l

no = constant (26)

This is the simplest form of constitutive equation in fluid dynamics


and has been thoroughly investigated. Most numerical schemes for
creeping flow (i .e ., with neglect of inertia) give direct solutions,
and no iterative procedures are necessary. Hence, it is an important
flow to understand^ because it can serve as the first step for solving
flows for higher-level constitutive equations.
One of the quantities of practical use is the total pressure drop
and its dependence on the contraction ratio and contraction angle.
The pressure drop is usually expressed in terms of an entrance
length defined as follows:

Ap - Ap
fd
L = (27)
e A.

where Ap is the total pressure drop over the domain, Apf^ is the fully
developed pressure drop and As is the fully developed pressure gra­
dient in the small channel or tube. Let us denote the half-gap thick­
ness of the small channel by b in the planar case, and the small-tube
radius by R in the axisymmetric case, with an average velocity U.
The fully developed pressure drop calculation is based on the fully
developed velocity profile at every cross section. The fully devel­
oped velocity profile and corresponding fully developed pressure drop
can be written as follows:

planar

U " 2[^ ^ ( b ) ] ’ ~ ^ s [^ s (28)

axisymmetric

8noU (29)

where Xs and Xl are the lengths of the small channel and the large
channel or tube, respectively.
Flow of Polymeric Melts 181

Calculating Ap for the Newtonian case based on a finite-element


formulation for the geometry in Fig. 28(a)^ and using Eqs. (2 7 )-
(29) gives the results for Le/b and Lg/R as functions of different
contraction ratios for the planar and axisymmetric cases shown in
Fig. 28(b). In both cases entrance length increases with 3 and ap­
pears to approach an asymptotic constant value for 3 > 9.
One of the concerns in this type of calculation is the corner re­
gion where the pressure is singular and a very fine mesh is required.
Because of the simplicity of the Newtonian fluids, analytical solutions
can be obtained for the corner region, which can be matched with
FEM or FD schemes as has been done by Shen, Morjaria, and Upadhyay
164], Ladveze and Peyret [65], and Holstein and Paddon [66]. As an
example^ the nature of the pressure singularity for 90° and 135° cor­
ners is shown in Fig. 29. The effect of the corner is localized and
does not influence the overall pressure drop very much.

P o w e r-L a w M odel

n-1
n - mY (30)

where m and n are constants characterizing a particular fluid. This


empirical relation does not describe the viscosity near y = 0, but it
represents the linear shear-thinning region of the log n —versus—log y
curve. It is clearly the first step toward including the shear-thinning
effect and, without doubt, the most well-known and widely used em­
piricism in polymer processing. Clearly, when n = 1 and m = rio> one
obtains the Newtonian fluid. To calculate the entrance length as de­
fined by Eq. (27), we can calculate the fully developed pressure drop
in closed form as follows:
planar

u ^ mri + 2n u y
U 1+n L bL n bj

^Pfd = ^s [^s ^ ,1 + n \ (31)


]
axisymmetric

ri + 3n
U 1+nL^ \r ) ] ’ R L n rJ

APfd = (32)
T T -n \ ]
182 Isayev and Upadhyay

Fig. 29 Pressure distribution near (a ) a 90° corner in a symmetric


geometry and (b ) a 135° corner in a converging flow; planar flow in
both cases.

Figure 30 shows the entrance length nondimensionalized Le as a


function of 1/n for different contraction ratios for the planar and
axisymmetric cases. The plotted results have been obtained by
Hieber [67] and verified with available results [68,69]. Qualita­
tively, the results are similar for the planar and axisymmetric cases
and show that Le is an increasing function of 1/n and contraction
ratio 3- In addition, Le seems to be a linear function of 1/n as
Flow of Polymeric Melts 183

Non-dim ensional normal


distance from wall
1 - 0 .0 (w a ll)
2 - 0.05
3 -0 .1 5
4 - 0.25
5 - Centerline

C a rre a u M odel

n = noti + (33)

where no is the zero-shear-rate viscosity, X is a characteristic time


equal to the reciprocal of the shear rate at which shear thinning
184 Isayev and Upadhyay

Fig. 30 Entrance length as a function of the inverse of power-law


index, 1/n, for a power-law fluid in planar (solid curves) and axi-
symmetric (dashed curves) geometries for different contraction ra ­
tios .

becomes significant, and n - 1 is the power-law slope of the viscosity


at high shear rates. Clearly, this model describes the viscosity be­
havior near Y = 0 and simplifies to a power law at high shear rates.
However, no closed-form solution for fully developed flow and the
pressure gradient can be obtained in this case.
Kim, Brown, and Armstrong [70] have studied axisymmetric flows
according to this model for a 4:1 abrupt contraction (3 = 4). Figure
31 shows their results for the entrance length as a function of power-
law index for different Carreau numbers defined as

Cu = Ay (34)
R

Increasing the Carreau number for a given n increases Le, being


bounded by the power-law value which is reached at Cu ~ 100.
Flow of Polymeric Melts 185

Fig. 31 Le/R in a tube as a function of power-law index, n, for a


Carreau fluid in a 4:1 sudden contraction. Solid curves from Ref.
70 and symbols from Ref. 71.

E llis M odel

1/n - 1
-Ha = 1 + ( J l l ) (35)

The Ellis model is an example in which the viscosity is not an ex­


plicit function of y. That is, the value of n must be obtained by solv­
ing an algebraic equation for a given value of y. In particular, no is
the zero-shear-rate viscosity and Ti / 2 is the value of shear stress at
which n = no/2. At large y the model represents a power-law fluid
with exponent n. A time constant no/'1^1 / 2 can be constructed from
the model parameters, and a nondimensional Ellis number, El, similar
to Cu, can be defined as

El ^ noU (36)
bXi/2
186 Isayev and Upadhyay

Fig. 32 Dependence of Le/b for the Ellis fluid on 1/n and Ellis num­
ber, El, for a planar 4:1 contraction. Results from Ref. 71.

Values of Le for the flow of an Ellis fluid through a sudden axisym-


metric contraction with 3 = 4 have been obtained by Vrentas, Duda,
and Hong [71], as shown in Fig. 32. For a given 1/n, Le increases
with El. This means that Le increases not only with increasing shear
rate but also as the effective range of the Newtonian region is mini­
mized. Comparisons of these Le values with a power-law model are
shown in Fig. 33. Unfortunately, the results for El = 5000 are slightly
higher than the power-law limit. Kim, Brown, and Armstrong [70]
have attributed this increase to the ’’numerical inertia” in the calcu­
lations by Vrentas, Duda, and Hong [71].

C. Viscoelastic Models

Viscoelastic models exhibit normal-stress effects during pure shear


flow and, in general, the stress-strain relation is described by a dif­
ferential or an integral equation. Since the stresses cannot be deter­
mined explicitly in terms of strain rates and substituted into the mo­
mentum equation, numerical methods used for generalized Newtonian
fluids cannot be employed. Although there is no consensus regard­
ing a particular numerical scheme, the following types have been
used. The displacement method [72,73], in which the pressure and
Flow of Polymeric Melts 187

Fig. 33 Comparison of Le/b values for power-law (curve) and Ellis


(symbols) fluid. Results from ( o ) Ref. 69 and ( , □ ), Ref. 71.
a

velocity components are the unknown computational variables and the


stress components are calculated by means of an iterative technique;
the mixed method [74], in which the pressure, velocity components,
and stress components are the unknown computational variables ; the
streamwise-integration schemes [75,76], in which the stresses are de­
termined explicitly by a stream wise integration of the constitutive
equation and are treated as known quantities in the momentum and
continuity equations to determine the velocity and pressure fields by
standard FEM. The subject of numerical schemes for non-Newtonian
fluids has been thoroughly covered by Crochet, Davies, and Walters
[77,78].
At this point some discussion of nondimensional quantities related
to fluid elasticity is required. First, a nondimensional shear rate,
referred to as a Deborah number, is defined as follows :
188 Isayev and Upadhyay

planar

De (37)
fs
axisymmetric

4U ^
De = (38)
R -

where 3U/b and 4U/R are the characteristic shear rates, and is the
zero-shear-rate relaxation time defined as

_ lira T.11 '^22 lim


(39)
o)->0 o)G’^
2YTi 2

where Th - T2 2 is the first-normal-stress difference, T1 2 the shear


stress, Y the shear rate, G^ the storage modulus, G^’ the loss modu­
lus, and 0 ) the frequency. The number De depends only on the zero-
shear-rate properties of the fluid and is zero in generalized Newtonian
fluids.
Elasticity is usually measured in terms of high-elastic deformation
Ye^ also referred to as recoverable strain. For linear viscoelastic
fluids Ye = D e, but for real nonlinear polymeric fluids this relation
is valid only up to Ye ~ 0*4, as shown by Isayev [79], based on the
experimental data for many polymer melts and polymer solutions; at
high values of De, Ye more slowly, never exceeding a value of
approximately 10 for most polymer melts [79]. Clearly, the extent of
high-elastic deformation depends on the relaxation properties of a
particular fluid and thus on the constitutive equation which repre­
sents these properties. However, for practical purposes, once the
relaxation properties of the fluid are described by a constitutive
equation, the task is to be able to calculate flows up to large non-
dimensional shear rates, De, encountered in real flow situations such
as injection molding.
The other number frequently used to represent fluid elasticity is
the Weissenberg number. We, defined as

We = (40)
2t12
-

At small De, We, Ye> and De are all equivalent. However, at high De
these measures may be quite different.
We now consider some of the viscoelastic models applied to the en­
trance-flow problem, particularly the planar case.
Flow of Polymeric Melts 189

S e c o n d - O r d e r -F lu id M odel

2rioe - Vie + 4va e*e (41)

where rioj ^ 2 are the zero-shear-rate viscosity, and first- and

second-normal-stress coefficients. The superscript v represents the


contravariant derivative of tensor A, defined [80,81] as

V DA . .T
A - Y -A - A -V (42)

where D/Dt is the substantial time derivative.


The second-order fluid is the simplest form for expressing visco­
elastic behavior just as the power-law fluid represents the simplest
shear-thinning behavior. This fluid may be considered as the limit
of a simple fluid for small natural times, and it can be used to test
the numerical schemes for more complex constitutive equations. This
has been done by Mendelson et al. [82], who studied a 4:1 planar
contraction problem and observed that the flow field was essentially
unchanged from De - 0 to De = 0.5, but that the corner vortex started
to grow and intensify above De = 0.5, as shown in Fig. 34. In addi­
tion, a smaller recirculation appears near the edge of the contraction
at De = 1.0 and continues to grow with increasing De. Since the
finite-element calculations did not converge above De = 2.3, how­
ever, some of the results may not be physical, and the vortex growth
is perhaps a numerical artifact caused by excessive approximation
error in the finite-element interpolation of the stress field , which is
shown in Fig. 35. On the other hand, the authors contend that the
flow could be extended up to arbitrarily large De if sufficiently re­
fined meshes were used. On the contrary. Tanner [83], using a non­
linear stability analysis has suggested that attempts to refine the so­
lution for a second-order fluid by using a finer mesh will often de­
crease convergence such that no solution may be attainable in the
limit h 0 (h denoting the mesh siz e ).

U p p e r - C o n v e c t e d M a x w e ll M o d e l

T + Xt = 2rioe (43)

where no and X are the viscosity and the relaxation time, respectively,
and ^denotes the upper convected derivative defined in Eq. (42).
This simple differential model has dominated recent developments in
numerical non-Newtonian fluid mechanics. Although this model is
viscoelastic, it fails to give a shear-thinning viscosity or a shear-rate-
dependent, first-normal-stress coefficient observed in most polymeric
190 Isayev and Upadhyay

Fig. 34 Streamlines for flow of a second-order fluid through a 4:1


planar contraction for De = 0 (a ) , 1.0 (b)^ and 2.3 (c ). Results from
Ref. 82.

fluids. A more serious defect in the model, pointed out by Mendelson


et al. [82], is that the elongational viscosity becomes infinite at finite
strain rates. Further, according to this model. Ye = We = De for all
strain rates such that the recoverable strain is overpredicted at high
strain rates. Hence, the usefulness of this model to represent real
polymer behavior is limited to small De.
Flow of Polymerie Melts 191

2k10'

The Maxwell model is a special case of Oldroyd^s eight-constant


model [80,81,84]
V
T + XiT + iJo(tr T)e - Ui(T*e + e - T ) + V i ( T : e ) 6

2rio(ö 2y2e*e) + 2v2(e:e)6 (44)

There are several restrictions on the constants in Eq. (44) that have
to be imposed so that the derived results are not aphysical. Many
simplifications and special cases of this model have been used [85—89],
but all of them cannot be discussed here. A complete discussion can
be found in the book by Bird, Armstrong, and Hassager [89] . The
Maxwell model can be solved by successive substitution to give an ex­
plicit expression for t in terms of Y and its convected derivatives. A
second-order approximation obtained in this manner is identical to the
second-order fluid if we take V2 = 0 in Eq. (41).
The Maxwell model can also be represented in terms of an integral,
memory-type form, which is explicit in the stress tensor:

-s/X
f]e
^(s) ds (45)
* 0

where is the Finger strain. In this form the model is a special


case of Lodgers rubberlike elastic liquid [90].
192 Isayev and Upadhyay

DOWNSTREAM DISTANCE , x

Fig. 35 Calculated normal-stress - Qxx a function of position along


a line tangent to the downstream wall of the small slit in 4:1 planar
abrupt-contraction flow of a second-order fluid. Entrance to small
slit corresponds to x = 0.0, and x is positive in downstream direction.
The oscillations in the stress field get progressively worse with in­
creasing De and propagate upstream for De > 2.0. Results from Ref.
82.

Viryayuthakorn and Caswell [75] have used Eq. (45) to calculate


the stresses by numerically approximating the memory integral and
computing velocity and pressure fields from the momentum equations
in which the stresses enter as body forces. They found that such a
scheme converged at higher De than previously possible. In particu­
lar, they studied a 4:1 axisymmetric contraction- and expansion-flow
problem. Figure 36(a), (b ) shows the centerline velocity for contrac­
tion and expansion flows at De = 2 with corresponding Newtonian
curves. For the contraction an overshoot of about 3% occurs at about
one radius downstream from the entrance. This is evidently a result
of the elastic behavior of the fluid. When the flow is reversed, the
maximum shifts toward the entrance plane. In contrast, no overshoot
and effect of flow reversal occurs for the Newtonian fluid. Figure
36(c) shows the normal stress invariant , tr t- 3x00, along the cen­
terline for contraction and expansion flows for De = 0 and 2. The
Flow of Polymeric Melts 193

latter results again illustrate the irreversibility of the flow when elas­
tic effects are included. Although the computations were not success­
ful beyond E)e = 3, which has usually been attributed to extremely
large approximation errors in the corner region, this problem might
be tractable if a local analytical solution near the corner (as available
for Newtonian fluid) could be obtained and matched with the FEM
calculation. Other investigators would attribute the breakdown to
a bifurcation phenomenon [82] or to the possibility of an inherent
instability in the Maxwell-type models [76]. Recent work by Crochet
and Keunings [91] has shown that using an Oldroyd-B fluid that
contains a retardation time increases the upper limit of We by a fac­
tor of 4 in computing die-swell phenomenon as compared to the Max­
well fluid. Evidently, a great deal more work is needed to clarify
some of these questions.

Phan T h ie n — T a n n e r M od el

Y ( t) t + 1 ^ ^ ] = 2rio (46)

with

Y ( t ) = 1 + — - tr T (47)
rio

where £ is a small parameter and, in a generalized form,

Y ( t ) = e x p (^ ^ tr T) (48)
\ Ho /

The superscript a represents the covariant derivative [counterpart


of contravariant form given by Eq. (42)] of a tensor A, defined as

A DA T T
A = ^ + V - A + A *V (49)

This constitutive equation was developed by Phan Thien and Tanner


[92,93]. Simultaneously, Johnson and Segalman [94] proposed a con­
stitutive equation which is the same as Eq. (46) with £ = 0, allowing
nonaffine deformations of the molecular network in its single-relaxa­
tion-time mode. The scalar parameter Ç in Eq. (46) governs the ratio
of the first- to the second-normal-stress difference in viscometric
flows. In this model the value of £ controls the elongational viscos­
ity, and Ç controls the shear viscosity; when both are zero, the Max­
well model is recovered.
194 Isayev and Upadhyay

Fig. 36 Centerline axial velocity distribution of (a ) die entry and (b )


die exit flow and (c ) normal-stress invariant, tr a - 3aee, along the
centerline for entry and exit flow of Maxwell fluid at De = 0 and 2.
Results from Ref. 75.

Some of the computational results obtained by Keunings and Crochet


[95] using the Phan Thien-Tanner model are shown in Figs. 37 and 38.
Figure 37 shows the development of the axial velocity along the center-
line for different values of XiU/2R. Note that a maximum in the ve­
locity occurs within about a half-radius downstream of the contraction.
Flow of Polymeric Melts 195

The large overshoot at high flow rates is evidently spurious. Note


that the fully developed downstream value increases above a certain
flow rate. Such a behavior is due to the retardation time in the con­
stitutive equation, which admits a Newtonian behavior at large flow
rates. Figure 38 shows the entrance length [defined by Eq. (27)]
for the Phan Thien-Tanner fluid (e = 0.015), the Johnson-Segalman
fluid (e = 0, and the Maxwell fluid (e = 0 and 5 = 0). We see that
Le decreases up to AiU/2R = 0.2 and then increases monotonically for
the first two fluids, but decreases for the Maxwell fluid. However,
in light of the extremely large overshoots in velocity (up to twice the
fully developed downstream valu e), which have not been observed ex­
perimentally, the physical significance of these results is unclear.

L e o n o v M odel

The Leonov model is a differential-type model proposed by Leonov


[58,96], which is based on irreversible thermodynamics and makes
use of the classical potential function from the network theory of
elasticity. This model has been developed to describe the flow be­
havior of polymeric fluids over a wide range of elastic strains, with
the governing equations being as follows :

N
T = 2rioSe + 2 I (W
Ik k k ,2 )
k = l^

(50)
k k
196 Isayev and Upadhyay

Fig. 37 Development of centerline velocity for various values of We in


4:1 axisymmetric contraction for Phan Thien—Tanner fluid. Results
from Ref. 95.

det C = 1
k

k 2y.

8W
(50)
k >3

where x, e, 6, and tiq defined as usual, e^ is the irreversible


strain tensor, the elastic potential, the elastic strain tensor
Flow of Polymeric Melts 197

Fig. 38 Le/R as function of We for Phan Thien—Tanner fluid (e


0.015), the Johnson-Segalman fluid (e = 0), and the Maxwell fluid.
Results from Ref. 95.

(Finger measure) in the kth relaxation mode, and and Ik ^ 2


its basic invariants. Further, 03 is the vorticity tensor, ( )^ rep­
resents the Jaumann tensor derivative with respect to time, nk
0k are the shear viscosity and the relaxation time of the kth mode,
]ik “ ^k/20k is the modulus of the kth mode, and s is a rheological
parameter between 0 and 1.
Further , it is assumed that the classical potential function of the
network theory of elasticity is applicable for viscoelastic fluids, such
that the elastic potential Wk can be taken as

W = u (51)
’k ^k ('‘k , r ^ )

Therefore the governing equation for Ck in Eq. (50) can be written


as

c
S
^ + 26,
— (52)
198 Isayev and Upadhyay

with

N n,
T = 2rioSe + 2 C (53)
k=l \ ^

This model has been used in many investigations [14—16,97—104]


and has good predictive capability and is quite readily amenable to
numerical implementation. The viscometric functions of this model
can be written as follows [96]:

N
n (Y ) = Tios + 2
NL (54)
k = i' " \

N /2(X, - 1)
'^22 — S (55)
k=l /I + K
k

where X]^ = / l + 4Y^0k^.


Upadhyay and Isayev [76] have recently used this model to solve
a 2:1 planar contraction problem for a polyisobutylene Vistanex em­
ploying a streamwise-integration procedure. The parameters of a
two-mode model were evaluated from a least-square fit of the visco­
metric functions with the experimental data. The resulting fit is
shown in Fig. 39, and the corresponding parameters are s = 0.01,
= 3.58 X 10\ 2,95 x lOVPavsee, 0k = 6.07, 0.47 sec. The char­
acteristic relaxation time (^ ) of this material, based on Eq. (39), has
been calculated to be 3.51 sec. Further, Fig. 40 shows results for
We versus De. The results are based on the present two-mode Leonov
model and, for comparison, corresponding results based on the Max­
well model with the same 6. For the Leonov fluid We first increases
with De and then decreases after attaining a maximum value of about
4.2. This decrease occurs because the two-mode viscosity eventually
reaches a constant value at high shear rates. This second-Newtonian
region and corresponding decreasing We could be pushed out further
if the underlying data were available to higher shear rates and more
modes were employed. For the present fit, however, it seems that
the real material behavior can be modeled only up to De ~ 250, cor­
responding to the maximum We in Fig. 40.
Based on the two-mode Leonov model and a finite-element code
which incorporates a streamwise-integration procedure [76], results
are shown in Fig. 41 for the calculated centerline velocity in a planar
Flow of Polymeric Melts 199

Fig. 39 Apparent viscosity n and first normal-stress difference Ni


vs. shear rate y for polyisobutylene Vistanex LM-MH (Enjay) at 27°C.
Symbols: experiment; curves: theoretical fit based on two-mode
Leonov model. (From Ref. 76.)

2:1 contraction for various values of the Deborah number. In par­


ticular, these results clearly demonstrate the shear-thinning effect
as evidenced by a decreasing fully developed (upstream and down­
stream) velocity with increasing De (or flow ra te ). Further, the elas­
tic effects are evidenced by the velocity overshoot, particularly at
the highest De. In addition. Figs. 42 and 43 show the nondimen-
sional shear stress t^2 and the first-normal-stress-difference ( th -
T2 2 ) profiles, respectively, at different cross sections (x/b = -8.0,
-2.0, -1.0, -0.5, 0.0, 0.5, 2.0, and 10.0) corresponding to De =
0, 5, and 50. The quantity x is measured from the start of the small
channel, and thus the positive values of x correspond to positions
along the small channel, negative values along the large channel.
Note from Figs. 42 and 43 at x/b = -2.0 and -1.0 that the recircula­
tion starts further upstream as De increases. The size of the recir­
culating region as a function of De is shown in Fig. 44 and indicates
a substantial increase with De. The lines defining the extent of the
recirculating zone have been obtained on the basis of zero net cross
flow. On the other hand. Fig. 45 shows the nondimensional normal-
200 Isayev and Upad hy ay

log
(f«)
Fig. 40 Weissenberg number, We, as a function of nondimensional
average shear rate De for the polyisobutylene Vistanex described by
a two-mode Leonov model. (From Ref. 76.)

stress difference, ( t^x " T2 2 ) / (noU/B) , along the centerline, which


passes through a maximum near the entrance plane and thereafter
relaxes in the small channel. The position of this maximum and the
extent of relaxation is pushed farther into the small channel as De
is increased.
The computations were also carried out for higher values of D e,
and the numerical scheme had no difficulty in converging at any De.
However, the maximum Weissenberg number for the parameters used
was approximately 4.2 (see Fig. 40) with a corresponding value of
De « 270 and a wall shear rate of about 120 sec“^. This shear rate
represents the limit of experimental data used to evaluate the model
parameters (see Fig. 39). The range of shear rate, however, is
Flow of Polymeric Melts 201

x/b

Fig. 41 Predicted nondimensional velocity u/U along centerline of 2:1


planar contraction for De = 0, 5, and 50; based on two-mode Leonov
fit given in text. (From Ref. 76.)

sufficiently high to fall in the range of most experiments and thus fa­
cilitate experimental comparison in realistic flows.
Further, Upadhyay and Isayev [76] adapted the numerical scheme
to compute flows with an upper - convected Maxwell model. In particu­
lar, Fig. 46 presents results for the normalized centerline velocity
and first-normal-stress difference for the 2:1 planar contraction with
De = We = 0 and 3. In Fig. 46(a) the fully developed upstream and
downstream centerline velocities for both values of De is 1.5 times the
corresponding average velocities, indicating the absence of shear thin­
ning. The model does predict an overshoot in the centerline velocity
for De = 3, as expected for viscoelastic fluids. Further, the trend
between De = 0 and De = 3 in Fig. 46(b) is similar to that for the
Leonov fluid in Fig. 41. However, efforts to extend the preceding
202 Isayev and Upadhyay

x/b = - 8.0 - 2.0 - 1.0

0.8 - 0.8 0.0 - 0.8

^ly^pU

0.0 0.5

-2

2.0 10.0
De = 0
5
50

Fig. H2 Predicted distribution of nondimensional shear stress T1 2 /


(ngU/b) at cross sections corresponding to x/b = -8.0, -2.0, -1.0,
-0.5, 0.0, 0.5, 2.0, and 10.0 for De = 0, 5 and 50; same conditions
as in Fig. 41. (From Ref. 76.)

Maxwellian results to higher De were not successful because the nu­


merics fail to converge. The reasons for this failure, which has oc­
curred in almost all published work for this type of problem, cannot
be determined unambiguously from these limited results and further
x/b = 8.0 - 1.0

J
0.8 -

0.0

- 1

(T1 1 -T2 2 )/-


2.0

(tii-T22)/:!^
b
10.0

De = 0
/. 5
// 50

(ti,-T22)/5^

Fig. 43 Predicted distribution of nondimensional first normal-stress


difference ( th - T2 2 )/(noU/b) at cross sections corresponding to x/b =
-8.0, -2.0, -1.0, -0.5, 0.0, 0.5, 2.0, and 10.0 for De = 0, 5, and 50;
same conditions as in Fig. 41. (Note: Fully developed values for
'^11 ~ '^22 identically zero in De = 0 case.) (From Ref. 76.)
204 Isayev and Upadhyay

Fig. 44 Predicted recirculating region for De = 0, 5, and 50; same


conditions as in Fig. 41. (From Ref. 76.)

Fig. 45 Predicted nondimensional normal-stress difference along the


centerline for different De; same conditions as in Fig. 41. (From Ref.
76.)
Fig. 46 Predicted nondimensional centerline velocity distribution (a)
and normal-stress difference (b ) corresponding to De = 0 and 3 for
Maxwell fluid. (From Ref. 76.)
206 Isayev and Upadhyay

work is required. However, the Leonov fluid does not exhibit any nu­
merical instabilities under comparable elasticity and mesh configura­
tion. This would indicate that the nature of the fluid as described by
the constitutive equation has some influence on the numerical stabil­
ity. A complete understanding of this influence can only be achieved
via a stability analysis using different constitutive equations.

IV . P R E D IC TE D R ESU LTS COMPARED


W ITH E X P E R IM E N TA L RESULTS

The last section showed that a great deal of work has been done in
simulating two-dimensional viscoelastic flows, although little direct
comparison with experiment was made. The most extensive compari­
son available in the literature seems to have been done for the hole-
pressure flow problem [105]. Although generalized Newtonian mod­
els have been found to predict the measured pressured drops fairly
well in some cases, they cannot even qualitatively describe some of
the viscoelastic effects observed in the experiments (see Section I I).
Most viscoelastic models used are not able to describe real polymers
and, at this time, are restricted to low strain rates. However, even
if the numerical problems were completely solved, the results would
be only of academic interest unless more suitable models were used.
Isayev and Upadhyay [17] have tried to address these problems.
They have extensively compared theoretical predictions for birefrin­
gence, based on a three-mode Leonov model with the experimental
data reported in Section II for planar contraction and expansion flows
(see Fig. 5 for the flow geometry). The stresses in the flow (corre­
lated with birefringence via stress-optical law) are the least accurate
quantities in the computations [76], such that a comparison on the
basis of birefringence represents a stringent test for the model.
In particular, the parameters of the Leonov model for polyisobutyl­
ene Vistanex were determined to be s = 0.006, nk = 3.20 x lo**, 1.80 x
10^, 1.58 X 10** Pa-sec, and = 7.025, 1.553, 0.182 sec. This fit de­
scribes the data [17] up to about 250 sec"^. Further, we can get an
idea of the elasticity in the flow from Fig. 47, which shows We as a
function of the characteristic shear rate, U/b, for these parametric
values. The value of We increases with U/b and attains a maximum
value of approximately 4.9, corresponding to U/b ~ 70 sec"^, and
then decreases. Thus the real material behavior can be modeled up
to U/b « 70 s e c '^
Figures 48-50 show the gapwise birefringence distribution at dif­
ferent cross sections, x/b (b is the half-gap thickness), from the en­
trance to the small channel at various flow rates for working cells A,
Flow of Polymeric Melts 207

log U/b, sec -1

Fig. 47 Weissenberg number, We = ( i n - T22)/2t i 2» as a function of


U/b for poly isobutylene Vistanex using a three-mode fit. (From Ref.
17.)

D, and E (see Fig. 5), respectively, with the flow directed from the
large channel to the small one. Figure 48(a) has only one cross-sec­
tional location because of limited optical resolution. The birefrin­
gence scale is the same for all the curves in a given figure; the sym­
bols indicate the experimental data, and the solid curves are the cor­
responding theoretical predictions. Although the birefringence be­
havior is quite different in the various cells, the predictions describe
the data quite well in all cases. Similar comparisons for the diverging
and abrupt-expansion flows are presented in Figs. 51—54, with the
predictions again agreeing reasonably well with the measurements.
Figures 55 and 56 show the corresponding centerline birefringence
as a function of dimensionless length x/b for the converging and di­
verging flows (Fig. 55, corresponding to cells A, B, and C ) and for
the abrupt-contraction and abrupt-expansion flows (Fig. 56, corres­
ponding to cells D and E ). Each curve in these figures corresponds
to a specific flow rate, Q; x is measured from the start of the small
208 Isayev and Upadhyay

CELLA
CONVERGING
(a)
3r ^ x/b = -5.34
2 yP Q = 0.0148 cm^/sec
y/b

:[/
(b )

Q = 0.0401 cm^/sec

y/b
x/b = -3.37

Fig. 48 Gapwise birefringence distribution at different cross sections


for converging flow at different flow rates in cell A. Symbols: ex­
periment; curves: theory. (From Ref. 17.)
Flow of Polymeric Melts 209

CELL D
CONTRACTION

Fig. ^9 Gapwise birefringence distribution at different cross sections


for contraction flow at different flow rates in cell D . Symbols: ex­
periment; curves: theory. (From Ref. 17.)

channel, with positive values corresponding to positions along the


small channel. In the converging and abrupt-contraction cases, the
birefringence starts to build up in the large channel, continues to in­
crease in the converging part for the converging cases and in the
preentrance region for the abrupt-contraction cases, and then passes
through a maximum very near the entrance to the small channel ( i . e . ,
x/b = 0). Thereafter, the relaxation of birefringence occurs in the
small channel. Depending on the flow rate, complete relaxation may
or may not occur within the length of the small channel. In the di­
verging and abrupt-expansion cases the birefringence along the cen­
terline seems to start building up just before the exit from the small
channel, passes through a maximum after the exit, and then relaxes,
respectively, in the diverging part and the large channel. The posi­
tion of the maximum is pushed farther downstream, and the extent of
210 Isayev aiid Upadhyay

CELL E
CONTRACTION

(a)
Q = 0.0956 cm^/sec

y/b

0.298 x/b = 1.846

Q = 0.0238 cm^/sec
(b)

2.386

(c) Q = 0.0716 cm^/sec

An X 10'*

Fig. 50 Gapwise birefringence distribution at different cross sections


for contraction flow at different flow rates in cell E. Symbols: ex­
periment; curves; theory. (From Ref. 17.)
Flow of Polymeric Melts 211

C ELLA
DIVERGING

Q = 0.0233 cm^/sec

x/b = - 0 . 9

(b)

Q = 0.0452 cm^/sec
-1 7 .4

x/b = - 2 .4
JLLL.

(C)
Q = 0.0878 cm3/sec
- 22.2
-1 2 .4

An X 10^

Fig. 51 Gapwise birefringence distribution at different cross sections


for diverging flow at different flow rates in cell A. Symbols: ex­
periment; curves: theory. (From Ref. 17.)
212 Isayev and Upadhyay

CELL B DIVERGING

(a)
-18 .1 6 - 11.2
Q = 0.0912 Cm^/Sec

~ = -4 .3 4
b

-8 .4 4 Q = 0.1445 Cm^/Sec

Fig. 52 Gapwise birefringence distribution at different cross sections


for diverging flow at different flow rates in cell B. Symbols: experi­
ment; curves; theory. (From Ref. 17.)

channel length needed for relaxation increases with flow rate. In all
cases the agreement between the predictions and measurements is rea­
sonable .
To help us compare the birefringence data from different cells,
Fig. 57 shows Anjnax ^ function of average shear rate, U/b (U is
the average velocity in the small channel), for all the cells. Com­
paring the abrupt-contraction and abrupt-expansion cases (cells D
and E ), we see that the birefringence maximum is significantly in­
fluenced by the contraction ratio. For fixed U /b, An^ax is higher
Flow of Polymeric Melts 213

CELL D
EXPANSION

y/b

Fig. 53 Gapwise birefringence distribution at different cross sections


for expansion flow at different flow rates in cell D. Symbols: ex­
periment; curves: theory. (From Ref. 17.)
214 Isayev and Upadhyay

CELL E
EXPANSION
(a)

-2 .7 4 x/b = -2 .3 2

Q = 0.0104 cm^/sec
y/b

- 6.8 - 4.26 ' - 2.94


o

y/b

2 -

(c) Q = 0.0538 cm3/sec

/-1 8 .1 4J - “

4 -i
y/b

Fig. 54 Gapwise birefringence distribution at different cross sections


for expansion flow at different flow rates in cell E. Symbols: ex­
periment; curves: theory. (From Ref. 17.)
o
CELL B CELL C

(a) (a) “D

3
a>

a>

(b)
^nX10*

Fig. 55 Birefringence along centerline for converging (a ) and diverging (b ) flow at different flow
rates in cells A , B, and C. Symbols: experiment; curves: theory. (From Ref. 17.)
216 Isayev and Upadhyay

CELL D

(a) C ELL E

(b) (b)

Fig. 56 Birefringence along centerline for contraction (a ) and expan­


sion (b ) flow at different flow rates in cells D and E. Symbols: ex­
periment; curves: theory. (From Ref. 17.)

for the larger contraction ratio (cell E ) , and this difference increases
with increasing U/b. Further^ the effect of contraction angle or ex­
pansion angle, a, on An^^ax demonstrated in Fig. 58, which shows
Anmax versus a for two different values of U /b. Only the test cells
A , D, and E, which have approximately the same contraction ratio,
are used in Fig. 58. The birefringence maximum increases mono-
tonically with a, with the slope of the curves being larger at the
smaller angle.
Flow of Polymeric Melts 217

• CONVERGING
O DIVERGING

-1 0
log U/b, sec’^
Fig. 57 Birefringence maximum along centerline vs. average shear
rate, U/b, for converging and diverging flows (cells A, B, and C)
and abrupt-contraction and -expansion flows (cells D and E ). Sym­
bols: experiment; curves: theory. (From Ref. 17.)
218 Isayev and Upadhyay

• CONVERGING

**U/b = 6 see“ “*

^ *}u/b = 2 sec-''

a, degrees

Fig. 58 Maximum centerline birefringence as a function of contrac­


tion angle, for U/b = 2 and 6 sec"^. Symbols: experiment; curves:
theory. (From Ref. 17.)

Figure 59 shows the dependence of volumetric flow rate Q on the


pressure drop Ap for the converging and abrupt-contraction flows
and also for the diverging and abrupt-expansion flows. The rever­
sal in flow direction influences the Q-versus-Ap curves in all test
cells. The most pronounced difference appears in cell E, corres­
ponding to abrupt-contraction and abrupt-expansion flows with a
large contraction ratio of 6 . 8 . In cell D, which has a contraction
ratio of 2 . 0 , only a slight difference appears at the higher flow rates.
A major conclusion from Fig. 59 is that the converging and abrupt-
contraction flows of polymer melts typically require a smaller pres­
sure drop than the diverging and abrupt-expansion flows in order
to maintain the same volumetric flow rate. Theory clearly predicts
this observation^ although the magnitude of the pressure difference
between the converging and diverging cases is not well predicted.
In general, the foregoing comparisons are encouraging and sug­
gest the desirability of further investigations, such as for axisym-
metric flows and more general geometric configurations.

V. IM PLICATIONS FOR JUNCTURE-TYPE


FLOW IN MOLD DESIGN

Currently, cavity-filling simulations typically assume that the pres­


sure losses due to juncture regions are negligible. This assumption
• CONVERGING
O DIVERGING
CELL
A

0.0 r

o - 1.0
0

>
E
o

D)
O
- 2.0

-3.0
6.0 6.5 7.0 5.5 6.0 6.5 6.0 6.5 7.0 5.5 6.0 6.5

log A p, Pa

Fig. 59 Pressure drop vs. volumetric flow rate for test cells A, B , C, B , and E in converging and diverging
flow cases. Symbols: experiment; curves: theory. (From Ref. 17.)
220 Isayev and Upadhyay

is reasonable as a first approximation and is generally valid for flat


cavities or cavities with gradual variation of cross section in the flow
direction. However, in the presence of an abrupt contraction and
abrupt expansion or ribs and bosses, the relative contribution of such
juncture regions to the overall pressure drop in the cavity may b e ­
come significant. Based on experimental results, we have shown in
the preceding sections that pressure losses due to entry- and exit-
type flows can be quite high, especially at high strain rates typical
of injection molding, where the juncture loss for typical polymers can
be about 1 0 0 bar.
A region of particular concern is in the vicinity of the gate, where,
due to a small cross-sectional area, the local rate of deformation is ex­
tremely high, but, since the gates are relatively short, the flow is far
from fully developed. Further complicating matters is the possibility
of significant viscous heating due to associated high strain rates.
Hence, both memory and nonisothermal effects can be expected to be
important in the gate region.
On the other hand, the state of the art in simulating the two-di­
mensional flow of polymeric melts presented in the previous section
shows that, despite significant achievements, such simulations are
still in an initial stage of development even for isothermal flow. In
particular, in many cases, available simulation approaches are un­
able to overcome computational difficulties arising when treating vis­
coelastic flows at high strain rates. Some success has been achieved
when the Leonov constitutive equation has been employed. In p ar­
ticular, a good capability has been demonstrated for predicting the
stress distribution in juncture areas. However, the model has not pre­
dicted quantitatively the additional pressure drop arising in expan­
sion-type flows of viscoelastic melts compared with contraction flows,
although the simulations do show a trend in the right direction.
A possible solution for handling the extra pressure loss arising in
the juncture region is to undertake a semiempirical approach based
on determining the extra pressure losses in terms of master curves
similar to that in Fig. 23. For example, the strain rate in the gate
could be calculated by using the assumption of fully developed iso­
thermal flow, and the extra pressure loss could be determined from
the master curve constructed for the particular juncture geometry.
Such pressure loss could be added to the total pressure when the
melt front reaches the particular location of the abrupt contraction
or abrupt expansion in the gate area. Kwon et al. [44] successfully
used a similar approach.
If cavities have ribs and bosses, extra pressure losses due to the
melt entry into these areas can alter the melt-front advancement in
the cavity. Therefore if one neglects this additional resistance in
Flow of Polymeric Melts 221

the computation, the melt-front location at any instant during filling


of the cavity will be predicted incorrectly, the extent of the inaccu­
racy depending on the level of the extra pressure losses and the
number of ribs and bosses. As a first approximation, one could im­
prove this situation by assuming that the melt enters the rib or boss
when the pressure generated at its entry reaches a level sufficient to
overcome the extra pressure loss. This pressure could be taken to
be equal to the extra pressure loss in the particular rib or boss at
shear rates corresponding to channel flow of the polymeric melt un­
der isothermal conditions. This pressure could be determined from
master curves for the extra pressure loss similar to those in Fig. 23.
Such curves would have to be generated for particular geometric con­
figurations .
The present results for simulating the two-dimensional flow of vis­
coelastic fluids in juncture regions can be viewed as an important
step toward the eventual goal of improved mold design. At present,
available cavity-filling simulations are based on inelastic fluid mod­
els. Extensions of such simulations to handle viscoelastic fluids will
be the necessary next logical step. In addition to the usual informa­
tion generated by inelastic cavity-filling modeling, viscoelastic simu­
lations will give important information concerning flow stress develop­
ment. An area of specific interest concerns the juncture region in the
mold, which, as the preceding sections have shown, is an area of
high stress level. However, to assess these stresses quantitatively,
we must extend the two-dimensional viscoelastic simulation to incor­
porate nonisot her mal flow and more general geometric configurations.

ACKNOWLEDGMENTS

The authors would like to thank Drs. C. A. Hieber and D. C. Bogue


for their review of the manuscript and useful comments.

REFERENCES

1. Perry's Chemical Engineers Handbook, R. H. Perry, D. W. Green,


and J. O. Maloney, Eds., McGaw-Hill, 6 th e d ., 1984.
2. C. D. Han, Rheology in Polymer Processing, Academic Press,
New York, 1976.
H. Janeschitz-Kriegl, Polymer Melt Rheology and Flow Birefrin­
gence, Springer-Verlag, Berlin, 1983.
4. A. S. Lodge, Nature, 176, 838 (1955).
5. W. Philippoff, Nature, 178, 811 (1956).
222 Isayev and Upadhyay

6. H. Janeschitz-Kriegl, Adv. Polym. Sci,, 6 , 170 (1969).


7. D. C. Bogue and F. N. Peeples, Trans, Soc, RheoL , 6 , 317
(1962).
8 . E. B. Adams, J. C. Whitehead, and D. C. Bogue, A A .C h .E ,
J,, 11, 1026 (1965).
9. T. R. Fields and D. C. Bogue, Trans. Soc. RheoL, 12, 39
(1968).
10. C. D. Han and L. H. Drexler, J. Appl. Polym. Sci., 17, 2329
(1973).
11. H. Y. Yoo and C. D. Han, J. RheoL, 25, 115 (1981).
12. V. I. Brizitsky, G. V. Vinogradov, A. I. Isayev, and Y. Y.
Podolsky. J. Appl. Polym. Sci., 22, 751 (1978).
13. J. D. Ferry, Viscoelastic Properties of Polymers, 3rd e d ., Wiley,
New York, 1980, p. 352.
14. A. I. Isayev, R. K. Upadhyay, and S. F. Shen, SPE Tech.
Papers, 28, 298 (1982).
15. R. K. Upadhyay and A. I. Isayev, SPE Tech. Papers, 29, 714
(1983).
16. A. I. Isayev and R. K. Upadhyay, in Advances in Rheology,
Eds. B. Mena, A. Garsia-Rejan and C. Rangel-Nefaile, Univer­
sity of Mexico, V . 3, 309, 1984.
17. A. I. Isayev and R. K. Upadhyay, J. Non-Newt. Fluid Mech.,
19, 135 (1985).
18. J. L. White and A. Kondo, J. Non-Newt. Fluid Mech., 3, 41 (1978).
19. D. V. Roger, Adv. Transport Processes, 2, 43 (1982).
20. E. B. Bagley, J. AppZ. P hys., 28, 624 (1957).
21. E, B. Bagley, Trans. Soc. RheoL, 5, 355 (1961).
22. W. Philippoff and F. N. Gaskins, Trans. Soc. RheoL, 2, 263
(1958).
23. G, Barr, Viscometry, Oxford University Press, London, 1931.
24. J. Schurz, RheoL Acta, 4, 107 (1965).
25. L. L. Sulzhenko and E. V. Kuvshinsky, Polym. Sci. USSR, llA ,
2363 (1969).
26. D. R. Oliver, W. Macsporran, and B. M. Hiorns, J. Appl.
Polym. Sci., 14, 1277 (1970).
27. T. Araiand H. Aoyama, Trans. Soc. RheoL, 7, 333 (1963).
28. A. Ram and M. Narkis, J. AppZ. Polym. Sci., 10, 361 (1966).
29. T. Hayahara and J. Takao, J. Appl. Polym. Sci., 11, 735 (1967).
30. B. Philipp and K. Wulf, RheoL Acta, 5, 93 (1966).
31. H. L. Weissberg, Phys. Fluids, 5, 1033 (1962).
32. H. L. La Nieve and D. C. Bogue, J. Appl. Polym. Sci., 12, 353
(1968).
33. J. S. Chong, E. B. Christiansen, and A. D. Baer, J. Appl.
Polym. Sci., 15, 369 (1971).
Flow of Polymeric Melts 223

34. V. Z. Volkov, V. D. Fikhman, and G. V. Vinogradov, J. Engg,


Phys., 31, 1464 (1975).
35. D. B. Holmes, Dissertation, Delft, 1967.
36. E. B. Christiansen and M. E. Lemmon, A .i.C h .E . J., 11, 995
(1965).
37. M. Collins and W. R. Schowalter, Phys. Fluids, 5, 1122 (1962).
38. E. M. Sparrow and S. H. Lin, Phys. Fluids, 7, 338 (1964).
39. Y. L. Wang and P. A. Longwell, A .i.Ch.E . J., 10, 323 (1964).
40. N. D. Sylvester and S. L. Rosen, A .i.C h .E . J., 16, 964
(1970).
41. N. D. Sylvester and S. L. Rosen, A .i.C h .E . J., 16, 967 (1970).
42. R. L. Boles, H. L. Davis, and D. C. Bogue, Polym. Eng. Sci.,
10, 24 (1970).
43. F. N. Cogswell, Polym. Eng. Sci., 12, 64 (1972).
44. T. H. Kwon, D. Chu, and K. K. Wang, SPE. Tech. Papers, 31,
809 (1985).
45. T, H. Kwon, S. F. Shen, and K. K. Wang, Polym. Eng. Sci.,
26, 214 (1986).
46. D. V. Boger and M. M. Denn, J. Non-Newt. Fluid Mech., 6, 163
(1980) .
47. L. Choplin and P. J. Carreau, J. Non-Newt. Fluid Mech., 9, 119
(1981) .
48. H. M. Laun, Rheol. Acta, 22, 171 (1983).
49. F. P. La Mantia, A. Valenza, and D. Acierno, Rheol. Acta, 22,
299(1983).
50. A. I. Isayev and B . Chung, Polym. Eng. Sci., 25, 264 (1985).
51. M. Yamada and R. S. Porter, J. AppL Polym. Sci., 13, 1711
(1974).
52. R. J. Crpwson and M. J. Folkes, Polym. Eng. Sci., 20, 934
(1980).
53. L. Czarnecki and J. L. White, J. Appl. Polym. Sci., 25, 1217
(1980).
54. C. D. Han, Trans. Soc. Rheol., 18, 163 (1974).
55. A. S. Lodge and L. de Vargas, Rheol. Acta, 22, 151 (1983).
56. R. D. Pike, D. G. Baird, and M. D. Read, SPE Tech. Papers,
29, 312 (1983).
57. C. Rauwendaal, SPE Tech. Papers, 30, 282 (1984).
58. A. I. Leonov, Rheol. Acta, 15, 85 (1976).
59. J. L. Duda and J. S. Vrentas, Trans. Soc. Rheol., 17, 89
(1973).
60. R. E. Nickell, R. I. Tanner, and B. Caswell, J. Fluid Mech.,
65, 189 (1974).
61. D. J. Paddon and K. Walters, Rheol. Acta,18, 565 (1979).
62. R. W. Williams, Rheol. Acta, 19, 548 (1979).
224 Isayev and Upadhyay

63. A. L. Halmos a n d D . V. Roger, Trans. Soc. Rheol., 20, 253


(1976) .
64. S. F. Shen, M. A. Morjaria, and R. K. Upadhyay, Numerical
Methods in Laminar and Turbulent Flow, C. Taylor, K. Morgan,
and C. A. Brebbia (e d s .), Pentech Press, London, 1976, p.
459.
65. J. Ladveze and R. Peyret, J. Mécanique, 13, 367 (1974).
6 6 . H. Holstein an d D . J. Paddon, J. Non-Newt. Fluid Mech., 8, 81
(1981) .
67. C. A. Hieber, private communications, 1984.
6 8 . D. V. Roger and R. Binnington, Trans. Soc. Rheol., 21, 515
(1977) .
69. D. V. Roger, R. Gupta, and R. Tanner, \i. Non-Newt. Fluid
Mech., 4(1978).
70. M. E. Kim-E, R. A. Brown, and R. C. Armstrong, J. Non-Newt.
Fluid Mech., 13, 341 (1983).
71. J. S. Vrentas, J. L. Duda, and S. Hong, J. Rheol., 26, 347
(1982) .
72. P. Chang, T. W. Patten, and B. A. Finlay son, Comput. Fluids,
7, 267 (1979).
73. P. Chang, T. W. Patten, and B. A. Finlayson, Comput. Fluids,
7, 285 (1979).
74. M. Kawahara and N. Takeuchi, Comput. Fluids, 5, 33 (1977).
75. M. Viryayuthakorn and B. Caswell, J. Non-Newt. Fluid Mech.,
6, 245 (1980).
76. R. K. Upadhyay and A. I. Isayev, Rheol. Acta, 25, 80 (1986).
77. M. J. Crochet and K. Walters, Ann. Rev. Fluid Mech., 15, 241
(1983) .
78. M. J. Crochet, A. R. Davies, and K. Walters, Numerical Simula­
tion of Non-Newtonian Flow, Elsevier, Amsterdam, 1983.
79. A . I. Isayev, J. Polym. Sci. : Phys. Ed., 11, 2123 (1973).
80. J. G. Oldroyd, Proc. Roy. Soc., A 245, 278 (1958).
81. J. G. Oldroyd, Rheol. Acta, 1, 337 (1961).
82. M. A. Mendelson, P. W. Yeh, R. A. Brown, and R. C. Arm­
strong, J. Non-Newt. Fluid Mech., 10, 31 (1982).
83. R. I. Tanner, J. Non-Newt. Fluid Mech., 10, 169 (1982).
84. R. R. Huilgol, Continuum Mechanics of Viscoelastic Liquids,
Wiley, New York, 1975, p. 191.
85. M. G. N. Perera and K. Walters, J. Non-Newt. Fluid Mech., 2,
49(1977).
8 6 . M. G. N. Perera and K. Walters, J. Non-Newt. Fluid Mech., 2,
191(1977).
87. A. R. Davies, K. Walters, and M. F. Webster, J. Non-Newt.
Fluid Mech., 4, 325 (1979).
Flow of Polymeric Melts 225

8 8 . T. Cochrane, K. Walters, and M. F. Webster, J. Non-Newt.


Fluid Mech., 10, 95 (1982).
89. R. B . Bird, R. C. Armstrong, and O. Hassager, Dynamics
of Polymeric Liquids. Vol. 1: Fluid Mechanics, Wiley, New
York, 1977, p. 366.
90. A. S. Lodge, Rheol. Acta, 7, 379 (1968).
91. M. J. Crochet and R. Keunings, J. Non-Newt. Fluid Mech.,
10, 339 (1982).
92. N. Phan Thien and R. I. Tanner, J. Non-Newt. Fluid Mech.,
2, 353 (1977),
93. N. Phan Thien, J. Rheol., 22, 259 (1978).
94. M. W. Johnson and D. Segalman, J. Non-Newt. Fluid Mech.,
2, 255 (1977).
95. R. Keunings and M. J. Crochet, J. Non-Newt. Fluid Mech.,
14, 279 (1984).
96. A. I. Leonov, E. H. Lipkina, E. D. Paskhin, and A. N.
Prokunin, Rheol. Acta, 15, 411 (1976).
97. A. I. Isayev and C. A. Hieber, J. Polym. Sci.: Phys. Ed.,
20, 423 (1982).
98. R. K. Upadhyay, A. I. Isayev, and S. F. Shen, Rheol. Acta,
20, 443 (1981),
99. R. K. Upadhyay, A. I. Isayev, and S. F. Shen, J. Rheol.,
27, 155 (1983).
100. R. K. Upadhyay and A. I. Isayev, Rheol. Acta, 22, 557
(1983).
101. R. K, Upadhyay and A. I. Isayev, J. Rheol., 28, 581 (1984).
102. A. I. Isayev and C. A. Hieber, Rheol. Acta, 19, 168 (1980).
103. A. I. Isayev, C. A. Hieber, and D. Crouthamel, SPE Tech.
Papers, 27, 110 (1981).
104. A . I. Isayev, J. Rheol., 28, 411 (1984),
105. N. A. Jackson and B. A. Finlayson, J. Non-Newt. Fluid Mech.,
10, 55 (1982); 10, 71 (1982).
Orientation, Residual Stresses, and
Volumetric Effects in Injection
Molding
Avraam I. Isayev
Polymer Engineering Center
The University of Akron
Akron, Ohio

I. IN T R O D U C T IO N

Injection molding, one of the most widely used methods of polymer pro­
cessing, is characterized by high production rates and accurately sized
products. In the process polymer melt flows through a runner system
and gates; it is then injected into a cold mold, packed under high pres­
sure, and cooled until solid. During the process the polymer under­
goes simultaneous mechanical and thermal influences in the fluid, ru b ­
bery, and glassy states. For semicrystalline polymers this process is
further complicated by crystallization, which in turn determines crys­
talline morphology. These effects introduce orientation, residual
stresses, and shrinkage in the injection-molded products, which af­
fect the physical and mechanical properties, dimensional stability and
appearance of the finished product. So it is quite important to better
understand the factors governing the orientation, residual stress,
and shrinkage during molding.
This chapter briefly describes the current development of the ori­
entation, residual stresses, density, and shrinkage during injection
molding of thermoplastics. It also describes available experimental
techniques for measurements and physical models developed in the
modern literature, including testing their ability to describe available
experimental data. Some techniques and models can be applied to
amorphous and semicrystalline polymers, but we deal mainly with amor­
phous polymers.

227
228 Isayev

11. DEVELOPMENT OF O R IE N T A T IO N
IN IN JE C T IO N M O LDING

We consider the experimental and theoretical aspects of orientation de­


velopment in injection molding of amorphous polymers. The develop­
ment of the microstructure of injection-molded semicrystalline poly­
mers will not be covered. For excellent reviews of this area, see
[ 1 and 2 ].

A. E xperim ental Techniques

Experimental techniques used to characterize orientation, crystallin­


ity, and morphology in injection-molded articles include birefringence,
trifringence, Fourier transform infrared, wide-angle and low-angle
x-ray diffraction, and heat shrinkage. Some techniques character­
ize articles molded from amorphous and semicrystalline polymers,
whereas others are used only for semicrystalline polymers. We briefly
describe these methods.

B ire frin g e n ce

The most widely used technique for characterizing molecular orienta­


tion in amorphous and semicrystalline polymers is birefringence. A l­
though the method is restricted to transparent polymers, it does pro­
vide a general view of the development of molecular orientation dur­
ing polymer processing, including opaque polymers.
During molding the orientation at every point of the molded article
can be characterized by the following symmetric tensor of refractive
index [ 3] :

nil ni2 ni3\

ni2 ^22 H23 ( 1)

ni3 H23 1133/

For one-dimensional flow this tensor can be simplified to

n il n i2
“ \
n.. n i2 n22 (2)
1] «
0 0 n33/

The simplest way to define the state of the frozen-in orientation in the
molded part is to measure the three mutually independent components
of birefringence,
Orientation, Residual Stresses, and Volumetric Effects 229

An = / (n il ■ 112 2 )^ + 4 n i 2 ^, nn-ngj, and n 2 2 -ns 3

corresponding to the three mutually perpendicular planes xy, xz, and


yz, respectively (with 1 or x, 2 or y, 3 or z corresponding to the
flow, gap wise, and width directions of the cavity). A polarizing mi­
croscope with a calibrated compensator is usually used to measure the
various components of birefringence. Compensational and isochromatic
pattern methods are used. Thin slices are removed from the molded
part at different distances from the gate in order to determine gap-
wise birefringence distribution at various cross sections. The slices
are removed by a low-speed diamond saw, a microtome, or any other
way as long as the state of the orientation is not distorted during
cutting. (In particular, for a diamond saw, Isayev found [4] that a
cutting procedure with rotating speed about 135 rpm and cutting
weight about 0.25 N did not distort the optical picture and therefore
did not require additional finishing of the sample surface.) The slice
is placed between a cross polarizer and analyzer with the light path
going through its thickness. The birefringence components An, n ^ -
n 3 3 , and n 2 2 "H3 3 , respectively, are measured on thin slices, where
the thickness is along the 3, 2, and 1 directions of the cavity, as
shown in Fig. 1. We can calculate the value of birefringence, accord­
ing to the compensational method by the formula A = R/d, where R is
the retardation and d is the thickness of the slice. The lowest-order
fringe in the isochromatic pattern method is determined from the mea­
surement by the compensational method; higher-order fringes are

GATE

Fig. 1 Schematic illustration of slices removed from molded parts used


for measurements of various components of birefringence.
230 Isayev

counted off from the lowest-order fringe that is known. The orders
of neighboring fringes are determined by gradually changing the re­
tardation value introduced by the compensator and observing the
fringe movement since, at some points, the birefringence passed
through a maximum. The birefringence value, measured by the
fringe method, is found by the formula A = nA/d, where n is the or­
der of the fringe and A is the wavelength. Birefringence values cor­
responding to the maximum are found by using the fringe and com-
pensational methods simultaneously. In this case, a compensation be­
tween two neighboring fringes is carried out until these two fringes
coincided. Birefringence values are calculated by the formula A =
(nA + R)/d, where n is the order of fringes between which the maxi­
mum takes place and R is the retardation corresponding to the coales­
cing of the two neighboring fringes by compensation.

T rifrin g en ce

Birefringence is a measure of the difference in refractive indices


and is used as an indicator of the anisotropy of molded parts. How­
ever, it does not indicate the absolute values of the refractive in­
dices. In general, the refractive index can be related to the den­
sity and to the dielectric constant of the material. We see from Eq.
( 1 ) that the refractive index of the molded parts becomes dependent
on direction. Hence, to completely describe the refractive index dis­
tribution, we use the trifringence technique [5 ,6 ]. The trifringence
technique uses the Abbe refractometer with rotatable polarized ana­
lyzer in the eyepiece. Originally the technique was developed to mea­
sure the refractive index in the directions n ^ , n 2 2 , ngg of thin films.
It was later extended to thick samples, such as molded parts. There­
fore the Abbe refractometer has been modified to accommodate thick
samples removed from the molded part. The sample rotates from 0°
to 360°, and readings of the refractive index are taken at various ro­
tation angles. The flow direction is designated as 0° angle. The
measured refractive indices are fitted to the indicatrix equation

n(c|)) = n33 + (nil “ n33>cos^ ( 3)

When the flow direction is not known a priori, the symmetry axis is
found by a best fit of this equation to the data. The refractive in­
dex n 2 2 is obtained as the average of the many values measured
through direction.

Fo u rier Transform In fra red

Another technique for investigating orientation in molded parts is


Fourier transform infrared (F T IR ). It gives valuable information
Orientation, Residual Stresses, and Volumetric Effects 231

about the orientation of particular bonds and groups in the polymer


chain [ 6 ]. Basically, the technique is identical to infrared dichroism
measurements using polarized infrared radiation of a wide spectrum of
frequencies that have been known for a long time. However, interest
in this technique was recently increased due to advances in computer
analysis of the absorption spectrum. Infrared radiation of various
frequencies is directed to the polymer sample in the thickness direc­
tion. Depending on polymer structure and orientation, some fre ­
quencies pass through the sample with scarcely a loss in intensity,
whereas other frequencies undergo significant absorption. Absorp­
tion frequencies in thin films are usually the main objects of study
by infrared dichroism. However, in thick molded samples absorption
is strong. Therefore only weak absorbing bands are transmitted.
These absorbing bands are used for the quantitative characteriza­
tion of orientation in thick samples. A quantitative definition of or­
ientation is given by the Herman orientation function

D +2
3 cos^0 - 1 _ /D-1 o
(4)
\D+2/ \D + li
\ o 'y

where p can be replaced by c or am, respectively, for the crystalline


or amorphous phase, 0 is the angle between the molecular and stretch­
ing axes^ Dy is the measured dichroic ratio of the sample at frequency
Y , Do is the dichroic ratio of a perfectly oriented sample at frequency
Y -i.e .,

D = 2 cot a (5)
o y

where ay is the molecular transition moment angle at frequency Y. By


knowing the values of ay and Do, we can determine the orientation
function (fo and fam)* Then the average orientation function of the
sample is

V f + (1-V ) f ( 6)
c c c am

where Vc is the volume fraction of crystals in the sample.

X - R a y D iffra ctio n

In the study of semicrystalline polymers wide-angle and low angle x -


ray diffraction is frequently employed to determine orientation [ 7 - 1 2 ]
The distribution of the orientation and the sizes of the unit cells are
usually measured. In particular, the distribution of the orientation
232 Isayev

of the unit cell axes with respect to the alignment of the molecular
axis is usually sought. For uniaxial orientation the Herman orienta­
tion factor is defined by Eq. (4 ). For parallel alignment f = 1, for
perpendicular alignment f = -0.5, and for an isotropic system f = 0 .
However, in dealing with injection molding the orientation direction
has to be defined with respect to some direction. Unlike uniaxial
elongation, in injection molded parts the orientation direction is un­
known. Thus, it has to be determined with respect to some unique
direction or plane. In injection molding it is natural to assume that
the flow direction is the orientation direction.

H e a t-S h rin k a g e M eth o d

The degree of molecular orientation in molded parts can also be de­


termined by the heat-shrinkage technique [13—15]. Heat shrinkage
occurs because during molding some recoverable deformation becomes
frozen in the molded part. After the sample is heated above the glass
transition temperature Tg or melting point Tm, the oriented polymeric
molecules return to their original coiled state, leading to decreases in
dimensions of the molded sample. The samples for heat-shrinkage
measurements are removed from various sections of the part by a mi­
crotome. To determine the distribution of shrinkage in the direction
of the thickness of molded parts, we make cuts parallel to the su r­
face. The shape of the microtome cut is usually taken to be rectangu­
lar. In this case biaxial dimensional changes are observed. Accord­
ingly, shrinkage of the sample in one direction is influenced by dilata­
tion in the perpendicular direction. Thus, the correct values of the
experimental shrinkages in the flow direction and perpendicular to
the flow direction are

1 + L /L - /I - S ,
o i

B ^ -1 (7)
1 + w / w - / I - s ,,
o 11

where Lq and Wq are the initial length and width of the microtome
cuts, respectively, and L, W are their final values. These equations
are solved by iteration.
The gap wise average heat shrinkage can be measured in a sim­
pler way. Specimens with a large length-to-width ratio (more than
1 0 ; 1 ) and a thickness equal to that of the molded article are removed

at different distances from the cavity entrance in directions both


Orientation, Residual Stresses, and Volumetric Effects 233

parallel and perpendicular to the flow. A large length-to-width ratio


allows a simple formula from uniaxial elastic recoil after extension to
be used. Such samples are heated above Tg or Tm, and the shrink­
age is the ratio of the initial specimen length to the final (equilibrium)
specimen length.
Two effects must be considered in heat-shrinkage experiments.
First, the heat shrinkage in a molded part for a given polymer is a
function of the heating temperature and exposure time. To charac­
terize the shrinkage quantitatively, we must determine its equilibrium
value. On the other hand, the surface tension effect in the late stages
of shrinkage experiments can accelerate shrinking. This effect can
easily be recognized by observing the time dependency of the length
variation during shrinkage experiments.

B. Modeling O rie n ta tio n in Term s o f B ire frin g e n c e

An important assumption in modeling orientation in terms of birefrin­


gence is that the stress-optical rules can be applied to unsteady, non-
isothermal flow and relaxation [ 16]. About 30 years ago it was shown
that two stress-optical rules can be applied to isothermal flow of poly­
meric fluids [17,18]. The first rule states that the birefringence An
is linearly proportional to the principal stress difference Aa; i.e.
An = CAa, where C is the stress-optical coefficient. The principal
stress difference Aa = + 4aiz^, where = On - 0 22 » is fh® first
normal-stress difference and 0^2 is the shear stress. The second rule
states that the orientations of the optical axes, Xq , and the principal
stresses, Xs coincide; Xo = Xs = X, where X is the extinction angle.
This has been experimentally verified for isothermal steady-state flow
[18,19]. It was later extended to unsteady isothermal flow and re ­
laxation £20]. The applicabüity of the stress-optical rules to non-
isothermal flow has not yet been verified. However, in the range of
melt temperatures above Tg or Tm, we can assume that these rules
apply. In particular, the rheo-optical measurements at various tem­
peratures and frequencies support this assumption [16,21], because
above Tg or Tm a polymer is in the rubbery state in which, accord­
ing to rubber elasticity theory, birefringence and stresses are gov­
erned by the molecular chain orientation [ 2 2 ].
Birefringence, as stated already, at any point of the molded part
is also characterized by the components n n ~ and 1122 ~ 1 ^ 3 3 .
The stress-optical rule can also be applied to them [16,23]. The fol­
lowing relationships between stresses and these components have been
experimentally verified for temperatures above Tg or Tj^:

^11 " ” C N3, ÏI22 ~ 1^33 ” C N 2 > An - C (N i - 4 Qi 2 2 )n 0 . 5 ( 8)


234 Isayev

where N 2 = O22 ~ ^ 3 3 is the second normal-stress difference and N 3 =


Oil - 0 3 3 . These relations are plausible evidence for extending the
stress-optical rules to injection molding. Due to the nonisothermal
nature of the injection molding process, however, a temperature-de­
pendent stress-optical coefficient has to be incorporated in these
rules. Fortunately, the temperature-dependency of C is usually
weak if temperatures lie above Tg or Tm* Accordingly, the afore­
mentioned discussion indicates that by knowing the stress-optical
coefficient for a particular polymer melt and evaluating various com­
ponents of stresses frozen in molded parts, one can easily determine
three components of birefringence and quantitatively characterize
frozen-in orientation. This approach is only valid for amorphous
polymers.
Janeschitz-Kriegl was the first to derive a simple theoretical de­
scription of birefringence in molded parts [24]. Using Newtonian
fluid theory, he proposed a dynamic model for determining the thick­
ness of the frozen polymer layer in injection molding. His approach
is based on a global energy balance in terms of transient conduction
together with thermal convection and viscous heating. The develop­
ment of the frozen layer at the cavity entrance is governed by a hot
polymer melt flowing through the duct with wall temperature below
Tg. In this case the thickness of a steady vitrified layer grows pro­
portionally as the cubic root of the distance from the gate. Another
vitrified layer develops along the wall behind the advancing melt
front, which is governed by one-dimensional transient conduction
that reaches a thickness proportional to the square root of the dis­
tance from the end of the cavity. Thus, resulting predictions indi­
cate that the thickness of the frozen layer assumes a maximum near
the gate region, with zero values occurring at the inlet and at the
end of the cavity. The distance of the birefringence peak from the
strip surface is considered as a measure of the frozen-layer thick­
ness. This peak occurs due to a continuous increase of the shear
stress near the wall, which is due to the increase of the viscosity in
that region and the freezing effect of the penetration of Tg into the
sample during cavity filling.
Dietz and White [25] extended the approach of Janeschitz-Kriegl.
Their approach is based on the formation of a frozen layer of poly­
mer in the wall region and the use of an isothermal power-law model
in the core region. The thickness of the frozen layer is based on
the transient one-dimensional conduction problem for a semi-infinite
solid. The first normal-stress difference required for calculating b i­
refringence in Eq. ( 8 ) is based on an empirically determined correla­
tion with shear stress as obtained in steady shear flow :

Ni = Aai2^ (9)
Orientation, Residual Stresses, and Volumetric Effects 235

where A and a are empirical constants. In the post-filling stage they


consider nonisothermal stress relaxation, using a Maxwellian-type
model in which the relaxation time is independent of shear rate and
is the same for both shear and normal stresses. These stresses are
determined under continued cooling according to the relations

Oi2 = a ii“ e x p / -r

Ni = expM ( 10)

where are the shear stress and the first normal-stress


differences at the time of filling t f , and 0 is the relaxation time (which
is temperature - dependent ). Thus, Dietz and Whitens theory can han­
dle only the birefringence component An.
Recently, Greener and Pearson [26], following the procedure de­
veloped by Janeschitz-Kriegl [24] and Dietz and White [25], analyzed
a problem of frozen-in orientation in molded strips by treating the
nonisothermal stress relaxation process in the post-filling stage on
the basis of the viscoelastic constitutive equation of Marucci [27].
Their approach leads to better quantitative results than previously
obtained. However, there are some deficiencies in the above-men­
tioned approaches. In particular, two different constitutive equa­
tions for fluid have been used in order to treat one process: a
power-law model in the filling stage, and a viscoelastic equation in
the post-filling stage. Moreover, only the An birefringence compo­
nent has been obtained, although orientation in molded parts has to
be characterized by three components. Thus, a three-dimensional
approach to the problem of orientation is required. According to
Eq. ( 8 ) a three-dimensional viscoelastic model is needed to describe
the development of the shear stress O12 and the various difference
normal stresses, N i, N 2 , and N 3 . This was done by Isayev and Hieber
[28] on the basis of the Leonov viscoelastic constitutive equation. [29].
The applicability of this equation to various fundamental flow situa­
tions has been verified in Refs. 30—32. The idealized problem in
which a polymer melt at uniform temperature T q and fully developed
Poiseuille-type flow between parallel plates under constant average
melt velocity has been considered. The temperature of the plates is
abruptly lowered to below Tg, and the effect of the developing tem­
perature field in terms of transient one-dimensional conduction upon
kinematic and stress fields is determined until the cavity is füled at time
236 Isayev

0.5 1.0
y/b

Fig. 2 Gapwise distributioi> of (a ) linear velocity, (b ) shear rate,


and (c ) r at following times, in seconds: 1—0; 2—0.05; 3—0.2;
4—0.5; 5—1.0; 6—1.5; 7—2.0; 8—2.5; U = 12 cm/sec, b = 0.1 cm,
To = 230®C, Tw = SO'^C. (From Ref. 28.)

t = tf. Subsequently, for t > tf nonisothermal stress relaxation is


carried out until stress is frozen in the polymeric article. Figure 2
shows the resulting gapwise distributions of velocity u, shear rate
Y, and nondimensional shear rate, V = Y0 at various times during
flow with cooling. The velocity and shear rate markedly decrease
with increasing time in the vicinity of the cold wall. Hence, to main­
tain a constant average velocity, the flow increases in the hot core
region, with the associated maximum Y moving continually inward
from the wall with increasing time. Despite the small shear rates at
the wall, the product r = Y 0 increases with time, the maximum value
of r remaining always at the wall, because the relaxation time of the
polymer increases with cooling time faster than the shear rate de­
creases. This result is important since it apparently indicates that
the polymer continues to deform in the wall region.
Orientation, Residual Stresses, and Volumetric Effects 237

Fig. 3 Gapwise distribution of (a ) shear stress, (b ) first normal-


stress difference, (c ) second normal-stress difference, and (d ) bi­
refringence An for same flow conditions and flow times as in Figure
2. (From Ref. 28.)

The corresponding predictions for the developing stress distribu­


tions during flow with cooling are presented in Fig. 3. In particular,
the shear-stress distribution remains linear in y with the slope pro­
portional to the increasing pressure gradient as time progresses. On
the other hand, the results for Ni and -N 2 indicate that the normal-
stress components become frozen in the vicinity of the wall with the
peaks continually increasing in magnitude and moving inward from
the wall with increasing time.
Wales [23] has found that the stress-optical coefficient of polysty­
rene is essentially independent of temperature and equal to 4.8 x 10"
cm^/dyne. Thus, from this value for C and the above results for O12
and N i, we obtain the corresponding distribution of An from Eq. ( 8 );
238 Isayev

Fig. 4 Gapwise distribution of (a ) shear stress, (b ) first normal-


stress difference, (c ) second normal-stress difference, and (d ) b i­
refringence An at flow time t = 2.5 sec (1) , at time 0.05 sec follow­
ing cessation of flow (2 ), and at end of relaxation (3 ). Same flow
conditions as in Fig. 2. (From Ref. 28.)

see Fig. 3 (d). Although O12 varies monotonically with y , the com­
posite result for An exhibits a peak away from the wall due to the
dominant effect of N^. In addition, a local minimum appears near
the wall as a result of the frozen normal stress and the continually
increasing shear stress.
The corresponding predictions for the gapwise distributions of
^i 2 > Ni, - N 2 , and An during relaxation (i .e ., t > tf) are shown in
Fig. 4. These results indicate that the relaxation occurs mainly with­
in 0.05 sec after the flow stops. Also, O12 relaxes dramatically over
the entire cross section. In fact, the value of changes instanta­
neously at t = tf due to a corresponding discontinuous change in r .
Subsequently, O12 remains frozen in the wall layer but relaxes in the
core region. The appearance of a maximum in the frozen O12 distri­
bution is as would be expected on the basis that the value of O12
Orientation, Residual Stresses, and Volumetric Effects 239

increases as the solidified layer moves in from the wall. In contrast


to a12 , we note from Fig. 4 that the normal stresses remain unchanged
in the wall layer. Therefore since An is a composite of 0^,2 and N i, it
is clear that the An distribution changes throughout the cross section
during relaxation. Also note that the peaks in the normal stresses
and An shift in toward the wall and decrease in magnitude due to the
relaxation in the core region and because the initial peaks ( i . e . , at
t = tf) correspond to temperatures above Tg.
To apply the foregoing idealized problem to injection molding, we
use a result of Janeschitz-Kriegl [24] and Dietz and White [25], which
shows that during the cavity-filling process a thermal layer develops
along the wall behind the advancing melt front that, in leading ap­
proximation, is governed by one-dimensional transient conduction
when viewed from a frame fixed to the cavity. Therefore the thick­
ness of this thermal layer is proportional to , where t is here mea­
sured from when the melt front passed the given point. This model­
ing neglects viscous-heating effects and the steady-state thermal
boundary layer that develops from the cavity entrance ; in particu­
lar, due to the neglect of thermal convection, the present predic­
tions overestimate the thickness of the cooled layer near the cavity
entrance, as shown in [24] and [25] and as borne out in the com­
parison with experimental data [28]. In addition, the present model­
ing evidently neglects the complicated elongational-type flow in the
vicinity of the advancing melt front. Recently Lafleur and Kamal
tried to incorporate all of these neglected effects into a viscoelastic
modeling of injection molding [33,34]. However, their results are
incomplete, and it would be premature to discuss them here.
Despite these drastic simplifications, the comparisons indicate that
predictions based on the theory [28] do correlate in many respects
with the experimental results available in the literature, namely, the
injection-molding investigations of Isayev [4 ], Wales [23], Dietz
and White [25], and Kamal and Tan [35] for the amorphous poly­
styrene .
During nonisothermal flow the behavior of the polymer is unsteady
and the question arises as to what effect this unsteadiness has upon
the orientation development. Addressing this issue, Isayev and Hieber
[28] repeated the calculation of the normal stresses for the ’’inelastic^’
model that results from neglecting the time derivatives in the Leonov
model [29] such that the stress field responds instantaneously to the
changing temperature and velocity fields. Concerning the shear stress
and the associated pressure drop, the preceding model is similar to the
various inelastic models commonly employed in polymer processing sim­
ulations. The difference in pressure gradients between the viscoelas­
tic model or the ’Tnelastic’’ model [28] is negligible. However, the gap-
wise distribution of the first normal-stress difference at different flow
240 Isayev

Fig. 5 Gapwise distribution of first normal-stress difference accord­


ing to viscoelastic (solid) and ’’inelastic^’ (dashed) models at same
conditions and flow times as in Fig. 2. (From Ref. 28.)

times (Fig. 5) shows that the behavior according to the two models dif­
fers qualitatively in the wall region. For the inelastic model the max­
ima is always at the wall. However, in the viscoelastic model case
does not develop in the wall region due to a large relaxation time and
consequent freezing out at the lower values associated with the earlier
times. In the core region, on the other hand, the viscoelastic and
inelastic models give the same normal stresses as a result of the small
relaxation times associated with the high melt temperature. These
dramatic differences in the normal stresses for the viscoelastic and
inelastic models introduce a profound effect upon the gapwise bire­
fringence distribution in molded parts. Figure 6 shows the theoreti­
cal gapwise birefringence distributions for inelastic (curve 1 ) and
viscoelastic models (curves 2 and 3). Curve 2 corresponds to the
instant of fill time right before the start of relaxation, and curve 3
corresponds to the instant of the end of relaxation. The two models
give the same birefringence behavior in the core region but differ
qualitatively in the wall region. For the inelastic model the maximum
Orientation, Residual Stresses, and Volumetric Effects 241

Fig. 6 Experimental and theoretical gapwise birefringence distribution


in molded strip having dimensions W x 2b x L = 0.051 x 0,00254 x 0.48
(m) ; processing conditions are mold temperature 40°C^ melt tempera­
ture 223°C, average velocity 0.465 m/sec. Curve 1—’’inelastic” model;
Curves 2 and 3—viscoelastic model before and after relaxation, respec­
tively. (From Ref. 4.)

of birefringence is always at the wall, whereas the viscoelastic model


indicates that the maximum moves inward with the birefringence freez­
ing out at the wall. The latter is due to the large relaxation times as­
sociated with the low cooling temperature. Correspondingly, the two
models give the same birefringence distribution in the core region due
to small relaxation times at the high melt temperature. These results
indicate the importance of unsteadiness upon birefringence develop­
ment during cavity filling.
The aforementioned deficiency of the inelastic model prediction
can, however, be corrected if one assumes that there is no flow in
the boundary region where T < Tg. Then the flow mainly exists
in the core region, which forms a channel of smaller dimension where
T > Tg. Such an attempt has been described by Dietz and White
[25]. However, within the framework of this model a few additional
assumptions have to be made about the relationship existing between
242 Isayev

normal and shear stresses during nonisothermal flow and during their
relaxation after the cavity is filled.

C. T h e rm a l- and F lo w -In d u c e d B ire frin g e n c e

Frozen-in birefringence in the injection molding of amorphous poly­


mers appears from two main sources. The first is flow-induced b i­
refringence, which is a consequence of shear and normal stresses
developing during cavity filling and leads to orientation of the mo­
lecular chains. During the subsequent cooling stage these orienta­
tions do not completely relax and appear as frozen-in birefringence
in the molded part. In particular, the approaches to modeling of
this flow-induced birefringence have been described in Section II.B .
The second source of frozen-in birefringence is the nonequilib­
rium density or shrinkage change and the viscoelastic behavior of
the polymer during the inhomogeneous rapid cooling through the
glass-transition temperature, Tg, which results in thermal birefrin­
gence. Such thermal- and flow-induced birefringences are essen­
tially coupled, and the nature of this coupling is unknown. There
exist a few papers where thermally induced birefringence in freely
quenched polymer strips (without melt contacting solid wall) has
been measured [4,36—40] . Such quenching gives rise to thermal
birefringence in planes parallel to the thickness direction. Experi­
ments have shown that no birefringence appears when light is passed
perpendicular to the thickness of the strip, thus indicating that the
values of birefringence in the other two planes are equal. Various
polymers have been used in quenching experiments, including poly­
styrene (PS) [4,38,39], polymethylmethacrylate (PMMA) [4,37—39],
and polycarbonate (P C ) [36,37,40] . Figure 7 shows the typical dis­
tribution of birefringence through the thickness of a freely quenched
PS strip for three different initial temperatures and a constant quench­
ing temperature. Two distinctive features about the development of
the thermal birefringence can be observed. The birefringence has
positive and negative values near the surface and in the extensive
core region, respectively, with a parabolic shape and nonequal areas
of the positive and negative regions, the difference between these
areas being seen to increase with initial temperature. The position
of zero birefringence moves toward the surface with increasing ini­
tial temperature. The latter also is true for quenched PC and PMMA
strips. However, in contrast to the measurement for PS, the results
for PMMA and PC indicate negative and positive values of birefrin­
gence in the surface and core regions, respectively (Figs. 8 and 9).
This difference between the PMMA and PC birefringence distribution
in comparison with PS birefringence is largely due to the correspond­
ing stress-optical coefficient, C, behavior. The value of C for PS is
Orientation, Residuai Stresses, and Volumetric Effects 243

Fig. 7 Gapwise birefringence An distribution in freely water-quenched


polystyrene strips (thickness 2b = 0.26 cm). Curves 1—3 correspond
to quenching temperature 30°C vdth initial temperatures 130°^ 150®,
and 170®C, respectively.

temperature dependent and, respectively, negative and positive above


and below Tg [23,41] . On the other hand, for PMMA the value of C
is positive for the melt and negative at room temperature [42,43],
whereas for PC the value of C is positive above and below Tg [44—
46]. Therefore, when stresses in the quenched sample are compres­
sive in the surface and tensile in the core region, one would expect
the sign reversal in the thermal birefringence for the PS as compared
with the PMMA and PC. On the other hand, inequality of the area of
negative and positive birefringence cannot be explained by consider­
ing the temperature dependency of C. As indicated in Ref. 4, it can­
not also be explained by the concept of residual birefringence intro­
duced in Ref. 37. Apparently, a major contributor to the inequality
in birefringence is time dependent behavior of the stress-optical co­
efficient observed for various polymers [16,41—43,47—57], including
PMMA and PS [41,42].
244 Isayev

Fig. 8 Gapwise birefringence An distribution for freely water-


quenched PMMA strips (2b = 0.3 cm); curves 1—3 correspond to
quenching temperature 0°C with initial temperatures 130®, 150®,
and 170®C, respectively.

Fig. 9 Gapwise birefringence distribution for polycarbonate sheet


(2b = 0.4 cm), quenched from 180®C into different cooling media:
( • ) water at 20®C; ( o ) copper platens at 20®C; ( ) copper platens
a

at 120®C; ( a ) air at 20®C. (From Ref. 37.)


Orientation, Residual Stresses, and Volumetric Effects 245

y /b

F ig . 10 Effect of annealing on gapwise birefringence distribution


across quenched strip (2b = 0.16 cm); ( • ) quench from 140°C to 70°
followed by slow cool to 25°C ( ■ ) • + annealing at 85°C for 24 hr.
(From Ref. 26.)

The thermal birefringence may gradually relax after quenching,


since it was observed for quenched PMMA strips [4 ]. The relaxa­
tion process proceeds very slowly such that the value of An decreases
by about 50% in a time span of approximately one year. On the other
hand, in the quenched PS strips the thermal birefringence is frozen-
in permanently. The reason for the occurrence or absence of such a
relaxation phenomenon for different polymers is presently unknown.
However, there exists an appreciable effect of thermal-stress relaxa­
tion for many polymers [58,59]. Annealing of a quenched part close
to Tg also lowers the thermal birefringence level [26,37], as shown
in Fig. 10 for a PS strip.
The gapwise distribution of thermal birefringence introduced in
PS strips by constrained and free quenching is totally different [4 ].
Constrained quenching is performed with the sample located between
two metal plates, whereas free quenching is performed without any
restrictrion at the surface of the sample. In particular, the thermal b i­
refringence in constrained quenching of PS strips is negative through­
out the cross section, but in free quenching the birefringence changes
sign near the strip surface. Further, constrained quenching can be
performed at an initial temperature close to the melt temperature of
246 Isayev

injection molding. Thus, constrained quenching more closely repre­


sents the cooling process in the mold. Qualitative differences be­
tween constrained quenching or cooling in the mold and free quench­
ing apparently appear as a consequence of the interaction between
the polymer melt and the solid surface during constrained quench­
ing. However, for other polymers constrained and free quenching
also show the gap wise thermal birefringence distribution [4,37]. In
particular Fig. 9 presents such data for a PC strip quenched in wa­
ter, air, and between copper platens. Quenching in water and b e ­
tween copper platens gives only a slight difference in the gapwise b i­
refringence distribution. This is probably because before quenching
the sample did not adhere strongly to the itietal wall.
As for the effect of quenching conditions upon the thermal bire­
fringence distribution, it has been shown that air quenching reduces
the level of the thermal birefringence in comparison with water quench­
ing, as shown in Fig. 9 and Refs. 37, 39, and 46. In addition, an in­
crease in the initial melt temperature and a decrease in the bath tem­
perature increase the level of thermal birefringence [4,39,46].
In practice, it is difficult to separate the effect of thermal and flow
birefringences since there is no frozen-in flow birefringence in manu­
factured plastic parts without thermal birefringence also being pres­
ent, However, one can perform an annealing experiment near Tg in
which the thermal birefringence will be fully released without appre­
ciable relaxation of the flow birefringence. So far such an experiment
has not been performed. What has been done is partial annealing,
where the thermal birefringence is partly released from the molded
strip [26]. Since the separation of flow-induced birefringence from
thermal birefringence is difficult, the distribution of flow birefrin­
gence in molded parts can be understood on the basis of the model
described in Section II.B and by comparing the model prediction with
experimental data on birefringence in molded parts. In addition, a
contribution of the thermal birefringence in the molded part can be
realized by comparing various birefringence components in the molded
parts with the thermal birefringence.

D. O rien tatio n in Molded P arts o f Amorphous Polymers

Frozen-in orientation in molded parts of amorphous polymers can be


characterized by measuring various components of frozen-in birefrin­
gence An, n ii-n 3 3 , and n£2 "n 3 3 [16]. Correlation between these bire­
fringence values strongly depends on the distance from the gate and
processing variables. Apparently, during shear flow the main bire­
fringence is An, which includes an orientation caused by the maximal
tangential stresses composed of shear stress and the first normal-
stress difference. For this reason many authors have measured this
Orientation, Residuai Stresses, and Volumetric Effects 247

value of birefringence [4,16,23,25,26,35]. However, in some investi­


gations it is difficult to even determine which component of birefrin­
gence was actually measured. Spencer and Gilmore [60] and Ballman
and Toor [61] were apparently the first to observe birefringence in
injection-molded PS parts. In particular, the measurements in Ref.
61 have been performed apparently in the 1-3 plane (1, 2^ and 3 cor­
responding to flow, gapwise, and width directions, respectively).
They discovered a local maximum in the birefringence with zero val­
ues at the center and surface. The maximum birefringence increased
with injection pressure, fill time, and gate size, but decreased with
increasing mold temperature. Ballman and Toor attributed the bire­
fringence maximum to the solidified layer near the wall, which in­
creases the shear stress in the adjacent intermediate zone^ whereas
the zero birefringence at the surface was thought to be due to the
relatively unoriented polymeric material which flows from the hot
core to the wall along the zero-shear-stress melt front. Further in­
vestigations [3,4,23,25^26^28,62—64] have confirmed the existence
of a local birefringence maximum in the 1-2 plane; in addition, a bi­
refringence minimum has been observed [3,4,25,35,63] between the
surface and position of the birefringence maximum with large bire­
fringence values at the surface and small (but nonzero) values at
the center. Furthermore, it has been found in these later investi­
gations that the birefringence decreases with the distance from the
gate. Measurements in the 2-3 plane [3,4,35] have also shown the
occurrence of either a local maximum or a monotonic distribution of
birefringence, depending upon the position from the gate, with the
birefringence values being less than in the 1-2 plane.
Unfortunately, most of these experimental works have been per­
formed without considering the rheological properties of the materi­
als, which play an important role in the residual stresses and orien­
tation of the final product. Thus, among the numerous experimental
investigations in this area, the work of Wales et al. [23,64] is espe­
cially important, since it involves injection molding of materials with
well-characterized rheological properties. For example, for one PS
material, Wales has measured the steady-shear flow birefringence in
all three planes, has experimentally determined the dynamic and
steady-shear rheological properties, and has conducted injection­
molding experiments in which cavity pressures were monitored at
various locations and birefringence measurements of final parts were
made in the 1-2 and 1-3 planes.
Various rheological investigations have been concerned with flow
visualization of the injection-molding filling process, with special at­
tention being given to the kinetics of the advancing melt front and
the occurrence or absence of jetting [65—69]. Han and Villamizar
[70] have tried to measure stress birefringence development during
248 Isayev

the filling and cooling stages of injection molding. They point out the
influence of melt elasticity upon the residual stresses but do not pre­
sent any quantitative results. Later investigations [4,25,26,28] show
the predominant effect of normal stresses upon the development of b i­
refringence in the injection-molded parts.
We will discuss the development of various components of birefrin­
gence along with their gap wise distribution and the effect of process­
ing conditions. In addition, some comparisons of experimental results
with theory [28] , outlined in Section II.B , will be presented. First,
let us consider a general picture of the gapwise birefringence distribu­
tion in a molded strip at various cross sections from the cavity en­
trance (Fig. 11). It is seen that An passes through a maximum near
the surface, has a nonzero value at the center line, and a significant
value in an intermediate zone. This is typical birefringence behavior
and agrees with most investigations [3,4,23,25,26,28,35,64]. At cross
sections near the cavity entrance, a plateau region appears whose level
of birefringence decreases with distance from the entrance. At the
end of the cavity, a maximum in An could not be observed because it
was difficult to measure near the surface. Figure 11 also presents
results of the theoretical predictions of the birefringence based upon
the theory developed in [28]. Figure 11 shows that the theory is in
good agreement with experiment and shows a significant birefringence
at the wall with a sharp maximum in An which diffuses further in from
the wall with increasing distance from the cavity end. However, ex­
periment gives a significant birefringence in the core region, whereas
the theory shows a zero value. This discrepancy can be partly ex­
plained by the neglect in the present calculation of thermal birefrin­
gence, which appears in the molded strip during the cooling stage
and which has been considered in Section II.C . In addition, there is
possible contribution of packing pressure upon frozen-in birefrin­
gence which has also been neglected in the present calculation. In
particular, the packing pressure effect upon birefringence in the
molded objects will be described later.
The gapwise distribution of the birefringence in the molded strip
is affected by the shape of the entry to the cavity and by any changes
in cross-sectional area. In particular. Fig. 12 shows the gapwise b i­
refringence distribution in the preentrance region to the cavity having
nonsymmetric entry. The nature of the nonsymmetricity is shown in
the field of Fig. 12. It is evident that this nonsymmetric flow region
introduces a nonsymmetric birefringence distribution. The arrow per­
pendicular to the abscissa axis indicates the position of the surface of
the molded part. In particular, at the cross section corresponding to
the entrance of the cavity, the birefringence maximum is seen to be
higher near the converging wall (see curve 1). This effect might be
expected on the basis of a stress concentration in the flow near the
Orientation, Residual Stresses, and Volumetric Effects 249

Theoretical ( - - - ) and experimental ( ----- ) gap wise birefrin­


F ig . 11
gence An distribution in molded strip having dimensions W x 2b x l =
0.051 X 0.00254 x 0.48 (m) at different distances x from cavity en­
trance. Tw = 40®C, To = 223°C, U = 0.465 m/sec.

corner, which could introduce additional frozen-in orientation in the


molded part. However, the situation is changed at cross sections
further upstream (curves 2—4), where the birefringence maximum
is slightly higher near the flat surface. The cause of such behavior
is not clear. Still further upstream (curve 5), in the region pre­
ceding the converging section, the two birefringence maxima become
equal, as might be expected.
The influence of melt temperature and flow rate on the maximum
birefringence An^ax the gapwise direction can be seen from Fig.
13. Generally, increasing the melt temperature decreases An^ax
(compare curves 1 and 2). This is due to the fact that increasing
the melt temperature drastically reduces the effective relaxation
250 Isayev

y X 10 , m

F ig . 12 Gapwise birefringence An distribution in preentrance region


of molded strip at different distances x from cavity entrance. Pro­
cessing conditions and cavity dimensions are same as in Fig. 11. x —
isochromatic pattern method, o —method of compensation.

time of the melt. Hence it creates favorable conditions for relaxation


of stresses and birefringences after the cavity is filled. Curve 3 in
Fig. 13 shows the data of Dietz and White [25] for the same melt tem­
perature as curve 1, with a slightly lower mold temperature but a
flow rate 35 times smaller. It is seen that decreasing the flow rate
increases the value of Anmax. There seem to be two main reasons
for this effect. First, an increased flow rate creates favorable con­
ditions for stress relaxation after cavity filling. That is, it is well
known that increasing the shear rate (flow rate) leads to truncation
of the long-term tail of the relaxation spectrum of the polymer melt,
thus decreasing the effective relaxation time [71]. It means that
after the cavity is completely filled, stresses will relax faster follow­
ing higher flow rates. Second, at the high flow rate the thermal con­
vection dominates over the thermal conduction to the cold mold, thus
again creating favorable conditions for the relaxation process and
Orientation, Residual Stresses, and Volumetric Effects 251

0 ^ 8
X X 102,m

Fig. 13 Value of maximum birefringence Anmax molded strips ver­


sus distance from cavity entrance. Curves 1 and 2 correspond to cav­
ity dimensions W x 2b x L = 0.051 x 0.00254 x 0.48 m. Tw = 60°C, T q =
210°C, and 244°C^ U = 0.56 and 0.82 m/sec. Curve 3 corresponds to
cavity dimensions W x 2b x l = 0.027 x 0.00254 x 0.081 (m ), Tw = 50°C,
To = 210®, and U = 0.0159 m/sec. (Data depicted on curve 3 are taken
from Ref. 25.)

also retarding the growth of the frozen surface layer during the fill­
ing stage.
Figure 14 shows the dependence of the position of maximum bire­
fringence upon the distance from the entrance for different processing
conditions. It is clear that the position of this maximum characterizes
252 Isayev

X X 10^,m
0 ^ 8

F ig . U Position of maximum birefringence, (y/b)Anniax» molded


strips vs. distance from cavity entrance for processing conditions and
cavity dimensions being same as in Fig. 13.

the thickness of the highly oriented surface layer. In particular,


curves 1 and 2 correspond to essentially the same conditions except
for a melt temperature of 210°C and 244°C, respectively. Therefore,
a higher melt temperature gives a thinner oriented surface layer.
Further, the thickness of this layer is fairly constant in the entrance
region but decreases to zero at the end of the cavity. Curve 3 in
Fig. 12 shows experimental data by Dietz and White [25] obtained in
a mold with the same thickness as in our experiments and with the
same melt temperature as for curve 1 and only a slightly different
mold temperature, but with an average flow rate about 35 times smaller
than that for curves 1 and 2.
This comparison shows that decreasing the flow rate increases the
thickness of the oriented layer. For example, the thickness of the
oriented layer with an average velocity of 0.0159 m/sec is 4.4 x 10"^ m
near the cavity entrance in comparison with 1.8 x 1 0 " with a velocity
of 0.56 m/sec. Following Wales et al. [23] the literature defines the
thickness of the oriented layer to be the distance between the sur­
face and the gapwise position of the maximum An [16]. Fujiyama and
Azuma [72] have defined it as the thickness of the outer bright layer
on a polymer slice placed between cross niçois. However, it is evident
Orientation, Residual Stresses, and Volumetric Effects 253

Fig. 15 Comparison of gap wise birefringence An distribution for cross


sections near and far from cavity entrance for molded strips having
different thicknesses. Processing conditions for thin and thick cav­
ity [W X 2b X L = 0.0051 X 0.00254 x 0.48 (m) and 0.076 x 0.00381 x
0.456 (m )] are = 40°C, T q = 223'^C, and 224°C, U = 0.465 and
0.36m/sec, (From Ref. 4.)

that in the latter case the thickness of the brighter outer layer will
be slice-thickness-dependent.
The effect of the cavity thickness on the birefringence An distri­
bution in the gapwise direction at cross sections located near and far
from the gate can be seen from Fig. 15. A change in the cavity thick­
ness has a minor influence on birefringence far from the gate but a
significant effect near the gate. Increasing thm cavity thickness
254 Isayev

lowers the maximum frozen-in birefringence near the gate and shifts
its relative location toward the wall. Moreover, at the cavity en­
trance, values of An in the thick strip pass through a local minimum
near the surface, which is evidence of the presence of a highly ori­
ented thin surface layer (with a thickness equal to approximately 4%
of the cavity thickness) in addition to the highly oriented subsurface
layer usually observed in molded parts [3,4,23,25,28,35]. The bire­
fringence minimum has also been observed in Refs. 3, 4, 25, and 35.
According to theoretical work [ 28] , the thin surface layer may be a
result of the frozen-in normal stresses and the continually increasing
shear stresses near the wall during nonisothermal ñow. Another ef­
fect which may contribute to the formation of such a layer is the de­
velopment of extensional flow of the melt front during cavity filling:
the ’’fountain effect” [73,74]. For the thin strip at about 7 x 10"^ m
from the surface, difficulties usually arose in the birefringence An
measurement caused by problems in distinguishing isochromatic pat­
terns. Apparently, a surface layer is present in the thin strip which
has even a higher orientation than the thick strip. Two main charac­
teristics of this layer are maximal level of birefringence An^ax and
its thickness (y/b)Anmax* interest to observe how these
characteristics can be predicted by the theory described in [28] and
outlined in Section II.B . In particular. Fig. 16 shows the theoretical
and experimental values of An^ax and its position (y/b)Anmax versus
distance from the cavity entrance as shown in Fig. 16(a) and (b ) for
cavities having different thicknesses. The agreement between theory
and experiment is seen to be very good for An^ax* However, the
theory overpredicts the thickness of the oriented layer at the cavity
entrance, due to the neglect of thermal convection in the present cal­
culations. Moreover, the data indicate a jump in Ahmax at the cavity
entrance with a sharp decay in the preentrance region [Fig. 16(a)] ,
which is due to a contracting flow area there. In experimental and
theoretical investigations dealing with two-dimensional isothermal flow,
a stress or birefringence maximum has been observed during flow of
polymer solutions and melts through an area of contraction [75—78]
(see also Chapter 2). The present results show a similar effect for
frozen-in birefringence. This indicates that any contraction on the
cavity wall will be a source of additional frozen-in orientation.
Many years ago, Spencer and Gilmore [60] found that increasing
the packing pressure increases the gap wise-averaged birefringence
component < n n -n 3 3 >. More recently, Koita [79] made qualitative ob­
servations of the photoelastic stress patterns in a molded telephone
housing. From his observations he concluded that an increase in
packing pressure gives rise to frozen-in stresses due to molecular
orientation in the gate area. However, quantitative measurements
were not performed. In our recent work [80] , measurements of the
Orientation^ Residual Stresses, and Volumetric Effects 255

•Q

Fig. 16 Theoretical (dashed line) and experimental (solid line) results


for gap wise maximum birefringence, Aniy^^x position yAnmax»
in molded strip vs. distance x from cavity entrance; a and b corres­
pond to thin and thick cavities and processing conditions given in
Fig. 15, (From Ref. 4.)

gapwise distribution of birefringence An in molded parts have been


performed which directly indicate the effects of packing pressure.
In particular, Fig. 17 presents the gapwise distribution of bire­
fringence An at various levels of packing pressure for two different
cross sections corresponding to 0.01 m and 0.11 m from the gate, re­
spectively. Near the gate the increase in packing pressure introduces
major changes in the gapwise birefringence distribution. In particu­
lar, two birefringence maxima appear across the gap, one near the
surface and the other in the core region. In particular, the position
256 Isayev

y/b

Fig. 17 Gapwise distribution of birefringence An at cross section lo­


cated 1 cm from the gate at various packing pressures: □ —200; — a

500; O-IOOO psi. T q = 227°C, Q = 36 cm^/sec, Tw = 30®C. The cav­


ity dimensions are W x 2b x l = 0,04 x 0.00254 x 0.12 (m ). (From Ref.
80.)

of the maximum near the surface is practically independent of packing


pressure, whereas its height significantly increases with packing pres­
sure. On the other hand, the position of the maximum in the core re­
gion slightly moves toward the center with increasing packing pres­
sure with its height also increasing. The appearance of this second
maximum may be explained as follows. On the one hand, the birefrin­
gence maximum near the wall is formed during cavity filling. It is
this maximum which has been treated in Refs. 4, 25, 26, and 28. On
the other hand^ as cooling continues after cavity filling, the material
in the cavity undergoes contraction, thus creating space for additional
material to flow into. If packing pressure is imposed after filling, this
additional material wül get supplied to the cavity. Accordingly, the
deformation process in the material continues during this stage. Such
deformation proceeds under high stresses, thereby introducing high
orientation. Whereas the material adjacent to the frozen-in surface
layer has a large relaxation time due to a further decrease in tem­
perature, the material further toward the core region is still hot.
Hence development of normal stresses in the layers adjacent to the
frozen-in surface layer should be retarded relative to the core. This
Orientation, Residual Stresses, and Volumetric Effects 257

would therefore introduce a secondary maximum inside the core re­


gion, which would not undergo relaxation when packing pressure is
released. In the absence of the packing pressure, this maximum
would not appear.
Although there is some effect of packing pressure upon the gap-
wise distribution of birefringence at the cross section located far from
the gate, this effect is difficult to interpret due to the closeness of
this cross section to the cavity end.
The foregoing results indicate that the birefringence An is in­
fluenced by the melt temperature, flow rate, and packing pressure.
Increasing melt temperature or flow rate noticeably reduces the bire­
fringence maximum and the thickness of the frozen surface layer.
Decreasing packing pressure sharply reduces the overall level of b i­
refringence and its maximum. In addition, a contraction area in the
cavity and the thickness of the cavity are important factors. The
contraction area increases the level of the frozen-in birefringence.
On the other hand , increasing the cavity thickness reduces the value
of the frozen-in birefringence.

M l. DEVELOPMENT OF R ES ID U A L STRESS
IN IN JE C T IO N M O LDING

A. Experim ental Techniques fo r R e s id u a l-S tre s s Measurements

A variety of experimental techniques has been described in the litera­


ture related to measurements of the residual-stress distribution in
molded or quenched plastic parts. Among them are the hole drilling
method, the stress relaxation method, photoelasticity, and the layer-
removal technique.

H o le D r i l l i n g M e th o d

Numerous examples of using the hole drilling method for residual-


stress measurements appear in the literature. This method was first
described by Mathar [81] in 1934 and later improved and used in many
investigations of residual stresses in metals [82—84]. In particular,
drilling a hole disturbs the stress field equilibrium in the sample having
residual stresses. Deformations which develop on the surface during
hole drilling in the sample can be recorded by four strain gages lo­
cated around the hole in directions mutually perpendicular to each
other. The small incremental change in surface strain As for a small
incremental increase in hole depth Ay is proportional to the magnitude
of the average residual stress in the layer corresponding to this depth.
A disruption of the residual stresses in a biaxial planar stress field
can be determined according to the following equations based upon
pure elasticity:
258 Isayev

I — 5 3 ^ [ K ^ ( A £ ^ ) + v K , ( A e ^ ) ]
X K i - V K2

E
I ^ [ K ^ ( A e , ) + v K ,(A e^)] ( 11)
z Ki-v^K

For an equal biaxial stress field (a x = az) such as occurs in the


quenching of strip specimens , these equations can be simplified as
follows:

E(Ae)
a = o ( 12)
X z Ki + VK2

where and K 2 are constants which can be determined from calibra­


ting specimens with a known stress field.
The hole drilling method has not received much attention in the
polymer literature. Ito [85] has reported some experience with ap­
plying this method to polymers. The main difficulties are calibra­
tion, the need of special equipment, including the attachment of
strain-gage rosettes to the plastic surface, and a precise drilling
method that does not introduce a stress during machining. Other
limitations of this method concern the development of cracking in
plastic molded parts during hole drilling [86].

Stress R e la xa tio n M e th o d

The stress relaxation method for determining residual stresses in


polymers has also evolved from similar work with metals, in which
it has been shown that the kinetics of creep and stress relaxation
depend on the level of the residual stresses. Two approaches [87,
88] are available for applying this method. Both are based upon
the following power-law relation between stress, a, and strain rate,
e, in uniaxial extension:

(13)

where B and n are empirical constants. For stress relaxation, e = 0


such that

-EBa^ (14)

When a residual stress a is present in the plastic part, the effective


stress is Oeff = a - aj. If flow has no effect on the residual stress,
Eq. (14) can be rewritten as
Orientation, Residual Stresses, and Volumetrie Effects 259

-E B (a - (15)

By solving this equation and eliminating time t , we can find the slope
of the relaxation curve da/d In t:
n -1 .
a. 0-0.
da ____ 1 1
1- (16)
d In t n

where Oq is the initial stress.


The method proposed by Li [87] uses the intercept of the plot
-(da/d In t) versus a with the a axis to give ai- Therefore, a single
relaxation experiment is sufficient to determine ai. Since the value
of -(da/d In t) may depend on ao (or deformation) , in some cases it
may be necessary [89] to repeat the experiment at different values of
Cq and then extrapolate the dependence of a^ versus Cq to Cq = 0. The
intercept with the a axis gives the residual stress. However, Li^s
method is difficult to implement when the curvature of the stress re­
laxation curve a versus In t is low or practically linear. To avoid
these difficulties, Kubat and Rigdahl [88] have proposed a method
based upon the measurement of the maximum slope (-da/d In t)max
for a number of relaxation curves corresponding to various initial
stresses Oq (or strains). Then, by extrapolating the plot of (-da/
d In t)max versus ag to a zero value of (-da/d In t)niax>
find ai as the intercept with the Cq axis.
The methods of Li and of Kubat and Rigdahl have both been ap­
plied to residual-stress measurement in plastic slabs produced by in­
jection molding [90,91—94] and by other processing operations [95—
98]. The main disadvantage of these methods seems to be the impos­
sibility of measuring the residual-stress distribution in the gap wise
direction of plastic slabs. Therefore, the meaning of the residual-
stress values determined by these methods is somewhat unclear. It
seems that the most suitable means for determining the gapwise re ­
sidual-stress distribution is the layer-removal method [48], which
has been employed in various investigations and will be described
later.

P h o to e la stic M e th o d

Photoelasticity is a well-known technique for measuring stresses in


transparent models subjected to a strain or stress [99—101]. When
a transparent plastic part is subjected to a strain or stress, changes
in optical properties take place which are directly proportional to the
stress developed. One can see these in a circular polariscope by ob­
serving the photoelastic color sequences called isochromatic patterns.
260 Isayev

When residual stresses occur in molded or quenched articles, the iso-


chromatic patterns are permanent, with constant color fringes corres­
ponding to contours of constant principal stress difference. There­
fore, the residual-stress distribution can be studied by finding the
order and location of fringes if stress-optical rules can be applied.
It was mentioned in Section II.B that the stress-optical rules involve
the hypothesis that the principal axes of the stress tensor coincide
with those of the refractive index tensor. Thus, if the difference in
principal refractive indices An is proportional to the difference in
principal stress , the following formula is satisfied :

An = h i - n 2 = C ( O i - 0 2 ) (17)

where C is the stress-optical coefficient. Hence, the residual-stress


distribution can be calculated from measurements of the residual b i­
refringence distribution and the known value of the stress-optical
coefficient, Although the photoelastic method is used in the litera­
ture for measuring residual stresses (see, for example. Ref. 102), it
has two limitations when used with molded or quenched articles. One
is that residual stresses are generated in molded or quenched sam­
ples during a highly nonisothermal process in which material passes
through glass transition. At temperatures corresponding to Tg or
below, the stress-optical coefficient can become highly temperature
dependent, leading to a frozen-in birefringence distribution that dif­
fers from that for a material with a constant stress-optical coefficient.
The second limitation is due to the contribution of the memory ef­
fect to the stress-optical coefficient, such that the latter becomes
time dependent at temperatures corresponding to the transition from
the rubbery state to the glassy state and in the glassy state [47—57] .
This time dependency of the stress-optical coefficient leads to a sit­
uation in which the above-mentioned simple relation is not satisfied.
Therefore the photoelastic method should be restricted to materials
having constant stress-optical coefficient in the glassy and rubbery
states, and a transition zone between them.

Layer-R em oval Tech niq u e

Residual stresses in quenched and molded strips can be measured by


the layer-removal technique of Treuting and Read [48]. In this anal­
ysis thin layers of uniform thickness are removed from one surface of
a rectangular specimen, thus upsetting the residual-stress equilib­
rium in the piece. To reestablish equilibrium, the specimen warps to
the shape of a circular arc. We can calculate the gapwise residual
stresses in the specimen prior to layer removal by measuring the re­
sultant curvature as a function of depth of material removed, using
the general biaxial stress relation:
Orientation, Residual Stresses, and Volumetric Effects 261

-E Í f d P x (y i ) ^ d p z(y i)l
6 (l-v

+ 4 (b + y i)[P x (y i) * ^Pz^yi^^

- 2J [ p x ( y ) + vp^C y)] d y
Ì (18)
yi ^

where Ox is the longitudinal stress, y = ±b are the initial upper and


lower surfaces of the specimen, y = yi is the new surface formed af­
ter each removal, E is the elastic modulus, v is Poisson’s ratio, and
and p2 are the measured longitudinal and transverse curvatures,
respectively. o^Cyi) can be derived by exchanging the x and z sub­
scripts in Eq. (18).
Residual molding stresses vary with direction and position in the
strip due to the influence of flow stresses, so specimens parallel and
perpendicular to the flow direction in molded strips are removed. A
modified Tensilkut machine is used to mül layers in increments of
0.00003 m from these specimens. The resulting curvatures Px(y)
and pz(y) for specimens respectively parallel and perpendicular to
the flow direction are measured with an optical microscope immediately
after every so many layer removals. Curvature data are fitted by
polynomial regression, and gapwise stress profiles ^ ^ (y ) and az(y )
were calculated from Eq. (18).
Thermal stresses from quenching are isotropic: a (y ) = o x (y ) =
Oz(y) and p (y ) = P x(y) = P z (y )‘ Therefore, based on the measured
curvature p (y ), the thermal stress profile is calculated from the fol­
lowing simplified form of Eq. (18) :

E
o (y) Xb+yi 2)x d p ( y i ) + 4 (b + y i)p (y i)
6 (l-v ) dyi

- 2J p (y) dy| (19)


yi ’

Rigorously speaking, the method of Treuting and Read is applicable


to a linear elastic material with isotropic elastic constants. It has
been shown in several investigations [103,104] that Young’s modulus
of oriented injection-molded polymer sheets is independent of the
particular direction of the molecular alignment in the test specimen,
be it parallel or perpendicular to the specimen’s longitudinal axis.
Such modulus values are gapwise averaged, and some variations of
262 Isayev

modulus with gap wise position would be expected. These gapwise va­
riations of modulus have been found [105—107]. In many investiga­
tions, however, Young^s modulus has been assumed to be constant,
irrespective of direction in the sheet or gapwise location.
The polymeric materials studied in the residual-stress analysis
most definitely do not follow a linear elastic constitutive model. In
particular, viscoelastic phenomena of both creep and relaxation have
been observed for these materials [58,108,109].
With regard to creep, it has been observed that for a stress speci­
men reduced to half its thickness ( i . e . , at the end of the layer-re­
moval operation), the magnitude of the curvature increases with time
to some equilibrium value. Thus, the response of the specimen to
the moment induced by removing layers can be separated into two
components: (1) an instantaneous, elastic response, resulting in an
initial level of curvature, and (2) a time-dependent, creeping re ­
sponse, whereby the degree of curvature grows to some final value.
This behavior can be represented mathematically by the epxression

p ( y i . t ) = Pi(yi. O) + P2(yi,t) ( 20)

where p(yi,t) is the curvature of a machined specimen at time t after


the upper surface has been milled down to the plane y = yi, Pi(yi,0)
is the curvature immediately after machining, and p2 (yi>t) is the time-
dependent additional term.
This creeping action has an important consequence with regard to
the application of the layer-removal method of Treuting and Read [48]
to a viscoelastic material. Since their formulation is based on a linear
elastic material, the first term of Eq. (20), namely p i(y i,0 ), must be
used in applying the curvature-stress relationships to polymers. For­
tunately, the creep phenomenon occurs very slowly , and measurement
of the curvature within the first few minutes following machining is
soon enough to insure an accurate value for the time-independent
term.
Two opposing opinions on the subject of the time-dependent curva­
ture measurement have been expressed in the literature. Thakkar
and Broutman [108] feel that the curvature of a specimen should be
measured within the first two minutes following machining in order to
avoid what they refer to as stress relaxation. In this instance, stress
relaxation is a misnomer, since the observed time-dependent behavior
is that of creep. However, regardless of the terminology used, the
procedure of measuring the curvature soon after machining is proper
in light of the preceding discussion. That is, the elastic response of
the material is what one needs to measure to calculate the residual
stresses.
Siegman et al. [109], on the other hand, have expressed the opin­
ion that three days should elapse between machining and curvature
Orientation, Residual Stresses, and Volumetric Effects 263

measurement in order to allow the specimen to warp to its equilibrium


configuration. The curvature that they measured, then, was the to­
tal value, not just the elastic time-independent component. For this
to be proper a new formulation of the stress-curvature relationships
would have to be developed, one based on long-time viscoelastic be­
havior as opposed to the elastic formulation of Trenting and Read
[48]. Because they give no indication that any such formulation was
developed, it can be assumed that the curvature values used by those
authors in their residual stress calculations were large (long-time in­
stead of instantaneous values). Such an error should have manifested
itself in the results by overpredicting the residual-stress magnitudes.
A check of the results supports this assertion. For 1/4-in. sheets
of PMMA receiving the same thermal processing (quenching from 130°
to 0°C ), Siegman et al. [109], who waited three days following ma­
chining for each curvature measurement, show residual stresses in
the surface layers which are 55% higher than those determined by So
and Broutman [110], who allowed only two minutes to elapse between
machining and measurement. This difference in magnitude is far
greater than the estimated experimental errors of 5% to 15%. Thus,
it seems likely that the higher residual stresses observed by Siegman
et al. [109] can be attributed to the fact that they used the larger
equilibrium values of curvature rather than just the elastic compo­
nent.
Relaxation of the residual-stress levels has also been observed in
the quenched polymeric sheets [ 58]. A specimen tested immediately
after quenching will indicate higher residual stresses than one tested
several weeks after quenching. Thus, the experimentally determined
stress profiles depend on the time that has elapsed between when the
stresses are introduced and when the layer-removal procedure is car­
ried out. To draw any conclusions about the effect of processing con­
ditions on residual stresses, it is imperative that one measure the
stresses as soon as possible following processing. In addition, it
would be of value to monitor the decay of the stresses with time in
order to quantify the relaxation process in the particular material.
It is important that the layer-removal technique not introduce ad­
ditional stresses and that the size of the test specimen not effect the
observed results. With regard to the method of layer removal, two
machining procedures were examined: a vertical-spindle Bridgeport
milling machine and a modified Tensilkut milling machine [58]. The
curvatures observed when the Bridgeport milling machine was used
are approximately an order of magnitude higher than those detected
when the Tensilkut machine was used. This additional deflection can
be attributed to tensile stresses introduced by the machining process.
Thus, the Bridgeport milling machine is not an acceptable means of
removing layers for residual-stress analysis since the deflections
caused by machining damage could mask the deflections due to the
264 Isayev

residual stresses present in the piece before machining. The Tensilkut


machine, on the other hand, introduces negligible stresses. Also, the
machined surface produced by the Tensilkut remained transparent ,
whereas that produced by milling on the Bridgeport was matted. With
regard to the size of the test specimen, it has been found that pro­
vided the length/width ratio is large (> 6) , there is no appreciable
effect of size on the observed curvature [58].

B. T h e rm a l- and F lo w -In d u c e d S tresses

During the injection-molding process the polymer undergoes simulta­


neous mechanical and thermal influences in the fluid, rubbery, and
glassy states. Such effects introduce residual stresses and strains
into the final product [60,61], resulting in highly anisotropic mech­
anical behavior [62,105,111—115] and warpage and shrinkage [13,
116--118] . Thus, understanding the factors governing the residual-
stress development during molding is of great importance.
Residual stresses in an injection-molded plastic part can be attrib­
uted to two main sources. The first is due to the shear and normal
stresses that develop during the nonisothermal flow of the polymer
in the mold cavity during filling and packing [16, 25,26, 28]. These
stresses do not completely relax but get frozen-in due to rapid in­
crease in relaxation time during cooling. Such stresses will subse­
quently be referred to as residual flow stresses. The second type of
residual molding stress arises during the rapid cooling stage [119,
120]. The viscoelastic behavior of the polymer during its passage
through the glass-transition temperature, Tg, coupled with non­
equilibrium density changes accompanying the rapid, inhomogene­
ous cooling, results in the development of residual thermal (cooling)
stresses. A rather complex picture of the development of the resid­
ual-stress field in an injection-molded article thus arises in which the
flow-induced and thermally induced stress fields are coupled. The
precise nature of this coupling is not presently understood, and no
single constitutive model seems capable of describing the full behavior
of the polymer through the injection, packing, and cooling stages of
the injection-molding process.
Flow stresses appear mostly as a result of the cavity-filling stage
with an additional contribution apparently due to packing under high
pressure. High strain rates in the melt during injection introduce
shear stresses and first and second normal-stress differences [16,
25,26,28] together with extensional stresses [73,74]. The shear
stresses arise in any polymer fluid (Newtonian or non-Newtonian)
during injection behind a propagating melt front. The extensional
stresses develop mainly at the melt-front location and are caused by
an extensional flow of the melt in the ’’fountain region,” where a hot
Orientation, Residuai Stresses, and Volumetric Effects 265

melt from the core area moves toward the cold mold surface and sub­
sequently solidifies. Moreover, the extensional stresses also develop
in the region of contraction or expansion in the mold where the flow
suddenly or gradually changes. First and second normal-stress dif­
ferences arise as a consequence of the viscoelastic nature of polymer
melts having large relaxation time. After the cavity is filled, relaxa­
tion of all the above-mentioned stresses is hindered because the re­
laxation time of the melt rapidly (exponentially) increases when the
melt approaches its solidification temperature. Hence the flow stress­
es become locked into the polymer article.
In general, residual flow stresses cannot be measured directly.
However, since the flow stresses give rise to molecular orientation,
it has been assumed that measurement of frozen-in birefringence is
an indirect indication of the level of the residual flow stresses. It
is clear that flow stresses develop when the polymer is in the fluid
and rubbery states, i.e ., above the melting point or glass-transition
temperature, such that the linear stress-optical rule relating birefrin­
gence and stresses is usually valid [16]. Many investigations of
frozen-in birefringence in molded articles [4,16,23,25,26,28,64,121]
have been made with PS for which the stress-optical coefficient, C,
in the fluid and rubbery states is almost three orders of magnitude
larger than in the glassy state [4 ]. Data obtained by Wales [23]
show a remarkable correlation between gapwise-averaged frozen-in
birefringence and shear stress at the wall for molded PS slabs.
Few attempts have been made to theoretically predict residual flow
stresses and correlate them with frozen-in birefringence. All ap­
proaches are one dimensional as described in Section II.B . There
are other possible means of determining the residual flow stresses.
For example , one can release the cooling stresses by annealing a
molded article near Tg and then slow cooling to room temperature.
The residual birefringence can then be measured , and a distribution
of the residual flow stresses can be constructed by the stress-optical
rule. However, such type measurements of flow stresses do not seem
to have been carried out yet.
The area of thermal or cooling stress investigation has paralleled
similar developments about thermal-stress effects in inorganic glasses
and metals. Presently, there exists an extensive literature devoted
to measurements of thermal stresses in polymeric slabs. A brief sum­
mary of the results obtained by various authors has been presented
in Refs, 40, 58, 89, and 122. Here we would like to elaborate on
those findings.
So and Broutman [110] seem to have been the first to experimen­
tally investigate stresses introduced by free quenching of PMMA and
PC sheets by using the layer-removal technique of Treuting and Read
for metals [ 48]. A milling machine was used to cut the layers.
266 Isayev

Application of this technique to polymers has been reviewed in depth


above. In all cases of quenching, the thermal-stress profiles were
found to be parabolic, with a maximum compressive stress at the su r­
face and a maximum tensile stress at the center. For both materials
the surface residual-stress levels decreased with increased specimen
thickness. Residual stresses in annealed specimens were negligible.
Later, Siegman et al. conducted thermal-stress investigations on
quenched sheets of PMMA and PS [109] and polyphenylene oxide
(PPO) [122]. Stress profiles in the PMMA and PPO sheets were de­
termined by layer removal at 20,000 rpm with a tensile cut milling ma­
chine , and surface stresses in the PS sheets were qualitatively char­
acterized by observing crazing behavior in n-neptane. The parabolic
stress profile characteristic of thermal stresses was observed in the
PMMA samples. An increase in the initial temperature before quench­
ing from 120° to 130°C increased the level of compressive surface
stresses in the PMMA. The experimental procedure was deemed sat­
isfactory since residual stresses found in annealed samples were neg­
ligible. In the later experiment with PPO [122], the compressive
stresses at the surface decreased with increasing initial temperature.
Such results cannot be explained within the framework of the ther­
moelastic theory to be described here. On the other hand, their ex­
perimental results agree with the thermoelastic theory in showing that
increasing the quenching temperature reduces the level of compres­
sive and tensile stresses. At constant initial temperature, T q, a
linear dependence has been found between the logarithm of the re ­
duced surface compressive stress as/CTg-Tc») and the logarithm of
the reduced temperature (To-Too)7(Tg-Too), where'Q s is the surface
stress and T^« the quenching temperature. The slopes of such curves
decrease exponentially with increasing initial temperature. At an in­
itial temperature T q - Tg a unique value of as/(Tg-Too) has been
found which is independent of T q. The dependence of reduced thermal
stresses on the reduced temperature has been indicated by numerical
calculations of Struik [119]. Moreover, Struik has presented experi­
mental results on the measurements of surface compressive strains
in quenched PMMA sheets with thicknesses varying from 0.003 m to
0.01 m. He did this by measuring the distance between surface marks
before and after quenching. It has been found that the compressive
strain depends on the Biot number, Bi = hb/X, where h is the heat
transfer coefficient at surface, X is the thermal conductivity, and
b is the half-gap thickness of the sheet. Using the layer removal
method, Mandell et al. [115] have measured the thermal-stress dis­
tribution in constrained quenched polysulfone sheets having a thick­
ness of 3.2 X 10"^ m, and they have also determined their effect upon
the fatigue. In this case, quenching was performed between two alu­
minum sheets with thicknesses of 3.2 x 10“^ m. The initial tempera­
ture was 200°C, and the water bath temperature was 13°C. As for
Orientation, Residual Stresses, and Volumetric Effects 267

free quenching, the thermal-stress distribution was parabolic with a


surface compressive stress of -14.7 MPa and a tensüe stress at the
center of 7.4 MPa.
Some authors attempted to use birefringence measurements in con­
junction with the stress-optical rule to determine the thermal-stress
values in quenched polymer sheets. Broutman and Krishnakumar
[36] subjected PC sheets to various thermal treatments to introduce
thermal stresses. Then they calculated surface stresses in these
sheets with the stress-optical rule, using birefringence measure­
ments in conjunction with an experimentally determined fringe con­
stant. In addition, surface stresses were calculated by measuring
the surface strain resulting from annealing a sheet inscribed with a
square grid pattern. Results from the two methods agreed to within
a few percent. The birefringence profües through the sheet thick­
nesses were seen to have unequal positive and negative areas, and
thus utilization of the stress-optical rule would result in a violation
of stress equilibrium. To explain this, they proposed that a stress
component perpendicular to the plane of the sheet existed that was
zero at the surface and tensile in the interior.
Later, Saffell and Windle [37] investigated thermal residual stress­
es in quenched sheets of PMMA and PC using birefringence measure­
ments and the stress-optical rule. Cooling was performed by one of
three methods: (1) exposure to still air at 20°C, (2) immersion in a
water bath at 20°C, or (3) sandwiching between copper platens at
various temperatures. The parabolic birefringence profües, seen
for all three quenching procedures, were observed to have unequal
positive and negative areas. Since equüibrium dictates equal areas
for the stress profiles, it was obviously impossible to calculate resid­
ual stresses from the birefringence profiles by direct application of
the stress-optical rule. To alleviate this problem, Saffell and Windle
introduced the concept of ’’residual birefringence.” In their mea­
surements, birefringence values varied with sample thickness for
pieces larger than 0.001 m in a direction perpendicular to light path,
but they became constant for samples thinner than 0.001 m. This
constant profüe of residual birefringence, attributed to molecular or­
ientation induced by flow during cooling, was subtracted by the au­
thors from the measured birefringence profiles to yield a net profüe
which was then used to calculate residual stresses by the stress-op­
tical rule with a constant coefficient C. The surface compressive
stresses were seen to increase with increased sheet thickness, con­
trary to the results of So and Broutman [110].
Recent evidence [4] casts some doubt on the validity of the re­
sidual birefringence concept. It has been demonstrated that the bire­
fringence profile is independent of sample thickness, even for samples
of thickness considerably smaller than the critical value of 10"^ cited
in [37]. As an alternative explanation for the differences between
268 Isayev

the birefringence and cooling stress profiles, the time dependence of


the stress-optical coefficient in the glassy state as well as the tran­
sition zone from the glassy to the rubbery states has been proposed
in [4 ]. However, when the initial temperature prior to quenching is
not far above Tg, one can still derive reasonable results for the re­
sidual-stress distribution in slabs from birefringence measurements,
provided the gap wise birefringence distribution is symmetric. Greener
and Kenyon [38] have used the stress-optical rule to relate the frozen-
in birefringence distribution in quenched PS and PMMA disks to the
stress distribution as calculated from the theory of Aggarwala and
Saibel [123]. For PMMA disks at a critical temperature of 125°C, ex­
perimental results agreed well with theory, whereas for PS disks the
birefringence profile did not. In these calculations the stress-optical
coefficient has been taken to be constant and independent of time and
temperature and assumed to correspond to the solid material.
Bogue et al. [39,46] have also used birefringence technique, mea­
suring the gapwise birefringence distribution in quenched PS and PC
strips. To relate the experimental frozen-in birefringence distribu­
tion to the calculated thermal-stress distribution based upon the
theory of Lee, Rogers, and Woo [124], they used the strain-optical
rule with both a temperature-dependent and a temperature-indepen­
dent strain-optical coefficient. Both approaches give satisfactory re ­
sults even though a time dependence of the strain-optical coefficient
has been disregarded. However, an initial temperature of 120°C has
been used, which is not far above Tg. In both investigations [38,39,
46], the procedure for separating molecular and residual orientation
[37] has not been used. Recently, Mills [40] has also used the bire­
fringence technique for evaluating the residual-stress distribution in
quenched PC sheets of different thicknesses. This data has shown
unbalanced areas of compressive and tensile stresses, indicating that
use of photoelastic technique is incorrect for the case considered.
It is of interest to evaluate the effects of strip thickness, initial
temperature, and bath temperature on thermal stresses determined by
the layer-removal technique. Results are shown in Fig. 18 for ther­
mal stresses versus fractional gapwise coordinates for PS samples
freely quenched from 130° and 150° to 23°C and for PMMA samples
freely quenched from 130° and 170° to 0°C. These results have been
generated from fresh specimens for which the layer removal has been
performed approximately 1 hr after quenching. The stress distribu­
tions are parabolic, with maximum compressive stress at the surface
and maximum tensile stress at the center. For the quenched PS sam­
ples, surface compressive stresses of -9.8 and -8.2 MPa and center-
line tensile stresses of 2.6 and 3.3 MPa can be seen for initial tem­
peratures of 130° and 150°C, respectively. For the quenched PMMA
specimens these stresses are significantly higher and lie in the range
Orientation, Residual Stresses, and Volumetric Effects 269

Fig. 18 Thermal-stress profiles for 0.3-cm thick PMMA quenched


from 170^0 (curve 1) and 130°C (curve 2) to 0°C and for 0.0026-m-
thick PS quenched from 150°C (curve 3) and 130*^0 (curve 4) to 23°C.

from -29.8 to 7.2 MPa for the strip quenched from 170°C and from
-26.3 to 9.2 MPa for the strip quenched from 130°C. The parabolic
profile is characteristic of thermal stresses and has been observed
in both polymers [40,89,110,116,120,122] and inorganic glass [125—
129], It can be seen for both materials that there is little effect of
initial temperature on the magnitude of thermal stress. In Fig. 19
the effect of the bath temperature on thermal stress is shown. In
this case the PMMA strips of 0.003 m thickness were quenched from
170° to either 22° or 0°C. The bath temperature has a more signifi­
cant influence on the thermal-stress level than the initial tempera­
ture does, with surface stresses increasing about 40% for a 22°C
drop in bath temperature. Finally, the thermal-stress profiles for
PMMA strips of thicknesses 0.003 m and 0.0071 m quenched from
130° to 0°C are shown in Fig. 20. Note that the thickness increase
has very little effect on the observed compressive stress but notice­
able effect upon the tensile stress.
The viscoelastic phenomena of creep and stress relaxation have
been observed for quenched PS and PMMA specimens [58]. When a
specimen was released from the substrate following machining, the
resulting curvature was seen to creep with time to some equilibrium
270 Isayev

Fig. 19 Thermal-stress profiles for a 0.3-cm thick PMMA quenched from


170° to 22°C (curve 1) and 0°C (curve 2).

value. With regard to the creeping, it is important to measure the in­


itial elastic response of the material during bending after layer re ­
moval. It is this value of curvature that is needed to calculate the
residual stresses by the method of Treuting and Read. It was found
that the additional curvature due to creep decreases exponentially
with time. For both the quenched PMMA and PS the initial level of
curvature was approximately 85% of the equilibrium value that re ­
sulted after about three days following machining. Thus, by mea­
suring the curvature within the first few minutes following layer re ­
moval, the experimental error attributable to this creeping phenome­
non can be kept negligible.
A substantial degree of relaxation of the residual thermal stresses
has been observed for both PS and PMMA. For the results discussed,
the stresses were measured as soon after quenching as possible, typ­
ically within 60 to 90 min. However, when a longer period was al­
lowed to elapse between quenching and stress measurement, the ob­
served stress profiles were smaller in magnitude, presumably due to
stress relaxation.
In Fig. 21 the corresponding stress profiles for PS quenched from
150° to 23°C are shown, respectively, for elapsed times between quench­
ing and measurement of 90 min (curve 1) and 2 x 10** min (curve
2). It can be seen that the stresses have relaxed approximately 23%
during the time between these measurements. Similar results have
been observed for quenched PMMA sheets. Shown in Fig. 22 are the
Orientation, Residual Stresses, and Volumetric Effects 271

Fig. 20 Thermal-stress profiles for 0.3-cm (curve 1) and 0.71-cm


(curve 2) thick PMMA quenched from 130® to 0®C.

corresponding stress profiles as functions of time elapsed after quench­


ing for PMMA strips quenched as follows: (1) 0.003 m thickness, 170®
to 0®C, (2) 0.003 m thickness, 130® to 0®C, (3) 0.0071 m thickness,
130® to 0®C^ (4) 0.003 m thickness, 170® to 22®C. Relaxation by as
much as 50% of the initial stress levels can be seen in several cases

Fig. 21 Thermal-stress profile for 0.26-cm thick PS quenched from


150® to 23®C for elapsed time between quenching and measurement of
90 min (curve 1) and 2 x 10** min (curve 2).
272 Isayev

Fig. 22 Thermal-stress profiles for PMMA at various times after


quenching: (a) 0.3 cm thick, 170° to 0°C, curves 1 and 2 corres­
pond to times 90 and 3.3 x lO^ min, respectively; (b ) 0.3 cm thick,
130° to 0°C, curves 1 and 2 correspond to times 90 and 3.1 x lO^
min, respectively; (c ) 0.71 cm thick, 130° to 0°C, curves 1 and 2
correspond to times 90 and 3.2 x 10^ min, respectively; (d ) 0.3 cm
thick, 170° to 22°G, curves 1, 2 and 3 correspond to times 90, 2.9 x
lO"*, and 1.1 X 10^ min, respectively.

for elapsed times of 3.2 x lO^ min (24 days). As shown for the case
of the sheet quenched from 170° to 22°C [Fig. 2 2 (d )], the majority
of the stress relaxation has occurred during the first few weeks af­
ter quenching.
Stress relaxation is a very important concept with regard to cor­
relating processing parameters with residual stresses. To do so,
stresses must always be analyzed at the same time following process­
ing for all samples. If this rule is not followed, a specimen with
Orientation^ Residual Stresses, and Volumetric Effects 273

relatively low levels of residual stresses tested soon after process­


ing might give the illusion of possessing greater residual stresses
than a sample initially having higher levels of stresses, but tested
laterj giving a longer period for relaxation to take place. Ideally,
the residual stresses should be measured immediately after process­
ing, but this is not feasible in practice since specimen preparation
and the layer-removal procedure require a finite amount of time
(^3 h r).
The exact time to measure the stresses must be determined by the
experimenter. If stresses are measured as soon after processing as
possible, errors are introduced because the residual stresses are re­
laxing during the layer-removal process, since this initial period
marks the greatest rate of stress decay. On the other hand, if a
long period is allowed to elapse following processing so that the stress
state is essentially static over the duration of the test, then the ob­
served stress profiles can only imply the magnitude of the initial
stresses if the relaxation process has been quantified for the ma­
terial .

C. Modeling o f Residual Stresses

An extensive literature for the theoretical prediction of thermal


stresses in quenched inorganic glass exists. Since an effort is made
later in this chapter to test the accuracy of several of these theories
in predicting thermal stresses in polymer slabs, a review of some of
the approaches is in order.
Bartenev [125,126] developed a theory which incorporates the
thermoelastic response of the glass together with instant freezing
during cooling through Tg. This introduces the assumption that the
temperature gradient gets frozen in at Tg, which separates the p er­
fectly fluid and perfectly elastic regimes of the glass behavior. That
is, above Tg the glass is plastic, and plastic strains proceed instan­
taneously with internal stresses being absent. Below Tg the glass
is elastic, with plastic strains being residual and internal stresses
unrelaxed. Bartenevas theory leads to the following equation for re­
sidual-stress evaluation in an infinite slab:

a (y ) ( 21 )
l-v [<Ky) - <l>avl

where y is the gap wise coordinate, a the normal stress in the xz


plane, 3 the linear coefficient of thermal expansion, and E and v are
Young^s modulus and Poisson^s ratio of the glassy state, respectively;
(t> denotes a fictitious temperature, defined to within an arbitrary ad­
ditive constant :
274 Isayev

ST / ( 22)
' 87 ' '

where T (y , t*) Tg; i.e ., t * (y ) is the time at which the tempera­


e

ture passes through Tg for given y, with T (y , t) being the actual


gap wise temperature distribution at time t and (fav being the gap-
wise-averaged value of (().
For glass quenched relatively slowly from temperatures sufficiently
above Tg, the Bartenev theory has been shown to closely predict ex­
perimentally observed residual stresses in works by Bartenev [125]
and Garden [127]. For initial temperatures close to Tg and also for
more rapid cooling rates^ however, the theory has been shown to
yield much poorer correlation with experiment [127].
Subsequently, Indenbom [128,129] has criticized the Bartenev
theory on a number of counts. First, Bartenev’s hypothesis of the
freezing of the fictitious temperature gradient only takes into account
the temperature gradient in the layer undergoing solidification and
ignores the state of the layers that have already hardened. Also,
Indenbom has pointed out that Bartenev’s formulation is not based
on a consideration of the magnitudes of the strains on both sides
of the transition zone. According to Indenbom, the total strain in
the x-z plane, e, is uniform across the plate thickness and is com­
posed of three components :

£ (t) = £ .(y , t) + G ^(y, t) + 3 [T (y , t) - T q] (23)

where ee is the elastic strain, which is zero above T g, £p is the


plastic strain, which is time-independent below Tg, and T q is the in­
itial temperature. For the slab region where T < T g, the average
elastic strain Eq is zero, and the corresponding average total strain
e is

£(t) = £p(y) + 3 [T (y , t) - T q] (24)

Hence since is independent of y, such that G ( t ) = G ( t ) , the p re­


g

ceding two expressions can be equated at y = y * (t ) , where T (y * , t)


E T g and Ge(y*, t) = 0, giving

1 ^
e ; p ( y * , t ) + B(Tg-To) t ) +

6 [T (y , t) - T o ]} dy (25)
Orientation, Residual Stresses, and Volumetric Effects 275

which can be used to generate Cp, given T (y , t) and assuming that


£p remains independent of t for y > y * ( i . e . , T < T g ). Finally,
when the temperature has uniformly reached its asymptotic value,
Tooi setting £ = £ with èe = 0 gives

£ ( y ) + EgCy) (26)

Hence with £e(y) determined, the final stress can be calculated based
upon the following relation from elasticity:

E£e(y>
cr(y) = (27)
1-v

In particular, Indenbom^s theory has demonstrated [128,129] a bet­


ter correlation with experimental glass-tempering data than that of
Bartenev, particularly for the higher rates of cooling associated
with forced convection.
The two previously described theories of Bartenev and Indenbom
are based upon thermoelastic formulations and thus cannot predict the
procession of the stress profiles through the transient cooling period
(recall that Indenbom’s £p was time independent for T < T g ). In
order for a theory to have such a capacity, a constitutive model based
on viscoelasticity must be employed. Aggarwala and Saibel [123] have
developed a theory for predicting residual thermal stresses in an in­
finite plate based on a four-parameter viscoelastic constitutive model.

A (£ -e) + B (£ -e ) = C (a -a ) + D((i -á ) (28)


A X X

where the plate’s midplane is the x-z plane with surfaces at y = ±b,
and

e = Ka + £ (29)

where

1 ,
e 3 (^ x ^ ^y ^ (30)

a = 3 (a ^ + Oy + a^) (31)

/■a dT (32)
276 Isayev

in which £x strain and stress, respectively, in the


plane of the sheet, K is the compressibility, A, B, C, and D are
the constitutive parameters, a is the coefficient of linear thermal
expansion, and T is temperature measured relative to ambient. A p ­
plying appropriate equilibrium equations, the residual-stress profile
Ox can be expressed as

I
2AK+C 3dT
exp dx
/ 2BK+D L 2BK+D

2AK+C
(A e^ + Béjj. -Ae- B é ) exp dx (33)
L 2BK+D

where Zx unknown time function determined from the stress


equilibrium of

u
/ dy (34)
0
Aggarwala and Saibel illustrated their theory by solving it for a Max­
well material with a temperature-dependent expansion coefficient,
with the viscosity coefficient assumed to be zero for T > Tg and in­
finite for T < Tg. Such a formulation corresponds exactly to the
inviscid fluid and perfectly elastic states assumed in Indenbom’s
theory [128,129] and thus, not surprisingly, the two solutions co­
incide if the expansion coefficient is taken as temperature indepen­
dent.
Muki and Sternberg [130] have developed a viscoelastic theory
for predicting residual thermal stresses based upon a time- and tem­
perature-dependent modulus. They have treated the case of a slab
with complete lateral restraint and thus zero lateral strain during
cooling. Lee and co-workers [124,131] have subsequently modified
the approach of Muki and Sternberg [130] for that of an infinite slab
with free lateral expansion and zero lateral resultant force. The
theory in Refs. 124 and 131 is based upon nonisothermal linear vis­
coelasticity for the deviatoric stress-strain fields coupled with a
purely elastic treatment of the trace components of the stress-strain
fields. For an infinite slab of thickness 2b, the resulting governing
equations reduce to

j(y , t) = 3 f R [X (y , t) X(y, t ) ] ^ [e (t ) - 3T(y, t )] dt (35)


0
Orientation^ Residual Stresses, and Volumetric Effects 277

subject to the zero lateral resultant force constraint

„b
f a (y , t) dy = 0 (36)
0
where e (t) is the isotropic elongational strain in the xz plane, X is
the reduced time defined by

I
dt
X(y, t) = / (37)
K (T )

where K (T ) is a temperature-dependent characteristic relaxation time,


and R(X) is an auxiliary modulus function satisfying

,X
4(l+2v) dR(X)
f G(X-X) dX + R(X) = 2G(X) (38)
dX

where G(x) is the relaxation shear modulus.


Narayanaswamy and Gardon [132] have applied the theory of Ref.
124 to calculate the transient and residual stresses during quenching
of inorganic glass by using a modified numerical integration scheme.
The stress-optical law holds for such materials, and stresses were de­
termined experimentally by birefringence measurements. It was shown
that the model correlates well with experiment, provided the initial
temperature is significantly above Tg. At initial temperatures below
Tg, it is necessary to consider the contribution of structural effects
to the residual-stress field. Narayanaswamy [133] has extended the
theory to include these structural effects by evaluating the thermal
strain using the fictive temperature concept of Tool [134], The fic-
tive temperature has been calculated from the real temperature and
reduced time. It has been shown [133] that structural relaxation
can introduce about 25% of the total residual stress. Ohlberg and
Woo [135] further modified the theory in Ref. 124, where a tempera­
ture-dependent expansion coefficient has been incorporated into both
fictive temperature concept and also the pseudotemperature approach
of Muki and Sternberg [130].
This review of the theoretical predictions of residual stresses in
inorganic glass suggests the possibility of applying the theories to
residual thermal stresses in polymers. Only a few authors have at­
tempted this. In particular, Struik [119] has modified the theory of
Aggarwala and Saibel [123] as developed for a Maxwell material with
zero viscosity above Tg and infinite viscosity below Tg by neglecting
the temperature dependency of the expansion coefficient. A reasonable
278 Isayev

correlation was demonstrated between the theoretically predicted and


experimentally observed surface stresses for quenched sheets of
PMMA. Later, use will be made of the thermoelastic theory of Inden-
bom [128,129] (which is the same as Struik^s modification [119] of
the theory of Aggarwala and Saibel [123] and the viscoelastic theory
of Lee, Roberts, and Woo [124] in regard to present experiments with
quenched PS and PMMA.
Two other approaches [116,136] have been developed for predict­
ing the thermal-stress distribution in quenched polymer slabs. A l­
though the authors have referred to stress distribution in molded
parts, their approaches are valid only for thermal-stress prediction
since the flow stresses have been neglected in their formulation.
Menges et al. [116] have used the thermoelastic formulation for
shrinkage, Ae, with a temperature-dependent linear expansion co­
efficient a and Young’s modulus E. Therefore the residual stress
value at the ith layer of the slab can be derived from the formula

E.
1
Aq,* ( Ae-a^ATj) (39)
1-v

where the shrinkage Ae can be determined from a force balance at


every instant of time according to the equation

Ea.AT.E.
1 11
Ae = (40)
ZE.
1

Rigdahl [136] performed calculations for residual stresses by a


finite element technique, assuming that the residual stresses arise
from a complete hindering of thermal shrinkage during cooling. Hence
a formula from elasticity theory for the residual stress, a = EaAT,
has been used and coupled with the computed transient temperature
field obtained from the energy equation.
It is of interest to compare the experimentally determined thermal-
stress profiles for quenched PS and PMMA specimens with those p re­
dicted by theory. In particular, use is made of the instant freezing
linear thermoelastic theory of Indenbom [128] and the linear viscoelas­
tic theory of Lee, Rogers, and Woo [124], both of which have been
taken from the inorganic glass-tempering literature. Table 1 shows
the assumed properties used to calculate the theoretical stress pro­
files ,
In particular, the temperature field required to conduct the ther­
mal-stress calculations have been determined from transient one-di­
mensional conduction with a convective heat transfer boundary condi­
tion at the surface, namely
Orientation, Residual Stresses, and Volumetric Effects 279

T a b le 1 Mechanical and Thermal Properties of PS and PMMA


Used to Calculate Theoretical Stresses Profiles^

PS PMMA

Young’s modulus 2.11 X 10® 3.5 X 10®


E (N/m^)
Poisson’s ratio 0.33 0.33
V

Density 1.15 X 10® 1.18 X 10®


p (kgym^)
Glass-transition temperature 373. 373.
Tg (K )
Thermal expansion coefficient 7.0 X 10‘ ® 4.5 X 10’ ®
3 (1/K)
Thermal diffusivity 6.13 X 10'® 1.095 X 10'^
a (m^/sec)
Thermal conductivity 1.30 X 10'^ 1.9 X 10"^
k (J/sec*m*K)
Heat capacity 1.844 X 10® 1.47 X 10®
Cp (]7kg-K )
Heat transfer coefficient 1.47 X 10® 1.47 X 10®
h (J/sec-m^-K)

^Data other than h were taken from Encyclopedia of Polymer


Science and Technology, Interscience, New York, 1970.

dT 8^T
(41)
3t

subject to the following conditions on T (y , t) :

T T
T (y , 0) = To, — (0, t) - 0, -k — (b , t) = h [T (b , t) - T ^ ] (42)

with Too being the ambient temperature of the bath and a the thermal
diffusivity. Equation (41) has been solved by a finite-difference
method with standard tridiagonalization. The heat transfer coefficient
280 Isayev

h has been based upon steady natural convection for a vertical plate
in water at representative temperatures [137]:

h = 0.555 ^ (Gr Pr)^/“ (43)


Li

where k is the thermal conductivity of the water , L is the length of


the plate, Gr is the Grashof number, and Pr is the Prandtl number.
The time dependence of the relaxation modulus G (t) and the tempera­
ture dependence of the characteristic relaxation time K (T ) for PS, re ­
quired in the calculation based upon the theory of Lee, Rogers, and
Woo [124], have been taken from Aklonis and Tobolsky [138] for mono-
disperse PS S-111, which has a molecular weight = 2.39 x lo^ (prac­
tically the same as that of PS Styron 678 as reported by Wales [2 3 ]).
Although these two PS samples have different molecular mass distribu­
tions (MMD), it is well known that such differences would affect G (t)
only in the fluid and rubbery states and not in the glassy state and
transition zone where, as we will show, the value of G (t) is most im­
portant for determining the residual thermal stress. Figure 23 shows
the functions G (t) and K (T ) according to Aklonis and Tobolsky [138]
and the corresponding calculated auxiliary modulus function R(X) sat­
isfying Eq. (38). In particular , calculations for the thermal stress
have been performed with R(X) represented in tabulated form over
the entire domain indicated in Fig. 23, as given by

Rn
R(X) = Rq 10^ (44)
Pa
(1+5.5X)^

Resulting predictions based upon these two representations for R(X)


have been found to be essentially the same, thus indicating that the
rubbery and fluid states, corresponding to where Eq. (44) does not
represent R(X) well at all, contribute little to the final stress field.
That is, the glassy state and transition zone are mainly responsible
for the thermal stresses; on the other hand, as shown in Ref. 28,
the fluid and rubbery states characterize the residual flow-induced
stresses in injection-molded parts. Figures 24 and 25 show the gap-
wise experimental thermal-stress profiles and corresponding theoreti­
cal predictions for quenched PMMA and PS strips. In particular, ex­
perimental results in Fig. 24 for the 0.003-m thick PMMA strips
quenched from 170° and 130° to 0°C [(a ) and ( b ) , respectively] and
from 170° to 22°C (c ) together with 0.0071-m thick PMMA quenched
from 130° to 0°C (d ) are compared with theoretical predictions based
upon the Indenbom theory [128]. Indenbom’s theory adequately p re­
dicts the entire residual-stress profile for quenched PMMA. Struik
Orientation, Residual Stresses, and Volumetric Effects 281

T, "C
50 80 100 120 UO 150 180 200 220 240

O)

Fig. 23 Relaxation G and auxiliary R moduli versus reduced time X


and characteristic relaxation time K versus temperature T for PS S-111.
G(X) and K (T ) from [138] ; R(X) from Eq. (38).

[119] has demonstrated a similar correspondance between experiment


and his thermoelastic theory for quenched PMMA, but only the sur­
face stresses were measured. On the other hand, Fig. 25 shows ex­
perimental results for 0.0026-m thick PS strips quenched from 130°
and 170° to 23°C [(a ) and ( b ) , respectively] together with theoreti­
cal predictions based upon the theories of Indenbom [128] and Lee,
Rogers, and Woo [124]. Both theories substantially overpredict the
level of the experimentally measured thermal stresses. There are
two major effects which may contribute to this observed discrepancy:
(1) nonlinear behavior of the polymer during quenching in the tran­
sition zone between the rubbery and glassy states and (2) stress re ­
laxation after quenching. Since the layer removal was initiated ap­
proximately 1 hr after quenching, there was not enough time for sub­
stantial stress relaxation. Thus, for PS it appears that the above-
mentioned nonlinear behavior mainly accounts for the lower than
282 Isayev

Fig. 24 Experimental thermal-stress profile (curve 1) and prediction


based upon theory of Indenbom (curve 2) for 0.3-cm thick PMMA
quenched from 170°C (a ) and 130°C (b ) to 0°C and from 170°C to
22°C (c ) and for 0.71-cm thick PMMA quenched from 130° to 0°C
(d ).

predicted experimental stress values. Such a nonlinear behavior may


reduce the value of the shear modulus of the polymer and, corres­
pondingly, the residual stresses. It is well known from creep [139]
and oscillatory deformation experiments [140] that a maximum non­
linearity effect for polymers occurs in the region of the glass-tran­
sition temperature. Some indication of the possible importance of
such an effect for residual thermal stresses in plastics can be found
in Weitsman [141].
The effect of initial temperature on the stresses predicted by the
theories of Indenbom [128] and of Lee, Rogers, and Woo [124] are
shown in Fig. 26(a) and ( b ) , respectively, for a 0.0026-m thick sam­
ple of PS quenched at 23°C from either 200° or 130°C. Both theories
Orientation, Residual Stresses, and Volumetric Effects 283

Fig. 25 Experimental thermal-stress profiles (curve 1) and predic­


tions based upon theories of Indenbom (curve 2) and Lee, Rogers,
and Woo (curve 3) for 0.26-cm thick PS quenched from 130°C (a)
and 150°C (b ) to 23°C.

Fig. 26 Effect of initial temperature on thermal-stress profiles in


0.26-cm thick PS as predicted by the theories of Indenbom (a ) and
Lee^ Rogers, and Woo (b ) . Quenching temperature 23°C.
284 Isayev

0*C
30*C
60*C

a.

-3

-4

Fig. 27 Effect of quenching temperature on thermal-stress profiles


in 0.26-cm thick PS as predicted by the theories of Indenbom (a ) and
Lee, Rogers, and Woo (b ) . Initial temperature 150°C.

indicate that the stresses are nearly independent of initial tempera­


ture, Indeed, the experimental data for both PMMA and PS samples
evidenced only a small dependence of residual stress on initial tem­
perature. On the other hand, both theories predict a significant ef­
fect due to temperature of the quenching medium, as shown in Fig.
27, in which a PS strip of 0.0026-m thickness is again used as the
example with the quenching being from 150° to 60°, 30°, or 0°C.
The heat transfer coefficient h in Eq. (42) has been evaluated
based upon an empirical equation (43) from the heat transfer litera­
ture [137]. Such a procedure may introduce some error in calculating
the value of h, which in turn could lead to incorrect predictions of
the thermal stresses. Therefore, it is appropriate to assess the sen­
sitivity of the present thermal-stress predictions to the value of the
heat transfer coefficient. In this regard. Fig. 28 shows results for
the predicted gap wise thermal-stress distribution based upon the
Indenbom theory [128] with two different values for the heat trans­
fer coefficient. Changes by a factor of 2 in h in either direction
from the assumed value in Table 1 introduced a maximal change in
stress of only about 20%.
All the foregoing modeling attempts deal with thermal stresses
and disregard a consideration of flow stresses arising during injection
Orientation, Residual Stresses, and Volumetric Effects 285

2.94 X 10^
7.35 X 10^

Fig. 28 Effect of heat transfer coefficient on predicted thermal-stress


profile according to theory of Indenbom in 0.26-cm thick PS quenched
from 150° to 60°C.

molding. Only in recent years has modeling of flow stresses in injec­


tion molding received some attention. Attempts there were to simu­
late frozen-in shear and normal stresses. These simulation attempts
have been initiated in order to predict the observed frozen-in bire­
fringence profile in molded articles, and Section II.B gives a brief
account of these investigations. Problems of development of frozen-
in stresses are inherently linked to the solution of the problem of a
frozen thermal boundary layer arising during nonisothermal cavity
filling of the viscoelastic polymeric melt. In this case simultaneous
consideration of the thermal and flow stresses is required. This is
a challenge for future research efforts.

D. E ffe c t o f Processing C onditions on Residual S tresses

The experimental determination of residual stresses in injection-molded


parts has been performed in several works [58,86,90,115,116,118,
286 Isayev

142—145] . In most of them the layer-removal method of Treuting and


Read [48] was used to evaluate the stress profiles. Unlike residual
stresses of a purely thermal origin, stresses in a molded part are
anisotropic in nature^ owing to the contribution of flow stresses in
the flow direction. However, this effect has not been taken into ac­
count in earlier investigations. Coxon and White [142] have investi­
gated the effect of cross-linking on the residual stresses in injection-
molded high-density polyethylene. Layer removal showed the resid­
ual-stress profiles for all of the specimens to be compressive near
the surface, with a subsurface compressive maximum, and tensile in
the interior. The residual-stress profile for the cross-linked speci­
men was similar in shape to that of the un-cross-linked specimen,
with a slightly greater compressive stress maximum. Thus, the char­
acteristic shape of the residual-stress profile was determined by the
processing variables, and subsequent cross-lipking served only to
slightly increase the level of compressive stresses near the surface.
In a separate investigation, Coxon and White [143] looked at the
effect of aging on residual stresses in injection-molded polypropylene.
Specimens were stored 8 to 10 months following molding before analy­
sis at room temperature, 40°C (deep freeze), and -196°C (liquid ni­
trogen). Stress analysis was by the layer-removal method. For the
sample aged in liquid nitrogen, the residual-stress profile was marked
by a compressive maximum at the surface and tensile stresses in the
interior. Storage in liquid nitrogen was believed to reduce stress re ­
laxation or aging to a negligible level, and thus the observed stress
profile for this sample was thought to be that contained in a freshly
molded part. A significant reduction in the residual-stress levels for
the sample aged at room temperature was observed. The surface
stress was near zero, followed by a subsurface compressive maximum
and a tensile interior.
Kindle et al. [118] have reported on residual stresses in injection-
molded glass-filled polypropylene. Stresses were measured by the
layer-removal method in specimens parallel to the flow near the gate
and perpendicular to the flow far from the gate. For the parallel-
orientation specimens, stresses near the surface were compressive,
with a maximum value some distance below the surface and the in­
terior in tension. Residual stresses measured in the samples trans­
verse to the flow direction were compressive at the surface, tensile
in a region below this, and compressive again in the center of the
sheet. Varying the mold temperature between 30° and 80°C was
shown to have a negligible effect on the residual-stress profiles.
SandRands and White [90] have investigated the effect of injection
pressure and crazing on the residual stresses in molded PS bars by
layer removal. Varying the pressure from 37 to 143 MPa was shown
to have essentially no effect on the observed residual stresses.
Orientation, Residual Stresses, and Volumetric Effects 287

which were parabolic in shape. The introduction of surface crazes


by bending around a cylindrical form increased the levels of resid­
ual stresses in the molded bars.
Schmidt et al. [86] have studied the effect of processing parame­
ters on the residual stresses in injection-molded PS strips. Stresses
were evaluated qualitatively by drilling holes through the sheets and
immersing them in n-heptane. In this manner the magnitude and gap-
wise location of tensile stresses could be determined by observing
craze formation. Strips molded at low wall temperatures developed
crazes in the center of the strip, and those subjected to high pack­
ing pressures crazed at the surface. Variations in melt temperature
and injection rate produced no discernible crazing upon immersion.
This method of stress analysis gives no information about the com­
pressive stresses in the molded part, however. No crazing occurred
in any of the sheets if 21 days were allowed to elapse between mold­
ing and immersion in n-heptane, presumably due to relaxation of the
residual stresses.
Russel and Beaumont [144] have investigated the residual stresses
in Nylon-6 injection-molded bars. According to them, the stress dis­
tribution in the molding is parabolic with a compressive stress at the
surface and a tensile stress at center. The magnitude of these stress­
es was shown to be inversely proportional to the mold temperature.
For example, the maximum compressive and tensile stresses were,
respectively, 6.5 and 3.5 MPa for a mold temperature of 25°C, com­
pared to stresses of 1.2 and 0.8 MPa for a mold temperature at 100°C.
Apparently, the residual flow stresses were negligible for these N y­
lon-6 moldings. Applying the Aggarwala-Saibel theory for thermal
stresses [123] to predict the distribution of residual stresses in these
moldings, showed that the model overestimates the experimental data.
The authors have explained this discrepancy as a consequence of the
stress relaxation in the molding, since the model assumes the absence
of stress relaxation below the freezing temperature of the polymer.
However, no experimental evidence of relaxation has been presented.
Further, an exposure of the molding to boiling water reversed the re­
sidual-stress distribution—a tensile stress at the surface and a com­
pressive stress at center. They proposed that this effect was due to
water-induced crystallization in Nylon-6.
Mandell et al. [115] studied residual stresses in injection-molded
polysulfone bars, particularly their effect on fatigue. Again, in
molded as well as quenched bars, the residual-stress distribution
was parabolic with the level of compressive and tensile stresses being
about four times lower in the molded bars than in the quenched bars.
In contrast to annealed samples, fatigue cracks in the molded and
quenched samples initiated in the interior due to the presence of com­
pressive stresses at the surface.
288 Isayev

Siegmann et al. [145] made an extensive study of the effect of in­


jection-molding conditions on the residual stresses in molded square
slabs from PPO (N o ry l). The measurements were made at different
locations from the gate in a direction perpendicular to the flow di­
rection. Complicated stress distributions were found. The main ef­
fect of melt temperature was observed in the surface layer, where the
compressive stresses ranged up to 20 MPa but decreased markedly
with increasing melt temperature. In the interior the tensile stresses
had a peak value of 5 MPa, depending slightly on melt temperature
below 250°C. Above 250°C these tensile stresses decreased and even
became compressive. At increasing distance from the gate the su r­
face compressive stresses decreased and approached the level of
stresses in the interior. The compressive stresses in the outer lay­
ers were approximately constant up to mold temperatures of 60®C.
Above this mold temperature a marked decrease in the stresses was
observed. In the interior the tensile stresses decreased gradually
with increasing mold temperature. However, with increasing distance
from the gate, the dependence of stresses on mold temperature be­
came weak. The injection rate has been shown to have a dominant
effect on the residual stresses. At a low injection rate the surface
stresses are tensile, whereas at a higher injection rate, they become
compressive and eventually pass through a maximum with subsequent
decay. In general, this maximum decreases with increasing distance
from the gate. In the interior the dependence of residual stress on
injection rate was less pronounced. As for the influence of injection
pressure, the surface residual stresses first increased with increasing
pressure, reaching a maximum and then decreasing. The range of
injection pressure was 35—140 bar, with the value and position of
the maximum depending on the distance from the gate. A reversal
of the surface stresses from compressive to tensile was found near
the gate at higher pressures. In addition, at high pressures a high
compressive stress was detected in the core region near the gate.
The holding pressure duration was also found to have a significant
effect upon residual stresses. According to Menges et al. [116], in­
creasing the duration of holding pressure on PS moldings increased
surface compressive stresses. Moreover, the level of the holding
pressure itself had less influence on the residual stresses.
Based upon pure bending theory , Jacques [117] has presented a
thermoelastic analysis of warpage in injection-molded slabs due to un­
balanced cooling caused by asymmetric cooling and geometry. He
found that warpage is more highly dependent upon the design condi­
tion of the asymmetric cooling fluid temperature than on differences
in the metal thickness between cooling line and cavity surface.
In most of the previously cited work on stresses in molded parts,
the stress profiles have been calculated based solely on the curvature
Orientation, Residual Stresses, and Volumetric Effects 289

component along one axis of the specimen, with the curvature in the
other direction assumed to be zero or equal. This treatment is valid
only for the uniaxial or equal biaxial formulations. For proper analy­
sis of stresses in the molded part, curvature profües of specimens
oriented parallel and perpendicular to the flow direction must be mea­
sured for samples taken from the same location in the mold, and the
general biaxial formulation in Ref. 48 must be applied to determine the
stress profiles parallel and perpendicular to the flow. Moreover, an
incorrect treatment could account for the peculiar stress profües ob­
served in many cases.
Residual stresses in molded parts arise from the coupled effects of
flow and thermal stresses. Although the planar thermal stresses are
directionally independent with equal areas of tensüe and compressive
stresses across the thickness for symmetrically cooled slabs, the flow
stresses are tensüe and highly dependent on the direction. Thus,
residual stresses in molded parts are expected to be highly aniso­
tropic with a complex distribution in the thickness direction. An at­
tempt has been made in what follows to elucidate the effects of mold­
ing processing variables on residual stresses and to qualitatively ex­
plain the nature of their gapwise distribution on the basis of knowl­
edge available from residual, thermal, and flow stresses.
Let us first consider the effect of melt temperature. Stress pro­
fües are shown in Fig. 29, where o\\ and aj_ indicate the residual
stresses parallel and perpendicular to the melt flow direction, re ­
spectively. Curve 1 for a strip molded with a melt temperature of
244®C has a parabolic stress distribution everywhere, with higher
compressive surface stresses near the end of the cavity. The stress­
es in a strip molded with a melt temperature of 210°C (curve 2), how­
ever, differ dramatically from those in the other strip (curve 1), par­
ticularly in the flow direction near the gate. In this region curve 2
shows tensüe stresses at the surface of the piece. Near the end of
the mold, surface stresses are compressive but smaller in magnitude
than those found in curve 1. The trends observed here can be ex­
plained in a qualitative sense based on the contributions of the ther­
mal and flow stresses to the total residual-stress picture. Flow
stresses have been shown [120] to be tensüe and maximum near the
surface of the molded strip. Purely thermal stresses, on the other
hand, are parabolic in proflle with compressive surface stresses and
tensüe stresses in the interior. A molded part contains both of these
stress sources, and the nature of stress near the surface can be re­
garded as the result of the coupled competition of the tensüe flow
stress and the compressive thermal stress. For a high melt tempera­
ture the flow stresses are small, and the thermally induced stresses
should predominate. For low melt temperatures, on the other hand,
the flow stresses are much higher, and the possibility exists that
290 Isayev

z
2

Fig. 29 Residual-stress profiles parallel (a ) ) ) and perpendicular (aj_)


to the flow direction for molded PS strips [W x 2b x l = 0.051 x 0.00254
X 0.48 (m )]; (a ), (b ) near the gate; (c ), (d ) near the end. Curves
1 and 2 correspond to Tw = 60°C, T q = 244°C and 210°C, and U = 0.82
and 0.56 m/sec, respectively.

they could predominate over the thermal compressive stresses near


the surface and result in a net tensile stress. This behavior is be­
lieved to be responsible for the stress profiles seen in the region
near the gate for strips molded at different melt temperatures. It
was shown in Ref. 4 that the flow stresses are maximum near the
gate and drop off rapidly toward the end of the mold. Thus, the
effect of the tensile flow stresses should be most evident near the
gate, and this is precisely what is indicated in Fig. 29.
Orientation, Residual Stresses, and Volumetric Effects 291

NEAR GATE

FAR FROM GATE

Fig. 30 Residual-stress profñes parallel (a n ) and perpendicular (ajL)


to the flow direction for molded PS strips [W x 2b x L = 0.076 x 0.00381
X 0.456 (m )] . Curves 1 and 2 correspond to = 60^C, T q = 247°C
and 215°C, and U = 0.70 and 0.71 m/sec, respectively.

The stress profiles are shown in Fig. 30 for samples taken from
the 0.00381-m thick strips. Melt temperature has again been varied.
Just as for the 0.00254-m thick strips (Fig. 29), the higher melt tem­
perature (curve 2) resulted in parabolic stress profiles, whereas ten­
sile stresses are seen at the surface in the flow direction near the
gate for a strip molded at a lower melt temperature (curve 1).
The effect of cavity thickness can be seen by comparing the
stress profiles of a thin strip (curve 1) in Fig. 29 and a thick
292 Isayev

Near gate

Far from gate

Fig. 31 Residual-stress profiles parallel (o\\) and perpendicular (a_[)


to the flow direction for molded PS strips [W x 2b x L = 0.076 x 0.00381
X 0.456 (m )] . Curves 1 and 2 correspond to Tw = 40°C and 60°C, T q =
224°C and U = 0.36 m/sec, respectively.

strip (curve 2) in Fig. 30. Molding conditions for both were very sim­
ilar. The surface compressive stresses in the thicker strip are larger
than those in the thinner strip, indicating that either the thermal
stresses (see Fig. 20) are higher or that the flow stresses are smaller
(see next section), or both for the thicker sample.
The residual-stress profiles for strips molded with variable mold
wall are shown in Fig. 31. A change in the wall temperature from 40°
to 60°C produces no appreciable change in the shapes or magnitudes
of the stress profiles, which are close to parabolic in shape. Thus,
within the range investigated the mold wall temperature does not have
a significant influence on the residual molding stresses.
O rie n ta tio n , Residuai S tre s s e s , and V olu m etric Effects 293

Near gate

F ig . 32 Residual-stress profiles parallel (a||) and perpendicular ( q_l)


to the flow direction for molded PS strips [W x 2b x L = 0.076 x 0.00381
X 0.456 (m )]. Curves 1 and 2 correspond to Tw = 40^C, T q = 224°C
and 228°C, U = 0.36 and 0.83 m/sec, respectively.

The effect of injection rates on stress profiles in molded strips are


shown in Fig. 32. All the profiles are close to parabolic in shape,
with the surface compressive stresses found in the strip molded at
higher flow rates (curve 2) larger than those in the strip molded at
a slower rate (curve 1). This trend can be explained in terms of the
effect of flow rate on flow stresses. Birefringence measurements [4,
120] have shown that the stresses induced by flow are reduced for
larger flow rates, presumably due to the enhanced relaxation process.
Thus , the lower flow stresses would allow the thermal stresses to pre­
dominate, giving rise to greater compressive surface stresses for a
piece molded at higher injection rates. This is consistently shown for
stresses in molded strips (Fig. 32).
294 Isayev

E. Comparison o f Flow, T h e rm a l, and Residual


Stresses in Molded P arts

The question which naturally arises concerns the relative contribu­


tions of flow- and thermally induced stresses to the total residual
stresses in molded parts. To address this issue, we should first de­
termine the level of the flow stresses in molded parts. The flow
stresses can be calculated from the experimental birefringence dis­
tribution in the gapwise direction An(y) of molded parts by using
the stress-optical rule A a(y) = An(y)/C, where A a(y ) is the gapwise
distribution of the difference in principal flow stresses. A stress-
optical coefficient C = -4800 Brewsters for the PS melt has been taken
from the shear-flow measurements of Wales [23] , whereas An(y) has
been measured experimentally as described in Ref. 4. Such a calcu­
lation of the flow stresses is justified, since, according to investiga­
tions [4,28,146] which deal with predicting the maximum value of
birefringence in injection-molded parts, it is sufficient to use a con­
stant value of C found from the flow behavior of the given material.
Moreover, it has been shown in Refs. 4 and 120 for molded PS strips
that the contribution of thermal birefringence to the gapwise maxi­
mum in the total frozen-in birefringence An is very small near the
gate. Thus, at least the gapwise peak value of the flow stresses
can be correctly evaluated from frozen-in birefringence measure­
ments in molded PS strips. Therefore the effect of melt and mold
temperature on the gapwise flow stress distribution near the gate
can be seen from Figs. 33 and 34, respectively. In general, in­
creasing the melt temperature decreases the frozen-in flow stresses
significantly (Fig. 33). This is mainly due to lower flow stresses
during cavity filling and also faster relaxation of the stresses after
mold filling at higher temperatures. On the other hand, increasing
the mold temperature slightly decreases the frozen-in flow stresses
(Fig. 34). These findings agree with theoretical results [28] which
have been developed for predicting the flow stresses in molded parts
(see also Section II.B ).
Finally, it is interesting to compare the residual and flow stresses
in molded parts with the thermal stresses in quenched parts. Figure
35 shows the result of such a comparison. A main conclusion from
the comparison in Fig. 35 is that thermal stresses for the quenched
sample are about 10^ N /m^ and are much larger than the flow stresses
or residual stresses for the molded part, which are about 10® N/m^.
Although frozen-in flow stresses in the molded part are always posi­
tive, the residual stresses are positive in the core region and nega­
tive in the wall region. Since the residual stresses in the molded
part are due to the coupled effects of flow and thermal stresses, it
follows from Fig. 35 that these two components are of comparable mag­
nitude and that both are extremely important for the injection molding
Orientation, Residual Stresses, and Volumetric Effects 295

Fig. 33 Gapwise distribution of the frozen-in flow stresses in PS


molded strip near the gate at different melt temperatures. Curves
1—4 correspond, respectively, to the following processing conditions:
Tw = 60°C, To = 244°C, 210°C, 247°C, 215°C, U = 0.82, 0.56, 0.70,
0.71 m/sec, respectively. Curves 1 and 2 correspond to cavity with
W X 2b X L = 0.051 X 0,00254 x 0.48 m. Curves 3 and 4 correspond to
cavity with W x 2b x L = 0,076 x 0.00381 x 0.456 m.

of polymers. This also means that the residual stresses in the molded
part cannot be easily related to the value of shear stresses during
cavity filling, as has been proposed in Ref. 147, but it seems to be
a much more complicated problem. On the other hand, since the level
of thermal stresses in the molded part is apparently much less than
that in the quenched strip, this indicates that the cooling process in
quenching is more severe than that in injection molding. Furthermore,
it is clear that the flow and thermal stresses are apparently the main
cause of undesirable warpage and dimensional instability effects ob­
served in molded parts. Therefore, in the injection molding of
296 Isayev

F ig . 34 Gapwise distribution of the frozen-in flow stresses in PS


molded strip [W x 2b x L = 0.076 x 0.00381 x 0.456 (m )] near the gate
at different mold temperatures. Curves 1 and 2 correspond, respec­
tively, to Tw = 40°C and 60°C, T q = 224°C and U = 0.36 and 0.34 m/
sec.

polymers, special attention should be addressed to the design of ap­


propriate cooling systems which would result in reduced residual
stresses or which would provide a symmetrical thermal-stress dis­
tribution .

IV . DEVELOPMENT OF D E N S IT Y AND
SHR IN KA G E IN IN JE C T IO N
MOLDING

The mechanical properties of plastic parts can be strongly affected by


the presence of frozen-in nonequilibrium density and residual-stress
distributions. Nonuniform density distributions also affect the optical
properties of the material, since density and refractive index are di­
rectly related. In previous sections, experimental and theoretical re ­
sults for the frozen-in orientation and stresses in quenched and molded
parts were reported. In this section the measurement of density dis­
tributions and shrinkage and the results of their theoretical modeling
Orientation, Residual Stresses, and Volumetric Effects 297

Fig. 35 Comparison of gapwise distribution of frozen-in flow stresses


(curve 1), total residual stresses in directions parallel to flow (curve
2) at cross sections near (a ) and far (b ) from the cavity entrance for
molded PS strip (W x 2 b x l = 0.051 x 0.00254 x 0.48 (m ), Tw = 40°C,
T q = 223°C, U = 0.47 m/sec) and gapwise thermal-stress distribution
(curve 3) in 0.26-cm thick PS strip quenched from 170° to 30°C.

are presented. In particular, the density and shrinkage at any point


of a molded part become functions of space and time; presently, there
is no approach available for evaluating a priori the shrinkage of a
molded part after demolding. Thus, mold makers rely heavily upon
experience and intuition to correctly design molds and compensate for
unavoidable shrinkage of the part. Evidently, in doing so a few ex­
perimental trials involving an alteration of the mold will usually be
needed. This is especially critical for molding high-precision parts
for which dimensional specifications are to be strictly satisfied. Hell-
meyer and Menges [148] were the first to approach the modeling of
shrinkage by using the equation of state. However, their multilayer
model for shrinkage during the holding-pressure phase has been found
to be only in qualitative agreement with experimental data.
Concerning the measurement of density distributions in quenched
or molded polymers, we note that only a few investigations have been
reported. Further, some of the available data seem to be contradic­
tory. For example, Siegmann et al. [105] measured the density dis­
tribution in quenched strips of modified amorphous PPO (N o ry l).
Quenching from 145°C, with the sample constrained between two
298 Isayev

plates, he observed a steep density gradient near the surface, with


the density at the surface being lower than that at the center of the
sample. The overall density variation was from 1.058 to 1.116 g/cm^,
representing a total spread of about 5%. Greener and Kenyon [38],
on the other hand, observed quite the opposite effect for amorphous
PS. In particular, they freely quenched PS under a variety of condi­
tions. In all cases the density at the surface was higher than that
at the center; also^ the surface density was higher for higher cool­
ing rates. The total spread in density was only about 0.15%, Kamal
and Moy [107,149] have made density measurements in injection-molded
strips of polyethylene (PE) and found larger densities at the center.
For this material, however, partial crystallization can occur during
molding. Since the center experiences a lower cooling rate than the
surface, crystal growth should be more pronounced at the center with
a larger density there. The density variation in the flow direction
was found to be small relative to that in the thickness direction, and
the effect of melt temperature was small over the range investigated
(176°-206°C),
The nonuniform density distribution in quenched plastics is due
to the combined effects of rapid cooling through the glass-transition
temperature, causing frozen-in nonequilibrium density, together with
the effects of residual stresses. Garden [150] has attempted to sep­
arate these two effects in tempered inorganic glass. In particular,
refractive-index measurements, related to the density, were made in
unbroken specimens and in broken fragments of glass. In the former
case the density was hypothesized as being due to the sum of the two
effects, with the latter case corresponding to the stress-free density.
In particular, for the unbroken specimen the surface was found to
have a higher density than the center, whereas in the broken fra g ­
ments the reverse was true,
Wust and Bogue [39] investigated the volume aging of homogene­
ous (no variation of properties in the gapwise direction; i , e . , the
strip was thin enough to disregard any cooling rate variation across
the thickness) strips of PS quenched at different rates. The density
of samples quenched rapidly was smaller than that of slowly quenched
samples. The rapidly quenched samples also aged faster.
Aging refers to the variation of properties over a long time [151].
It is due to the kinetic nature of the glass transition. That is, above
Tg the molecular mobility is higher so that the relaxation time is very
short compared to the time scale of the experiment. Therefore, any
change in temperature results in an almost instantaneous change in
volume. As the polymer is cooled through the glass-transition tem­
perature, the molecular mobility decreases rapidly with an accom­
panying increase in viscosity and modulus. At glass transition the
relaxation time is of the same order as the experimental time scale.
Orientation, Residual Stresses, and Volumetric Effects 299

So the polymer molecules cannot rearrange themselves into their equi­


librium configuration at that temperature. That is, they ’’lag” behind
and are frozen into a nonequilibrium state of higher volume. This
serves as a driving force for attaining the equilibrium configuration.
However , due to the highly reduced molecular mobility below glass
transition, this results in a long-term variation in the physical and
mechanical properties of the plastic. We give a brief review of the
glass-transition theory and the theories of nonequilibrium behavior.
Struik [151] and Tant and Wilkes [152] have given more details.
Alfrey, Goldfinger, and Mark [153] have used the free-volume or
hole idea, originally proposed by Eyring [154] to describe the diffu­
sion from a liquid to a gas, to describe glass transition. The free
volume is a way of describing the intermolecular distance, which in
turn affects the transport properties such as viscosity and diffusiv-
ity. The free volume V f can be defined as

V f - V - YAT) (45)

where V is the actual volume and V q is the volume occupied by the


molecules. Turnbull and Cohen [155] have hypothesized that the
free volume is distributed over holes of various sizes, and that only
holes which are larger than a certain critical volume constitute the
free volume that contributes to the molecular mobility. Thus, mobil­
ity depends on the distribution of free volume and not solely upon
the total free volume.
Rusch [156] has described free volume in terms of Fig. 36 as fol­
lows, The free volume, Vf, is defined as that part of the thermal
expansion which can be ’’freely” redistributed. A temperature rise
causes an increase in the free volume as well as in the occupied vol­
ume. The total free volume V f consists of the equilibrium (v f) and
nonequilibrium (w f) free volumes. T 2 is the lower limit of the range
of Tg. At T 2 there will be no equilibrium free volume.
The free volume V f determines the molecular mobility, which in
turn determines the rate at which equilibrium is reached, i . e . ,
dVf/dt. This in turn determines Vf. This can be expressed as
[151]
M

f dt
where M is the molecular mobility. The closed-loop scheme implies
that volume relaxation is a nonlinear process [that is, M, hence Vf,
affects the relaxation time, as in Eq. (46)].
300 Isayev

F ig . 36 Volume-temperature curve to illustrate the free-volume con-


cept,

Rate equations have been proposed to describe the nonequilibrium


behavior of amorphous polymers. For a review of the theories see
Ref. 152. In particular, Kovacs [157] has proposed a first-order
rate equation to describe volume relaxation:

d (V -V co') (V -V ^ )
(46)
dt T(V^, T)

where V is the actual volume, the equilibrium volume, and x (V f, T )


the relaxation time, which depends on Vf and T. Kovacs has extended
this to the case where the temperature changes at a constant rate:

d (V -V 00) (V
^ 0-V0 )
-----—----- = AaV q + ---------- (47)
dt oo T
Orientation, Residual Stresses, and Volumetric Effects 301

where q = dT/dt is the rate of cooling or heating and Aa is the dif­


ference in the expansion coefficient between the rubbery and glassy
states. The free-volume theory, which is a single-parameter model,
uses the free volume as the ordering parameter in addition to tem­
perature and pressure to specify the state of the system. This the­
ory is useful in providing a qualitative description of nonequilibrium
behavior.
Later, Kovacs et al. [158] proposed a multiparameter model. That
is, a distribution of relaxation times rather than a single relaxation
time is used, namely.

d/(V-V„)/Voo).
= Aa.q + (48)
dt 1 V T .
1

where i refers to the order.


The Williams, Landel, and Ferry (WLF) equation [159] can be used
to describe the temperature dependence of the relaxation times. That
is, a shift factor a is introduced, such that

(49)

where and tare the relaxation times at some arbitrary reference


temperature (T r ) and given temperature (T ) , respectively, and ax is
the shift factor at the given temperature. The WLF equation has the
form

- C i (T - T r )
log = (50)
Ca + (T - T r )

where Ci and Cz are constants which depend on T r and the polymer


under consideration. It has been found that if the reference tem­
perature is chosen to be Tg, then = 17.44 and C 2 = 51.6°C for
a wide range of amorphous polymers [159] with Eq. (50) applying in
the temperature range Tg < T < Tg + 100°C.
Below the Tg, the WLF equation fails to predict experimental val­
ues of a [156]. Accordingly, Rusch [156] has derived a modified
form of the WLF equation based on the assumptions that the relaxa­
tion mechanism is controlled primarily by free volume and that the
total (equilibrium + nonequilibrium) free volume V f determines the
relaxation time. This gives
302 Isayev

- Ci(Te-TR)
(51)
" C2 + ( T e -T R )

where

Wf
T +— for T > To
Aa ^
T =
e
w.
T, + for T < To
Aa

In particular, T 2 denotes the glass-transition temperature observed in


experiments of infinite time scale, and w f(T ) is the nonequilibrium
free volume at temperature T. Further, Te is defined as the ^’effec­
tive temperature,” which is the hypothetical temperature at which the
free volume of equilibrium glass is equal to the total free volume of
the nonequilibrium glass at temperature T . Thus, relaxation will pro­
ceed as long as wf is nonzero.
Wust and Bogue [39] used the model of Rusch [156] to fit their vol­
ume-aging data of homogeneous samples and have obtained good agree­
ment. However, they had to use different values of the relaxation
time for quenching and aging.
In the sequel experimental and theoretical results are reported on
the density distribution and relaxation in freely water-quenched PS
and PMMA strips and injection-molded PS strips. In addition, the re ­
sults on volumetric shrinkage modeling and experimentation in injec­
tion-molded strips are presented.
The density distributions in PS samples of 0.381-cm thickness,
quenched from 170^ and 130^ to 23°C, are shown in Fig. 37. A steep
density gradient occurs at the surface, with the gradient being steeper
for the 170°C 23°C quench. On the other hand, the density away
from the surface is almost constant. The average density of the un­
cut specimens was also measured^ as shown by the horizontal lines.
The total variation in density is only about 0.1%. Further, the sam­
ple quenched from 130° to 23°C has only a slightly lower density than
the one quenched from 170°C, indicating that the initial temperature
is not an important factor affecting the density distribution.
In Figs. 38 and 39 the effect of the bath temperature and sample
thickness are presented, respectively. As for the initial temperature,
these two variables also do not appreciably influence the density dis­
tribution in the quenched PS slabs. However, there is a noticeable
trend indicating that lower bath temperature and thinner thickness
lead to a lower density.
Orientation, Residuai Stresses, and Volumetric Effects 303

o
(O

Fig. 37 Lines of average density and gap wise density distribution in


0.38-cm thick PS quenched from 170°C (line 1, o) and 130°C (line 2,
□ ) to 23°C, measured 24 hr after quenching.

For PMMA the effects of initial and bath temperatures are shown
in Fig. 40. The distributions are those observed 24 hr after quench­
ing. For the 170°C 23^C quench the density varies from 1.1896 to
1.1926 g7cm^, that is, by about 0.25%. Again, a similar variation in
density, as for PS, can be observed, i.e ., a steep gradient near the
surface and a nearly constant density in the interior. The sample
quenched from 130^0 has a higher density than that quenched from
170°C. The same dependence on initial temperature has also been ob­
served for PMMA samples 0.28 cm thick. In addition, we note from
Fig. 40 that the samples quenched from 170° to 23° and 0°C have al­
most exactly the same density profile .
The density of the uncut PMMA samples was observed to be small­
er than the density of the cut slices, as can be noted from Fig. 40.
The reason for this is not clear. A possible explanation is that the
slices, when cut, are not constrained any more by the material
304 Isayev

o
(O

Fig. 38 Lines of average density and gap wise density distribution in


0,38-Gm thick PS quenched from 170° to 23°C (line 1, o) and 0°C
(line 2, □ ), measured 24 hr after quenching.

adjacent to them and are, therefore, able to relax faster than they
can in the uncut state. But the same reasoning does not hold for
PS, in which case the measured and calculated values of the average
density were the same. Figure 41 shows the effect of thickness on
the density of quenched PMMA strips, from which it is seen that there
is a slight tendency toward a lower density in the strip of smaller
thickness.
Relaxation of density with time was found for both PS and PMMA.
However, for PS, even though the density of the quenched sample
was less than the annealed sample density, very little relaxation was
observed (density of the annealed sample for PS was 1.0481 g/cm^).
As shown in Fig. 42, a more pronounced relaxation of the density oc­
curs in the case of PMMA. A very steep gradient can be seen in the
layer near the surface at an elapsed time of 8.5 hr after quenching,
with the layer thickness being about 10% of that of the slab. With an
elapsed time of 24 hr, however, the interior density of the sample
tends to catch up with the surface density , resulting in a less steep
Orientation, Residual Stresses, and Volumetric Effects 305

o
(D

y/b

Fig. 39 Lines of average density and gapwise density distribution in


PS with thicknesses of 0.38 cm (line 1, o) and 0.25 cm (line 2 , □)
quenched from 170° to 23°C, measured 24 hr after quenching.

density gradient. In particular, the density gradient persists after


24 hr for the sample quenched from 170°C, whereas, when quenched
from 130°C, the gradient disappears after 24 hr.
The density distribution after quenching and the density variation
during aging have been modeled with a first-order rate equation based
on the free-volume approach. According to this theory, nonequilib­
rium free volume is frozen into the polymer as it is cooled through the
glass-transition temperature, the amount of free volume depending on
the rate of cooling. This frozen-in free volume determines the density
after quenching and also acts as the driving force toward the equilib­
rium volume at room temperature (a g in g ). The rate equation employed
is the one used by Rusch [156] :

Wn
(52)
dt
v / r Vj'^R
306 Isayev

o
VO

y/b

Fig. 40 Lines of average density and gap wise density distribution in


PMMA strip of 0.60-cm thickness quenched from 170® to 23®C (line 1,
0 ) , 170® to 0®C (line 2, □ ) , and 130® to 23®C (line 3, ) ; data taken
a

24 hr after quenching.

where Ve and V are the equilibrium and actual volumes at temperature


T, and the difference between them is the nonequilibrium free volume
wf. That is, the rate of change of volume is proportional to wf , which
is the driving force, divided by the relaxation time, which depends on
the temperature and the amount of free volume present. The relaxa­
tion time is obtained by multiplying t r , the relaxation time at a refer­
ence temperature T r , by a factor ax, Vf- The shift factor aT, V f
can be determined from a modified form of the WLF equation (51).
The equilibrium volume is computed from the following relations:

Vi - aL(Ti - T) T > T2
(53)
Vi - a^CTi - T 2 ) - - T) T < T2
Orientation, Residual Stresses, and Volumetric Effects 307

lO
0>

y/b

Fig. 41 Lines of average density and gapwise density distribution in


PMMA strips of thickness 0.60 cm (line 1, o) , and 0.28 cm (line 2, □ )
quenched from 170 to 0°C; data taken 24 hr after quenching.

where aL and aQ are the coefficients of volumetric expansion for the


rubbery and glassy states, respectively. Further, Ti is the initial
temperature, and Vi is the volume at Ti. In particiilar, Vi has been
taken to be 0.9950 g/cm^ (0.8685 g/cm^) for PS (PMMA) at Ti = 130°C.
The effective temperature, Te, is then defined by Eq. (51).
The spatial and temporal variations of temperature, T (y , t ) , dur­
ing quenching have been obtained by solving Eq. (41). The values of
thermal conductivity, k, and thermal diffusivity, a, for PS (PMMA)
have been taken to be 0.13 (0.19) J/(sec*m‘ K) and 6.13 x 10"® (1.095
X 10"^) m^/sec, respectively. The heat transfer coefficient h [1470
J/(sec*m^-K)] has been computed based on steady natural convection
for a vertical plate in water [137]. Solving the heat diffusion equa­
tion by using an explicit finite-difference procedure, we used the out­
put from this program as the input to the free-volume modeling.
The first-order rate equation (52) has been solved along with (51)
and (53) by using a fourth-order llunge-Kutta scheme with automatic
adjustment of time step (the time step used by the Runge-Kutta
308 Isayev

C7)

y/b

Fig. 42 Lines of average density and gap wise density distribution in


PMMA strip of 0.60-cm thickness, quenched from 170° to 23°C; data
taken 8.5 hr (line 1, o) and 24 hr (line 2, □ ) after quenching.

scheme was halved until the specified tolerance was satisfied). Since
the temperature values were available only at fixed intervals of time,
the temperature was assumed to vary linearly within any time step.
In the initial stages, i.e. , at high temperature, when the relaxation
time is very small, the time step in the Runge-Kutta scheme had to
be reduced to no more than half the relaxation time.
Values of aL, aG> and T 2 for PS have been calculated from the
P -V -T data of Hellmeyer and Menges [148] , resulting in = 0.00052
gy(cm^-K), aq = 0.00019 g/(cm^-K) and T 2 = 70°C. For PMMA, the
values have been taken from Ref. 156, namely, = 0.00047 g/(cm^*K),
aq = 0.00012 g/(cm^«K), and T 2 = 50°C. The time constant for PS was
calculated at 130°C from stress-relaxation data presented by Aklonis
and Tobolsky [138] , whereas the relaxation time at any temperature
has been taken to be the time at that temperature required for 3G(t)
(the shear-stress relaxation modulus) to reach a value of 1 0 ^ dyne/cm^
[160]. The resulting calculated value for of 0.0021 sec agrees well
with the relaxation time of 0.002 sec measured by Patterson et al. [161]
via photon-correlation spectroscopy. The relaxation time for PMMA
Orientation, Residual Stresses, and Volumetric Effects 309

has been taken to be 0 .008 sec, based upon the results pre­

sented by Patterson et al. [162] .


Values for the constants Ci and C 2 are given by Rusch [156] as
follows: Cl = 12.4 (16.4) and C 2 = 41°G (55'^C) for PS (PMMA) at a
reference temperature of 100°C (105®C). Since the present calcula­
tions are based upon a reference temperature of 130°C, the preceding
values have been appropriately converted as follows [163]:

Cl1 ^2
c
Ci =
(C , + T r - T r )
(54)
C, = C, 4- T r - T r

where ( ~ ) denotes values cited from Ref. 156 such that now Ci = 7.16
(11.28) and C 2 = 71°C (80°C) for PS (PMMA).
The prediction of the gapwise density variation in the PS slab at
different instants of time during quenching is shown in Fig. 43. In

y/b
Fig. 43 Theoretical gapwise distribution at 0 sec (line 1), 10 sec
(line 2)^ 100 sec (line 3), and 210 sec (line 4) during quenching
of PS strip of 0.38-cm thickness from 130^ to 23°C. (From Ref. 80.)
310 Isayev

y/b

Fig . 44 Same as Fig. 43 but for PMMA of 0.60-cm thickness at 0, 15,


150, 300 sec. (From Ref. 80.)

particular, the variation of density at the exterior is the fastest due


to the large cooling rate at the surface. Therefore, more free vol­
ume should be frozen-in at the surface. From the figure it is seen
that the center of the slab has a lower density than the surface. This
is due to the higher interior temperature. The model predicts that
the densities at all points in the slab eventually converge to the same
value at the end of quenching. Similar results can be seen for PMMA
(Fig. 44). Although not shown, the model predicts that the initial
temperature has essentially no effect and that the thickness has a
very small effect, with the density being slightly (about 0 . 1 %) higher
for the larger thickness for both PS and PMMA. These results are
to be expected, since at high temperatures the response of the ma­
terial will be almost instantaneous. Accordingly, free volume will be
frozen-in only at lower temperatures. This implies that the initial
temperature will not be a governing factor.
At this point we can make a qualitative assessment of the density
distribution in quenched slabs. Based on the free-volume approach
alone, one would expect a larger density at the center. However,
Orientation, Residual Stresses, and Volumetric Effects 311

the experimental results indicate just the reverse. The reason for
this is that thermal stresses arise in the slab during quenching; fu r­
ther, these stresses are nonuniform in the thickness direction. In
particular, it is well known (see Section III) that these stresses
are compressive at the surface and tensile in the interior of freely
quenched amorphous polymers. The density distribution is, there­
fore, due to the combined effects of these two factors. Accordingly,
the form of the final distribution will depend upon which is the domi­
nant factor. For free quenching then we can conclude, based upon
the present experimental results, that the residual stresses dominate
over the free-volume effect. That is, the compressive stresses at
the surface cause a larger density there.
To model aging, the density at the end of quenching has been used
as the starting density. Since there is no temperature variation, the
free volume is the only driving force. Figure 45 shows the experimen­
tal and theoretical results for the aging of PMMA strips 0.28 cm and
0.60 cm thick, respectively, after quenching from 130 to 23°C. These
results are for the gapwise-averaged density. The prediction is lower

Fig. 45 Predictions (curves) and experimental data (symbols) for


gap wise-averaged density variation with elapsed time after quench­
ing PMMA slabs of thicknesses 0.28 cm (line 1, A ) , and 0.60 cm
(line 2, □ ) from 130® to 23®C. (From Ref. 80.)
312 Isayev

than the experimental results in both cases. Moreover, the experi­


mental results show a faster density rise than the theoretical results.
For PS very little relaxation was observed, whereas the model p re­
dicts that the density should increase. In this regard the theory
does not take account of the presence of residual stresses. This is
believed to be the reason for the discrepancy between the experimen­
tal and theoretical results. Further, although the free-volume model
fails to predict the density variation during aging, this does not im­
ply that the model is incorrect. In particular, the average free vol­
ume will not be a good measure of mobility, particularly not for the
nonequilibrium states below Tg. However, the free-volume concept
is a valuable heuristic tool [151].
Now let us consider the effect of the processing condition upon
the density variation and the volumetric shrinkage in molded parts.
In these experiments a 50-ton BOY injection-molding machine has
been used. The cavity dimensions were 1 2 x 4 x 0.254 (cm^) with a
gate size of 0.5 x 0. 6 x o . l cm^. Figure 46 shows experimental re ­
sults from four injection-molded strips corresponding to different

X/L
Fig. 46 Dependence of gapwise-averaged density of PS on distance
from gate for To = 180°C ( □ ) , 200°C ( o ) , 220°C ( ) , and 240°C (X )
a

Q = 36 cm^/sec, Pp^ck ~ psi, = 30°C.


Orientation, Residual Stresses, and Volumetric Effects 313

injection temperatures (T q) with flow rate and packing pressure fixed.


In particular, it is seen that the density is almost constant in the
longitudinal direction for injection temperatures of 180°, 2 0 0 °, and
220°C. For T q = 240°C, however, the density near the gate is de­
cidedly higher than that at the other end. This can be explained on
the basis that at higher temperatures the polymer remains in the fluid
state for a longer time such that more material can be forced through
the gate. Meanwhile, at the other end the polymer would have cooled
and started to solidify. Due to a larger degree of packing, the den­
sity will therefore tend to be higher near the gate. At lower tem­
peratures this is not observed, apparently due to a lower degree of
fluidity at the lower temperatures. Similar results have been observed
at other flow rates, with the only difference being that a higher flow
rate causes the density to be higher. Further, the maximum differ­
ence between the curves for highest and lowest melt temperatures is
only about 0 . 1 %.
Figure 47 shows results of the effects of packing pressure. As
expected, a higher packing pressure causes increased densities by
forcing more material into the cavity. On the other hand, low pack­
ing pressures have very little or no effect on the density distribu­
tion. There is apparently a threshold pressure above which the

C\J

Fig. 47 Dependence of gap wise-averaged density of PS on distance


from gate at packing pressures of 200 psi ( ) , 500 psi ( □ ) and 1000
a

psi ( 0 ) ; To = 200°C, Q = 7.2 cm^ sec, T ^ = 30°C.


314 Isayev

CM
Tf

Fig. 48 Dependence of gap wise-averaged density of PS on distance


from gate at flow rates of 7.2 cm^/sec ( o ) , 36 cm^/sec ( □ ) , and 72
cm^/sec ( A ) ; To = 220^^0, Ppack = 200 psi, = 30®C.

effect of packing pressure will be significant. From the data ob­


tained it also appears that this threshold pressure might depend on
the flow rate—being higher at a higher flow rate or at a higher in­
jection temperature.
Data presented in Fig. 48 indicate that the flow rate does not sig­
nificantly affect the density distribution in injection-molded strips.
Therefore of those process parameters considered, the packing pres­
sure and injection temperature have the largest effect on the frozen-
in density.
As for the density distribution in injection-molded strips, the total
volumetric shrinkage for the part has also been plotted against pro­
cessing variables in Figs. 49, 50, and 51. The shrinkage in the in­
jection-molded strips has been defined as (V n ■ V s )/V n > where
is the nominal volume of the cavity and Vg is the actual volume of
the strip, as determined via the Archimedes principle. Since an ex­
act calculation of the nominal cavity volume is somewhat ambiguous,
the nominal volume has been taken to be the volume of the strip made
under processing conditions corresponding to (T q, Q, Ppack) =
(240°C, 72 cm^/sec, 1000 psi) such that the shrinkage for that case
is zero.
Figures 49-51 show that the shrinkage is mainly affected by the
temperature and packing pressure. In particular, the shrinkage
Orientation, Residual Stresses, and Volumetric Effects 315

Fig. 49 Measured volumetric shrinkage vs. injection melt temperature


in molded PS strips for Ppack ” rates of 7.2 cmV
sec ( 0 ) and 36 cm^/sec ( □ ) ; = 30°C.

Fig. 50 Measured volumetric shrinkage vs. packing pressure in


molded PS strips for T q, Q = 220®C, 72 cmVsec ( o ) , and 240°C,
7.2 cmVsec ( □ ) ; = 30°C.
316 Isayev

Fig. 51 Measured volumetric shrinkage vs. flow rate in molded PS


strips for To, Ppack = 2 2 0 ®C, 2 0 0 psi ( o ) , and 240°C, 1 0 0 0 psi ( □ ) ;
Tw^SO^^C.

decreases with increasing temperature. In this regard, although the


polymer cools through a larger temperature difference at higher tem­
peratures, the viscosity is lower such that the additional material
packed into the cavity more than offsets the contraction in volume
due to cooling. The largest shrinkage is seen to occur at the low­
est injection temperature, other conditions remaining the same.
An increase in the packing pressure causes the densities to in­
crease and the shrinkage to decrease. On the other hand, the flow
rate has almost no effect on shrinkage, as can be seen from Fig. 51.
An ability to predict the shrinkage for different processing con­
ditions would evidently be useful. Such an attempt has been made
at predicting the volume shrinkage for the processing conditions
used in the injection-molding experiments, thus allowing direct com­
parison with the measurements [80] . To this end, volume shrinkage
has been defined as
V.
1
Shrinkage (55)
V.
1

where v^ = initial specific volume (of melt) and Vf = specific volume at


room temperature.
Orientation, Residual Stresses, and Volumetric Effects 317

During injection molding, the polymer melt undergoes severe pres­


sure and temperature changes in a short time. Since the initial spe­
cific volume does not correspond to a uniform temperature and pres­
sure, it will have to be taken as some average volume. Accordingly,
Vi has been calculated via

(56)
pack

where the lower limit of integration corresponds to the end of filling,


whereas tpack denotes the time at which the pressure goes to zero.
To calculate the specific volume variation with time, i . e . , v ( t ) , the
temperature and pressure variations must be known as the polymer
melt cools. In particular, the temperature profile has been obtained
by solving Eq. (41), with the following conditions on T (y , t) :

(0, t) = 0 T (b , t) = T w
9y

The ’’initial” distribution, T (y , 0), corresponding to the end of fill­


ing, has been obtained by solving the cavity-filling simulation pro­
gram (see Chapter 1). In particular, the runner-cavity system has
been represented as a series of circular runners and an end-gated
rectangular strip. With T (y , t) so determined and p (t ) being mea­
sured at two locations within the cavity, the corresponding values
of v (t ) in (56) have been obtained by using P -V -T data for PS from
the literature [148] and averaging over the gap thickness. Hence
the value of vj in (56) has been evaluated by trapezoidal quadrature
at the streamwise locations corresponding to the two flush-mounted
pressure transducers (located 1.5 cm and 10.5 cm from the gate).
The resulting comparison with the measured total volumetric shrink­
age has then been done by averaging the foregoing calculated results
at the two transducer locations.
The results of this computation and experimental values have been
compared in Figs. 52, 53, and 54. Figure 52 shows that at low pack­
ing pressures, where the shrinkage is high, the prediction is quite
good. However, at higher pressures the prediction is larger than
the measured value. Concerning the effects of melt temperature
and flow rate, shown in Figs. 53 and 54, the predictions are con­
sistently higher than the experimental results but stñl in reason­
able agreement. For all three processing variables the trend fol­
lowed by the predicted shrinkage is seen to be the same as that of
the measurements. In particular, shrinkage is mainly affected by
packing pressure, with injection temperature and flow rate having
relatively small effects.
318 Isayev

Fig. 52 Measured ( o ) and predicted (solid line) volumetric shrink­


age vs. packing pressure in molded PS strips; T q = 220°C, Q = 72
cm^/sec, Tw = 30°C. (From Ref. 80.)

Fig. 53 Measured (o ) and predicted (solid line) volumetric shrink­


age vs. melt temperature in molded PS strips; Ppack ” Q ~
36 cm^/sec, = 30°C. (From Ref. 80.)
Orientation, Residual Stresses, and Volumetric Effects 319

Fig. 54 Measured ( o ) and predicted (solid line) volumetric shrinkage


vs, flow rate in molded PS strips; T q = 2 0 0 ®C, Ppack = 500 psi, =
30®C. (From Ref. 80.)

V. CONCLUDING REMARKS

In this chapter theoretical and experimental aspects of the orientation,


density, shrinkage, and residual-stress development in the injection
molding of polymers have been considered, including effects of pro­
cessing conditions, material, and cavity thickness. Since these re­
sidual stresses appear as coupled flow- and cooling-induced stresses,
both injection-molding and quenching literature have been reviewed
in order to elucidate the respective contributions of thermal and flow
orientations and stresses in the overall orientation and residual-stress
development and to make comparisons with the corresponding theoreti­
cal predictions.
Experimental techniques suitable for measuring orientation and re­
sidual stresses in molded plastic parts have been critically reviewed.
Birefringence, trifringence, FTIR, x-ray diffraction, and heat-shrink-
age methods have been discussed for measuring the orientation in
molded parts. Among methods for residual-stress measurements are
hole drilling, relaxation, photoelasticity, and layer removal. The last
method is simple to implement and gives the most extensive information
about the residual-stress field. Although this method needs to be
320 Isayev

further formulated for viscoelastic materials, the approach based on


pure elasticity has been employed in most cases where thermal stress­
es in quenched slabs and residual stresses in molded slabs have been
measured.
For those polymers for which the stress-optical coefficient in the
glassy state is much smaller than in the rubbery and fluid states, the
level of flow-induced stresses can be derived from measurements of
frozen-in birefringence by the stress-optical rule. The resulting dif­
ference between the principal frozen-in flow stresses in molded plastic
slabs typically does not exceed 2 MPa and is tensile in nature. These
flow stresses are due to the development of shear and normal stresses
in viscoelastic polymeric melts and can be predicted by using visco­
elastic constitutive equations. In particular, the fluid and rubbery
states of the polymeric melt mainly contribute to these flow stresses
and orientation. An increase in melt temperature and cavity thick­
ness and a decrease in packing pressure significantly decrease such
frozen-in flow stresses and orientation. In contrast, an increase in
mold temperature was found to have little effect on the flow stresses
and orientation. On the other hand, the level of compressive thermal
stresses determined in quenched slabs by the layer-removal technique
has typically been found to be about 10 MPa.
Calculation of the thermal stresses on the basis of the instant-
freezing elastic theory of Indenbom and the viscoelastic theory by
Lee, Rogers, and Woo have been performed and compared with ex­
perimental data. These comparisons show that the theories of Inden­
bom satisfactorily predict the experimental data for the gap wise dis­
tribution of thermal stresses in quenched PMMA slabs. However, for
quenched PS slabs both theories systematically overpredict the ex­
perimentally measured stresses. This latter discrepancy is apparently
due to the neglect of nonlinear polymeric behavior during quenching
through the glass-transition temperature and to the phenomenon of
thermal-stress relaxation following quenching. In particular, it has
been found that within a month the stresses might reduce to one-half
their initial values. In addition, two representations of the relaxation
modulus necessary for calculation of the thermal stresses according to
the viscoelastic theory have been used. In one case the relaxation
modulus data covering the fluid, rubbery , and glassy states of poly­
mer has been used, and in the second case the relaxation modulus has
been based upon data covering only the glassy state and the transi­
tion zone from the glassy to rubbery states. Since both representa­
tions give the same predicted thermal-stress distribution, it follows
that the glassy state and the transition zone are mainly responsible
for the thermal stresses. Moreover , it is well known that the relaxa­
tion modulus in the glassy state and transition zone are independent
Orientation, Residual Stresses, and Volumetric Effects 321

of molecular weight and molecular-weight distribution for polymers


having molecular weight larger than critical molecular weight cor­
responding to formation of the rubbery plateau. Thus, the ther­
mal stresses wül also be independent of these molecular character­
istics of polymers and will only depend on the nature of the poly­
mer chain.
Unlike thermal stresses and birefringence, which are the same in
two mutually perpendicular directions, residual stresses and birefrin­
gence in molded slabs have been found to depend on orientation rela­
tive to the flow direction during cavity filling and on the distance from
the gate. Although gap wise distributions of thermal stresses are para­
bolic in nature with compressive and tensüe stresses at the surface
and interior, respectively, residual stresses in molded slabs are, gen­
erally, nonparabolic. Characteristics of their gapwise distribution de­
pend on processing conditions. Increasing the flow rate reduces the
level of orientation in molded products. On the other hand, within
the range considered the injection rate and mold temperature have
been found to have only a minor effect on the level and gapwise dis­
tribution of molded stresses, whereas melt temperature has a signifi­
cant effect. Lowering the melt temperature increases the level of or­
ientation and reduces the level of compressive stress or even results
in a tensile stress at the surface of the molded part. The maximum
level of residual stress is typically less than 2 MPa. This value is
comparable with the level of flow stresses in molded parts but is much
lower than the level of thermal stress in quenched samples. This
means that the cooling procedure in quenching is more severe than
that in injection molding. However, remember that molding stresses
were obtained on molded parts aged over one year after molding. Def­
initely, during this period, a residual-stress relaxation may occur in
molded parts essentially simñar to that observed for thermal stresses
in quenched slabs. That is, one would expect somewhat larger resid­
ual stresses in fresh molded samples than that measured in the pres­
ent work. However, this requires further investigation.
Density distribution in freely quenched slabs of PS and PMMA and
molded strips of PS has been reported. The effects of the initial and
bath temperatures and of sample thickness have been investigated for
quenching, whereas the effects of injection temperature, flow rate,
and packing pressure have been studied for molded strips. In quench­
ing, large density gradients were observed in a layer near the sur­
face, the layer thickness being about 25% of the slab thickness for PS
and 10% for PMMA. Density at the surface is higher than at the cen­
ter. This is in contrast to what one would expect based on the free-
volume concept alone, and it has been attributed to the formation of
of compressive residual stresses on the slab surface during quenching.
322 I sayev

The total variation in density from the center to the surface is only
about 0.1% for PS and 0.25% for PMMA. Increasing the bath tempera­
ture or slab thickness increases the final density, whereas the initial
temperature has essentially no effect. Relaxation of density with time
has been monitored for PS and PMMA. Very little change in density
with time has been observed for PS, whereas the density of PMMA in­
creases considerably during aging.
Modeling the density has been done by using a first-order rate
theory for specific volume in conjunction with solving the transient
one-dimensional heat cohduction equation with a convective heat trans­
fer boundary condition at the surface. In contrast to experiment,
this simple modeling shows no gapwise distribution of density in the
quenched strips at large times because the modeling has not accounted
for the development of residual stresses during quenching. On the
other hand , the modeling has been found capable of describing an
experimentally observed increasing density during the aging of PMMA
strips.
Measurements of density and volumetric shrinkage have been car­
ried out for molded PS strips. Increasing the packing pressure and
melt temperature increase the density and decrease shrinkage, with
the flow rate having little effect. Maximal variation in the measured
density and shrinkage have been 0.12% and 6 %respectively. Model­
ing of shrinkage has been carried out by using the equation of state
for specific volume and by simultaneously solving the governing equa­
tions for one-dimensional mold filling followed by the one-dimensional
transient conduction equation during the packing and cooling stages.
Predicted and experimental results for shrinkage have been found to
be in quantitative agreement.

ACKNOWLEDGMENT

This work has been initiated as part of the Cornell Injection Molding
Program, which is supported by the NSF and by Industrial Consor­
tium . It was completed after the author joined The University of
Akron. Thanks are due to my co-workers over the years. Dr. K. K.
Wang, Dr. S. F. Shen, and Dr. C. A. Hieber, Mr. D. L. Crouthamel,
and Mr. T. Hariharan.

REFERENCES

1. S. S. Kattiand J. M. Schultz, Polym. Eng, Sci., 22: 16 (1982).


2. J. L. White, Pure Appi, Chem, , 55: 765 (1983).
3. M. Fleibner, Kunststoffe, 63: 597 (1973).
Orientation, Residual Stresses, and Volumetric Effects 323

4. A. I. Isayev, Polym. Eng, Sci., 23: 271 (1983).


5. R. J. Samuels, J. Appl, Polym, Sci,, 26: 1383 (1981).
6 . R. J. Samuels, Polym, Eng, Sci,, 23: 257 (1983).
7. G. L. Wilkes, Adv, Polym, Sci,, 8: 91 (1971).
8 . Z. Mencik and A. J. Chompff, J, Polym, Sci,, Phys, Ed,, 12:
977 (1974).
9. M. R. Kantz, N. D. Newman, Jr.^ and F. H. Stigale, J, Appl,
Polym, Set , 16: 1249 (1972).
10. E. S. Clark, AppZ. Polym, Symposia, 20: 325 (1973); 24: 45
(1974).
11. J. C. Vyvoda, M. Gilbert, and D. A. Hemsley, Polymer, 21: 109
(1980).
12. J, L. White and J. E. Spruiell, Polym, Eng, Sci,, 23: 247
(1983).
13. G. Menges and G. Wubken, SPE Tech, Papers, 79: 519 (1973).
14. N. Heron, S. Pedersen, and L. L. Chapoy, Rheol, Acta, 15: 379
(1976).
15. M. Gilbert, D. E. Marshall, J. C. Vyvoda, and C. J. Copsey,
Plast, Rubber Process,, Sept.: 96 (1979).
16. H. Janeschitz-Kriegl, Polymer Melt Rheology and Flow Birefrin­
gence, Springer-Verlag, New York, 1983.
17. A. S. Lodge, Nature, 176: 838 (1955).
18. W. Philippoff, Nature, 178: 811 (1956).
19. H. Janeschitz-Kriegl, Adv, Polym, Sci,, 6: 170 (1969).
20. F. H. Gortemaker, M. G. Hansen, B. de Cindio, H. M. Laun,
and H. Janeschitz-Kriegl, Rheol, Acta, 15: 287 (1976).
21. G. V, Vinogradov, A. I. Isayev, D. A. Mustafayev, and Y. Y.
Podolsky, J, Appl, Polym, Sci,, 22: 665 (1978).
22. L. R, G. Treloar, The Physics of Rubber Elasticity, 3rd ed..
Clarendon Press, Oxford, 1975.
23. J. L. S. Wales, The Application of Flow Birefringence to Rheo­
logical Studies of Polymer Melts, Delft University Press, 1976.
24. H. Janeschitz-Kriegl, Rheol, Acta, 16: 327 (1977).
25. W. Dietz and J. L. White, Rheol, Acta, 17:676 (1978).
26. J. Greener and G. Pearson, J, Rheol,, 27: 115 (1983).
27. G. Marucci, T. Titomanlio, and G. C. Santi, Rheol, Acta, 12:
269(1973).
28. A. I. Isayev and C. A. Hieber, Rheol, Acta, 19: 168 (1980).
29. A. I. Leonov, Rheol, Acta, 15: 85 (1976).
30. R. K. Upadhyay, A. I. Isayev, and S. F.Shen, Rheol, Acta,
20: 443 (1981).
31. R. K. Upadhyay, A. I. Isayev, and S. F. Shen, J, Rheol,, 27:
155(1983).
32. R. K. Upadhyay and A. I. Isayev, Rheol, Acta, 22: 557 (1983).
33. P. G. Lafleur and M. R. Kamal, Polym, Eng, Sci,, 26: 92 (1986).
324 Isayev

34. M. R. Kamal and P. G. Lafleur, Polym. Eng. Sei., 26: 103


(1986).
35. M. R. Kamal and V. Tan, Polym. Eng. Sei., 19: 558 (1979).
36. L. J. Broutmann and J. M. Krishnakumar, Polym. Eng. Sei.,
16: 74 (1976).
37. J. R. Saffell and A. H. Windle, J. Appl. Polym. Sei., 25: 1117
(1980).
38. J. Greener and P. M. Kenyon, Proceedings of the Annual Meet­
ing of SPIE, San Diego, 1981.
39. C. Wust an d D . C. Bogue, J. Appl. Polym. Sei., 28: 1931 (1983).
40. N. J. Mills, J. Mater. Sei., 17: 558 (1982).
41. T. F. Rudd and E. F. Gurnee, J. Appl. P h ys .,28: 1096 (1957).
42. R. S. Stein, in Die Physik Der Hoehpolymeren,Vol. 4, H. A.
Stuart, ed.. Springer-Verlag, Berlin, 1956.
43. H. Peukert, Kunststoffe, 41: 154 (1951).
44. M. M. Qayym and J. R. White, J. Appl. Polym. Sei., 28: 203
(1983).
45. M. M. Qayym and J, R. White, Polymer, 23: 129(1982).
46. S. M. Lee, J. De La Vega, and D. C. Bogue, J.Appl. Polym.
Set , 31: 2791 (1986).
47. W. Retting, Colloid Polym. Sei., 257: 689 (1979).
48. R. G. Treuting and W. T. Read, J r., J. Appl. Phys., 22: 130
(1951).
49. E. G. Gurnee, L. T. Patterson, and R. D. Andrews, J. Appl.
Phys., 26: 1106 (1955).
50. B. E. Read, Polymer, 3: 143 (1962).
51. B. E. Read, Polymer, 5: 1 (1964).
52. R. S. Stein, S. Onogi, and D. A. Keedy, J. Polym. Sei., 57:
801 (1962).
53. T. Kyu, S. Suehiro, S. Nomura, and H. Kawai, J. Polym. Sei.,
Phys. Ed., 18: 951 (1980).
54. L. S. Priss, I. I. Vishnuakov, and I. P. Pavlova, Intern. J.
Polym. Mater., 8: 85 (1980).
55. I. V. Yannas and M. J. Doyle, J. Polym. Sei. A-2, 10: 159
(1972).
56. P. S. Theocaris, J. Appl. Polym. Sei., 8: 399 (1964).
57. Y. Fukui, T. Sato, M. Ushirokawa, T. Asada, and S. Onogi,
J. Polym. Sei. A-2, 8: 1195 (1970).
58. A. I. Isayev a n d D . L. Crouthamel, Polym. Plast., Teehnol.
Eng., 22: 177 (1984).
59. L. D. Coxon and J. R. White, Polym. Eng. Sei., 20: 230 (1980).
60. R. S. Spencer and G. D. Gilmore, Modern Plasties, 28: 97
(1950),
61. R. L. Ballman and H. L. Toor, Modern Plasties, 38: 113 (1960).
62. L. Hoare a n d D . Hull, Polym. Eng. Sei., 17: 204 (1977).
Orientation, Residual Stresses, and Volumetric Effects 325

63. Z. Bakerdjian and M. R. Kamal, Polym. Eng, Sci,, 17: 96 (1977).


64. J. L. S. Wales, T r. J. van Leeuwen, and R. van der Vijgh,
Polym, Eng, Sci., 12: 358 (1972).
65. J. L. White and W. J. N. B. Dee, Polym, Eng, Sci., 14: 212
(1974).
6 6 . L. R. Schmidt, Polym, Eng, Sci,, 14: 797 (1974).
67. M. R. Kamal and S. Kenig, Polym, Eng, Sci,, 12: 294 (1972).
6 8 . L. R. Schmidt, Polym, Eng, Sci,, 17: 667 (1977).
69. K. Oda, J. L. White, and E. S. CLark, Polym, Eng, Sci,, 16:
585 (1976).
70. C. D. Han and C. A. Villamizar, Polym, Eng, Sci,, 18: 173
(1978) .
71. A. I. Isayev, J, Polym, Sci,, Phys, Ed,, 11: 2123 (1973).
72. M. Fujiyama and K. Azuma, J. Appl, Polym, Sci,, 23: 2807
(1979) .
73. Z. Tadmor, J. Appl, Polym, Sci,, 18: 1753 (1974).
74. Z. Tadmor and C. Gogos, Principles of Polymer Processing,
Wiley, New York, 1979.
75. T. R. Fields and D. C. Bogue, Trans, Soc, Rheol,, 12: 39
(1968).
76. C. D. Han, Rheology in Polymer Processing, Academic Press,
New York, 1976.
77. V. I. Brizitsky, G. V. Vinogradov, A. I. Isayev, and Yu. Ya.
Podolsky, J, Appl, Polym, Sci,, 22: 751 (1978).
78. A. I. Isayev and R. K. Upadhyay, J, Non-Newt, Fluid Mech,,
19: 135 (1985).
79. Y. T. Koita, Polym, Eng, Sci,, 14: 840 (1974).
80. A . I. Isayev and T. Hariharan, Polym, Eng, Sci,, 25: 271 (1985).
81. J. Mathar, Trans. ASME, 56: 249 (1934).
82. R. A. Kelsey, Proc, Soc, Exp, Stress Anal,, 14: 181 (1956).
83. N. J. R e n d le ra n d l. Vigness, Exp, Mech,, 6: 577 (1966).
84. A. J. Bush and F. J. Kromer, ISA Trans,, 12: 249 (1973).
85. K. Ito, Japan Plastics Age, 15: 36 (1977).
8 6 . L. Schmidt, J. Opfermann, and G. Menges, Polym, Eng, Reviews,
1: 1 (1981).
87. J. C. M. Li, Cañad, J, Phys,, 45: 493 (1967).
8 8 . J. Kubat and M. Rigdahl, Intern, J, Polym, Mater,, 2: 287
(1981).
89. B. Haworth, C. S. Hindle, G. J. Sandilands, and J. R. White,
Plastics and Rubber Processing and Applications, 2: 59 (1982).
90. G. J. Sandilands and J. R. White, Polymer, 21: 338 (1980).
91. B. Haworth and J. R. White, J, Mater, Sci,, 16: 3263 (1981).
92. B. Haworth, G. J. Sandilands, and J. R. White, Plast, Rub,
Int,, 5: 109 (1980).
93. J. Kubat and M. Rigdahl, Polym, Eng. Sci,, 16: 792 (1976).
326 Isayev

94. J. Kubat and M. Rigdajil, Polymery 16: 925 (1975).


95. J. Kubat, J. Petermann, and M. Rigdahl, Mater, Sei. Eng.,
19: 185 (1975).
96. J. Kubat, J. Petermann, and M. Rigdahl, J. Mater. Sei., 10:
2071 (1975).
97. J. Kubat, M. Rigdahl, and R. Seiden, J. Appl. Polym. Sei.,
20: 2799 (1976).
98. J. Kubat, R. Seiden, and M, Rigdahl, J. Appl. Polym. Sei.,
22: 1715 (1978).
99. M. M. Frocht, Photoelastieity, Wiley, New York, 1948.
100. A. J. Durelli and W. F. Riley, Introduetion to Photomeehanies,
Prentice-Hall, Englewood Cliffs, N .J ., 1965.
101. A. Kuske and G. Robertson, Photoelastie Stress Analysis, Wiley,
New York, 1974.
102. R. K. Mittal and V. Rashmi, Polym. Eng. Sei., 26: 310 (1986).
103. G. B. Jackson and R. L. Balman, SPE Teeh. Papers, 6: 18-1
(1960).
104. H. Keskkula, G. M. Simpson, and F. L. Dicken, SPE Teeh.
Papers, 12: 15-2 (1966).
105. A. Siegmann, A. Buchman, and S. Kenig, Polym. Eng. Sei.,
21: 997 (1981).
106. A. Siegmann, A. Buchman, and S. Kenig, J. Mater. Sei., 16:
3516(1981).
107. N. Kamal and F. Moy, in Rheology, Vol. 3, G. Astarita, G.
Marucci, and L. Nicolais, Eds., Plenum Press, New York,
1980, p. 143.
108. B. S. Thakkar and L. J. Broutman, SPE Teeh. Papers, 25:
554(1979).
109. A. Siegman, M. Narkis, a n d N . Rosenzweig, Polym. Eng. Sei.,
19: 223 (1979).
110. P. So and L. J. Broutman, Polym. Eng. Sei., 16: 785 (1976).
111. T. T. Tee and C. Y. Yap, SPE Teeh. Papers, 24: 361 (1978).
112. S. E. Meacham, SPE Teeh. Papers, 24: 808 (1978).
113. B. S. Thakkar and L. J. Broutman, Polym. Eng. Sei., 20:
1214(1980).
114. D. P. Russell and P. W. R. Beaumont, J. Mater. Sei., 15: 216
(1980).
115. J. F. Mandell, K. L. Smith, and D. D. Huang, Polym. Eng.
Sei., 21: 1173 (1981).
116. G. Menges, A. Dierkes, L. Schmidt, and E. Winkel, SPE Teeh.
Papers, 26: 300 (1980).
117. M. St. Jacques, Polym. Eng. Sei., 22: 241 (1982).
118. C. S. Hin die, J. R. White, D. Dawson, W. J. Greenwood, and
K. Thomas, SPE Teeh. Papers, 27: 783 (1981).
119. L. C. E. Struik, Polym. Eng. S e t , 19: 223 (1979).
Orientation, Residual Stresses, and Volumetric Effects 327

120. A. I. Isayev, C. A. Hieber, and D. L. Crouthamel, SPE


Tech, Papers, 27: 110 (1981).
121. M. R. Kamal and V. Tan, SPE Tech. Papers, 24: 121 (1978).
122. A. Siegmann, A. Buchman, and S. Kenig, Polym. Eng. Sci.,
22: 40 (1982).
123. B . D. Aggarwala and E. Saibel, Phys. Chem. Glasses, 2: 137
(1961).
124. E. H. Lee, T. G. Rogers, and T. C. Woo, J. Amer. Cer.
Soc., 48: 480 (1965).
125. G. M. Bartenev, J. Tech. Phys., 18: 383 (1948).
126. G. M. Bartenev, J. Tech. Phys., 19: 1423 (1949).
127. R. Gardon, Proceeding Vllth Intern. Congress on Glass,
Brussels, Belgium, paper 79.
128. V. L. Indenbom, J. Tech. Phys., 24: 925 (1954).
129. V. L. Indenbom, Fizika Tverdogo Tela, Sbornik Statey, 1:
236(1959).
130. R. Muki and E. Sternberg, J. Appl. Mech., 28: 193 (1961).
131. E. H. Lee and T. G. Rogers, J. Appl. Mech., 30: 127 (1963).
132. O, S, Narayanaswamy and R, Gardon, J. Amer. Cer. Soc.,
52: 554 (1969),
133. O, S, Narayanaswamy, J. Amer. Cer. Soc., 61: 146 (1978),
134. A, Q, Tool, J. Amer. Cer. Soc., 29: 240 (1946),
135. S. M. Ohlberg and T. C. Woo, Rheol. Acta, 12: 345 (1973).
136. M. Rigdahl, Intern. J. Polym. Mater., 5: 43 (1976).
137. F. Kreith, Principles of Heat Transfer, 2nd e d.. International
Textbook Co., 1965.
138. J. J. Aklonis and A. V. Tobolsky, J. Appl. Phys., 36: 3483
(1965).
139. I. V. Yannas, J. Macromol. Sci., Phys,, B6: 91 (1972).
140. A. I. Isayev and D. Katz, Intern. J. Polym. Mater., 8: 25
(1980).
141. Y. Weitsman, J. Appl. Mech., 47, 1 (1980).
142. L. D. Coxon and J. R. White, J. Mater. Sci., 14: 1114 (1979).
143. L. D. Coxon and J. R. White, Polym. Eng. Sci., 20: 230
(1980).
144. D. P. Russell and P. W. R. Beaumont, J. Mater. Sci., 15: 208
(1980).
145. A. Siegmann, A . Buchman, and S. Kenig, Polym. Eng. Sci.,
22: 560 (1982).
146. J. L. White and W. Dietz, Polym. Eng. Sci., 19: 1081 (1979).
147. B. Miller, Plastics World, July: 71 (1979).
148. H .-O . Hellmeyer and G. Menges, SPE Tech. Papers, 22: 386
(1976).
149. F. H. Moy and M. R. Kamal, SPE Tech. Papers, 25: 108 (1979).
150. R. Gardon, J, Amer. Cer. Soc., 61: 143 (1978).
328 Isayev

151. L. C. E. Struik, Physical Aging in Amorphous Polymers and


Other Materials, Elsevier, New York, 1978.
152. M. R. Tant and G. L. Wilkes, Polym, Engr. Sci., 21: 874 (1981).
153. T. A^frey, G. Goldfinger, and H. Mark, J. AppL Phys., 14:
700 (1943).
154. H. Eyring, J. Chem. Phys., 4: 283 (1936).
155. D. Turnbull and M. H. Cohen, J. Chem. Phys., 34: 120 (1961).
156. K. C. Rusch, J. Macromol. Sci., B2: 179 (1968).
157. A. J. Kovacs, Fortschr. Hoschpolym. Forsch., 3: 394 (1963).
158. A. J. Kovacs, J. J. Aklonis, J. M. Hutchinson, and A . R.
Ramos, J. Polym. Sci., Phys. Ed., 17: 1097 (1979).
159. M. L .. Williams, R. F. Landel, and J. D. Ferry, J. Am, Chem.
Soc., 77: 3701 (1955).
160. A. V. Tobolsky, Properties and Structure of Polymers, Wiley,
New York, 1962, pp. 160-166.
161. G. D. Patterson, P. J. Carroll, and J. R. Stevens, J. Poly.
Sci., Phys. Ed., 21: 605 (1983).
162. G. D. Patterson, P. J. Carroll, and J. R. Stevens, J. Poly.
Sci., Phys. Ed., 21: 613 (1983).
163. J. D. Ferry, Viscoelastic Properties of Polymers, Wiley, New
York, 1971.
4
Thermoset Injection Molding

Musa R . Kamal Michael E. Ryan


Department of Chemical Department of Chemical Engineering
Engineering State University of New York at
McGill University Buffalo
Montreal, Canada Buffalo, New'York

1. IN T R O D U C T IO N

Thermosetting plastics are distinguished by the fact that they are ca­
pable of forming an infusible cross-linked or network structure by an
irreversible chemical reaction. As a consequence, thermosets inher­
ently possess excellent mechanical and heat-resistance properties.
Their performance characteristics at elevated temperatures are gen­
erally superior to those of thermoplastic polymers. Despite these pos­
itive attributes, thermosets currently occupy about 18% of the plas­
tics market in the United States [1 ]. Undoubtedly, difficulties asso­
ciated with the processability of these materials have limited their
penetration in many conventional markets.
Thermosets are generally processed in a molding operation, such
as compression, transfer, or injection molding, although other pro­
cesses such as extrusion, coating, and pultrusion are receiving in­
creased interest and attention. Injection molding is an automated
process capable of producing close tolerance, three-dimensional parts
in high volume. These desirable features often account for the selec­
tion of injection molding as the preferred process for the manufacture
of a thermoset part.
During the past decade, the technology of reaction injection mold­
ing has been successfully developed for a variety of commercial ap­
plications. Reaction injection molding is characterized by in-mold
polymerization from monomeric or oligomeric components by a fast
polymerization reaction. Due to the lower capital investment and less

329
330 Kamal and Ryan

energy-intensive nature of the process, reaction injection molding has


received considerable attention and is discussed elsewhere. The fo­
cus of our treatment is directed toward the injection molding of ther­
mosetting materials which are solid or highly viscous under ambient
conditions and which are frequently highly filled.
An accurate quantitative understanding of both material behavior
and process history is required in order to enhance the design, op­
eration, control, productivity, and cost effectiveness of the thermo­
set injection-molding process. The following sections describe the
theoretical considerations and experimental methodology relating to
the characterization of reaction kinetics, rheological behavior, and
thermal properties of thermosetting systems. The mathematical mod­
eling of the thermoset injection-molding process is considered in de­
tail, and comparisons with experimental data are provided and dis­
cussed.

II. THERMOSET RESIN C H A R A C T E R IZ A T IO N

The reactive nature of thermosetting systems necessitates a basic un­


derstanding of the kinetics governing the cure reaction. A knowledge
of the heat of reaction and variation of the degree cure with the ther­
mal history of the material is required in order to interpret the struc­
tural and physical property changes that accompany the cross-linking
reaction. In addition, the analysis, optimization, and control of ther­
moset molding operations require reliable information about the ther­
mal and rheological behavior of the material. Thus, the technical as­
sessment and economic evaluation of the injection molding of a thermo-
set part is critically dependent on accurate kinetics and physical prop­
erty data.

A. Kinetic C haracterizatio n

Kinetic characterization of the curing reaction has long been recog­


nized as an essential element for the mathematical modeling of the in­
jection-molding process [2 ,3 ]. Since most thermosetting reactions
are exothermic, a knowledge of the rate of heat generation, as a func­
tion of the time and temperature history experienced by the material
during processing, is necessary for the proper quantitative determin­
ation of the heat transfer requirements. Since dramatic increases in
the viscosity of the material can occur at some critical conversion, a
reliable predictive capability of the extent of reaction is required in
order to ascertain if a mold can be filled under an imposed set of op­
erating conditions.
Thermoset Injection Molding 331

E x p e rim e n ta l T e ch n iq u es

A variety of experimental methods have been employed for monitoring


or characterizing thermoset cure reactions. The various techniques
are briefly described along with their relative merits and limitations.
Chemical Analysis
Thermosetting polymers are generally synthesized from reactants
containing multifunctional species possessing a functionality larger
than 2. A quantitative measure of the extent of the cure reaction
may be obtained by monitoring the disappearance of a reactive group
or species or the appearance of a reaction product. Titration or
chromatographic, spectroscopic, or chemical analyses may be applied
in order to follow the conversion of the reaction species. Occasion­
ally, the analysis is performed on the hydrolytic or pyrolytic degra­
dation products of the cross-linked polymer. Apart from the iden­
tification of a suitable procedure, the main difficulty with chemical
analysis is that the measurements are generally tedious and time con­
suming .

Sol-Gel Analysis
Un-cross-linked polymers may be dissolved in a suitably chosen
solvent. The extent of swelling of the cross-linked network may be
used to ascertain the gel fraction and extent of cure of the system.
The point of incipient formation of a three-dimensional network struc­
ture (gelation point) may be determined by measuring the gel content
as a function of time and extrapolating to zero gel fraction. Equilib­
rium swelling measurements are also tedious and time consuming and
are generally limited to low degrees of cross-linking. At high cure
levels the system may vitrify, or prolonged exposure times to the sol­
vent may become necessary.
Infrared Spectrometry
Infrared spectrometry can be used to follow a cure reaction by mon­
itoring the change in peak intensity at some specified frequency. The
selection of an appropriate band depends on good isolation or minimal
interference from other bands and reasonable intensity. Overlapping
bands in the spectra and low intensities limit the accuracy and resolu­
tion and render interpretation of the data exceedingly difficult. From
a knowledge of the chemistry or mechanism of the curing reaction,
changes in the infrared spectra can be related to the appearance or
disappearance of particular groups. For thermosets very thin sam­
ples are generally employed. Different band choices may be prefer­
able, depending on the form and manner of preparation of the sample.
Fourier transform infrared analysis (FTIR ) permits rapid analysis of
the entire infrared spectrum.
332 Kama! and Ryan

Spectroscopic Methods
Various spectroscopic techniques (NMR, ESR, EPR, Raman, etc.)
have been devised to serve as analytical tools for the determination
of compositional or configurational changes in a polymer specimen. In
general, these techniques are based on the detection of discrete fre ­
quencies where energy absorption occurs. The different methods cor­
respond to the different frequency ranges used. In the context of
polymer analysis, many of these techniques are relatively new and
can hardly be regarded as simple and routine.
Physical Property Changes
The curing reaction of a thermosetting polymer is often accompanied
by changes in a variety of physical properties such as density, elec­
trical conductivity, refractive index, viscosity, and dynamic mechani­
cal properties. Certain physical properties (e . g . , viscosity) or ex­
perimental methods (e . g . , dilatometry) are more suitable for thermo­
setting systems prior to gelation, whereas others (e . g . , dynamic mech­
anical properties) may be more suitable for samples in the post-gela­
tion phase. The application and assessment of the relative merits and
limitations of each of these techniques have been extensively discussed
in the literature. The attraction of these methods is that they can di­
rectly provide a continuous measure of the progress of the cross-link­
ing reaction. However, the accuracy, reliability, and sensitivity of
the measurements may be deleteriously influenced or limited by fac­
tors such as impurities, moisture content, and testing rate.
Differential Thermal Analysis
The study of reaction kinetics by differential thermal analysis in­
volves the comparison of the temperature at the center of the thermo­
set sample with the temperature of an inert reference material sub­
jected to the same heating history. For an exothermic reaction, the
sample temperature will somewhat exceed the reference temperature.
This differential temperature is indirectly related to the exothermic
heat of reaction of the sample and consequently to the reaction kinet­
ics of the system. The sample and reference are usually heated at a
uniform rate.
Differential Scanning Calorimetry
Differential scanning calorimetry is based on the principle of dif­
ferential enthalpic analysis and involves providing heat to the sample
and inert references at a varying rate to maintain their temperatures
equal. The calorimeter may be employed in either an isothermal or
nonisothermal (scanning) mode. Thus, the sample temperature is held
constant in the isothermal mode or varied linearly with time in the non­
isothermal mode. The endothermic or exothermic heats of reaction can
therefore be directly measured as functions of time or sample tempera­
ture, depending on the mode of operation. By assuming that the rate
Thermoset Injection Molding 333

of heat evolution is directly proportional to the reaction rate, one can


use calorimetry to provide information about the kinetic behavior of a
thermosetting system. Information can be simultaneously obtained
about physical transitions and thermal properties. If volatile compo­
nents are liberated during the course of the cross-linking reaction,
the accuracy and interpretation of results in a kinetic sense may be
obscured. In some cases, the problem can be suppressed by operating
the calorimeter cell at an elevated pressure. Differential scanning
calorimetry has been widely used for characterizing thermosetting re­
actions because it is simple and easy to use.

Iso th e rm a l S t u d i e s : K in e tic M odels

As mentioned in the previous section, the differential scanning calor­


imeter provides a convenient tool for obtaining kinetic data under iso­
thermal conditions. In this mode of operation a small sample is intro­
duced into the calorimeter cell^ which is continuously maintained at a
constant temperature. For an exothermic cross-linking reaction the
rate of heat generation is measured as a function of time. Typical
isothermal data are shown in Fig. 1 [4] for an epoxy resin (Dow, DER-
332) cured stoichiometrically with m-phenylene diamine (A ld rich ).
These data were obtained from a Perkin-Elmer differential scanning
calorimeter (Model DSC -1). As the isothermal cure temperature is
increased, the maximum value of heat evolution rate increases and
occurs at shorter times. However, the duration of measurable cure
is more protracted at the lower temperatures. Similar cure exotherms
have been reported for a variety of other thermosetting systems. The
cumulative heat generated during the cross-linking reaction may be
obtained from numerical integration of the rate data. Figure 2 shows
the cumulative heat evolved as a function of time corresponding to
the isothermal rate data shown in Fig. 1 [41. At all times, the cumu­
lative heat evolved is higher at higher temperatures. Also, these
data indicate that the total heat evolved during isothermal cure in­
creases with increasing temperature.
The development of a kinetic model for a cross-linking reaction is
inherently concerned with the accurate correlation of experimental
data to an expression describing the dependence of the rate of cure
on time and temperature. It is generally assumed that the exothermic
heat liberated during isothermal cure is proportional to the number of
bonds that have reacted in the system [5—7]. Therefore, the cumu­
lative heat evolved at any time is directly related to the extent of re ­
action at that particular time. Based on these considerations, a rela­
tive degree of cure is normally defined as follows:

( 1)
Qx
334 Kamal and Ryan

F ig . 1 Rate of heat evolution as a function of time for an epoxy resin


cured stoichiometrically with m-phenylene diamine. (From Ref. 4.)

where Q denotes the cumulative heat evolved isothermally up to some


given time t, and Qt denotes the total heat evolved during the course
of the entire isothermal reaction. It has already been mentioned that
Qx increases with increasing temperature.
The simplest kinetic model for describing the isothermal cure of
thermosets is given by the familiar nth-order rate expression

( 2)
(ti k ( l - a)^

where k is the temperature-dependent rate constant and n is the ki­


netic exponent. However, the kinetic expression given by Eq. (2)
is not capable of realistically describing the progress of the curing
reaction, where a maximum in the isothermal rate is usually observed
as shown in Fig. 1. For a variety of thermosetting systems and mold­
ing compounds, including epoxies, unsaturated polyesters, and
Thermoset Injection Molding 335

Fig. 2 Cumulative heat evolution as a function of time for an epoxy


resin cured stoichiometrically with m-phenylene diamine. (From Ref. 4.)

phenolic systems, the following kinetic expression has been found to


correlate isothermal kinetic data to a reasonable degree of accuracy
[4,8-14,75]:

(ir) * k z a ’^ j d - (3)

Usually the kinetic exponents m and n are assumed to be independent


of temperature, and the rate constants ki and k 2 are assumed to have
an Arrhenius temperature dependence. Thus

ki = A l exp (4)
336 Kamal and Ryan

k 2 = A 2 expl -
(4 i) (5)

where Ai and A 2 are pre-exponential constants, and are ac­


tivation energies associated with the rate constants and k 2 , respec­
tively, R is the gas constant, and T is the absolute temperature. A
comparison between the calculated reaction rate and cumulative heat
evolution with experimental data is shown in Fig. 3 [12]. The values
of the kinetic parameters are determined by standard fitting or re ­
gression techniques. It is often convenient to ”fix^’ the kinetic pa­
rameters in a sequential manner. This is done by correlating all of
the isothermal data obtained at different temperatures and determin­
ing the best values for A 2 and AEk2 * The value of the rate constant
k 2 is now fixed at any given temperature, and the best values for A^
and AEkj are next determined. Finally, the optimal values of the ki­
netic exponents are ascertained. Figure 4 shows the variation of the
logarithm of k 2 with reciprocal absolute temperature for an epoxy mold­
ing compound [12]. The kinetic activation energies obtained in this
manner are generally consistent with the values reported in the lit­
erature for similar systems. For some systems the reaction order may
change with temperature, and consequently the kinetic exponents will
give evidence of a variation with temperature. Also, for many systems
ki is effectively zero.
The time-consuming determination of the kinetic parameters by least-
squares regression methods can be facilitated or even circumvented if
the overall reaction order m + n is known or can be estimated. This
shortcut estimation procedure uses information from a single charac­
teristic point on the reaction exotherm, namely, the point of maximum
rate of cure [15].
Since the success and convergence of many nonlinear regression
schemes depends on the accuracy of the initial guesses of the re ­
gression parameters, the following estimation procedure can be profit­
ably used to provide reliable initial estimates. The method entails re ­
arrangement of Eq. (3) to the form

, a(l-a) ,
ID = i n — ------------- 'In a ( 6)

Initially, when t = 0, a = 0 and Eq. (3) yields

[(91.1 t= 0
= ki (7)
Thermoset Injection Molding 337

TIME (seconds)

200 400 600 800

16

a
H
W
S

12
g
S
H <
P O!
O t:^
> <!
w W
O
H
C
W

TIM E (seconds)

F ig . 3 Comparison of calculated and experimental reaction rate and


cumulative heat evolution for an epoxy molding compound. (From
Ref. 12.)

Thus, the kinetic constant ki is readily determined from isothermal


reaction rate data and, as was mentioned, is negligible for many sys­
tems. The maximum or peak of the reaction rate curve is defined by

= 0 ( 8)
3t^
33B Kamal and Ryan

1 - 1 2
- (K ) X 10
T

Fig. 4 Logarithm of kinetic rate constant, k 2, as a function of re ­


ciprocal absolute temperature for an epoxy molding compound. (From
Ref. 12.)
Thermoset Injection Molding 339

Applying this condition to the kinetic rate equation gives

1 -m
nk^ «p + k^(m + n )a p mk^ = 0 (9)

where ap denotes the extent of reaction at the peak. If it is assumed


that the overall reaction order is known, then

A = m+ n ( 10)

where A denotes the overall reaction order of the system. Using Eq.
(10) in Eq. (9) and rearranging gives

1 -m
(A - m)k^ ap
ko = ( 11 )
m - Aa

Using Eq. ( 6 ) expressed at the peak and substituting Eq. (11) gives

a p / (l - a p )^ - ki
m = In ---------------- —--------------- — /In a ( 12)
(A-m)k^ap ”^/(m - ^

The quantities k^ and 6 tp may be rapidly determined experimentally at


any given temperature. Equation (12) may be solved numerically to
determine m. The kinetic rate constant k 2 is found from Eq. (11).
This method has been successfully applied to the cross-linking of
epoxy [15] and unsaturated polyester systems [16—18].
The kinetic model given by Eq. (3) can alternatively be expressed
in terms of an absolute conversion or extent of reaction, defined
as

-0 . = ^ / a3(T) (13)
Qjj Q,p W n /

where Qn denotes some ultimate heat of cure or normalizing parame­


ter. Combining Eqs. (3) and (13) leads to the following kinetic ex­
pression in terms of the absolute degree of cure:

™ j. 1 „l-m -n
; + k 2 a„ )(6 - a . ) 3 (14)
(ir l =
340 Kamal and Ryan

Integrating Eq. (14) gives the variation of the absolute degree of cure
with time at a specified temperature.

N onisoth erm al S tu d ie s : Estim a tin g K in e tic P a ra m e ters

Isothermal experiments are usually used to precisely determine the ki­


netic parameters governing thermoset cure. The m^jor disadvantage
of the isothermal method is that several time-consuming series of mea­
surements are required at different temperatures. Another limitation
of isothermal measurements arises from the possibility of premature
reaction while the sample is elevated to the desired cure temperature.
Nonisothermal characterization techniques are generally more rapid
compared to isothermal experimental studies, which makes them more
attractive for providing comparable information in routine industrial
or quality control applications. Various techniques for treating d y­
namic data have been reviewed in the literature [19—23].
Several methods have been suggested for the analysis of the ki­
netics of a thermosetting material subjected to a nonisothermal tem­
perature history. Most of the published work refers to simple nth-
order kinetics or some equivalent form of E q. ( 2). Both derivative
methods [24—33] and integral methods [19,20,34—38] have been suc­
cessfully employed for determining the kinetic activation energy and
reaction order. For more complicated kinetic expressions, the extent
of cure could be calculated from an isothermal kinetic model by sub­
dividing the time interval into a sequence of small isothermal steps.
Incremental changes in the progress of the reaction are then readily
computed. However, the inverse problem of ascertaining the isother­
mal kinetic parameters from a nonisothermal experiment for a general
rate equation is not so readily apparent.
Since the fractional degree of conversion is a function of both time
and temperature, the total differential of the extent of reaction is ex­
pressed as follows:

da : dt (15)
(If). « ^ ( i ) .
Equation (15) can also be expressed as

da _ /8 a\ /8 a\
(16)
dt = \ 8 t;^ dt

where (aa/Bt)»!« denotes the isothermal rate equation and dT/dt is the
heating rate. The validity of Eq. (16) has been the subject of much
discussion and controversy [5,23,39—48]. The isothermal rate equa­
tion may be expressed in the general form
Thermoset Injection Molding 341

f(a , T ) (17)
(ifl
Integrating Eq. (17) gives

da
= g(a , T ) = t + (KT) (18)
/f
'f (a , T)

Since initially a = 0 at t = 0, then g(0, T ) = (t>(T) and

IÌK \ = à ± (19)
\3T/^ dT

The partial derivative on the left side of Eq. (19) may be expanded to
give

(M \ - + IlK ] ( 20 )

Combining Eqs. (16)--(20) gives the dynamic rate expression

( 21 )

All terms on the right side of Eq. (21) may be related to the iso­
thermal kinetic expression. The major difficulty arises in determin­
ing the function g (a , T ) from the integral in Eq. (18). Analytical
expressions for the dynamic rate equation have been derived for
various isothermal rate equations (48), including that given by Eq.
(3 ).

H eats o f R e a ctio n a n d S p e c ific H eats

An important application of the scanning mode of the differential scan­


ning calorimeter relates to determining the specific heat of the resin
as a function of temperature and degree of cure. The measured dif­
ferential heating rate between an uncured or partially cured sample
and an empty reference pan represents a combination of the heat of
reaction and the sensible heat required to raise the sample tempera­
ture. Assuming that heat losses are negligible and that no other
phenomena can give rise to changes in the enthalpy of the system,
one may use a simple heat balance to relate the specific heat to the
initial degree of cure, the reaction rate expression, and the mea­
sured calorimeter signal [ 4]. For a liquid epoxy resin the specific
342 Kamal and Ryan

H
< U
WO
W
GO
CJ\
M rH
W ctJ
MO
U
W
W &
WU

Fig. 5 Variation of specific heat with temperature and level of cure


for an epoxy resin cured stoichiometrically with m-phenylene diamine.
(From Ref. 4.)

heat remained essentially constant at low degrees of cure (a < 0 . 2 ).


Similarly, at high conversions (a > 0.8) the specific heats showed
little variation. The values were consistent with independent mea­
surements reported on the uncured and fully cured systems. How­
ever, at intermediate conversions, the specific heat varied signifi­
cantly and continuously from the uncured value to the fully cured
value. Figure 5 shows the specific heat of this system as a function
of temperature and degree of cure [4 ]. To a first approximation, the
specific heat can be considered to vary linearly with temperature
over the temperature range considered. Thus

Cp = a + bT ( 22 )

Remember that a and b depend on the extent of conversion. Such


information can be particularly useful for the modeling of thermoset
Thermoset Injection Molding 343

processing operations. Similar results have been obtained by Pusat-


cioglu et al. [76] for a general-purpose unsaturated polyester sys­
tem.

Residual Cure
From Fig. 2 it is evident that the total or cumulative heat evolved in­
creases with increasing temperature. Therefore, the ultimate state
of cure attained in the material is dependent on the thermal history.
If a sample is initially cured isothermally at some temperature T , not
all of the reactive sites or species may participate in the cross-link­
ing reaction, and the sample will then possess some residual reactiv­
ity or residual cure. This residual reactivity can be measured ex­
perimentally by subjecting the sample to a series of isothermal cure
experiments at successively higher temperatures. Alternatively, the
partially cured sample could be scanned in the calorimeter until ho
further reactivity is evident. For a liquid epoxy resin, the residual
reactivity has been found to be inversely proportional to the cure
temperature, as shown in Fig. 6 . Sidi [49] found that, for a glass-
and mineral-filled, injection-mol din g grade unsaturated polyester
resin, the residual extent of cure decreased exponentially with in­
creasing cure temperature. If the initial curing temperature is suf-
ficiehtly low, prolonged post-cure does not necessarily force the

0
H
<3
W
ÎX!

1
Q
O'
<
M
zn
W

F ig . 6 Variation of residual extent of cure as a function of isothermal


cure temperature for an epoxy resin cured stoichiometrically with m-
phenylene diamine. (From Ref. 4.)
344 Kamal and Ryan

cross-linking reaction to ultimate completion. Thus, to some extent


the initial thermal history of the system determines the network struc­
ture and cross-link density and controls the level of accessibility of
reactive segments during later stages of the reaction.
Calorimeter determination of the residual cure provides a useful
method for measuring the cure distribution in molded thermoset parts
[50-52].

B. Rheological C h a ra c te riza tio n

Determining the rheological behavior of thermosetting materials is


made difficult because the resin undergoes reaction during measure­
ment. Many tests have been devised to characterize and investigate
the rheological behavior of thermoset-molding compounds, including
ASTM cup flow, orifice flow, and spiral flow. Other commonly used
empirical techniques include the Brabender torque rheometer and the
Monsanto disk rheometer. The major limitation of the above methods
is that they cannot be employed to measure fundamental rheological
properties which can subsequently be used in the mathematical model­
ing of thermoset processing operations. The main utility of these
tests is that they provide a consistency standard that the resin sup­
pliers can meet or guarantee in a reasonable and economical manner.
For the determination of fundamental rheological properties, carefully
defined and controlled flows are necessary.

E x p e rim e n ta l T ech n iq u es

As mentioned, a variety of empirical standardized tests have been de­


veloped for evaluating the flow characteristics of thermosetting ma­
terials. Two of the more commonly used ASTM procedures are the
instrumented spiral flow test [53,54] and the Brabender torque rhe­
ometer [55—57].
The spiral flow of thermoset-molding compounds involves an in­
strumented spiral flow mold in conjunction with a transfer press. A
standardized procedure is followed, and the spiral flow length of the
material is measured. Although this measurement may be of some
value in comparing different grades of material, it is of limited util­
ity for the analysis or understanding of processing problems. In
the Brabender torque rheometer, the temperature of the mixing head
is controlled. The material is compacted, melted, and cross-linked,
and the torque is measured as a function of time. The maximum load­
ing torque, the maximum torque, the final torque peak, the time re­
quired to reach the minimum torque, and the time required to reach
the final torque peak are all reported.
To systematically investigate the influence of time, temperature,
and shear rate and to obtain a more fundamental understanding of
Thermoset Injection Molding 345

the rheological behavior of reactive systems involves a variety of


methods. The most common technique involves shearing a sample of
the thermosetting material using cone and plate fixtures [8,12,58—67].
The two most widely used commercial instruments are the Weissenberg
rheogoniometer and the Rheometrics mechanical spectrometer. These
instruments are capable of controlling temperature by an environmen­
tal chamber surrounding the test fixtures. In addition to determin­
ing the viscosity as a function of time at a particular temperature and
shear rate, these instruments can also yield information with regard
to the first normal-stress difference by measuring the normal force
or thrust on the platens. The cone and plate geometry restricts the
measurements to relatively low shear rates due to problems associated
with loss of sample and flow instabilities. Also, the time required for
the sample to attain thermal equilibrium limits meaningful results to
systems that do not polymerize rapidly or are maintained at low tem­
perature. Typical viscosity-time isotherms are shown in Fig. 7. Once
the sample has been introduced between the test fixtures, a period of
approximately one minute is required for the sample to attain the de­
sired temperature. During this initial period the measured viscosity
will decrease slightly with time. As the cross-linking reaction pro­
gresses, the viscosity continuously increases until the melt trans­
forms to an infusible solid. Occasionally, material may be observed
to flow out of the test fixtures when the sample reaches a high cure
level and, as a result, the calculated ’’viscosity level” would decrease
due to the loss of material. This observation dictates an upper limit
for meaningful viscosity measurements.
Other methods for measuring the rheology of thermosetting systems
have been used, including falling sphere [ 6 8 ] , parallel plate [69,70] ,
squeezing flow [71], and eccentric rotating disks [72] . These meth­
ods are generally subject to the same limitations as the cone-plate mea­
surements. For liquid resins the samples are generally charged di­
rectly to the rheometer, whereas for molding compounds the samples
are generally premolded to the appropriate geometrical configuration.
To characterize the material at high shear rates, one can use a
capillary rheometer [12,14,49]. A split-barrel design facilitates clean­
ing of the fully cured material. Two limitations of the capillary rhe­
ometer are the short measurement interval dictated by the reservoir
capacity and the generally extended time period for the material to
achieve thermal equilibrium.
An on-line parallel-plate viscometer has been designed for fast
polymerizing systems [73].

Effects of the Degree of Cure on Rheology


The rheological behavior of reactive systems is dependent on sev­
eral factors. A primary consideration is that the rheological or flow
u>
J?
ai

o
Pu

QJ
3
5L
03
3
a
73
03
Fig. 7 Variation of viscosity as a function of time for an epoxy resin cured stoichiometrically with m-phenyl- 3
ene diamine. (From Ref. 74.)
Thermoset Injection Molding 347

properties are intimately influenced by the extent of reaction or the


degree of cure of the cross-linking or polymerizing system. In ad­
dition, the rheological response is also influenced by a variety of
other factors such as pressure, temperature, particulate loading
level, and the rate at which the system is being deformed. For sim­
plicity it will be assumed in this section that the deviatoric stress
tensor depends only on the instantaneous level of the rate of defor­
mation. Thus, a general constitutive relation for the deviatoric stress
tensor t may be expressed as

T = f (A , P, T, (1), a ) (23)

where t denotes the deviatoric stress tensor, A is the rate of defor­


mation tensor, P is the pressure, T is the temperature, cj) is the load­
ing level, and a is the degree of cure. The rate of deformation ten­
sor is defined in terms of the velocity gradient as follows:

A = Vv + (V v) (24)

where (Vv)*^ is the transpose of the velocity gradient tensor. Other


measures of the deformation rate could be used in E q. (23). It is
generally assumed that no significant interactions exist between the
primary independent variables.
For the purpose of analyzing various polymer processing opera­
tions, numerous specific forms of Eq. (23) have been suggested. An
important special case corresponds to steady simple shear flow. In
this situation three independent material functions may be defined as
follows:

^ 12
(25)

'^11 (26)

^22 ~ ^33 (27)


y2

n, and ^ 2 are the shear viscosity, first normal-stress coefficient,


and second normal-stress coefficient, respectively, Ti 2 is the shear
stress. Til, '^22 and T3 3 are normal stress components, and Y is the
shear rate.
The variation of viscosity with reaction time t has been described
with some success by the relation [8,59—61,77]
3U8 Kamal and Ryan

AE
__ n
In = n = Cl + ■
RT
+ C 2 t exp
\ (-^ )
RT/
(28)

where denotes the flow activation energy of the uncured system,


AEk is the kinetic activation energy, R is the gas constant, and Ci
and C 2 are constants. Equation (28) has been used to correlate rea­
sonably well typical viscosity time isotherms, such as the data in Fig.
7. Time may be eliminated as the independent variable in favor of the
extent of reaction or degree of cure by using an appropriate expres­
sion for the cure kinetics, such as that given by Eq. (3 ). Figure 8
shows the viscosity data given in F ig. 7 plotted directly as a function
of the degree of cure. The apparent viscosity of the curing system
increases continuously during the cross-linking reaction. As the for­
mation of a network structure is approached, the viscosity increases
in a very pronounced manner. Other expressions relating viscosity
to the extent of reaction have been developed from a detailed consid­
eration of the chemical or structural changes that accompany the cross-
linking process [14,78—81].
Relatively few systematic studies have been conducted for the
variation of other rheological properties with the extent of reaction.
Measurements of the first normal-stress difference with cure time have
been reported by Ryan and Kamal [74,123] and Han and Lem [16—18] ,
among others.
Negative normal stresses may be exhibited by an undeformed sam­
ple or by a sample subjected to low shear rates as a result of the
shrinkage that accompanies the curing process. At higher shear
rates, viscoelastic effects predominate, and positive normal stresses
are developed. These positive normal stresses increase markedly with
cure time, as the material transforms to a network structure.
Dynamic measurements in oscillatory shear have the advantage of
being conducted very close to, or even beyond, the point of gelation.
The dynamic viscosity n ’ and storage modulus exhibit similar pro­
nounced increases with cure time. However, the dynamic measure­
ments generally appear to be more dependent or sensitive to struc­
tural variations or resin chemistry.

Temperature Effects
The variation of viscosity with time and temperature for a nonisother-
mal curing history may be described in terms of a modified form of
Eq. (28):

AE I
___n
In n = c i + RT / C 2 exp dt (29)
Thermoset Injection Molding 349

Fig. 8 Variation of viscosity as a function of degree of cure for an


epoxy resin cured stoichiometrically with m-phenylene diamine. (From
Ref. 74.)

More recently, the Williams-Landel-Ferry (WLF) equation has been


found to accurately describe the variation of viscosity under noniso-
thermal conditions [82,83]. In this case the WLF equation is
350 Kamal and Ryan

-c r + (T - T ^)
In (30)
C’^ + (T - T )

where nx is the viscosity of the reactive system at temperature T ,


is the viscosity at the glass-transition temperature Tg, and
Cy and C ’2 are the WLF parameters. An expression for the variation
of the glass-transition temperature with degree of cure must also be
introduced and will be discussed in more detail later. In addition,
the parameter Ci is permitted to be dependent on Tg.
Plots of viscosity as a function of reciprocal absolute temperature
at constant shear rate and degree of cure may be readily constructed
from experimental measurements of viscosity as a function of time or
degree of cure under various isothermal conditions. Typical results
are shown in Fig. 9. Over a moderate temperature range, the data
generally follow a linear relationship from which a flow activation en­
ergy may be determined. Although relatively little experimental data
of this type are available, the published results indicate that the flow
activation energy is independent of the shear rate but may increase
with advancing levels of cure. Thus, it becomes apparent that Eq.
(29) may be somewhat limited in terms of its predictive ability, since
the flow activation energy is considered to be independent of the de­
gree of cure.

Non-Newtonian Behavior
In a similar fashion ^ the non-Newtonian behavior of reactive systems
may be systematically investigated from viscosity—degree of cure or
viscosity-time isotherms by cross-plotting the viscosity as a function
of shear rate at constant temperature and degree of cure. Some il­
lustrative results are shown in Fig. 10. The viscosity of some liquid
thermosetting systems at low levels of cure are approximately New­
tonian. At high conversion levels or for high-viscosity-molding com­
pounds or filled systems, the viscosity behavior may be approximated
by a simple power-law expression over a moderate range of shear
rates. The consistency index will depend on the temperature and
degree of cure, as discussed in the preceding sections. The power-
law index is essentially independent of temperature but may depend
slightly on the extent of reaction for some thermosetting systems.

C. Therm al P ro perties

The heating and solidification phases generally constitute the most


significant part of a molding cycle or processing operation. Thus,
thermal properties such as thermal conductivity and thermal diffu-
sivity may have a direct bearing upon or even dictate the process
Thermoset Injection Molding 351

2.50 2.55 2.60 2.65 2.70 2.75

1 3
i (K ) X 10
T

F ig . 9 Variation of viscosity as a function of reciprocal absolute


temperature at constant shear rate and degree of cure. (From
Ref. 74.)

economics. In addition, the glass-transition temperature is an impor­


tant structural parameter for both reactive and nonreactive systems
and has proven to be useful in various correlations, such as the WLF
equation cited in a previous section.
fjl
ro

u
I
iz

7s
D)
3
9L
01
D
Ql

0 }
3
Fig. 10 Variation of viscosity as a function of shear rate at constant temperature and degree
of cure. (From Ref. 74.)
Thermoset Injection Molding 353

Thermal Conductivity and Diffusivity


The thermal conductivity k is defined in terms of the steady-state
heat conduction in an isotropic medium as given by Fourier’s law;

kVT (31)

q is the heat flux vector, and VT is the temperature gradient. The


Thermal conductivity depends on factors such as the temperature and
degree of cure of the material. Although fairly extensive data exist
relating to the thermal conductivity of nonreactive systems, very little
information is available about the thermal conductivity of thermoset­
ting materials.
Ueberreiter et al. [84,85] measured the thermal conductivity of
polystyrene cross-linked with various amounts of divinyl benzene.
They found that the thermal conductivity increased with increasing
cross-linked density. The effect of radiation-induced cross-linking
on the thermal conductivity of polyethylene has been investigated
by Tomlinson et al. [ 8 6 ]. Above the melting point they observed
that the thermal conductivity increased with increasing radiation
dose due to cross-linking. Similar observations were also made by
Yamamoto and co-workers [87—89] , Berman et al. [90], and Nielsen
[91],
The thermal conductivity of several thermosetting systems has
been measured [4,76,92—98,121]. In general, it increases linearly
with the extent of cure during the initial stages of the cross-linking
reaction. A more pronounced increase in the thermal conductivity
may occur as a result of a glass transition due to cure. The varia­
tion of thermal conductivity with degree of cure for an epoxy resin
cured stoichiometrically with m-phenylene diamine is shown in Fig.
11 [4,12].
The thermal diffusivity at is particularly important with regard
to the transient temperature distribution within a material and is re ­
lated to the thermal conductivity by

(32)
t = pC
TT

p is the density, and Cp is the specific heat at constant pressure.


Although the thermal diffusivity may be calculated directly from a
knowledge of these properties, it is frequently determined by ex­
perimental techniques from the dynamic temperature response of the
material. These experimental methods include periodic heat flow,
pulse and flash methods, peak-time measurements, and quasi-sta-
tionary methods. A review of these experimental techniques has
been given by Sourour [ 4].
354 Kamal and Ryan

M
>
M
H
O
P
Q
O
u

w
¡II
H

AMOUNT OF CURE

Q (cal/g)

F ig . 11 Variation of thermal conductivity as a function of degree of


cure for an epoxy resin cured stoichiometrically with m-phenylene
diamine. (From Ref. 4.)

Figure 12 shows the variation of thermal diffusivity with cure time


for a liquid epoxy resin system [4,121]. For this system, the ther­
mal diffusivity increases sigmoidally with cure time, attains a maxi­
mum level, and then decreases slightly to a final value.

Glass-Transition Temperature
The glass-transition temperature is an important structural parame­
ter that depends on the internal rigidity of the polymeric chains,
the molecular weight, the thermal history of the material, and the
H
3*
(D
-s
3
O
{fi
a>

a>
n
O
3

a
5’
CQ

CURE TIME, t (minutes)

U)
Fig. 12 Variation of thermal diffusivity as a function of degree of cure for an epoxy resin cured stoichiomet- U1
Ü1
rically with m-phenylene diamine. (From Ref. 4.)
356 Kama! and Ryan

w
!=)
H

H
EH
O

cn
IH
O

EXTENT OF CURE,

Fig. 13 Variation of the glass-transition temperature as a function of


degree of cure for an epoxy resin cured stoichiometrically with m-
phenylene diamine. (From Ref. 4.)

cross-link density of the thermosetting material. During isothermal


cure the glass-transition temperature Tg increases and may approach
or even surpass the isothermal cure temperature [99—103]. Figure
13 shows the variation of Tg with degree of cure for an epoxy resin
[4,122]. Initially, the increase in Tg is due primarily to the increase
Thermoset Injection Molding 357

in molecular weight that accompanies the cure reaction. After gela­


tion, Tg increases markedly due to the increase in the cross-link den­
sity of the network and ultimately levels off to the final value for the
cured material.

D. E ffect o f F ille rs on K inetics and Rheological


and Therm al P ro perties

Particulate fillers are frequently incorporated into thermosetting res­


ins to improve the dimensional stability, ultimate mechanical proper­
ties, and performance of the finished product. However, the addi­
tion of particulates or fibers to the resin may have a pronounced in­
fluence on the curing kinetics, rheology, and thermal properties of
the system.
The effect of fillers on the reaction kinetics has been studied by
Kubota [104], Dutta and Ryan [13], and Lem and Han [17,18] , among
others. The curing process will be influenced by the presence of spe­
cific chemical constituents or complexes on the particulate surface.
Thus, the dependence of cure kinetics on filler concentration wül be
a function of the particular matrix-filler combination being consid­
ered. Dutta and Ryan [13] found that the kinetic rate constants
given by Eqs. (4) and (5) initially decreased, went through a mini­
mum, and then gradually increased with increasing filler concentra­
tion for a liquid epoxy system containing either a carbon black or a
silica filler. The effects were more pronounced at higher tempera­
ture and were more strongly influenced by the carbon black filler.
The rate constants were modified by letting the preexponential fac­
tors Ai and A 2 be described by simple third-order polynomial func­
tions in filler concentration. Lem and Han [17] found that the rate
of cure increases with increasing particulate content for an unsat­
urated polyester resin. Although it is difficult to make quantitative
generalizations or to elucidate the detailed mechanisms responsible
for the observed behavior, the presence of certain chemical complexes
or reactive groups on the particulate surface may have a marked cata­
lytic effect on the cross-linking reaction or may even participate in
the curing reaction.
Numerous theoretical and empirical expressions have been proposed
for describing the variation of viscosity with filler concentration or
loading level. Many of these relationships have been recently re ­
viewed in the context of polymer composites [104,105]. Some of the
more successful or commonly used expressions are summarized in
Table 1. Unfortunately , most expressions do not generally give ac­
curate quantitative results at high loading levels which are typical of
commercial filled molding compounds. Very few systematic studies
358 Kama! and Ryan

T a b le 1 Some Commonly Used Expressions for Describing


the Dependence of Relative Viscosity on Loading Level

Thomas [106]
= 1 + 2.5(j) + 10.05(j)^ + A exp (B(J))
R

Mooney [107]

( 2

Eilers [108]
1.25(1)
1 +
‘R

Frankel and Acrivos [109]

9
'’r 8 1 -

Chong et al. [110]

0 .7 5 (^/< l.

^R = F" 1 -

Kitano et al. [ I l l ]

Ml -

Tanaka and White [112,113]

nR - 1 ]^ -^ "

riR = relative viscosity, cj) = loading level, ci)]vi = maximum


packing fraction, A = empirical constant (0.00273), B =
empirical constant (16.6), C = empirical constant (0.68
or (l)ivi)i = empirical constant, n = empirical constant.
Thermoset Injection Molding 359

have been made relating to the effect of filler concentration on the


rheological behavior of reactive thermosetting systems [65—67]. How­
ever, some general observations regarding the dependence of viscos­
ity on particulate loading may be readily ascertained from the behav­
ior of nonreactive systems. The viscosity increases with loading level
particularly at high volume fractions. These systems may even ex­
hibit a yield stress. Particle-particle interactions and the breakup
of particle clusters or agglomerates may cause the system to exhibit
non-Newtonian shear thinning behavior even where the matrix is New­
tonian. At elevated shear rates the viscosity of the filled system or
composite tends to approach that of the matrix.
Fillers are generally inert and act as both a diluent and a heat
sink. Consequently, the exothermic heat generation and tempera­
ture rise in a thermosetting system are reduced.
Various additivity rules, mixing laws, or inclusion models may be
used to estimate the thermal properties of the composite material.
These methods are discussed in detail in the literature and will not
be reviewed here.

III. THER M O SET IN J E C T IO N -M O L D IN G S IM U L A T IO N

Simulation is an essential engineering tool for the development, de­


sign, analysis, operation, and control of innovative processing strat­
egies. The technology of computer hardware and systems software
for simulation purposes has made amazing strides during the past
decade. The current state-of-the-art consists of user-friendly CAD/
CAM packages which permit the assessment of trends with regard to
changes in material properties, process conditions, or mold geometry.
These developments have been almost exclusively focused on the mold­
filling behavior of commodity thermoplastic materials. Application to
thermosetting systems has not been extensively evaluated due to in­
adequacies in material characterization and process models incorpor­
ating chemical reaction.

A. In jectio n -M o lding System s: Equipm ent and Resins

Thermosetting materials were successfully injection molded on plunger-


driven machines during the 1940s. The in-line reciprocating-screw
injection-molding machine was subsequently developed during the
1950s. However, the reactive nature and complex rheology of ther­
mosetting systems posed numerous difficulties initially and hampered
the development and application of the screw injection process for
360 Kamal and Ryan

these materials. Despite these difficulties, injection molding permits


greater flexibility for part geometry or design, better control of the
process, no need for special preheating equipment or procedures,
short cycle times, and greater adaptability for automation. Process
control units, robots, and automatic mold change equipment have had
a dramatic impact on production efficiency, scrap reduction, data
acquisition, part consistency, conformance to tight tolerances, and
automated part removal, sorting, and finishing.
In a typical injection-molding process, the molding compound is
usually gravity fed from a hopper to the flights of the reciprocating
screw. The material is conveyed forward by the rotation of the screw.
A schematic diagram of a typical in-line reciprocating-screw machine
is shown in Fig. 14. The mechanical action of the screw as well as ex­
ternally supplied heat causes the polymer to transform to a viscous
molten state. The screw is capable of axial movement in a reciproca­
ting manner in order to force a predetermined amount of material into
the mold cavity. Once the mold cavity has been completely filled, ad­
ditional material is packed into the cavity under pressure to compen­
sate for shrinkage that occurs when the part solidifies. An elevated
pressure is maintained to minimize any dimensional changes that might
otherwise occur. Solidification is accomplished by a heated mold which
causes the material to react, cross-link, and transform to an infusible
solid. Once solidified, the molded article is injected, and the molding
sequence is repeated cyclically. The clamping unit provides the re ­
quired force to keep the mold closed during the filling and packing of
the cavity, and opens to permit ejection of the finished part.
More productive molding techniques have been developed, includ­
ing hot-cone molding, runnerless injection/compression (R IC ), and
lost-core hoUow-part injection [114]. These innovations have re ­
sulted in dramatically shorter molding cycle«, reduced scrap, and
improved finished part properties. They also permit molding of com­
plicated part configurations.
Hot-cone or hollow-sprue molding involves placing a simple man­
drel or core into the sprue tooling to obtain a hollow sprue and there­
by reduce the sprue weight. Cure times maybe substantially reduced
since the sprue may be one of the thickest sections. The technique
was developed by General Electric and has been licensed to Rogers
Corporation and Durez Resins & Molding Materials.
Runnerless injection/compression molding is a patented process
developed by Durez and licensed to more than 40 molders. A cold
manifold plate and cooled bushing are employed with modified injec­
tion/compression tooling to minimize the production of scrap in the
form of cured sprues and runners. Plasticized material is injected
into a partially closed mold without creating back pressure on the
cold manifold sprue. The mold is closed to compression mold the
part having no sprue or runner system.
3-
fl>
"5
3
o
w
a>

a>
n
o
3
O.
a
5*
(Q

Fig. U Schematic diagram of a reciprocating-screw injection-molding machine.

cr>
362 Kamal and Ryan

Lost-core molding involves the molding of a part around a eutec­


tic alloy insert, typically tin/bismuth. The insert is automatically
melted out and recovered during the part post-bake operation.
A variety of thermo set-molding compounds have been developed,
including allyls, aminos, phenolics, epoxies, silicones, and thermo­
set polyesters.
Diallyl ortho-phthalate and diallyl isophthalate molding compounds
exhibit superior electrical insulating properties under severe environ­
mental conditions involving elevated temperatures and humidity.
Molded parts also exhibit excellent dimensional stability, chemical re ­
sistance, mechanical strength, and heat resistance. These materials
are frequently used in demanding electrical and electronic applica­
tions, such as connectors, housings, switches, relays, circuit break­
ers, and terminal strips, for communication, computer, and aerospace
systems. Molding grades are available in mineral, organic fiber, and
glass-fiber-füled formulations.
Amino resins are produced through a controlled reaction of formal­
dehyde with various compounds containing an amino (NH 2 ) group.
Urea and melamine formaldehyde are the two most widely known and
commonly used examples. The ureas are commonly used in wiring de­
vices, closures, control housings, knobs, etc. They exhibit excel­
lent surface finish in a wide variety of color. The alp ha-cellulose
melamine grades are typically used for dinnerware, knobs, buttons,
ashtrays, etc. The molding compounds may contain minerals, glass
fibers, chopped cotton flock, or wood flour as filler materials.
Phenolic resins are produced from the reaction of phenol with form­
aldehyde, Single-stage resins (resoles) consist of phenol, formalde­
hyde, and an alkaline catalyst. Two-stage resins (novolocs) consist
of an excess of phenol, formaldehyde, an acid catalyst, and a cross-
linking agent such as hexamethylenetetramine. Good physical strength,
heat resistance, dimensional stability, and moderate cost make them
attractive for transportation, appliance, and electrical component ap­
plications. Phenolic molding compounds may be reinforced with min­
eral, wood flour, cotton flock, chopped fabric, glass fibers, Teflon,
and nylon. General-purpose wood-flour filled compounds are exten­
sively used for knobs, handles, and appliance housing. Mineral- and
glass-filled molding compounds are used in high-performance products
such as wiring devices, connectors, switches, automotive components,
and commutators. New injection-moldable phenolic alloys have been
developed with much greater heat-deflection resistance and thermal
stabüity than conventional phenolics have. These materials are suc­
cessfully penetrating high-performance automotive, aerospace, elec­
trical, and other metals-replacement markets [114].
Epoxies are frequently used for encapsulation of capacitors, semi­
conductor devices, and inductors for the electronics industry. They
are also used in molding applications where high thermal stability may
Thermoset Injection Molding 363

be required. Molding compounds may be filled with glass fibers, sil­


ica, or other minerals.
Rigid silicone molding compounds are employed for encapsulation
of semiconductor devices, capacitors, resistors, etc. Semirigid elas­
tomer molding compounds are used for connectors, gaskets, seals,
0 -ring, plugs, etc.

Unsaturated polyesters include bulk molding compounds (BM C),


sheet molding compounds (SMC), and thick molding compounds (TM C).
These compounds are used for large circuit breakers, machine hous­
ing, automotive panels, appliance housings, etc. Reinforcement sys­
tems include glass fibers, organic fibers, and minerals.

B. Mathematical Modeling and C om puter Simulation

For thermoset injection molding, the pathway from the original design
or conception to the production of the finished part can be time con­
suming, expensive, and economically risky. Computer simulation can
be advantageously used as an engineering tool to expedite, analyze,
or optimize most facets of this process. Computer-aided design can
be used to ascertain the appropriate location for gates, parting lines,
weld lines, etc., to permit balancing of multicavity or family molds,
to enable the diagnosis of molding problems such as short shots or
overpacking, and to obviate the need for many costly tooling changes
or modifications. The basis of a computer simulation or program is
typically a mathematical model or description of the process. The es­
sence of mathematical modeling seeks to achieve a reasonable com­
promise between a rigorous mathematical description of the physics
of the process, on the one hand, and computational feasibility and
ease, on the other.

General Equations
The mathematical description of the fluid dynamics and heat transfer
characteristics associated with the various stages of the inject ion-mold­
ing cycle is based on the fundamental conservation principles of mass,
linear momentum, and energy applied to a continuum. In general, the
tensorial form of these relationships may be expressed as follows:
Continuity

I f + V.(pv) = 0 (33)

Motion

Dv
p —= = - V P + V •T + pg (34)
Dt -----
364 Kamal and Ryan

Energy

DT
pC = V -(k V T ) - (V -v ) + t ;Vv + pQ (35)
V Dt

where p is the density, P is the pressure, Cy is the heat capacity


at constant volume, T is the temperature, k is the thermal conduc­
tivity, V is the specific volume, Q is the rate of heat generation
due to chemical reaction, v is the velocity vector, g is the gravi­
tational force vector, y is the del or vector differential operator,
T is the deviatoric stress tensor, and D/Dt is the material or sub­
stantial derivative.
A balance on chemical species or reactive groups results in the
following expression:

3C.
—r + v-N. = R. (36)
at - -1 1

where ci denotes the concentration of species, i, Ni is the total flux


of species i including both diffusion and convection, and Ri is the rate
of generation of species i.
These relations are not sufficient in themselves to completely de­
termine the variables of interest, and additional constitutive equations
are required. They include a rheological equation of state, such as
that given by Eq. (23), to relate the deviatoric stress tensor to the
rate of deformation tensor, and a kinetic expression, such as that
given by Eq. (2) or (3 ), to describe the reactive nature of the sys­
tem .

A vailab le Models
For convenience, the injection-molding cycle may be divided into a
series of sequential steps: plastication, filling, packing, and solidi­
fication. The thermomechanical history of the material during plas­
tication determines the degree of cure at the time of injection into
the mold. Plasticating extrusion theory may be used to determine
the residence time of the resin within the barrel and, consequently,
the extent of reaction [115,116]. Relatively little reaction occurs
during this part of the molding cycle in order to minimize filling
problems. As a first approximation, a rigorous analysis of the ther­
mal and deformation history of the material in the barrel is typically
ignored.
The early mathematical analyses of the filling stage for thermo-
sets were developed by modifying mold-filling theory for nonreactive
Thermoset Injection Molding 365

systems. For most production cavities, the flow may be regarded as


two-dimensional. The relevant governing equations for a generalized
Newtonian fluid, where inertial and gravitational forces are negligible,
may be written as follows:

Continuity

. 3(pV ) 3(pV )
+ ^ + -------5L = 0 (37)
3t 3x 3y

Motion

^ = A (38)
dx 3z

i? = ^ !!z) (39)
3y 3 z \ 3z /

Energy

^ /3 T 33T . „
T_^.^ 33T\
T\ _ 8 /, 3 X\
" S I h ~ 3z

3V
(40)

Integrating Eqs. (38) and (39) twice with respect to z and making use
of no-slip conditions at the mold surface and symmetry at the midplane
of the cavity give

V =^ I (41)
X 3x

V = (42)
y 9y

where

1=1 - dz (43)
^wall ^
366 Kamal and Ryan

Substituting Eqs. (41) and (42) into Eq. (37) gives

3p rs P 3 PI t F^
^p I p + I p] +
3t X 3x 3y 3 yJ

I I + IP i i ] (44)
^ [sx 3 x ^ 3y 3yJ

During the filling, the flow may be considered incompressible, so Eq.


(44) becomes

iT— 1 !E] + i? ^ + ^ ii= o (45)


Lsx^2 + ^ 1
3y^J
3y^J 3x 3x 3 y 3y

If the variation of viscosity in the plane of the cavity (x and y direc­


tions) is considered negligible in comparison to the variation across
the thin gap, Eq. (45) reduces to the familiar expression

3^P . 3-^P _ ^
(46)
3x^ 3y^

This result was originally derived by White [117] and Kamal et al.
[118,119] for nonreactive systems and by Ryan and Kamal [12,124]
for thermosetting systems. Since the governing equation and asso­
ciated boundary conditions may be formulated without explicit refer­
ence to the rheological nature of the material, the pressure distribu­
tion and flow patterns are independent of the rheological characteris­
tics of the system. However, the filling time, velocity profiles, and
other details of the filling process are sensitive to the rheological
properties of material.
The general governing equation for the analysis of the packing
stage is Eq. (44). The major difficulty for the mathematical model­
ing of the packing stage rises from the requirement for an equation
of state relating density, temperature, and pressure. Very little ex­
perimental information is available pertaining to the compressibility of
reactive materials during cure, and few analyses of the packing be­
havior of thermosetting systems are available [12,14,124].
The cure and heat transfer requirements are often the dominant
characteristics of the overall cycle. Since a relatively long cure time
is required compared to the time required for filling, the effect of flow
on the temperature and degree of cure distribution can generally be
neglected as a first approximation. Since many injection-molded parts
are relatively thin compared to other characteristic mold dimensions,
the mathematical description of the solidification step reduces to a
Thermost Injection Molding 367

single unsteady, one-dimensional heat transfer analysis. Thus, Eq.


(40) becomes

pq + pQ (47)
at 3z 3z/
The thermal properties are usually assumed to be constant. Equation
( 4 7 ) has been analyzed for different kinetic models and various bound­
ary conditions, such as adiabatic, isothermal, and constant heat flux
conditions, at the mold wall.

C o m p a rin g M odel P r e d ic tio n s to E x p e r i m e n t a l D a t a

The analysis of production mold cavities may be accomplished by de­


composing or representing the complex geometry by a combination of
one-dimensional flow configurations. Many of the early CAD/CAM
simulation packages for thermoplastic polymers were based on this
approach. The current generation simulators employ a finite element
numerical scheme to analyze the actual cavity configuration directly.
A commercial mold-filling simulation program for thermosets has been
recently developed along similar lines [120]. In addition to the pre­
dictive capability for the filling time, pressure, stress, and tempera­
ture distributions, these simulation programs are effective in deter­
mining weldline location and balancing runners for multicavity or fam­
ily molds. Most thermoset mold-filling analyses neglect the fountain
flow at the moving front , which has a significant influence on the tem­
perature and degree of cure distribution across the part.
The filling time and front progression in simple mold cavity geome­
tries can be predicted with a reasonable degree of accuracy. Figure
15 gives a typical comparison between the predicted and experimental
melt-front position for a phenolic molding compound injected into a
thin mold cavity. At the onset of filling, the progression of the melt
front is essentially independent of the initial state of cure of the ma­
terial. Ultimately, the filling rate decreases due to the advancing
level of cure of the system.
The temperature within the cavity plays a prominent role in the
injection molding of thermosetting materials due to the dependence
of viscosity and reaction kinetics on temperature. Initially, the tem­
perature dependence of viscosity dominates the filling behavior. The
viscosity is relatively low initially due to the elevated mold tempera­
ture, viscous dissipation, and exothermic heat evolution. The filling
is consequently rapid. However, near the end of filling the curing
process dominates and gives rise to an increase in viscosity and a
resulting decrease in filling rate.
Figure 16 shows the axial pressure distribution in a thin rectangu­
lar cavity for a phenolic molding compound at various times. The
368 Kamal and Ryan

TIME (s)

Fig. 15 Melt-front position as a function of time for a phenolic mold­


ing compound. (From Ref. 14.)
3’
fD
3
o
w
(0

a>
n
o
D

a
5*
ID

AXIAL DISTANCE (cm) cy>

Fig. 16 Normalized pressure as a function of axial position along the cavity. (From Ref. 14.)
370 Kamal and Ryan

s'
w
Pi

w
Pi
p-<

TIME (s)

Fig. 17 Cavity pressure as a function of time. (From Ref. 14.)

pressure tends to become more uniform throughout the cavity during


packing. Similar behavior has been reported for the case of an epoxy
molding compound [12,124] . Figure 17 gives the pressure variation
as a function of time, and reasonably good agreement is observed be­
tween the predicted results and the experimental data.
A systematic experimental study of the cure distribution and ten­
sile properties in a filled unsaturated polyester molding compound
Thermoset Injection Molding 371

Surface Center

DISTANCE FROM SURFACE (MM)

Fig, 18 Experimental and computed cure distribution for a filled un-


saturated polyester molding compound. (From Ref. 52.)

has been undertaken by Sidi and Kamal [49—51]. A comparison be­


tween the experimental and predicted cure distribution is shown in
Fig. 18. The calculated results were obtained by a simple one-dimen­
sional heat transfer analysis [52]. Curing continues after the molded
part has been ejected from the mold, and good agreement is obtained
by accounting for this post-molding reactivity.
372 Kamal and Ryan

IV . C O N CLUSIO N S

To significantly advance the current state-of-the-art requires strength­


ening the scientific base in the areas of material property characteriza­
tion and process modeling procedures. This includes a better under­
standing and description of the actual process environment as well as
careful and systematic evaluation of process models by experimental
verification.

REFERENCES

Modern Plastics Magazine, January 1986.


M. R. Kamal, Polym. Eng. Sci., 14: 231 (1974).
M. R. Kamal, Proc. First International Conference on Reactive
Processing of Polymers, Pittsburgh, Pennsylvania, October 28—
30, 1980.
S. Sourour, Thermal and Kinetic Characterization of Thermoset­
ting Resins During Cure, Ph.D. Thesis, McGill University,
Montreal, Canada, 1978.
R. B. Prime in Analytical Colorimetry, Vol. 2, R. S. Porter
and J F. Johnson, Eds., Plenum Press, New York, 1970, p.
201.
K. E. J. Barrett, J. AppL Polym. Sci., 11: 1617 (1967).
S. Y. Choi, Soc. Plastics Engrs. J., 26: 51 (1970).
M. R. Kamal, S. Sourour, and M. Ryan, Soc. Plastics Engrs.
Tech. Papers, 19: 187 (1973).
M. R. Kamal and S. Sourour, Soc Plastics Engrs. Tech. Pa-
pers, 18: 93 (1972).
10 . M. R. Kamal and S. Sourour, Polym. Eng. Sci., 13: 59 (1973).
11 . A. Thakar, A. C. Hettinger, and K. C. Guyler, Soc. Plastics
Engrs. Tech. Papers, 24: 295 (1978).
12. M. E. Ryan, The Injection Molding of Thermosets, Ph.D. Thesis,
McGill University, Montreal, Canada, 1978.
13. A. Dutta and M. E. Ryan, J. Appi. Polym. Sci., 24: 635 (1979).
14. T. S. Chung, The Injection Molding of Phenolic Resins, Ph.D.
Thesis, State University of New York at Buffalo, Amherst, New
York, 1980.
15. M. E. Ryan and A. Dutta, Polymer, 20: 203 (1979).
16. C. D. Han and K.-W. Lem,J. Appi. Polym. Sci., 28: 3155 (1983).
17. K.-W. Lem and C. D. Han,J. Appi. Polym. Sci., 28: 3185 (1983).
18. K.-W. Lem and C. D. Han, J. Appi. Polym. Sci., 28: 3207 (1983).
19. J. F. Flynn and L. A. Wall, J. Res. Natl. Bur. Stand. , 70A: 487
(1966). ^
20. J. F. Flynn and L. A. Wall, Polymer Letters, 4: 323 (1966).
21 . R. N. Rogers and L. C. Smith, Thermochimica Acta, 1: 1 (1970).
Thermoset Injection Molding 373

22. B. Carroll and E. P. Manche, Thermochimica Acta, 3: 449


(1972).
23. R. B. Prime, Polym. Eng, Sci., 13: 365 (1973).
24. H. J. Borchardt and F. Daniels, J, Amer, Chem, Soc,, 79: 41
(1957) .
25. E. S. Freeman and B. J. Carroll, J. Phys, Chem,, 63: 394
(1958) .
26. G. O. Piloyan, I. D. Ryabchikov, and O. S. Novikova, Nature,
212: 1229 (1966).
27. H. L. Friedman, J, Macromol, Sci, -Chem,, 1: 57 (1967).
28. S. M. Ellerstein in Analytical Calorimetry, Yol, 1, R. S. Porter
and J. F. Johnson, Eds., Plenum Press, New York, 1968, p.
279.
29. H. L. Friedman, Polymer Letters, 7: 41 (1969).
30. H. E. Kissinger, Anal, Chem,, 29: 1702 (1957).
31. L. J. Taylor and S. W. Watson, ACS Polymer Preprints, 12: 457
(1971) .
32. J. M. Barton, J, Macromol, Sci, -Chem,, 8: 25 (1974).
33. T. Olcese, O. Spelta, and S. Vargin, J, Polym, Sci,, Symp, No.
53: 113 (1975).
34. C. D. Doyle, J. Appl, Polym, Sci,, 5: 285 (1961).
35. A . W. Coates and J. P. Redfern, Nature, 201: 6 8 (1964).
36. O. R. Abolafia, Soc, Plastics Engrs, Tech, Papers, 15: 610
(1969) .
37. J. R. MacCallum and S. Tanner, Europ, Polym, J,, 6: 1033
(1970) .
38. V. Satava, Thermochimica Acta, 2: 423 (1971).
39. J. R. MacCallum and J. Tanner, Nature, 225: 1127 (1970).
40. A. L. Draper, Proc. Third Toronto Symp. on Thermal Analy­
sis, 1969, p . 63.
41. R. A. W. Hill, Nature, 227: 703 (1970).
42. V. M. Gorbatchev and V. A. Logvinenko, J, Thermal Analysis,
4:475(1972).
43. E. L. Simmons and W. W. Wendlandt, Thermochimica Acta, 3: 498
(1972) .
44. J. R. MacCallum, Nature Phys, Sci,, 232: 41 (1971).
45. J. Sestak and J. KratochvH, J, Thermal Analysis, 5: 193 (1973).
46. J. Kratochvil and J. Sestak, Thermochimica Acta, 7: 330 (1973).
47. G. Gynlai and E. J. Greenhow, Thermochimica Acta, 5: 481
(1973) .
48. A. Dutta and M. E. Ryan, Thermochimica Acta, 33: 87 (1979).
49. S. Sidi, The Dynamics of Thermoset Injection Molding and the
Anisotropies of Molded Parts, M. Eng., McGill U n iv ., Montreal,
Canada, 1980.
50. S. Sidi and M. R. Kamal, Soc, Plastics Engrs, Tech, Papers,
26: 419 (1980).
37iJ Kamal and Ryan

51. S. Sidian dM . R. Kamal, Polym. Eng, S c i., 22: 349 (1982).


52. M. R. Kamal, V. Tan, and S. Sidi, Proc. 2nd World Congress of
Chem. Eng., Vol. 6 , 1981, p. 471.
53. D. L. Kerr and A. J. Dontje, Modern Plastics, 44: 147 (1966).
54. Spiral Flow of Low-Pressure Thermosetting Molding Compounds,
ASTM D3123-72, 1982, Annual Book of ASTM Standards, Part 35,
Philadelphia, 1982, p. 849.
55. S. Y. Choi, Soc. Plastics Engrs. J,, 26: 51 (1970).
56. P. E. Willard, Soc. Plastics Engrs, Tech, Papers, 19: 210
(1973).
57. Thermal Flow and Cure Properties of Thermosetting Plastics by
Torque Rheometer, ASTM D3795-79, 1982 Annual Book of ASTM
Standards, Part 35, Philadelphia, 1982, p. 968.
58. C. W. Macosko and F. G. Mussatti, Soc, Platics Engrs, Tech,
Papers, 18: 73(1973).
59. F. G. Mussatti and C. W. Macosko, Polym, Eng, Sci,, 13: 236
(1973).
60. R. P. White, J r., Soc, Plastics Engrs, Tech, Papers, 19: 192
(1973).
61. R. P. White, J r., Polym, Eng, Sci,, 14: 50 (1974).
62. L. T. Kale and K. F. O’Driscoll, Polym, Eng, Sci., 22: 402
(1982) .
63. Y. A. Tajima and D. Crozier, Polym. Eng. Sci,, 23: 186
(1983) .
64. J. S. Osinski, Polym. Eng. Sci,, 23: 756 (1983).
65. C. D. Han and K.-W. Lem, J. Appl, Polym. Sci., 28: 743
(1983).
6 6 . C. D. Han and K.-W . Lem, J. Appl. Polym. Sci., 28: 779
(1983).
67. K.-W. Lem and C. D. Han, J. Appl. Polym, Sci., 28: 79
(1983).
6 8 . R. E. Cuthrell, J. Appl, Polym. Sci., 12: 955 (1968).
P.. M. D. Hartley and H. L. Williams, Polym. Eng. Sci., 21: 135
(1981).
70. R. J. Farris and C. Lee, Polym. Eng. Sci., 23: 586 (1983).
71. R. J. Silva-Nieto, B. C. Fisher, and A. W. Birley, Polym,
Eng. Sci., 21: 499 (1981).
72. D. W. Sundstrom and S. J. Burkett, Polym. Eng. Sci., 21:
1108 (1981).
73. S. J. Perry, J. M. Castro, and C. W. Macosko, to be pub­
lished .
74. M. E. Ryan, Rheology of Polyester and Epoxy Liquids Dur­
ing Cure, M. Eng., McGill University, Montreal, Canada,
1973.
75. S. Y. Pusatcioglu, A . L. Fricke, and J. C. Hassler, J. Appl.
Polym. Sci., 24: 937 (1979).
Thermoset Injection Molding 375

76. S. Y. Pusatcioglu, A. L. Fricke, and J. C. Hassler, J. Appl.


Polym. Sei., 24: 947 (1979).
77. K. M. Hollands and I. L. Kalnin in Epoxy Resins^ H. Lee,
Ed., Advances in Chemistry Series, Vol. 92, American Chem­
ical Society, Washington, D. C ., 1970, p. 60.
78. C. W. Macosko and D. R. Miller, Macromolecules, 9: 199
(1976).
79. D. R. Miller and C. W. Macosko, Macromolecules, 9: 206
(1976).
80. G. P. Schmitt and J. P. Wiley, Soc. Plastics Engrs. Tech.
Papers, 30: 270 (1984).
81. S. A. Bidstrup and C. W. Macosko, Soc. Plastics Engrs.
Tech. Papers, 30: 278 (1984).
82. y . A. Tajima and D. G. Crozier, Polym. Eng. Sei., 23: 186
(1983).
83. T. H. Hou, Soc. Plastics Engrs. Tech. Papers, 31: 1253
(1985).
84. K. Ueberreiter and S. Nens, Kolloid-Z, 123: 92 (1951).
85. K. Ueberreiter and E. Otto-Laupenmuhlen, Laupenmuklen,
Kolloid-Z, 133: 26 (1953).
8 6 . J, N. Tomlinson^ D. E. Kline, and J. A. Sauer, Soc. Plas­
tics Engrs. Trans., 5: 44 (1965).
87. W. Knappe and O. Yamamoto, Kolloid-Z. Z. Polymere, 240:
775 (1970).
.
8 8 O. Yamamoto, Polymer J., 2: 509 (1971).
89. O. Yamamoto and H. Kambe, Polymer J., 2: 623 (1971).
90. B. L. Berman, R. P. Madding, and J. R. Dillinger, Phys.
Lett., 30A: 315 (1969).
91. L. E. Nielsen, J. Macromol. Sei., Revs. Macromol. Chem.,
C3: 69 (1969).
92. R. P. Krealing and D. E. Kline, J. Appl. Polym. Sei., 13:
2411 (1969).
93. D. E. Kline, J. Polym. S e t , 50: 441 (1961).
94. W. Knappe, Kunstoffe, 51: 707 (1961).
95. K, Kanari and T. Ozawa, Polymer J., 4: 372 (1974).
96. L. N. Cherkasova, Russ. J. Phys. Chem., 33: 1928 (1959).
97. M. Hattori, Kolloid-Z. Z. Polymere, 202: 11 (1965).
98. M. Hattori, Kobunshi Kagaku, 19: 32 (1962).
99. T. G. Fox and S. Loshaek, J. Polym. Sei., 15: 371 (1955).
100. K. Shibayama, Chem. High Polymers, Japan, 19: 219 (1962).
101. A. F. Lewis, Soc. Plastics Engrs. Trans., 3: 201 (1963).
102. R. A. Fava, Polymer, 9: 137 (1968).
103. P. G. Babayevsky and J. K. Gillham, J. Appl. Polym. Sei.,
17:2067(1973).
104. C. D. Han, Multiphase Flow in Polymer Processing, Academic
Press, New York, 1976.
376 Kamal and Ryan

105. A. B . Metzner, to be published.


106. D. G. Thomas, J. Colloid Sci,, 20: 267 (1965).
107. M. Mooney, J. Colloid Sci., 6: 162 (1951).
108. N. Eilers, Kolloid Z . , 97: 313 (1941).
109. N. A. Frankel and A. Acrivos, Chem. Eng, Sci., 22: 847 (1967).
110. J. S, Chong, E. B . Christiansen, and A. D. Baer, J. Appl.
Polym, Sci., 15: 2007 (1971).
111 . T. Kataoka, T. Nishimura, T. Katooka, and T. Sakai, Rheol,
Acta, 19: 671 (1980).
112. H. Tanaka and J. L. White, Polym, Eng, Sci,, 20: 949 (1980).
113. H. Tanaka and J. L. White, J, Non-Newt, Fluid Mech,, 7: 333
(1980).
114. J. A. Sneller, Modern Plastics, 63: 7, 48 (1986).
115. B. Elbirli and J. T. Lindt, Proc. Second International Confer­
ence on Reactive Processing of Polymers, p. 269, Pittsburgh,
Pennsylvania, November 2—4, 1982.
116. H. C. Hiner, M.S. Thesis, State University of New York at
Buffalo, Amherst, New York, 1984.
117. J. L. White, Polym. Eng. Sci., 15: 44 (1975).
118. M. R. Kamal, Y. Kuo, and P. H. Doan, Polym, Eng, Sci,, 15:
863(1975).
119. Y. Kuo and M. R. Kamal, AIChE J., 22: 661 (1976).
120. T. Smith, Soc. Plastics Engrs, Tech. Papers, 31: 1260 (1985).
121. S. Sourour and M. R. Kamal, Polym. Eng. Sci., 16: 480 (1976).
122, S, Sourour and M. R. Kamal, Thermochimica Acta, 14: 41
(1976).
123. M, E. Ryan and M. R. Kamal, Proc. Seventh International Con­
gress on Rheology, Gothenburg, Sweden, August 23-27, 1976,
pp, 290-292.
124. M. R. Kamal and M. E. Ryan, Polym, Eng. Sci. , 20: 859 (1980).
Rheological Behavior and Molding
Technology of Elastomers

H en ry S . - Y . Hsich and R ich ard J. Ambrose


Corporate Research Center
Lord Corporation
Cary, North Carolina

I. IN T R O D U C T IO N

The rheological properties of elastomers can be extremely valuable


when selecting elastomers for use as engineering materials and when
designing molds to process those elastomers into final product form.
The study of rheological properties can be divided into three distinct
branches: flow rheology of the uncured state, chemorheology (cure
kinetics), and relaxation spectra of the cured and uncured state.
The first branch deals with structural aspects of the uncured state,
the second involves the chemical transformation of molecules from the
uncured state to the cured state, and the third considers relaxation
aspects of the cured and uncured states.
The use of flow rheology and chemorheology information for elas­
tomer processing and molding operations can reduce operating costs
and improve production rates and product quality. In addition, the
use of chemorheology data and mechanical spectra of elastomers for
engineering material design allows one to optimize end-product per­
formance. Figure 1 is a scheme illustrating the interrelationships
between elastomer rheology and the use of elastomers as engineer­
ing materials. Dynamic modulus, mechanical damping, fatigue life,
and mechanical strength are important mechanical properties for con­
sideration in engineering material design. By manipulating the mech­
anical spectrum through compounding and chemorheology, one can
achieve any set of desirable mechanical properties. Such tailoring
of property sets is important to end-product performance.

377
378 Hsich and Ambrose

Fig. 1 Rheology of elastomers and elastomer applications.


Rheological Behavior and Molding Technology 379

This chapter reviews and discusses recent progress in elastomer


molding. In particular, the effects of filler on the flow rheology,
chemorheology, and mechanical properties are discussed. Such ef­
fects are pertinent because most elastomers used as engineering ma­
terials contain filler reinforcement. In addition, the relaxation spec­
trum of elastomers is given some attention because it provides infor­
mation about the dynamic mechanical properties, which in turn deter­
mine material applications. Finally, an attempt is made to describe
how the relaxation spectrum, flow rheology, and chemorheology of
elastomers can be effectively used in molding technology.

II. R E L A X A T IO N SPEC TR A OF POLYMERS

The relaxation spectrum of polymers is represented in Fig. 2. At


very high temperatures in the ’liquid flow” region un-eross-linked
polymers behave as viscous liquids. The temperature in this region
is commonly used in processing plastics. The temperature slightly

Fig. 2 The relaxation spectrum of polymers. (From Ref. 12.)


380 Hsich and Ambrose

below this region is the "rubbery flow" region. This is the process­
ing temperature region of un-cross-linked elastomeric materials. With
a further decrease in temperature, the "rubbery state" (or rubbery
plateau) region will appear. Recently, Hsich et al. [1,2] gave de-
tañed discussions about the relaxation times of cured and uncured
elastomers. As temperature further decreases from the "rubbery
plateau" region, the mobñity of the polymer chains decreases. In
this glass-transition region the relaxation time increases, and the
mechanical loss factor, tan 6 , of the polymers in this region is also
high. At the low-temperature end of the glass-transition region,
there is a glass-transition temperature, Tg, in which the relaxation
time is about 1800 sec [3—7]. The glass-transition temperature can
be determined as the intersection point between the slope of the mod­
ulus in the glassy state region and the slope in the glass-transition
region. Below the glass-transition region polymeric materials will be
in the "glassy state," and their relaxation times are very high.
Understanding the relaxation spectrum and the effective use of
relaxation times can lead to improvements in material processing. It
is common practice in polymer processing to increase the mechanical
strength of the polymer in question by inducing molecular orienta­
tion through stretching the molecules. However, in order to taüor
this supermolecular structure, one must be able to predetermine an
optimum relaxation time which is equivalent to determining a particu­
lar operating temperature. In this case the relaxation time during
molecular stretching must be longer than the time required to freeze-
in the desired structure during the cooling or quenching stages. If
this condition is met, once the structure is formed, the molecules can­
not reorient or relax back to their original unstretched state during
cooling. On the other hand, the relaxation time must be short enough
(this is equivalent to a particular temperature or modulus as shown in
Fig. 2) so that the molecules can be easily stretched without perman­
ently fracturing the material. The actual value will depend on the
operating speed of the process. As Fig. 2 shows, the optimum re ­
laxation time for molecular stretching is at the high-temperature end
of the glass-transition region and at a temperature somewhat above
the glass-transition temperature.
Another example where relaxation times are important is thermal
stress. To relieve thermal stress, one should anneal glassy poly­
mers at long relaxation times. Typically, a temperature just below
Tg is most satisfactory. In this way the thermal stresses in glassy
polymers can be easüy released without fear of reforming those
stresses during quenching to room temperature.
Mechanical spectra of polymers can also be used in engineering
material design. Snowdon [ 8 ] has discussed the use of mechanical
spectra of elastomers for designing systems needed to control
Rheological Behavior and Molding Technology 381

vibration and shock. Kelley and Williams [9] have proposed the ’’in­
teraction matrix” method for tailoring the mechanical properties and
the use of mechanical spectra in predicting the fatigue life of poly­
meric materials. More recently, Hsich et al. have used the mechani­
cal spectra of polymers to preprogram cure conditions to provide de­
sirable mechanical properties [ 1 , 2 , 1 0 , 1 1 ] and to vary filler content
to control mechanical properties [12,13].

A. T h e R u b b e ry S ta te

Flory [14] describes the behavior of rubbery materials as capable of


sustaining large deformations without rupture and of recovering spon­
taneously to very nearly their original dimensions after removal of an
applied stress. These two behavioral characteristics of rubber exist
because of two unique structural features. The first is the presence
of long polymer chains. In the unstressed state these chains exist as
random coil arrangements capable of rearranging to new configurations.
Thus, when rubber is subjected to an externally applied stress, a
large deformation can be accommodated merely through rearrangement
of the configurations of the individual polymer chains. The second
structural feature unique to rubber is the presence of a network struc­
ture. The network structure exists because of physical (entangle­
ments) or chemical (covalent bonds) cross-links. The physical en­
tanglements allow chain slippage, which in turn disrupts the network.
Chemical cross-links are required for the formation of a permanent
network structure. It is this permanent network structure that pro­
vides the rubber with the ability to recover spontaneously to very
nearly its original dimensions. This occurs primarily because of the
high restoring force associated with the permanent network structure.
If one accepts the criteria of ability to recover to its initial con­
figuration after undergoing a large deformation as defining the ru b ­
bery state, then unvulcanized elastomers would not be viewed as part
of the ideal rubbery state. However, the statistical theory of rubber
elasticity does not distinguish between physical entanglements and
chemical cross-links in its treatment of rubber networks. This is just
another example of the difficulty in attempts to define a rubbery state.
Rubbery materials resemble ’’viscous” liquids regarding their deform-
ability so as to slow down rupture but act like solids in their capacity
to recover from large strain deformations.

B. A H y b rid Model fo r Mechanical S pectra


o f Filled and U n filled Elastomers

Elastomers have been used as engineering materials in applications such


as shock and vibration isolators for many years. In spite of this we
382 Hsich and Ambrose

still lack a comprehensive understanding of the mechanical properties


of elastomers at various filler loadings. The statistical theory of ru b ­
ber elasticity does describe the variation of modulus with temperature
adequately. According to this theory, the modulus of ideal rubbers
increases with increasing temperature. An ideal rubber is defined as
a rubber which is capable of storing all energy during deformation
with no mechanisms for energy dissipation. However, most elasto­
meric compounds do not behave as ideal rubbers over a broad serv­
ice temperature range. Therefore, a study of the mechanical spec­
tra of elastomers is needed to determine the temperature dependence
of mechanical properties.
Flory [14] has stated that elastomeric materials can be considered
to be in a nonrelaxation state and to have a perfect network structure.
The perfect network is defined as having no free chain ends; i . e . ,
the primary molecular weight M is infinite. Then from the statistical
theory of Gaussian networks in an ideal rubbery state [17], Young’s
modulus E of a rubbery material can be expressed as

E T ( 1)
Me

where Me is the molecular weight between cross-links, p is the den­


sity, R is the gas constant, and T is the absolute temperature.
For an ideal rubber, E is equal to three times the shear modulus
G. However, in practice, all networks have free chain ends which
may be regarded as flaws in the structure. By considering these
network defects, Flory [14,18] assumed that any real network must
contain terminal chains bound at one end to a cross-link and termi­
nated at the other by the end (free end) of a primary molecule. Then
we have

/ 2M \
E = ^ T ll (2)
\ M

As we can see from Eqs. (1) and (2 ), the modulus of an ideal rubber
is linearly proportional to the temperature. In real elastomers ideal
behavior is not exhibited untR the temperature is much higher than
Tg.
Hsich et al. [12] investigated the mechanical spectra of both filled
and unfilled natural rubber compounds. Their study illustrated the
behavior of natural rubber to be ideal at low filler loadings [modulus
was predictable from Eqs. (1) and (2 )] but nonideal at medium to high
filler contents.
Rheological Behavior and Molding Technology 383

The mechanical properties of elastomers depend on the environ­


mental temperature, the glass-transition temperature, and the shape
of the relaxation spectrum. Therefore, the phenomenon of glass tran­
sition must be included in the interpretation of the mechanical spec­
tra in the environmental temperature range of interest.
In the glass-transition region thermodynamic, physical, mechani­
cal, electrical, and chemical properties of polymers undergo striking
changes. The most obvious changes occur in mechanical properties.
For example, modulus changes by a factor of 10^ over the glass-tran­
sition region. There are many theories and hypotheses which are de­
rived either from thermodynamic or relaxation aspects on the glass
transition [3--7,19]. None of these explanations by themselves can
be used to fully characterize glass transition. A complete explana­
tion probably lies in a blend of these theories. The most popular
theory based on the glass transition is the Williams-Lan del-Ferry
(WLF) equation [20]. In this equation, a,p is the shift factor, de­
fined as the ratio of the relaxation time at temperature T to the re­
laxation time at Tg for a polymer in the temperature range from Tg
to Tg + 100 K. The expression for ax can be written as

-C i(T -T )
In a = — ^ (3)
^T C, + (T -T ^ )

In Eq. (3 ), Ci and C 2 are constants which relate to the free volume


and the difference of thermal expansion between the liquid and the
glassy state of the polymer, respectively. In fact, the WLF equation
can be correlated with Doolittle’s free-volume theory on viscosity in
the glass-transition region [21]. Then the viscosity n at any tem­
perature T is

-C i(T -T ^ )
n(T)
In (4)
n(T„) C 2+(T-T^)

At temperatures above Tg and for nonideal rubbery behavior, G* and


G', the complex and storage dynamic moduli, decrease with increasing
temperature similar to the steady flow viscosity. Therefore the WLF
form was modified by Hsich et al. [12] for the prediction of G* or
G'.

-C i(T -T )
G (T )
In (5)
G(T ) C2 + (T-T )
384 Hsich and Ambrose

where G (T g ) is the modulus at Tg. From Eq. (5) we see that the
modulus decreases with increasing temperature in the glass-transi­
tion region. However, for elastomeric materials that behave as an
ideal rubber, the modulus would increase with increasing tempera­
ture, as shown in Eqs. (1) and (2 ). It was this type of reasoning
that led to the development of the following hybrid model [ 1 2 ] , which
includes aspects of both glass transition and rubber elasticity.

-C i(T -T )
Tl
G (T ) + G (T ^ ) - G ( 6)
C 2 + (T -T )
- g J

where G q is the shear modulus at a reference temperature T q, the


lowest temperature for an ideal rubbery state. In this hybrid model,
if G (T g ) is much larger than G q (or G (T g)/ T » G q/Tq), then Eq. ( 6 )
becomes the form of the WLF equation shown in E q. ( 5). On the other
hand, if the temperature in question is higher than the glass-transi­
tion temperature ( i . e . , T > T g) and Ci is large, then Eq. ( 6 ) has the
form of the elastic modulus expression in Eqs. (1) and (2 ). This is
true because the exponential term in Eq. ( 6 ) becomes small. When
the values of G (T g)/T to G q/Tq and T to Tg are intermediate to those
extremes discussed above, the WLF and/or elastic modulus equations
are inappropriate by themselves for describing the polymer mechanical
spectrum. It is in these instances when a combination of the two (the
hybrid equation) is needed.
The results of calculations from the hybrid model [ 6 ] are plotted in
Figs. 3—6 along with the experimental data which were measured at
an angiilar frequency of 10 rad/sec. The parameters used to make
these theoretical calculations shown in Figs. 3—6 are listed in Table
1. There is excellent agreement between the experimental results and
those calculated from the hybrid model at all filler loadings. Results
obtained from the modified WLF equation [Eq. (5 )] also are shown in
Figs. 3—6 for the purpose of comparison. The modified WLF equa­
tion, even though not intended to describe the mechanical spectra of
elastomers, does very well at high filler loadings (Fig. 6 ). This is
probably due to the broadening effect that high filler loadings have
on the mechanical spectrum. In Figs. 3 and 4, at temperatures con­
siderably above Tg, the rubber behaves as if in the ideal rubbery
state. The hybrid equation [Eq. ( 6 )] models such behavior and shows
an increasing modulus in the rubbery region. An expanded tempera­
ture scale in Fig. 3 illustrates this more clearly. When filler loading
increases to 30 phr, the transition region from glassy to rubbery b e ­
comes broad (Fig. 5). At this point the rubbery plateau region is
1 0 ^0 - I I .. ■_L..1-L_l 1 . I I I I I I I I ..1.. I I ...1 -L . 1 1 I III I J I I I I I I I I I I I I I I I I I1
73
zr
a>
o_
o
o*
9L
03
n>
IT
1 0 ^- 0)
<
o*
Q)
3
a
Cvn o_
o a
5’
ID
^ 1 Q8 -
>- H
Q a>
n
CD
zr
3
O,
o
CD

10 ^ —

TEMPERATURE (°C)
1—7—T
—I I ( I I I I f I I I I I I I I—I I I I I I—I I I I j I r I I I I I I I I I I I I I I I r i I I I TI I r
-6 0 0 100 160 ^

Fig. 3 G W s . T for natural rubber at 10 phr filler loading. (From Ref. 12.)
<T>

X
i£.
n‘
X
D>
D
a
>
3
XT
-s
O
-6 0 100 160 fP
Fig. 4 vs. T for natural rubber at 20 phr filler loading. (From Ref. 12.)
IT
0
1
n
9L
00
n>
D“
Q)
<

0)
3
a
o,
a
5*
H
fl>
n
3-
3
O
O
CQ

CO

Fig. 5 G' vs. T for natural rubber at 30 phr filler loading. (From Ref. 12.)
10 10 - _l_1— 1_L I I_I_1— 1_I I I ! _J_Lj_I_L_J_L I I I I I I I I I I I I I I I I I I I I I I I I I I
($) 00
00
4 50 phr
(•) Experimental Data
...... WLFEq.
----- Hybrid Eq.

109-
\
«è

O
O)
^ 108-
>
Q
.m.
h

107 —

o*
□r
0)
3
a
TEMPERATURE (»C)
>
3
106- I II II I I tI I I I I I { r.r.I..[ [ I I 1 I j~ T ~ r I i pi i i i | i i i.i I I I I I I I I I I I [ I I I T- D*
O
-6 0 0 100 160 0)
fD

Fig. 6 vs. T for natural rubber at 50 phr filler loading. (From Ref. 12.)
Rheological B ehavior and Molding Technology 389
^

T a b le 1 Parameters for Hybrid Model

Hybrid equation WLF equation


Filler Go/ To
loading ^ 2 (dyn-cm"^ Ca
(p h r) Cl (K -i) K -i) Cl (K '^ )

0 11.59 18.40 3.0 X 10“ 6.45 3.52


1 0 10.45 16.81 3.7 X 10“ 6.32 4.36
2 0 8.53 15.66 5.0 X 1 0 “ 5.93 5.44
30 7.45 14.84 7.4 X 10“ 5.50 7.28
40 5.66 11.39 4.4 X 10“ 5.20 9.20
50 4.99 10.84 0 4.99 10.84

overshadowed due to the broadening phenomenon. Indeed, at high


filler loading (50 p h r ), the hybrid equation is essentially the same as
the WLF equation. The data in Fig. 6 substantiate this phenomenon.
The broadening of the relaxation spectrum with increasing filler load­
ings for cured rubbers agrees with previous results on the viscosity
of uncured rubbers [22]. A broadening of the relaxation spectrum by
increasing filler loadings also has been found in studies of the cure
reaction [ 1 , 2 , 1 0 ].

C. Composite Mechanics o f Filled Elastomers

We have discussed the moduli of filled and unfilled elastomers as a


function of temperature. There is also a practical interest that the
mechanical properties of filled elastomers be expressed as a function
of filler loading. There have been many studies on mechanical prop­
erties of filled polymers. Recently, Chow [23] has given an exten­
sive review on the models proposed in these studies. These models
generally give reasonable correlations between modulus and filler for
specific temperatures; however, these correlations rarely reflect tem­
perature dependence or quantify polymer-filler interactions over a
broad temperature range. The hydrodynamic model of filler effects
on viscosity, which was developed by Einstein [24] and later modified
390 Hslch and Ambrose

by Guth and Gold [25], has been extended to explain these same ef­
fects on modulus [14,26]. The hydrodynamic model is

G = God + 2.5(i)) (7)

or

G = God + 2.5(|) + 14.1(j)^) ( 8)

where G and Go are the moduli of the filled and unfilled elastomer , re ­
spectively, and (j) is the volume fraction of filler.
In a study of the modulus of filled elastomers, Hsich et al. [13]
found that a linear relationship exists between modulus and filler load­
ing at all temperatures when the data are plotted semilogarithmically.
They expressed the shear storage modulus G’ as a simple function of
filler loading ^ by

GH(j))
In K((j) - (|)o) (9)
G’(cj)o)

where G’(({)) and G’((j)o) are the shear storage moduli at a filler load­
ing (j) and a reference filler loading cf)o, respectively, K is the slope
obtained from plots of In G’ versus and measures the degree of poly­
mer-filler interaction. Experimental measurements were made at a fre ­
quency of 10 rad/sec. Values of K obtained from plots of In G’ v e r­
sus [13] are replotted against temperature in Fig. 7, which shows
that K is much more temperature sensitive in the glass-transition re­
gion (-60° to -45°C) than in either the rubbery or the glassy regions.
It was concluded that the degree of polymer-filler interaction in the
glassy region is much less than the degree of polymer-filler interac­
tion in the rubbery region. This large difference in polymer-filler
interaction between these two regions is reflected by the high sensi­
tivity of K in the glass transition region.
The slope K was modeled effectively by considering all three tem­
perature regions separately. For example. Fig. 7 shows that K de­
creases slightly with increasing temperature in the rubbery region
(-40° to +120°C). An equation describing this behavior is

K = a - b In T ( 10 )

where a and b are constants and T is the temperature in kelvins. In


the glass-transition region the model of K is more complicated because
of its high temperature sensitivity. The modulus depends on tempera­
ture much the same as viscosity does. In an earlier report [22] Hsich
showed the model of viscosity for filled elastomers at constant shear
rate or frequency to be
73
IT
O

n
9L
00
a>
3“
Q)
<
O*
"5
0)
3
a

2 .
a
5’

H
fD
O
3-
3
2 .
o

w
co
392 Hsich and Ambrose

In ^ = C((^ - (T - T ) ( 11)
Ho J- ^ &

where n and no are viscosities cj) and c|)o, respectively, and the other
symbols have their usual meanings. Similarly, a model for K in the
glass-transition region is

K = (T - T ) ( 12)
g

where c is a constant. Finally, in the glassy region

K = d (13)

for some constant d. Values of constants a, b , c and d are listed in


Table 1. Values of K calculated from Eqs. (10), (12), and (13) from
-100 to +120®C are also plotted in Fig. 7. Figure 7 shows that there
is good agreement between the fitted and experimental values for K.
Using the constants in Table 1, we obtain a K value of about 0.8 in
the temperature region from -100 to -60®C. K increases dramatically
from 0.8 to 5,0 in the glass-transition region from -60 to -40°C, but
this slowly drops to 4.0 in going from -40° to 120°C.
This study [13] indicates markedly different effects of filler on
modulus, depending on the temperature region in question. The e f­
fect of filler on modulus is much greater in the rubbery region than
in the glassy region. This is exemplified by the high-tern per ature
sensitivity of K in the glass-transition region. These filler effects
have been successfully modeled according to the temperature range
of interest, i.e ., glassy region, glass-transition region, and ru b ­
bery region. There is excellent agreement between the filler effects
measured experimentally and those predicted from these models.

III. FLOW RHEOLOGY OF ELASTOMERS

As injection molding of elastomers becomes an increasingly important


molding process, an understanding of elastomeric rheological behavior
is paramount. Effective mold designs as well as the ability to optimize
molding conditions will depend on such an understanding.
Flow rheology of elastomers has been studied in connection with the
quality control of elastomers and in the design of processes and equip­
ment. Experimental studies on the rheological behavior of elastomers
during mixing and milling have been given by White [27—29], Tokita
and White [30], and Bolen and Colwell [31]. The flow of elastomers
Rheological Behavior and Molding Technology 393

during extrusion has also been studied by Vila [32] and Zamodits and
Pearson [33].

A. V isco sity T h e o ry o f Polymers

Many models have been proposed for explaining the viscosity of liq­
uids [34]. All of these models fail to explain the viscosity behavior
of complex liquids. Eyring [35] proposed a rate theory for viscosity
in which viscous flow is considered to be a rate process controlled by
the free energy required to overcome a potential barrier. According
to this theory, each molecule in the liquid medium is located in a po­
tential ’’well.” Holes exist in the liquid medium due to irregularities
in arrangements of the molecules. The primary flow process is con­
sidered to be the ’’jump” of a molecule from its potential energy well
into a neighboring hole, i . e . , from one equilibrium position to an­
other. Eyring’s theory can be expressed mathematically as

hN fASl rE 1
(14)
=V ^^p L r -J ®^p Lrt J
where h is Planck’s constant, N is Avogadro’s number, V is the molar
volume, R is the gas constant, T is the absolute temperature, AS is
the entropy, and E is the enthalpy. Equation (14) demonstrates a
marked increase in viscosity as the temperature is lowered. This in­
crease in viscosity is related to the reduction in the molecular mobil­
ity of the flowing units. This mobility is determined by the time nec­
essary for the molecules to accumulate sufficient thermal energy to
overcome the energy barrier separating them from nearby holes. A
major success of this theory is that it leads to a viscosity proportional
to exp(E/RT). This is the Arrhenius equation which has been long
observed to hold for many liquids.
Fox and Flory [36—39] discussed the dependence of the molecular
mobility near the glass-transition temperature, using free-volume con­
cepts. In their approach the probability of a jump by a polymer mole­
cule into a neighboring hole is not determined by the rate with which
that molecule can overcome an energy barrier. Instead, the probabil­
ity of a jump is determined by whether or not sufficient free volume
exists to accommodate the jumping molecule. Based on this free-vol­
ume concept, Doolittle [21] proposed the empirical equation for vis­
cosity

In n = In (15)
394 Hsich and Ambrose

where n is viscosity, A and B are empirical constants, V is the molar


volume, and Vf is the free volume. Later, Williams, Landel, and Ferry
[ 2 0 ] modified the free-volume model to give what is now known as the
WLF equation:

-Ci(T-T )
_n ^ ^ g
In (16)
n„ C2 + ( T - T )
g g

where rig is the viscosity at Tg and Ci and C 2 are constants. In 1925


Fulcher [40] proposed a model similar to Eq. (16) for predicting the
viscosity of glasses.
According to Eyring [35], E in Eq. (14) is often about one-third
to one-fourth of the energy of vaporization. In 1938 Ewell [41] dis­
covered that the value of E for the flow of raw rubber between 70°C
and 140°C was much lower than one-third to one-fourth of the energy
of vaporization predicted by Eyring. Ewell suggested that for long-
chain molecules, the elementary process for viscous flow consisted of
the displacement or jump of only a small segment of the molecule rather
than the displacement of the entire molecule as a unit.
Recently, Litt [42] developed a modified viscosity equation which
fits the data both for experimental free volume and viscosity with a
minimum number of adjustable parameters.
In 1952 Bueche [43] proposed that the viscosity of polymers is a f­
fected by intramolecular valence bonds and intermolecular chain en­
tanglements connecting the segmental flow units. Based on this en­
tanglement coupling, Bueche concluded that the viscosity has an
abrupt change in slope at a critical molecular weight Mq, which can
be identified as the point at which network formation through chain
entanglements becomes possible. At low molecular weights (M < Mq)
the viscosity is proportional to the first power of the molecular weight
(ri oc M). At high molecular weights (M > Mq) the viscosity is propor­
tional to the 2.5 power of the molecular weight (n « M^*^). Later,
Berry and Fox [44] found that the viscosity indeed is proportional
to the first power of the molecular weight at M < Mq; however, at
M > Mo the viscosity is proportional to the 3.4 power of the molecu­
lar weight (n “ M^'^).
According to thermodynamic fluctuation theory, the properties of
materials can be expressed as a relaxation function in the Fourier
components of wave-vector space [45,46]. Since polymer fluids are
relaxational liquids, the shear modulus and viscosity can be written
as [46-48]

oc

G'(o)) = Go + (G (17)
L 1 +
H i
Rheological Behavior and Molding Technology 395

dx (18)
G "<") = ( « . - “ • > / „

„ . < „ , = 2 :1 « and (19)


Ü) 0 )
00

He = lim n*' = (G - Go) f — - V - 2 g / — \ dx ( 20)


“ -/q I + u t \^t^I
o)->0

where GKoo), G”(o)), ri^(w)» and n’^(ijo) are the real and imaginary com­
ponents of the shear modulus and viscosity, respectively, ns is Ihe
steady-state viscosity, is the relaxation time, and g(x/xp) is the
normalized relaxational distribution function. G q and Gc» are shear
modiili at frequencies of zero and infinity, respectively. Equations
(17) —(20) allow us to study the modulus and viscosity as a function
of frequency under dynamic conditions. In steady-state measure­
ments of viscosity, it has been found that the viscosity also depends
on the shear rate, y [48]. According to Bueche [49], the dynamic
viscosity n’(w) is equal to the steady shear viscosity n(Y) at a shear
rate Y = o). Coleman and Markovitz [50] have shown that

lim n’(oo) = lim h( y ) ( 21)


->0
(jo y -^0

More recently, Kulicke and Porter [51] have found that the rheo­
logical behavior of polystyrenes and polyacrylamides satisfies the re­
lationship of Cox and Merz [52]

n*(co) = [n^(w)^ + n’’( c o ) ^ ] = n(Y) ( 22 )

It is convenient to express viscosity and shear rate in a power law for


practical applications [53,54]:

n- 1
(23)
n = Ay

where A represents the temperature dependence of viscosity and n rep­


resents the power-law index for this viscosity model. This power law,
though convenient, is an accurate measure of the shear-rate depen­
dence of viscosity only in narrow temperature and shear-rate ranges.

B. V iscosity S tudies o f Elastomers and S h e a r-F lo w -


Induced S tru c tu ra l Changes o f Molecules

Capillary rheometry is a common method used to study the viscosity of


elastomers. A detailed discussion of capillary rheometry is given by
396 Hsich and Ambrose

Walters [53] and Han [54]. The viscosity characterization of elasto­


mers is difficult because of the high elasticity, long-chain structure,
filler dispersion problems, and scorch tendencies of these materials.
Complicating these measurements further are factors such as viscos­
ity corrections due to capillary entrance pressure drops, elastomer
slippage at the capillary wall, and viscous heating effects that de­
pend largely on the die configuration. In spite of these problems,
however, capillary rheometer measurements remain useful as a means
to carry out the rheological characterization of elastomers for a vari­
ety of applications.
Recently, Hsich et al. [55] studied the viscosity of silicone elas­
tomers using both capillary and dynamic measurement techniques.
They found that the steady-state viscosity of unfilled silicones as
measured by capillary rheometry approached values of the complex
viscosity n* at 80°C. As temperature was increased to 160°C, the
steady-state viscosity approached values of the dynamic viscosity
x]\ For filled elastomers this trend was nearly reversed. As tem­
perature was increased from 80° to 160°C, the steady-state viscos­
ity approached values near those of n*. These data are shown in
Figs. 8—11 where ri*, ri’ , and the steady-state viscosity are all plot­
ted. Hsich et al. [55] explain that the elastic behavior contributed
from the unfilled polymer decreases with increasing temperature.
The unfilled polymer appears to behave as a viscous liquid at ele­
vated temperatures. At high filler loadings this behavior is not
observed.
Shear rate is an important variable that must be considered in
polymer processing. Hsich et al. [55] observed further that melt
fracture is a dominant factor at high shear rates for unfilled sili­
cone elastomers. At increasing filler contents^ however, the ten­
dency for melt fracture is reduced. Plots of die swell versus shear
rate and extrusion pressure versus shear rate illustrate these phe­
nomena in Figs. 12—17. It can be seen that die swell decreases
drastically above a certain minimum shear rate in unfilled silicone
elastomers. This decrease represents the onset of melt fracture
(Figs. 12-13). This tendency for melt fracture is reduced dramati­
cally at increased filler loadings (Figs. 14—17). In addition, un­
usually large increases in extrusion pressure were observed above
a certain minimum shear rate for filled silicone elastomers. We have
no specific explanation for these increases at this time. Folt et al.
[56] observed similar increases in extrusion pressure in their stud­
ies using natural rubber. Folt explained these increases by shear-
induced crystallization.
Finally, an interesting phenomenon for silicone elastomers was ob­
served [55] while measuring both extrusion pressure and die swell at
Rheological Behavior and Molding Technology 397

Fig. 8 Viscosity vs. frequency or shear rate for silicone rubber with
0 phr filler loading at 80°C. (From Ref. 55.)

moderate shear rates (near 100 sec"^). In this shear-rate range in­
creases in extrusion pressure were minimal (Figs. 18—20). In fact,
extrusion pressure actually decreased slightly in one experiment (Fig.
20). In this same shear-rate range, die swell showed abrupt increases
(Figs. 18—20). Such reductions in extrusion pressure are sometimes
explained in terms of slippage. However, when slippage occurs, we
would expect a reduction in die swell as well. In fact, we observed
an accompanying increase in die swell. This fact, as well as the ex­
treme reproducibility of these data over a very narrow shear-rate
range, led us to believe that a molecular phenomenon is involved.
Molecules uncoiling would lead to a reduced extrusion pressure. Re­
coiling of these molecules, once the applied stress is removed by ex­
trusion through the die, would give an increase in die swell.
398 Hsich and Ambrose

F ig . 9 Viscosity vs. frequency or shear rate for silicone rubber with


0 phr filler loading at 160°C. (From Ref. 55.)

C. E ffects o f F ille r on th e V isco sity o f Elastomers

As mentioned in Section II.C , Einstein [24] studied the effect of filler


on viscosity. Later Guth and Gold [25] modified Einstein^s model to
include a second-order power of (|). Lee [57] extended the hydrody­
namic interactions between spheres to include three-body collisions
of spheres. The viscosity model then becomes a power series expan­
sion up to a third-order power of cj).

T] = noCl 2.5(j) + K 2 (I> + Kscj)^) (24)

where K 2 and K 3 are constants, and no is the viscosity at zero filler


loading. Landel et al. [58] developed the following empirical equa­
tion by studying the suspension of particles in Newtonian liquids:
Rheologica! Behavior and Molding Technology 399

Fig. 10 Viscosity vs. frequency or shear rate for silicone rubber with
33 phr filler loading at 80°C. (From Ref. 55.)

(25)
■ - m '

where (j)m is a packing factor. Recently, Pliskin and Tokita [59] gen­
eralized LandeFs equation by considering the degree of orientation of
the system. The modified equation becomes

■N
n = rio(l - (26)

where cj) 0 is the volume fraction of rigid material and N is the parame­
ter describing the degree of orientation of the system. For a com­
pletely laminated material, N = 1.0, and for random suspension of
400 Hsich and Ambrose

SHEAR RATE (1/Sec) or FREQUENCY (rad/sec)

Fig. 1 1 Viscosity vs. frequency or shear rate for silicone rubber with
33 phr filler loading at 160°C. (From Ref. 55.)

discrete particles, N = 2.5. The above equations work well for some
liquids at limited temperatures and shear rates.
Hsich [22] studied the effects of fillers on the viscosity of natural
rubber. He found the dependence of elastomer viscosity on filler was
much more pronounced at high temperatures compared to low tempera­
tures. This behavior suggests that the glass transition may play an
important role in controlling the rheological behavior of elastomer-
filled composites. His studies indicate further that Tg was not af­
fected by filler loading, but the transition zone from the glassy to
rubbery state on the temperature scale is broadened by increasing
filler contents. Hsich also concluded that elastomer viscosity de­
pends not only on the temperature at which the viscosity is mea­
sured but also on the glass-transition temperature of the elastomer
Rheological Behavior and Molding Technology 401

Fig. 12 Extrusion pressure and die swell vs. shear rate for silicone
rubber with 0 phr filler loading at 110°C. (From Ref. 55.)

in question. These studies resulted in a viscosity model for elasto­


mers that described both the effects of filler and temperature:

T
In C (( (T - T ) (27)

where n is the viscosity of the filled elastomer at a temperature T, C


is a constant which depends on the surface of the filler, (})(j^3 is the
weight ratio of filled to unfilled polymer, (j)r is this weight ratio at a
reference filler loading, and rip is the reference viscosity at filler
loading
Experimental data (points) and theoretical predictions (solid lines)
from Eq. (27) are given in Fig. 21. The experimental curves were
402 Hsich and Ambrose

Fig. 13 Extrusion pressure and die swell vs, shear rate for silicone
rubber with 0 phr filler loading at 135°C. (From Ref. 55.)

obtained at a frequency of oo = 10 rad/sec. The predicted curves were


obtained for C = 8.89 x 10“^ K"^. There is excellent agreement be­
tween experiment and theory.

IV . CHEMORHEOLOGY OF ELASTOMERS

As discussed earlier, a characteristic of rubberlike behavior is the ca­


pability of the material to recover spontaneously to its initial dimen­
sions with no appreciable fraction of deformation remaining perman­
ently after removal of the stress. The magnitude of this restoring
force in rubbery materials is proportional to the number of network-
supporting polymer chains per unit volume of rubber. A supporting
chain is a segment of polymer backbone between network junctures.
Rheological Behavior and Molding Technology 403

-20 Jt!

F ig . 14 Extrusion pressure and die swell vs. shear rate for silicone
rubber with 33 phr filler loading at 110°C. (From Ref. 55.)

An increase in the number of network junctures increases the number


of supporting chains. Vulcanization produces network junctures by
the insertion of chemical cross-links between polymer chains. These
cross-links may be chains of sulfur atoms, carbon-carbon bonds,
polyvalent organic radicals, or polyvalent metal ions.

A. Chemical Reactions of C u re

The cure of elastomers has been reviewed recently by Wise [60] and
Morrell [61]. Coran [62] gave an extensive review on various cross-
linking agents for different elastomers. The most commonly used
cross-linking agents and the elastomers that these agents cure can be
summarized as follows:
404 Hsich and Ambrose

Fig. 15 Extrusion pressure and die swell vs. shear rate for silicone
rubber with 33 phr filler loading at 135®C. (From Ref. 55.)

Sulfur: Sulfur was the first agent used to vulcanize the first
commercial elastomer (natural ru bb e r). The elastomers suit­
able for sulfur vulcanization include natural rubber, synthetic
polyisoprene, polybutadiene, styrene-butadiene rubber, ni­
trile rubber, butyl rubber, and ethylene-propylene terpoly-
mers.
Peroxides: Most elastomers can be cured by the action of or­
ganic peroxides. However, many of the properties are some­
what inferior to those cured by accelerated sulfur. Neverthe­
less, peroxide curatives are most suitable to silicone rubber,
ethylene-propylene copolymer, and diene elastomers in appli­
cations where creep resistance at elevated temperatures is re ­
quired.
Rheological Behavior and Molding Technology 405

80- . -40

60PHR FILLER - IIQOC

60- ■30

40- -20 g

20- -1 0

n—I—r'"'"r'T-r.[. “T'T 'i.r.


10^ 102 103
SHEAR RATE (1/Sec)

Fig. 16 Extrusion pressure and die swell vs. shear rate for silicone
rubber with 60 phr filler loading at llO^C. (From Ref. 55.)

3. Metal oxides: Polychloroprene, fluorocarbon elastomers, and


chlorosulphonated polyethylene are generally cured by the ac­
tion of metal oxides.
4. Phenolic resins, quinone derivatives, and maleimides: Diene
elastomers can also be cured by these curatives or a combina­
tion of these curatives with sulfur, depending on the desired
application.

Two of the most interesting elastomers of those mentioned are nat­


ural rubber (Hevea) and silicone elastomers. The chemistry of nat­
ural rubber cure is quite complex and therefore difficult to completely
understand. Nevertheless, in the last two decades rubber scientists
have unraveled some of the major reactions that take place during
406 Hsich and Ambrose

Fig. 17 Extrusion pressure and die swell vs. shear rate for silicone
rubber with 60 phr filler loading at 135°C. (From Ref. 55.)

cure. For example, the accelerator reacts with sulfur to give mono­
meric poly sulfides, Ac--Sx—Ac, where Ac is an organic radical de­
rived from the accelerator. The poly sulfides can interact with ru b ­
ber to give polymeric poly sulfide s , R u bber-S x “ A c . The rubber
polysulfides then react, either directly or through a reactive inter­
mediate, to give cross-links or rubber poly sulfides. Rubber—Sx“ Rub­
ber [61-65].
Vulcanization by sulfur alone is no longer commercially significant
because of the long cure time involved. Instead, different accelera­
tors are used in combination with sulfur. These accelerators will pro­
vide different scorch rates (premature cure) and cure rates during
cross-linking. The cross-link types and chain modifications that form
during the vulcanization of natural rubber are present as monosul-
fidic, disulfidic, polysulfidic, cyclic monosulfidic, and cyclic disul-
fidic moieties.
Rheological Behavior and Molding Technology 407

Fig. 18 Extrusion pressure and die swell vs. shear rate for silicone
rubber with 33 phr filler loading at 160®C (L7D = 30). (From Ref.
55.)

If desulfuration proceeds rapidly, the final network will be highly


cross-linked with mainly monosulfidic bonds. Such a network is
termed efficiently cross-linked. On the other hand, if desulfuration
proceeds slowly, there will be opportunities for thermal decomposi­
tion. These networks are inefficiently cross-linked. During the vul­
canization of natural rubber , if the temperature rises above 160°C,
reversion, degradation of polymer chains or network, and inefficiently
cross-linked networks are possible.
Silicone elastomers are widely used in the aerospace, automotive,
and electronic industries. These elastomers are commonly cured with
organic peroxides. They have exceptional mechanical and electrical
properties under extreme temperature conditions. The common types
of silicone resins include poly (dim ethyl siloxane) , poly (dimethyl vinyl-
sRoxane), and poly(dimethylvinylphenyl) silicone rubber [66]. When
408 Hsich and Ambrose

SHEAR RATE (1/Sec)

Fig. 19 Extrusion pressure and die swell vs. shear rate for silicone
rubber with 40 phr filler loading at 135°C (L/D = 16). (From Ref.
55.)

less than one-half of 1% of the methyl groups are replaced by vinyl


groups, the resulting polymer makes efficient use of peroxide vulcan­
ization agents. The resulting shoxane network resists reversion and
has low compression set. Consequently, nearly all commercial silicone
gums contain the vinyl-modified polymers. Pure dimethyl siloxane
elastomers also tend to become stiff below -50°C because of crystal­
lization. Phenyl groups are frequently used to substitute for some
of the methyl groups to improve the low-temperature flexibility. Re­
placement of only 5% to 10% of the methyl groups by phenyl groups will
lower the crystallization temperature and extend the useful service
temperature range to below -100°C. For these reasons, dimethylvinyl-
phenyl and dimethylvinyl are the most popular commercial siloxane
elastomers.
Rheological Behavior and Molding Technology 409

-2 0 S

Fig. 20 Extrusion pressure and die swell vs., shear rate for silicone
rubber with 40 phr filler loading at 160^C (L/D = 16). (From Ref.
55.)

The effects of curing on the properties of these elastomers is shown


qualitatively in Fig. 22. In general, an increase in cross-link density
will increase modulus and hardness, but decrease hysteresis, compres­
sion set, and damping. However, tear strength, tensile strength, and
fatigue life have maximum values at an optimum cross-link density.
Therefore, the compounder must keep his end-use requirements in
mind in order to optimize the ’’key property” and make the best trade­
offs of the other properties as well.

B. K in etic Model o f C u re as an A id to
P ro p e rty and Processing C ontrol

The cross-link density is controlled primarily by thermal history and


the chemistry of the sulfur/acceleration system employed. Because
410 Hsich and Ambrose

Fig. 21 Viscosity vs. filler loading at various temperatures for nat­


ural rubber. (From Ref. 22.)

of the profound effect of cure on the mechanical properties of elasto­


mers, it is important to take a close look at the kinetics of the curing
reaction.
Differential scanning calorimetry (DSC) has been used in the study
of cure kinetics of thermosets [67—71]. In these studies a quasi-theo-
retical model is proposed to describe the dependence of the cure rate
on the cure temperature and time by measuring the amount of heat
Rheological Behavior and Molding Technology 411

Fig. 22 Cross-link density vs. elastomer properties.

evolved during cure under isothermal conditions. The kinetic model


of cure is [71]

^ 0^_ /-T T , TT ^\/1


= ( K i + K 2a ) ( 1 - a ) (28)

where Ki and K 2 are rate constants and m and n are constants inde­
pendent of temperature. The relative degree of cure a is defined by
Q
R
(29)
m
412 Hsìch and Ambrose

where Qj^ is the amount of heat evolved isothermally as a result of


the cure reaction from time t = 0 to t and Q ^, which is temperature
dependent for some materials, is the total amount of heat generated
or the isothermal heat of cure. Kamal et al. [71] were able to use
the above model satisfactorily for describing data from DSC experi­
ments for both epoxy and unsaturated polyester systems. However,
the degree of cure, given in Eq. (29), does not correlate directly
with the physical or mechanical properties of the polymer at speci­
fic points during the isothermal cure cycle. In addition, cure ki­
netics of polymeric materials not only depend on the type of cura­
tive but also on filler loading. Since the physical and mechanical
properties of thermosets depend on the thermal history of cure, it
is of practical interest to have a kinetic model which can predict
both physical and mechanical properties at various times during the
cure cycle.
Craig [72], Mussatti and Macosko [73], and Roller [74] have de­
veloped a kinetic model for predicting viscosity and modulus during
the cure reaction. The model does not include polymer-filler inter­
actions and is limited to first-order (single relaxation function) chem­
ical reactions.
Hsich et al. have developed a general relaxation function for ex­
plaining cure kinetics in which polymer-filler interactions are taken
into account [1,2,10,11]. The following equation expresses the re ­
lationship of a particular property to cure time during any cure re­
action :

P (t)
= exp (30)
i m
and

To exp (31)
( i)
where 3 is a constant describing the width of the relaxation spectrum.
To is constant, R is the gas constant, T is the absolute cure tempera­
ture, and H is the activation energy of the cure reaction. P is a par­
ticular property of the elastomer such as modulus or viscosity. The
subscripts ~ and 0 denote maximum and minimum values of the prop­
erty during the cure cycle.
Remember that properties such as modulus and viscosity decrease
to minimum values at the beginning of the cure cycle before they start
to increase because the modulus or viscosity of the sample at room tem­
perature is higher than at the cure temperature before cross-linking
Rheological Behavior and Molding Technology 413

begins. Furthermore, it also may require a retardation time before


cross-linking onset. Therefore, in fitting cure curves the induction
time to (which is the time for the torque to reach a minimum value)
must be subtracted from t in Eq. (30). Then Eq. (31) is modified
to

P (t) =Po + (P - P o ) U - exp (32)

If the cure model is extended to include nonisothermal cure kinetics,


the modification can be expressed as

P q exp
= '■e " [ ''•

1 - exp (33)

and

T (t) dt + T(0) (34)


dt

where Pe represents the value of the property at temperatures ap­
proaching infinity, E is the activation energy of viscous flow, Tc is
the equilibrium or final cure temperature, and tf is the final cure
time. Under isothermal cure conditions, t is a function of tempera­
ture only. However, under nonisothermal conditions t is a function
of temperature, which in turn is a function df time, as shown in Eq-
(33). In Eq. (33), Pe exp[E/RT] describes the effect of tempera­
ture on the property of concern in the absence of cure. The rest of
Eq. (33) describes the effect of the curing reaction itself on the
property.
Hsich et al. [1,2,10,11] tested these models by measuring the mod­
ulus and viscosity curves of filled and unfilled natural rubber com­
pounds at a frequency of 10 rad/sec. These curves measured under
isothermal conditions are shown in Figs. 23 and 24 and under non­
isothermal conditions in Figs. 25 and 26. The parameters used to
make these theoretical calculations shown in Figs. 23—26 are listed
in Table 3. It can be seen that both models [Eqs. (32) and (33)]
predict the modulus and viscosity values of filled and unfilled elas­
tomers quite well.
Rheological Behavior and Molding Technology 415

Fig. 24 Shear storage modulus vs. cure time for natural rubber with
40 phr filler loading. (From Ref. 1.)

Values for the relaxation time tq and the width of the relaxation
spectrum , 3, will depend on the particular property measured as well
as on the filler loading. However, values for the activation energy of
the cure reaction, H, are relatively insensitive to filler loading and
the particular property measured. Plots of tq and 3 for shear mod­
ulus and viscosity of natural rubber are shown in Figs. 27 and 28,
respectively. The relaxation time tq decreases as filler loading in­
creases. In addition, 3 decreases with increasing filler loading. This
is equivalent to broadening the relaxation spectrum with increasing
filler loading. This broadening of the relaxation spectrum was also
observed in a previous study on viscosity—elastomer-filler interac­
tions [10] and dynamic mechanical properties of cured rubbers [12,
13].
From the nonisothermal cure studies the activation energy E of
viscous flow decreased with increasing filler loading. This result
416 Hsich and Ambrose

Fig. 25 Shear storage modulus vs. cure time for natural rubber un­
der nonisothermal cure conditions. (From Ref. 1.)

is not surprising because the presence of filler tends to reduce the


intermolecular forces between polymer chains; therefore the flow be­
havior and mechanical properties of the polymer system become less
temperature dependent than in the absence of filler.
In a similar study [2] on the cure of filled and unfilled silicone
elastomers, Eqs. (32) and (33) were used to predict the modulus and
viscosity of these compounded stocks. In this case the effect of filler
on the shape of the relaxation spectrum was markedly different from
that found for natural ru b b e r. This illustrates a different degree of
polymer-filler interaction in silicone rubber compared to the natural
rubber system. It was also noted in this silicone cure study [2] that
for filled elastomers, the cure behavior proceeded as a simple first-
order chemical reaction. This was not the case for unfilled silicone
ru bber.
Rheological Behavior and Molding Technology 417

U)
O

Fig, 26 Complex viscosity vs. cure time for natural rubber under
nonisothermal cure conditions. (From Ref. 1.)

Table 2 Polymer-Filler Interaction


Constants

Constant Value

a 11.700
b 2.910
c (°K “^) 0.248
d 0.830
4
mT

00

Table 3 Parameters for Natural Rubber

Filler Cure 3 To (X 10" ^ min)


loading tempera­ H E to Modulus Viscosity Modulus Viscosity
(p h r) ture (° C ) (kcal/mol) (kcal/mol) (min) data (G ') data (n *) data (G ') data (n *)

10 130 18 5.60 15.8 4.09 3.26 3.56 3.18


150 6.4

20 130 18 13.9 3.04 3.16


150 6.2

30 130 18 4.09 12.5 2.89 2.51 3.05 2.73


150 5.9

40 130 18 12.3 2.45 3.00


150 5.5
X
0)
50 130 18 2.94 11.7 2.40 1.74 2.95 2.36 o’
3“

150 5.4 0)
3
a
>
3
cr
o
V)
a>
Rheological Behavior and Molding Technology 419

Fig. 27 To vs. filler loading for natural rubber. (From Ref. 1.)

In conclusion, the kinetic model discussed above can be used to de­


scribe the process of cross-linking. It can predict physical and mech­
anical properties at various times during the cure cycle. The model
can also be used to predict scorch times (premature c u re ). This
latter parameter is vital in polymer processes such as mixing and
molding.
420 Hsich and Ambrose

Fig. 28 3 vs. filler loading for natural rubber. (From Ref. 1.)
Rheological Behavior and Molding Technology 421

V. M O LDING TEC H NO LO G Y OF ELASTOMERS

Conventionally, the molding of elastomeric materials is carried out by


transfer or compression processes which involve extensive hand labor.
Because of this, they are relatively slow operations. Injection mold­
ing of elastomers offers automation and speed but is not yet widely
used because injection molding machines are expensive and the tech­
nology of injection molding is less well developed than that of trans­
fer molding or compression molding.
Regardless of which molding operation is chosen, the ability to un­
derstand and predict both the flow rheology and chemorheology of
elastomers is the key to successfully molding these materials. Chemo­
rheology provides a means of predicting cure behavior. Such pre­
dictions allow one to overcome scorch problems and optimize cure con­
ditions to obtain desired end-product properties. Flow rheology per­
mits optimization of variables such as molding pressure and tempera­
ture and aids in the actual design of runners, sprues, and gates.

A. Elastomers as E n g in eerin g M aterials

The versatility and adaptability of elastomers have been responsible


for their use in engineering applications for decades. Unlike metal,
elastomeric materials possess inherent damping, which is particularly
beneficial when resonant vibrations are encountered. In addition,
elastomers can store more elastic energy than steel and can be chem­
ically bonded to metal, which provides a means for location and fix­
ing, in elastomer-metal composites.
A major application which employs the unique features of rubber
elasticity is the spring. Uses of elastomeric materials in engineering
applications have been discussed by Snowdon [8 ], Bindley [75], and
Freakley and Payne [76]. These include suspension systems, flex­
ible couplings, bridge bearings, piers, fenders, noise controllers,
structural dampers, and, extensively antivibration mountings.
Antivibration mounting systems are required to provide small val­
ues of transmissibility at aU frequencies contained in the Fourier spec­
trum of the force applied to the item of equipment or machinery that
they support. An effective antivibration mounting system should pro­
vide a low natural mounting frequency, wq, a low transmissibility at
resonance, and a transmissibility that decreases rapidly with fre ­
quencies greater than wq*
The natural frequency and transmissibility of a simple mounting
system are [8,76]
422 Hsich and Ambrose

0)0 (35)
m ' ‘

and
1/2
1 + tan^ 6
(36)
( l - (oa^/w^^KGQ/G^)) ^ + tan^ 6o

where K is a shape factor, M is the mass of the system, GJ and Gj) are
the shear storage moduli of the elastomer at natural frequency O3o and
forced frequency oa, respectively, tan 6^j is a measure of elastomer
damping at the forced frequency, and T is the transmissibhity of the
elastomeric mounting system.
As was mentioned earlier, both dynamic modulus and damping are
functions of temperature and frequency. These mechanical properties
can be manipulated by tailoring the molecular structures of elastomers
through compounding and cure control. To that end, one can achieve
a low natural-frequency vibration system by reducing the shear stor­
age modulus of the elastomer in question [see Eq. (3 5 )]. In addition,
one can suppress the resonant transmissibility by choosing elastomers
with high damping factors [see Eq. (3 6 )].
These examples serve to illustrate how effective engineering ma­
terial design requires the integration of chemistry, physics, mechan­
ics, and engineering. Further integration of design with materials
and processes is necessary for sophisticated materials management.
This is now possible by analytical modeling techniques and modern
computers. Both computer-aided design (C AD ) and computer-aided
manufacturing (CAM) are emerging from the ability to develop mathe­
matical models which are deduced from inseparable multidisciplinary
fields of science and engineering. A scheme for completing the tri­
angle of material, processing, and design is shown in Fig. 29. The
highest end-product performance to cost ratio can be achieved by
using such a scheme.

B. Elastomer Molding

The advent of the microprocessor has made it possible to control man­


ufacturing processes under predetermined conditions automatically.
This has helped injection molding to become an important method for
producing plastic parts. Injection molding has not achieved the same
degree of popularity and importance in the rubber industry. This
might be because the rheological characteristics of elastomers are far
more complicated than those of thermoplastics. Perhaps this explains
Rheological Behavior and Molding Technology 423

Fig. 29 A scheme for completing the triangle of material, design, and


processing.

the limited research in the field of elastomer rheology. Some studies


have been undertaken on the injection molding of elastomers in the
last decade. Wheelans [77—79] has discussed the modification of cure
systems to improve the processibility of elastomers during the injec­
tion-molding operation. More recently, Byam et al. [80,81] have de­
veloped a practical computer program for the simulation of elastomer
injection molding. In their program they calculated pressure drop,
viscous heating, degree of scorch, and degree of cure during the
molding operation. The key to simulating any elastomer molding pro­
cess accurately is to successfully model the changes in the flow rheol­
ogy and the chemorheology of the elastomer during the molding oper­
ation. During molding, elastomer melts experience change in shear
forces and shear rates at different locations and times. The thermal
424 Hsich and Ambrose

history of these elastomer melts needs to be accumulated. This is


true because the cure state of the elastomer continues to change, as
does the viscosity and modulus throughout the entire molding pro­
cess. A nonisothermal cure kinetic model can be used to calculate
the state of cure, viscosity, and modulus of the elastomer during the
molding operation. Figure 30 shows a scheme for elastomer molding
in which the rheological properties of the elastomer are first charac­
terized, followed by modeling the elastomer viscoelastic flow, cure ki­
netics, and thermal properties. These models are required in the
fluid and heat transport equations. Once these fluid and heat trans­
port equations have been properly set up , one can use a computer to
simulate the molding operation. From the geometric configuration of
the end-product part, a mold can be designed by doing fluid and heat
flow analysis during the molding simiilation. Once an optimum design
is achieved, this design can be used to obtain injection-molding con­
ditions. Elastomeric properties can be predicted and cure conditions
optimized from proper use of the kinetic model.
Two commercial computer programs have been widely used for mold
flow analysis. MOLDFLOW was developed by Colin Austin of Australia.
The second was developed at Cornell University as part of the Cornell
Injection Molding Program (CIM P). Both programs were developed
primarily for molding thermoplastics. The program developed by
CIMP discloses the details of the program derivations. This gives
users great benefit in evaluating and modifying the program to suit
their own applications. Some details of the program developed by
CIMP are given in Chapters 1 and 2.

C. T h e Use o f Elastomer Rheological P ro p e rtie s in


Molding Laminated Elastomer Composites

Antivibration mountings and other bonded mechanical rubber parts


have complicated geometric shapes. These parts often contain sev­
eral elastomeric layers separated by metal plates. If one is to use
transfer or injection-molding processes to make such parts, a mold
design that provides balanced pressure between each elastomer layer
is necessary to avoid displacement or distortion of the metal plates.
Different elastomers will display different rheological behavior.
For example, one would expect the viscosity, elasticity, and die swell
properties of natural rubber to differ markedly from those of a sili­
cone rubber counterpart. If this is true, a mold designed to balance
the pressure between natural rubber layers would not be suitable for
balancing the pressure between silicone rubber layers.
To test this hypothesis, a simple four-cavity mold was designed
and balanced employing the rheological properties of a natural rubber
compound 1. The mold, illustrated in Fig. 31, had two identical long
runners and two identical short runners. This allowed a cross-check
Rheological Behavior and Molding Technology 425

Fig. 30 A scheme for elastomer molding.

on tooling errors within two pairs of identical runners. The mold con­
sisted of three plates having a sprue and the four semicircular runners
in the top plate, four gates in the center plate, and four identical cir­
cular cavities in the bottom plate.
Rheological Behavior and Molding Technology 427

A series of short-shot experiments were run at various injection


times and the weights of the natural rubber compound 1 in each mold
cavity were recorded at each time. The weights of rubber in each of
the four cavities were almost identical at a given injection time. When
these experiments were repeated, however, with a different natural
rubber compound (designated 2) or a silicone rubber compound, the
weights of rubber in each of the four cavities were different at a
given injection time. This indicates that a pressure imbalance existed
in the runner system when the natural rubber compound 2 and the
silicone rubber compound were employed. The results are summarized
in Table 4. Cavities 1 and 3 are joined to the long runners, and cavi­
ties 2 and 4 connect to the short runners. These results illustrate
quite vividly the importance of considering the rheological properties
of the elastomer when designing a mold.
The molding of parts containing several thin elastomeric layers sep­
arated by metal plates may first require hand laying-up the elastomer
stocks in each separate layer of the mold cavity. This operation is
followed by transferring an elastomer fill stock to each layer. Metal
plate distortion, elastomer knit lines, and uneven cure are common

Table 4 Weight of Rubber in Cavities (g )

Injection
Cavity
time
Rubber stock 1 2 3 4 (sec)

Natural rubber 0.96 1.37 1.01 1.41 12


(compound 1) 5.24 5.97 5.38 5.89 18
10.36 10.51 10.53 10.44 24
15.35 15.22 15.23 15.22 30
29.50* 29.64* 29.59* 29.63* 55

Natural rubber 2.38 2.95 2.76 2.94 12


(compound 2) 11.66 13.16 11.37 13.82 24
16.86 17.71 16.01 19.40 30
21.12 27.08 20.68 26.95 36

Silicone rubber 7.84 7.75 7.82 8.33 12


16.56 20.22 15.67 17.85 18
24.12 29.71 22.46 30.57 24
29.51 33.98* 29.96 33.96* 30

*Complete fill.
428 Hsich and Ambrose

Fig. 32 Mold for five-layer laminated composite.

problems associated with this process. Overcoming this may require


tedious trial-and-error mold development. By using the rheological
properties of elastomers along with mold flow analysis, one can design
a mold which permits the injection or transfer molding of laminated
composites without conventional hand lay-up procedures. Figure 32
illustrates a mold used to make a five-layer rubber-metal composite
Rheological Behavior and Molding Technology 429

(rod end used in helicopters). The mold has a balanced runner sys­
tem that permits the injection molding of five different rubber stocks
simultaneously into five separate layers of the mold without any dis­
tortion of the metal shims. In Fig. 33 the finished part (far right)
is shown along with the separated metal parts.
The above mold designs were achieved by taking into account
the changes in flow rheology from one elastomer to another. Such
considerations make possible the molding of parts having complex
U30 Hsich and Ambrose

F ig . 33 Finished laminated elastomer composite and individual metal


parts.

geometric shapes. Without such considerations, molding these parts


would be extremely difficult and perhaps impossible.

D. Summary

The molding of thermosetting elastomers is a complicated process. Not


only is it necessary to understand the flow behavior of these elasto­
mers at different temperatures and filler loadings, but one must also
be aware of the chemorheology, since this impacts on the flow behav­
ior. The elastomer rheology is highly sensitive to the elastomer heat
history and cure kinetics. For this reason models are necessary to
describe the changes in flow behavior at various temperatures and
times if one wishes to use rheological characteristics to aid in the
molding process.
We have attempted to use the rheological behavior of thermosetting
elastomers to help us design molds. We have concentrated our efforts
on designing molds with multiple cavities that fill simultaneously. In
so doing, we have considered the theory of the rubbery state, the
viscosity theory of polymers, and the chemorheology of elastomers to
create models which predict both chemorheological and mechanical b e ­
havior of elastomers. It is through such models that we have designed
runner and gate systems for a number of laminated elastomer compos­
ites.
Rheological Behavior and Molding Technology 431

REFERENCES

1. H. S .-Y . Hsich, R. M. Zurn, and R. J. Ambrose, in Chemo-


rlneology of Thermosetting Polymers, C. A. May, Ed., ACS
Symposium Series, No. 227, Chapter 16, 1983.
2. H. S .-Y . Hsich, L, C. Yanyo and R. J. Ambrose, J. Appl.
Polym. Sci., 29: 2331 (1984).
3. H .S .-Y . Hsich, J. Mater. Sci., 15: 1194 (1980).
4. R. O. Davies and G. O. Jones, Adv. P h ys., 2: 370(1953).
5. R. O. Davies and G. O. Jones, Proc. Roy. Soc. ,217A: 26
(1953).
6. H. S .-Y . Hsich, J. Mater. Sci., 13: 750 (1978).
7. H. S .-Y . Hsich, J. Mater. Sci., 14: 2581 (1979).
8. J. C. Snowdon, Vibration and Shock in Damped Systems, John
Wiley, New York, 1968, Chap. 2.
9. F. N. Kelley and M. L. Williams, Rubber Chem. Technol., 42:
1975 (1969).
10. H. S .-Y . Hsich, J. Appl. Polym. Sci., 27: 3265 (1982).
11. H. S .-Y . Hsich and R. J. Ambrose, Organic Coatings & Appl.
Polym. Sci. Proc. , 47: 649 (1982).
12. H. S .-Y . Hsich, L, C. Yanyo, and R. J. Ambrose, J. Appl.
Polym. Sci. , 29: 413 (1984).
13. H. S .-Y . Hsich, M. J. Hronas, H. E. Hill, and R. J. Ambrose,
J. Mater. Sci., in press.
14. P. J. Flory, Principles of Polymer Chemistry, Cornell University
Press, Ithaca, 1971, Chap. XI.
15. I. Prigogine and R. Defay, Chemical Thermodynamics, trans­
cribed by D. H. Everett, John Wiley, New York, 1972, pp.
174-310.
16. D. Turnbull, in Solid State Physics, Vol. 3, F. Seitz and D.
Turnbull, Eds., Academic Press, New York, 1956, pp. 225—306.
17. L. R. G. Treloar, The Physics of Rubber Elasticity, Oxford Uni­
versity Press, 1958, Chap. IV.
18. P. J. Flory, Chem. Revs., 35: 51 (1944).
19. G. E. Roberts and E. F. T. White, in The Physics of Glassy
Polymers, R. N. Haward, Ed., John Wiley, New York, 1963,
Chap. 3.
20. M. L. Williams, R. F. Landel, and J. D. Ferry, J. Amer. Chem.
Soc., 77: 3701 (1955).
21. A. K. Doolittle, J. Appl. Phys., 22: 1471 (1951).
22. H. S. -Y . Hsich, J. Mater. Sci., 17: 438 (1982).
23. T. S. Chow, J. Mater. Sci., 15: 1873 (1980).
24. A. Einstein, Ann. Phys., 19: 289 (1906).
25. E. G ut h a n d O. Gold, Phys. R ev., 53: 322 (1938).
26. L. H. Cohan, India Rubber World, 117: 343 (1947).
432 Hsich and Ambrose

27. J. L. White, in Science and Technology of Rubber, F. R. Eirich,


Ed., Academic Press, New York, 1978, Chap. 6.
28. J. L. White, Rubber Chem. TechnoL, 42: 257 (1969).
29. J. L. White, Proc. Int. Rheol, Congr,, 5th, 3: 17 (1970).
30. N. Tokita and J. L. White, J. AppL Polym, Sci., 10: 1011 (1966).
31. W. R. Bolen and R. E. Colwell, SPE Tech Papers, 4: 1004
(1958).
32. G. R. Vüa, Ind. Eng. Chem., 36: 113 (1944).
33. H. Zamodits and J. R. A . Pearson, Trans. Soc. Rheol., 13: 357
(1969).
34. T. G. Fox, S. Gratch, and S. Lashaek, in Rheology, Vol. I,
F. R. Eirich, Ed., Academic Press, New York, 1956, Chap. 12.
35. H. Eyring^ J. Chem. Phys., 4: 283 (1936).
36. T. G. Fox and P. J. Flory, J. Am. Chem. Soc., 70: 2384 (1948).
37. T. G. Fox and P. J. Flory, J. Appl. Phys., 21: 581 (1950).
38. T. G. Fox and P. J. Flory, J. Phys. Chem., 55: 221 (1951).
39. T. G. Fox and P. J. Flory, J. Polym. Sci., 14: 315 (1954).
40. G. S. Fulcher, J. Am. Ceram. Soc., 8: 339 (1925).
41. R. H. Ewell, J. Appl. Phys., 9: 252 (1938).
42. M. H. Litt, Trans. Soc. Rheol., 20: 47 (1976).
43. F. Bueche, J. Chem. Phys., 20: 1959 (1952).
44. G. C. Berry and T. G. Fox, Adv. Polymer Sci., 5: 261 (1968).
45. T, A . Litovitz and C. M. Davis, in Physical Acoustics, Vol. IIA,
W. P. Mason, Ed. , Academic Press, New York, 1965,Chap. 5.
46. P. A . Egelstaff, An Introduction to the Liquid State, Academic
Press, London, 1967, pp. 1--161.
47. W. Philippoff, in Physical Acoustics, Vol. I I B, W. P. Mason,
Ed., Academic Press, New York, 1965, Chap. 7.
48. J. D. Ferry, Viscoelastic Properties of Polymers, John Wiley,
New York, 1970.
49. F. Bueche, Physical Properties of Polymers, Wiley-Interscience.
New York, 1962, pp. 168-221.
50. B. D. Coleman and H. Markovitz, J. Appl. Phys., 35: 1 (1964).
51. W. M. Kulicke and R. S. Porter, Rheol. Acta, 19: 601 (1980).
52. W. P. Cox and E. H. Merz, J. Polym. Sci., 28: 619 (1958).
53. K. Walters, Rheomeiry, John Wüey, New York, 1975, Chap. 5.
54. C. D. Han, Rheology in Polymer Processing, Academic Press,
New York, 1976, Chaps. 5, 6.
55. H. S. -Y . Hsich, R. S. Engel, and R. J. Ambrose, to be sub­
mitted to Rub. Chem. & Tech.
56. V. L. Folt, R. W. Smith, and C. E. Wilkes, Rubber Chem. Tech-
nol., 44: 1 (1971).
57. D. I. Lee, Trans. Soc. Rheol., 13: 273 (1969).
Rheological Behavior and Molding Technology 433

58. R. F. Landel, B. G. Moser, and A. J. Bauman, in Proceedings


of the Fourth International Congress of Rheology, Providence,
R I., Part 2, E. H. Lee, E d., Wiley-Interscience, New York,
1965.
59. I. Pliskin and N. Tokita, J. AppL Polymer Sci,, 16: 473 (1972).
60. R. W. Wise, in Rubber Technology, M. Morton, Ed., Van Nos­
trand Reinhold, New York, 1973, Chap. 4.
61. S. H. Morrell, in Rubber Technology and Manufacture, C. M.
Blow, Ed., Novnes-Butterworths, London, 1977, Chap. 5.
62. A . Y. Coran, in Science and Technology of Rubber, F. R.
Eirich, E d ., Academic Press, New York, 1978, Chap. 7.
63. R. H. Campbell and R. W. Wise, Rubber Chem, Technol., 37:
635(1964).
64. R. H. Campbell and R. W. Wise, Rubber Chem. Technol., 37:
650(1964).
65. M. M. Coleman, J. R. Skelton, and J. L. Koenig, Rubber Chem.
Technol., 46: 938 (1973).
66. W. J. Bobear, in Rubber Technology, M. Morton, E d., Van
Nostrand Reinhold, New York, 1973, Chap. 15.
67. R. A. Fara, Polymer, 9: 137 (1968).
68. M. A . Acitelli, R. B. Prime, and E. Sacher, Polymer, 12: 335
(1971).
69. R. B. Prime, Polym. Eng. Set., 13: 365 (1973).
70. E. Sacher, Polymer, 14: 91 (1973).
71. M. Kamal, R. S. Sourour, and M. Ryan, SPE Tech Papers, 19:
187(1973).
72. D. Craig, SPE Tech Papers, 18: 533 (1972).
73. F. G. Mussatti and C. S. Macosko, Polym. Eng. Sci., 13: 236
(1973).
74. M. B. Roller, Polym. Eng. Sci., 15: 406 (1975).
75. P. B. Lindley, Engineering Design with Natural Rubber, NR
Technical Bulletin, MRPRA, London, 1967.
76. P. K. Freakley and A. R. Payne, Theory and Practice of Engi­
neering with Rubber, Applied Science, London, 1978, pp.
319-637.
77. M. A. Wheelans, Prog. Rubber Technol., 32: 57 (1968).
78. M. A. Wheelans, NR Technol., 7: 86 (1976).
79. M. A. Wheelans, Rubber Chem. Technol., 51: 1023 (1979).
80. J. D. Byam and G. P. Colbert, presented at the 111th Meeting
of the Rubber Division, ACS, Chicago, 111., May 1977.
81. J. D. Byam, G. P. Colbert, and K. D. Ziegel, presented at the
Golden Jubilee Meeting, Society of Rheology, Boston, M ass.,
October 1979.
Injection Molding of Rubber
Compounds

Avraam I . Isayev
Polymer Engineering Center
The University of Akron
Akron, Ohio

I. IN T R O D U C T IO N

Since the Hyatt brothers received a patent for a ram injection-mol ding
machine in 1872, injection molding of polymers has been an important
processing operation [1 ]. A variety of complex and sophisticated
parts are injection molded in the modern plastics industry from a wide
range of thermoplastic materials. At the same time the rubber indus­
try has traditionally used compression and batch blow molding to fab­
ricate commercial products [2 ]. These processes require hand opera­
tion and are highly time consuming and difficult to automate. The in­
jection molding of rubber by a ram injection machine emerged 40 or
more years ago, with the first user apparently being Chrysler [3 ].
Screw injection molding of rubber was first reported in the 1950s.
Ram and screw injection-molding machines have been actively mar­
keted since the 1960s. The most important feature of injection mold­
ing is that it allows introduction of a high degree of automation in the
rubber industry. In addition, rubber compound preparation and cure
time can be reduced, and trimming of molded product can be practi­
cally eliminated.
At the present time injection molding of rubber is rapidly growing.
There is a gradual shift from compression to injection molding of ru b ­
ber products. The automotive, machinery, shoe, and pharmaceutical
industries are typical examples of rubber injection-molding business.
Nowadays, rubber injection molding is entering into the rubber man­
ufacturing business.

U35
436 Isayev

Although there is extensive fundamental research in injection mold­


ing of thermoplastics, there is very little dealing with elastomers. In
particular, different scientific teams are presently working on various
aspects of injection molding of thermoplastics and liquid and thermoset
molding. The rubber injection-molding industry has been neglected
and relies essentially upon experience, lacking a scientific basis for
the technology. Of course, technology developed in injection molding
of other materials can be partially extended to rubber molding. How­
ever, there are fundamental differences in the molding of thermoplas­
tics and thermosets on the one hand and rubber on the other. Plas­
tics are injected into a cold mold where they solidify as they flow into
the mold, whereas rubber is injected into a hot mold where it is sub­
sequently vulcanized. In reaction injection molding, low-viscosity re ­
active liquids or a mixture of reactant fluids (usually Newtonian) are
injected into a hot mold where they are polymerized to solidification.
The inherent complexity of various rubber compounds and their
peculiar flow behavior during molding requires special attention. An
integral approach to injection molding of rubber is needed which will
take into account all major aspects of rubber behavior, including mod­
eling of flow in the extruder, nozzle, sprue, runner system, and cav­
ity, curing behavior, packing, mold design, rheology, and effect of
processing upon structure and properties of molded products. A de­
scription of this integrated concept is the major thrust of this chap­
ter.

II. PROGRESS IN IN JE C T IO N MOLDIN G


OF R U B BE R

A. In tro d u c to ry Remarks

During the last decade significant progress has been made in the area
of the injection molding of thermoplastics. The development of the in­
jection molding of rubber, despite many advantages, has been slow.
The main reason for this is premature scorch, incomplete mold filling,
and other failures due to the absence of science-based technology.
Presently, the injection molding of rubber is a widely employed pro­
cess due to the tremendous developments in rubber compounds, ma­
chine design, and process control. Since the early 1970s injection
molding of rubber in many instances has gradually replaced compres­
sion molding. According to Wheelans [3 ,4 ], any rubber product can
be injection molded. However, a decision has to be made on the basis
of expected capital investment, the number of runs, and a detañed
analysis of expenses involved in changing from the existing molding
technology to a new one. Faced with the complexity of the injection
Injection Molding of Rubber Compounds 437

molding process of rubber, many manufacturers have decided to stay


with the known cumbersome technology. In particular, available in­
jection-molding processes of various rubber products were developed
by trial and error.
Injection molding of rubber has been defined as the automatic feed­
ing of heated and plasticated rubber stock into an injection chamber
at a temperature below the vulcanization temperature and injection
through a nozzle, sprue, runner, and gate into a cavity having a wall
temperature high enough to initiate vulcanization and subsequently
vulcanize the rubber inside the cavity and eject the molded part.
Clearly , the nature of injection molding requires a contribution from
various disciplines, such as machiue design, polymer rheology and
fluid mechanics, heat transfer, reaction kinetics and structure-prop­
erty relationships of polymers. In general, these disciplines are in­
herently linked in the process and have to be considered simultaneously
in an integrated approach. The literature contains many instances in
which some of these disciplines have separately been considered with
regard to applications to injection molding.

B. Injection Machines

The injection-molding operation is usually carried out by a single ma­


chine. There are three major types of injection machines in the ru b ­
ber industry (Fig. 1): the ram type, the in-line reciprocating-screw
type, and the out-line nonreciprocating-screw type [4—7]. At the
present time ram-type machines are not widely used, due to difficul­
ties of achieving homogeneity and uniform temperature distribution.
The in-line screw machines are commonly used because of uniformity
of feeding, mixing, and heating by screw motion. The out-line screw
machines have a separate screw and injection chamber and can develop
the high pressure needed to fill large-volume molds.
Screw extruders for rubber were introduced over 100 years ago
[8 ]. However, applications of the screw extruders in rubber injec­
tion molding only date to the last few decades. There have been ma­
jor developments in the understanding of the flow of rubber in an ex­
truder during molding. The flow of rubber in the barrel is a com­
bination of drag transverse and pressure flow and leakage. Gener­
ally speaking, the flow of rubber in the screw extruder should be
easier to analyze than the flow of thermoplastics. This is mainly due
to the absence of a melting zone. Despite this evident simplicity, most
theoretical analyses of rubber flow in extruders are based upon New­
tonian behavior and can be considered as qualitative [8—10]. In prac­
tice, rubber compounds are highly non-Newtonian fluids with shear-
rate and temperature-dependent viscosity. Depending on the feeding
438 Isayev

(a ) Nozzle
Feed
Throat Mold
Ram
Injection
B arrel Chamber

B arrel
Heating
Fluid

(b)
Nozzle

Fig, 1 Various types of injection-molding machines for rubber com­


pounds. (a ) ram machine; (b ) screw machine; (c ) a combination of
ram and screw machine.

arrangement, rubber extruders can be divided into hot- or cold-feed


extruders [7,8,11,12]. Hot-feed extruders are usually fed with p re ­
warmed milled strips, and cold-feed extruders are fed with pellets or
with strips at ambient temperatures. Presently, cold-feed extruders
are more extensively used due to their lower capital investment and
production costs and short heat history [7 ]. As reviewed by Johnson
[7] , several screw designs have been proposed in order to introduce
complex flow patterns needed to achieve high dispersion levels and
thermal homogeneity of the rubber compound component. Extensive
literature exists for the rational design of screw extruders for pro­
cessing of thermoplastics where non-Newtonian flow of molten poly­
mers has been incorporated [10,13—20]. This literature has to be
Injection Molding of Rubber Compounds 439

utilized and applied to rubber injection molding. The main difference


of rubber extruders from extruders for thermoplastics is the absence
of a pellet/powder conveying and melting region. In rubber extrusion
there is only metering, though part of the metering region near the
hopper resembles melting solid conveying regions. This significantly
simplifies the situation with extruder design. However, the rheologi­
cal properties of rubber compounds introduce additional complications
into the screw design.

C. Rheology of R u b b e r Compounds

In general, the rheology of rubber compounds is very complex and


quite different from thermoplastics. Molten thermoplastics have been
extensively studied in shear and elongation over a wide range of de­
formation rates [21,22]. In many cases they are viscoelastic non-New­
tonian fluids obeying the usual no-slip boundary conditions. The
theory of viscoelastic non-Newtonian fluids has been extensively em­
ployed to analyze a wide spectrum of polymer processing operations
[23—32]. As far as rubber compounds are concerned, presently it
has been well established that they are a non-Newtonian media with
inherent yield stress and high modulus [33-37]. In general, rubber
compounds are highly filled with carbon black. They are highly thix­
otropic. Due to the presence of yield stress, rubber compounds
440 Isayev

F ig . 2 Viscosity “ Shear stress behavior of the carbon black—filled iso­


prene rubber (Cariflex 305/Shell) at various volume loading of carbon
black (N326-HAF) and temperature 100°C. (From Ref. 37.)

respond as solid bodies or gels at stresses lower than yield stress and
like viscoelastic fluids above it. Figure 2 shows a typical rheological
behavior of rubber compounds. Some highly filled particulate thermo­
plastics also behave similarly [38—42]. There have been some efforts
at developing a quantitative theory of this behavior with comparison
to experiment [43—45]. In particular, initial efforts have been made
Injection Molding of Rubber Compounds 441

to apply the White theory [43] to describe the rheological behavior of


rubber compounds. As an example, Fig. 3 shows a comparison be­
tween the experimental and theoretical data for carbon-black-füled
polystyrene. A comparison is given for a narrow range of strain
rates and stresses, and an agreement of the theory with experiment
is found to be qualitatively correct. In addition, rubber compounds
and some pure elastomers exhibit slippage at the solid boundary which
highly complicates the flow situation during processing [36,46—49].
Typically , the slip effect manifests itself in an experiment involving
flow in capillaries of different diameters. In particular, pressure
drop—flow rate curves obtained in capillaries with a small diameter
are shifted toward higher flow rate and lower pressure drop com­
pared with the curves obtained in capillaries with a large diameter
(Fig. 4). This result indicates that one needs to determine bound­
ary conditions to analyze the processing of rubber, in general, and
the injection molding process, in particular.
Little has been reported on the flow behavior of rubber compounds
at injection-molding conditions [7 ]. To do so, one should perform
rheological tests in the range of shear rates comparable to molding.
Many rubber molders rely on the Mooney viscosity test to ascertain
the processibüity of rubber compounds. However, the Mooney vis­
cosity is not always reliable as a processibüity indicator since the
viscosity is measured at one particular shear rate. Certainly, vis­
cosity determined at one shear rate cannot characterize the viscous
behavior of rubber [7,50]. In particular. Fig. 5 schematically dem­
onstrates the case of two rubber compounds having the same Mooney
viscosity and exhibiting remarkably different rheological behavior.
This situation with Mooney viscosity test of rubber compounds is
analogous to the melt index test of molten thermoplastics which also
indicates the viscosity at only one particular shear rate.
Presently, in studies of thermoplastics injection molding, sophis­
ticated rheological techniques have been employed to characterize the
viscous behavior. In particular, rotational, capülary, and slit rhe-
ometry are employed to determine the flow curve at a wide range of
shear rates [21,42,51]. Even a capillary die attached to the screw
extruder of an injection-molding machine has been used to determine
the viscosity of thermoplastics in the range of shear rates corres­
ponding to injection molding [52]. Recently, researchers dealing
with rubber molding have also presented convincing evidence of the
use of capillary rheometry for predicting the flow behavior of rubber
compounds during injection molding [50,53]. In particular, the Instron
capülary rheometer has been employed. The recent development of
the Monsanto processibüity tester, specifically designed for handling
rubber compounds, may become a practical tool for measuring the vis­
cous properties of rubber compounds in various rubber processing
u
(U
C/1

fO
CL

Fig. 3 Steady-state shear viscosity (a ) and first normal-stress dif­


ference (b ) as a function of shear stress for carbon black-filled poly­
styrene (Styron 678U/Dow) at various volume loading of carbon black
(N542) and temperature 180®C. (From Ref. 45.) Symbols are experi­
mental data: lines are theoretical predictions.
Injection Molding of Rubber Compounds 443

L /R = CONST = 20

Fig. 4 Average effective shear rate of sodium-butadiene rubber vs.


shear stress for capillaries of different radii (o -0.505; a -0.95; □
-2.06 mm) and different temperatures. Dashed lines show flow curves
of rubber in bulk corrected for slippage. (Data from Ref. 48.)

operations including injection molding [7,54—56]. In addition to vis­


cosity, extrudate swell is an important characteristic of a rubber com­
poundes behavior. According to Oda et al. [57], the extrudate swell
gives a criteria for determining the allowable gate size to avoid unde­
sirable jetting during cavity filling.
444 Isayev

Fig. 5 Schematic representation o’f flow curves of two rubber com­


pounds having the same Mooney viscosity.

D. Scorching and Moldabllity

Many failures of rubber injection molding have occurred due to the


premature scorching of rubber compounds before the mold is com­
pletely filled [3 ]. Thus studies of vulcanization kinetics are of para­
mount interest in successful injection molding. For various thermo­
sets, curing kinetic studies include the method of calorimetry, such
as differential scanning calorimetry or differential thermal analysis,
infrared and other spectroscopic techniques, electrical conductivity,
and dielectric methods [58—73] . Although the state of cure measured
Injection Molding of Rubber Compounds 445

by these techniques is not directly related to the rheology of the cur­


ing system, measurements are extremely important and easy to per­
form beyond the gel point. However, from the viewpoint of deter­
mining the scorch index and time-dependency of rheological proper­
ties, they are very useful. Measurements of vulcanization character­
istics of rubber compounds have been performed by using a Mooney
viscometer, oscillating disk rheometer, and mechanical spectrometer
[7,50,53, 74,75]. Scorch time measurements have also been performed
by using a capillary rheometer at a shear rate close to or typical to
the injection-molding condition of the rubber compound [50]. In p ar­
ticular, any rubber compound has its own ’’operating window” which
defines a combination of processing parameters for a successful mold­
ing. A typical example of constructing such a window in terms of in­
jection temperature and time is given in Fig. 6. Evidently, the width

INJECTION TIME

Fig. 6 Schematic representation of operating window for moldability


of rubber compound.
nn6 Isayev

INJECTION MOLDING MACHINE

OIL HEATED
SLIT CHANNEL MOLD

PRESSURE DISPLACEMENT
THERMOCOUPLES
TRANSDUCERS TRANSDUCER

Fig. 7 Schematic representation of experimental setup to study mold-


ability of rubber compounds.

of this window depends on compound properties, machine size, mold


geometry, and temperature conditions. Moldability studies have also
been performed by filling a spiral mold which roughly indicates whether
a particular compound can be molded under given processing condi­
tions [6 ]. Other possible methods for studying moldability of rubber
compounds have been described [76] in which the test mold has been
installed to allow injection into an oil heated open-end slit channel.
It has been designed as a two-part mold clamped to an inject ion-mold­
ing machine (Fig. 7). Temperature and pressure in the slit channel
and injection pressure and screw feed have been measured. Similar
devices have been described elsewhere [52,116].

E. Modeling of the Injection-Molding Process


Modeling of the injection-mol ding process of rubber compounds in­
volves consideration of the unsteady, nonisothermal flow of a non-
Newtonian thixotropic media having a yield stress. This includes
simiilation of the flow in the screw extruder, nozzle, sprue, runner.
Injection Molding of Rubber Compounds 447

gate, and cavity until the cavity is completely filled. After filling is
completed, further consideration includes modeling of vulcanization
kinetics combined with packing under pressure in order to evaluate
the extent of cure and volume shrinkage. This is especially impor­
tant for determining the time suitable to demold and eject the part
having sufficient strength. So far the literature does not contain
instances at which all of the foregoing aspects of the modeling have
been described for rubber molding. The most notable progress in
this direction during the last decade has been made in the modeling
of injection molding of thermoplastics [77]. Success attained there
will eventually become a useful tool in applying it to molding of ru b­
ber compounds. However, even for thermoplastics molding, consid­
eration has been given for different stages of injection molding, with
each stage being considered separately. To the best of our knowl­
edge, all modeling attempts of thermoplastic injection molding disre­
gard the flow in the extruder, the starting point being the cavity or
runner. However, there exists a tremendous amount of literature
considering simulation of thermoplastic flow in screw extruders [13—
20,78]. In thermoplastic molding, simulation of flow in the screw may
not be as crucial as for rubber compounds, which should not be over­
heated in the extruder. This is imperative in order to avoid prema­
ture scorch.
In earlier modeling studies [79,80] for thermoplastics, attempts
have been made to simulate cavity filling based upon one-dimensional
flow of a Newtonian fluid. Later, it has been extended to cavity fill­
ing of non-Newtonian power-law fluids [81—101]. Examples consid­
ered include a rectangular-cavity [81—83,88,89,92—95] and a center­
gated or half-disk [84—87,90,91]. Cavity filling has been described
in terms of generalized Hele-Shaw flow. Solutions of the governing
equations of motion, continuity, and energy have been based upon
numerical methods in terms of finite-difference representation. Fur­
ther, as a step toward handling cavities of arbitrary geometry fre ­
quently encountered in injection molding, more accurate solutions
based upon two-dimensional flow models have been obtained [96—98].
Numerical methods with finite-difference or finite element approxima­
tions, or hybrids of both numerical schemes have been applied. These
methods have been employed for filling narrow gap cavities of any
planar shape with constant gap thickness or with variable gap thick­
ness. In many cases the results of the computations have been com­
pared with experiments. Computations usually give the advancement
of the melt front, position of the weldline, temporal development of
the gate pressure, required injection pressure and clamp force, gap-
wise velocity, and temperature distributions at any particular time
during cavity filling. This valuable information is especially impor­
tant for molding rubber compounds in order to proceed further with
the modeling of the vulcanization in the mold.
448 Isayev

In general, two-dimensional cavity-filling calculations are complex


and time consuming. Thus, in recent years modeling of injection mold­
ing has been evolved in which a system of the nozzle, sprue, runner,
gate, and cavity is decomposed into a simplified network of basic units
(primitives) of simple geometries [99—101]. These primitives can be
of the following types (Fig. 8): (a ) a circular tube of constant di­
ameter; (b ) an end-gated strip of constant thickness; (c ) a center­
gated sector of disk of constant thickness; (d ) a center-gated sector
of disk of constant thickness with nonzero inner radius. The flow in
each of these basic units is considered one dimensional. In addition,
similar primitives with thickness varying linearly along the flow axis
can be considered. For many cases such decomposition, though not
unique, is relatively straightforward and leads to a tree structure
(Fig. 9). The main complication arises when the cavity is filled by
more than one gate, leading to formation of a weldline in the cavity.
In this case the position of the weldline has to be guessed, and its
actual position has to be found by repeating the simulation until as­
sumed and simulated weldlines coincide. For handling of three-di­
mensional cavities, a lay-flat approach to flow analysis is usually em­
ployed. In particular. Fig. 10 shows a typical example of the lay-flat
of a three-dimensional plastic part. Flow analysis of this unfolded
part can be made by using one- or two-dimensional cavity-filling sim­
ulation. Applying simulation techniques developed for the cavity fill­
ing of thermoplastics is important for extending to the molding of ru b ­
ber compounds.
The flow of rubber compounds through the runner and gate during
molding is complicated in terms of'a theoretical description. In p ar­
ticular, runners and gates in injection molding are usually prepared
as ducts having noncircular cross section. Thus, the simulation of
nonisothermal flow of non-Newtonian fluids in noncircular ducts is
important for designing molding process. However, this problem
seems to have received little attention in the literature, although
there does exist an extensive literature dealing with fully developed
non-Newtonian-flow in noncircular ducts under isothermal conditions
[102—113]. Simulation of a power-law fluid in a square duct has re ­
cently appeared without including the coupling effect of the nonuni­
form temperature distribution upon the velocity and pressure fields
[114]. Further, there is another paper describing an empirical ap­
proach to this problem [115]. Recently, more general consideration
of the steady nonisothermal flow of shear-rate, temperature- and
pressure-dependent generalized Newtonian fluids in noncircular run­
ner of various shapes has been given [116]. In particular, the char­
acteristic dimension of the noncircular runner needed for simulation
has been expressed in terms of an ’’equivalent" circular runner in
which its radius is chosen based upon various recipes available from
the literature for fully developed isothermal flow. The predicted
449
Injection Molding of Rubber Compounds

(a)

Fig. 8 Primitives used to simplify cavity-fflling simulation.


450 Isayev

(a )

(b)

Fig. 9 Schematic representation of runner-gate-cavity system in terms


of primitives (a ) and various flow paths (b ).
Injection Molding of Rubber Compounds 451

(b)

Fig. 10 Three-dimensional plastic part (a ) and its lay-flat approxima­


tion ( b ).

results have been compared with experiments for various thermoplas­


tics. Specific examples of such comparisons are given in Fig. 11 for
pressure losses versus volumetric flow rates during nonisothermal
452 Isayev

E
u
'aJ
c
>N
•D
o

- 2

40 80
Q (cm /sec)

Fig. 11 Experimental (symbols) and predicted (curves) steady-state


pressure vs. volumetric flow rate for (a ) semicircular, (b ) trape­
zoidal, and (c ) circular arc runners (see Fig. 12) during nonisother-
mal flow of polystyrene melt (Styron 678U/Dow) at melt temperature
218°C, wall temperature 27°C. Solid and dashed curves are respec­
tively based upon hydraulic radius approach and radius determined
from cross-sectional area.
Injection Molding of Rubber Compounds 453

6 . 4 d ia . r u n n e r

(a ) (b ) (c)

Fig. 12 Plan view and end views of three noncircular runners of (a )


semicircular, (b ) trapezoidal, and (c ) circular arc cross section. The
location of pressure transducer is indicated as ( • ). Dimensions are
in millimeters.

flow of a polystyrene melt through runners of semicircular, trapezoi­


dal, and circular arc cross sections. The geometry of the runners is
shown in Fig. 12. Calculations of pressure drops are based upon re ­
placing noncircular ducts by circular tubes with the same hydraulic
radius or cross-sectional area. Using this approach to rubber com­
pound molding is desirable.
Available attempts of flow simulation in injection molding are based
upon the neglect of extra pressure drops arising from contraction or
expansion flow in the so-called junction region. In the range of strain
rates encountered in injection molding, such pressure losses have been
found to be about 100 bar [117]. In particular. Fig. 13 shows the
454 Isayev

Fig. 13 Dependence of extra pressure losses on shear rate at the wall


of the small tube during abrupt contraction flow of polystyrene melt
(Styron 678U/Dow) at temperature 192°C. Geometry of the die is
given in field of the figure. Dimensions are in millimeters.

dependences of extra pressure losses arising during flow of a poly­


styrene melt in an abrupt contraction from a large tube to a small
tube on shear rate at the wall of the small tube. As can be seen in
the range of shear rates covering three orders of magnitude, extra
pressure losses increase about 100 times. A major contribution from
this type of flow occurs at gates which are short ducts. The prob­
lem of evaluating pressure drops in short ducts is more complicated
in terms of a theoretical description. Local pressure losses at the
edges, additional pressure losses at the entrance and exit, and a de­
veloping velocity field and pressure gradient due to the short flow
paths and small residence times involved are all major features of
end effects during the flow of melts in a tube [118]. In dealing
Injection Molding of Rubber Compounds 455

with the flow of polymer melts in the gate and juncture regions of in­
jection-molding systems, it is important that one consider the visco­
elastic properties of the material. Although actual situations may be
three-dimensional and nonisothermal, initial efforts in this area are
being limited to isothermal two-dimensional flow situations. However,
even with these limitations, the problem is still difficult. Despite
many experimental and theoretical studies [119—143], at present
there does not seem to exist any systematic approach to solving the
problem of extra pressure losses. Theoretical investigation of such
flows has been performed on the basis of differential and integral
Maxwellian-type constitutive equations by using finite element analy­
sis [130—132,137,138]. However, due to numerical stability problems,
these simulation attempts have been restricted to rather limited elas­
ticity, the product of the characteristic relaxation time of fluid, and
the characteristic strain rate of process, whereas actual values of
about 1000 can be encountered in injection molding [30]. Agreement
between theory and experiment is, in any case, poor for pressure
losses. Much better agreement has been found in recent investiga-
tons [142,143] based on the Leonov constitutive equation. This is,
however, for viscoelastic polymer melts. There is another way to in­
corporate extra pressure losses in the juncture into the modeling of
injection molding—namely, by constructing master curves correlating
extra pressure drop with the structural, rheological, and processing
characteristics of the polymers and the geometrical dimensions of the
juncture. An initial attempt in this direction has already been made
[117], and further work is needed to extend this approach to other
polymers, including rubber compounds.
After the cavity is filled, the packing and vulcanization stages fol­
low simultaneously. In the packing stage, due to the presence of hold­
ing pressure, the weight and density of the molded part are formed.
In addition, the holding pressure partly affects the shrinkage, which
for molding of rubber compounds is dominated by the vulcanization
process and the temperature history. Fluid-mechanical treatments
of the packing stage are available for thermoplastics [77,144,145].
In particular, the approach is the same as for the filling stage,
with equations of continuity, motion, and energy being simplified and
coupled with the rheological equation. Since in the packing stage the
compressibility of the melt is dominant, the equation of state has to
be incorporated and solved simultaneously with other equations. The
above-mentioned papers do not present specific results for applying
this approach to the shrinkage in molded parts. The first attempt to
model shrinkage by using the equation of state was given in Ref. 146.
In particular, the multilayer model for shrinkage during the holding
pressure phase has been found to be only in qualitative agreement
with experimental data. Recently, investigators [147] have presented
456 Isayev

Fig. 14 Volume shrinkage versus volumetric flow rate for injection-


molded polystyrene (Styron 678U/Dow) strip (120 x 40 x 2.54 mm^)
at following processing conditions: o-melt temperature 220°C, pack­
ing pressure 200 psi; n-melt temperature, 240°C, packing pressure
1000 psi. Mold temperature 30°C.

theoretical and experimental results on shrinkage of thermoplastics in


injection molding. In Fig. 14 are shown the experimentally determined
dependencies of volume shrinkage on volumetric flow rate for two lev ­
els of the packing pressure in the cavity with a polystyrene melt. As
can be seen, there is little dependence of shrinkage on the flow rate.
At the same time, effect of the packing pressure on shrinkage is highly
pronounced; namely, a sharp decrease in shrinkage has been observed
with an increase in the packing pressure. In particular, theoretical
volume shrinkage has been determined by averaging the specific vol­
ume through the cavity space and time and subtracting the specific
volume at room temperature. Predicted and experimental results for
the shrinkage have been found to be in quantitative agreement (Fig.
15). An extension of this approach to the molding of rubber com­
pounds needs to be made .
Injection Molding of Rubber Compounds 457

Fig. 15 Experimental (dashed curve) and theoretical (solid curve)


shrinkage vs. volumetric flow rate for injection-molded polystyrene
(Styron 678U/Dow) strip (120 x 40 x 2.54 mm^) at following process­
ing conditions: melt temperature 200°C, mold temperature 30°C, and
packing pressure 500 psi. (From Ref. 147.)

Additional information has to be generated in order to take into


account shrinkage caused by the vulcanization process. So far,
the available literature does not consist of such information. The vul­
canization process is a transient heat conduction problem with inter­
nal heat generation due to the exothermic reaction of vulcanization.
In differential scanning calorimeter studies of the cure kinetics of
thermosets, the kinetic model has been proposed to describe the
dependence of the cure rate on the curing temperature and time by
<
u
Q
LU
>
-J
o
>
LU
H
<
LU
X
-J
<
0^
u
LU
h-
z

10 20 30 40 50 60 70

TIME MINUTES

Fig. 16 Calculated and experimental isothermal cure results for epoxy


at 94°C. (From Ref. 73.)

analyzing the amount of evolved heat [73,148—150]. It has been found


that the cure kinetics can be represented by the form

= (Ki + K^a^Xl-a)’^ (1)

where a is the relative degree of cure and K i(T ) and K 2 (T ) are func­
tions and m, n are constants. These two temperature functions and
two constants can be fitted from calorimetry measurements. The rel­
ative degree of cure has been defined as the ratio of the evolved heat
at any given time to the total heat evolved. Results in Fig. 16 indi­
cate that the proposed model fits well with experimental data obtained
on an epoxy system. Recently, the kinetic model for cure reactions
Injection Molding of Rubber Compounds 459

has been proposed from the concept of the nonequilibrium thermody­


namic fluctuation theory [75]. The model has successfully been applied
to describe the time-dependent torque in the Monsanto rheometer and
the storage modulus in the Rheometrics mechanical spectrometer ob­
tained during the vulcanization of a natural rubber compound under
isothermal conditions at different temperatures. However, vulcaniza­
tion in the mold is a highly nonisothermal process. Thus, to predict
the state of cure of rubber compound in the mold, one must calculate
the temperature distribution as a function of time. Based on the as­
sumptions of one-dimensional heat conduction with convective bound­
ary conditions, several authors have solved this equation coupled with
the reaction kinetics equation by using finite-difference numerical
techniques [151—157]. Recently, Schlanger [157] has given evidence
of successfully using this approach to cure tires in a compression
molding press, but he did not consider heat generation during the
cross-linking reaction. Even with this limitation the author has ob­
tained a good comparison between experimental and theoretical tem­
perature distributions inside the tire at different times (Fig. 17).
The long curing time is probably responsible for the good agreement
with theory. During injection molding of rubber, the curing time is
much shorter; therefore heat evolution by chemical reaction plus pro­
cessing energy will play a dominant role. Moreover, such approaches
to rubber injection molding are significant for determining rigidity of
the parts before demolding. A knowledge of the state of cure before
demolding helps to prevent excessive warpage and porosity in thick
sections.

F. C ha ra c te riza tio n o f R u b b e r Molding

In many applications injection-molded rubber products are used as


critical parts whose failure would result in great losses. Thus, the
success of the molding operation is also linked to the performance of
the molded part. At the present time the performance of the part in
a particular environment is mainly evaluated on the basis of the na­
ture of the compound by measuring physical properties on laboratory
sheeted vulcanizate. Such an approach is not sufficient for injection-
molded parts. For injection-molded thermoplastic parts it has been
generally recognized that their performance characteristics are highly
anisotropic and dependent upon molding conditions [1].
Frozen-in orientation and residual stresses [158,159], which are a
combination of flow and thermal stresses and orientation introduced
during the molding operation, affect the performance of the final prod­
uct. For rubber-molded parts, frozen-in flow and curing stresses
will probably dominate over thermal stresses. The latter is negligible
due to the absence of a transition to the glassy state during rubber
460 Isayev

PROBE

Fig. 17 Schematic diagram of tire section (a ) and its equivalent model


(b ) , (c ) Experimental (symbols) and calculated (solid curves) tem­
perature distribution inside tire at different times during curing. In­
itial steam temperature 201^0, mold temperature 140°C. (Data from
Ref. 157.)
Injection Molding of Rubber Compounds 461

(c)

molding. In particular, higher flow stresses will be frozen-in in ru b ­


ber parts than have been observed in thermoplastic parts because ru b ­
ber compounds have high viscosity and relaxation time, which cause
a slow release of the stresses introduced during cavity filling.
Curing stresses are introduced due to rapid contraction during the
vulcanization of the rubber in the mold. In addition to the residual
stresses, rubber injection molding introduces anisotropy of physical
properties in the final products. This is especially relevant for thick
462 Isayev

(a)

F ig . 18 (a ) Center-gated disk injection molded from SBR 1805 and lo­


cations of the dumbbell samples, (b ) The stress-strain curve for the
dumbbell samples parallel ( ||) and perpendicular to flow direction,
(c ) The 100% modulus after 10 min relaxation as a function of degree of
cross-linking for the dumbbell samples cut parallel ( |j) and perpendic­
ular i\) to flow direction. Barrel, nozzle, and mold temperatures are
70°C, 90°C and 165°C, respectively. Volumetric flow rate is 70 cm^/
sec. (Data from Ref. 163.)

moldings. In general, there is expected to be a distribution of cross­


links density through the thickness of the rubber molding caused by
various temperature histories in the core and wall regions. This will
introduce performance characteristics different from those obtained
in the laboratory. Despite this obvious effect of injection molding
upon product performance, studies of the anisotropy of the physical
and mechanical properties of rubber-molded parts have received lim­
ited attention. In particular, for rubber compression molding, flow-
induced anisotropy measurements have been carried out in Refs. 160—
162. For rubber injection molding some recent measurements of the
degree of cross-linking, stress-strain dependencies, and stress re ­
laxation have been performed both parallel and perpendicular to the
flow [163]. In particular, a center-gated circular disk made from sty­
rene-butadiene rubber compound has been prepared [Fig. 1 8 (a )].
Injection Molding of Rubber Compounds 463

It has been concluded that orientation rather than the state of cure
is responsible for the observed anisotropic behavior of the stress-
strain curve and relaxation modulus [Fig. 1 8 (b ,c )]. On the other
hand, the degree of cross-linking has been found to be independent
of direction, injection speed, and shot size; it depends only on cure
time. Further, the effect of cure times and various acceleration com­
binations in natural, nitrüe-butadiene, and styrene-butadiene rubber
compounds and ethylene-propylene-diene terpolymers have been re ­
ported. Injection and compression moldings have been studied. Ten­
sile modulus, hardness, ultimate tensile stress, and elongation have
464 Isayev

been determined [164]. Tensile strength and hardness for injection


molding are found to be, respectively, higher and lower than for com­
pression molding. Based upon measurements of the properties of in­
jection-molded parts, some correlation between the cure time deter­
mined from an oscillating disk rheometer and the best cure time ob­
served in the injection molding has been found.
Results of earlier studies of physical properties of injection-molded
vulcanizates have been reviewed in Ref. 3. However, these inves­
tigations have not paid particular attention to anisotropy of physical
properties.

III. SUGGESTIONS FOR F U T U R E WORK

A. General

The scientific basis for injection molding of rubber compounds can be


developed by considering the process as one integrated system. In­
itially, the major thrust of the research has to be directed toward the
measurements of the properties of the rubber compounds, including
rheological and thermal characteristics and pressure-volume-tempera­
ture relationships incorporated with vulcanization kinetics. Accord­
ingly, these measurements will allow the formulation of theoretical ex­
pressions necessary to carry out simulation of the flow dynamics and
thermal history of rubber compounds in the screw, nozzle, sprue, run­
ners, gates, and cavity. Moreover, the equation of state and models
for the vulcanization kinetics and the extent of cure are extremely im­
portant for the theoretical and experimental investigations of the post­
filling behavior of the rubber compound in the mold. In addition, ex­
perimental and theoretical studies of the structure-property relation­
ships of the cured-in orientation and residual stresses and anisotropy
of properties and shrinkage are necessary to characterize and predict
the performance of the molded rubber products. This outlined pro­
gram for the basic study of injection molding of rubber compounds
will lead to the formulation of the fundamentals of the process. We
next give a description of specific tasks to be undertaken in future
research.

B. Characterization o f R u b b e r Compounds

Characterization of rubber compounds should include extensive mea­


surements of rheological properties, such as the dependence of the
apparent viscosity, normal stresses, and extrudate swell on the shear
rate and their time-dependent behavior. The determination of rheo­
logical properties of various rubber compounds as a function of con­
centration and structure of carbon black fillers and their surface area
Injection Molding of Rubber Compounds 465

and the effect of other ingredients have to be performed. Special at­


tention has to be given to rheological changes associated with the thix­
otropic nature of filled rubber systems. Rheological properties at low
deformation rates can be measured in a sandwich rheometer, and high­
er deformation rates in a capillary rheometer. This will supply suffi­
cient experimental data to develop constitutive equations.
Rubber compounds have a peculiar behavior—a tendency to slip
along the solid boundary during flow. Slip velocity depends on pres­
sure, temperature, and shear stress. Slip velocity can be measured
by the Mooney capillary method. To investigate the effect of pres­
sure upon slip velocity, one must construct a modified capillary rhe­
ometer. The major difference between this rheometer and a conven­
tional one is the presence of a second pressurized chamber where the
pressure is kept below that imposed in the first reservoir with the
compound. This pressurized chamber is schematically shown in Fig.
19. It is essential that the pressure drop during flow in a capillary

Fig. 19 Schematic drawing of pressurized die system for slip velocity


measurements.
466 Isayev

between two pressurized chambers be kept at a level significantly low­


er than the absolute pressure level. In this case a slip velocity Vs at
various temperatures and shear stresses at the wall and pressures
p can be determined by the expression Vs = 9(32Q/ttD^)/9(1/D) -p ^
where Q is volumetric flow rate and D is the capillary diameter. An­
other way to carry out experimental slip studies is to develop a pres­
surized coaxial cylinder viscometer which can be used to determine
slip by comparing torque as a function of rotor radius and smooth­
ness. However, this requires a development of a highly complicated
instrument.
An experimental determination of the extra pressure loss in the
juncture and gate regions of the injection-molding system would also
have to be performed. Their relation to elastic properties of rubber
compounds has to be established. These pressure losses account for
significant portions of the total pressure drop. Consideration should
be given to junctures having converging, diverging, abrupt-contrac­
tion, and abrupt-expansion areas and T and elbow shapes. Some
semiempirical correlations between extra pressure drops and the struc­
tural, rheological, and processing characteristics of rubber compounds
and the geometry of various junctures have to be established.
Experimental data have to be obtained on the effect of vulcaniza­
tion upon the rheological properties of various compounds by using a
Monsanto processibility tester in the range of temperatures, times,
and shear rates relevant to injection molding. Other conventional
techniques for rubber compound characterization have to be used,
particularly, extensive measurements of the scorch time by an oscil­
latory disk rheometer and differential scanning calorimeter.
Kinetic parameters characterizing the extent of cure and the heat
of vulcanization have to be determined together with data on the pres­
sure-volume-temperature relationships of rubber compounds during
vulcanization. Also, thermal characteristics such as specific heat and
thermal diffusivity of rubber compounds and their vulcanizates have
to be determined. From these values thermal conductivity can be cal­
culated. These properties are paramount for describing heat balance
during molding. The proposed basic studies of physical properties of
rubber compounds are necessary steps toward creating an extensive
material data bank for various injection-molding rubber compounds .
This information will be applied for simulation of the molding process.

C. E xperim ental In v e s tig a tio n and Modeling


o f th e Molding Process

Experimental investigation and modeling of the injection-molding pro­


cess for rubber compounds in a screw extruder during plastication
and in the nozzle, sprue, runner and gate system, and cavity during
Injection Molding of Rubber Compounds 467

filling stage have to be carried out. In addition, theoretical and ex­


perimental investigations of the post-filling stage, such as packing
and vulcanization, should be considered.

Initial Conditions for Rubber Entering Mold


As a first approximation, temperatures and pressures may be mea­
sured for the rubber compounds at the entry to the mold system. It
is preferable, however, to develop these from first principles. Such
an approach is outlined in the following paragraphs, where principles
of the modeling of the screw part of the machine are described. Rub­
ber compounds are usually fed into the extruder in the form of a con­
tinuous strip at a temperature significantly above the glass transition.
Therefore in rubber extrusion a melting zone is not present. Thus,
the flow of rubber compound in the extruder can be treated as the
flow of the melt in the metering zone. The main differences in the
analysis of rubber flow in the metering zone arise, however, due to
the presence of the slip condition on the solid boundary and yield
stress in the material. The temperature of the material in the screw
is usually kept below the vulcanization temperature. Therefore the
effect of vulcanization can be neglected at this stage. The major
source of the heat supply and the corresponding temperature rise
in the rubber extrusion appear from the viscous dissipation produced
by the flow of the high-viscosity rubber compounds. Hence the ex­
trusion process is highly nonisothermal such that even cold-feeding
arrangements may require a cooling system. The major thrust must
be to determine the temperature distribution and bulk temperature of
the rubber compound in the barrel before injection. The temperature
level is a crucial factor in avoiding scorching during molding.
When considering screw extrusion of rubber compounds, one can
assume that the melt flow is in a rectangular channel with the screw
surface stationary and the barrel considered as a top plate moving
along and perpendicular to the channel.
The flow of a non-Newtonian fluid in such a channel is governed
by the following equations of motion:

0= ( 2)
ax ay\ ay j

iP + i. I 9w\
(3)
3z 35 r 3y)

where x, y, z denote width, depth, and length directions, respectively,


p is pressure, u (y ), v = 0, w (y ) are the velocity components in x, y,
and z directions and n ( t , T ) is the dependence of the steady-state
468 Isayev

viscosity on the shear rate. The heat balance is governed by ther­


mal conduction, convection, and dissipation. By assuming that heat
conduction is only dominant in the y direction over heat convection
and by neglecting transverse convection, we can write the energy
equation as

pC w — nY (4)
p 3z

where p, Cp, k, T are, respectively, the density, specific heat,


thermal conductivity, and temperature of the rubber compound, and
Y is the strain rate, determined as y = [O u /3 y )^ + ( 3w/3y)^] °
The temperature and velocity boundary conditions are

u = Vg sin 6, w Vg cos 0, T at y = 0

V -Vc, sin e, w = V cos 6, T = T, at y = H (5)


X S z S b

where Tp is the barrel temperature, Vs the slip velocity, Vx =


ttDN sin 0 , and Vz = irDN cos 0 are the velocity components on the
moving barrel, and H , N , D are the channel depth, screw rotational
speed, and diameter of the barrel, respectively.
Neglecting leakage over the flights means the transverse flow rate
in the channel is zero:

H
r.
u dy = 0 ( 6)
L

The output flow rate is

Q =B i w dy (7)

where B is the channel width. The specific form of the rheological


constitutive equation included in Eqs. (2) —(4) and the dependence
of Vs upon temperature, shear rate, and pressure in Eq. (5) has to
be experimentally determined. The necessary procedure for doing so
was outlined earlier.
Equations (2) —(6) can be solved by finite differences. Therefore,
pressure-output functions, temperatures, and velocity profiles can be
determined. The bulk temperature of the material in the barrel will
be subsequently calculated and used in further modeling steps.
Injection Molding of Rubber Compounds 469

F l o w D y n a m i c s in N o z z l e , Spru e, R unner, a n d G ate

The flow of rubber compounds in the nozzle, sprue, runner, and gate
can be considered to be one dimensional. Our goal here is to deter­
mine the pressure drop, temperature and velocity distributions, and
the bulk temperature at the entrance to the cavity. The cross-sec­
tional areas of the nozzle and sprue are usually circular. The run­
ner and gate are frequently made with a noncircular cross section and
can be approximated by a hydraulic radius equal to twice the area di­
vided by the wet perimeter. This approach has been found to be the
most reliable in assessing pressure drops during the flow of non-New­
tonian fluids in channels of noncircular cross section. The equations
of motion, energy, and continuity for flow in the nozzle, sprue, run­
ner, and gate in cylindrical coordinates are

0 = -3P + l l / r ( 8)
3z r 3r\ 3r j

_ /3T ^ 3T\ _ K 3 /.
r + R H (9)
^ p\ 3t ^ 3z) r 3r(' 3r / V V

R
Q = 2u ru dr ( 10 )

where r, z are the radial and axial coordinate, respectively, u is the


axial velocity, T is the temperature, p is the pressure, n is the ap­
parent voscosity versus shear rate, y = 3u/3r, and temperature, Ry
and Hv are the rate and heat of vulcanization, respectively, R is the
radius or hydraulic radius, and Q is the volumetric flow rate. The
boundary conditions on u and T are

3T
= 0 at r = 0
3r 3r

u = V s ,T at r = R ( 11 )
w

where Vg is the slip velocity and T ^ is the runner wall temperature.


The specific form of functions n(Y, T, t ), Ryip, T, t ), and V s (p ,
T, t) has to be determined from the experiment outlined in Section
III.B .
The inlet temperature at the entrance to a subsequent section can
be taken to be uniform and equal to the bulk temperature of the pre­
vious section. The inlet temperature at the entrance to the nozzle
will be equal to the bulk temperature in the barrel before injection.
470 Isayev

The advancing front of the rubber compound can be assumed to be fla t,


with the temperature there equal to the temperature at the inlet.
Extra pressure drops arising at the juncture area, such as occurs
in the transition region from the barrel to the nozzie, from the sprue
to the runner, from the runner to the gate, and from the gate to the
cavity, can be added to the total pressure losses. These extra pres­
sure losses can be determined on the basis of the experimental inves­
tigation outlined in Section III.B . The development of pressure and
bulk temperature is the major thrust of this simulation.

Cavity-Filling Simulation
The one- and two-dimensional cavity-filling simulations have to be
performed. The modeling approach used in the injection of the ther­
moplastics can, with some modification, be applied to the cavity-fill­
ing simulation of the rubber compounds. It is customary to employ
the classical Hele-Shaw approximation for such flows. Equations of
motion, continuity, and energy for filling thin cavities with inelastic
non-Newtonian fluids under nonisothermal conditions are

( 12)
3x 3y\ dyj

0 = - 3P + ± L (13)
3z dy\ 3y/

(14)
è * Tz “

„ /9T 8T
pc Hr: + U T— + V — = K ^ + nY^ + H R (15)
p\ 3t 8x 8z / 3y2 ' V V

where x, y, z are the width, gapwise coordinate, and flow coordi­


nate, u and V are the velocity components in the x and z directions,
respectively, u and v are the gapwise average velocity components,
and n(Y, T, t) is apparent viscosity versus shear rate, temperature,
and time. The shear rates are

Boundary conditions for Eqs. (12) —(15) are

V V , T at z = ±b
w
8u _ 8y _ 8T
= 0 at z = 0 (16)
8y 8y 8z
Injection Molding of Rubber Compounds 471

where Vs is the slip velocity and is the mold temperature. In p ar­


ticular, the equation of motion is formulated for shear-dominated flow
in which the shear stresses are balanced against the pressure gradi­
ent. The equation of energy includes convection, conduction, vis­
cous dissipation, and heat of vulcanization terms. The specific forms
of functions n(Y» T, t ), V s (T , p, t ), and RyCp» T, t) have to be de­
termined from experiments. The distribution of the temperature, ve­
locity, extent of cure, rubber front location, and pressure trace in
the cavity can be predicted.

Simulation of Post-Filling Stages in the Mold


The post-filling stage is started when the cavity is completely filled.
This stage also may be called the vulcanization stage. A few major
effects have to be considered : the distribution of temperature and
extent of cure in the cavity space and time and the shrinkage and
residual stresses. Since at the post-filling stage the flow effect is
negligible, the convection and viscous dissipation terms in Eq. (15)
can be omitted. The energy equation for the post-filling stage can
then be written as

pC ^ = K ^ + HR (17)
P 3t 3y2 V V

with the initial conditions for the gap wise temperature distribution in
the cavity at the post-filling stage taken to be equal to those deter­
mined at the end of the cavity filling. This equation along with the
equation for the rate of vulcanization have to be solved simultaneously
to get the distribution of temperature and extent of cure in the cavity
space and time. In particular, computations of these quantities will
give a guideline for determining the time for successful ejection of the
molded part.
The volume shrinkage S can be determined by

V .-V .
S - ^ ^ (18)
V.
1

where Vf is the specific volume of rubber compound at room tempera­


ture , Vi is the specific volume averaged through the cavity space and
time such that

v . = : i - J P V (t ) dt (19)
0
472 Isayev

where V (t) is the specific volume averaged through the cavity space
at any time t , and tp is the time at which the pressure in the cavity
becomes zero. The equation of state is needed to find the current
specific volume at any pressure and temperature. This equation has
to be determined experimentally.

Experimental Studies of the Molding Process


The experimental study of the molding process should have two as­
pects. The first is to investigate the mold filling and test the mod­
els from the previous sections. These experiments should be p er­
formed at various injection rates and pressures, mold and melt tem­
peratures, etc., for rheologically and cure-characterized compounds.
The second is to prepare molded parts for structural and mechanical
characterizations. These molded parts should be prepared under dif­
ferent molding conditions by varying injection speed, stock and mold
temperatures, and level and duration of packing pressure and time.
Experimental traces of the nozzle, runner, gate, and cavity pres­
sures have to be determined in order to be compared with the pre­
dictions following from the proposed simulation. Distributions of the
extent of cure and shrinkage in molded parts obtained at various
processing conditions should be measured and compared with results
of the simulation. Specific attention has to be paid to the presence
of anisotropy of the mechanical properties, including relaxation mod­
ulus, ultimate strength, and elongation to break. Orientation and
residual stress in the rubber moldings have to be determined. In
particular, residual stresses and orientation in the rubber-molded
parts are a coupled effect from two main sources. The first is the
flow-induced stresses, shear and normal stresses, which develop
during the cavity filling and lead to orientation of the macromolecu-
lar chains and carbon black network. These stresses do not com­
pletely relax due to the vulcanization process, and they appear as
frozen-in flow stresses in the molded part.
The second source is due to the contraction of rubber in the cav­
ity caused by vulcanization. These stresses may be called vulcaniza­
tion or curing stresses. In contrast to the molding of thermoplas­
tics, in rubber-molded parts, thermal stresses have negligible con­
tribution to the total residual stresses. This is due to the absence
of solidification in rubber injection molding. A rheo-optical technique
is most suitable for measuring residual stresses in rubber moldings
prepared from transparent rubber models. Comparison of experimen­
tal results for the residual stresses has to be made with the theoreti­
cal prediction following from previously developed theories for ther­
moplastics modified for rubber molding.
Injection Molding of Rubber Compounds 473

IV . CONCLUDING REMARKS

A brief review of the previous studies related to injection molding of


rubber compounds has been presented along with some suggestions
for future work. The main thrust of future research has to be di­
rected toward developing a scientific basis for injection-molding tech­
nology of rubber compounds by considering the process as an inte­
grated system. An interdisciplinary approach has to be taken, which
involves the contribution of various engineering disciplines, including
polymer rheology, reaction engineering, fluid mechanics, heat trans­
fer, and polymer science. Specific areas for future research must in­
clude (1) quantitative characterization of the rubber compounds by
measurements of rheological and thermal properties, vulcanization ki­
netics, and slip boundary conditions at the solid surface, (2) formu­
lation of rheological constitutive relations and equation of state, (3)
creation of an extensive material data bank for various rubber com­
pounds, (4) experimental and theoretical investigations of flow dy­
namics in the screw, delivery system, and cavity, including the post­
filling stage, with consideration given to packing and vulcanization in
the cavity, (5) experimental measurements and development of quanti­
tative theory for characterizing the shrinkage in molded parts and
frozen-in orientation and residual stresses, (6) experimental study
of anisotropy of mechanical properties of molded parts effected by
processing conditions and the nature of the rubber compound.

ACKNOWLEDGMENTS

The author would like to thank Mr. F. Weissert and Dr. J. L. White
for their helpful comments. This work was supported by a grant
from the National Science Foundation, Division of Engineering.

REFERENCES

1. I. I. Rubin, Injection Molding, Theory and Practice, Wiley, New


York, 1972.
2. R. R. Barnhart, in Encyclopedia of Chemical Technology, Vol.
20, Wüey, New York, 1982, p. 445.
3. M. A. Wheelans, Injection Molding of Rubber, Butterworth,
London, 1974.
4. M. A. Wheelans, Prog. Rubber TechnoL, 42: 71 (1979).
5. M. A. Wheelans, European Rubber J., 163: Feb. (1981).
6. K. Palit and M. Barel, Plast. Rubber Processing, 37: 37 (1979).
474 Isayev

7. P. S. Johnson, Elastomerics, May: 22 (1983).


8. B. G. Crowther, Prog, Rubber Technology, 54: 55 (1981).
9. E. G. Kontos, Rubber Chem, TechnoL, 43: 1082 (1970).
10. J. J. Boguslawski, Rubber Chem, TechnoL, 45: 1421 (1972).
11. D. H. Smith and R. L. Cristy, Rubber Chem, TechnoL, 45: 1434
(1972).
12. R. L. Cristy, Rubber World, 180: 30 (1979).
13. R. M. Griffith, lEC Fund, 1: 180 (1962).
14. F. W. Kroesser and S. Middleman, Polym, Eng, Sci,, 5: 230
(1965).
15. M. Narkis and A. Ram, Polym, Eng, Sci,, 7: 161 (1967).
16. H. J. Zamodits and J. R. A. Pearson, Trans, Soc, RheoL , 13:
357 (1969).
17. Z. Tadmor and I. Klein, Engineering Principles of Plasticating
Extrusion, Van Nostrand Reinhold, New York, 1970.
18. R. T. Fenner, Extruder Screw Design, Iliffe, London, 1970.
19. R. T. Fenner and J. G. Williams, J, Mech, Eng, Sci,, 13: 65
(1971).
20. R. T. Fenner, Polymer, 16: 298 (1975).
21. C. D. Han, Rheology in Polymer Processing, Academic Press,
New York, 1976.
22. G. V. Vinogradov and A. Ya. Malkin, Rheology of Polymers,
Springer-Verlag, New York, 1980.
23. R. I. Tanner, J, Polym, Sci, A2, 8: 2067 (1970).
24. J. L. White, RheoL Acta, 4: 600 (1975).
25. K. R. Reddy and R. I. Tanner, J, Rheol,, 22: 661 (1978).
26. J. L. White and Y. Ide, J. AppL Polym, Sci,, 22: 1061, 3057
(1978).
27. W. Dietz and J. L. White, Rheol, Acta, 17: 676 (1978).
28. A . S, Wineman, J, Non-Newt, Fluid Mech,, 4: 249 (1978).
29. M. J. Crochet and R. Kuenings, J, Non-Newt, Fluid Mech,, 7:
199 (1980); 10: 339 (1982).
30. A. I. Isayev and C. A. Hieber, Rheol. Acta, 19: 168 (1980).
31. B . Bernstein, M. K. Kadivar, and D. S. Malkus, Computer
Methods in AppL Mech, Eng,, 27: 279 (1981).
32. B. Caswell and M. Viriyayuthakorn, J, Non-Newt, Fluid Mech,,
12: 13 (1983).
33. N. V. Zakharenko, F. S. Tolstukhina, and G. M. Bartenev, Rub,
Chem, TechnoL, 35: 326 (1962).
34. G. V. Vinogradov, A . Ya. Malkin, E. P. Plotnikove, O. Y. Sabsai,
and N. E. Nikolayeva, Int, J, Polym, Mater,, 2: 1 (1972).
35. L. F. Ramos de Valle, and F. Aramburo, J, Rheol,, 25: 379
(1981).
36. S. Toki and J. L. White, J, AppL Polym, Sci,, 27: 3171 (1982).
Injection Molding of Rubber Compounds 475

37. S. Montes and J. L. White, Rubber Chem. TechnoL , 55: 1354


(1982) .
38. F. M. Chapman and T. S. Lee, Soc. Plastics Eng, J,, 26: 37
(1970).
39. V. M. Lobe and J. L. White, Polym. Eng. S ci., 19: 617 (1979).
40. H. Tanaka and J. L. White, Polym. Eng. Sci., 20: 949 (1980).
41. Y. Suetsugu and J. L. White, J. Appl. Polym. Sci., 28: 1481
(1983) .
42. C. D. Han, Multiphase Flow in Polymer Processing, Academic
Press, New York, 1981.
43. J. L. White, J. Non-Newt. Fluid Mech., 5: 179 (1979).
44. J. L. White and H. Tanaka, J, Non-Newt. Fluid Mech., 8: 1
(1981).
45. Y . Suetsugu and J. L. White, J. Non-Newt. Fluid Mech., 14:
121 (1984).
46. M. Mooney, in Rheology, F. R. Eirich, E d., Academic Press,
New York, 1958.
47. M. Mooney, Rubber Chem. TechnoL, 35: 27 (1962).
48. G. V. Vinogradov and L. I. Ivanova, Rheol. Acta, 6: 209 (1967).
49. J. L. Den Otter, Rheol. Acta, 14: 329 (1975).
50. G. P. Colbert, Rubber Plastics News, 22: April 2 (1979).
51. R. W. Whorlow, Rheological Techniques, Wiley, New York, 1980.
52. R. J. Crow son, A. J. Scott, and D. W. Saunders, Polym. Eng.
Sci., 21: 12 (1981).
53. J. D. Byam and G. P. Colbert, ACS Rubber Division Meeting,
Chicago, May 1977, Paper #21.
54. R. H. Norman and P. S. Johnson, Rubber Chem. TechnoL,
54: 493 (1981).
55. J. L. Leblanc and E. A. Pintens, Kautschuk, Gummi, Kunst,
34:34 (1981).
56. J. A . Sezna, ACS Conference on Advances in Polymer Pro­
cessing, Akron, Ohio, May 1984.
57. K. Oda, J. L. White, and E. S. Clark, Polym. Eng. Sci., 16:
585 (1976).
58. J. C. Illman, J. Appl. Polym. Sci., 10: 1519 (1966).
59. R. Jenkins and L. Karre, J. Appl. Polym. Sci., 10: 303
(1966).
60. T. Imai, J. Appl. Polym. S e t, 11: 1055 (1967).
61. T. F. Saunders, M. F. Levy, and J. F. Serino, J. Polym. Sci.
Al, 5: 1609 (1967).
62. P. Fojolka and Y. Shabab, J. Polym. Sci. A l, 6: 1217 (1968).
63. R. A. Fava, Polymer, 9: 137 (1968).
64. G. J. Learmouth and G. Pritchard, J. Appl. Polym. Sci., 13:
2119 (1969).
476 Isayev

65. M. A. Acitelli, R. B. Prime, and E. Sacher, Polymer, 12: 335


(1971).
66. M. G. Rogers, J. Appi. Polym. Sci., 16: 1953 (1972).
67. J. I. Koenig and P. T. K. Shih, J. Appi. Polym. Sci. A2, 10:
721 (1972).
68. R. B . Prime, Polym. Eng. Sci., 13: 365 (1973).
69. E. Sacher, Polymer, 14: 91 (1973).
70. M. R. Kamal and S. Sourour, Polym. Eng. Sci., 12: 59 (1973).
71. M. R. Kamal, S. Sourour, and M. Ryan, SPE Tech. Papers, 19:
187(1973).
72. S. Sourour and M. R. Kamal, Thermochimica Acta, 14: 41 (1976).
73. M. R. Kamal and M. E. Ryan, Polym. Eng. Sci., 20: 859 (1980).
74. J. D. By am, G. P. Colbert, and J. F. Hagman, Elastomerics,
Nov. : 26 (1981).
75. H. S. Y. Hsich, J. Appi. Polym. Sci., 27: 3265 (1982).
76. F. Buschhaus and W. Benfer, Advances Polym. Technol., 2: 283
(1982).
77. M. R. Kamal and P. G. Lafleur, Polym. Eng. Sci., 22: 17 (1982).
78. R. T. Fenner, Principles of Polymer Processing, Chemical Pub­
lishing, New York, 1980.
79. S. Richardson, J. Fluid Mech., 56: 609 (1972).
80. S. Middleman, FundameniaZs of PoZymer Processing, McGraw-Hill,
New York, 1977.
81. H. L. Toor, R. L. Ballman, and L. Cooper, Modern Plastics,
38:117(1960).
82. V. N. Grimblat, Soviet Plastics, 2: 24 (1970).
83. D. H. Harry and R.G. Parrot, Polym. Eng. Sci., 10: 209 (1970).
84. M. R. Kamal and S.Kenig, Polym. Eng. Sci., 12:294 (1972).
85. M. R. Kamal and S.Kenig, Polym. Eng. Sci., 12:302 (1972).
86. J. L. Berger and C.G. Gogos, Polym. Eng. Sci., 13: 102 (1973).
87. P. C. Wu, C. F. Huang, and C. G. Gogos, Polym. Eng. Sci.,
14: 223 (1974).
88. Z. Tadmor, E. Broyer, and C. Gutfinger, Polym. Eng. Sci.,
14: 423 (1974).
89. J. L. White, Polym. Eng. Sci., 15: 44 (1975).
90. G. Williams and H. A. Lord, Polym. Eng. Sci., 15: 553 (1975).
91. G. Williams and H. A. Lord, Polym. Eng. Sci., 15: 569 (1975).
92. E. Broyer, C. Gutfinger, and Z. Tadmor, Trans. Soc. Rheol.,
19: 423 (1975).
93. M. R. Kamal, Y. Kuo, and P. H. Doan, Polym. Eng. Sci., 16:
585 (1976).
94. Y. Kuo and M. R. Kamal, AIChE J., 22: 661 (1976).
95. J. R. A . Pearson, Mechanical Principles of Polymer Melt Pro­
cessing, Pergamon Press, Oxford, 1976.
Injection Molding of Rubber Compounds 477

96. W. L. Krueger and Z. Tadmor, Polym. Eng, Sei., 20: 426


(1980) .
97. C. A. Hieber and S. F. Shen, J. Non-Newt, Fluid Mech. , 7:
1 (1980).
98. C. A. Hieber, L. S. Socha, S. F. Shen, K. K. Wang, and
A. I. Isayev, Polym. Eng. Sei., 23: 20 (1983).
99. S. M. Richardson, H. J. Pearson, and J. R. A. Pearson,
Plast. Rubber Proeessing, 55: June (1980).
100. A. M. Varner, SPE Teeh. Papers, 27: 290 (1981).
101. C. A. Hieber, SPE Teeh. Papers, 28: 356 (1982).
102. R. S. Schechter, AIChE J . , 7: 445 (1961).
103. J. A. Wheeler and E. H. Wissler, AIChE J . , 11: 207 (1965).
104. D. W. McEachern, AIChE J., 12: 328 (1966).
105. W. Kozicki, C. H. Chou, and C. Tiu, Chem. Eng. Sei., 21:
665 (1966).
106. N. Mitsuishi, Y. Kitayama, and Y. Aoyagi, Int. Chem. Eng.,
8: 168 (1968).
107. N. Mitsuishi and Y. Aoyagi, Chem. Eng. Sei., 24: 309 (1969).
108. T. Arai and H. Toyoda, Proc. Fifth Intern. Congress on Rhe­
ology, Vol. 4, 1970, p. 461.
109. F. Ramsteiner, Kunststoffe, 61: 943 (1971).
110. C. Miller, Ind. Eng. Chem. Fun dam., 11: 524 (1972).
111. R. W. Hanks, Ind, Eng. Chem. Fundam. , 13: 62 (1974).
112. A . I. Isayev, K. D. Vachagin, and A. M. Nähereznov, J. Eng.
P hys., 27: 998 (1974).
113. J. L. White and D. C. Huang, Polym. Eng. Sei., 21: 1101
(1981) .
114. A . R. Chandrupatla and V. M. K. Sastry, Numerieal Heat
Transfer, 1: 243 (1978).
115. J. B. Smith, Plast. Rubber Proeessing, 17: March 1979.
116. C. A. Hieber, R. K. Upadhyay, and A. I. Isayev, SPE Teeh.
Papers, 29: 698 (1983).
117. A. I. Isayev and B. Chung, SPE Teeh, Papers, 30: 433
(1984).
118. J. M. McKelvey, Polymer Processing, Wüey, New York, 1962.
119. W. Phüippoff and F. H. Gaskins, Trans. Soe. Rheol., 2: 263
(1958).
120. E. B. Bagley, Trans. Soe. Rheol., 5: 355 (1961).
121. T. Arai and H. Aoyama, Trans. Soe. Rheol., 7: 333 (1963).
122. H. L. La Niewe and D. C. Bogue, J. Appl. Polym. Sei., 12:
353 (1968).
123. R. L. Boles, H. L. Davis, and D. C. Bogue, Polym. Eng.
Sei., 10: 24 (1970).
124. N. D. Sylvester and S. L. Rosen, AIChE J., 16: 964 (1970).
478 Isayev

125. T. Arai, Proc. 5th Intern. Congress on Rheology, Vol. 4,


1970, p. 497.
126. T. F. Balienger and J. L. White, J. Appl, Polym, Sci,, 15:
1949(1971).
127. A. L. Halmos and D. V. Boger, AIChE J ., 21: 550 (1975).
128. V. Z. Volkov, V. P. Fikhman, G. V. Vinogradov, and A. I.
Isayev, J. Eng, Phys., 32: 51 (1977).
129. J. L. White and A. Kondo, J, Appl, Polym, Sci,, 21: 2289
(1978).
130. P. W. Chang, Th. W. Patten, and B . A. Finlay son. Comp.
Fluids, 7: 267 (1979).
131. M. Viriyayuthakorn, and B . Caswell, J, Non-Newt, Fluid
Mech, , 6: 245 (1980).
132. C. J. Coleman, J, Non-Newt, Fluid Mech,, 7: 265, 289 (1980).
133. D. V. Boger and M. M. Denn, J. Non-Newt, Fluid Mech,, 6:
163(1980).
134. K. Geiger, SPE Technical Papers, 27: 693 (1981).
135. H. J. Yoo, and C. D. Han, J, Rheol,, 25: 115 (1981).
136. L. Choplin and P. J. Carreau, J. Non-Newt. Fluid Mech,, 9:
119 (1981).
137. M. A. Mendelson, P. W. Yeh, R. A. Brown, -ind R. C. Arm­
strong, J, Non-Newt. Fluid Mech., 10: 31 (1982).
138. N. R. Jackson and B. A. Finlayson, J. Non-Newt, Fluid Mech,,
10: 85 (1982).
139. H. M. Laun, Rheol, Acta, 22: 171 (1983).
140. D. Curto, F. P. La Mantia, and D. Acierno, Rheol, Acta, 22:
197 (1983).
141. F. P. La Mantia, A . Valenza, and D. Acierno, Rheol, Acta, 22:
299, 308 (1983).
142. A . I. Isayev, R. K. Upadhyay, and S. F. Shen, SPE Techni­
cal Papers, 28: 298 (1982).
143. R. K. Upadhyay and A. I. Isayev, SPE Technical Papers, 29:
714 (1983).
144. Y. Kuo and M. R. Kamal, M .I.T . Polymer Processing Confer­
ence, 1977, p. 329.
145. T. S. Chung and M. E. Ryan, Polym, Eng, Sci,, 21: 271 (1981).
146. H. O. Hellmeyer and G. Menges, SPE Tech, Paper, 22:386
(1976).
147. A. I. Isayev and T. Hariharan, SPE Tech, Papers, 30: 765
(1984).
148. M. R. Kamal, Polym, Eng, Sci,, 14: 231 (1974).
149. R. C. Progelhof and J. L . . Throne, Polym, Eng, Sci,, 15: 690
(1975).
150. S. Y. Pusatcioglu, A. L. Fricke, and J. C. Hassler, J, Appl,
Polym, Sci, , 24: 937 (1979).
Injection Molding of Rubber Compounds 479

151. G. A. Prentice and M. C. Williams, Rubber Chem, Technol,,


53:1023(1980).
152. L. J. Lee and C. W. Macosko, ini. J. Heat Mass Transfer, 23:
1479 (1980).
153. J. M. Castro and C. W. Macosko, AIChE J ., 28: 250 (1982).
154. E. Broyer and C. W. Macosko, AIChE J., 22: 268 (1976).
155. L. T. Manzione, SPE Tech, Papers, 27: 338 (1981).
156. H. A. Nied, SPE Tech, Papers, 27: 344 (1981).
157. H. P. Schlanger, Rub, Chem, Technol,, 56: 304 (1983).
158. A. I. Isayev, Polym, Eng, Sci,, 23: 271 (1983).
159. A. I. Isayev and D. L. Crouthamel, Polymer Plast, Technol,
Eng,, 22: 177 (1984).
160. W. A. Gurney and V. E. Gough, Rubber Chem, Technol,, 20:
863(1947).
161. C. M. Blow, K. B. Demirli, and D. W. Southwart, J, Inst,
Rubber Ind,, 8: 244 (1974).
162. W. V. Chang, P. H. Yang, and R. Salovey, Rubber Chem,
Technol,, 54: 449 (1981).
163. B. C. Tsai, Rubber Chem, Technol,, 51: 26 (1978).
164. W. W. Paris, J. R. Dillhoefer, and W. C. Woods, Rubber Chem,
Technol,, 55: 496 (1982).
Compression Molding of Polymers
and Composites

C harles L . T u c k e r III
Department of Mechanical and Industrial Engineering
University of Illinois at Urbana-Champaign
Urbana^ Illinois

I. IN T R O D U C T IO N

A. Technological S ig n ifican ce o f Compression Molding

Compression molding is one of the oldest techniques for processing


polymers, though it has not received a great deal of study on the
fundamental level. For many years the standard technique for mold­
ing phenolics and similar thermosets, it has to some extent been re­
placed by thermoset injection molding. By comparison, injection mold­
ing offers advantages in materials handling and ease of automation.
However, compression molding retains a distinct advantage when pro­
cessing reinforced polymers. Compression molding flows involve mod­
est amounts of deformation, and there are no regions of very high
stress, such as at the gate of an injection mold. Consequently, re­
inforcing fibers are not damaged by the flow during mold filling as
often happens in injection molding. Thus, higher concentrations of
reinforcing fibers and longer fibers can be included in compression-
molded materials. Recent interest in polymer matrix composites for
strong, lightweight structures has renewed interest in compression
molding. Many developments in compression molding are concerned,
either directly or indirectly, with the processing of composite ma­
terials .
A wide variety of composite materials are compression molded.
Composite materials with thermoset matrices and short reinforcing
fibers, such as sheet molding compound (SMC), are particularly in­
teresting because the properties of the finished products are so
strongly affected by processing. Other compression-molded materials

481
482 T ucker

include thermoset matrices with continuous aligned reinforcing fibers


(^’high-performance” composites, typical of aerospace applications)
and thermoplastic matrices with continuous random fiber reinforce­
ment ( ’’stamping” materials). Traditional laminates, usually with
woven cloth reinforcement, are also compression molded.

B. C ritic a l Issues

Recent work in compression molding has focused on the critical issues


of mechanical property control, surface finish, cycle time, mold de­
sign, and process automation.
Mechanical property control is a problem for short-fiber-reinforced
polymers like sheet molding compound. Properties such as elastic mod­
ulus and tensile strength can vary greatly even in carefully controlled
parts made in a laboratory environment. Denton [1] has shown that
the tensile strength and elastic modulus of SMC samples have a stand­
ard deviation equal to about 50% of their mean values. In effect, this
means that parts have to be designed as if the material is half as
strong and stiff as it really is. There are several probable causes
for this variability. In part, it is due to the inherent inhomogeneity of
the material; some spots simply have more reinforcement than others.
This could be quantified by repeating the tests with different size
samples. For instance, tensile specimens which are 50 mm wide and
250 mm long should have a lower variation in strength than samples
25 mm wide and 100 mm long. Another source of property variation
is orientation of the fibers. It is well known that flow during mold­
ing can orient fibers and that fiber orientation affects mechanical
properties. Only recently have researchers tried to quantify this
effect. Since fiber orientation tends to produce a variation in prop­
erties which depends on location within the part, but is consistent
from part to part, it must be understood in order to separate it from
the inherent variability in the material. The overall goal in this area
is to develop the ability to produce parts with reliable, repeatable
mechanical properties.
A second critical issue is the surface finish of parts which con­
tribute to the appearance of a product. Automotive exterior body
parts such as hoods and doors are excellent examples of products
where surface finish is important. The glass-fiber-reinforced poly­
ester SMC which is used for these parts develops porosity, surface
waviness, and sink marks—shallow depressions on the smooth appear­
ance side corresponding to ribs and bosses on the back side. Ap­
pearance problems have been attacked by altering the composition of
the polymer matrix (leading to low-profile re sin s), by combining lay­
ers of long and short glass fiber reinforcements, and by adding a fi­
ber-free coating to the part (in-mold coating).
Compression Molding of Polymers 483

A third major concern is cycle time—the time it takes to produce a


part from a single mold and press. Cycle time controls the cost of
tooling and production equipment and is particularly important in
high-volume applications. Considerable empirical effort has been ex­
pended to reduce cycle time. Consequently, it is important to un­
derstand compression molding from the viewpoing of fundamentals,
so that fruitful avenues for experimentation can be found and so
that no effort is wasted seeking improvements where none are avail­
able .
Mold and part design is a topic closely related to the preceding is­
sues, since it controls mechanical properties, cycle time, and possibly,
surface finish. For example, thermal design of the mold--the layout
and size of heating channels--determines the minimum cycle time
available. Construction of molds is expensive and time consuming,
so any analytical tools which can show that a mold design will work
before the tool has been built are extremely useful.
The last critical issue is process automation—the reduction of the
amount of human effort required to produce parts. This subject cen­
ters around devices which load and unload the mold, since the opera­
tion of the press is relatively easy to automate. The goal here is to
reduce labor content so as to be more competitive with processes like
injection molding, which handle material in bulk.

C. Scope o f T h is C h a p te r

All of the critical issues (with the possible exception of process auto­
mation) are related to the flow of molding compound, transfer of heat,
and chemical reaction which occur in the mold. Flow causes fiber or­
ientation, which controls mechanical properties. Heat transfer inter­
acts with curing to determine cycle times. The design of the mold and
its heating system influences flow and heat transfer, and variations in
mechanical properties and curing influence surface finish. Hence, the
fundamental study of flow, reaction, and heat transfer in the mold are
discussed extensively in this chapter.
Certain fairly simple models for mold-filling flow have proven use­
ful for thin compression-molded parts, so they are discussed first.
Next, heat transfer and curing are considered, followed by a discus­
sion of some more complicated rheological effects. The problem of fi­
ber orientation by flow and its quantitative description is then con­
sidered .

D. Stages in th e Compression Molding Process

For the sake of clarity and convenience, the stages of compression


molding are defined as follows:
484 T ucker

1. Material preparation is the stage in which the molding com­


pound or prepreg is made. This may involve compounding a
resin, combining a resin with fillers or fibers, or impregnating
a reinforcing cloth or fibers with a resin. Subsequent thick­
ening or B-staging of the resin is also part of this phase. Ma­
terial preparation often controls the rheological properties of
the molding material and, in reinforced polymers, the bonding
between fibers and resin.
2. Prefill heating is often carried out to speed processing. With
some thermosets this is done outside the mold, using dielectric
heating. With SMC some preheating may take place after the
charge of molding compound is placed in the mold, but before
the mold is closed and flow begins. The temperature field
which results from this stage is part of the initial condition
for the mold-filling stage.
3. Mold filling begins when the polymer begins to flow, and ends
when the mold is full. The amount of flow in compression mold­
ing may be small, but it is critical to the quality of the part.
This flow controls the orientation of the reinforcing fibers in
short-fiber-reinforced materials, so it has a direct effect on
the mechanical properties of the part. Even in unreinforced
materials, the flow plays an important role in heat transfer
and so controls the curing of the part. In some processes,
especially those involving lamination, there is essentially no
flow; the initial charge completely fills the mold. For these
materials there is no mold-filling stage.
4. In-mold curing describes the stage that follows mold filling,
where the part cures in the mold. Some curing takes place
during mold filling, whereas the final stages of curing may be
completed during a ’^post-cure” heating cycle after the part is
removed from the mold. Most often in-mold curing must con­
vert the polymer from a liquid (which can flow to fiU the mold)
into a solid (which is strong enough to be removed from the
mold). Considerable heat transfer by conduction also takes
place during this stage, and the interaction between heat trans­
fer and curing is important to understand.
5. Part removal and cool-down is the final phase. It is seldom
analyzed, but is included here because it plays a role in dis­
tortion of the parts and the development of residual stresses.
One source of residual stress is the difference in thermal ex­
pansion between different portions of the part. If the part
is stress free at the molding temperature where it was first
formed, then residual stresses and distortions will arise as it
cools down to room temperature. The temperature distribution
and rate of cooling are important in determining how these
stresses may relax in a viscoelastic polymer.
Compression Molding of Polymers 485

Fig. 1 Example of thin part which can be treated by the generalized


Hele-Shaw flow model.

II. FLOW MODELS FOR T H IN C A V IT IE S

A. A pproxim ations fo r T h in P arts

This section covers models for the mold-filling flow in parts which are
thin compared to their lateral dimensions, such as the part shown in
Fig. 1. Thin parts comprise a large percentage of all compression-
molded parts, since they can be cooled or cured quickly to get low
cycle times. Mold-filling flow in thick parts is discussed in Section
IV. A.
The goal of this type of modeling is to predict the mold-filling pat­
tern , starting with the initial charge shape and finishing with the full
mold. Useful results include knit-line locations, velocity distributions,
and pressure distributions. An understanding of the filling pattern
is also a prerequisite to understanding heat transfer and curing.
The first step in modeling a thin part is to use the lay-flat approx­
imation. The part is conceptually ’’unfolded” so that it lies in the x -y
plane, with z being the thickness direction. (Note the local coordi­
nates attached to the part in Fig. 1.) This is similar to making a pat­
tern for a part to be made from sheet metal. This approximation trans­
forms the three-dimensional shape into an equivalent flat part, though
the geometry in the x -y plane may be complex.
At this stage we assume negligible inertia (because of the high vis­
cosity of the molding compound) and adopt an inelastic, isotropic model
for the rheology of the molding compound. In fact, most molding com­
pounds are viscoelastic to some extent, and those with fibrous rein­
forcement are anisotropic. However, for many cases one can still
486 T ucker

Flow Front Charge Mold Boundary

Fig. 2 Coordinate system and nomenclature for compression mold-fill-


ing analysis.

predict the mold-filling pattern quite accurately by neglecting these*


Also, since discussion of the heat transfer problem is postponed to
Section III, for the present we assume that the temperature field is
known. The only contribution that the temperature field makes to
the flow problem is to cause the viscosity of the molding compound to
vary with position, so for the moment we take viscosity to be a known
function of X , y , and z.
As shown in Fig. 2, flow in the x -y plane is caused by the squeez­
ing motion of the upper half of the mold, which moves downward with
velocity s. The instantaneous thickness of the part is h (x , y ), and
the velocity components of the fluid are denoted by u, v, and w in
the X , y, and z directions, respectively. Taking L as the character­
istic dimension of the part in the x -y plane, an order of magnitude
analysis shows that

w s ( 1)

L
U Sr- (2)
h
L
V Sr- (3)
h

where the symbol means ^’has the order of magnitude o f .” These


relations will hold except near the flow front of the charge (the ’’foun­
tain flow” region ). As long as the part is thin (h << L ) , the flow may
be treated as being approximately planar, and the velocity in the gap-
wise direction w can be neglected. Two useful quantities are the ve­
locity components averaged across the part thickness, u and v. They
are defined as
Compression Molding of Polymers 487

1 n
u{x,y)=-J u (x ,y ,z ) dz (4)

and

1 n
v (x , y ) = r- i v (x ,y ,z ) dz (5)

The average velocities must obey a mass balance equation, which


can be derived by integrating the continuity equation in z from the
lower to the upper mold surface. The result (assuming that h is a
continuous, differentiable function of x and y ) is

4 (uh) + ^ (vh ) = s ( 6)

(Note that s = -3h/3t.) Equation (6) will be referred to here as the


continuity equation, and it must be satisfied by any solution for u
and V .
Under the assumptions listed so far, the equations of motion re­
duce to

3T 3T 3t
XX yx (7)
8x 3x ay 3z

3T 3T 3t
1? xy
3X
yy
ay 3z
zy ( 8)
ay

From this point there are two ways to simplify the problem, which
lead to two very different models. First, one can adopt a lubrica­
tion-type model in which the right sides of Eqs. (7) and (8) are
dominated by the shear stress terms 3i2x/az and 3Tzy/3z. This
leads to the generalized Hele-Shaw model, which is discussed in the
next section. Or one can assume that a thin lubricating layer near
the mold wall prevents significant shear stresses from arising so
that the dominant stresses are xxx» ^^xy» leads to
the lubricated squeezing flow model, discussed in Section II.C .
These models represent two extreme limiting cases of thin compres­
sion molding flows. As such, they bracket other, more general so­
lutions .
488 T ucker

B. G eneralized H ele-Shaw Flow Model

Governing Equations
If a no-slip boundary condition exists at the upper and lower surfaces
of the mold, then an order of magnitude analysis shows that the ve­
locity gradients in the z direction are far larger than in the x or y
directions:

9u c
(9)
9x L

u
3U c
(10)
3z IT

with Uq being a characteristic value of u. For an inelastic fluid in


which the viscosity n does not vary much with z, the terms in the
equation of motion take on magnitudes like

3T nu
XX
3x ( 11 )

3t nu
( 12)
3z

so that terms of the former type can be neglected, reducing the equa­
tions of motion to

3T
3P zx
(13)
3x 3z

- lln (14)
3y 3z

Hieber and Shen [2] have shown that even in a non-Newtonian, non-
isothermal case, this type of flow has the character of a Hele-Shaw
flow [3 ]. That is, velocity is a vector function of x, y , and z, but
at any (x ,y ) location the direction of the velocity vector is not a func­
tion of z. This offers the possibility for enormous simplification to the
problem, since one can track ü and v as a function of x and y, making
the mathematical problem two dimensional.
Compression Molding of Polymers 489

Following Hieber and Shen, the average velocities are

- _ S 8P
(15)
^ ' h 3x

S ^
V = (16)
h ay

where S is a measure of the ease with which fluid flows locally, and
is given by

h
(z - X)^ dz
(17)
/0
Here, n = n(z) is the viscosity, and X is the value of z at which the
shear stresses tzx and izy vanish. For problems in which the ther­
mal boundary conditions are the same at the upper and lower mold
surfaces, X = h/2. These results may be substituted into the con­
tinuity equation, Eq. (6 ), to give a single governing equation for
the pressure distribution [4,5]

-^(s
ax\
—)
ax/
+ -^/s
ay\
— ^
ay/
(18)

This equation may be solved for P (x ,y ), subject to the boundary con­


ditions that P = 0 on the flow front, and aP/an = 0 when the charge
contacts the mold boundary . This latter condition is equivalent to re­
quiring that there be no flow across the boundary. This formulation
is identical to Hieber and Shen^s formulation for injection molding in
thin cavities, except that the squeezing motion of the mold, s, appears
from the continuity condition.
Once the pressure distribution has been found by solving Eq. (18),
the entire velocity distribution may be found from

(z - X) dz
u (x ,y ,z ) = (19)
ax

aP r (z - X) dz
v (x , y ,z ) = — (x ,y ) J (20)
490 T ucker

These equations may also be used to find X when it is not known a


priori from some symmetry condition, by meeting the no-slip bound­
ary condition at the upper plate (u and v = 0 at z = h ) . This leads to

( 21 )
r\
-'o V-^0

A special case of interest is an isothermal Newtonian fluid in a mold


where the thickness h is uniform in x and y at any given time. Then

hf
u = ( 22 )
12y m

(23)
12y\ 3y/

and Eq. (18) becomes Poisson’s equation

^ ^ -12ys
(24)
h"

The governing equations, Eq. (18) or, where applicable, Eq. (24),
apply at any instant of time and determine the pressures and veloci­
ties at that instant. Thus, the problem is quasi-static in nature. The
actual solution is complicated because the flow front, where a bound­
ary condition is applied, is moving. This introduces the time depen­
dence of pressure and velocity.

Lim ita tio n s

The generalized Hele-Shaw model is restricted to parts for which


h « L; however, other restrictions apply as well.
First, the assumptions of negligible w and negligible in-plane stress­
es are not valid near the edges of the part. Near the flow front
there is presumably some type of fountain flow, where w is compara­
ble to u or V . Also , where the charge contacts the edges of the mold
there is a no-slip boundary condition along the vertical edge, leading
to significant values of 9iyx/3y and/or 3Txy/9x. Both of these e f­
fects are confined to a narrow border around the edge of the charge,
which is only about as wide as h. The Hele-Shaw model sacrifices ac­
curacy in this small region to gain tremendous simplification when solv­
ing for the remainder of the flow. As a result, the calculated veloci­
ties, u and V , slip along the boundaries of the mold in the x -y plane.
Compression Molding of Polymers 491

Second, the Hele-Shaw model will only be valid when stress terms
like 3t2x /3z are large compared to stress terms like 9txx /9x . The
shear stresses are largest near the mold walls, while the extensional
stresses are largest near the midplane of the charge. In a noniso-
thermal flow the viscosities in these two regions can be quite differ­
ent. Using nw ^c represent the viscosities near the wall and
near the center of the charge, respectively, the orders of magnitude
of the stress gradients are

rv2 n U
3^u w c
^ n — 0 — 5— (25)
3x 'w

and

3T n u
c c
(26)
3x ^c 3x^

For the Hele-Shaw model to be valid, we must have

» 1 (27)

This is never a problem in injection molding, since a hot polymer is


flowing into a cold mold, and nw > ^c* But in thermoset compression
molding a cold polymer is molded in a hot mold, and nw < Tic* The
applicability of the Hele-Shaw model is then determined by the value
of the dimensionless group defined in Eq. (27). Although all ther­
moset compression molding flows have a low viscosity at the wall and
a high viscosity at the midplane, it is always possible for L to be
large enough and h small enough that the generalized Hele-Shaw model
is valid. Certainly one would expect this to be true of large, thin
parts like the hood of an automobile, for which L is about 2 m and
h roughly 2 mm.

Numerical Implementation
The geometry of real compression molding problems is complex enough
that the governing equations must be solved numerically. However,
solving Eqs . (18) or (24) (usually by finite differences or finite ele­
ments) is only half the task. Some procedure must also be chosen
for keeping track of the moving flow front and the shape of the charge
and for modifying the boundary conditions as portions of the charge
front encounter the boundaries of the mold.
492 T ucker

Fig. 3 Charge shapes calculated so the flow front reaches the edges
of a square mold everywhere at once. (One-quarter of the charges
and mold are shown.) (Solid lines from Ref. 5, dashed line from
Ref. 6.)

Silva-Nieto, Fisher, and Birley [6] solved the isothermal Newton­


ian problem [Eq. (24)] using finite differences. They reported cal­
culations for square and trapezoidal molds. Rather than modify the
boundary conditions on the flow front as it encountered the mold waU,
they started with the final shape of the mold, applied a boundary con­
dition of P = 0, and ran the simulation backwards in time. The result
of this calculation is an initial charge shape which reaches the bound­
aries of the mold everywhere at the same instant. A sample result for
a square mold is shown in Fig. 3. One reason for using such a charge
is to avoid the formation of knit lines (where two flow fronts meet).
However, there is no particular reason to believe that this is the best
charge to use for a given part.
A finite element implementation has been developed by Lee, Folgar,
and Tucker [4 ,5 ]. They performed both isothermal Newtonian and
isothermal power-law calculations. Six-node triangular elements with
straight sides were used, since a quadratic interpolation function is de­
sired when the governing equation is second order. Their calculation
Compression Molding of Polymers 493

starts with the definition of a mesh which covers the initial charge
shape. Finite element equations are solved to find the pressure dis­
tribution, from which the average velocities, u and v, can be calcu­
lated at any point. The flow front is then advanced by moving each
vertex node on the flow front a distance uAt in the x direction and
vAt in the y direction. The flow front is approximated by a series of
straight lines connecting the vertex nodes. A new mesh, which cov­
ers this new shape, is then automatically generated by stretching the
old mesh as if it had been drawn on a rubber sheet, and new pres­
sure and velocity distributions are found for the next time step. This
stepping process is repeated until the mold has been filled and the
simulation is complete. Figures 4(a) through 4(d) show steps dur­
ing a typical simulation. The successive positions of the flow front
are summarized in Fig. 5.
As the flow front encounters the mold boundary, boundary condi­
tions are modified automatically. This is done by reducing the time
step, if necessary, so that only one vertex node on the flow front
reaches the mold boundary during any given time step. The bound­
ary conditions on that node are altered before the calculation for the
next time step. This procedure is costly in terms of computation time,
since many extra time steps must be taken. However, it makes the
simulation run automatically without operator intervention.
The mesh stretching scheme used by Lee et al. [5] is a key ingre­
dient in an automatic simulation, but it is only suitable for simple
geometries. The mesh cannot stretch into a complicated shape with­
out becoming too distorted to give accurate results. One alternative
has been developed by Brown [7 ]. This method begins by covering
the entire mold with a finite element mesh, using eight-node isopara­
metric quadrilateral elements. The initial charge shape is described
separately by specifying the location of its boundary. The program
proceeds by time-stepping as before, but only those elements con­
taining part of the charge are used. Thus, elements are either
^’empty" or ’’fu ll,” with the proportion of full elements increasing as
the simulation proceeds. Partially filled elements are temporarily dis­
torted during any given time step to make their boundaries coincide
with the flow front. This method has shown some promise. A com­
parison between predicted and experimental results for a simple mold
with a knit line is shown in Fig. 6. Developments are moving in the
direction of a simulation technique which can handle any part shape
and will run without operator intervention.

Comparison with Experiments


The generalized Hele-Shaw model shows excellent agreement with ex­
perimental flow front shapes, both for Newtonian and non-Newtonian
fluids. Figures 6 and 7 show comparisons between experimental
494 T ucker

Fig. 4 Successive mesh shapes using mesh stretching technique for


finite element simulation.
Compression Molding of Polymers t»95

Fig. 5 Calculated flow front positions for rectangular charge filling


rectangular mold.

shapes for silicone oil (Newtonian) in a rectangular mold and the


predictions of Lee, Folgar, and Tucker’s finite element model [4 ,5 ].
The excellent agreement substantiates the model and the calculation
method.
The simple isothermal Newtonian model also does an excellent job
of predicting flow front shapes for non-Newtonian molding compounds,
provided that the parts are thin enough to make the Hele-Shaw model
valid [Eq. (2 7 )]. Example comparisons between experiments with
SMC and the finite element calculations are shown in Figs. 8 and 9.
Figure 10 shows what happens in a charge for which the Hele-Shaw
model is not valid. This is the same experiment as Fig. 9 but with a
charge which is three times thicker, and the result is a much dif­
ferent set of flow fronts.
The fact that a Newtonian model so accurately predicts the features
of a non-Newtonian flow is quite remarkable. In fact, for isothermal
flow in a flat gap [ i . e . , h = h (t) o n ly ], the flow fronts predicted by
496 T ucker

Fig, 6 Calculated and experimental flow front positions for filling mold
with obstacle (T . A. Osswald and C. L. Tucker, unpublished research
resu lts).

Newtonian and power-law calculations are nearly identical (Fig. 11),


though the pressure distributions are quite different. Silva-Nieto
et al. [6] report pressure measurements in SMC molding which gen­
erally agree with the Newtonian predictions.
Compression Molding of Polymers 497

F ig . 7 Comparison of predicted and experimental (dashed) flow front


shapes for silicone oil. (From Ref. 59.) This experiment also pro­
duced the fiber orientation distributions in Fig. 39.

C. L u b ricated S queezing Flow Model

An alternative formulation for thin parts is to model the flow as a lu­


bricated squeezing flow. This model is based on the assumption that
there is no shearing across the thickness of the charge and that the
charge slips along the upper and lower surfaces of the mold.
498 T ucker

Successive Predicted Shapes

------ Actual Molded


Shape

Initial

/'777777777777777777T//
Fig, 8 Comparison of predicted and experimental flow front shapes for
triangular charge of SMC.

Barone and Caulk [8] have suggested that this model is appropriate
for SMC on the grounds that the network of fibers inhibits shearing
across the sheet thickness. Another argument for this model would
be that a thin layer of resin near the surface serves as a lubricant
for the rest of the charge, absorbing the apparent slip velocity while
transmitting only a very small shear stress. The lubricating layer
could arise because the rigid mold surface constrains the arrangement
of nearby fibers, providing a resin-rich region with lower viscosity,
and/or because the mold has heated a thin film of molding compound
on the surfaces of the charge.
With the assumption that Bu/9z = 9v/8z = 0, it is clear that u = u
and V = V . The equations of motion reduce to

<>-D ____ ^
3P
9x 9x
^ __ yx
9y
( 28)
and
9t 9t
^ ^ xy __ yy (29)
9y 9x 9y
499

------------------ Prediction
------------------ Experiment

Fig. 9 Comparison of Hele-Shaw predictions and experimental flow front


shapes for thick charge of SMC. The charge is 75 mm x 150 mm x 3 mm
thick.

with the boundary conditions that velocity normal to a mold boundary


is zero, and total stress normal to the free flow front is zero.
For many simple flows the solution is trivial, resulting in either b i­
axial extensional flow or pure shear flow. In either case pressure is
constant over the entire charge. As an example. Fig. 12.shows the
flow fronts predicted by this model for a rectangular charge in a rec­
tangular mold. Initially the flow is a biaxial extension, with the charge
stretching equally in the x and y directions. The charge shape is geo­
metrically similar to the original shape and expanded in size. Once the
right-hand edge of the charge encounters the mold boundary, only
500 T ucker

Experiment

Fig. 10 Comparison of Hele-Shaw predictions and experimental flow


front shapes for thick charge of SMC. The charge is 75 mm x 150
mm X 9 mm thick.

flow in the y direction is possible. The charge is now in pure shear


flow (i.e ., two-dimensional extensional flow in the y-z plane) and is
being stretched uniformly in the y direction. These flow fronts con­
trast sharply with the predictions of the generalized Hele-Shaw model
in Fig. 5. The behavior of the thick SMC charge in Fig. 10 is very
close to what the lubricated squeezing flow model would predict. This
points out the importance of the model: it represents an extreme of
Compression Molding of Polymers 501

Fig. 11 Predicted flow fronts for Newtonian and power-law fluids with
isothermal Hele-Shaw model.

behavior which is opposite to the generalized Hele-Shaw flow. Thus,


in a sense the two models provide upper and lower bounds on the ac­
tual results.
Lubricated squeezing flow is often thought of as providing biaxial
extension in the plane of the sheet, but this is not generally true.
For example, the charge and mold configuration shown in Fig. 13
clearly produces a flow which is neither biaxial extension nor pure
shear. No problems of this type have been solved, but the formula­
tion is straightforward. For example, consider isothermal Newtonian
flow. The equations of motion reduce to

3P 1 3^U'
(30)
ay".
and1

3P 3V
r ^ + (31)
3y ^ L 3x^
502 T ucker

F ig . 12 Flow fronts of rectangular charge in rectangular mold for lu­


bricated squeezing flow.

and the continuity equation becomes

(32)
3x 8y h

provided that h does not vary with x or y at fixed time. This problem
is two-dimensional, just as the Hele-Shaw problem, but the equations
cannot be integrated to eliminate any of the variables. One is left with
a two-dimensional problem with three variables, u, v, and P. This is
Compression Molding of Polymers 503

Fig. 13 Example of charge and mold configuration for which succes­


sive flow positions in lubricated squeezing flow are not obvious.

still a considerable simplification over the general problem, which is


three-dimensional in four variables. Solution of Eqs. (30), (31), and
(32) by finite elements would be straightforward, and the same tech­
niques of flow front advancement and automatic meshing that have
been applied to the Hele-Shaw problem could be used here as well.
One boundary condition on Eqs. (30) —(32) is that velocity normal
to the boundary is zero where the charge touches the mold boundary.
On the free flow front the boundary condition is that total stress nor­
mal to the boundary, ann> is zero. Using Vn to denote the velocity
component normal to the boundary and n to denote the normal direc­
tion, we obtain

a = 2y - P = 0 (33)
nn dn

or

8v
__ n
2y (34)
an
504 T ucker

on the free flow front. Note that P is the pressure inside the charge,
which in this problem does not equal the pressure outside.

D. C o n tro llin g Flow in th e Mold

Having seen how one might predict the flow of molding compound dur­
ing mold filling, one might reasonably ask, ^’What variables are avail­
able to control the flow .’’ The modeling work shows that the primary
factor determining mold flow is the shape and placement of the original
charge. Once this has been chosen, other factors have a secondary
influence on the flow. Experiments by Herman [9] on the influence of
processing variables on the properties of SMC support this idea.
The only other variable which is likely to have a major influence on
the mold-filling flow is the extent to which the press holds the mold
halves parallel. The pressures generated during squeezing flow are
often large and give rise to the large forces which a compression mold­
ing press must provide. It has become apparent that even a massive
press has enough compliance to permit the mold to cock slightly under
the influence of an off-center force. Even a few thousandths of an
inch can be significant for the mold-filling flow, but a typical mold
which is cocked at an angle of only 1 min will have a difference in
gap height of several millimeters between the two ends. Note from
Eq. (22) that average velocities vary as the square of the gap height,
so that small errors in h become much larger errors in velocity.
To counteract this problem, press designers have taken the con­
ventional arrangement [Fig. 14(a)] and added position sensors and
individually controlled short-stroke hydraulic cylinders at each cor­
ner of the press [Fig. 1 4 (b )], When the mold closes, the upper
platen contacts the corner cylinders, which are controlled to main­
tain a parallel mold and which can also control the closing speed of the
mold or the total force applied to the charge. Newer model presses,
designed from the beginning with this capability in mind, have al­
ready appeared [10]. The effect of this development has been to
make mold-fillmg flow more consistent from part to part, thereby im­
proving the repeatability of the process.

III. HEAT T R A N S FE R AND C U R IN G

Most thermosetting polymers processed by compression molding cure


by a thermally activated reaction (as opposed to a mixing activated
cure, like urethane RIM). Also, the curing reaction is exothermic.
The combination of these two features gives heat transfer a double
role: Heat from the mold must be conducted into the polymer to be­
gin the cure, but heat from the curing reaction must be conducted
Compression Molding of Polymers 505

(a)

(b)

Fig. 14 Press arrangements for compression molding, (a ) traditional


press; (b ) press with short-stroke cylinders and feedback control to
maintain parallelism and control motion during filling.
506 T ucker

out before it damages the part. Many thermo set-molding compounds


liberate enough heat to ^’burn themselves up” if curing is not man­
aged properly. A slow, careful cure avoids this problem entirely,
but does not produce parts fast enough to be economical. Instead,
real production situations operate, or would like to operate, close to
the limit on thermal degradation so as to maximize production rate.
A clear understanding of how parts cure and how heat transfer in­
teracts with curing is critical to the rapid production of high-quality
parts.
Emphasis here will be placed on materials which cure by addition
polymerization, such as unsaturated polyesters. (Polymers which
cure by condensation reactions are discussed in the chapter on re ­
action injection molding.) Addition-curing polymers tend to under­
go a much sharper transition from liquid to gel during cure than do
condensation-curing polymers, so resins are typically formulated to
avoid any curing during mold filling. In unsaturated polyesters this
is done by adding an inhibitor, a small molecule which reacts prefer­
entially with the free-radical initiator, avoiding the initiation of any
chains until the inhibitor is consumed. This makes it possible to con­
sider heat transfer without curing during the mold-filling stage and
then separately examine heat transfer and curing without flow during
the closed curing stage. Following this pattern, heat transfer during
mold filling is discussed first. Then, after reviewing reaction kinet­
ics for polyesters and similar materials, the curing problem is consid­
ered. This section concludes with a discussion of the thermal design
of molds and a brief consideration of residual stress problems.

A. Heat T ra n s fe r D u rin g Mold F illin g

The basic heat transfer problem during mold filling concerns a charge,
initially at T q, placed in a hot mold. Since the thermal conductivity of
the mold is much greater than that of the polymer, it is customary to
assume that the mold wall temperature is a constant, T ^ The prob­
lem is to find the temperature history in the charge as it flows to form
the part. In what follows, we assume that the velocity field for mold
filling is known, though in a realistic case the two problems must be
solved simultaneously. For thin parts, order of magnitude analysis
shows that conduction in the x and y directions is much smaller than
conduction in the thickness (z ) direction. If viscous dissipation is
negligible, the energy equation becomes

/3T 8T 3T 3T^ , 3^
+ u — + V -— + w k —2 (35)
3x 3y 3z j 3z^

The initial condition is that T = T q everywhere in the charge, and the


boundary conditions are that T = T ^ at z = 0 and z = h (t ). This
Compression Molding of Polymers 507

continuum model can also be applied to molding compounds with fibers


and filler, provided that one is willing to ignore details on the order
of the fiber or particle diameter in size. Materials like SMC do have
anisotropic thermal conductivity because of the arrangement of the fi­
bers [8 ]. However, the z axis is an axis of material symmetry so the
problem is easily resolved by making k in Eq. (35) the conductivity in
the thickness direction.
The solution of Eq. (35) by numerical techniques could be carried
out in conjunction with a mold-filling flow simulation, but few such
calculations have been performed to date. The special case of lubri­
cated squeezing flow has been solved analytically by Barone and Caulk
18], and their results provide considerable insight.
For lubricated squeezing flow the velocity components u or v do not
vary with z:

— = IZ. = 0 (36)
8z 3z

Differentiating the equation of continuity of an incompressible fluid


with respect to z and substituting Eq. (36) show that

(37)

Satisfying the boundary conditions that w = 0 at z = 0 and w = -s at


z = h yields

3W -s
(38)
8z “h

w (39)
h

That is, the motion in the z direction is a uniform compressive strain­


ing, since 8w/3z is not a function of z. If one were to draw vertical
lines in the initial charge, during molding these lines would move about
the mold always remaining straight and vertical along their length as
the mold closed. With this in mind, we can define a new coordinate
Cby

z
(40)
hiT)

Although fluid particles change z coordinates during mold filling, they


remain at the same position c throughout the flow. Thus, c is a co­
deforming coordinate. Since there is no x -y variation in initial or
508 T ucker

boundary conditions, no temperature gradients arise in the x or y di­


rections. Consequently, the energy equation can be rewritten as

pc (41)
9t h (t )"

where the 3T/8t in Eq. (41) represents the time derivative at a fixed
position c* Equation (41) looks exactly like the classical conduction
problem of a slab initially at T q, exposed to a wall temperature of Tw
at t = 0, except that the slab is shrinking across its thickness with
time. The solution to Eq. (41) is [8]

n+1 -a ^ (t )a
( - 1) n
= 1 + 2 2 exp cos(a c) (42)
w - To n
n=0

where a is the thermal diffusivity of the molding compound, the co­


efficients an are

.(n + l ) (43)

and t (t ) is an equivalent time,

L
t (t ) = f h
f
dt'
h ^ (t ')
(44)
0

Here t^ is a dummy integration variable, and hf is the final thickness


of the charge. (The preceding results are for a part with uniform
thickness, but Barone and Caulk^s original development is general
enough to deal with a gap height h that varies slowly in x and y . )
The result in Eq. (42) is identical to the classical solution for a solid
slab when time is replaced by the equivalent time, t. That is, the
changing thickness of the sheet warps the time scale, accelerating
the conduction of heat as the sheet becomes thinner.
The temperature profiles given by Eq. (42) are not surprising—the
charge is warmed by conduction from the outside to the inside. One
noteworthy point is that this heat must be supplied from the mold, and
that much more heat must be supplied from places on the mold covered
by the initial charge than by the last parts of the mold to fill, as
shown in Fig. 15. This has important implications for the thermal de­
sign of molds, and will be discussed later.
No filling temperature histories for flows other than lubricated
squeezing flow have been calculated, but it is clear that changes in
Compression Molding of Polymers 509

F ig . 15 Profile of heat removed from the mold during filling. (From


Ref. 8.)

the flow can introduce some differences in heating and curing. The
most important difference is caused by the fountain effect at the flow
front. In a generalized Hele-Shaw flow the maximum velocity will be
at the midplane of the charge, and near the flow front cool material
from the center of the charge is convected outward to the walls [Fig.
1 6 (a )]. Consequently, the regions of the mold not covered by the or­
iginal charge are filled by cooler material than in the original charge
area. In other words, the cool center part of the charge flows out­
ward preferentially [Fig. 1 7 (b )].
510 T ucker

77777777777777777777777777
(a)

f
^ /7 7 7 7 7 7 7 7 7 7 7 7 /Z ////////
(b )

F ig . 16 Gapwise velocity profiles during compression molding, (a )


generalized Hele-Shaw flow; (b ) reverse fountain flow.

The opposite behavior has been observed in experiments by Marker


and Ford [11]. They saw the heated layers near the mold surface flow
outward preferentially while the cool inner layers remained nearly sta­
tionary [Fig. 1 6 (b )]. This is a different type of mold-filling flow, in­
termediate between the Hele-Shaw and lubricated squeezing flows. The
fluid mechanics of this type of flow will be discussed in Section IV .A .
With respect to cure, the preferential flow of hot material causes the
part to cure first at the edges. In contrast, a part filled by Hele-
Shaw flow would cure first at the center.
These examples point out the importance of the flow field in deter­
mining the temperature and cure pattern. Although heat transfer and
filling flows are usually independent of the later curing reaction, the
curing process depends strongly on the flow during mold filling.

B. K inetics o f th e C u rin g Reaction

Before dealing with heat transfer during curing, we consider the


kinetics of the curing reaction itself. Unsaturated polyesters and
Compression Molding of Polymers 511

(a)

(b)

(c)
Fig. 17 Effect of transverse velocity profile on filling pattern, (a)
transverse section of initial charge; (b ) final part when filled with
Hele-Shaw velocity profile; (c ) final part when filled with reverse
fountain velocity profile.

related materials cure by free-radical copolymerization between un­


saturated bonds in the short polyester chains and monomeric styrene
in the resin. The presence of more than one unsaturated site per
polyester chain leads to a network or cross-linked polymer. In ad­
dition to polyester and styrene, a resin contains a free-radical in­
itiator (such as ieri-butyl perbenzoate), thickening agents, and an
inhibitor. Decomposition of the initiator produces free radicals, which
react preferentially with the inhibitor when it is present. Once the
inhibitor has been consumed, each free radical initiates the growth of
a chain of styrene monomers and polyester chains, which grows until
it meets another growing chain or another free radical or until no re ­
actants are available.
The thickening agent (typically MgO or isocyanate) dramatically
increases the viscosity of the resin during a ^’maturation” phase,
which lasts from the time the compound is formulated until it is mold­
ed. Thickening is almost certainly relevant to the rheology of the
material, but it is outside the scope of the present discussion.
Reaction kinetics are often measured in a differential scanning cal­
orimeter, where the rate of heat released by the reaction can be mea­
sured under isothermal or scanning conditions. The amount of heat
liberated by the reaction is usually taken as a measure of the degree
512 T ucker

CM
E

F ig . 18 Isothermal curing behavior of unsaturated polyester resin.


(From Ref. 17.)

of cure, on the assumption that fractional conversion is proportional


to the number of bonds formed in linking chains together, and that
each bond releases the same amount of heat. This leads to a defini­
tion of the fraction conversion ( ’’degree of cure”) c as

Q
(45)

where Q is the amount of heat released up to the current time and Q»p
is the total heat of reaction. Thus c is a field quantity which convects
with the fluid and increases monotonically with time from zero (uncured)
towards unity (totally c u re d ).
Figure 18 shows typical cure-versus-time curves for isothermal cur­
ing of an unsaturated polyester [17]. The dwell time, an initial period
where no appreciable curing takes place, is due to the inhibitor and
can be adjusted by varying the inhibitor concentration. Curing then
begins slowly, accelerates, proceeds most rapidly around the halfway
point, slows down, and levels off when incompletely cured. Complete
curing does not occur at low temperatures because the developing net­
work structure has a glass-transition temperature (T g ) which increases
as the reaction proceeds. When Tg rises to the temperature of the
Compression Molding of Polymers 513

reacting material, then the reacted portion of the polymer becomes


glassy, inhibiting the molecular motions which bring reactants to­
gether and preventing further reaction [12]. The degree of cure
at equilibrium increases as the curing temperature increases. A cure
temperature above the glass temperature of the completely cured poly­
mer will give complete cure. In some polymers this temperature will
also cause some thermal degradation, complicating the interpretation
of experiments.
Simple nth-order kinetics do not describe this type of reaction,
since they cannot model a reaction rate which first increases and then
decreases with extent of reaction. An empirical model which has proved
useful was proposed by Kama! and Sourour [13,14]. Their equation
has the form

dc
(di + d2c"^)(l - c)^ (46)
dt

The dependence of reaction rate on temperature is incorporated by


allowing

-bi/RT
di = a^e (47)

-b2/RT
d 2 = a20 (48)

where ai, a 2 , b i, and b 2 are constants, T is absolute temperature, and


R is the gas constant. Note that the rate of heat liberated per unit
mass is

Q = Q (49)
^ ^ T dt

Since Eq. (46) is empirical, the six constants ai, a 2 , b i, b 2 , m, and n


are found by adjusting them to fit experimental data. Figure 19 shows
that the model fits experimental data quite well. Typical values of the
parameters are given in Table 1. Note that Eq. (46) predicts an asymp­
totic approach to complete cure for any curing temperature, in contrast
to the experiments. Apparently, this is not a serious problem in simu­
lating molding problems, probably because the typical temperatures
provide nearly complete cure.
A more detailed mechanistic model of thermoset cure kinetics has
been developed recently by Stevenson [15,16]. Rather than consider
’’cure” as a single scalar quantity, he included the kinetics of each in­
dividual part of the reaction, such as initiator decomposition, inhibi­
tor consumption, the various chain growth reactions, and the chain
514 T ucker

Time, min

Fig. 19 Kamal-Sourour kinetic model and Stevenson model fitted to


DSC data on unsaturated polyester resin (S . C. Claudin and C. L.
Tucker, unpublished research results).

termination reaction. Modeling each of these reactions with simple


kinetics and relating the different reactions with a mole balance re ­
sults in an overall model for the reaction. Stevenson^s full model is
quite complicated, but it offers considerable insight into the curing
reaction. Since most of the parameters can be measured or estimated
and are directly related to the properties of the various constituents

T a b le 1 Example Values of Kinetic Parameters


[Eqs. (46) “ (48)] for Unsaturated Polyester
Sheet Molding Compound [8]

a i = 4 .9 X 10^** s " ^ b i = 140 kJ*mor^

az = 6.2 X 10^ s"^ hz = 51 kJ-mor^

m = 1. 3 n = 2.7
Compression Molding of Polymers 515

of the resin ( i . e . , initiator, inhibitor, e tc.)» this model can be used


to explore the effects of different resin recipes on the curing reac­
tion .
For molding simulations a simplified version of Stevenson*^s model
[16a, 16b] has been used by Lee [17]. His model deals with a single
initiator (typical of SMC) and assumes that no monomer reacts until
enough radicals have been created by decomposing initiator to con­
sume all the inhibitor. The effect of this is to decouple inhibition
from polymerization, The induction time, t 2 » at which polymerization
begins is found from

qZo = 2fIo 1 - exp dt (50)

Z q and I q are the initial concentrations of inhibitor and initiator, q


and f are their respective efficiencies, and kd is the rate constant
for initiator decomposition, assumed to have an Arrhenius tempera­
ture dependence

-E^R/T
(51)

For most commercial materials the induction time is longer than the
cure time at processing temperatures.
After the inhibition phase ends at time tz, the remaining initiator
concentration I q is

lo = lo exp (52)

For t > Î 2 , polymerization proceeds according to

1- c 1 - e x p / -/ ^ k dt'j (53)
_ \ tz /J

Here, kp is the rate constant for polymerization,

-E R/T
k =A e P (54)
P P

Although this model requires more information than the Kamal-Sourour


model, it is based on the chemistry of the process. This means that
changes in the resin formulation can be directly reflected by changes
516 T ucker

in the parameters of the model, an ability which is extremely useful


when developing resin formulations.

C. Heat T ra n s fe r and C u rin g

Curing involves a balance between heat conduction (primarily across


the thickness of the part) and heat generated by the curing reaction.
The rate of heat generation per unit volume is pQx(9c/3t), so heat
transfer during the curing process is governed by an equation like

T r) 3c
(55)
3t

At the same time, the curing rate is controlled by the local tempera­
ture and degree of cure:

= f(c , T ) (56)
3t

Here, f(c , T ) represents any appropriate expression for the curing


rate, such as (46) or (53).
Since curing begins as soon as the molding compound becomes
warm, Eq. (56) should be integrated during the mold-filling process
as well as during closed curing. For lubricated squeezing flow, equiv­
alent time (44) proceeds at the same rate as real time once the mold is
closed. That is, 3t/3t = 1 once h = h f. The only effect of the closing
cycle is to offset the origin of the time axis [ 8] . This means that the
curing problem can be solved as if the sheet were introduced into the
mold at a uniform temperature T q and with a thickness hf. The only
effect of the lubricated squeezing flow is that the times in Eq. (52)
and (53) are related to real time by adding a constant. This offset
time is given by

tf h^ dt’
(57)
^offset ^f /0 h^(t')

where tf is chosen so that h = hf for t > t f .


Equations (55) and (56) have been solved by Barone and Caulk [8]
Lee [17], and Stevenson [16], using finite differences for (55) and a
numerical integration technique for (56). The results are most easily
understood if one first nondimensionalizes the equations. Since these
workers have considered a thermally symmetric case, it is convenient
to move the origin to the midplane of the part and define H = hf/2.
Appropriate dimensionless variables are
Compression Molding of Polymers 517

T - To
(58)
AT 0

t* = — (59)

(60)
H

where

A T q ~ T ij^ Tq (61)

is the operating temperature difference, and

pc
(62)

is a characteristic time for heat conduction. Note that c is a dimen­


sionless quantity, but that the reaction may occur either much faster
or much slower than heat conduction, so 3c/3t* will not be of order 1.
Instead, we define a characteristic time for the reaction, t^, as the
characteristic isothermal cure time at the wall temperature, T w The
dimensionless reaction rate is then defined as

1. ^ (63)
t 3t
r

which is of order 1. The value of tj* is based on the kinetic equation


being used. For example, with the Kamal-Sourour model [Eq. (46)]
Barone and Caulk suggest tr = l/d 2 . Also note that the adiabatic tem­
perature rise due to the heat of reaction is

Qt
ATr = (64)

Equation (55) can now be written in dimensionless form as

3T* 3^*
(65)
3t*

The dimensionless group Gr is known as the Damkohler number. Group


IV [16,16b, 18] , and is given by
518 T ucker

Fig. 20 Temperature and curing profiles of an unsaturated polyester


molding compound with Gr = 0.22. (From Ref. 16.)

tcAT^
Gr
t AT ac t (T - T )
( 66)
r 0 p r w 0

This represents the ratio of rates of temperature rise due to reaction


(ATr/tr) and conduction (ATo/tc). When Gr is small, conduction far
outweighs reaction, and the reaction has little effect on the tempera­
ture profile. When Gr is large, the reaction dominates the tempera­
ture history, and when Gr is close to unity, there is considerable in­
teraction between reaction and conduction.
These effects are shown in Figs. 20—22. Each figure gives cal­
culated profiles of temperature and degree of cure at various times.
These calculations use Stevenson^s kinetic model, with parameters and
operating temperatures characteristic of SMC. The part thickness is
varied far out of its usual range to vary Gr. Figure 20 shows pro­
files for a small Damkohler number. The temperature history is ap­
proximately the same as transient conduction in a slab, with a slight
Compression Molding of Polymers 519

C
o
o5
I.
o
>
c
o
Ü

75
c
o
+-»
Ü
CO

Fig. 21 Temperature and cure profiles of an unsaturated polyester


molding compound with Gr = 3.6. (From Ref. 16.)

excursion above near the end of the cycle. Curing takes place
nearly uniformly across the thickness of the part and does not begin
until late in the cycle.
A thicker part is shown in Fig. 21. Now there is a considerable
rise above the mold temperature in the center of the part. Curing
starts at the outside, then proceeds inward. In Fig. 22 a stül thick­
er part is shown, with Gr = 57. Now there is a very sharp tempera­
ture spike in the center of the part, and curing proceeds inward from
the surface like a wavefront. The parts shown in these two figures
would likely experience severe themal degradation if they were actu­
ally molded, due to the high centerline temperatures reached.
Stevenson [16] has shown that many features of the curing process
depend mainly on the Damkohler number. These include maximum tem­
perature and time to cure. For Gr « 1 the cure time is about 2tr (d e ­
pending on its definition) ; for Gr » 1 the ratio of cure time to t^ goes
520 T ucker

Fig. 22 Temperature and cure profiles for an unsaturated polyester


molding compound with Gr = 57. (From Ref. 16.)

up linearly with Gr. Barone and Caulk [8] also point out that as the
final part thickness increases (increasing Gr) thermal degradation
from the high maximum temperature becomes a problem. This must be
offset by reducing the mold temperature (reducing Gr by reducing
ATq and increasing ti*), a correction which is costly because it in­
creases the cycle time. Without some means of absorbing heat in the
part this is an unavoidable feature of thermoset molding.

D. Therm al Design o f Molds

All molds for thermoset compression moldings are heated, usually by


pumping steam or oil through channels drilled in the mold. Conven­
tional practice is to distribute the channels evenly throughout the
mold and to place them a uniform distance from the cavity surface,
but this is not the best design. Section III.A discussed the fact that
heat is removed from the mold nonuniformly during mold filling. More
Compression Molding of Polymers 521

i^SHEAT TRANSFERRED
' igS lN T O MOLD

-0.5

DISTANCE ALONG CAVITY SURFACE

Fig. 23 Net heat lost or absorbed by the mold during one cycle as a
function of position. (From Ref. 19.)

heat is removed from the area covered by the original charge than from
areas which fill later (Fig. 15). Additional heat is removed uniformly
from the mold at the start of closed curing; then heat liberated by the
curing reaction flows back into the mold uniformly. The result is that
some portions of the mold see a net loss of heat, and others see a net
gain (Fig. 23). If the heating channels are distributed uniformly in
the mold, then heat must be conducted within the mold body to bal­
ance the thermal loads. This imbalance causes spatial variations in
the mold surface temperatures, which can cause substantial problems.
The thermal time constant of a massive steel mold is much larger
than the length of one cycle, so the variable heat load from the mold­
ing compound looks to the mold like a high-frequency disturbance.
After many cycles the mold surface approaches a steady-state tem­
perature distribution. In this condition areas under the original
charge are cooler than average , and the last parts of the mold to fill
are hotter than average. The temperature variation is proportional
to L^/htcycle tends to increase as parts become larger and
thinner and are produced more rapidly. A typical SMC auto body
panel molded with a 60-s cycle time can have a 25°C variation in the
mold surface temperature with a conventional mold design [19]. This
522 T ucker

leads to a substantial variation in cure time along the part, reducing


productivity. A rule of thumb is that curing time increases 50% for
each 10°G drop in mold temperature [19]. Variations in mold surface
temperature may also cause surface waviness and warping.
The problem cannot be solved by simply increasing the overall tem­
perature of the mold, because this will cause the hotter spots in the
part to cure too quickly and/or experience thermal degradation. A
high mold temperature can also leave the pai’t too weak to be ejected
from the mold. The solution is to design the tìnold heating system to
compensate for the nonuniform heat load that is imposed.
This problem has been solved by Barone ahd Caulk [19] . Their
method starts by solving for the heat transferred to the part during
filling and curing, assuming a uniform mold surface temperature.
(Their calculations assume lubricated squeezing flow, but this is not
a requirement of the method. ) From this solution the thermal load
(heat removed or added per cycle) as a function of position on the
mold surface is found. One also includes heat lost by radiation and
convection, accumulated while the mold is open for part ejection and
loading the next charge. The steady-statè temperature field within
the body of the mold can then be solved as a straightforward heat
conduction problem by treating the heat to/from the part as constant
in time and variable in space. Barone and Caulk solve this by a
boundary integral technique and use the solution in an optimization
program which finds heating channel locations to minimize the varia­
tion in steady-state mold surface temperature.
Figure 24 shows a simple mold with a conventional heating design
and the attendant surface temperatures. The tulip-shaped tempera­
ture profiles are a consequence of the extra heat removed by the
charge in the center of the mold, a problem which becomes more pro­
nounced as the cycle time decreases. The overall downward trend in
temperature from left to right along the mold occurs because the heat­
ing channels are connected in series, so the oil temperature drops
steadily across the mold. The temperature variations for the 1-min
cycle are too large to produce an acceptable part : the center does
not cure, and the edges are too weak to demold. Some improvements
can be made by controlling each heating line individually, but the
best design is to optimize the location of the heating channels and
control the oil temperature in each channel independently. Figure
25 shows the optimized design and the resulting steady-state tem­
perature distribution. Note the dramatic improvement in mold tem­
perature profile for a 1-min cycle time compared to the conventional
design. The procedure is also conservative, in that flat temperature
profiles are easily maintained for cycle times longer than the design
cycle times.
Compression Molding of Polymers 523

CYCLE TIME (min) O IL TEMPERATURE(°C) FLOW RATE (//m in )


IN OUT

1 180 175.4 30.3


3 180 171.9 10.5
6 180 170.6 7.7

Fig. 24 Traditional mold heating circuit design and corresponding


steady-state distribution of mold surface temperature. (From Ref. 19.)

Nonuniform cavity surface temperatures occur in almost all types


of molding processes, but their importance has not been widely rec­
ognized. The type of analysis developed by Barone and Caulk can
not only alleviate the overall problem, but also saves on expensive
trial-and-error solutions to mold heating problems.

E. Residual S tresses

Virtually all molded articles contain residual stresses, and these


stresses affect the dimensional accuracy, dimensional stability, and
524 T ucker

LINE NUMBERS OIL TEMPERATURE (°C) FLOW RATE ( //m in )

3,4 200 11.7

2,5 200 12.7


1,6 170 9.2

(a)

- ________ ______ ________

* A è 1 A ¿ ^
_______ 1------------- 1 ---------------- 1 1

Fig. 25 Optimized mold heating circuit design and corresponding


steady-state distribution of mold surface temperature. (From Ref.
19.)

mechanical performance of the finished part. Some work has been


done on residual stresses in thermoplastics; it is known that resid­
ual stress can improve the toughness of molded parts and that un­
even cooling can cause parts to warp. The situation for thermoset­
ting materials is less well understood, though Marker and Ford [11]
convincingly demonstrated the existence of such stresses. Conse­
quently, the following discussion is general, and, in the absence of
any thorough studies, attempts to outline and organize the problem.
Compression Molding of Polymers 525

Residual stresses can be divided into three types according to the


mechanisms by which they arise [20,21]. The first type is stress due
to molecular orientation, and it arises whenever a polymer in the ru b ­
bery state is rapidly deformed and then quickly quenched below its
glass-transition temperature. Strictly speaking, this is not a residual
stress, since there is no physical stress in the material where no exter­
nal loads are applied. However, transparent polymers in this state do
show birefringence, and molecular orientation is often referred to as
”frozen-in stress” or, more confusingly, as ’’residual strain.” In fact,
what is frozen-in is the spatially nonuniform orientation of molecules.
Physical stresses do arise in these materials when they are heated back
above their glass-transition temperature. For example, a rod which is
twisted above Tg and quickly quenched to below Tg will untwist upon
heating [22]. This type of stress is common in thermoformed arti­
cles and contributes to the birefringence in injection moldings [21,
23]. However, it is probably not important in thermosetting polymers,
where forming occurs before chains of a significant length are formed.
It is important to understand this form of residual ’’stress” in order to
separate it from truly physical residual stresses.
A second type of residual stress is due to nonuniform thermomech­
anical properties. If a part is cured at T^ and is stress free at that
temperature, then stresses wñl arise as it cools to room temperature
if any location has different thermal expansion coefficients from the
rest of the part. In fiber-reinforced composites this can easily hap­
pen as a result of the fiber orientation which occurs during molding.
Preferential orientation of the fibers creates an anisotropic thermal
expansion, so differences in direction or degree of fiber orientation
over the part can cause stresses and distortion upon cooling. Fig­
ure 26 shows an extreme example of this phenomenon. This part was
molded in a flat rectangular mold and was flat when it was removed
from the mold. It consists of four layers of SMC with continuous,
aligned fibers. The two layers on the near side have fibers running
across the short direction of the plaque, and the two layers on the
back side have fibers running along the long direction of the plaque,
90^ to the other two layers. Upon cooling, the part curls up much
like a bimetallic strip, because the thermal expansion coefficient per­
pendicular to the fiber direction is much larger than the expansion
coefficient along the fiber direction.
If the distribution of thermal and mechanical properties in a part is
known, then estimating the stresses and distortions from this effect is
a straightforward thermoelasticity problem. The part may be assumed
to be stress free at Tm, and one simply calculates the elastic stresses
present at room temperature. This ignores the fact that the material
is viscoelastic and that some of the stresses can relax out. However,
Compression Molding of Polymers 525

Residual stresses can be divided into three types according to the


mechanisms by which they arise [20,21] . The first type is stress due
to molecular orientation, and it arises whenever a polymer in the ru b ­
bery state is rapidly deformed and then quickly quenched below its
glass-transition temperature. Strictly speaking, this is not a residual
stress, since there is no physical stress in the material where no exter­
nal loads are applied. However, transparent polymers in this state do
show birefringence, and molecular orientation is often referred to as
"frozen-in stress" or, more confusingly, as "residual strain." In fact,
what is frozen-in is the spatially nonuniform orientation of molecules.
Physical stresses do arise in these materials when they are heated back
above their glass-transition temperature. For example, a rod which is
twisted above Tg and quickly quenched to below Tg wñl untwist upon
heating [22], This type of stress is common in thermoformed arti­
cles and contributes to the birefringence in injection moldings [21,
23] . However, it is probably not important in thermosetting polymers,
where forming occurs before chains of a significant length are formed.
It is important to understand this form of residual "stress" in order to
separate it from truly physical residual stresses.
A second type of residual stress is due to nonuniform thermomech­
anical properties. If a part is cured at T^ and is stress free at that
temperature, then stresses wül arise as it cools to room temperature
if any location has different thermal expansion coefficients from the
rest of the part. In fiber-reinforced composites this can easily hap­
pen as a result of the fiber orientation which occurs during molding.
Preferential orientation of the fibers creates an anisotropic thermal
expansion, so differences in direction or degree of fiber orientation
over the part can cause stresses and distortion upon cooling. Fig­
ure 26 shows an extreme example of this phenomenon. This part was
molded in a flat rectangular mold and was flat when it was removed
from the mold. It consists of four layers of SMC with continuous,
aligned fibers. The two layers on the near side have fibers running
across the short direction of the plaque, and the two layers on the
back side have fibers running along the long direction of the plaque,
90° to the other two layers. Upon cooling, the part curls up much
like a bimetallic strip, because the thermal expansion coefficient per­
pendicular to the fiber direction is much larger than the expansion
coefficient along the fiber direction.
If the distribution of thermal and mechanical properties in a part is
known, then estimating the stresses and distortions from this effect is
a straightforward thermoelasticity problem. The part may be assumed
to be stress free at Tm, and one simply calculates the elastic stresses
present at room temperature. This ignores the fact that the material
is viscoelastic and that some of the stresses can relax out. However,
Fig. 26 Laminated SMC sheet, molded flat and curled by residual stresses during cooling.
528 T ucker

rate of reaction were ignored. This type of modeling does offer some
promise of estimating residual stresses in thermoset moldings.
Two general types of residual stress from nonuniform hardening
can arise in molded thermosets. The first, involving a stress that
varies through the thickness of the part, is an inherent feature of
temperature-activated cure. As the cure and temperature profiles
become increasingly nonuniform (i.e ., as Gr increases), the severity
of these stresses will increase. A second type of stress, varying
across the plane of the part, will arise if mold-filling flow causes some
locations to cure before others. This will not occur in a part with uni­
form thickness which fills the mold by lubricated squeezing flow, but
will occur in virtually every other part or with any other mold-filling
flow. Marker and Ford [11] showed this type of stress in disk-shaped
moldings where the flow caused the edges to cure before the center.
In summary, few good methods exist to estimate the magnitude of
residual stresses in compression-molded thermosets. All physical
residual stresses arise from spatial nonuniformities of thermomech­
anical properties or temperature and cure time.

IV . MORE COMPLEX R H E O LO G IC A L EFFECTS

There are two reasons to investigate more thoroughly the rheology of


reinforced thermoset-molding compounds. The first is that rheology
may affect the mold-filling pattern, with consequent effects on cur­
ing, fiber orientation, mechanical properties, etc. The second is that
rheology controls the relationship between mold closing speed and
closing force, and thus is important when designing presses or con­
sidering short-shot problems.

A. E ffect on Mold F illin g P a tte rn

Temperature-Dependent Viscosity and Thick Charges


As mentioned before, some experimenters have observed compression
moldings in which the outer layers of the charge flow faster than the
central layers of the charge [11]. This problem has been treated
analytically, by Lee, Denn, Crochet, and Metzner [31], who ana­
lyzed a Newtonian liquid with stratified viscosity being squeezed be­
tween parallel disks. Their model consisted of a central layer of high-
viscosity fluid sandwiched between two layers of low-viscosity fluid.
This simulates nonisothermal compression molding, where the fluid
nearest the plates is warmed while the central core remains cool. So­
lution was by finite elements, and a closed-form solution was also pre­
sented for some special cases.
Compression Molding of Polymers 529

The resiilts show that the dominant parameter in determining the


flow pattern is S = nmaxH^/iiminR^» where nmax and virnin are the
maximum and minimum viscosities, and H and R are the gap height
and disk radius, respectively. When this parameter is small, the ve­
locity profile has a single maximum halfway between the plates, and
in the region covered by the initial charge there is parallel squeez­
ing. Parallel squeezing means that fluid planes initially normal to
the z direction remain flat and parallel to one another. Note that
the condition for this type of squeezing to occur is exactly the same
as the condition for the generalized Hele-Shaw model to be valid [Eq.
(2 7 )]. (Generalized Hele-Shaw flow does not always involve parallel
squeezing, but for Newtonian fluids the parallel squeezing flow is a
generalized Hele-Shaw flow.)
When S is large, the type of velocity profile observed by Marker
and Ford occurs. Velocity is maximum in the low-viscosity layers,
which are squeezed out preferentially. The central core of high-
viscosity fluid experiences very little shearing and deforms pri­
marily by biaxial extension. In addition, parallel squeezing does not
occur. The low-viscosity surface layers become thinnest near the
edge of the disks and thick in the center, and the central portion of
the surface layer may eventually be completely surrounded by the in­
ner layers (Fig. 27). This effect has been observed in SMC moldings
by Denton [32] and produces the patterns in Fig. 28.
As noted, thick SMC charges do not expand in the mold in the
same way as thin charges do. Since the parameter S depends on
charge thickness and controls the velocity profile when viewed from
the side, it might also control the development of the charge shape
when viewed from the top. This has been verified by Lee and
Tucker [33], who performed a finite element simulation of a rec­
tangular charge with layered viscosity, much like the disk calcula­
tions of [31]. Although their calculation neglects some details of
the fountain flow at the flow front, it does show the right trend
in charge shape development. Figure 29 shows how the tendency
of a thin charge to become round during flow is lost as the charge
thickness (and S) increases. Instead, the flow tends towards nearly
biaxial extension and, in the extreme cases of very large S, b e ­
comes a lubricated squeezing flow.
Nonisothermal flows where S is large cannot be modeled by the gen­
eralized Hele-Shaw model, which ignores important stresses due to in­
plane stretching. However, models including the complete equations
of motion for creeping flow appear to do a good job. Thus, these
flows can be modeled adequately at the cost of doing a more compli­
cated calculation. Calculations of this type, including heat transfer
as well as flow, are now being developed [34].
530 T ucker

(a)
I ^ 1
fluid 1

fluid 2

(b) H = I.O
t = 0.

H = 0.75
t = 0.33

f
fluid J2i_
fluid 2,

H =0.50
t =1.02

Fig. 27 Shape change of layered Newtonian fluids with differing vis­


cosities in squeezing flow between disks, (a ) initial configuration;
(b ) configuration after squeezing. (From Ref. 31.)

A n iso tro p ic E ffe cts

A complicating effect is that the addition of fibers to a viscous liquid


introduces anisotropy into the viscous response. This has been dem­
onstrated experimentally for SMC by Lee, Marker, and Griffith [35].
They measured shear viscosities in drag flow between parallel plates
Compression Molding of Polymers 531

(a)

(b )

Fig. 28 Cross section of SMC molding made with colored plies, (a)
initial charge; (b ) final molded part. (From Ref. 32.)

and extensional viscosity in lubricated squeezing flow between paral­


lel disks. For a Newtonian fluid one would expect the extensional
viscosity n to equal three times the shear viscosity n. (Note that
biaxial extensional flow is the classical Trouton uniaxial extension
experiment run in reverse.) However, Lee et al. measured exten­
sional viscosities which were 50 to 200 times higher than the shear
viscosities, in a strain rate range where the resin paste is Newton­
ian. The reason is that in SMC all the fibers are oriented in the
plane of the sheet and they resist stretching in the plane of the
sheet much more than they resist shearing across the sheet thick­
ness. Lee, Marker, and Griffith use the analogy between elasticity
and viscous flow together with theories predicting the moduli of com­
posite materials to explain the difference between shear and exten­
sional viscosities.
In a more complex flow situation the effect of this anisotropy
should be to emphasize the importance of extensional stresses. This
would limit the applicability of the generalized Hele-Shaw model,
which neglects those extensional stresses. There have been no the­
oretical or experimental studies of the effects of anisotropy on mold-
filling patterns. One stumbling block to such a study is the need
for a representative constitutive equation. Possible choices include
Ericksen^s transversely isotropic fluid [36,83], Hand’s model for
anisotropic fluids [37], or the ’’structural” model of Dinh and Arm­
strong [38]. The anisotropy of materials like SMC is major , not mi­
nor, and will eventually have to be accounted for.
532 T ucker

Fig. 29 Flow fronts of rectangular charge predicted by nonisothermal


model. Compare to experimental results in Fig. 10.
Compression Molding of Polymers 533

B. E ffect on Mold Closing Force

P u re ly V isco u s S u b s ta n c e s

Rheologists have long been interested in relating the applied force and
gap closure rate in squeezing flow. Scott [39] solved the problem for
an isothermal power-law fluid in a narrow gap between parallel disks,
using a quasi-steady lubrication model and assuming parallel squeez­
ing. (His solution is a special case of generalized Hele-Shaw flow .)
His result relating the applied force F to the rheological parameters
m and n, the instantaneous gap height 2h, and the disk radius R has
been widely used in interpreting rheological measurements. It is often
called the Scott equation
n + 3
„ _ (- h )^ /2 n + l \n TrmR^
( 68 )
2n + 1 \ 2n / n + 3

Here h is the time rate of change of h. Recently, Brindley, Davies,


and Walters [40] presented a solution of squeezing flow between disks
for a viscous fluid with an arbitrary relationship between stress and
strain rates. In this case parallel squeezing does not occur, as it
does in the power-law case. Many researchers have used squeez­
ing flow between disks as a technique for rheological measurement.
Ghandhi and Burns [41] measured the viscosity of fiber-reinforced
polyester molding compounds this way and found power-law behavior.
Their materials contained low loadings (10% or less by weight) of rela­
tively short (3-mm) glass fibers. As the loading and/or aspect ratio
of the fibers was increased, the materials became more viscous and
more shear thinning (i .e ., m increased and n decreased). This is
consistent with most other measurements on fiber suspensions (e .g .,
see the review in Ref. 42).

V isco ela stic E ffe c ts

When squeezed rapidly, viscoelastic fluids depart from the predictions


of purely viscous models like the Scott equation. All experimental evi­
dence is that elasticity makes a fluid a better lubricant [40,43]. That
is, a greater force arises for the same closing rate (or a slower closing
rate results from the same force) in a viscoelastic fluid than in a vis­
cous fluid with equal viscosity. When the force is provided by a
weighted upper disk the disk may ’’bounce” ; i.e ., gap height shows
a decaying oscillation with time instead of decreasing monotonically.
534 T ucker

The most common experiment is to fix the applied load (by releas­
ing a weighted upper disk at t = 0 ) and measure ti/ 2 , the time for the
gap height to fall to one-half of its original value. Leider [43] showed
that elastic effects are negligible and that the Scott equation correctly
predicts t i / 2 when t i / 2 > X, a time constant for the fluid. When t i / 2
< A, viscoelastic effects become important, and the gap closes more
slowly. He defined X as
l/(n^-n)
(69)
(fi)
where both the shear viscosity and the first normal-stress difference
of the fluid are fit by power-law models of the form

3v
T
xy
=m
ay (a (70)

3v
T
XX
- T
yy
m'
ay ft) (71)

Brindley et al. [40] had similar success predicting t i / 2 at low loads


from viscosities measured in simple shear, but they were not success­
ful in correlating the high-load (fast squeezing) cases with any such
time constant.
Are viscoelastic effects important in compression molding? Prob­
ably not for thermoset polyesters. Lee, Marker, and Griffith [35]
give the viscoelastic time constant for an SMC paste (no fibers) as
a function of temperature. At 25°C they measure X = 1.27 sec, and
at 70°C X is 0.005 sec. Only at room temperature is X close to t i / 2
(usually about 2—4 sec); at temperatures representative of the mold
wall X is extremely small compared to ti^ 2 * The same may not be true
of thermoplastic stamping operations, where closing speeds are as
rapid as possible and the polymers have higher molecular weights
and longer time constants. In addition, viscoelastic effects may be
evident in the mold closing force.
From a theoretical standpoint, viscoelastic squeezing flow is a d if­
ficult problem. Despite the fact that experiments invariably show
that elastic liquids are better lubricants than their inelastic counter­
parts, all theoretical calculations to date predict that elastic fluids
should be worse lubricants. The list of theoretical calculations in­
clude Tanner’s lubrication calculation with an upper convected Max­
well model [44], Kramer’s calculation with Lodge’s rubberlike liquid
[45], Brindley, Davies, and Walters’ calculation with a second-order
Compression Molding of Polymers 535

fluid [40] and Phan-Thien and Tanner^s exact solution (including in­
ertia as well as elasticity) for an upper convected Maxwell fluid [46].
The problem seems to be that these constitutive equations are inade­
quate to model this type of flow. Leider and Bird [47] suggested that
stress overshoot on startup of simple shear flow is responsible for the
improved load-carrying capability of viscoelastic fluids, and they did
an approximate calculation to support their claim. Another possibility
is that the constitutive equations do not adequately represent the be­
havior of the fluid under the extensional deformation inherent in
squeezing flow. Whatever the reason, transient viscoelastic squeez­
ing flow remains as a challenge to theoretical rheologists.

F i b e r C om p a ctio n

Molding compounds with high concentrations of fibers show an unex­


pected force response in squeezing flow. For example, in constant
velocity squeezing the Scott equation ( 6 8 ) predicts that F is propor­
tional to gap height to the power - ( 2 n + 1 ), giving a force-versus-
time curve which increases monotonically and always has positive
curvature. However, using SMC with as little as 25% glass fibers by
weighty Süva-Nieto, Fisher, and Birley [48] observed a force-versus-
time curve with two inflection points (Fig. 30). Ghandhi and Burns
[41] also noted a departure from power-law behavior in squeezing
glass fiber—polyester compounds when the gap height fell below 5
mm. This distance is comparable to the 3-mm fiber length in their
experiments. Similar effects can be seen in the data of Lee, Marker,
and Griffith [35] in lubricated squeezing flow. Their experiments
used constant force and show the strain rate decreasing as strain in­
creases. Zentner [49] has found that applying a constant squeezing
force will cause SMC to squeeze down to an equilibrium thickness.
This equilibrium thickness is nearly independent of applied force
(Fig. 31), though the time to reach equilibrium is much longer for
lower forces. These results all show that SMC and similar concen­
trated suspensions can behave very differently from Newtonian,
power-law, or viscoelastic fluids.
One explanation for some of these phenomena is that the fibers
form a mechanical network. This network can support part of the
applied load and carries more load as the material becomes thinner.
In SMC the fibers are oriented nearly in the plane of the sheet. It
seems plausible that fibers may overlap or pile up to form short mech­
anical load paths from the top plate to the bottom plate. The strength
and/or density of these pñes should increase as the gap height de­
creases.
This hypothesis is supported by an experiment performed by Zent­
ner [49,50]. Here SMC is tested in lubricated squeezing flow between
536 T ucker

Fig. 30 Squeezing force vs. time for SMC squeezed between parallel
disks. (From Ref. 48.)

parallel disks. A small-amplitude sinusoidal load is applied, and the


amplitude and phase shift of the strain response are used to compute
a dynamic modulus. Figure 32 shows the magnitude of the modulus
and the phase shift as a function of compressive strain. At low
strains the phase shift shows a significant viscous component due to
the resin paste. As the sheet is squeezed down, the magnitude of
the modulus increases and the phase shift decreases, because the
elastic fiber network comes into play. Finally, the modulus takes a
dramatic rise and the phase shift drops to zero, as the fiber net­
work becomes stiff enough to dominate the response.
Another interesting phenomena, which may be related, is that
when SMC is sheared across its thickness it deforms in discrete lay­
ers, like a deck of cards [35]. The layers are several times thinner
than the SMC sheet.
In extreme cases the fibers can be trapped, and the resin squeezed
out. This was observed by Gandhi and Burns [41] at very low
Compression Molding of Polymers 537

Fig. 31 Equilibrium thickness as a function of squeezing force for


SMC R-30 in lubricated squeezing flow. (From Ref. 50.)

squeezing rates, when presumably the viscous drag forces could not
move the fibers against restraining frictional forces.
Unprocessed SMC is also highly porous. Presumably these voids
are squeezed out during the early stages of compression, whether in
molding or in rheological tests. One would expect this to make the
apparent viscosity lower in the initial stages of squeezing and higher
once the voids had disappeared. This effect has not been studied
quantitatively, though it is known that cured SMC is still slightly
porous.
The thickening reaction used to increase resin viscosity before
processing is also poorly understood. Presumably the thickening
creates some type of effective cross-link between the short polymer
chains in the resin, for the thickened resin has elasticity as well as
a high viscosity. These bonds may rupture as the SMC begins to
flow, causing a decrease in viscosity (i.e ., a thixotropic effect).
This has been suggested by SRva-Nieto, Fisher, and Birley [48].
Because of the complexity of these effects, it is extremely difficult
to predict force response when squeezing SMC or similar reinforced
materials. Interestingly, this phenomena does not seem to disturb the
predictions of the mold-filling pattern discussed in Section III. A l­
though in principle the pressure distribution calculated from the
538 T ucker

Fig. 32 Dynamic modulus and loss tangent at room temperature as a


function of compressive strain for SMC R-30 in lubricated squeezing
flow. (From Ref. 49.)

generalized Hele-Shaw model [Eqs. (18) or (24)] can be integrated to


find the total mold closing force, in practice the answers are not at
all accurate. Prediction of mold closing force will require a more
thorough understanding of the fiber compaction phenomena and its
role in squeezing flow.
Compression Molding of Polymers 539

V. FIB ER O R IE N T A T IO N

Reinforcing fibers become oriented by flow during processing. This


is true in compression molding and in many other processes as well.
The fiber orientation pattern in a composite material is a primary fea­
ture of the microstructure. The composite is stiffer and stronger in
the direction of maximum orientation, weaker and more compliant in
the direction of minimum orientation. For many years flow-induced
fiber orientation was treated as an undesirable by-product of process­
ing. An effort was made to produce materials with random orienta­
tion of fibers having isotropic mechanical properties. However, many
workers are now trying to predict and control fiber orientation in re­
inforced polymers. By controlling the fiber orientation pattern in
molded parts, one can tailor the characteristics of the material to suit
each individual application, much in the same way that mechanical
properties and microstructure of steels are controlled through heat
treatment.
This section introduces the basic phenomena of fiber orientation,
then discusses quantitative models to describe fiber orientation and
the ways in which one can predict the orientation pattern in molded
parts. Related phenomena such as fiber length degradation during
processing are discussed at the end of the section.

A. Phenomenology

To understand how fibers are oriented by flow, consider a single


rigid fiber immersed in a viscous fluid. Since polymeric fluids al­
most always have high viscosities, and since most fibers are small,
we will consider the case of very low Reynolds number flow (i.e .,
creeping flo w ). Assume that the fluid is undergoing steady simple
shear flow in the x direction ; i.e .,

V = Yy V =v =0 (72)
X ^ y z

as shown in Fig. 33.


If the center of the fiber is at the origin, then the fiber experi­
ences no net translational force, and its center remains stationary.
However, the two ends of the fiber see the fluid moving by in op­
posite directions, so a stationary fiber experiences a net torque.
Since no external moments are applied to the fiber, it must rotate in
such a way that the net moment exerted on the fiber by the fluid is
zero. In Fig. 33 the fiber rotates clockwise in response to the shear­
ing motion of the fluid. This simple example shows that fibers are or­
iented in response to deformation of the suspending fluid. If there
is no deformation, then fiber orientation does not change.
540 T ucker

F ig . 33 Rigid fiber in simple shear flow.

The orientation behavior of a single rigid ellipsoidal particle im­


mersed in a Newtonian fluid was first solved by Jeffery in 1923 [51].
He solved for the flow field around a particle with arbitrary transla­
tion and rotation in a fluid which is being deformed homogeneously.
By applying the conditions of zero net force and moment on the p ar­
ticle^ he derived equations for the translational and rotational ve­
locities of the particle. Equations appropriate for fibers are found
by considering the case of a prolate spheroid of revolution, i . e . , a
body whose surface is described by

= 1 (73)
a^ b" b^

Here, Xi, X 2 , and X3 denote directions in an orthogonal coordinate


system attached to the fiber, a and b are constants, and a/b is on
the order of 1/d, the length-to-diameter ratio of the fiber. For a
fiber lying in the x -y plane and flows which are planar in the x -y
plane Jeffery’s equation becomes

3v 9v
d(j) _ ___ X
•sin cos sm
dt + 1 8 x 9y

8 v 3v
sin cos y
+ cos
3x ay
Compression Molding of Polymers 541

dv 8 v
X 2 ___ X
■sin cb cos —— + COS^
r ^+ 1 8x 9y
e

8 v 8 v
-sm _ y + sm d> cos y (74)
3x ay ]

The quantity is called the equivalent ellipsoidal axis ratio and


equals a/b for an ellipsoid of revolution. The angle cj) is defined in
Fig. 33.
Two examples suffice to show how flow affects the orientation of
individual fibers.

P la n a r E x t e n s i o n a l Flo w

Planar extensional flow is a two-dimensional stretching flow with a ve­


locity field like

V = ex V = -ey (75)
X y

Specializing Eq. (74) to this flow gives

r ^- 1 ( \
d(j) _ _e___
- 2 e sin d) cos (76)
dt r ^+ 1
e

which is easily solved to give

r -h1
tan (j) = tan cj)q exp (77)

where cj)o is the initial orientation angle. In this flow all fibers rotate
monotonically toward c{) = 0, which is the direction of stretching. Fi­
bers oriented with = 0 are in stable equilibrium, while those at cj) =
7t / 2 are in unstable equilibrium. This example demonstrates the fa­
miliar rule that in a stretching flow fibers orient in the direction of
stretching.

S im p le S h e a r Flo w

In a simple shear flow described by

V = yy V =0 (78)
X ^ y
542 T ucker

Equation (74) reduces to

d(|) -Y
(79)
dt

This integrates to give

tanci) = r tan (80)


^ e

when the initial condition is cf) = 0. Equation (80) shows that the p ar­
ticle rotates periodically. According to Eq. (79), the angular velocity
of the particle is greatest at (f) = tt/2 , when it lies across the stream­
lines, and least at (|) = 0, when it aligned with the streamlines. The
ratio of the angular velocities at these two points equals re^. Thus,
long particles such as fibers seem to rotate rapidly towards the posi­
tion (j) = 0 , dwell close to that position for a long time, then rotate
rapidly again until aligned with the streamlines once more. This ex­
ample shows the second familiar rule: in shear flows the fibers align
with the streamlines.
Extensive experiments by Mason and co-workers ( e . g . , [52]) have
verified Jeffery^s equations and shown that they apply to cylindrical
fibers as well as to ellipsoidal particles. Bretherton [53] has shown
theoretically that Jeffery’s equation applies to any axisymmetric rigid
particle. This is simply a consequence of the linearity of the equa­
tions of motion for creeping flow of a Newtonian liquid and of the
symmetry of the particle.
When applying Eq. (74) to cylinders, one must use an appropriate
value of Vq . For example, Harris and Pittman [54] correlated their
experimental results with

0 .
(81)
V

in the range of 2 0 < 1 /d < 400.

B. Models fo r C on cen trated Suspensions


Since Jeffery’s equation was derived for a single fib e r, it applies only
to suspensions which are so dilute that fibers never approach one an­
other. If C is the concentration (volume fraction) of fibers in the
suspension this requirement is
-2
C « ■■ '
n
Compression Molding of Polymers 5U3

Fig. 34 Motions of individual fibers in simple shear flow in a dilute


suspension (C = 0.01, 1/d = 15, nylon monofilaments in 10,000 cs
silicone oil).

For example, if 1/d = 10 (a small value for reinforcing fibers), then C


muât be less than 1% for Jeffery^s equation to apply. Not many use­
ful composites are made with such a small amount of reinforcement.
Experiments in Ref. 55 show that interactions between fibers alter
their orientation behavior, though the qualitative trends of the
Jeffery model still apply. This effect can be seen in Figs. 34 and
35. The data were taken on individual fibers in simple shear flow.
The solid line is the prediction of Jeffery^s Eq. (80). The fibers in
the more dilute suspension in Fig. 34 follow Jeffery^s result closely,
but in the more concentrated suspension in Fig. 35 they orient more
slowly, and the motion is somewhat random.
A recent theory of two-body interactions by Lee and Springer
[56] has successfully modeled these dilute suspension experiments.
However, molding compounds for composite material typically contain
a high concentration of fibers. One can define the number of ’’near­
est neighbors” a test fiber has as the number of fibers whose cen­
ters lie within a distance 1 of the center of the test fiber. This
number, N is

N (83)
■ è i
For example, a typical reinforced molding compound might have C =
0.30 and 1/d = 20. For this composite, N = 120. A theory with only
544 Tucker

Fig. 35 Motions of individual fibers in simple shear flow in a moder­


ately concentrated suspension (C = 0.08, 1/d = 15, nylon monofila­
ments in 1 0 , 0 0 0 cs silicone oil).

two-body interactions is unlikely to provide accurate predictions for


such a case. Also, the random nature of the fiber motions in Fig. 35
suggest that a deterministic model such as Jeffery*s equation is not
appropriate for concentrated suspensions.
Instead, one can take a statistical approach [57,58]. Rather than
calculate the orientation angle of individual fibers, one can calculate
a probability function for orientation. For planar orientations this
function is 4 j((|)), and is defined such that the probability of a fiber
being oriented between the angles c|)i and (j) 2 is

P(<i)l < c{) < (j)2 ) = J Ip(ct)) d(i) (84)

(Similar developments can be made for three-dimensional orientation


[5 7 ].) The symbol ^ can be thought of as an orientation distribution
function or as a probability density function. The angle where is
maximum is the angle of maximum orientation, and the angle where ijj
is a minimum is the angle of minimum orientation. The function ij^((j))
equals a constant for a randomly oriented structure. Note that real
parts may have one type of orientation in one location and another or­
ientation state somewhere else. Thus, ifj is also a function of position
within the part.
Compression Molding of Polymers 545

Since a fiber oriented at angle c() cannot be distinguished from one


oriented at cj) + tt , the function ijj must be periodic with period tt :

ip((f)) = 4;((|) + tt) (85)

Because of this periodicity, one need only keep track of ip in the in­
terval from 0 to TT. Since all fibers must lie in this interval, ip must
also satisfy the normalization condition

/■
( (|)) d(j) ( 86)
0

That is, the area under the curve of versus cp equals unity. If i|j((j))
changes with time, this requirement stUl holds. Thus, ip is also sub­
ject to a conservation equation. In differential form this condition is

(87)
at

Equation (87) is often called the continuity equation because of its sim­
ilarity to the continuity equation in fluid mechanics. In this equation
(f represents the average angular velocity of the fibers and can be a
function of (p and time. For example, one could substitute Eq. (74)
for i in Eq. (87). The resulting equation would give the evolution
of the orientation distribution function ip in a dilute suspension, where
Jeffery^s equation is valid.
For concentrated suspensions Folgar and Tucker [57,58] have pro­
posed that the angular velocity of fibers can be approximated by

8 v 8 v 9v
X - sin
• 2^
. --— X 9 V
= \ - sin cos + cos d)
ay ^ ax

av c, Y
__ y a\ p
+ sin d) cos ( 88 )
9y a ({)

for two-dimensional problems. The term in braces on the right side


is Jeffery’s equation for a fiber with infinite aspect ratio. The other
term on the right side is introduced to approximate the randomizing
effect of interaction between fibers. In this term y is the magnitude
of the strain-rate tensor, and Cj is a phenomenological interaction co­
efficient. The terms in Jeffery’s equation proportional to l/(re^ + 1)
are assumed to be negligible in relation to this term.
546 Tucker

Combining Eqs. (87) and ( 8 8 ) gives a governing equation for the


orientation distribution function. Because fibers move with the aver­
age velocity of the fluid, ijj is a convected quantity. To account for
this, we replace the partial time derivative in Eq. (87) by the ma­
terial derivative, which is

= + v -^ + v (89)
Dt 3t X dx y dy

for the two dimensional case. The resulting equation is

Bv Bv
ri Y
• _d_ ___X ___ X
—- = C ip^-sin cos sin"
Dt r d(p^ 3 ( j) 3x

Bv 3v
+ cos (|)
2 y
+ sm COS
y
(90)
oX 3y

This equation can be solved for the time evolution of ijj, provided the
velocity gradients on the right side are known. One must have an in­
itial condition ipCcj), t = 0). The boundary conditions in cf) are provided
b y E q . (85):

i|;((j)) = ipCcj) + tt) (91)

Ml Bip I
3^ B(p\ (92)
'0

Several interesting consequences of this theory seem to be borne


out by experiments. First, for any homogeneous flow there exists a
steady-state orientation distribution which (for nonzero values of C i)
is not perfectly aligned. This is different from dilute suspension b e ­
havior. A dilute suspension in simple shear flow never reaches steady
state, but has a periodic orientation state. A dilute suspension in
planar extensional flow has a steady state in which all the fibers are
aligned in the direction of stretching. Figure 36 shows steady-state
orientation in simple shear flow. The Folgar-Tucker model also does
an excellent job of fitting this data.
Figure 37 demonstrates the role of the interaction coefficient Cj.
This coefficient multiplies the interaction term, which represents a
randomizing effect much like diffusion. In contrast, the other terms
represent the aligning forces of flow. As Cj increases, the random­
izing effect becomes stronger and the steady state is less oriented.
It is also widely believed that elongational flows are better at
orienting fibers than shear flows. Figure 38 shows the predicted
Compression Molding of Polymers 547

Fig. 36 Steady-state fiber orientation in simple shear flow fit by Fol­


la r and Tucker^s theory. The suspension has 8 % by volume nylon
monofilament fibers in 1 0 ^ 0 0 0 cs silicone oü.

steady-state distributions for simple shear flow and planar extensional


flow. The Folgar-Tucker theory indeed predicts better alignment in
the extensional flow than in the shear flow.
Although some predictions of this theory do not quantitatively agree
with experiments, it is a substantial improvement over dilute suspen­
sion theories and does an excellent job in many cases. With such a
theory in hand one should be able to make useful prediction of fiber
orientation in molded parts.

C. P red ictin g F ib e r O rie n ta tio n

Having chosen some model for fiber orientation, one must still relate
this to the particular processing situation. As can be seen from Eqs.
(74) and (90) the effect of processing on fiber orientation enters the
548 Tucker

Fig. 37 Effect of the interaction coefficient Ci on steady-state dis­


tributions of fiber orientation in simple shear flow.

equations through the velocity gradients. Thus, one must solve for
the flow field in the process in order to find the fiber orientation pat­
tern .
Two factors complicate the computation. The first is that the rhe­
ological properties of a reinforced molding compound are a function of
the fiber orientation state. Consequently, the calculations for the
flow field and for fiber orientation are coupled. All calculations to
date have made the idealization that the flow field is independent of
fiber orientation, decoupling the two calculations. As mentioned be­
fore, rheological models which account for fiber orientation have only
recently been developed, so this is an area which will demand work in
the future. There are some instances where the idealization may be
quite reasonable, such as the Hele-Shaw flow of an SMC with planar
fiber orientation.
Compression Molding of Polymers 549

Orientation Angle, $

Fig. 38 Relative orienting ability of simple shear flow and planar


elongational flow according to the Folgar-Tucker model.

The second difficulty with fiber orientation calculations arises be­


cause fibers are convected with the fluid. This permits two different
schemes of calculating orientation. Either one may follow the trajec­
tories of individual fluid particles, integrating the fiber orientation
equations for the velocity gradients experienced by the particle. Or,
one may use the material derivative to formulate an equation for the
orientation state at fixed points. Both of these approaches have been
used.
The calculations by Givler, Crochet, and Pipes [62a] for dilute sus­
pensions and by Jackson, Folgar, and Tucker [59,60] for compres­
sion-molded SMC are examples of the particle tracing method. Jack-
son et al. used the Folgar-Tucker model for fiber orientation and a
550 T ucker

generalized Hele-Shaw finite element calculation for the flow field.


Since the fibers in SMC are much longer than the gap height, the fi­
ber orientation pattern was taken to be planar. The nonisothermal
nature of SMC mold-filling flow leads to a flat gapwise velocity pro­
file for Hele-Shaw flow. Therefore, the gradients of the average ve­
locities, Ü and V , were used in the fiber orientation equation, and
the fiber orientation was assumed to be constant through the thick­
ness of the part. To calculate the orientation pattern, they chose
individual points in the part and followed their paths through the
mold-filling process. Since the mold-filling flow is assumed to be
decoupled from the fiber orientation pattern, the mold-filling simu­
lation was completed before doing the fiber orientation calculations.
An example result is compared to experiments in Fig. 39. The ex­
periments used a suspension of nylon monofilament fibers in silicone
oil and were performed in a mold with a transparent upper platen.
Figure 7 shows the predicted and experimentally observed flow fronts
during the molding, and the calculated paths of the fluid particles
that were followed. Photographs taken during filling were analyzed
to find the fiber orientation distributions. The initial orientation con­
dition at point 2 in Fig. 7 is shown in Fig. 39(a). Subsequent orien­
tation states are shown in Figs. 39(b) and (c ). The predictions com­
pare favorably with the experiments, though some improvement could

Fig. 39 Experimental (histograms) and predicted (lines) fiber orien­


tation distributions for point 2 in Fig. 7. (a ) initial condition; (b )
second intermediate step; (c ) final condition.
Compression Molding of Polymers 551

be made. The major source of error in this particular calculation is


inaccuracy in the velocity gradients provided by the mold-filling cal­
culation. Recall from Section II.B that in the Hele-Shaw formulation
one solves numerically for the pressure field. The velocities are then
proportional to the pressure gradients, so the velocity gradients are
552 T ucker

second derivatives of the pressure field. The calculations in Fig. 39


simply used the second derivative of the numerical results for pres-
sure to find the velocity gradients, a procedure which can introduce
sizable errors for generalized Hele-Shaw flows. An improved method
for calculating velocity gradients has been developed by Lee [61] and
should make this particular calculation much more accurate.
The second method, solving for orientation at fixed points, has
been developed by Givler [62,63]. His flow calculations are also done
with finite elements, but they are for general steady two-dimensional
Newtonian flow. Orientation is assumed to follow Jeffery^s equation
as the fiber aspect ratio goes to infinity. Givler does not calculate
the entire fiber orientation distribution but instead finds a single or­
ientation angle at each point. Thus, his calculations are strictly ap­
plicable only to a dilute suspension in which the fibers are completely
aligned at every point. However, the results do provide useful in­
formation about the effect of flow on fiber orientation. Example re ­
sults are shown in Figs. 40 (streamlines) and 41 (orientation field)
for plane flow past a square slot. Note that the well-known effects
of converging flow, diverging flow, and shear flow in orienting fibers
can all be observed in this example.

Fig. 40 Streamlines for planar flow past a square slot, analyzed by


Givler. (From Ref. 62.)
Compression Molding of Polymers 553

Fig. 41 Fiber orientation patterns for planar flow past a square slot.
Time increasing clockwise from the upper left. (From Ref. 62.)

Each of the two approaches has its advantages and disadvantages.


The particle tracing method can provide the entire distribution of fi­
ber orientation at any given point. This is the information needed at
each point in the part to make detailed predictions of the mechanical
properties of reinforced composites. However, the analysis of a re ­
alistic part would require the tracing of many points, which would be
a tedious task. There are also difficulties which arise near stagnation
points in the flow [63]. The approach which uses fixed points in space
provides an excellent overall picture of the orientation pattern. How­
ever, it provides only limited information about the orientation state
at each point. If one were to extend this approach to calculate the
554 T ucker

entire orientation distribution at every point (which is the information


that is needed), the size of the calculation would become unmanage­
able. For example, accurate representation of the distribution func­
tion would require at least 2 0 points at various values of cj) for every
point in space (every node point). The simple problem in Figs. 40
and 41 has 667 nodes, so each time step would involve solving finite
element equations for more than 13,000 variables. A more realistic
part with complex geometry would have many more nodes, making the
problem much larger. Although calculations of this size are not out
of the question for modern finite element programs, they are prob­
ably not worth the time and expense. A simplification of the problem
that would provide the complete orientation field in both space and
angle for a reasonable amount of computation would be extremely de­
sirable.

D. M easuring and C h a ra c te riz in g F ib e r O rie n ta tio n

Experimental techniques to measure fiber orientation are essential to


test the accuracy of fiber orientation calculations. As industry b e ­
comes more sensitive to the effects of fiber orientation in composites,
good measurement techniques will be demanded for quality control and
failure diagnosis work. This section also discusses alternative ways
of quantifying the fiber orientation state.

Orientation Measurement
The first step in measuring fiber orientation is to obtain a picture of
the fibers. When the matrix is transparent, as are some polymers and
most model systems used for laboratory experiments, standard photog­
raphy or photomicrography may be used. In some cases it is helpful
to add opaque tracer fibers [57—60] , since the picture cannot be in­
terpreted if all the fibers are visible.
For thermoplastic matrix materials a technique called contact micro­
radiography gives good results [64—66]. A slice of the part approxi­
mately 0 . 1 mm thick is placed on a high-resolution x -ray plate and ex­
posed with an x-ray source. The plate may then be viewed and photo­
graphed with an optical microscope. This technique is particularly well
suited to materials with small fibers.
For composites where the fibers are larger and remain in bundles, a
radiographic tracer technique has been developed [60,67]. Special
glass fibers are made, using a glass with a high lead content. A small
percentage of these fibers are incorporated into the molding compound.
The lead glass fibers are much more absorbent to x-rays than ordinary
E-glass, so they are easily distinguished on radiographs. An advan­
tage of this technique is that it is nondestructive. Also, the lead glass
Compression Molding of Polymers 555

tracer fibers are representative of the rest of the fibers in the part,
whereas other tracers such as metal wires may not be.
Once a picture has been obtained, it must be analyzed. One ap­
proach is to simply measure the orientation angle of every fiber on
the photograph. A histogram of the results, properly normalized, be­
comes an experimental measurement of the orientation distribution func­
tion A relatively easy way to do this is to place the photographs
on a digitizing tablet connected to a computer. One can quickly digi­
tize the endpoints of each fiber, and let the computer calculate the
angles and store the data. Advantages include the ability to manip­
ulate the data with the computer and redraw the picture to quickly
check the validity of the data [57—59]. This method is direct and
subject to little or no distortion, but it is very time consuming and
extremely difficult to automate.
A more rapid technique uses the picture of the fibers as a diffrac­
tion mask in a Fraunhoefer diffractometer [68—69]. The apparatus is
shown in Fig. 42. The picture of the fibers is made into a high-con­
trast slide in which the background is opaque and the fibers are
transparent (or vice versa). Placed in the diffractometer, the slide
creates a diffraction pattern which is related to the distribution of
fiber orientation on the slide. Figures 43 and 44 show examples of
different fiber orientation patterns and the diffraction patterns they
produce. A quantitative relationship between the orientation of fi­
bers on the slide and the diffraction pattern has been derived [ 6 8 ,
69]. As an excellent first approximation the intensity of the dif­
fraction pattern at a fixed radius from its center is proportional to
the fiber orientation distribution, shifted by 90^. That is, the in­
tensity of the diffraction pattern, I, is

I (r , (})) = f(r)i|; ( i - * ) (93)

Fig. 42 Schematic of Fraunhoffer diffraction apparatus for measuring


fiber orientation. (From Ref. 68.)
556 T ucker

Fig. 43 Simulated fiber orientation distributions used to produce Fig.


44. (From Ref. 6 8 .)

where f ( r ) is a function depending on the intensity and wavelength of


the light source, the open area on the slide and the focal length of the
focusing lens. Note that f ( r ) need not be measured, since must sat­
isfy the normalization condition. A random fiber orientation produces
a round diffraction pattern, whereas fibers oriented primarily in the
vertical direction produce a diffraction pattern with horizontal lobes.
To date this device has been applied in a semiquantitative manner.
However, it offers the opportunity to develop an automated measure­
ment technique, and such equipment is now under development [81].

C h a ra c te riz in g O rien ta tio n

Up to this point we have discussed fiber orientation at a point as being


characterized by the distribution function i|j((()). However, one might
ask if there are other, more convenient ways of describing fiber ori­
entation .
One description of this type is the orientation parameters proposed
by McCullough [70,71]. For planar orientations they are defined as

f = 2 / cos^(j)ip( (j)) d(j) - 1 (94)


0
Compression Molding of Polymers 557

0.0

f = 0.3
P

f = 0.6
p

0.9

Fig. 44 Experimental and predicted diffraction patterns for the orien­


tation distributions shown in Fig, 43. (From Ref. 6 8 .)

I
g cos^^cpijjCcj)) d(j) (95)
L0
L
Although at first glance these definitions may be seen arbitrary, they
are derived from a theory which predicts the mechanical properties of
a short-fiber composite as a function of fiber orientation. When the
distribution function is symmetric about = 0 —i . e . , when
558 T ucker

(96)

- f and g contain all the information needed to compute the influence


of fiber orientation on the elastic modulus of the composite [68,70].
The parameter f is a planar version of the familiar Hermans orienta­
tion factor [72].
The factors f and g are scaled to provide convenient interpretation
of the amount of fiber alignment. For a random orientation, f = g = 0.
For perfectly aligned fibers in the (|) = 0 direction, f = g = 1. Thus
all realistic planar orientation states have f and g between zero and
unity, with smaller values representing less alignment and larger val­
ues representing more alignment.
The orientation factors f and g are easily interpreted, but they are
awkward quantities to predict. One is forced to assume that the or­
ientation distribution is symmetric, Eq. (96), the direction of maxi­
mum orientation must be found separately, and f and g related to this
direction. For predictive calculations a coordinate-independent quan­
tity is much more useful. One set of quantities which fit this require­
ment is that of tensors which describe the orientation distribution [82].
To define these tensors, first denote the direction of any fiber by the
unit vector p. In this notation 4 ^(cj)) becomes ip(p), and Eq. ( 8 6 ) be­
comes

filj(p ) dp (97)

Using Pi to denote the scalar component of p in the xi coordinate di­


rection, we can define the second-order orientation tensor by

ay = /p.p. ^ (p ) dp (98)

where i and j can take on any of the values of 1, 2, or 3 (1 or 2 for


planar orientations). The fourth-order orientation tensor is

^iki = ^PiPfkPi^(p) <ip (99)

and higher even-order tensors can be defined analogously. There are


no odd-order tensors because is an even function. That is.

^ (p ) = ^ (-p ) ( 100)
These tensors have properties which include symmetry

a.. = a.. (101)


1] ]i
Compression Molding of Polymers 559

and an equivalent of the normalization condition [Eq. ( 8 6 )]


3
( 102 )
i=l

Because of the symmetries of ay and there are a limited number


of independent components. In a two-dimensional problem (planar or­
ientation distribution) there are only two independent components of
aij, and the combination of ayj^j and ay have but four independent
components. Thus, the orientation tensors provide a compact descrip­
tion of the orientation state, which is easily transformed into other co­
ordinate systems. One can also derive equations for the evolution of
these tensors with time from a model of fiber orientation such as Eq.
( 8 8 ) [37,82].
The tensors defined in Eqs. (98) and (99) appear in rheological
models for fiber suspensions [38,83] and in mechanical property the­
ories for short-fiber composites [68,70]. They seem to be intimately
related to many features of composite materials. Because of this and
because of their efficiency in describing orientation, they may well
become the primary variables used to predict fiber orientation in mold­
ed parts.

E. F ib e r Damage D u rin g Flow

Many experiments have established the fact that fibers are damaged,
usually being broken into shorter lengths, during processing [74—78].
Compounding, extrusion, and molding can all degrade the length of
reinforcing fibers, causing a substantial drop in the strength and
stiffness of the material. Relevant questions are: What is the mech­
anism of damage, and how can it be avoided?
The dominant mechanism of fiber breakage seems to be elastic buck­
ling during shear flow. This phenomena was first analyzed by Forgacs
and Mason [78] and has been verified for single glass fibers by Salinas
and Pittman [79]. As discussed previously, fibers rotate in a simple
shear flow. At the position cj) = +45° (refer to Fig. 33) the fiber is in
tension, whereas at cf) = -45° the fluid acting on the fiber puts it in
compression. A single filament in this position can be analyzed as a
column under compression by elastic buckling theory. The result is
that buckling occurs at a critical shear stress in the fluid, given by
[79]

E^[ln(2 1/d) - 1.75]


(103)
crit 2 ( 1 /d)“
560 T ucker

for fibers lying in the plane of motion. Here Ef is the elastic modulus
of the fiber. The important feature of this result is that bulk shear
stress determines whether or not the fibers break. For common val­
ues of 1 /d for reinforced plastics the critical stress varies little with
fiber length. This is borne out by the common experience of proces­
sors: "No matter what I do, the fibers always come out the same
length."
The effect of fiber volume fraction on fiber breakage has been
studied by vonTurkovich and Erwin [77]. They found that glass fi­
bers in polystyrene had the same length distribution at the exit of an
extruder for volume fractions ranging from 1% to 20%. This suggests
that abrasion between fibers is unimportant, and that even in highly
concentrated suspensions fibers are free to rotate. These workers
identified the melting zone as causing the most damage to fibers in the
extruder. This is reasonable since extruder melting zones involve
shearing a thin polymer film (high shear rate) at temperatures just
above melting (high viscosity), generating some of the highest shear
stresses possible. A rough estimate [77] places this stress level at 3 x
10® Pa, and Eq. (103) suggests that 10-ym-diameter glass fibers could
only have 1/d about 10 and resist this stress. Longer fibers will buckle
and break at this stress. Fibers with 1/d of 100 buckle at a stress of
approximately 10® Pa. Given this dramatic effect, there is no easy cure
for fiber degradation during processing. Reducing the stress level
helps, but a reduction of two orders of magnitude or more is needed
to have a significant effect. High-performance fibers (carbon, aramid)
with their higher stiffness and strength should be more resistant to
breakage, but at present the "high stress" process of plasticating
extrusion and injection molding seem limited to short fiber lengths.

IV . SUMMARY AND C O N CLUSIO N S

Many of the key problems in compression molding of thermosets and


composites are related to the fluid flow, heat transfer, and curing
that take place in the mold. Significant developments in modeling
and analysis of these phenomena have been made and will contribute
to the solution of the key problems.
For mold-filling flow the generalized Hele-Shaw model provides an
efficient and accurate representation for thin parts. The model does
have limitations, and these are now clearly identified. When the Hele-
Shaw approximation is not valid, more general finite element models
can be used, albeit at a higher computation cost. Automated numeri­
cal schemes for finite element mold-filling simulations have been de­
veloped and are being improved. Mold-filling simulations provide in­
formation about knit line locations and possible unfilled regions of the
Compression Molding of Polymers 561

mold, and are the starting point for heat transfer-curing and fiber
orientation calculations.
Along with better tools for analyzing mold filling has come better
equipment for molding parts. Accurate control of press parallelism
and closing speeds improves the repeatability of the process as well
as the correspondence between computer simulations and reality.
Heat transfer and curing analyses can be used to minimize cycle
time without degrading the part from high temperatures. Changes in
mold temperature, part thickness, and even resin and catalyst formu­
lation can be explored and optimized with these simulations. More
work remains to be done on the effect of mold-filling flow on heat
transfer and curing.
Uneven heat load on the mold surface is now recognized as a major
problem. By properly designing the mold heating system, one can
minimize variations in the mold surface temperature and reduce the
time to mold and cure a part.
The source and importance of residual stresses in thermoset poly­
mers is not well understood. This stands as an area needing further
investigation, as does the rheology of fiber-reinforced molding com­
pounds.
Fiber orientation is an area that has seen major advances. For the
first time, the effect of flow on fiber orientation can be predicted in
nontrivial geometries with useful accuracy. This is a promising area
needing further development. Future work is likely to center on ef­
ficient descriptions of the fiber orientation state, better numerical
schemes for predicting orientation, and extending comparisons between
predictions and experiments.
Although each of these types of analysis answers specific useful
questions, even greater potential will be realized when all of the com­
ponents are integrated into a unified system. One such set of pro­
grams has already been developed [80], and others will doubtless fol­
low. Eventually, designers will be able to specify the shape of the
part, simulate mold filling and curing, calculate the fiber orientation
pattern, develop detailed mechanical property data for the p art, and
perform a stress analysis to see if the part performs as required. If
so, they can build the tooling with confidence; if not, they can change
the design, material, or processing conditions until a suitable part is
found. This type of development should lead to improved quality and
productivity for compression-molded parts.

ACKNOW LEDGMENTS

My own work in this field has been carried out in conjunction with
many fine graduate students at the University of Illinois at Urbana-
562 T ucker

Champaign: Suresh Advani, Peter Brown, Che-Yang Chen, Francisco


Folgar, William Jackson, Ching-Chih Lee, Mark Lovrich, Tom Osswald,
and Martin Zentner. Financial support for the work was provided by
the National Science Foundation, Owens-Corning Fiberglas Corpora­
tion, and the Ford Motor Company. I also wish to thank many friends
in industry who have provided ideas, comments, criticism, perspec­
tive, and support. They include Dr. Douglas Denton, Dr. Richard
Griffith, Dr. George Hartt, Mr. Ed Herman, Dr. Alan Isham, Dr. Carl
Johnson, Dr. Stuart Munson-McGee, Dr. O. Narayanaswamy, Dr.
Seymour Newman, Mr. Don Sage, Dr. Sandy San zero, and Dr. Jim
Stevenson.

NOTE ADDED IN PROOF

Since the manuscript was written a new formulation for the flow of SMC
during mold filling has appeared: M. R. Barone and D. A. Caulk, J,
AppL Mech., 53: 361 (1986). This model contains some features of
the anisotropy observed in SMC and, roughly speaking, interpolates
between the generalized Hele-Shaw model and the lubricated squeezing
flow model. Early comparisons with experiments are promising.
In the context of the Hele-Shaw model there has also been consid­
erable progress with numerical methods for mold filling, including a
boundary element method for thin, flat parts and a finite element/con­
trol volume simulation for thin parts with three-dimensional shapes.
See T. A. Osswald and C. L. Tucker, ”A Boundary Element Simula­
tion of Compression Mold Filling,” Polym. Eng. Sci. , to appear, and
T. A. Osswald, ’’Numerical Methods for Compression Mold Filling Sim­
ulation,” Ph.D. Thesis, Department of Mechanical and Industrial En­
gineering, University of Illinois at Urbana-Champaign, Urbana, 111.,
1986.

REFERENCES

1. D. L. Denton, Proc. SPI RP/C Inst., 16-A (1981).


2. C. A. Heiber and S. F. Shen, J. Non-Newt. Fluid Mech., 7:
1 (1980).
3. H. Schlichting, Boundary-Layer Theory, McGraw-Hill, New
York, 1968, p. 114.
4. C, L. Tucker and F. Folgar, PoZym. Eng. Sci., 23: 69 (1983).
5. C .-C . Lee, F. Folgar, and C. L. Tucker, J. Eng. Industry,
106: 114 (1984).
6. R. J. Silva-Nieto, B. C. Fisher, and A. W. Birley, Polym.
Compos., 1: 14 (1980).
Compression Molding of Polymers 563

7. P. F. Brown, M.S. Thesis, Department of Mechanical and In­


dustrial Engineering, University of Illinois at Urbana-Cham-
paign, Urbana, 111., 1983.
8. M. R. Barone and D. A. Caulk, Int. J. Heat Mass Transfer,
22: 1021 (1979).
9. E. A. Herman, Plast, Eng., 36: 31 (1980).
10 . D. Andersen and A. Lawrence, Tech. Pap. No. 830487, Soc.
Automotive E n g rs., 1983.
11 . L. Marker and B. Ford, Proc. SPI RP/C Inst., 16-E (1977).
12 . J. K. Gillham, Polym. Eng. Sci., 19: 67 (1979).
13. M. R. Kamal and S. Sourour, Polym. Eng. Sci., 13: 41 (1976).
14. S. Sourour and M. R. Kamal, Thermochim. Acta, 14: 41 (1976).
15. J. F. Stevenson, SPE Tech. Papers, 26: 452 (1980).
16. J. F. Stevenson, Polym. Process Eng., 1: 203 (1983-84).
16a. J. F. Stevenson, Proc. First Int. Conf. on Reactive Polymer
Processing, Pittsburgh, Pa., 1980.
16b. J. F. Stevenson, Proc. Second World Congress on Chem. Eng.,
Montreal, 1981, p. 447.
17. L. J. Lee, Polym. Eng. Sci., 21: 483 (1981).
18. D. F. Boucher and G. E. Alves, Chem. Engr. Progr. , 55: 55
(1959).
19. M. R. Barone and D. A. Caulk, Polym. Eng. Sci., 21: 1139 (1981).
20 . A. I. Isayev and D. L. Crouthamel, Polym. Plast. Technol. Eng.,
22: 177 (1984).
21 . A. I. Isayev, Polym. Eng. Sci., 23: 271 (1983).
22 . L. C. E. Struik, Polym. Eng. Sci., 18: 799 (1978).
23. Z. Tadmor, J. Appl. Polym. Sci., 18: 1753 (1974).
24. B. D. Aggarwala and E. Saibel, Phys. Chem. Classes, 2: 137
(1961).
25. R. Muki and E. Sternberg, J. Appl. Mech., 28: 193 (1961).
26. E. H. Lee, T. G. Rogers, and T. C. Woo, J. Amer. Ceramic
Soc., 48: 480 (1965).
27. O. Narayanaswamy and R. Garden, J. Amer. Ceramic Soc., 52:
554 (1969).
28. B. W. Shaffer and M. Levitsky, J. Appi. Mech., 41: 652 (1974).
29. M. Levitsky and B. W. Shaffer, J. Appi. Mech., 41: 647 (1974).
30. M. Levitsky and B. W. Shaffer, J. Appi. Mech., 42: 651 (1975),
31. S. J. Lee, M. M. Denn, M. J. Crochet, and A. B. Metzner,
J. Non-Newt. Fluid Mech., 10: 3 (1982).
32. D. L. Denton, Owens-Corning Fiberglas Technical Center,
Granville, Ohio, personal communication.
33. C . -C . Lee and C. L. Tucker, SPE Tech. Papers, 29: 740 (1983).
34. C .-C . Lee and C. L. Tucker, ’’Flow and Heat Transfer in
Compression Mold Filling,” J. Non-Newt. Fluid Mech., to ap­
pear.
564 T ucker

35. L. J. Lee, L. F. Marker, and R. M. Griffith, Polym. Comp,, 2:


209(1981).
36. J. L. Ericksen, Roll. Zeit., 173: 117 (1960).
37. G. L. Hand, J, Fluid Mech., 13: 33 (1962).
38. S. M. Dinh and R. C. Armstrong, J. Rheology, 28: 207 (1984).
39. J. R. Scott, Trans. Inst, Rubber I n d ., 7: 169 (1931).
40. G. Brindley, J. M. Davies, and K. Walters, J. Non-Newt. Fluid
Mech., J: 19 (1976).
41. K. S. Gandhi and R. Burns, Trans, Soc, Rheology, 20: 489
(1976) .
42. R. O. Maschmayer and C. T. Hill, Adv, in Chem. Series, 134:
95 (1974).
43. P. J. Leider, Ind. Eng. Chem. Fund., 4: 342 (1974).
44. R. I. Tanner, ASLE Trans., 8: 179 (1965).
45. J. M. Kramer*, Appi. Sci. Res., 30: 1 (1974).
46. N. Phan-Thien and R. I. Tanner, J. Fluid Mech., 129: 265
(1983) .
47. P. J, Leider and R. B. Bird, Ind. Eng. Chem. Fund., 13: 336
(1974) .
48. R. J. Silva-Nieto, B. C. Fisher, and A . W. Birley, Polym. Eng.
Sci., 21: 499 (1981).
49. M. Zentner, M.S. Thesis, Department of Mechanical and Indus­
trial Engineering, University of Illinois at Urbana-Champaign,
Urbana, lU. , 1984.
50. M. Zentner and C. L. Tucker, University of Illinois, Urbana,
111., unpublished research.
51. G. B . Jeffery, P. Roy. Soc., A 102: 161 (1923).
52. S. G. Mason and R. St.J. Manley, P. Roy. Soc., A 238: 117
(1957).
53. F. P. Bretherton, J. Fluid Mech., 14: 284 (1962) .
54. J, B. Harris and J. F. T. Pittman, J, Coll. Inter. Sci., 50: 280
(1975) .
55. P. A. Arp and S. G. Mason, J. Coll. Interf. Sci., 59: 378
(1977) .
56. W. I. Lee and G. S. Springer, J. Reinf. Plast. Comp., 1: 279
(1982).
57. M. Folgar, Ph.D. Thesis, Department of Mechanical and Indus­
trial Engineering, University of Illinois at Urbana-Champaign,
Urbana, 111., 1983.
58. F. Folgar and C. L. Tucker, J. Reinf. Plast. Comp., 3: 98
(1984) .
59. W. C. Jackson, S. G. Advani, and C. L. Tucker, J. Compos.
MaVls., 20: 539 (1986).
60. W. C. Jackson, F. Folgar, and C. L .. Tucker, Polymer Blends
and Composites in Multiphase Systems, C. D. Han, E d ., Adv.
Chem. Ser. No. 206, American Chemical Society, Washington,
D. C . , 1984, p. 279.
Compression Molding of Polymers 565

61. C .-C . Lee, University of Illinois, Urbana, 111., unpublished re ­


search.
62. R. C. Givler, in Transport Phenomena in Materials Processing,
AS ME, New York, 1983, p. 99.
62a. R. C. Givler, M. J. Crochet, and R, B. Pipes, J. Compos»
MaVls., 17: 330 (1983).
63. R. C. Givler, Ph.D. Thesis, Department of Mechanical and Aero­
space engineering. University of Delaware, Dewark, Del., 1983.
64. M. W. Darlington and P. L. McGinley, J. Mater» Sci», 10: 906
(1975).
65. M. W. Darlington, P. L. McGinley, and G. R. Smith, J» Mater»
Sci», 11: 877 (1976).
66. M. W. Darlington, B . K. Gladwell, and G. R. Smith, Polymer,
18: 1269 (1977).
67. D. L. Denton and S. H. Munson-McGee, ASTM STP 873, D. Wil­
son, E d., ASTM, Philadelphia, 1985, p. 23.
68. S. H. McGee, Ph.D. Thesis, Department of Chemical Engineer­
ing, University of Delaware, Newark, Del., 1982.
69. S. H. McGee and R. L. McCullough, J. AppL Phys», 55: 1394
(1984).
70. R. L. McCullough, in Treatise on Materials Science and Technol­
ogy, Vol. 10, Part B, Academic Press, New York, 1977.
71. R. B. Pipes, R. L. McCullough, and D. G. Taggart, Polym»
Compos», 3: 34 (1982).
72. P. H. Hermans, Contributions to the Physics of Cellulose Fibers,
Elsevier, New York, 1946.
73. R. Schweizer, Proc» SPI RP/C Inst» (1978).
74. R. Burns and D. Pennington, Proc» SPI RP/C Inst» (1979).
75. R. J. Crowson and M. J. Folkes, Polym» Eng» Sci», 20: 934
(1980).
76. L. Czarnecki and J. L. White, J. AppL Polym» Sci., 25: 1217
(1980) .
77. R. vonTurkovich and L. Erwin, Polym» Eng. Sci», 23: 743
(1983).
78. O. L. Forgacs and S. G. Mason, J» Coll» Interf. Sci», 14: 457
(1959).
79. A. Salinas and J. F. T. Pittman, PoZym. Eng» Sci», 21: 23
(1981) .
80. J. J. Quigley and R. B . Pipes, Plast» Polym» Tech» Eng», 20:
103(1983).
81. M. L. Lovrich and C. L. Tucker, S»P»E» Tech» Paper, 31,
1119 (1985).
82. C. L. Tucker and S. G. Advani, The Use of Tensors to De­
scribe and Predict Fiber Orientation in Short Fiber Compos­
ites, unpublished manuscript.
83. G. G. Lipscomb, R. Keunings, G. Marrucci, and M. M. Denn,
Proc» Int» Cong» Rheology 1984.
8
Design of Mold Cooling System

Kamar J . S in g h *
Corporate Research and Development
General Electric Company
Schenectady, New York

1. IN T R O D U C T IO N

Until recently the task of mold design had been an art which, unfor­
tunately, was a specialty of few mold designers. Even the standard
of this art was not very refined since it was based on a long experi­
ence of failures and successes—more of the former than the latter.
Now the development of efficient computer software and high-speed
computers is transforming this art of mold design into a complete sci­
ence. There are numerous advantages from computer-aided analysis
and design which help to

Optimize the location of cooling lines


Produce defect-free parts
Design the mold before the drawings are finalized
Increase productivity
Reduce time from part design to production
Provide the production rate far in advance of production, aiding
the planning of machines and the scheduling
Eliminate the time wasted in modifying the mold when it does not
perform as expected

These advantages can be transformed into tremendous savings to


the manufacturer because of the reduction in the overall cost of the
part. The time saved from part design to actual production can some­
times mean the success of the project. Delays can be eliminated which

*Current affiliation: Aircraft Engine Business Group, General Electric


Company, Cincinnati, Ohio.

567
568 Singh

might have caused missed deadlines and lost markets. Moreover, pro­
ducing a defective part may also place the manufacturer in the same
unpleasant situation. Further, the knowledge of the production rate
helps in planning the resources needed in the plant. These resources
can be the number of machines and operators needed, the amount and
pressure of coolant needed to cool the molds, or the need, if any, for
robots to accelerate the cycle time.
Faultless design of molds, therefore, becomes very important for
the manufacturer. The availabiLity of computer software makes the
task very easy for the designers who can sit at a computer terminal
and complete the entire mold design in a matter of hours. Mathemati­
cal simulation of the process forecasts the future quality of the mold
and any shortcomings. With the present technology the computer can
plot all the results of mold performance so even a layman can under­
stand it easily.
Computer software for the mold design has to account for the heat
transfer from the hot plastic into the mold and, further, from the
mold into the coolant which flows under pressure in the cooling lines.
The faster the heat is removed from the plastic, the sooner the mold
can be opened to eject the part and, therefore, make it available for
the next shot of molten plastic. The duration of the cycle is conse­
quently dependent upon the efficiency of the cooling system.
Usually, the locating and sizing of the cooling lines in the mold are
based on the whim of the designer. He or she is typically unaware of
the relative advantages of different sizes and locations of the cooling
lines. If a computer program is available to the designer for optimi­
zing these cooling line parameters, experience and analysis can then
go hand in hand in increasing the productivity of the part.
The defects in plastic parts are generally due to a defective cool­
ing design. The defects can be in the form of a deformed part due
to warpage or residual stresses. Deformed parts coming out of a mold
may be useless. In that case, a mold may need redesigning to take
care of the problem. A part with high residual stresses may look nor­
mal from the outside, but, once put to use, it may crack at the loca­
tion where the stresses are locked in. These defects are very com­
mon in real life. Fortunately, with correct mold design they can be
avoided rather easily.

II. HEAT T R A N S FE R IN MOLDS

Before discussing mold design concepts, it is pertinent for us to study


the different modes of heat transfer and their relative effects on the
overall cooling of the plastic part. The usual design practice is to
provide cooling channels in the mold at locations convenient to the
designer. The size of these channels is arbitrarily chosen based on
Design of Mold Cooling System 569

common practice in the industry. This approach may be called a one-


shot design process. There is no effort made to improve the design
based on heat transfer fundamentals. The reason for this is the lack
of computer software usage as a design tool.
The cooling of a plastic part from the injection temperature to the
ejection temperature, the temperature when it is cool enough for the
mold to open, is a function of the rate of heat transfer through the
plastic, from the plastic to the mold, through the mold, and there­
after to the coolant in the cooling lines or to the atmosphere. O b­
viously, the modes of heat transfer are conduction, convection, and
radiation. The object, therefore, is to improve these modes of heat
transfer wherever possible so that the cooling time is as short as pos­
sible.
The cooling time of plastic is a function of the following variables:

Part thickness
Part shape
Plastic properties
Injection and ejection temperatures of plastic
Mold material properties
Number, location, and size of cooling lines
Coolant properties
Coolant flow rate and temperature

It is important to list these variables at this stage for the sake of


understanding their impact on the cooling process. Some of these pa­
rameters are constant variables. For example, thickness and shape
of part, properties of plastic, and injection temperature are already
fixed at the time of the mold design process. The relationship of
these variables to the mode of heat transfer provides a basis for the
ultimate design of molds.

A. Conduction

One mode of heat transfer is the process of molecular transport of


heat in a body or between bodies due to temperature variation in the
domain considered. The conduction can be steady state, where the
temperature of a body remains constant in time, or it can be in the
transient state, in which case the temperature in the domain not only
varies from point to point but also with time. Transient conduction
takes place in plastic- and glass-parts manufacturing, rubber vulcan­
izing, and metal casting.
For unidirectional conduction of heat, Fourier [1] derived the law

= kA Ì - - Ì ( 1)
dt \ dx /
570 Singh

where dQ/dt is the instantaneous rate of heat flow, k is the thermal


conductivity of the material, A is the section area and -dT/dx is the
temperature gradient. For a steady-state case, dQ/dt is independent
of time and is normally referred to as q. Equation (1) can then be
written as

q = -kA ( 2)
dx

The rate of heat flow by conduction is directly proportional to the


thermal conductivity of the material. Use of a mold material with high­
er thermal conductivity does not necessarily mean the plastic part will
cool faster since the cooling of plastic is also a function of its own
thermal properties. Thermal diffusivity, the property of a material
which determines the rate at which it wül gain or lose heat, is the
limiting factor in determining the cooling time of a part. Thermal dif­
fusivity, a, is given by the relation

k
(3)
pc

where p is the density, Cp is the specific heat, and k is the thermal


conductivity of the material. From this it follows that the larger the
value of a, the higher will be the rate of change of temperature at
any point in the body. Table 1 shows the values of thermal conduc­
tivity, specific heat, and thermal diffusivity for some common unfilled
plastics and mold materials. In Section III the behavior of plastic cool­
ing will be discussed in more detail.

B. Convection

Another mode of heat transfer in plastic molds is heat transport oc­


curring through movement of the macroparticles of the coolant flow­
ing in the cooling lines. There is a combined action of convection and
conduction taking place in the fluid. The convection between the
coolant and the surrounding mold will be referred to here as heat
transfer by convection.
The calculation of convective heat transfer is based on Newton^s
law

q = h (T -T )A (4)
m e

where q is the heat flow rate, h is the convective heat transfer coef­
ficient, Tm is the temperature of the mold, T q is the temperature of
Design o f Mold Cooling System 571

T a b le 1 Thermal Properties of Materials

k Cp a
Material (W/km) (kJ/kg K) (m^/hr X 10^)

Plastics
ABS 0.14-0.353 1.26-2.10 0.480
Acetal 0.23-0.37 1.68 0.456
Acrylic 0.18-0.28 1.26 0.557
Epoxy 0.17-1.23 0.84-1.26 1.320
Nylon 0.17-0.25 1.68 0.398
Phenylene oxide 0.20-0.22 1.26 0.555
Polypropylene 0.10-0.17 2.10 0.254
Polycarbonate 0.18-0.20 1.26 0.452
Polyester 0.17-0.30 1.26-2.10 0.489
Polyethylene 0.30-0.51 2.10-2.52 0.681
Polystyrene 0.10-0.14 1.26 0.312
Vinyl, PVC 0.14-0.21 0.84-1.68 0.389

Mold material
Aluminum 130.0 0.924 187.0
Beryllium copper 151.0 1.890 155.0
H13 29.5 0.462 29.4
P6 46.9 0.462 46.7
P20 36.5 0.462 36.3
S7 24.3 0.462 24.2
414SS 25.0 0.462 25.1
420SS 25.0 0.462 25.1

the coolant, and A is the area over which the convective heat transfer
takes place.
The heat transfer coefficient, h, is a variable dependent upon the
physical properties, velocity, and temperature of the fluid, shape and
size of the surface, and the nature of fluid flow. There are two kinds
of convective heat transfer: free or natural, and forced. Natural
convection is created by motions due to heterogeneity of the mass
forces acting upon the fluid, whereas forced convection takes place
due to motion from external kinetic energy, as from pumps or fans.
Natural convection takes place on the external surface of the mold
where the surface is cooled by the ambient air. On the other hand,
the flow of coolant in the cooling lines provides the forced-convection
cooling of the mold.
572 Singh

There is no easy way to determine the value of h, the convective


heat-transfer coefficient. A common method of evaluating h is by di­
mensional analysis. Using this approach for flow in a pipe, we can
determine independent dimensionless parameters from the physical
variables of the problem. These are

1. The Nusselt number

XT ™
Nu = —
k

where D is the diameter of the pipe and k is the thermal con­


ductivity of the fluid.
2. The Reynolds number

R e = ^

where p is the density of the fluid and ]i is dynamic viscosity


of the fluid.
The Prandtl number

Pr - Vk
where Cp is the specific heat of the fluid.

The functional relationship can be written as

Nu = F(Re, P r) or h = — F(Re, Pr)

Experimental data can now be correlated to obtain the relationship


between these variables. A number of investigators have done exten­
sive work to calculate the value of h. Depending upon the Reynolds
number of flow, different empirical equations are valid for laminar,
transient, and turbulent flows of coolant.
There is a distinct difference in heat flow by convection if the
flow is turbulent on one hand and laminar on the other. For example,
the effect of cooling line length is greater for laminar flow than for
turbulent flow.
Laminar or streamline flow is generally demarcated by a Reynolds
number less than 2300. The empirical relationship for laminar flow
[2] is
Design of Mold Cooling System 573

1, ,0.1^
D\^/^
Nu = i.se/— ] Pr (5)
*'■(?)
where is the bulk viscosity of the fluid, yw is the viscosity of the
fluid at the wall temperature, and L is the heated length of cooling
line.
In laminar flow there is no mixing of the warmer and cooler fluid
particles and the heat transfer takes place mainly by conduction
through the liquid. Therefore, the heat transfer coefficient in this
region is relatively small.
The transient region lies between laminar and turbulent flows and
is bounded by Reynolds numbers of 2300 and 10,000. The common
empirical formula [3] for calculating the heat transfer coefficient is

Nu = 0.116(Re 2 / 3 125) Pr 1 / 3 1+ ( 6)

Because this relationship is very sensitive to the L/D ratio, the


relationship between laminar and turbulent regions is not smooth for
large values of L/D. The flow pattern in the transient region is not
as well defined as that for laminar or turbulent regions, although it
is known that some mixing of warm and cold fluids takes place in this
region.
In the turbulent region the flow has a Reynolds number greater
than 10,000. The commonly used empirical formula [4] for heat trans­
fer is

Nu = 0.023 Re°*® Pr 1 / 3 (7)

At a high Reynolds number the flow inside the cooling line is tu r­


bulent except for a very thin layer near the wall surface. In this
layer the turbulence is damped by viscous forces, and the heat flow
near the surface is mainly by conduction. Because of the turbulent
eddies, there is very effective heat transfer and rapid cooling of the
mold surface. An increase in the flow velocity, which produces tur­
bulent eddies, is accompanied by an increase in frictional pressure
drop in the cooling line. Therefore, a bigger pump is needed to main­
tain the flow. Whenever possible, however, laminar flow should be
avoided because of the low heat transfer coefficient in this region.
The Pr an dtl number, which appears in all these equations, is a
function of fluid properties alone. It relates the temperature dis­
tribution to the velocity distribution. For a given Reynolds number.
574 Singh

the temperature gradient is steeper in a fluid with large Prandtl num­


ber and, therefore, provides better heat transfer.
Heat transfer in noncircular cooling lines can be determined by use
of an equivalent cooling line of a circular cross section with a diameter

( 8)
eq P

where A is the cross section and P is the perimeter of the noncircular


cooling line. This method of using the equivalent diameter or hydrau­
lic diameter is only approximate, and it is recommended only for tu r­
bulent flows.
For cooling lines with annular cross section the coefficient of heat
transfer can be calculated from the equation [5]

Nu = 0.017 Re°‘ ® Pr^ (9)

where Pr^ is the Prandtl number at the wall temperature, di is the in­
ner diameter, and d 2 is the outer diameter of the cooling line. How­
ever, a limitation on this relationship is that do/di should be between
1.2 and 1.4.

C. Radiation

The phenomena of conduction and convection, as seen earlier, are af­


fected primarily by temperature difference and very little by tempera­
ture level. In radiation the heat transfer increases rapidly with in­
crease in temperature level. It is immediately obvious that for low
temperatures, conduction and convection are dominant; for high tem­
peratures radiation is the controlling factor. In particular, radiation
from the external surface of the mold to the atmosphere is not signifi­
cant because the temperature of the mold surface is not very high.
The rate of heat transfer from a gray body to its surroundings, as
applicable to the mold, is given by

q = a A F i ( T i “ - T a “) ( 10 )

Where a is the Stefan-Boltzmann constant with a value of 5.669 x 10”®


W/m^-K“, A is the surface area, T i is the temperature of the body, Ta
is the temperature of the surrounding medium in Kelvin, and is the
emissivity factor.
Keeping these basic concepts of heat transfer in mind, we can now
study the actual thermal history of plastic through a complete cooling
cycle.
Design of Mold Cooling System 575

111. C O O LING OF P LA S TIC

It is imperative for the mold designer to understand the cooling pat­


tern through the thickness of the plastic. The thermal history, if
known, can help in deciding when the mold should be opened for
ejecting the part. If the mold is opened before the temperature in
the plastic reaches a certain minimum value, the part is susceptible
to mechanical deformation which may not be acceptable. On the other
hand, if the part is allowed to cool more than necessary, the extra
time thus spent is wasted and would decrease the productivity of the
machine.
Cooling of plastic in a mold is due to heat transfer by conduction
through the plastic and thereafter by conduction through the mold.
The contact temperature between the mold and the plastic is a func­
tion of the thermal properties of the mold and plastic materials. As
discussed earlier, thermal diffusivity of plastic is much lower than
that of the mold material. Therefore, the overall cooling is mainly a
function of the rate at which the plastic would transfer heat from its
inner region to the outer surface.
The cooling characteristics of plastic can be best understood if we
first study the thermal distribution in a plastic plate being cooled in­
side a mold without any cooling line and then inside the mold with an
extremely efficient cooling-line system. The actual process would lie
somewhere between these two cases. In particular, the second case
would provide the partes theoretical minimum cooling time (TM CT),
which would be mainly a function of the plastic properties and part
thickness.
But before we study these two cases, we must define a parameter
known as the Biot number. It is a dimensionless group which repre­
sents the ratio between internal thermal resistance to conduction and
external thermal resistance to convection :

d/k ^ M
Bi = ( 11)
1/h k

where d is the half-thickness of the plate, h is the heat transfer co­


efficient and k is the thermal conductivity of plastic material.
The temperature distribution in a plate being cooled depends on
the Biot number. If the initial temperature of the plate is T q, the
cooling pattern at time t^, t 2 , . .. will be given by curves T^, T 2 ,
T3, . . . , respectively, as shown in Fig. 1. All these curves seem to
pass through a common point at a distance H from the plate surface.
In fact.

H= ( 12 )
Bi
576 Singh

Fig. 1 Temperature variation in a plate during cooling.

If Bi is too small (less than 0.5), as for metals, the value of H will
be very large. Therefore, the whole thickness of the plate will cool
at almost a uniform temperature. On the other hand, if Bi is too large
(greater than 50), as in plastics, the value of H will be very small and
the temperature distribution during cooling will be as shown in Fig. 2.

Fig. 2 Temperature variation in plastic plate during cooling.


Design of Mold Cooling System 577

At = 7.8 SECONDS

Fig. 3 Cooling of a polycarbonate plate—without cooling lines.

Consider the thermal history of a plastic plate being cooled in a


mold without cooling lines. A computer program [6] is used here to
calculate and plot the transient results for a complete cooling cycle.
As an example, a 0.5-cm thick polycarbonate plate is analyzed with
this program. For simplicity the thermal properties of the plastic are
taken as temperature independent. The effects of pressure and shrink­
age are also neglected. The initial temperature of the melt is taken as
282°C. Figure 3 shows the cooling pattern with the initial temperature
of the mold as 37.7°C. The program stops calculations as soon as the
maximum temperature at all points within the plate falls below a given
limiting value. In this example this limiting value is 65°C. For this
data the cooling time obtained from the program is 54 sec.
If cooling lines are now introduced in the mold at a distance of 1.25
cm from the plastic with coolant temperature of 4°C, a new cooling pat­
tern develops, as shown in Fig. 4. The time needed for the maximum
578 Singh

At = 6.4 SECONDS

Fig. 4 Cooling of a polycarbonate plate—with cooling lines.

temperature at all points to fall below 65®C is reduced to 40 sec. There


will be negligible improvement in this cooling time if the cooling system
is further improved, because of the relatively high thermal resistance
to heat flow in the plastic itself.
Note from these curves that there is a marked difference between
temperatures at the surfaces and at the center region of the plastic.
If surface temperature measurements are taken by thermocouples, it
should be realized that this would not be the true temperature of the
plastic because the inner region would still be at a much higher tem­
perature.
Computer software [6] can provide a good approximation of tem­
perature history and can help determine the TMCT for the plastic part.
Knowing the TMCT during the mold design process gives the designer
a fairly good idea of the minimum cooling time possible so that it can be
Design of Mold Cooling System 579

COOLING TIME SEC

Fig. 5 Thickness vs. cooling time for polycarbonate plate.

set as a goal of productivity. It will be impossible to have a mold with


cooling time equal to or less than TMCT. The calculation of TMCT de­
pends on the melt and ejection temperatures; it is the latter quantity
which is extremely difficult to measure or determine on the production
line.
For a given plastic the TMCT as a function of part thickness can
also be plotted by a computer software [6 ]. Figure 5 shows the TMCT
as a function of part thickness for polycarbonate, cooling from a melt
temperature of 282°C to an ejection temperature of 65°C. These types
of curves cannot be generalized in one plot because the ejection tem­
perature may vary from part to part, depending on its size and shape.
During these analyses we have not considered plastic shrinkage b e ­
cause what is happening at the interface of the plastic and the mold
during the cooling process is complex and not well understood. The
effect of the gap on the heat flow can be mathematically accounted for,
but it is difficult to verify. During shrinkage the contact may be main­
tained at some arbitrary regions. This can be a function of the geom­
etry, the thickness, and the location in the part.
580 Singh

IV . HEAT FLOW TH R O U G H MOLDS

If there are no cooling lines in the mold, heat transfer through the
mold can be approximated as unidirectional heat flow in a slab or
plate. The heat flow rate as given by Eq. (2) can be simplified to

-k A (x , T2 ) (13)

where Ti is the average temperature of the plastic , T 2 is the ambient


temperature, and L is the mold thickness in the direction of heat flow.
In Eq. (13), A/L is known as the conduction shape factor, S. It con­
tains only those variables that depend on the geometry of the mold.
As the plastic cools in the mold, decreases from the melt tempera­
ture to the ejection temperature. As an approximation, an average
value of Ti can be substituted for calculating the heat flow rate. How­
ever, to determine this average value, one has to know the ejection
temperature of the part. A better approach would be to numerically
solve the entire transient cooling cycle from the melt to the ejection
temperatures instead of using a steady-state model.
A computer program [7] has been developed which accounts for
the complete injection-molding cycle. Using transient thermal varia­
tion, it calculates the heat flow through the mold and the cooling lines.
Commencing from any given initial mold surface temperature, the pro­
gram undergoes a series of iterations till the temperature fluctuation
during any two consecutive cycles is identical. The program, there­
fore , does not need the ejection temperature as an input ; rather it is
an output parameter along with the temperature history of mold su r­
face. One can also obtain the temperature distribution along the thick­
ness of the plastic part from this program.
Without any form of cooling system, the mold will keep retaining
heat from cycle to cycle until it is too hot for molding parts. Cooling
channels must be provided in the mold to maintain the surface tem­
perature at an acceptable level.
If we introduce cooling lines in the mold, the rate of heat flow as
given in Eq. (13) will change, based on the shape factor of the new
geometry. Therefore, if we know the shape factors of different con­
figurations of cooling lines with respect to the plastic surface, we can
very conveniently find the rate of heat flow through the mold. Using
the boundary element method [8 ], the author has developed a library
of conduction shape factors. These geometries cover all the common
configurations found in molds, including inserts and curved parts as
shown in Fig. 6.
With cooling lines in the mold, T 2 will be the temperature of the
coolant in the cooling lines rather than the ambient temperature con­
sidered in Eq. ( 13). A major part of the heat will dissipate into the
Design of Mold Cooling System 581

cooling lines instead of the atmosphere. If the coolant temperature is


reduced, the heat flow rate will increase and thus improve the cycle
time. However, remember that (1) lowering the coolant temperature
may change the properties of the coolant and reduce the convective
heat transfer coefficient as discussed in Section VI, and (2) since
cooling the plastic is mainly a function of its thermal properties, re ­
ducing the coolant temperature beyond a certain value will not nec­
essarily reduce the cooling time.

V. COOLING SYSTEM
A. Cooling Lines
The basic cooling system in a mold is in the form of circular holes
drilled at convenient places so that the coolant flowing through these
passages will take the heat out of the hot regions close to the plastic
melt. For a plastic part which is not flat, the passages change direc­
tion to follow the configuration of the part. In such a case, for one
flow path, a number of holes are drilled and some of them are blocked,
forcing the coolant to flow through the desired circuit. An example
is shown in Fig. 7. For parts which have restricted regions where
circuits as shown in Fig. 7 are not feasible, use can be made of heat
pipes, bubblers, or baffles. All of these systems have projections
suitable for reaching the restricted areas. Figure 8 shows an example
where a heat pipe can be used very effectively.

B. Heat Pipes
The use of heat pipes is becoming more common. They are sealed,
round, or flat tubes filled with a fluid. The tube can be copper, alu­
minum, brass, nickel, mild steel, or any other suitable material. Many
fluids can be used, depending on the thermal range of application.
Some common fluids are water, ammonia, freon, pentane, or acetone.
The tube has a wick which transports the fluid in the liquid state from
the condenser end towards the evaporator end. At this end the liq­
uid evaporates, due to the heat, and changes into a gaseous state.
It travels back to the condenser end to repeat the cycle and thus con­
tinuously transports heat from one end to the other. At the conden­
ser end the heat is transferred to the coolant flowing in the cooling
lines, and the fluid condenses back to the liquid state. However, the
efficiency of the heat pipe depends on the angle of the axis of pipe
with respect to the horizontal axis. If the evaporator is below the
condenser, the efficiency of the heat pipe increases. It is at maxi­
mum efficiency when the heat pipe is vertical and the evaporator is at
the lower end. Obviously, the efficiency would be minimum if the evap­
orator is at the uppermost point. Variation in efficiency for the two
extreme cases can be about 400%.
582 Singh

1.
^ 7 .

WmMmii/MMMMMMP'
0 0 0 0 8-

0 O 0 0

J
0 0 0 0

Fig. 6 Geometry of conduction shape factors.

To calculate the amount of heat transferred through the heat pipe,


we need to consider the thermal resistance of the following: the heat
pipe tube at the evaporator end, the wick at the evaporator end, the
fluid throughout the pipe, the wick at the condenser end, and the heat
Design of Mold Cooling System 583

16.

13 . 17 .

14 . w
18 .

15 . 19 .

pipe tube at the condenser end, as well as the effect of angle of tüt
of the pipe.

C. B u b b lers

This system, also known as fountains, consists of concentric annuli


as shown in Fig. 9. The coolant flows through the inner tube and
584 Singh

COOLANT
IN

Fig. 7 Typical cooling line configuration.

returns via the annulus. For uniform flow of coolant, the internal di­
ameter Di should be

Di = 0.707D2 - t (14)

where t is the thickness of the inner pipe and D 2 is the diameter of


the drilled hole. For minimum pressure drop at the annular bend, the
distance S should be approximately 0.35D2.

Fig. 8 Heat pipe.


Design of Mold Cooling System 585

COOLANT
OUT

' J 02

COOLANT
IN

F ig . 9 Bubbler.

D. B affles

As the name suggests, in this system of cooling line, a baffle plate is


used inside a pipe to regulate the flow of coolant into the restricted
region as shown in Fig. 10. The baffle can be considered as two semi­
circular passages with the bend at the end. The effect of the bend on
the pressure drop is again a function of the clearance S.

COOLANT
OUT

SECTION A-A

IN

F ig . 10 Baffle.
586 Singh

Table 2 Thermal Properties of Water

Tempera­ Density Sp. heat Viscosity Therm, cond.


ture (°C ) (g/cm^) (cal/g °C ) (kg/m h r) (W/cm^C)

0.0 1.0023 1.0074 6.437 0.0055

20.0 1.0005 0.9988 3.590 0.0060

37.7 0.9946 0.9980 2.460 0.0063

65.5 0.9803 0.9990 1.560 0.0066

93.3 0.9627 1.0040 1.100 0.0068

V I. C O O LA NTS

Water is usually used to cool the mold. Due to its superior thermal
properties and availability, it is preferred over other coolants. Table
2 shows the properties of water as a function of temperature.
Convective heat transfer is a function of the coolant properties. As
seen from Table 2, the properties of water vary appreciably as the tem­
perature is changed. Consider the heat transfer coefficient for turbu­
lent flow as given by E q . ( 7). The equation can be written as

. 0 . 6 6 7 y O . 8 pO, 8 Q 0. 333

= 0.023 (15)
^ 0 . 2 ^ 0.467

If we keep the diameter and velocity constant and vary the tempera­
ture of water from 20° to 37.7°C, h would vary by +3.3% due to k,
-0.47% due to p, -0.26% due to C p , and +19% due to y. The overall
effect on the value of h would be +21%. It is obvious that as the tem­
perature of water increases, so does the convective heat transfer co­
efficient. However, the overall heat transfer from the mold to the
cooling line will decrease, as given by Eq. (4 ). Therefore, any de­
cision to change the coolant temperature has to be based upon the
overall heat transfer due to conduction and convection taking place
in the mold.
Design of Mold Cooling System 587

Another coolant commonly used is mineral oil. The normal working


range of oil is 0° to 200°C. It is usually used in closed circulating
systems and, consequently, has less problems.
Glycol mixed with water is sometimes employed for cooling molds.
It is primarily an antifreeze solution and is, therefore, used against
freezing in systems which are periodically subjected to low tempera­
tures .

VII. DESIGN OF CO O LING SYSTEM

As a part of the mold cooling design process, the designer has to ad­
dress the following design variables:

Location of cooling lines


Size of cooling lines
Type of cooling lines
Length of circuits
Flow rate of coolant

The location of cooling lines is not an independent decision. Be­


cause of other hardware in the mold plates, the drilling of cooling
lines goes hand in hand with these constraints. Therefore, the lo­
cation of cooling lines should be carried out at the initial stages when
the mold is being designed for ejection pins, angle pins, core pins,
cam slides, etc. In regions where a normal cooling line cannot be
placed, use of heat pipes, bubblers, or baffles must be considered.
If an insert is used in the mold, the usual practice is to locate the
cooling lines outside the insert. But, due to the high resistance to
heat flow through the interface of the inserts, the efficiency of cool­
ing lines decreases considerably. With the help of 0-rings the cool­
ing lines can be easily taken inside the insert and placed closer to
the hot plastic.
When deciding the number and size of cooling lines to be provided,
the designer must know the rate of coolant flow and the pressure
needed to maintain efficient heat transfer. The size of a cooling line
affects conduction of heat through mold and convection of heat through
the coolant. This can be best explained by a simple example.
A 20-cm-wide plate is to be cooled by a cooling-line system. We
will assume that there is no restriction from any ejection pins, bolts,
etc. Starting from three cooling lines of 1.0 cm diameter placed at
pitches of 10.0 cm and 2.0 cm away from the plastic, we keep adding
one cooling line at a time, thus reducing the pitch to 6.67, 5.0, 4.0
cm ,... for 4, 5, 6 ,... cooling lines, respectively.
588 Singh

NUMBER OF COOLING LINES

Fig. 11 Effect of diameter on heat flow rate by conduction.

If the mold has thermal conductivity of 25.0 W/km and the average
temperature difference between the plastic and the coolant is 100.O^C,
the heat flow rate in watts for a unit length can be plotted against the
number of cooling lines, as shown by curve A in Fig. 11. After the
number of cooling lines reaches 7 (pitch of 3.33 cm), the curve flat­
tens out. Therefore, the number of cooling lines to be provided
Design of Mold Cooling System 589

would be based on the optimum location on this curve. One has to


bear in mind, however, that the increase in number of cooling lines
would require a greater quantity of coolant. As the next design step,
if the diameter of these cooling lines is increased to 1.5 cm, the heat
flow rate through the mold would increase, and the results would be
as depicted by curve B .
The point of interest here is that though we have increased the
heat flow rate by increasing the diameter, there are two implications
of this. First we will need a much higher volume of coolant to flow
through the cooling lines if turbulent flow is to be maintained. Sec­
ond, there wül be a reduction in the convective heat transfer through
the coolant. The amount of this reduction depends on whether the
same velocity of coolant flow is maintained. If the flow rate is the
same, then the velocity will decrease, and the heat transfer coeffi­
cient for turbulent flow will be such that

1
Di.i (16)

That is, the value of h would be reduced by 52% because of an in­


crease in diameter from 1.0 cm to 1.5 cm. The heat flow rate, which
is proportional to hD, wiU consequently decrease by 27.7%.
On the other hand, if the flow rate is increased to maintain the
same velocity, then

(17)
D0 . 2

In this case h is reduced by 7.8%. However, the heat flow rate in­
creases by 38%.
Therefore, any change in the number and size of cooling lines
should be considered with respect to the overall heat transfer and
fluid flow variables involved. The variation in the shape factor value
of a cooling line due to modification of either the pitch, the diameter,
or the distance of cooling lines from the plastic surface, is depicted
in Fig. 12.
The connection of external hoses to cooling lines to form circuits
is another important design feature. As the coolant flows from the
inlet manifold, it experiences a pressure drop along the cooling line.
Darcy, Weisbach, and other investigators proposed the following re ­
lation for the head loss in pipes:

AP f/ J l) (18)
\2gD/
590 Singh

SHAPE
FACTOR

SHAPE
FACTOR

Fig. 12 Variation in conduction shape factor values.

where f is the friction factor, which depends on the smoothness of


pipe, the velocity, the diameter D, and the viscosity of the fluid. In
other words,
Design of Mold Cooling System 591

SHAPE
FACTOR

POINT A: X = 2 cm, P = 10 cm, D = 1 cm

(19)
("“• I )
where e is the mean height of the roughness of the pipe, If e is large,
then f will be large, so the pressure drop will be higher. Since there
is no closed-form solution for f, it is determined from experimental
results which have been compiled by investigators. The plots re­
lating the friction factor and Reynolds number for different rough­
nesses of pipes are called Stanton diagrams. From these diagrams it
becomes clear that in the laminar flow region the pressure drop is not
a function of the roughness of the cooling line. On the other hand,
in the turbulent flow region the pressure drop is more dependent
upon the roughness of the cooling lines than on the Reynolds num­
ber.
592 Singh

If a circuit is too long, there will be appreciable loss in pressure;


therefore the flow will not be turbulent. The circuits must be de­
signed in lengths appropriate for adequate pressure to be maintained
in the coolant. As the inlet pressure in cooling lines increases, so
does the flow rate and, consequently, the convective heat transfer
rate. The designer must look at the whole range of design variables
and choose the optimum parameters. If these optimum parameters are
maintained, the cycle time will be a minimum for those conditions.
The availability of a range of these coupled variables becomes a
necessity in the decision-making process. The only convenient method
for calculating and providing this information to the designer is by
computer software, which can evaluate the heat transfer and fluid
flow simulation of the process for a wide range of values and plot the
results for the designer. The selection of optimum parameters then
becomes a trivial task.

VIII. OPTIMUM PARAMETERS

We have looked at the heat transfer characteristics in the mold and


the coolant in cooling lines. The next step in the design process is
to evaluate the coolant flow rate or pressure needed so that the best
possible results are obtained for the capital investment in the equip­
ment, such as pumps, chillers, etc.
To evaluate this, one has to calculate results for a large range of
flow rates. For each discrete flow rate value, the corresponding pres­
sure drop, heat flow rate, and cooling time for the plastic can be cal­
culated. If the designer already has the equipment for the coolant
in the plant, the pressure rating of this equipment will determine the
coolant flow rate, heat flow rate, and the cooling time. A typical
plot showing the relationship between these parameters is shown in
Fig. 13.
Suppose a plant has cooling equipment with a pressure rating
P^. The actual process can then be designed for a pressure drop of
0.9Pi. The corresponding values of heat flow rate, coolant flow rate,
and cooling time are shown in Fig. 13 as H^, F^, and T i, respectively.
If new cooling equipment is to be installed, the designer has the
option of selecting the optimum parameters from these curves. Using
either the heat flow rate or the cooling time curve, one can select a
point where the payoff is maximum for a given coolant flow rate value.
From these two curves it is clear that as the coolant flow rate is in­
creased , the heat flow rate increases, and the cooling time decreases.
But, there is a marked change in the slope of these curves, and they
tend to become horizontal. Using the law of diminishing returns, we
can mark the optimum point as the coolant flow rate, F 2 , and corres­
ponding values of H2 , P 2 > and T 2 for the heat flow rate, pressure
Design of Mold Cooling System 593

CO O LAN T FLOW RATE

Fig. 13 Process parameter relationship.

drop, and cooling time, respectively. The pressure at the coolant


inlet manifold, therefore, should be approximately 1.1P2* To obtain
these curves, the designer has to have computer software which can
provide plots for a desired range of coolant flow rates or coolant pres­
sure drops. At the design stage the knowledge of predicted cooling
time is of profound value to the manufacturer. The entire planning
of personnel, scheduling, material requirements, and production can
be based on the forecast of the production rate of the part thus cal­
culated .

IX . DEFECTS IN PLASTIC PARTS

Defects in plastic parts can have several causes. Some defects are due
to improper filling and others are due to defective cooling. In this
section we discuss only those defects caused during the cooling pro­
cess.
594 Singh

A. Warpage

The most common defect created during the cooling process is warp­
age. On ejection from the mold, the part bends and deforms. The
part in that condition is unacceptable due to functional or aesthetic
requirements. Unbalanced cooling between the core and the cover
sides can cause this problem. If one side cools faster than the other,
it will freeze earlier and will undergo compression. The part will
bend away from the cooler side since it will be longer than the warmer
side.
The common remedy tried in plants is to provide ribs to eliminate
this problem. However, this introduces high stresses in the part.
Sometimes, the part deforms if subjected to a slightly warmer tem­
perature. Another remedy employed is to change the shape of the
mold so that when the part warps, it deforms towards the true shape.
This technique is based on what may be called reverse warpage in the
molds. Not only is this a very expensive method because of the iter­
ations involved in reshaping the mold, but it also introduces high
stresses into the part.
These remedies, which can be called superficial, are not recom­
mended because of their side effects. The correct way to remove this
defect is by redesigning the cooling system in the mold for balanced
heat flow out of the part. In other words, the heat taken out of the
core side should be approximately the same as that taken out of the
cavity side. Heat flow rates on both sides can be compared by plot­
ting the heat flow rate versus the pressure-drop curves. A typical
comparison is shown in Fig. 14. If the two curves are far apart, the
part is going through unbalanced cooling and most likely will warp
when molded. The aim of the designer would be to bring these curves
together by augmenting the cooling system on the deficient side. This
can be achieved by modifying the warmer side cooling lines in any of
the following ways;

Bringing the existing lines closer to the part


Introducing more cooling lines
Reducing the coolant inlet temperature
Modifying the diameter of cooling lines
Introducing heat pipes, e tc ., in restricted regions

These changes may in turn modify the process parameters. For ex­
ample, there might be an additional demand for the coolant due to mod­
ifications. Or there may be a change in the flow velocity, and thus
the Reynolds number would change. These improvements, if applied
judiciously, would reduce the molding cycle time, which would be an
added benefit.
Design of Mold Cooling System 595

F ig . 14 Comparison between core and cavity heat flow rates.

Sometimes it is not possible to improve the cooling system on the


deficient side. In that case, at the expense of increasing the cycle
time, the cooling system on the efficient side is modified to balance it
with the deficient side. This is a last resort, used only in extreme
cases when there is no other reasonable alternative for removing the
warpage.

B. Residual Stress

If the part comes out of the mold with high residual stresses, though
it apparently looks normal, it will crack or deform when put to use.
This type of defect is very common in plastic parts, and one can find
numerous examples of cracked parts in daily usage. For balanced
cooling of a part, the stresses are symmetric about the center of the
thickness; the surfaces are in compression, and the inside is in ten­
sion, as shown in Fig. 15. On the other hand, for unbalanced cool­
ing, the stress distribution becomes skew, as shown in Fig. 16.
596 Singh

Fig. 15 Stress distribution for balanced cooling.

The cooler side has high compressive stress while the warmer side may
have very low compressive stress or even some tensile stress in acute
cases. This nonuniform stress distribution may not cause any warp-
age if kept in the mold for a long time, but it can cause cracks in the
part. A little twist can buckle the part and snap the locked-in stress­
es. This defect can be avoided by balanced cooling in the mold.

Fig. 16 Stress distribution for unbalanced cooling.


Design of Mold Cooling System 597

X. COMPUTER-AIDED DESIGN OF
COOLING SYSTEMS

Because of numerous heat transfer and fluid flow parameters involved


in designing the mold cooling system, it is extremely useful to utilize
a computer-aided method to make life easy. The design process can
be divided into two distinct phases. The first phase consists of hav­
ing the geometry of the mold, including the part and the cooling lines,
in the computer data base. The second phase involves the heat trans­
fer and fluid flow analyses of the mold based upon mathematical model­
ing. Such analyses provide the process results which can be used
for improving the cooling of the part, if needed. Any changes made
to the cooling-line geometry would be entered into the data base, which
is shared by both engineering and manufacturing. There might be a
few iterations in this design process before a satisfactory cooling sys­
tem is attained. Within a matter of hours one can optimize the process
parameters and geometry to obtain the best possible cooling-time dura­
tion for the part. The final configuration can then be transformed in­
to working drawings for mold makers. This technology is currently
being used by a number of mold makers. The benefits accrued by
using a computer-aided design process have been found to be ex­
tremely high and cannot be ignored.
Unfortunately, quite often the computer analysis is used after the
mold has been fabricated and is found lacking in expected perform­
ance. At that late stage the advantage of the computer software is
greatly reduced because of the very few options left. Surprisingly,
even at that late stage, the performánce of many molds has been im­
proved considerably by making minor modifications to the cooling lines.
This is especially true for cases where warpage was a problem.
Mathematical methods are available for analyzing the mold for heat
transfer and fluid flow. The degree of complexity of these methods
can vary from one-dimensional steady state to a three-dimensional
transient analysis. However, the criteria are to keep the computer
costs reasonable and provide a fast user-friendly package for mold
designers. In other words, a software should become a tool which
can be used in a daily design environment.
Numerical methods such as the finite element method (FEM) or the
boundary element method (BEM) can be used. These methods need a
fairly good theoretical knowledge to make full use of their capabilities,
and they are too expensive for a three-dimensional analysis. Another
approach in mold cooling design is based on a discrete type of two-
dimensional heat transfer method, or shape factor concept, combined
with fluid flow analysis in a pipe. The cooling lines can be discreti­
zed into a finite number of sections. The entire length of a section
has the same geometrical property and same shape factor characteris­
tics. The heat transfer through the mold is then calculated using
598 Singh

the proper shape factor and the transient variation in temperature of


the plastic part from the melt to the ejection temperature. The con­
vective heat transfer through the coolant is included in the overall
heat transfer coefficient, and the cooling time of the part is evalu­
ated from these relations.

A. C om puter S oftw are

Extensive work has been carried out during the last decade to study
the behavior of plastic parts undergoing cooling [7,8—13]. Some of
the computer software programs for mold cooling analysis in use are
listed below. There are many new programs currently being devel­
oped, so the list may not be complete.

MCAP has been developed by Corporate Research and Development,


General Electric C o ., Schenectady, New York. It has been inte­
grated with the Calma CAD/CAM design system, where it is known
as POLYCOOL.
Graphics Technology C orp., Boulder, Colorado has developed soft­
ware called SIMUCOOL.
MOLDCOOL is software from Application Engineering C orp., Elk
Grove Village, Illinois.
DIEPAK has been developed by Scientific Process and Research
Inc., Somerset, New Jersey.

The structures of different computer software used for mold cooling


analysis may vary considerably. However, a typical flowchart of a
computer program which uses the shape factor technique of heat flow
calculations is shown in Fig. 17.

B. Example Problems

Example 7
The first example concerns an existing mold which was being replaced
by an identical one. Since the fabrication of the new mold was already
in progress, it was felt that major changes in the cooling design were
not possible. Because the cooling system was considered to be very
well designed, it was offered as a test case to test program MCAP.
The cycle time of the mold was 74 sec. The old cooling line configura­
tion is shown in Fig. 18. It consisted of three circuits marked as A ,
B , and C in the figu re. The initial run using MCAP evaluated the
performance of the existing cooling geometry. It was immediately
evident from the results that circuit A was too long because the tem­
perature rise in coolant was exceeding 2.5°C. Circuit A was, there­
fore, split into two circuits. This was easily accomplished by drilling
Fig. 17 Typical flowchart of mold cooling analysis programs.

599
600 Singh

F ig . 18 Example problem 1: original cooling-line geometry.

two holes for inlet and outlet. The original V-shaped sections of the
cooling lines were considered much less effective than rectangular
patterns and were modified at five places. With a few other minor
changes, the final configuration of the cooling lines is shown in Fig.
19.
MCAP analysis predicted an improvement of 35% in the cycle time
for the new design. A new mold was fabricated based on the new
Design of Mold Cooling System 601

F ig . 19 Example problem 1: redesigned cooling-line geometry.

cooling-line design and is currently in production with a cycle time


of 48 sec. The entire computer analysis effort for redesigning the
mold took about 12 hr, including a study of the drawings, data in­
put, and four design iterations. This increase in productivity of
54% was considered astonishing by the plant as they considered the
design to be almost perfect.
602 Singh

E x a m p le 2

Another mold was analyzed because the part was warping as soon as
it came out of the mold. Though the part was acceptable since the
defect could not be easily perceived, the plant wanted to get rid of
the warpage in their new mold which had just been ordei^ed.
The geometry of the part was in the form of a flat plate with cool­
ing lines drilled straight through the mold on either side of the part.
The results from the cooling analysis immediately showed ¿unbalanced
cooling in the core and cavity. The original and the final locations of
cooling lines with respect to the part are shown in Fig. 2,0. The cool­
ing lines in the core were brought closer, two extra cooling lines were
added in the core, and the circuit connections were modified to bal­
ance the cooling. Figure 21 compares the before and after results of

CIR CU IT 1 CIR CU IT 2

b o o o o o o o
6.0 cm

k
2.5cm
o o q o o o o o q—^
C IR C U IT S CIR CU IT 4 C IR C U IT S

C IRCUIT 1 CIRCUIT 2 C IRCUIT 3

b "o "b 6 o o b b "o b T


3.7cm

"T “
2.5cm
o o o o o o
V--------------- ^ ---------- ./
o o o—^
C IR CU IT 4 C IR C U IT S C IR C U IT S

Fig. 20 Example problem 2: Cooling-lines geometry—original and


modified.
COOLANT FLOW RATE (LITERS/MINUTE)

Fig. 21 Example problem 2: Heat flow rate comparison between core


and cavity.
604 Singh

heat flow rate versus the coolant flow rate in the core and the cavity.
After implementation of these modifications to the new mold, it was
put in production to replace the old mold. Not only was the warpage
completely eliminated, but there was a cycle time reduction of more
than 35%.

E xa m p le 3

The final example depicts the common daily experience of manufactur­


ing where the plastic parts do not meet the specifications and have to
be rejected. For such cases a computer analysis prior to fabrication
of the mold is extremely helpful, as confirmed by this example. A
company started manufacturing plastic road-marker housings, but due
to warpage of 2.75 mm the parts could not be fixed on the roads. The
tolerance limit on part warpage was 0.38 mm. Computer analysis of
the mold traced the problem to unbalanced cooling. Accordingly, the
cooling system was redesigned to remove the imbalance. The before
and after designs are shown in Fig. 22. The modified design reduced
the warpage to 0.27 mm which was well within the tolerance limits.
There has been no rejection of the part since the new design was in­
troduced by the company. This is an example where a complete fail­
ure was turned into success by minor modifications to the cooling­
line design with the help of computer analysis.
There are many such success cases where not only the cycle time
has been reduced considerably, but warpage in the parts has been
eliminated completely to make the part acceptable.

NEW LOCATION
OLD LOCATION

PLASTIC
PART V y —y I 2.2cm
1.6 cm

Fig. 22 Example problem 3: Cooling-lines geometry—original and mod­


ified .
Design of Mold Cooling System 605

More and more mold designers are getting on the bandwagon of


computer-aided design because of tremendous benefits in the form of
good quality parts produced at faster rates. There has been a grad­
ual increase in the use of plastic in all the industries. Cost savings
by using computer software will be extremely beneficial as it involves
millions of dollars.

REFERENCES

1. J. B. J. Fourier, Theorie Analytique de la Chaleur, Gauthier-


Vülars, 1822; English translation by Freeman, Cambridge, 1878.
2. E. N. Sieder and G. E. Täte, Ind. Eng, Chem,, 28: 1429
(1936).
3. H. Hausen, Z. Ver, deut, I n g ,, Beiheft Verfahrenstechnik,
4: 91 (1943).
4. A. P. Colburn, Trans, AIChE, 29: 174 (1933).
5. V. P. Isachenko and N. M. Galin, Izv, vuzov, Energetika,
6 : (1965).
6. K. J. Singh, MELTCOOL - Computer Program for Melt Cooling
Analysis, General Electric report to be published.
7. K. J. Singh, SPE ANTEC, 30: 962 (1984).
8. K. J. Singh, NASA Conference Publication, 2245: 141 (1982).
9. A. Prasad and E. W. Rapp, SPE ANTEC, 18: (1972).
10. E. Schurmann and G. Menges, SPE ANTEC, 23: 104 (1977).
11 . K. K. Wang, P. Khullar, and W. P. Wang, SPE ANTEC, 27:
285 (1981).
12 . K. J. Singh and H. P. Wang, SPE ANTEC, 28: 330 (1982).
13. K. R. Schauer, SPE ANTEC, 29: 865 (1983).
Computer-Aided Mold Design
and Manufacturing

K . K . Wang and V . W. W ang*


Sibley School of Mechanical and Aerospace Engineering
Cornell University
Ithaca, New York

I. IN T R O D U C T IO N

A. T y p ic a l Mold C o n fig u ratio n s

A mold is a mechanical apparatus which provides a rigid confined space


where parts are shaped through solidification of polymer melt with pre­
cise dimensions and other features as dictated by the mold cavity. De­
pending on the complexity of the part geometry, a mold could be a sim­
ple structure with a cavity situated between two base plates or as com­
plicated as a sophisticated machine by itself incorporated with power-
driven slides and other moving parts. They move in and out when the
mold opens and closes under the control, which can be integrated into
the control system of the entire molding machine. Despite its complex­
ity, however, the whole structure is always attached mechanically to a
pair of parallel surfaces: one on a stationary platen and the other on
a moving platen of a molding machine.
The most essential part of an injection mold is the cavity or cavities
connected through a delivery system network consisting of a sprue,
runners, and gates. Once they are properly laid out, a cooling (or
heating, depending on type of materials to be molded) system is usually
placed close to the cavity wall to control the heat transfer between the
polymer melt and the temperature-control medium. In a production mold

*Present affiliation: Advanced CAE Technology, Inc., Ithaca, New


York.

607
608 Wang and Wang

a part-ejecting mechanism is usually incorporated, which comprises an


ejector plate and a set of ejector pins strategically located to provide
forces to push the part out of the cavity without distorting it. De­
pending on mold design, parts can be separated automatically from the
delivery system at the gates when the mold opens. They can be car­
ried away by a material handling system, e . g . , on a conveyor, or
picked up by an unloading robot attached to the machine in a more
sophisticated operation.
In practice, only a relatively small number of molds have very com­
plex constructions; the majority of the molds used in industry are of
the so-called two- or three-plate type. These molds all have cavities
and their delivery system engraved between those base plates. In
some cases cavities or runners are machined on replaceable inserts
which are imbedded in the mold base plates. Such a design has the
advantage of flexibility in changing mold configurations and the con­
venience of only machining the inserts instead of the whole mold base
plate. This is particularly critical in precision molding, such as for
optical lenses [1] , in which special machine tools are needed to gen­
erate special contours to an extremely high degree of accuracy and
surface finish. However, except for the inserts, commercial stand­
ard mold bases [2] can still be used economically in these applica­
tions .

B. C onventional Mold Design and M a n u fa c tu rin g Methods

Mold design often starts from the design of a plastic part by a prod­
uct designer in the form of either an engineering drawing or a proto­
type part. The mold designer has to make a series of decisions [3]
such as the number of cavities, the connecting runner system on the
mold base, the type and number of gates, location of the sprue, the
number and location of parting lines, and so on. Based on previous
experience, the mold designer makes these choices by taking into ac­
count the material behavior, the machine capacity ( e . g . , clamp force
and shot size), and even economic factors such as lot size, tooling
cost, surface quality and dimensional requirements, etc.
With a few simple calculations the designer will lay out a cooling
system in which the size, number, and location of cooling channels
will be decided. Other design considerations would include location
of vents to prevent air trap s, a proper ejection system, part shrink­
age, and sometimes even the rigidity and alignment of mold bases.
Once all the design decisions are made, the designer (typically an
experienced draftsman or moldmaker) puts everything down on a set
of drawings sent to the mold shop for manufacture, usually making
use of commercial standard components.
Computer-Aided Mold Design 609

Besides the captive molders who may have a substantial mold shop in-
house, most molds are manufactured (or even designed in many cases)
by small tool-and-die shops. These shops are typically equipped
with conventional machine tools, which may include high-precision jig
borers, copy-mills, and possibly some electric-discharge machines
(EDM). In recent years, more tool-and-die shops have started to
add numerically controlled (N C ) machine tools to their facilities to im­
prove production efficiency and reduce the pressure due to the short­
age of skilled toolmakers. Unfortunately, mold-making by nature is a
labor-intensive task. Stiff requirements of high-quality molds (there­
by good final molded products) demands much manual work. Further,
rework on the same mold through many iterations due to faulty design
or other reasons is commonplace in industry. This makes the tool cost
very high and may offset the benefit of using the injection-molding
process for small or even medium sized production quantity.

C. C om puter-B ased Mold Design and M a n u fa c tu re

The injection-molding industry is one of the fertile areas which can


benefit if proper application software can be developed. Basically,
there are three areas in injection molding where computer techniques
can be introduced to improve the situation: mold design, mold man­
ufacture, and control of the injection-molding process. However, be­
fore any viable software can be developed, one must have a good un­
derstanding of the process, i . e . , the physics and chemistry involved
in the process. Since the 1960s and even earlier, numerous studies
on rheology, flow simulation, and other related problems in injection
molding have been carried out, as referenced in previous chapters.
A more concentrated effort using a system approach to all three areas
was initiated in 1974 at Cornell University with the Cornell Injection
Molding Program (CIMP) sponsored by the National Science Founda­
tion. A vast amount of literature [4] has been produced since then,
including a substantial part of this book.
Historically, the molding industry also began to use commercial
CAD software as a computer-aided drafting tool since the late 1960s.
As scientific understanding of the molding dynamics began to shape
up and computational methods to solve the governing equations were
developed, computer programs became available as a tool to help mold
designers make design decisions. The first such program, appearing
commercially in 1978, was MOLDFLOW, developed in Australia.
As the computational task becomes more elaborate for practical mold
design and manufacturing, use of computer graphics as a human/ma­
chine interface also becomes necessary to facilitate pre- and post-pro­
cessing of various programs. Many CAD/CAM software vendors now
610 Wang and Wang

offer graphic packages at various degrees of sophistication interfaced


to mold design programs. One of the most sophisticated software pack­
ages is C-FLOW, developed by Advanced CAE Technology, In c., which
simulates the mold-filling process for complicated parts in three-di­
mensional space with high-level graphics. In front of a graphics ter­
minal the user not only can get the results from flow simulation in vari­
ous forms, including color-shaded pictures for ’^animating’’ the molding
process, but also incorporate the process control parameters via a
graphic menu to find their respective effects on the filling process.
On the mold manufacturing side, use of wire-frame-based turnkey
CAD/CAM systems is on the rise in the injection-molding industry.
Generation of machining codes for NC machine tools is usually limited
to two- or three-axis machining operations. Tool paths generated pre­
viously can be displayed in line traces with different colors for verifi­
cation. NC machine tools are also used to machine electrodes for EDM,
which have become more commonly used to produce cavities of intricate
shape for better surface finish. Use of microprocessors for process
control is also gaining acceptance throughout the industry. In some
sophisticated molding shops, microcomputers are used to monitor and
control the operations of the whole shop, including directing the ro­
bot motion and gathering management information.
This chapter is dedicated to computer usage in the molding indus­
try, particularly for mold design and manufacture, providing the read­
er an overview of computer graphics and solid moldeling. A detailed
treatment of more recent developments of specific CAD/CAM software
is also included. Integration of design and manufacturing functions
incorporating various data bases will be briefly discussed.

II. O VERVIEW OF IN T E R A C T IV E COMPUTER


G R A PH IC S AND GEO M ETR IC M ODELING

As the power of computers at all levels continues to rise and cost de­
clines, use of computers in all aspects of engineering has become a
way of life. When a supercomputer can crank out numbers at an ex­
tremely high speed, bottlenecks still remain, depending on how fast
one can communicate with the machine and interpret the vast amount
of data. Interactive computer graphics (IC G ) is a natural and ideal
means for human/machine communication. From picture synthesis to
representing data graphically, interactive graphic techniques become
a necessary tool for engineers and scientists in handling certain com­
plex tasks.
The history of modern interactive graphics dates back to the early
1960s in Ivan Sutherland's work on the Sketchpad drawing system [5 ].
Computer-Aided Mold Design 611

He introduced data structures for storing symbol hierarchies and de­


veloped some fundamental interaction techniques for using the key­
board and light pen, which are still used. Since then, the enormous
potential for partially automating the labor-intensive drafting in de­
sign and machining code generation in manufacturing has been rap­
idly explored. This has developed into a multibillion dollar CAD/CAM
industry, which is drastically changing the way engineers work. This
section is designed to provide the non-ICG-oriented readers an over­
view of the main components involved in an interactive graphics en­
vironment .

A. D isplay Devices

A display unit is an essential device for human/machine communica­


tion. In addition to some devices which produce hard copies, such
as a line printer or a pen plotter, devices most commonly used in ICG
are either a cathode ray tube (CR T) or a liquid crystal display (LCD)
unit. Depending on the design of the device, the images can be gen­
erated on the screen either by random scan or by raster scan. In a
random scan device the image displayed on the screen can be in any
random order, just as a pen plotter can draw a picture in any se­
quence on a piece of paper. On the other hand, a raster scan device
scans the screen in a definite order from left to right and top to bot­
tom , like a line printer or a television set. The picture is composed
by the darkness (or color) ’’painted” at each picture element (or
pixel, a point on the screen) over the whole display area—hence the
higher the density of the pixels, the better the quality of the picture.
Typically, a broadcast television set in the United States has 525 x
340 pixels, whereas most raster graphic systems use between 256 x
256 pixels in the low end of resolution to 1024 x 1024 pixels in the
high end.
In addition to high resolution, ICG often requires the image on the
screen to be changed or moved around very quickly. For fulfilling
these requirements, CRT is by far the most commonly used display
device, and probably will be for many years. In general, CRT works
on the principle that a controlled electron beam emitted from an elec­
tron gun hits a specific point on the phosphor-coated CRT screen.
This produces an energy release at the spot in the form of light due
to the change of state of the electrons in the phosphorus atoms. This
light output decays exponentially with time after the beam moves away
from the spot; and the rate of decay (or the persistence of light) de­
pends on the characteristic of the phosphors to be used.
Therefore, a trace on the CRT produced by a single sweep of the
beam is only visible momentarily. To produce a steady picture, the
612 Wang and Wang

same pattern must be repeatedly traced (or refreshed) at an appro­


priate rate, without flicker, to the user^s eyes. The refresh rate for
a raster scan display, which has a fixed number of pixels to refresh,
is 30 to 60 per second, whereas that for a vector refresh unit depends
on the complexity of the picture ( e . g . , the number of lines and char­
acters involved): typically, the greater the complexity, the longer
the time required for a single refresh and thus the lower the refresh
rate. In this case phosphors with very long persistence would ap­
pear to be desirable.
For most applications three types of hardware display units are
commonly used;

Vector Refresh Displays


A display device developed in the mid-sixties and still used is em­
ployed to display points, vectors, strokes, or calligraphics on the
screen. It consists of a display processing unit (D P U ), a display
buffer memory, and a CRT with its associated electronics. The b u f­
fer memory stores the display list or program produced by the CPU
of a host computer, and the DPU interprets and then converts the
digital information to analog voltages, which in turn control the di­
rection of the electron beam writing on the C R T . To avoid flicker on
the screen, the DPU must cycle through the display list to refresh
the phosphor; hence the buffer holding the display list is often called
a refresh buffer. In general, the vector refresh display has the ad­
vantages of dynamic picture manipulation, but the requirements of
large refresh buffer and fast DPU made the devices very expensive
in the sixties.

Storage Tube Displays


In the late sixties the storage tube or the direct-view storage tube
(DVST) display technology was developed and considered as a major
step in making ICG affordable. In DVST the image is stored and
written on the screen only once until erased. This is done with a
relatively slow-moving electron beam on a storage mesh in which the
phosphor is embedded. Without needing buffer and refresh process,
DVST became a cost-effective alternative to the expensive vector dis­
play, and has had a great impact on the acceptance of ICG as a vi­
able tool in complex CAD applications. Storage tubes are still popu­
lar for applications requiring high resolution but not dynamic picture
manipulation. This small and self-sufficient device was ideal at the
time, particularly for an inexpensive, low-speed telephone interface
to a time-sharing system.
Computer-Aided Mold Design 613

Raster Scan Displays and Frame Buffer

Evolved from the commercial television technology , the relatively low-


cost raster scan displays were developed in the mid - seventies. Since
the image is formed from the raster, a matrix of pixels covering the
entire screen, the whole area is scanned sequentially 30 times per
second by only varying the intensity of the electron beam at each
pixel. This requires a large memory in the refresh buffer to store
the information on each pixel explicitly in a ^^bit map” form, which
has one-for-one correspondence to the points on the screen. A sim­
ple refresh buffer has 1 bit for each pixel and thus defines a two-
color (usually black and white) image. Modern refresh buffers (also
called frame buffers) often provide up to 8 bits per pixel in order to
generate satisfactory continuous shades of gray. For color displays,
8 bits are needed for each of the additive primary colors re d , blue,
and green, i . e . , a total of 24 levels. Such a large frame buffer be­
came economically viable only after the technology of the random-ac­
cess, solid-state memory was developed with its continuously declining
price. The raster graphics boost the applications of ICG in many areas
because it makes possible the display of color-shaded images, which is
an especially powerful means of communicating information.

B. In p u t Devices and Techniques

Along with the development of output technology has been improvement


in input devices. The bulky light pen touched on the screen used in
the early years has largely been replaced by a slim stylus moved on a
data tablet, by a transparent, touch-sensitive panel mounted on the
screen, or even by audio communication. This section provides an
overview of some hardware input devices typically used in an ICG en­
vironment today, as shown in Fig. 1. The numerous interaction de­
vices and techniques can be classified into five basic logical devices
[6] according to their functions: the locator, the pick, the valuator,
the keyboard, and the button. A brief description of each follows.

Locators
A locator is an input device for entering a single value to the computer
in the space of real numbers. The most commonly used locator is the
^tablet,” a flat surface over which a stylus (like a pencil) or hand cur­
sor is moved to locate the position of interest. Most tablets use an
electrical sensing mechanism to measure the stylus or hand cursor po­
sition. One such arrangement has a grid of wires embedded in the tab­
let surface. When the stylus or cursor is near or in contact with the
6U Wang and Wang

IBM 5081 Display

Fig. 1 A typical graphic workstation with a display unit and input de­
vices, as labeled. (Courtesy of IBM Corporation, White Plains, New
Y ork.)

tablet, the electromagnetic coupling of the electrical signals between


them induces an electrical signal in a wire coil in the stylus or cur­
sor. In the meantime, a ’’screen cursor” is usually displayed to track
the position being read from the tablet to provide a visual feedback to
the user. The tablet can also be used for digitizing drawings without
recording an excessive number of points.
The second commonly used locator is the ’’mouse,” which is a hand­
held de\dce with rollers on its base. Moving the mouse across a flat
surface causes the rollers to turn, which sense the relative movements
in two orthogonal directions via the coupled potentiometers. The mo­
tion is converted to digital values which represent the direction and
magnitude of the mouse’s movement. Unlike the tablet, which reports
absolute positions, a mouse reports relative movement because one can
pick it up, move it, and then put it down anyplace with no change in
the apparent position of its screen cursor.
The third commonly used locator is a ’’joystick,” which can be
moved left or right, forward or backward. Again, with the poten­
tiometers sensing the movement, a joystick can be used for control­
ling absolute position of a screen cursor.
Computer-Aided Mold Design 615

Picks
The light pen is the only pick device used in computer graphics. De­
veloped very early in the history of ICG, the light pen ^’sees” light
on the screen through a lens and a fiber-optic bundle and sends the
signal directly to the DPU^s control logic. In this sense, light ^’pen”
is a misnomer because it ’’reads” the light from screen rather than
’’writes” on the screen. A finger-operated switch is usually used to
enable and disable the pen. When implemented, it performs a ’’pick”
function on a vector display but acts as a ’’locator” on a raster dis­
play. When a pixel is detected, the image display system can make
the X and y coordinates of the raster scan’s current position available
for the associated computer.

Valuators
A valuator is usually a potentiometer which provides scalar values typ­
ically in terms of an analog voltage. In many ICG installations they
may have 8 or 10 rotary dials mounted as a group on a panel, as shown
in Fig. 1. Each potentiometer is associated with an analog-to-digital
converter and a power supply; the output from the converter goes to
a DPU register, which in turn can be read by the computer CPU.

Keyboards
The alphanumeric keyboard is the most common input text device. Its
important function is to create a code (e . g . , ASCII, EBCDIC, etc.)
uniquely by depressing a corresponding key. The closure of a key
can be achieved by mechanical contact, change in capacitance, or mag­
netic coupling.

Buttons
The most common button device is the programmed function keyboard
(P F K ). It is often built as a separate unit, as shown in Fig. 1, or it
can be integrated with the keyboard. Other configurations could be
keys found on many tablet cursors and on the mouse. Buttons are
generally used to enter commands or menu options to a graphics pro­
gram.

C. Geom etric Modeling

Since all physical objects have a shape, the design and manufacture
of these objects will require an efficient means of representing shapes.
The term ’’geometric modeling” often refers to the methodologies, tech­
niques, and systems focused on ’’informationally complete” representa­
tion of physical objects in a computer. Traditionally, this is done via
616 Wang and Wang

an engineering drawing, which is a collection of projected views of an


object represented by a composite of connected lines and associated
notes and special symbols. As a result, in the early stage of the de­
velopment, and even today to some people, the commercial CAD sys­
tems are often interpreted as automated drafting packages using the
’’wire-fram erepresentation. The evolution started from the mid­
sixties with two-dimensional systems, moved to three-dimensional wire
frames 10 years later, and extended to include bounded-surface capa­
bility with colored display and incorporated with better engineering
analysis packages in today’s ICG systems.
Though useful as an automated drafting tool, wire-frame systems
pose some serious deficiencies in order to achieve a higher degree of
automation. Information in the wire-frame representation of a solid
object sometimes may not be complete and unique. An incomplete rep­
resentation could result in an object which may not exist, whereas a
nonunique representation could cause ambiguity, which requires hu­
man interpretation. Hence, a wire-frame representation scheme can­
not be used reliably for automatic sectioning or volume/weight calcu­
lation of the object. Much low-level data are often needed for further
description and clarification. To cope with these problems, a new
technology, now known as ’’solid modeling,” began to evolve in the
late sixties [7 ]. The main distinction of solid modeling from wire­
frame is that the object is represented by unambiguous combinations
of solids instead of lines and points in three-dimensional space. This
section provides a brief overview of each system as background infor­
mation .

W ire-Fra m e R e p r e s e n ta tio n

Mathematically speaking, most wire-frame systems require only some


simple analytic geometry—the transformational algebra for translation,
rotation, scaling, and projection. However, the implementation of such
tasks in an ICG environment requires lots of programming; numerous
clever techniques and algorithms have been developed to perform these
functions with the capabilities of removing hidden lines and surfaces
and getting shaded pictures. One of the basic difficulties in display­
ing a three-dimensional object as a two-dimensional image on the screen
is to provide special depth information which depends on the position
of the viewpoint. Figure 2 illustrates a typical wire-frame drawing for
a mold assembly in which every edge of the components involved in the
assembly is displayed according to a chosen viewpoint and viewing di­
rection. To remove hidden lines from such a complex assembly requires
much more computation. Although a perspective display like this may
help visualize the object, a work drawing, as shown in Fig. 3, is still
preferred in practice to provide all detailed information for manufac­
turing.
Computer-Aided Mold Design 617

Fig. 2 Wire-frame drawing of a mold assembly displayed on the CRT


screen.

Fig. 3 Orthographic projections of a mold base displayed on the CRT


screen.
618 Wang and Wang

Fig. 4 Constructive solid geometry representation of a sprue bushing.

S o lid M o d elin g

The basic concept of solid modeling is to represent a physical object


by manipulating (adding, subtracting, etc.) simple but unambiguous
solids, often called ’’primitive solids” or simply ’’primitives.” These
’’building blocks” (typically including cylinders, blocks, spheres, cones,
wedges, etc.) can be employed to construct an object of complex shape
using a number of set operators, such as union for addition and differ­
ence for subtraction. There are two commonly used schemes for solid
modeling: constructive solid geometry (CSG) and boundary represen­
tation (B .R e p ). Figure 4 illustrates the principle of the CSG repre­
sentation of a sprue bushing typically used as part of an injection
Computer-Aided Mold Design 619

mold. The bushing can be constructed from four primitives: two cyl­
inders of different size, one sphere, and one cone. The symbols —
and U at the nodal point represent the Boolean set operators, ’’differ­
ence” (subtraction) and ’’union” (addition), respectively. After each
Boolean operation a higher level of object is constructed until the
whole ’’tree” of operations is completed. On the other hand, the
boundary representation scheme represents a solid in terms of its
boundary, i.e ., ’’enclosing surfaces.” The boundary of a solid is
usually represented as a union of enclosing faces, with each face rep­
resented by the union of its enclosing edges. Figure 5 shows the
boundary representation of the same sprue bushing, which is repre­
sented by the union of seven bounding faces.
620 Wang and Wang

F ig . 6 Shaded picture of an exploded mold assembly represented in


solids.

With solid modeling, one can readily obtain shaded pictures for
better visualization and calculating mass properties ( e . g . , volume,
weight, center of gravity, etc.) and possibly other applications. Fig­
ure 6 is the shaded picture of the same mold assembly as shown in Fig.
2 but modeled in solids. One can also get exploded views for each
component or take sections across any planes.

III. CAD /C A E /CAM SOFTWARE FOR


IN JE C T IO N M O LDING

From the viewpoint of industrial application, computer software be­


comes an important part in putting the scientific principles to work
in practice. In recent years a growing number of commercial soft­
ware packages especially for mold design became available. Packages
better known to industry are MOLDFLOW from Australia, MOLD COOL
from the Application Engineering Corporation, and OptiMold from
GRAFTER, Inc. MOLDFLOW (a flow analysis software) and MOLD-
COOL (a cooling system design program) were the first two commer­
cial packages available through a time-sharing computer network
Computer-Aided Mold Design 621

from General Electric Information Services. Many CAD/GAM turnkey


system vendors later acquired and interfaced MOLDFLOW into their
system as a module for mold design. On the other hand, GRAFTER,
as a member of the GIMP consortium, made use of CornelPs research
core programs and developed an integrated CAD/CAE/CAM system,
OptiMold, which includes both flow and cooling analyses. Under a
license agreement with Cornell University, OptiMold added the FLOW 3D
module to handle three-dimensional parts of any complexity. More re­
cently, as spinoffs of university research, two CAE software packages
also became commercially available: CADMOLD is based on the work
done at the Technical University of Aachen in West Germany, and C-
FLOW was developed by Advanced CAE Technology, Inc. under a li­
cense agreement with Cornell University.
Since all commercial software is proprietary, this chapter attempts
to provide adequate underlying theories and methodologies which could
be used to develop software on the following selected topics.

A. R u n n er System Design

In injection molding, polymer melt flows through a delivery system con­


sisting of a sprue, a runner system, and gates into individual cavities.
The gate, which is typically a small opening separating the runner and
part, can also be used for controlling the rate of flow into the cavity.
Typical runner shapes are either full round, half round, or trape­
zoidal cross section. Circular runners are the easiest to eject from
the mold, but they are difficult to manufacture. On the other hand,
trapezoidal runners, typically machined into one of the cavity plates,
are easy to make but require an ejection system. Based on thermal
conditions around a runner, the mold can have either a hot runner,
an insulated runner, or a cold runner system. Hot runner and insu­
lated runner systems are often useful in molding large parts or parts
with thin cross sections. Conventional cold runner molding will gen­
erally require longer cycle times simply to cool the runners. The hot
runner mold adds substantial fabrication costs and consumes more en­
ergy. However, the solidified runners from a cold runner mold be­
comes scrap, even though they might be reground for future use in
the case of thermoplastics.
The runner system design is very important to the quality of mold­
ed parts and to the production costs. Runners that are too large re­
sult in wasted material and a longer cooling cycle, whereas runners
that are too small will cause excessive pressure drop which may cause
insufficient injection pressure for filling the cavity. In addition, small
runners may degrade the polymer by excessive viscous heating. For
multicavity molds balancing the fill time and pack pressure are often
the key to successful mold design. If the cavities are not filled at the
same instant and under the same pressure, it may cause nonuniform
622 Wang and Wang

quality and part weight from various cavities. For a multigated cav­
ity, different gate openings and runner systems will result in differ­
ent weldline locations and, in turn , will affect the mechanical proper­
ties of the product.
The following design criteria should be kept in mind when design­
ing runner systems:

The total volume of the runner system should be as small as pos­


sible .
The pressure loss in the runner system should be tolerable; the
bigger the part, the smaller the percentage of total injection
pressure that should occur in the runners.
The viscous heating generated by the flow in the runner system
should not cause degradation of the material.
The gate should be as small as possible consistent with desired fill
time, orientation of the flow, and appearance of the p art.
The cooling effect in the gate should not freeze the polymer flow
during filling, but the gate should be effectively cooled after the
cavity has been filled.
For multicavity molds the runner system should be designed such
that each cavity gets filled at the same instant and receives the
same packing pressure.
For a multigated cavity the runner system should be designed so
that the weldlines are located as desired.

The traditional way of designing the runner system is by experi­


ence or trial-and-error through actual experiments. This is time con­
suming, ineffective, and costly. A rational approach to this problem,
presented in the next two sections, is based on an engineering analy­
sis to determine the dimensions of the runner system [8] . In the anal­
ysis the following assumptions are made:

The cross sections of all runner segments in a system are similar in


shape but not dimensions.
Runner segments are approximated by circular tubes with the same
hydraulic radius, which is defined by

( 1)
C
where A denotes the cross-sectional area, and C the circumference
(wetter perimeter).
The polymer melt is considered to be a power-law fluid without elas­
tic effects.
The polymer-melt flow in the runners is under isothermal conditions
so that no cooling or viscous heating effects are included.
Pressure losses at junctures are neglected.
Computer-Aided Mold Design 623

jqj noturolly
bolonced

noturolly
unbolonced

□ —r-€l O -r-Q
, V noturolly
IT ^ unbolonced

-□ O

F ig . 7 Various types of runner systems.

Basically, runner systems in a mold can be divided into two cate­


gories, based on their geometric arrangements, namely, naturally bal­
anced and naturally unbalanced runner systems, as illustrated in Fig.
7. Here, ’’naturally balanced” means that all the flow paths are iden­
tical in shape, size, and total length. If they are not identical in all
flow paths, the runner system is naturally unbalanced, because the
flow resistance in each flow path will not be identical. However, it
is desirable to make it balanced artificially such that every cavity in
a mold gets filled at the same instant in order to achieve uniform prod­
uct quality from all cavities.

D e sig n o f N a tu ra lly B a la n ce d R u n n e r S y s te m s

Because of its geometric symmetry , all flow paths in a naturally bal­


anced runner will impose identical resistance to the flow. As a result,
all cavities will be filled at the same instant if the cavities are also
identical. Therefore, the design criteria for naturally balanced run­
ner systems should be

To minimize the pressure loss of the flow in the runners with an


acceptable runner/cavity volume ratio
To size the runners such that the pressure gradient along the en­
tire runner path is uniform
624 Wang and Wang

The first criterion represents a compromise between making full use


of the machine capacity for filling the cavities and, on the other hand,
minimizing material wasted in the cold runner system. The second cri^
terion is employed to make the flow distribution in each path relatively
insensitive to the dimension of any runner segment which may be off
due to machining error or wear. Based on the above assumptions, it
can be shown that the pressure loss will be minimized, for a specified
total runner volume and fixed runner lengths, provided that the up­
stream and downstream runner radii, au and a^, respectively, at a
juncture are related by

ad ( 2)

where n denotes the number of downstream branches at a given junc­


ture. In addition, with a given total runner volume V, the individual
runner segment size can be determined by the relation

V = dE Li (3)

where D = C^/4A is a constant for a given runner shape and aj and Lj


are the radius and length of an individual runner segment. The gate
is considered to be small such that its volume is not included.
As an example, an eight-cavity mold with a naturally balanced run­
ner system, as shown in Fig. 7(a) has total cavity volume of 32 cm^.
Assume that the desired runner/cavity volume ratio is 1:2; the run­
ner lengths in each segment from upstream to downstream are 5, 4,
and 2 cm, respectively; the corresponding hydraulic radius of each
runner segment can be determined as follows:

a^^ = 2

8l 2 ^ = 2 as^

V = D(2 X ai^ X 5 + 4 X 8 .2 ^ X 4 + 8 X 8 s 2)

Since V and D are known and there are three equations for three un­
knowns, the values for a^, aa, and 8 ^ can be obtained by solving
these algebraic equations. For a full round runner, D = tt such that
ai = 0.440 cm, 82 = 0.349 cm, and 8 ^ = 0.277 cm.

Design of Artificially Balanced Runner Systems


There are at least three cases where a runner system has to be spe­
cially designed to balance the entire runner/cavity system: (1) a
naturally unbalanced system as shown in Fig. 7(b) and ( c ) , (2) a
Computer-Aided Mold Design 625

family mold which consists of cavities for different parts, and (3) con­
trol of the weldline position in a multi-gated cavity. In the simple
case of the naturally unbalanced runner system, the design task will
be simply to size the runner segments such that the flow front reach­
es the gate in each flow path at the same instant. The relations used
in determining the sizes of the various segments are based upon the
following considerations:

1. To maintain a constant volumetric flow rate in each runner seg­


ment such that the flow in any segment is unaffected by the
branching in other segments, we can derive a relation between
the radii of the upstream and downstream runner segments at
a given juncture in the same way as in Eq. (2 ):

a. (4)
1
i=l

where d denotes the total number of downstream branches at a


given junction.
2. A schematic representation of two runner paths is given in Fig.
8, where two segments with radii ar and as have a common up­
stream segment. Suppose that one flow path passing through
ar has m downstream runner segments after ar, whereas a
flow path passing through ag has n downstream runner seg­
ments after ag. If we require the flow front to reach the end
of these two runner paths at the same instant, it can be shown
that the following relation must be satisfied [4 ]:

^ ^ ^m r k /, ^d4-l \V^n
^ I'l=l(‘ * "lil ■^1,1) J^K
(5)
S 1=1 L ]=1\ 1=1 1,]/ J 1

where Li denotes the length of runner segment i, dj denotes the


total number of branches having the same upstream segment as
runner segment j, and

r. . = ;;— w, i = 1, 2, 3.......d^-l ( 6)
1,3 a(jj i

It has been tacitly assumed that Eq. (4) is satisfied at every


branch point such that the volumetric flow rate in each segment
does not vary with time once it gets entered by the polymer
626 Wang and Wang

/ =2

k =m /= n

runner _L
runner
^ P t.r Ap

T "

F ig . 8 Schematic representation of a runner restrictor system.

melt. From Eqs. (4) and (5) we have the relations for the
radii of all runner segments. That is, with specified runner
lengths and a given total volume of runner system, the radius
of each runner segment can be calculated.

Equations (2) and (4) correspond to the same relation between the
radii of the upstream and downstream runner segments. Also, Eq.
(5) is automatically satisfied for a naturally balanced runner system.
Accordingly, in either a naturally balanced or artificially balanced
runner system, the above design considerations will be met if Eqs.
(4) and (5) are satisfied.

D esig n o f R u n n e r R e s t r ic t o r s

The analysis of polymer flow through various types of gates ( a type


of juncture) is rather complex and has been partly treated in Chap­
ter 2. For simplicity, the present modeling scheme employs so-called
Computer-Aided Mold Design 627

runner restrictors in place of gates. The purpose is to impose a pres­


sure drop at the end of each runner path so that the volumetric flow
rates into the cavities are distributed in a manner that the designer
desires. The present mathematical formulation of runner restrictor
design is based on the following assumptions:

The flow front reaches the runner restrictor in each flow path at
the same instant.
The volume of the runner restrictors is very small compared with
the volume of the runners such that the time required to fill the
runner restrictor is negligible compared with that for the run­
ners .

In particular, the volumetric flow rate passing through each run­


ner restrictor can be evaluated such that each cavity (or domain be­
tween weldlines for a multigated cavity) fills at the same instant; that
is,

a corresponding gate-cavity volume to be filled (7)

where qj denotes the desired volumetric flow rate in flow path i when
the melt front leaves the runner restrictor and where the volumetric
flow rate passing through each runner segment is

(volumetric flow rate of flow paths that


‘*1 I pass through runner segment j)
( 8)

The required pressure-drop difference between runner restrictors r


and s in Fig. 8 is then given by

AP, - AP^ =D (9)


t ,r t,s

where AP^^ denotes the pressure drop across runner restrictor i.


Ri denotes the total number of runner segments along runner path
i, and D is a function of the power-law index n and material prop­
erties, which is a constant for specified material and melt tempera­
ture.
In particular, suppose we specify the same length, 1, for each of
these runner restrictors but allow their radii to differ in order to
satisfy the desired design condition. We then have
628 Wang and Wang

D = AP, - AP, ( 10)


3n+ll 3n+ll t ,r t,s
L\ t,r
«
tJ,s
.

where at,i denotes the radius of runner restrictor i.


From Eqs. (9) and (10) it follows that we can determine all at ,i once
we have specified the value of one such radius. In the present formu ­
lation this is done by specifying the smallest allowable runner restric­
tor radius, which, indirectly, is equivalent to specifying the maximum
allowable pressure drop across such runner restrictors. With the Qj
in Eq. (9) based upon Eqs. (8) and (7 ), it then follows that the melt
will enter each gate with a volumetric flow rate proportional to the
gate-cavity volume. Hence, with the present isothermal modeling,
each cavity will fill at the same time provided its flow resistance (per
unit filled volume) is the same.
This then constitutes the basis for generating the initial design
configuration. Nonisothermal effects and a more detailed modeling of
the cavity flow dynamics are then incorporated in succeeding iterations
by making use of the one-dimensional-flow analysis program described
in Chapter 1.

An In te ra ctiv e C om p u ter Program fo r R u n n e r - G a t e - C a v it y D esig n

Based on one-dimensional-flow analysis, an interactive computer pro­


gram for cavity-filling simulation and for runner system design has
been developed as an aid for mold design. The program has the for­
mat of interactive input and graphic output display. A three-dimen­
sional pressure or temperature plot can be drawn for each flow path,
with temperature here denoting the bulk temperature at a specified
streamwise location and time. The flow front position, fill time, in­
jection pressure, and pressure gradient in the cavity can be read
from the three-dimensional pressure plot, as shown in Fig. 9. In
Fig. 10 the temperature plot indicates the bulk temperature history
and its variation along one flow path and points out where and when
the lowest and the highest bulk temperatures occur during filling,
which relates to the possible occurrence of a short shot or of poly­
mer degradation due to overheating.
A flowchart of the procedure used to run the program for runner
system design is given in Fig. 11. In particular, the designer has to
find the best design of the cavity through iterations with the one-di­
mensional-flow program. The runner system design, however, will be
handled by the RUNNER and RESTIC design programs, which save time
and cost by determining the initial dimensions in the runner system di­
rectly within the framework of isothermal modeling.
GomputerT^Aided Mold Design 629

1806.0 PRESSURE

F ig . 9 Pressure distribution along the shortest flow path of an arti­


ficially balanced runner system, as shown in Fig. 7 (b ).

In the usual case nonisothermal effects in the runners and runner


restrictors will then be incorporated in successive iterations by using
the one-dimensional-flow program. ITERATEl in Fig. 11 is construc­
ted such that, under nonisothermal effects, the flow front will reach
the runner restrictor at the same instant, whereas ITERATE2 estab­
lishes the same fill time for every flow path, including runners, run­
ner restrictors, gates, and cavities, under nonisothermal conditions.
During these iteration procedures, the radius of the last runner seg­
ment (ITERATEl) or the radius of the runner restrictor (ITERATE2)
in every flow path will get updated based upon the relation

,1/m
B = (11)
^0 \ts/
630 Wang and Wang

533.0 TEMPERATURE

2.95

TIME

11.58 0.00
511.00

Fig. 10 Bulk temperature distribution along the shortest flow path of


an artificially balanced runner system, as shown in Fig. 7 (b ).

where íq is the system fill time, which is calculated as the total vol­
ume of all runner segments divided by the constant total volumetric
ñow rate (ITERATE 1); or ts is calculated as the total volume of sys­
tem , including runners, runner restrictors, gates, and cavities di­
vided by the constant total volumetric flow rate (ITERATE2); to is
the fill time of a given flow path in the last interation; ao is the old
radius of the last runner segment (ITERATE 1) or of the runner re ­
strictor (ITERATE2) for the given flow path; an is the new radius of
the last runner segment (ITERATE 1) or of the runner restrictor
(ITERATE2) for the given flow path; m is a relaxation factor pro­
vided by the user before each iteration; for stability reasons it is
suggested to use 1 (fast) to 3 (slow) for ITERATE 1 and 10 (fast) to
30 (slow) for ITERATE 2.
Moreover, the one-dimensional-flow program can be used to design
the runner system for a multigated cavity to control the weldline
Computer-Aided Mold Design 631

Fig. 11 Flowchart for a runner system design program.

positions. In this case the multigated cavity is imagined to be parti­


tioned based upon the desired weldline locations. Each portion, which
is filled through one runner restrictor, is then treated as an individual
single-gated cavity. A runner system is then developed such that
each of these individual single-gated cavities fills at the same instant,
thus assuring the desired weldline locations.
632 Wang and Wang

An example of an artificially balanced runner system is given in


Fig. 7 (b ), which includes three identical striplike cavities in one
mold base. The design criterion is that all cavities should be filled
at the same instant. In particular, all of the cavities are identical
in this example. Accordingly, each flow path will see the same cav­
ity flow resistance if the melt fills each cavity with the same volu­
metric flow rate, thus assuring that the injection pressure for each
cavity will also be the same at every instant. This is done by add­
ing a restrictor at the end of each flow path in place of a gate. The
material properties in this example have been taken to correspond to
polypropylene with the injection melt temperature taken as 533 K, the
wall temperature as 300 K, and the constant total volumetric flow rate
as 10 cm^/sec.
Figure 12 shows the calculated result of a naturally unbalanced
runner system where the radii of all runner segments are identical,

10 cm'^/sec

flow path

2 cm
- H ------ 1
1.87
— f — 3.13
5.01
time flow front
(sec) position 2.50
— 1 - -
0.44 3.32
1.11 — ____ 1 ____

1.52 ------— 4.18

fill time (sec) 1.70 1.90 1.98

cavity pressure at
time = 1.98 sec 2008 1792 860
(psi)

F ig . 12 Conditions of cavity filling in a naturally unbalanced runner


system.
Computer-Aided Mold Design 633

restrictors are all set to the minimum allowable value. The result in­
dicates the flow front positions at three time intervals, fill time for
each flow path, and the final pressure in each cavity at the time of
fill. The maximum difference in fill time of 0.28 sec may not be sig­
nificant, but the difference in cavity pressure of 2008 versus 860 psi
could cause inconsistent quality of the parts molded from the same
shot. On the other hand, by resizing the runner segments the arti­
ficially balanced runner system ensures that all flow paths are filled
at the same instant and the variations in cavity pressure are minimal,
as shown in Fig. 13.
It can be concluded that this simple interactive runner system de­
sign program can give a quick, initial design of runner systems, based
upon isothermal power-law fluid modeling, such that (1) the pressure
loss is minimized in the case of a naturally-balanced runner system or
(2) the polymer melt flow is artificially balanced in the case of a nat­
urally unbalanced system. Experience shows that only a few iterations

Fig. 13 Conditions of cavity filling in an artificially balanced runner


system.
634 Wang and Wang

are then required to balance the fill time with nonisothermal effects
included.

B. M o ld -F illin g Sim ulation Program s

Computer simulation of the mold-filling process has been recognized


as a powerful tool for mold designers to make rational design decisions
and achieve better results. Theoretical work dating to the early sev­
enties is fully described in Chapter 1. Early commercial versions of
the mold-filling simulation programs were based on one-dimensional-
flow analysis (one dimension in the sense that only one dimensional
variable in the flow field is used for computation). Such programs
include MOLDFLOW^s so-called Branching Flow, which treats each
flow path separately, and GRAFTEK^s Simuflow, which employs the
coupled flow path method [9] where all branched flow paths are
coupled fluid dynamically. In this section we present a computer
program developed at Cornell University which simulates the com­
plete mold-filling process for three-dimensional thin parts of any com­
plex geometry with high-level interactive graphics [ 10].

T h e o r e tic a l B a s is a n d N u m e rica i im p ie m e n ta tio n

The underlying mathematical formulation for fluid flow used in the pro­
gram is fully described in Chapter 1 (or see [16,17]). Essentially ,
the modeling is based on generalized Hele-Shaw flow for an inelastic,
non-Newtonian fluid under nonisothermal conditions. A hybrid nu­
merical scheme is employed in which the injection-molded part is de­
scribed by two-dimensional triangular elements, provided the cavity
thickness is relatively thin, and the gap wise and time derivatives are
expressed in terms of finite differences. The elements are flat but
can have any orientation in three-dimensional space to approximate
the surfaces of the molded p art. For cavities of more general planar
geometry, triangular elements are preferred to quadrilateral elements
[11] for the following reasons:

Highly irregular cavities can be more easily divided into triangu-


lars than into quadrilaterals.
Cavities with curved boundaries are better approximated by tri­
angular elements than by quadrilateral ones.
For the same number of nodes, triangular elements provide great­
er flexibility than quadrilateral elements in distributing the nodes
inside the calculation domain.
General quadrilateral elements necessitate the use of isoparametric
transformations [12] , but linear triangular elements are free from
this complication.
Computer-Aided Mold Design 635

F ig . 1Í» Illustrative three-node triangular elements and a polygonal


control volume.

In the present implementation, a control volume approach [13,14]


is used for the numerical formulation, which lends itself to a direct
physical interpretation of mass balance; it is discussed in the follow­
ing sections. The surfaces of an injection-molded part are first di­
vided into three-noded triangular elements. Following this, the cen­
troids of the elements are joined to the midpoints of the three corres­
ponding sides. This creates polygonal control volumes that surround
each vertex node. A sample domain discretization is shown in Fig. 14;
the solid lines denote the element edges, the dashed lines represent
the control volume boundaries, and the shaded area shows the control
volume associated with a node N. The thickness of an element is spe­
cified at the three nodal points and linearly interpolated within the
element. In order to accommodate a possible sudden change of cavity
thickness, the nodal thicknesses are assigned separately for each ele­
ment; in other words , the thickness may be discontinuous across in­
terelement boundaries.
Circular tubes are used to represent the delivery system. A tubu­
lar element is defined by two vertices, with the distance between these
two nodes being the length of the element and a radius. For polymer
636 Wang and Wang

melt in noncircular tubes, the same approximation is employed as in


the runner design described earlier, where a noncircular tube is rep­
resented in terms of an equivalent circular tube having the same hy­
draulic radius but with the volumetric flow rate scaled down to give
the same average velocity. It has been shown that this simple cor­
relation works reasonably well even in nonisothermal cases [15].
Pressure Field
It can be seen from Fig. 14 that the triangular elements adjacent
to grid point N accommodate portions of the control volume and the
control volume boundaries. Since the polymer melt is assumed to be
incompressible, the mass conservation for a control volume is derived
by adding the contributions of these elements to the total mass flow­
ing across the control volume boundaries. In particular, if the tem­
perature field and melt domain are known at a given instant, the pres­
sure within an element 1 can be approximated by three nodal pres­
sures and corresponding linear interpolation functions. Moreover,
for small elements it is reasonable to take a constant flow conductance,
S (see Chapter 1), over each element, evaluating it at the centroid of
the element. The gap wise-averaged volumetric flow rate then becomes
a constant for each element, and the mass that flows into each control
volume can be simply evaluated from a line integral along its bound­
aries. The net flow, qj, entering the control volume of node N from
an element 1 can be represented by

m=l
2 D.
im N’
i = 1, 2, or 3 ( 12)

where N = N ELN O D(i,l), N' = NELNOD(m ,l), and Dimd) is the influ-
ence coefficient of the nodal pressure to the net flow in element 1. In
addition, from the conservation of mass, the net flow to the control
volume N should be equal to zero (except for partially filled regions,
as described in the next section); i.e.

D. = 0 (13)
im N'
r V m=l

where 1’ is over all the triangular elements containing node N. It can


be further shown [13] that the control volume formulation for this
diffusion-type problem results in discretization equations identical to
those which arise from using Galerkin or variational techniques. How­
ever, the former scheme facilitates a physical interpretation of mass
Computer-Aided Mold Design 637

conservation within the control volumes which can be readily applied


to complex three-dimensional mold geometries. In particular, addi­
tional coordinate transformations are required to transform the nodal
coordinates of an element from a global system to a local (in-plane)
system, with the discretization equations of the net flows, qi^s in
Eq. (12), being derived for this latter system. In a similar manner
a tubular element can be divided into two halves, each of which is as­
sociated with the control volume of the vertex node at its end of the
element. For a linear pressure approximation and a constant flow
conductance, S , over such elements, the numerical equation of the
net flow in a tube can be represented in a form similar to Eq. (12),
where

%
(1) _ o (l)
s
m=l
R.
im ^ N'
1 or 2 (14)

In conjunction with two-dimensional flow the control volume asso­


ciated with a node is defined as the collection of all subdivided vol­
umes, whether triangular or tubular elements, which are connected
to that node. The mass conservation for one such control volume, at
vertex node N , can be written as

2 s«'> 2
m=l
D.
im
p
^N ,*2F’
2
m=l
R.
im N'
= 0 (15)

where V is over all the triangular elements and 1” is over all the tu­
bular elements containing node N. Equation (15) is a nonlinear equa­
tion since S (l) is itself dependent upon the pressure field and re­
quires iteration for its solution. In the present implementation suc­
cessive underrelaxation is employed for the solution of the pressure
field. Most importantly, the relaxation procedure requires storage
arrays of 0 (n ) in size, where n is the total number of vertex nodes,
compared with O(n^) for direct-solution methods. Thus, it is more
suited for this fixed-grid calculation in which a large number of ele­
ments may be required.
Melt-Front Advancement
It remains next to account for the advancement of the melt front.
Since the mold filling is a transient process in which the melt front
advances with time, all of the governing equations given above hold
only in the melt region. To determine the free boundary at a given
time, several investigators have developed various techniques. In
638 Wang and Wang

particular, Tadmor et al. [18-21] developed a simple approach in their


flow analysis network (FAN ) where the cavity geometry is described in
terms of a fixed grid, and a scalar parameter, f, is introduced for each
rectangular cell. The parameter f gives the ratio of the occupied vol­
ume to the total volume of each cell. Thus, if the pressure field is
known, one can calculate the net flow rate into each partially filled
cell and update the corresponding values of f for an elapsed time A t .
However, further approximation is required in handling those border
cells which contain part of the cavity boundary.
In the present calculation, the melt front is advanced along the
lines of the FAN approach, but the rectangular cells are replaced by
the control volumes described above. In particular, four kinds of
nodes (or their associated control volumes) are defined for the com­
putations, as illustrated in Fig. 15:

1. Entrance node: Node through which the polymer melt enters


the mold.
2. Interior node: Node for which the associated control volume
is completely filled with polymer melt (f = 1).
3. Melt-front node: Node whose control volume is partially filled
with polymer melt (0 < f < 1).
4. Empty node: Node whose control volume has not yet been
reached by the melt ( f = 0).

entrance interior melt-front empty


node nodes nodes nodes

fM 0 < f< l f =0

Fig. 15 Node (control volume) definition for fluid flow computation.


Computer-Aided Mold Design 639

At a given time the melt-front nodes are assumed to have zero gauge
pressure, whereas all interior nodes should satisfy Eq. (15) based
upon the conservation of mass principle. After solving the system
equations for the pressure field, one can calculate the net flow rate
into the control volume of each melt-front node from Eqs. (12) and
(14), and the volume fraction, f, can be updated for a given time in­
terval. The time step is chosen such that only one such melt-front
node gets filled per step, with all of its adjacent empty nodes be­
coming new melt-front nodes. Thus, starting with totally filled en­
trance nodes, the calculation is made such that one control volume
gets filled during each time interval and, from the successive calcu­
lation of pressure and temperature at each time step, the melt-front
nodes advance until the whole cavity is filled.
Apparently, this approximated melt front and its advancement de­
pend upon the number of elements and their shapes. However, from
a separate study it has been found that, for a ’’reasonable” number
of elements, the overall melt-front movement is essentially indepen­
dent of the element shape even though the front may spatially oscil­
late. From the same study it has been found that equilateral triangu­
lar elements give the least oscillation in the shape of the advancing
front; accordingly, fewer elements are required to give a ’’good” de­
scription of melt-front advancement when equilateral triangular ele­
ments are employed. Since the number of numerical calculations for
the pressure and temperature are on the order of the required time
steps (or the number of control volumes), minimizing the number of
control volumes and elements becomes very important in applying this
calculation to complex mold geometries.
Temperature Field
In the present formulation the temperature profile in the gapwise
direction has to be determined such that the flow conductance S in
Eq. (15) can be evaluated to solve the pressure field. Hence, a
finite-difference grid is employed in the gapwise direction (or radial
direction for tubes) to store the temperature profile. A variable mesh
is introduced at each vertex node by means of ZNDM(J) , J = 1 ,2 ,... ,
N2P1, which denotes the nondimensional transverse coordinate nor­
malized to zero at the centerline (J = 1) and 1 at the wall (J = N2P1).
Since it is difficult to describe the temperature profile at a vertex
node where there may be an abrupt change in cavity thickness or a
juncture between tubular and triangular elements, one temperature
profile is assumed for each vertex node, which can then be scaled for
each of its adjacent elements according to the respective thickness or
radius. In addition, the normalized temperature profile is linearly
interpolated within an element for calculation purposes.
In particular, the temperature profile is determined by solving the
energy equation at each vertex node. However, since the contribution
640 Wang and Wang

of the thermal conduction, convection, and viscous-heating terms to


a given vertex node may not be continuous across interelement bound­
aries, an averaging process for such contributions is necessary. The
current procedure is to obtain a weighted average of such energy
terms for each vertex node based upon the volume fraction which
each element contributes to the associated control volume. To eval­
uate the temperature at a normalized coordinate ZNDM (J), we divide
the control volume of a vertex node into N2P1 ’’layers” (or sub vol­
umes) lying between the midpoints of successive ZNDM(J) values.
The averaging process for a ZNDM(J) coordinate now becomes the
weighted average from the corresponding layer of all its adjacent ele­
ments based upon their effective volume fractions to the subvolume.
As an example, the heat conduction term at layer j of an element 1 is
evaluated at the centroid of the element, denoted by [k (3 ^T /3 z^)]j(l),
and weighted by the volume fraction of its associated subvolume.
Hence,

( 1’)
i!r \ (16)
N,j (V ) V
I . Id
1’

where N = NELNOD(i,l’) with 1’ spanning all the elements containing


node N. The term denotes the volume contribution from the
jll^ layer of element 1’ to the corresponding sub volume at node N.
On the other hand, in order that the numerics be stable, an ’’up­
wind” procedure is employed in handling the thermal-convection ef­
fect; i.e ., at each vertex node only the contributions coming from
the adjacent upstream elements are included in the averaging. There­
fore the following variable is introduced:

V. . ( 1) .if node N lies downwind of element 1 at t = tv


i ^
( 1)
(17)
Id
^0 otherwise

The thermal convections u(3T/3x) and v(3T/3y) are, again, evalu­


ated at the centroid of each element, and the contribution to the sub-
volumes is given in the same form as in Eq. (16), except that the
V i , j ( l ) ’ s are replaced by A similar procedure is employed in
handling the viscous-heating term.
In summary, an implicit method is used for the conduction term,
but the convection and viscous-heating terms are evaluated at the
Computer-Aided Mold Design 641

^ Part
Geometry
Model

----J
| “ Contour Plots
— ^
ECOLOR PCOLOR [TCOLOR -C u rv e Plots
File File F ile
-W ire -F ra m e Drawings

I I — Program
REPLAY — Shaded Color
Images f I -- Data File

Fig. 16 System’diagram for C IMP-FLOW 3D.

earlier time. In the present calculations a relaxation method has been


employed. Typically, a very small Atj^ is employed due to the melt-
front advancement procedure such that only one or two iterations per
time step are required for convergence of the temperature field.

System Description of CiMP-FLOW3D


The flow analysis program CIMP-FLOW3D developed at Cornell Univer­
sity has been implemented on several superminicomputers. As shown
in the system diagram (Fig. 16), the program starts with the part
geometry model which defines the shape of the cavities and associated
delivery system. Based on the geometric information, a preprocessor
generates finite element meshes automatically over the entire system.
The resulting mesh configuration includes the number of elements.
642 Wang and Wang

element types, and their nodal coordinates, together with the dimen­
sions of the elements, such as the thickness of a triangular element
and the perimeter and cross-sectional area of a tubular element. Other
input data fed into the program are the materials properties (melt den­
sity, thermal conductivity, and specific heat, all treated as constants,
together with shear viscosity as a function of shear rate, temperature,
and pressure in terms of the five-constant model described later) and
the process conditions, such as volumetric flow rate, injection tem­
perature, and mold temperature. A material data bank has been es­
tablished which supplies all material information needed for flow anal­
ysis, and it can be accessed through interactive graphics. To select
process conditions, C-FLOW, a commercial version of GIMP-FLOW3D,
displays a panel on the CRT so that the user can pick the process
conditions interactively from the panel on the screen.
The analysis program produces a great number of various outputs
at each time step, namely, the pressure and the temperature profile
at each node and the shear-rate, shear-stress, and velocity profiles
at the centroid of each element. Two device-independent graphic
post-processors have been implemented to display these stored re ­
sults. In particular, an interactive post-processor, PLOT3D, pro­
vides the wire-frame drawings of the molded part and performs two
functions :

It positions the angle and distance to view the part via rotation,
translation, and zoom operations.
It displays such calculated results as the melt-front advancement,
the pressure and bulk temperature contours on the three-dimen­
sional molded part as well as the pressure and bulk temperature
history at specified points, and the required clamp force versus
time.

All the operations in this interactive program are either driven through
key-in commands on the keyboard or by a menu displayed on the screen
in C-FLOW.
For three-dimensional complex geometries, a wire-frame drawing
would become ambiguous and hard to interpret; in addition, such
drawings cannot provide an overall display of the distributions.
Therefore, an alternative post-processor, COLOR3D, was developed
to generate shaded pictures on a raster display unit so that the pres­
sure and bulk temperature distributions could be ’’seen’’ in color on
the three-dimensional molded part surfaces. A color map with differ­
ent hues is defined to represent the various levels of pressure and
biilk temperature. The hue of a pixel within an element is determined
from a linear interpolation of the pressure or bulk temperature values
at the nodal points in order to give smooth contours of the results.
Computer-Aided Mold Design 643

Furthermore, a shading model is employed to determine the intensity


of each pixel. The simplest diffuse reflection model, based upon Lam­
bertos cosine law , is applied in this implementation to provide the
shaded effect on the images. Finally, replay of such images at se­
quential time steps gives an animationlike display of mold filling which
is very helpful in visualizing the filling process. However, to ana­
lyze a complex three-dimensional part could require many minutes or
hours of computational time on a superminicomputer. For fast re­
sponse a less accurate version, FLOW3DEZ, was developed to provide
a quick display of flow front movement by treating the problem as a
Newtonian fluid under isothermal conditions. This version gives rea­
sonable first approximation of weldline formation and, hence, can be
used as a convenient tool at the initial stage of mold design to estab­
lish the number and locations of gates.

Illustrative Examples
For illustration, two examples are given. One is of a flat plate with
three holes at various places; the other is of an instrument panel (or
dashboard) for an automobile, which has a very complex shape. Ex­
perimental results from the first example are used for comparison with
the predictions, but predictions only are given for the second example.
Rectangular Cavity with Two Gates and Three Inserts
Figure 17(a) is a series of superimposed short-shot sequences ob­
tained from a set of experiments by controlling ( i . e . , increasing) the
flow volume at each shot. The cavity is fed by an artificially balanced
runner system so that the flow front will reach both gates at the same
instant. The lines shown in the picture are flow front positions from
both gates at various time intervals. Both flow fronts were split by
the rectangular inserts and formed a total of three weldlines at the
central portion of the cavity; the fourth weldline was formed after
the circular insert.
Figure 17(b) shows the cavity and runner geometry, which has
been approximated by 569 triangular elements and 18 tubular ele­
ments, all shown in shrunken form. The trapezoidal runners have
been modeled in terms of circular tubes based upon their correspond­
ing hydraulic radii as discussed earlier; in the present case the flow
area of the left (right) arm is 0.75 cm^ (2.12 cm^) and the perimeter
is 3.508 cm (5.9 cm), resulting in a hydraulic radius of 0.428 cm
(0.719 cm). The cavity has planar dimensions of 74.3 cm x 11.l cm
with two fan gates at the inlets. The fan gates have been repre­
sented by triangular elements with variable thicknesses at the nodal
points, such that the thickness varies between 0.78 cm (1.38 cm) and
0.128 cm (0.128 cm) in the left (right) arm; all the remaining triangu­
lar elements have the same uniform cavity thickness of 0.336 cm.
Wang and Wang

Fig. 17 A detailed example: (a ) experimental short shots superim­


posed; (b ) triangular and tubular representation of molded part ge­
ometry; (c ) predicted melt-front advancement; (d ) comparison be­
tween experimental (cross symbols) and predicted (dotted symbols)
weldlines; (e ) comparison between measured (solid lines) and calcu­
lated (symbols) pressure traces.

The experiment was performed on an 800-ton Beloit injection-mold­


ing machine with three flush-mounted Dynisco pressure transducers
[locations indicated in Fig. 17(b)] in the cavity. ABS (Lustran Q-
714/Monsanto) was used in the experiment, with the viscosity being
represented by the following five-constant model:
C om pu ter-A id ed Mold Design 645

(d)

(e ) 4.00

Time ( sec )
646 Wang and Wang

no(T, p)
n(Y, T, p) =
1 + (noY/T*)

no(T, p ) = B ex p | ^ j exp(3p) (18)

where

n = 0.298, Tb = 11,400 K, B = 1.46 x 10'^ g/(cm-sec)

T* = 3.699 X 10^ dyne/cm^, 3 = 3.0 x 10“^ cm^/dyne (19)

with the thermal properties of the melt taken as

p = 1.02 gm/cm^, Cp = 2.4 x 10^ erg/(gm «k)

k = 1.8 X 10** erg/(sec-cm*K) ( 20)

The operating conditions for the calculation are

Q = 90 cmVsec, T q = 498 K, = 313 K ( 21 )

where the melt temperature, T q, was determined by means of ”air shot”


measurements, and the wall temperature, Tw, was taken to be the out­
let coolant temperature.
Comparison of Figs. 17(a) and (c ) indicates a very good qualita­
tive agreement between the experimental short shots and the predicted
melt-front advancement. In the actual experiments the velocity con­
trol ended after the lower right corner of the plate was filled; on the
other hand, the simulation assumes a constant volumetric flow rate
throughout, such that the flow rate is overestimated during the lat­
ter half of the filling process. Nevertheless, the melt fronts still show
good agreement in their shapes. Furthermore, Fig. 17(d) shows an
enlarged center portion of the plate in which the locations and shapes
of the three weldlines are seen to agree fairly well, even though the
shortest weldline at the lower left has some discrepancy in its loca­
tion .
For the pressure traces from the flush-mounted pressure transdu­
cers, comparisons are made in Fig. 17(e) for the period during which
velocity control was imposed in the experiments. The predicted pres­
sure traces for transducers 1 and 3 are seen to give good agreement
with the measured values, whereas the predictions for transducer 2
are seen to be consistently low by around 20%. This latter discrep­
ancy can be explained by comparing Figs. 17(a) and ( c ) , where it is
seen that, during the early stage of filling, the melt-front advancement
Computer-Aided Mold Design 647

from the left arm is underpredicted. In turn^ this discrepancy in the


distribution of volumetric flow rate between the two runner arms dur­
ing the early phase of cavity filling could be due to the neglect of
juncture losses in the sprue-runner-gate assembly. That is, where­
as the present results indicate that juncture losses may be important
for predicting the melt-front advancement during the early filling
stage, the fact that the final weldline locations are quite well pre­
dicted indicates that such juncture losses are relatively unimportant
during the latter phase of cavity filling when such juncture losses are
dominated by the cavity pressure drop. From a practical point of
view, then, the present model may be adequate for predicting the flow
pattern in three-dimensional thin parts even though the juncture
losses are not included.
Instrument Panel of an Automobile
The second example is given to demonstrate the ability of the sim­
ulation model to handle a part of great geometric complexity. This
particular automobile dashboard has four sprues leading through four
separate runner systems to a total of 17 gates. Figure 18 shows the

Fig. 18 Part model with automatically generated finite element meshes


for an automobile instrument panel.
648 Wang and Wang

Fig. 19 Simulated melt-front advancement.

finite element mesh of the dashboard. In particular, the cavity has


been discretized into about 1500 linear triangular elements, whereas
the sprues and runners have been approximated by circular rod ele­
ments. The cavity thickness is assumed to be uniform (0.3 cm).
The flow analysis program FLOW3D has been used to simulate the
injection-mold filling with ABS whose viscosity data is drawn from
the data bank. By making use of the graphic post-processor PLOT 3D,
the predicted melt-front advancement indicates the weldline locations,
as shown in Fig. 19. In addition, the program calculates the re­
quired inlet pressure to fill the mold, clamp force to prevent the
mold from flashing, and other useful information. For clarity the
pressure and bulk temperature distribution in the cavity during
any instant of filling can be displayed in color-shaded pictures.
Figure 20 is a black-and-white print of an originally colored image
of the temperature distribution in the cavity at the time of fill. A
color scale is given on the screen for convenience in interpreting
the results.
Computer-Aided Mold Design 649

Fig. 20 Pressure distribution in the mold cavity at the time of fill dis­
played in color-shaded picture.

C. Cooling System Design Program w ith T h r e e -


Dimensional B ou n dary Element Method

General Concept
A great many fundamental details regarding the cooling system design
were given in Chapter 8. It is well recognized that an optimal design
of a cooling system is of great importance because it affects the final
quality of the molded part, the cycle time, and hence the productiv­
ity. Therefore, two criteria are generally of major concern in the de­
sign of cooling systems: one is to achieve uniform cooling so that the
thermal residual stresses and warpage can be minimized; and the other
is to minimize the cycle time.
There are several commercial software packages available for cool­
ing system design, as described in Chapter 8. They are based on
either a one-dimensional heat transfer analysis using an analytic solu­
tion or a two-dimensional analysis using the finite element method. In
this section the program CCSDP (Cornell Cooling System Design Pro­
gram) which uses the three-dimensional boundary element method
(BEM) is briefly introduced [22] .
It is well known that the exact heat transfer analysis during con­
tinuous injection-molding operation requires a solution of three-di­
mensional diffusion equation for the mold, coupled with the transient
heat transfer for the polymer melt in a cyclic manner. Even though
such a rigorous numerical analysis would be possible, at least in
650 Wang and Wang

principle, it would cost too much to be used as a practical tool for


cooling system design in the injection-molding industry. However,
because a large number of design parameters are involved and the
criteria for optimum design are often unclear, the usual trial-and-
error procedure is also not desirable. Therefore, a computer-aided
cooling system design program, even based on a steady-state analy­
sis, should help reduce the time and cost required for designing a
satisfactory mold. The concept presented here is a three-stage de­
sign strategy described as follows [23]:
First stage: Initial design with one-dimensional approximation
Based on an idealized geometry and energy balance, this design
stage shall provide mold designers with the following rough initial de­
sign parameters.

Pitch and depth of cooling channels


Diameter of cooling channels
Coolant flow rate
Cooling time, etc.

One can also evaluate the head loss of coolant in cooling channels.
Second stage: Optimum design with two-dimensional approximation
With the initial design parameters known from the first stage, two-
dimensional BEM analysis can be interactively carried out to check the
uniformity of temperature and heat flux over the cavity surface of a
two-dimensional cross-sectional mold geometry. At this stage the de­
signer has the following options to choose for design modification if
the previous result is not satisfactory:

Change location of channels


Change size of channels
Add or delete channels
Change coolant flow rate in channels
Change coolant temperature in channels
Change injection melt temperature
Change ejection temperature
Change given wall temperature
No more change

Such modifications can be interactively made through a keyboard and/


or a graphic cursor. This iterative design process can help design­
ers improve the cooling system with the aid of interactive computer
graphics. If the mold is considered to be reasonably well approxi­
mated in two dimensions, then the designer can stop the design pro­
cess at this second stage.
Computer-Aided Mold Design 651

Third stage: Three-dimensional cooling analysis


The three-dimensional analysis using BEM would be used as a fi­
nalizing design step so that the design parameters determined by the
previous stages can be checked in terms of the total heat balance and
the uniformity of temperature and heat flux over the entire cavity sur­
face. This analysis requires three-dimensional mold geometry repre­
sentation , boundary element discretization over the mold surface, in­
cluding cooling channels, and three-dimensional presentation of the
results of analysis. Thus, in addition to the main analysis program,
several interface programs have been developed to incorporate some
existing CAD/CAM systems. Obviously, this final stage analysis re­
quires more computer time, but it provides the most realistic results
of the heat transfer analysis. This section describes CCSDP as a
third-stage design tool.

Mathematical Basis
After a certain transient period, the continuous injection-molding op­
eration may result in a cyclic heat transfer behavior in a mold. The
cycle-averaged temperature is assumed to be represented by a steady-
state heat conduction equation

V^T = 0 for X £ ( 22 )

with proper cycle-averaged boundary condition over the boundary


surface of a mold such as

for X e S'];«

3T 3T
for X e Sn (23)
9n 3n

-k = h (T T„) for X e Sh
3n

where T represents the temperature field, 3/3n denotes the outward


normal derivative on the mold surface, k is the thermal conductivity
of the mold material, and h is the heat transfer coefficient between
the mold and ambient environment.
In solving Laplace^s Eq. (22) with BEM, we can first transform the
differential equation into an integral equation by using Greenes second
identity :

/ TV^G dn = / GV^T da + J (t ^ - G ds (24)


652 Wang and Wang

Choosing G as the usual fundamental solution defined by

V^G = 6(x - xO (25)

i.e .,

1
G = (26)
4ïïr

where

r = / (x - O + (y - n)^ + (z - (27)

We can write Eq. (5) as

^ 3 (l/ r) ds (28)
a T (x )
3n an

where a denotes the internal angle formed by the surfaces at a given


point (2Tr for any point on a smooth surface and for any internal
point).
After discretization over the surface of the mold, Eq. (28) is ap­
plied to each element (constant quadrangles have been employed in
the present case study) to obtain the system of linear algebraic equa­
tions

[A ]{x } = {f } (29)

To obtain Eq. (29), we introduced boundary conditions such that { x }


represents either unknowns T or 8T/8n for each element. Once Eq.
(29) is solved, the temperature at internal points can be evaluated by
Eq. (28) if needed.
After solving Eq. (29), one can evaluate the total heat transfer
rate from the cavity surface to the mold (Q^) or that from the mold to
the cooling channels (Q 2 )- The effective shape factor, S, then can
be determined from the definitions:

Qi
(30)
^("^melt "^coolant^

. ^ ^ __________
(31)
^(^melt ' ^coolant)
Computer-Aided Mold Design 653

where Tmelt "^coolant represent the injection melt temperature and


coolant temperature, respectively. Any difference between Si and S 2
reflects the heat transfer rate through the surfaces exposed to the
environment. In particular, the shape factor is an indication of the
effectiveness of the cooling system design.
As defined in Eqs. (30) and (31), S has the units of length. In
the simpler two-dimensional case, one would customarily use the heat
transfer rate per unit depth rather than Q in Eqs. (30) and (31),
such that S would then be nondimensional. Therefore, for the three-
dimensional case it seems appropriate to introduce S = S /L, where L
is some characteristic length (e . g . , diameter of cooling channel), such
that the value of S is therefore independent of the overall scale of the
system geometry.

The Program Structure


The program CCSDP is designed to deal with as precise a geometry as
possible for a more rigorous analysis. Because of the complicated
three-dimensional mold geometry, including cooling channels, one
needs computer-aided tools to define and discretize the geometry (as
preprocessors) and to display results in three-dimensional space for
fast interpretation of the results (as post-processors). Therefore,
several interface programs together with a three-dimensional BEM
analysis main program have been developed for some existing CAD/
CAM programs. Figure 21 shows a schematic diagram for CCSDP.
Three-dimensional mold geometry is represented by the solid modeler
CATIA, which generates a boundary file for the mold geometry in its
data base. An interface program (CATBR) between CATIA and SU­
PERTAB has been developed to transfer the boundary file into an ap­
propriate data format for SUPERTAB (i .e ., a universal data format).
Then the universal file, which includes the boundary information for
the solid, can be read in by SUPERTAB, with which one can interac­
tively discretize the surfaces of the mold for boundary elements.
The discretization over the mold surface is evidently much easier
than that for the corresponding three-dimensional FEM solid-type ele­
ment mesh; hence, BEM requires much less effort for discretization.
SUPERTAB generates a final universal file, which includes informa­
tion on mesh configuration such as coordinates of nodal points and
element connectivity. CCSDP also consists of an interface program
for MOVIE.BYU, the post-processor^ and a BEM main program. The
interface program for SUPERTAB reads and interprets the universal
file, whereas the preprocessor draws the wire frame and hidden-line-
removed object of the three-dimensional mold geometry and also inputs
the prescribed boundary conditions. The BEM main analysis program
solves the three-dimensional heat transfer problem in terms of tem­
perature and its gradient over the boundary surfaces. As a result
654 Wang and Wang

Fig. 21 System structure of CCSDP, a three-dimensional BEM cooling


system design program.

of the analysis the effective shape factor is evaluated to indicate the


effectiveness of the cooling system design. Next, the interface pro­
gram for MOVIE.BYU (a color graphics program) prepares the nec­
essary input data for MOVIE.BYU ( i . e . , geometry data files and func­
tion data files). Finally, the color image of temperature and/or tem­
perature gradient over the surface of the three-dimensional mold can
be displayed on a color raster terminal by MOVIE.BYU. These color
images of results will help the designer interpret the result quickly
in terms of uniformity of temperature as well as the heat flux distribu­
tion over the cavity surface.

An Illu s tr a tiv e E xa m p le

To illustrate the three-stage design philosophy, we select a simple


cup-shaped mold cavity plate. Suppose that a one-dimensional pro­
gram has been used in the initial design stage so that a number of
design parameters are determined ( e . g . , the diameter of the cooling
channel, the distances between the cooling channels and that to the
cavity wall, etc. [2 4 ]). In the second stage a two-dimensional BEM
program is called to help the designer make a choice among several
alternatives. Figure 22 shows two different cooling channel arrange­
ments on half of the cavity plate. The temperature distributions along
Computer-Aided Mold Design 655

Fig. 22 Display of temperature distribution at the boundary of a cup­


shaped cavity plate: (a) Initial design with two cooling channels.
Temperature for Index = 1. (b ) Modified design with four cooling
channels. Temperature for Index = 6.

the boundary as a result of the analysis are also represented in the


figure by the relative height of the bars. A similar graphic display
can be obtained for the temperature gradient distributions, which will
show the relative magnitude and direction of heat flow through all
boundaries. Since the two-dimensional program treats a cross section
of the conical cavity as a slab, the temperature distribution on the
sectioned surface can be calculated and displayed as shown in Fig­
ure 23(a).
However, a real cup-shaped mold cavity plate has its cooling chan­
nels led out of the mold plate in the third direction, as shown in Fig.
2 3 (b ), where one-eighth of the three-dimensional mold plate is dis­
played. Also shown in Fig. 23(b) are the isotherms on the sectioned
surface calculated by the three-dimensional program CCSDP. The
temperature contours are designed to be displayed surface by sur­
face to facilitate interpretation in the case of a very complicated mold
geometry. The color image of the temperature field and heat flux over
the three-dimensional mold surface, which is not shown here, can help
the designer interpret the result much faster than the contour plots
can. In this particular cup-shaped mold the results show that two-
dimensional analysis leads to higher temperature distribution over the
656 Wang and Wang

TEMPERATURE CONTOUR ON SIDE NO. - 1

Fig. 23 Predicted isotherms on a surface of interest in a cup-shaped


cavity plate based on (a) two-dimensional approximation, (b ) three-
dimensional true representation. Temperature contour on side no.
(a) and (b ) = 1.
Computer-Aided Mold Design 657

cavity surface than that of the three-dimensional simulation by about


10°C. And temperature over the cavity surface is found to be more
uniform in the three-dimensional case than that in the two-dimensional
case. Such a discrepancy is obviously due to the poor geometry ap­
proximation by the two-dimensional treatment for a three-dimensional
cup-shaped mold. For precision molding where dimensional stability
and thermal residual stresses in the molded part are of prime concern,
a more accurate three-dimensional analysis should be used as a viable
tool for cooling system design.

D. Mold Assembly Design

In a typical mold design procedure, one has to select a proper type


of mold assembly and sizes of its components after the layout of the
runner-gate-cavity system and the cooling system design are com­
pleted. In general, this task can be accomplished by selecting com­
mercial standard mold assemblies which have the advantages of low
cost and ease of replacement. A few commercial packages are avail­
able from some turnkey CAD/CAM system vendors. Instead of build­
ing a huge data base to store a large variety of mold components, a
mold assembly design program, MOLD ASM [24], has been developed
using a special algorithm to generate the data base; it is briefly de­
scribed in this section.
Because the mold bases by the DME Co. are the most widely used
in the injection-molding industry, the MOLDASM program was devel­
oped based on DME product information. There are seven basic ser­
ies of mold assemblies in the DME catalog: The A-series, AR-series,
and B-series are so-called two-plate mold assemblies, whereas the X5-
series, X6-series, and AX-series are three-plate mold assemblies, and
the T-series is a four-plate mold assembly. Depending on the number
of mold plates, a mold assembly can have one, two, or three parting
planes. An industrial survey has shown that about 70% of the molds
used in industry are of the two-plate types. Among the two-plate
series, the A-series is the most frequently used, since in this case
it is often possible to machine cavities as inserts which can be re ­
tained firmly in the mold plates, where the cavity is located between
the A and B plates.
The selection procedure starts by specifying the desired series
type of mold assembly. The next step is to enter the overall length
and width of the smallest rectangular area that completely covers all
of the cavities in the mold. The size of mold base is selected with the
following two constraints:

The width of ejector plate must be greater than or equal to the


overall width of the runner-cavity system
658 Wang and Wang

The center-to-center distance of leader pins along the length of the


mold assembly must be greater than or equal to the overall length
of the runner-cavity system.

The first constraint accounts for the need to locate all ejector pins
within the ejector plate and, as such, the ejector plate should be just
large enough to cover the cavity areas where ejector pins may be lo­
cated. The second constraint ensures that the mold plates be large
enough to contain the cavity and the leader pins.
There are several possible mold base lengths for a given mold base
width in the DME catalog. However, for each mold base width, the
width of the ejector plate is fixed. Furthermore, the center-to-center
distance is fixed for a particular mold base. Therefore, once the width
of the ejector plate is determined, the width of the mold base can be
obtained. Based on this width, the length of the mold base can then
be selected to satisfy the second constraint. If the overall length of
the cavity is longer than all the possible center-to-center distances of
the leader pins for the selected mold base width, the selection process
is repeated with a larger mold base width, for which the corresponding
lengths are also larger.
Data tables that contain the standard ejector plate widths and cen­
ter-to-center distances are searched until the above two constraints
are satisfied. Since the dimensions are coded in an ascending order
of magnitude and the search always proceeds in the same direction,
the selection of the smallest possible mold base satisfying the above
two constraints is assured. The search also provides pointers which
locate the required width and length of the mold base from the data
tables. The same pointers are also used to select, from appropriate
data files, the sizes of other components that depend on the size of
the mold base.
Additional information required from the designer are the mold plate
thickness, which should be determined based upon the cooling system
layout, and the required strength of the mold plate. Possible thick­
nesses of the mold plates are displayed on the screen, and the de­
signer is asked to select the desired sizes. A stress-strain analysis
can be employed to determine the stresses and deflections for various
plate thicknesses. Once the thickness of all the plates has been de­
termined, the lengths of various pins and shoulder bushings are se­
lected such that they are less than the sum of the thicknesses of the
corresponding plates within which they are located.
For the ejector pins the designer has a choice among several ma­
terials. In addition, the various diameters available within the pin
type that has been selected are displayed, and the designer is asked
to choose one. The designer then enters the number of ejector pins
required and positions them at appropriate locations in the ejector
Computer-Aided Mold Design 659

plate. Iterations may be needed to modify the cooling-channel design


due to ejector pin location considerations. When all the selection pro­
cedures are completed, the program prints out a complete parts list,
the total price, and mold assembly drawings, as shown in Figs. 2
and 6.

E. C o m p u ter-A id ed Mold M a n u fa c tu rin g

When an NC machine tool is used to cut material from a mold base, a


set of specific commands has to be fed into the machine controller so
that correct cutter motions will be carried out. These commands, con­
sisting of geometric information and direction of cutter motions, are
generated through programming. To insure that this information (so-
called NC data) is correct, one must verify either by graphing the cal­
culated tool path or by actually cutting a real part with a soft material
( e . g . , a piece of wood or styrofoam) for visual inspection.

A u to m a tic N C C o d e G e n e ra tio n

Most CAD/CAM turnkey system vendors offer a module for generating


NC codes for machining. Such a module would use the common geo­
metric data base employed by CAD modules, which is typically in wire­
frame representation. Therefore, most systems can only handle two-
dimensional or 2-1/2-dimensional (break the third dimension into steps
rather than continuously) machining. For more complicated machin­
ing which may require three to five-axis motion (one rotational axis
on machine table and another on tool axis in addition to the three
primary axes in a cartesian coordinate system), the state of practice
in industry today is to use the APT (automatically programmed tools)
system, a dedicated computer software to generate NC codes.
Although APT is a powerful well-tested system under continuous
development since the early fifties, it requires manual preparation
(so-called part programming) by writing hundreds of lines of code in
the APT language. This task demands specially trained people and
is frequently subject to human error. Furthermore, since it is a ded­
icated system for NC machining, it is rather difficult to integrate in­
to a CAD/CAM system to achieve a higher degree of automation.
The technology of solid modeling provides an opportunity for im­
proving the situation. Conceptually, a machining operation is just
to remove the unwanted volume of material from a given solid object.
Therefore, if a Boolean difference operation (subtract a volume from
another volume) can be performed on those geometric entities, the ma­
chining operations can then be simulated by a computer. From another
viewpoint one can consider machining as an operation to generate a
new set of surfaces which bounds the volume of the desired machined
660 Wang and Wang

part. Topologically, those are ’’faces” of the solid object; they can
be well defined if the object is represented in the computer with a
unique boundary field.
Currently, research and development efforts are being made to
generate NC codes automatically by using solid modelers. One such
effort is illustrated in Fig. 24(a) in which a solid object (e . g . , a
mold cavity plate) is represented in a boundary file, or its bound­
ary representation can be obtained from a CSG representation [25].
Using a cursor (e . g . , a light pen), one can pick a face on the CRT
screen as the part surface in APT terminology which will control the
tool motion to generate the desired surface as a result. Instead of
generating NC codes directly, this system consists of an APT source
code generator, as shown in Fig. 25, in the form of a flowchart. An
optimization scheme is incorporated in this system for rough cutting
the pocket in a zigzag fashion. The finish cut is accomplished by
guiding the tool around all boundaries. APT source codes (geo­
metric and motion statements) are automatically generated; they are
then sent for post-processing to get machine-specific NC codes in a
normal way. Figure 24(b) is an example of an actual part cut on a
three-axis computer-controlled NC milling machine. This part was
produced without writing any NC codes.

G ra p h ica l V erifica tion o f N C Tool Path

Before the NC data, generated by whatever method, is employed to


run an NC machine tool, it is usually desirable (sometimes necessary)
to check its validity. The most positive way of verifying the result
is to run it through an actual NC machine tool by cutting a piece of
soft material and then checking the result. This is usually a very
expensive method, particularly if it must go through a few iterations.
A more practical way commonly used in industry is to plot the con­
secutive cutter locations (so-called CL data) on a CRT screen to
check its correctness. Some CAD/CAM system vendors also offer a
more elaborate graphical verification system by displaying the con­
secutive tool axis positions in a three-dimensional space. However,
such displays are often in wire-frame form, which cannot show the
real dimensions of the cutting tool as well as the part being machined.
The results are often ambiguous, hard to interpret, and thereby not
reliable.
New graphical methods to verify the tool path more reliably are
being developed by solid modeling techniques [26,27]. When the ma­
terial before machining (e . g . , a casting, forging, or just a simply-
shaped flat stock) and the swept volume generated by the programmed
cutter motion are both represented in solids, the machined part can be
obtained graphically by subtracting the cutter swept volume from the
stock by a Boolean operation. One such system has been developed
Computer-Aided Mold Design 661

(a)

(b)

Fig. 24 A sample cavity plate with three-dimensional feature: (a)


display of boundaries on the screen based on solid modeling, (b ) ac­
tual part machined on a CNC milling machine.

and is being used in production. It has achieved a real-time speed


in simulating NC machining graphically on a superminicomputer
[28]. The program, NCS (numerical control simulator), was first
662 Wang and Wang

A P T -C o d e
Generator

F ig . 25 System structure of the NC code generation program.

developed at Cornell University and later extended at the Corporate


Research Center of the General Electric Company. The current ver­
sion can verify three-, four-, and five-axic NC milling operations,
which are widely used in the mold/die and aerospace industries. It
also handles the three most commonly used tools: the flat-end, ball-
end, and fillet-end mills. Figure 26 shows a graphic display to verify
the NC machining operation for an injection mold. The machining con­
sists of more than 200,000 steps of cutter motion to finish the job. All
intermediate steps during the machining operation can be displayed
for examination.

F. In te g ra te d System fo r Mold Design and M a n u fa c tu re

C A D IC A E IC A M System Concept and S tru ctu re

As the basic knowledge regarding molding processes has improved and


various computer software has become more available, the establishment
Computer-Aided Mold Design 663

F ig . 26 Display of graphic verification of NC code for machining an


injection mold for a backpack frame at four intermediate steps.

of an integrated system involving all aspects of the molding operation


now seems more feasible. For injection molding, Fig. 27 suggests one
computer-based system for mold design and manufacture. Interacting
with the computer through a workstation, users can call upon any
software module at their fingertips for various purposes. In mold de­
sign, for instance, one would like to analyze the flow behavior in or­
der to help the designer make a rational decision on the layout of a
runner-gate-cavity system. Flow simulation programs are available
for this purpose, with the consideration of molding machine capacity
and material properties. Other programs, as described previously
in this chapter, can also be called through integrated program control.
To represent the shape of the part or mold bases, solid modelers or
other forms of representation best suit its functional requirement.
Output from such a system can be in many forms: tables, graphs,
wire-frame displays, or color-shaded pictures.
The system discussed above is only concerned with the design and
manufacture of molds, i.e. , the tooling for producing plastic parts.
Typically, the design of plastic parts itself is handled separately by
a product design engineer. For functionally critical parts, CAE tech­
niques , such as finite element methods for stress and thermal analyses,
664 Wang and Wang

are commonly used. However, such analyses are usually handled by


treating the material with homogeneous, isotropic, and linear proper­
ties. It is well known that the injection-mol din g process causes ori­
entation of polymer molecules due to shear and elongational flow. A
molded plastic part normally comprises a considerable amount of resid­
ual stress induced by the combined effect of material flow in the mold
and heat conduction through the cold cavity wall. Therefore, distri­
bution and orientation of these stresses depend much on the mold de­
sign and processing conditions. If the nonhomogeneous and anisotropic
properties of the molded part can be predicated either through simu­
lation of the molding process or by some empirical methods, it is then
conceivable that such properties can be incorporated early in the de­
sign stage of the part itself.
Conceptually, a higher-level integration of CAD/CAE/CAM tasks
can be instituted from the design of the plastic part to process control
of the molding machine, as illustrated in Fig. 28. This system starts
from representing part geometry in solids which can be readily used
for structure analysis. Applying Boolean operations, the geometric
Computer-Aided Mold Design 665

CAD CAM

Plastic /G e o m e tric \ Tooling


Parts 1 Modeling \ ^ _ Design &
Design y Systems j Manufacture

Injection Molding Machine

Machine
Data File

Fig. 28 Components of an integrated CAD/CAE/CAM system for de­


sign and injection mold of plastic parts.

data base established for part design can also be used to create the
cavities in the mold base. Solid modeling can be further used to con­
struct the mold assembly, design the cooling system, and analyze the
rigidity of mold bases if desired. Process simulation plays an impor­
tant role throughout the system, from determining mold design param­
eters, predicting mechanical properties, to providing information for
setting up the process controller. As mentioned earlier, the simula­
tion package C-FLOW, developed by Advanced CAE Technology, Inc.,
incorporates a graphic emulator of a process-control panel from which
the user can pick setup points graphically and get the simulated re­
sults immediately.
666 Wang and Wang

Most turnkey CAD/CAM system vendors offer systems with various


degrees of integration and capabilities. An experimental system using
a solid modeler and based on a similar concept as described above has
been developed at the General Electric Company. One example deal­
ing with a plastic backpack frame produced by injection molding was
reported as a news article in the August 1985 issue of Mechanical En­
gineering magazine. According to the report, G.E. mechanical engi­
neers started with a conceptual design of the backpack frame. In­
stead of making engineering drawings and blueprints, they created
a solid model using the solid modeling system TRUCE , developed at
the G.E. research center. TRUCE can handle not only regular ob­
jects with analytical surfaces but also those with sculptured surfaces
which are used extensively for modeling the backpack frame.
To define the mold, we subtract the part shape represented in the
solid from the solid mold block, which is then split into halves with a
parting plane. TRUCE used a finite element mesh of the surface to
refine the part and conducted a stress analysis to test structural
performance. From the geometric data base produced by TRUCE,
NC codes were generated for machining the mold cavity. G.E. en­
gineers used the NCS program to verify the tool path graphically,
as shown in Fig. 26, and also cut a piece of high-density foam for
the same purpose before the actual metal mold was machined.

D oto B a s e s

The importance of the data base and its efficient management for any
computer-based system cannot be emphasized enough. This is par­
ticularly true in an integrated system for a physical manufacturing
process like injection molding, which involves a vast amount of data
characterizing the complex, viscoelastic properties of the polymeric
materials. With reference to Fig. 28, there are at least three cate­
gories of data involved in injection molding. The geometric data base
consists of all geometric or topological information representing the
shape of the plastic part and the tooling to make it. Its format will
be dictated by the type of modeler used and its data structure. An­
ticipated applications of these data also play a role in designing the
data base.
The data base of material properties is the most important of all.
Without reliable property data in an appropriate form, the results
from the analysis would be erroneous and misleading. Depending on
the application, various types of data are needed. For instance, to
support flow simulation in injection molding, one example [29] contains
a data base involving hundreds of specific materials, and there are 12
attributes for each of them. These attributes consist of unique ma­
terial identification, generic name, manufacturer’s name, and rheo­
logical and thermal properties of the material. A dedicated data
Computer-Aided Mold Design 667

base management system was also developed in this case. As one com­
ponent in an integrated system, the data base manager can not only
retrieve the existing data from the data file but also add, delete, or
update any information in the data file. To interact with the data
base, the system provides a user-interface in this case to help the
user select the desired type, search for appropriate data, and dis­
play information automatically.
Other types of data bases may be needed for certain applications.
For instance, at a given injection-molding plant with a given number
and type of machines, the mold designer has to consider which ma­
chine or a number of machines on which the mold being designed can
be used without exceeding its capacity in terms of such values as
clamp force, shot size, or injection pressure, etc. In this case a
data file containing pertinent information regarding all facilities in
the plant should be available and accessible to the mold designer.

IV . C O N C LU D IN G REMARKS

It seems clear that the trend of rapid advancement of computer tech­


nologies will continue. The molding industry is facing the challenge
of how to take advantage of the powerful tools and make plastic
products even more attractive and cost effective. It has been gen­
erally recognized that an integrated CAD/CAE/CAM system as illus­
trated in Fig. 28 would be a logical approach to drastically enhance
the engineer's capability to handle all aspects of the production pro­
cess. However, to establish such a system requires great effort in
developing application software. Furthermore, developing viable
software, particularly for simulating physical processes, requires
basic scientific knowledge of the processes. Unfortunately, this
key point has been neglected by many people in practice who might
have been misled or deluded into thinking that ’’the computer will
do the magic.”
For the future it is our belief that even greater efforts are need­
ed to improve our scientific knowledge of materials and the methods
to process them . Such a knowledge base has to be resolved to such
a level that it can be implemented in a computer-based system. The
implementation could involve vigorous analyses of the interdisciplin­
ary nature as well as empirical methods. The latter may be needed
for those problems where physical laws do not exist or where such
laws are unable to characterize the problems. Under such circum­
stances the software will probably not only include extensive engi­
neering analysis packages as described in this chapter, but also the
knowledge base, which can be acquired and manipulated using arti­
ficial intelligence techniques. Another point the authors would like
668 Wang and Wang

to stress is the importance of computer graphics. Without a powerful


graphic user-interface, an integrated CAD/CAE/CAM system can never
be efficient and become a viable tool for industry.

REFERENCES

1. K. Ueda, Proc. Precision Eng. Conference, Raleigh, N .C ., p.


9, 1985.
2. DME Company Mold Base Components Catalog, Vol. 1, 1985.
3. E. C. Bernhardt, E d ., Computer-Aided Engineering for Injec­
tion Molding, Hauser Verlag, 1983.
4. K. K. Wang et a l., Computer-Aided Injection Molding System,
Progress Report Nos. 1—11, Cornell Injection Molding Program,
Cornell University, 1975—1985.
5. I. E. Sutherland, SJCC, 329 (1963).
6. J. D. Foley and A . Van Dam, Fundamentals of Interactive Com­
puter Graphics, Addison-Wesley, 1982.
7. A. A. G. Requicha and H. B. Voelcker, IEEE Computer Graph­
ics & Applications, 2: 9 (1982).
8. V. W. Wang, K. K. Wang, and C. A. Hieber, Proc. SPE-ANTEC,
29: 663 (1983).
9. C. A. Hieber, Proc. SPE-ANTEC, 28: 356 (1982).
10. V. W. Wang, C. A. Hieber, and K. K. Wang, Proc. SPE-ANTEC,
32: 97 (1986).
11. B. R. Baliga and S. V. Patankar, J. Numerical Heat Transfer,
3: 393 (1980).
12. O . C. Zienkiewicz, The Finite Element Method, McGraw-Hill,
New York, 1977.
13. A. M. Winslow, J. Comput. Phys., 2: 149 (1967).
14. S. V. Patankar, Numerical Heat Transfer and Fluid Flow, Hem­
isphere, 1980.
15. C. A. Hieber, R. K. Upadhyay, and A. I. Isayev, Proc. SPE-
ANTEC, 29: 698 (1983).
16. C. A. Hieber and S. F. Shen, Israel J. Technology, 16: 248
(1978).
17. C. A. Hieber and S. F. Shen, J. Non-Newtonian Fluid Mech.,
7: 1 (1980).
18. E. Broyer, Z. Tadmor, and C. Gutfinger, Israel J. Technol­
ogy, 11: 189 (1973).
19. Z. Tadmor, E. Broyer, and C. Gutfinger, J. Polym. Eng. Sci.,
14: 660 (1974).
20. E. Broyer, C. Gutfinger, and Z. Tadmor, Trans. Soc. Rheol.,
19: 423 (1975).
Computer-Aided Mold Design 669

21. W. L. Krueger and Z. Tadmor, J. Polym, Eng, Sci.y 20: 426


(1980) .
22. K. K. Wang et al. , Computer-Aided Design and Fabrication of
Molds and Computer Control of Injection Molding, Progress Re­
port No. 11, Cornell University, 1985.
23. T. H. Kwon, S. F. Shen, and K. K. Wang, Proc. SPE-ANTEC,
32: 110 (1986).
24. K. K. Wang, P. Khullar, and W. P. Viang, Plastics Eng., 37: 25
(1981) .
25. R. Ferstenberg, K. K. Wang, and J. Muckstadt, Proc. IEEE
Conference on Robotics and Automation, Vol. 1, p. 325, 1986.
26. R. Fridshal, K. P. Cheng, D. Duncan, and W. Zucker, Proc.
Numerical Control Society, 19th Annual Meeting, 1982.
27. W. P. Viang, Solid Geometric Modeling for Mold Design and Man­
ufacture, Ph.D. Thesis, Cornell University, 1984.
28. W. P. Wang and K. K. Viang, Proc. IEEE Conference on Robotics
and Automation, Vol. 1, p. 166, 1986.
29. C. A. Hieber, V. W. Wang, and H. H. Chiang, DATA BANK:
Polymer Melt Viscosity and Thermal Properties, Technical Man­
ual No. TM-5, Cornell Injection Molding Program, Cornell Uni­
versity, 1985.
Author Index

Aoyama, H ., 2, 98, 99, 107, 129,


159, 222, 455, 477
Abolafia, D. R ., 340, 373 Appledoorn, J. K ., 108, 135
Acierno, D ., 2, 98, 130, 162, Aral, T. , 2, 98, 99, 107, 129,
171, 223, 455, 478 159, 222, 448, 455, 477, 478
Acitelli, M. A ., 410, 422, 444, Aramburo, F ., 439, 474
476 Arman, J ., 94, 95, 135
Acrivos, A ., 358, 375 Armstrong, R. C ., 2, 131, 184,
Adams, E. B ., 139-141, 222 186, 189, 190-193, 224, 225,
Adamse, J. W. C ., 93, 96, 97, 455, 478, 531, 559, 563
134 Arp, P. A . , 543, 564
Andrews, R. D ., 243, 260, 324 Asada, T ., 243, 260, 324
Advani, S. G ., 558, 559, 565 Austin, C, 1, 36, 44, 129
Aggarwala, B . D ., 268, 275, 277, Azuma, K ., 252, 325
278, 287, 327, 527, 563
Aklonis, J. J ., 90, 134, 280, 281, B
301, 308, 327, 328
Akovali, G ., 90, 134 Babayevsky, P. G ., 356, 375
Alfrey, T ., 299, 328 Back, P . , 124, 128, 136
Alle, N ., 2, 130 Baer, A. D ., 160, 222, 358, 375
Alves, G. E ., 517, 563 Bagley, E. B ., 96, 135, 157, 222,
Ambrose, R. J ., 379, 380-383, 455, 477
385, 392, 396-409, 412-417, Baird, D . G . , 172, 223
419, 420, 431, 432 Baliga, B . R ., 634, 668
Andersen, D ., 504, 562 Ballenger, T. F ., 2, 92, 107, 129,
Aoyagi, Y. 448, 477 455, 478

671
672 Author Index

Ballman, R. L ., 1, 2, 13, 19, 20, Bolen, W. R ., 392, 432


129, 131, 132, 247, 261, 264, Boles, R. L ., 161, 223, 455, 477
324, 326, 447, 476 Borchardt, H. J ., 340, 373
Bakerdjian, Z ., 2, 92, 130, 247, Boucher, D, F ,, 517, 563
325 Boudreaux, E ., 2, 98, 130
Bankar, V. G ., 2, 130 Bowles, W. A ., 2, 129
Barel, M., 437, 446, 473 Bretherton, F. P ., 542, 564
Barnhart, R. R ., 435, 473 Brindley, G ., 533-535, 563
Barone, M. R ., 498, 507-509, Brinkman, H. C ., 108, 135
514, 516, 517, 520-523, 562, Brizitsky, V. I ., 140, 142-144,
563 151, 222, 254, 325
Barr, G ., 159, 222 Broutmann, L. J ., 242, 262-265,
Barrett, K. E. J. , 333, 372 267, 269, 324, 326
Bartenev, G. M., 269, 273, 274, Brown, P. F. , 493, 562
327, 439, 474 Brown, R. A ., 184, 186, 189, 190,
Barton, J. M., 340, 373 192, 193, 224, 455, 478
Bartos, O . , 2, 102, 103, 107, Broyer, E ,, 44, 53, 133, 447, 459,
129 476, 479, 638, 668, 669
Bauman, A. J., 398, 433 Buchman, A ., 262, 264-266, 269,
Beaumont, P. W. R ., 264, 286, 286, 288, 297, 326, 327
287, 326, 327 Bueche, F. , 394, 395, 432
Benfer, W ., 446, 476 Burkett, S. J ., 345, 374
Bergen, R. L ., 2, 129 Burns, R ., 533, 535-537, 559,
Berger, J. L ., 1, 23, 129, 447, 563, 565
476 Buschhaus, F. , 446, 476
Berman, B. L ., 353, 375 Bush, A. J. , 257, 325
Bernhardt, E. C ., 608, 668 Byam, J. D ., 423, 433, 441, 445,
Bernstein, B ., 439, 474 475, 476
Berry, G. C ., 394, 432
Bidstrup, S. A ., 348, 375
Binnington, R ., 182, 224
Bird, R. B ., 108, 126, 135, 136,
191, 225, 535, 564 Campbell, R. H . , 406, 433
Birley, A. W., 345, 374, 492, 496, Carreau, P. J., 104, 107, 109,
535, 537, 562, 564 135, 161, 162, 223, 455, 478
Blow, C. M. , 462, 479 Carrol, B ., 340, 373
Bobear, W. J ., 407, 433 Casale, A ., 2, 19, 22, 84, 122,
Boger, D. V ., 153, 161, 171, 172, 123, 130
179, 182, 187, 222-224, 455, Castro, J. M ., 345, 374, 459, 479
478 Caswell, B ., 179, 187, 192, 194,
Bogue, D. C ., 2, 92, 107, 129, 223, 224, 329, 344, 474, 478
131, 139-141, 160, 161, 222, Caulk, D. A ., 498, 507-509, 514,
223, 242, 243, 246, 254, 268, 516, 517, 520-523, 562, 563
298, 302, 324, 325, 455, 477 Chandrupatla, A. R ., 448, 477
Boguslawski, J. J., 437, 438, 474 Chang, P. , 186, 224
Author Index 673

Chang, P. W., 455, 478 Collins, E. A ., 2, 19, 23, 120,


Chang, W. V ., 462, 479 121, 131
Chapman, F. M., 440, 475 Collins, M., 160, 223
Chapoy, L. L ., 232, 323 Colwell, R. E. , 392, 432
Charles, M., 96, 98, 100, 135 Cooper, L ., 1, 129, 447, 476
Chattopadhyay, S ., 2, 85, 130 Copsey, C. J. , 232, 323
Chen, I. J., 2, 23, 92, 107, 129, Coran, A . Y. , 403, 406, 433
133 Cox, H. W ., 2, 13, 129, 130
Cheng, K. P ., 660, 669 Cox, W. P . , 87, 134, 395, 432
Cherkasova, L. N ., 353, 375 Coxon, L. D ., 245, 286, 324, 327
Chiang, H. H ., 100, 101, 135, Craig, D ., 412, 433
666, 669 Crawley, R. L ., 2, 105, 108, 131
Choi, S. Y . , 19, 23, 132, 333, Cristy, R. L ., 438, 474
344, 372, 374 Crochet, M. J., 187, 193, 194,
Chompff, A. J . , 231, 323 196, 197, 224, 225, 439, 474,
Chong, J. S ., 160, 222, 358, 375 528, 529, 549, 563, 564
Choplin, L ., 161, 162, 223, 455, Cross, M. M. , 2, 79, 104, 107,
478 109, 131
Chou, C. H., 448, 477 Crouthamel, D ., 198, 225, 245,
Chow, T. S. , 389, 431 262-265, 269, 289, 294, 324,
Christiansen, E. B ., 160, 222, 327, 459, 479, 525, 563
223, 358, 375 Crowder, J. W., 2, 92, 107, 129
Christmann, L ., 19, 23, 132 Crowson, R. J., 167, 223, 441,
Chu, D ., 41, 133, 161, 223 446, 475, 559, 565
Chuck, W., 36, 133 Crowther, B. G ., 437, 438, 474
Chung, B ., 2, 3, 100, 101, 105, Crozier, D ., 345, 349, 374, 375
131, 132, 163-165, 167, 168, Cuculo, J. A ., 2, 98, 130
174-177, 223, 453, 455 Curto, D. , 455, 478
Chung, T. S ., 61, 62, 134, 335, Custer, R . C . , 2, 130
345, 348, 366, 368-370, 372, Cuthrell, R. E. , 345, 374
455, 477, 478 Czarnecki, L ., 167, 223, 559, 565
Clark, E. S ., 231, 247, 323, 325,
443, 475
Coates, A. W. , 340, 373
Cochrane, T. , 191, 225
Cogswell, F. N ., 19-23, 41, 132, Dalai, M. D. , 2, 131
133, 161 Daniels, C . A . , 2, 131
Cohan, L. H., 390, 431 Daniels, F ., 340, 373
Cohen, M. H ., 299, 328 Darlington, M. W., 554, 564
Colbert, G. P ., 423, 433, 441, Daryanani, R ., 108, 135
445, 475, 476 Davis, A. R ., 187, 191, 224
Colburn, A. P . , 573 Davis, C . M ., 394, 432
Coleman, B. D ., 395, 432 Davis, H. L ., 161, 223, 224, 455,
Coleman, C. J., 455, 478 477
Coleman, M. M., 406, 433 Davies, J. M., 533-535, 563
674 Author Index

Davies, R. O ., 380, 383, 431 Durelli, A. J., 259, 326


Dawson, D ., 264, 285, 286, 326 Dussan, E. B ., 124, 128, 135,
De Cindio, B ., 233, 323 136
Dee, H. B ., 2, 130 Dutta, A . , 335, 336, 339, 340,
Dee, W. J. N. B . , 247, 325 357, 372, 373
Defay, R ., 431 Duvdevani, I. J ., 2, 19, 23, 130
De La Vega, J., 243, 246, 268,
324
Demirli, K. B ., 462, 479
Den Otter, J. L ., 93, 96-99, 102,
103, 134, 135, 441, 475 Egelstaff, P. A ., 394, 432
Denn, M. M., 161, 171, 172, 223, Ehrmann, G ., 2, 130
455, 478, 528, 529, 559, 563, Eüers, N ., 358, 375
565 Einstein, A ., 389, 398, 431
Denton, D. L ., 482, 529, 554, Elbirli, B . , 364, 375
562, 563, 565 Ellerstein, S. M ., 340
Deterre, R ., 66, 75, 134 Emelyanouv, D. N ., 12, 79, 132
De Vargas, L ., 172, 223 Engle, R. S ., 396-409, 432
Dicken, F. L ., 261, 326 Ericksen, J. L. , 531, 563
Dierkes, A ., 264, 269, 278, 285, Erwin, L ., 559, 560, 565
288, 326 Evdokimova, M. G ., 10, 12, 13,
Dietz, W. , 125, 126, 136, 234, 132
235, 239, 241, 247, 248, 250- Ewell, R. H . , 394, 432
252, 256, 264, 265, 294, 323, Eyring, H ., 299, 328, 393, 394,
327, 439, 474 432
Dijksman, J. F ., 23, 125, 126,
133
Dülhoefer, J. R. , 464, 479
Dülinger, J. R ., 353, 375
Dillon, R. E . , 2, 131 Fara, R. A ., 356, 375, 410, 433,
Dinh, S. M ., 531, 559, 563 444, 475
Doan, P. H., 366, 375, 447, 476 Farris, R. J . , 345, 374
Dogadushkina, A. Y ., 10, 12, Fenner, R. T ., 438, 447, 474, 476
13, 132 Ferry, J. D ., 18, 81-85, 132, 142,
Donkai, N ., 86, 134 222, 301, 309, 328, 383, 394,
Doolittle, A. K ., 383, 393, 431 395, 431, 432
Dontje, A. J ., 344, 374 Ferstenberg, R ., 660, 669
Doyle, C. D ., 340, 373 Fields, T. R ., 139, 222, 254, 325
Doyle, M. J . , 243, 2p0, 324 Fikhman, V. D ., 160, 222, 455,
Draper, A. L ., 340, 373 478
Drexler, L. H., 140, 222 Finger, F. L. , 2, 129
Duda, J. L ., 179, 185-187, 223, Finlayson, B . A ., 186, 206, 224,
224 225, 455, 478
Dumoulin, M. M., 2, 98, 130 Fisher, B. C ., 345, 374, 492, 496,
Duncan, D ., 660, 669 535, 537, 562, 564
Author Index 675

Fleibner, M., 228, 247, 248, 254, Gerrard, J. E ., 108, 135


322 Gilbert, M. , 231, 232, 323
Flory, P. J., 85, 86, 134, 382, Gillham, J. K ., 356, 375, 527,
390, 393, 431, 432 562
Flynn, J. F ., 340, 372 Gillmore, G. D ., 1, 61, 62, 64, 69,
Fojolka, P. , 444, 475 129, 133, 247, 254, 264, 324
Foley, J. D ., 613, 668 Givler, R. C., 549, 552, 553, 564
Folgar, F ., 492, 493, 495, 544- Gladwell, B. K. , 554, 564
552, 554, 555, 562, 564 Glasscock, S. D ., 2, 105, 108,
Folkes, M. J., 167, 223, 559, 131
565 Gleissle, W., 2, 92, 96, 102, 103,
Folt, V. L ., 396, 432 107, 130
Ford, B . , 510, 524, 528, 529, Goedhart, D. J . , 79, 134
562 Gogos, C. G ., 1, 2, 23, 129, 130,
Forgacs, D. L ., 559, 565 254, 264, 325, 447, 476
Fourier, J. B . J . , 569 Gold, O ., 390, 398, 431
Fox, T. G ., 85, 86, 134, 356, Goldblatt, P. H., 18, 82-84, 132
375, 393, 394, 432 Goldfinger, G ., 299, 328
Frankel, N. A . , 358, 375 Gorbatchev, V. M., 340, 373
Freakley, P. K ., 421, 433 Gortemaker, F. H ., 92-95, 103,
Freeman, E. S.> 340, 373 134, 233, 323
Fricke, A . L ., 335, 343, 353, Gough, V. E . , 462, 479
374, 375, 458, 478 Graessley, W. W., 2, 79, 87, 90,
Fridshal, R ., 660, 669 91, 105, 108, 131, 134
Friedman, H. L ., 340, 373 Grant, G. E ., 100, 101, 135
Frocht, M. M., 259, 326 Gratch, S. , 393, 432
Fujiyama, M., 252, 325 Green, D. W ., 137, 221
Fukui, Y ., 243, 260, 324 Greener, J. , 235, 242, 245-248,
Fulcher, G. S ., 394, 432 256, 264, 265, 268, 298, 323,
324
Greenhow, E. J ., 340, 373
Greenwood, M. J., 264, 285, 286,
Galili, N. , 108, 135 326
Galskoy, A ., 23, 133 Griffith, R. M., 438, 447, 474,
Galvin, G. D . , 108, 135 530, 531, 534-536, 563
Gandhi, K. C ., 533, 535-537, Grimblat, V. N ., 447, 476
563 Gupta, R ., 182, 187, 224
Garden, R ., 269, 274, 277, 298, Gurnee, E. F ., 243, 260, 324
327, 527, 563 Gurney, W. A ., 462, 479
Garfield, L. J., 2, 92, 131 Gutfinger, C ., 44, 53, 133, 447,
Gaskins, F. N ., 159, 162, 222, 476, 638
455, 477 Guth, E. , 390, 398, 431
Geiger, K ., 455, 478 Guyles, K. C . , 335, 372
Gendron, R ., 2, 98, 99, 103, Guzman, G. M., 101, 105, 135
107, 130 Gynlai, G ., 340, 373
676 Author Index

H 108-111, 126, 127, 133, 135,


136, 182, 198, 224, 225, 235-
Hagan, R. S . , 2, 101, 131 240, 247, 248, 254, 256, 264,
Hagler, G. E . , 2, 92, 107, 129 265, 269, 289, 294, 323, 327,
Hagman, J. F ., 445, 476 439, 447, 448, 455, 474, 477,
Halmos, A. L ., 179, 223, 455, 488, 489, 562, 622, 634, 636,
478 666, 668, 669
Han, C. D . , 2, 92, 96, 98-100, HÜ1, C. T. , 533, 563
102, 103, 107, 129, 135, 138, Hm, H. E ., 381, 390-392, 415,
140, 142, 171, 221-223, 247, 431
254, 325, 335, 345, 348, 357, HiU, R. A . W. , 340, 373
359, 372, 374, 375, 395, 396, Hindle, C. S ., 259, 264, 265, 269,
432, 439-441, 455, 474, 475, 285, 286, 325, 326
477 Hiner, H. C ., 364, 375
Hand, G. L ., 531, 559, 563 Hiorns, B. M., 159, 222
Hanks, R. W., 448, 477 Hoare, L ., 247, 264, 324
Hansen, M. G ., 233, 323 Hocking, L. M., 124, 128, 136
Hariharan, T ., 254, 256, 309, Hoftijzer, P. J . , 79, 134
310, 311, 316, 318, 319, 455, Holland, K. M., 347, 375
478 Holmes, D. B ., 160, 222
Harris, J. B ., 542, 564 Holstein, H ., 181, 224
Harry, D. H ., 447, 476 Hong, S ., 185, 186, 187, 224
Hartiey, M. D . , 345, 374 Hou, T. H ., 349, 375
Hattori, M., 353, 375 Hronas, M. J ., 381, 390-392, 415,
Hassager, O ., 124, 128, 136, 431
191, 225 Hsich, H. S. Y ., 379, 380-392,
Hassler, J. C ., 335, 343, 353, 396-410, 412-417, 419, 420,
374, 375 431, 432, 445, 459, 476
Haworth, B ., 259, 265, 269, 325 Huang, C. F ., 1, 23, 129, 447, 476
Hausen, H., 573 Huang, C. R ., 2, 92, 96, 98, 100,
Hayahara, T ., 159, 222 129
Hele-Shaw, H. S ., 46-48, 133 Huang, D. C ., 2, 92, 131, 448, 477
Hellwege, K. H., 15, 18-20, 132, Huang, D. D ., 264, 266, 285, 287,
133 326
Hellmeyer, H. O ., 297, 308, 317, Huebner, K. H . , 50, 133
327, 455, 478 Huügol, R. R. , 191, 224
Hemsley, D. A ., 231, 322 Hull, D. , 247, 264, 324
Herman, E. A ., 504, 562 Hutchinson, J. M., 301, 328
Hermans, P. H., 558, 565 Hutton, J. F . , 108, 135
Heron, N . , 232, 323 Hyun, K. S . , 2, 131
Herrmann, H. D ., 2, 15, 17-19,
22, 130 I
Hettingr, A. C ., 335, 372
Hieber, C. A ., 36, 37, 41, 43-45, Ide, Y ., 61, 134, 439, 474
50, 51, 53, 56-59, 100, 101, Illman, J. C ., 444, 475
Author Index 677

Imai, T ., 444, 475 K


Indenbom, V. L ., 269, 274-276,
278, 280-282, 284, 327 Kadivar, M. K . , 439, 474
Isachenko, V. P ., 574 Kale, L. T ., 345, 374
Isayev, A. I ., 2, 37, 44, 45, 56- Kalnin, I. L ., 347, 375
59, 100, 101, 105, 131, 133, Kamal, M. R ., 1, 2, 23, 44, 61, 76,
140, 142-144, 147, 151, 163- 92, 108, 129-131, 134, 135, 239,
165, 167, 168, 174-177, 187, 247, 248, 254, 262, 265, 298,
188, 193, 198-219, 222, 225, 323-327, 330, 335, 344, 345,
229, 233, 235-242, 245-248, 347, 348, 353, 354, 356, 366,
250, 253-256, 262-265, 267- 370-375, 410-412, 433, 444,
269, 282, 289, 290, 294, 309- 447, 455, 458, 476, 478, 513,
311, 316, 318, 319, 323, 325, 562
327, 439, 447, 448, 453, 455, Kambe, H ., 353, 375
459, 474, 475, 477-479, 525, Kanari, K ., 353, 375
563, 636, 668 Kantz, M. R . , 231, 323
Ishida, Y . , 21, 83, 133 Karre, L ., 444, 475
Ito, K ., 2, 19, 20, 23, 129, 132, Kataoka, T ., 358, 375
258, 325 Katooka, T ., 358, 375
Ivanova, L. I ., 441, 443, 475 Katti, S. S ., 228, 322
Katz, D. , 282, 327
Kawahara, M., 187, 224
Kawai, H . , 243, 260, 324
Keedy, D. A ., 243, 260, 324
Jackson, G. B ., 261, 326 Kelley, F. N ., 381, 431
Jackson, N. A ., 206, 225 Kelsey, R. A ., 257, 325
Jackson, N. R ., 455, 478 Kenig, S ., 1, 23, 61, 129, 247,
Jackson, W. C ., 497, 549-552, 262, 264-266, 269, 286, 288,
554, 555, 564 297, 326, 327, 447, 476
Jacobsen, J., 108, 135 Kerr, D. L ., 344, 374
Jacovic, M. S ., 2, 130 Keskkula, H ., 261, 326
Jacques, M. St., 264, 288, 326 Kenyon, P. M., 242, 268, 298,
Janeschitz-Kriegl, H ., 2, 91-99, 324
102, 103, 107, 108, 125, 126, Keunings, R ., 193, 194, 196, 197,
131, 134-136, 139, 221, 222, 225, 439, 474, 559, 565
233-235, 239, 243, 246, 247, Khullar, P ., 654, 657, 669
252, 254, 264, 265, 323 Kim, K. U ., 2, 92, 96, 98, 100,
Jeffrey, G. B ., 540-542, 564 129
Jenkins, R ., 444, 475 Kim-E, M. E . , 184, 186, 224
Johnson, M. W., 193, 225 Kissinger, H. E ., 340, 373
Johnson, P. S ., 437, 438, 441, Kitagaua, K ., 90, 134
443, 445, 474, 475 Kitayama, Y ., 448, 477
Jones, B ., 108, 135 Klein, I ., 2, 19, 23, 130, 438, 447,
Jones, G. O . , 380, 383, 431 474
Jung, A . , 19, 132 Kline, D. E. , 353, 375
678 Author Index

Knappe, W., 2, 15, 17-19, 20, 22, Landel, R. F ., 18, 81-85, 132,
23, 130, 132, 133, 353, 375 301, 328, 383, 393, 398, 431,
Knutsson, B. A ., 2, 130 433
Koenig, J. L ., 406, 433, 444, La Nieve, H. L ., 160, 222, 455,
476 477
Koita, Y. T . , 254, 325 Laun, H. M., 2, 17, 19, 23, 92-95,
Kondo, A ., 92, 134, 153, 222, 107, 130, 131, 162, 172, 175,
455, 478 223, 233, 323, 455, 478
Kontos, E. G . , 437, 474 Laurence, R. L ., 107, 135
Kotaka, T . , 86, 134 Lawrence, A ., 504, 562
Kovacs, A. J . , 300, 301, 328 Learmouth, G. J . , 444, 475
Kozicki, W., 448, 477 Leblanc, J. L ., 443, 475
Kramer, J. M., 534, 563 Lee, C ., 345, 374
Kratochvil, J . , 340, 373 Lee, C . -C ., 492, 493, 495, 529,
Krealing, R. P ., 353, 375 552, 562, 563, 564
Kreith, F ., 280, 284, 307, 327 Lee, E. H., 268, 276-278, 280-282,
Krier, C. A . , 2, 131 327, 527, 563
Krishnakumar, J. M., 242, 267, Lee, D. I . , 398, 432
324 Lee, L. J . , 459, 479, 512, 515,
Kroesser, F. W., 438, 447, 474 516, 530, 531, 534-536, 562,
Kromer, F. J., 257, 325 563
Krueger, W. L ., 44, 133, 447, Lee, S. J . , 528, 529, 563
477, 638, 669 Lee, S. M. , 243, 246, 268, 324
Kubat, J ., 258, 259, 325, 326 Lee, T. S ., 440, 475
Kulicke, W. M., 395, 432 Lee, W. I. , 543, 564
Kunii, D . , 2, 129 Lehmann, P ., 20, 133
Kuo, Y ., 2, 36, 44, 61, 76, 131, Leider, P. J . , 533-535, 563, 564
133, 134, 366, 375, 447, 455, Lem, K .-W ., 339, 345, 348, 357,
476, 478 359, 372, 374
Kuske, A . , 259, 326 Lemmon, M. E ., 160, 223
Kuvshinski, E. V ., 159, 222 Leonov, A . I ., 172, 195, 198, 223,
Kwon, T. H., 41, 133, 161, 220, 225, 235, 239, 323
223, 650, 669 Levitsky, M., 527, 563
Kyu, T ., 243, 260, 324 Levy, M. F . , 444, 475
Lewis, A. F ., 356, 375
Li, J. C. M., 258, 259, 325
Lightfoot, E. N ., 126, 136
Lin, S. H., 160, 223
Lindley, P. B ., 421, 433
Ladveze, J ., 181, 224 Lindt, J. T . , 364, 375
Lafleur, P. G ., 239, 323, 324, Lipkina, E. H. , 195, 196, 225
447, 455, 476 Lipscomb, H. H ., 559, 565
La Mantia, F. P ., 2, 98, 130, Litovitz, T. A ., 394, 432
162, 171, 223, 455, 478 Litt, M. H . , 394, 432
Lamont, R. R ., 102, 107, 135 Lobe, V. M., 440, 475
Author Index 679

Lodge, A. S ., 139, 172, 191, McGinley, P. L . , 554, 564


221, 223, 225, 233, 323 McGowan, J. C ., 19-23, 132
Logvinenko, V. A ., 340, 373 McCuliough, R. L ., 555-559, 565
Longwell, P. A ., 160, 223 Mckiélvey, J. M., 454, 477
Lord, H. A ., 2, 15, 19, 21, 23, Meacham, S. E ., 264, 326
130, 133, 447, 476 Mead, R ., 2, 111, 131
Loshaek, S ., 356, 375, 393, 432 Meissner, J., 107, 135
Loundes, J ., 124, 128, 136 Mëncik, Z ., 231, 323
Lovrich, M. L . , 556, 565 Mendelson, M. A ., 189, 192, 193,
Lupton, J. M ., 2, 130 224, 455, 478
Lyn gaae-Jorgen sen, J . , 2, 130 Mendëisoh, R. A ., 2, 129, 131
Menges, d . , 20, 85, 133, 232, 258,
264, 269, 278, 285, 287, 288,
M
297, 308, 317, 323, 325-327,
455, 478
MacCallum, J. R ., 340, 373 Mentzer, C. C ., 2, 130
Macosko, C. W ., 2, 129, 345, Merz, E. H . , 87, 134, 395, 432
374, 375, 412, 433, 453, 479 Metzger, A. P . , 2, 131
Macsporran, W., 159, 222 Metzner, A. B ., 357, 375, 528,
Madding, R. P ., 353, 375 529, 563
Malkin, A. Y ., 8, 9, 12, 14, Michaeli, W., 20, 85, 133
132, 439, 474 Middleman, S ., 13, 124, 132, li5,
Malkus, D. S ., 439, 474 438, 477, 474, 476
Maloney, J. L ., 137, 221 Mieras, H. J. M. A ., 90, 134
Manche, E. P ., 340, 373 Miller, B . , 295, 327
Mandell, J. F ., 264, 266, 285, Miller, C . , 37, 133, 448, 477
287, 326 Müler, D. R . , 348, 375
Manley, R. St. J ., 542, 564 Müls, N. J. , 79, 90, 134, 242,
Manzione, L. T ., 459, 479 265, 268, 269, 324
Marin, G ., 90, 94, 95, 134 Mütz, J. , 2, 87, 130
Mark, H., 299, 328 Minoshima, W., 2, 79, 131
Marker, L. F ., 510, 524, 528- Mitsuishi, N ., 448, 477
531, 534-536, 562, 563 Mittal, R. K ., 260, 326
Markovitz, H . , 395, 432 Monge, Ph., 94, 95, 135
Marrucci, G ., 235, 323, 559, 565 Montes, S ., 439, 440, 475
Marshall, D. E ,, 232, 323 Montfort, J. P ., 94, 95, 135
Maschmayer, R. D ., 533, 563 Mooney, M., 358, 375, 441, 475
Mason, S. G ., 542, 543, 559, Morjaria, M. A ., 181, 224
564, 565 Morrell, S. H. , 403, 406, 433
Masuda, T ., 90, 134, 135 Morris, H. L . , 2, 129
Mathar, J ., 257, 325 Moser, B. G ., 398, 433
Matsuoka, S ., 20, 21, 83, 133 Moy, F. , 262, 298, 326, 327
Maxwell, B ., 19, 20, 132, 133 Muckstadt, J . , 660, 669
McGee, S. H. - see Munson- Muki, R ., 276, 277, 327, 527,
McGee 563
680 Author Index

Munson-McGee, S. H ., 554-557, Oldroyd, J. G ., 189, 191, 224


559, 565 Ohlberg, S. M., 277, 327
Munstedt, H., 2, 97-101, 105, Oliver, D. R ., 159, 222
130, 131 Onogi, S ., 90, 134, 243, 260, 324
Mussatti, F. G ., 345, 347, 374, Ophermann, J., 258, 285, 287, 325
412, 433 O'Reüly, J. M., 18, 21, 86, 90,
Mustafayev, D. A ., 233, 323 132, 134
Myasnikova, L. I ., 14, 79, 132 O'Rourke, V. M ., 90, 134
Osinski, J. S ., 345, 374
Osswald, T. A ., 496
Otto-Laupenmuhlen, E ., 353, 375
Ozawa, T . , 353, 375
Naberezhnov, A. M., 37, 133,
448, 477
Nadkarni, V. M., 2, 85, 130
Nakajima, N ., 2, 19, 23, 120,
121, 131, 132 Paddon, D. J. , 179, 181, 223, 224
Narayanaswamy, O. S ., 277, 327, Palit, K ., 437, 446, 473
527, 563 Panova, G. D ., 14, 79, 132
Narkis, M., 159, 222, 262, 263, Paris, W. W ., 462, 479
266, 326, 438, 447, 474 Parrot, R. G ., 447, 476
Nelder, J. A . , 2, 111, 131 Pashkin, E. D ., 195, 198, 225
Nens, S . , 353, 375 Patankar, S. V ., 634, 635, 668
Nevin, A ., 79, 90, 134 Patten, T. W., 186, 224, 455, 478
Newman, N. D ., 231, 233 Patterson, G. D ., 308, 309, 328
Ngan, C. G ., 124, 128, 136 Patterson, L. T ., 243, 260, 324
Nicholas, M. O ., 125, 126, 136 Paul, F. , 15, 18-20, 132
Nickel, R. E . , 179, 223 Paulson, D. C ., 61, 134
Nied, H. A. , 459, 479 Pavlova, I. P ., 243, 260, 324
Nielsen, L. E ., 353, 375 Payne, A. R. , 421, 433
Nikolayeva, N. E ., 439, 475 Pearson, G. H ., 2, 92, 131
Nishimura, T ., 2, 98, 100, 101, Pearson, G ., 235, 245-248, 256,
105, 130, 358, 375 264, 265, 323
Nomura, S ., 243, 260, 324 Pearson, H. J., 23, 36, 133, 447,
Norman, R. H . , 443, 475 448, 477
Novikova, O. S ., 340, 373 Pearson, J. R. A ., 1, 23, 36, 108,
Nyun, H . , 108, 135 129, 133, 135, 393, 432, 438,
447, 448, 474, 476, 477
Pedersen, S ., 232, 323
O
Peeples, F. N ., 139, 222
Pennington, D ., 559, 565
Ockendon, H., 108, 135 Penwell, R. C ., 2, 18-20, 22, 84,
Oda, K ., 247, 325, 443, 475 122-124, 130, 132, 135
O'Driscoll, K. F ., 345, 374 Perera, M. G. N ., 191, 224
Olcese, T ., 340, 373 Perry, R. H . , 137, 221
Author Index 681

Perry, S. J ., 345, 374 R


Petermann, J., 259, 326
Peukert, H ., 243, 324 Rabinowitsch, B ., 4, 132
Peyret, R ., 181, 224 Racin, R ., 2, 92, 131
Phan Thien, N ., 193, 225, 535, Ram, A ., 2, 87, 130, 159, 222,
563 438, 447, 474
Philipp , B . , 159, 222 Rashni, V ., 260, 326
Phüippoff, W., 139, 159, 162, Ramos, A. R ., 301, 328
221, 222, 233, 323, 394, 432, Ramos, de Valle, L. F ., 439, 474
455, 477 Ramsteiner, F ., 2, 19, 20, 37, 96,
Pike, R. D ., 172, 223 98, 99, 101, 105, 107, 129, 132,
Püoyan, G. O . , 340, 373 448, 477
Pintens, E. A ., 443, 475 Rauwendaal, C ., 172, 223
Pipes, R. B ., 549, 556-558, 561, Read, B. E ., 243, 260, 324
564, 565 Read, M. D . , 172, 223
Pisatcioglu, S. Y ., 335, 343, 353, Read, W. T ., 243, 260, 262, 263,
374, 375, 458, 478 265, 286, 289, 324
Pittman, J. F. T ., 542, 559, 564, R eber, D. H . , 23, 133
565 Reddy, K. R . , 439, 474
Plazek, D. J., 90, 134 Redfern, J. P ., 340, 373
Pliskin, I. , 399, 433 Rendler, N. J . , 257, 325
Plotnikova, E. P ., 439, 474 Requicha, A. A. G ., 616, 668
Podolsky, Y. Y ., 140, 142-144, Retting, W., 243, 260, 324
151, 222, 233, 254, 323, 325 Richardson, S. M. , 23, 36, 125,
Pollock, D ., 2, 130 126, 133, 136, 447, 448, 476,
Porter, R. S ., 2, 18-22, 82-84, 477
86, 90, 122-124, 130, 132, Rigbi, Z . , 108, 135
134, 135, 167, 223, 395, 432 Rigdahl, M ., 258, 259, 278, 325-
Prasad, A ., 598 327
Prentice, G. A ., 459, 479 Riley, W. F . , 259, 326
Prest, W. M., 86, 90, 134 Roberts, G. E. , 383, 431
Prigogine, I ., 431 Robertson, G. , 259, 326
Prime, R. B ., 333, 340, 372, Robins, G ., 2, 130
373, 410, 433, 444, 476 Robinson, D ., 13, 132
Priss, L. S. , 243, 260, 324 Rogers, M. G ., 444, 476
Pritchard, G ., 444, 475 Rogers, R. N. , 340, 372
Progelhof, R. C ., 458, 478 Rogers, T. G ., 268, 276-278, 280-
Prokunin, A. N ., 195, 198, 225 282, 327, 527, 563
Prozorovskaya, N. V ., 2, 8, 9, Roller, M. B. , 412, 433
12, 100, 131, 132 Roman, J. F ., 2, 92, 130
Rose, W ., 124, 135
Rosen, S. L ., 180, 223, 455, 477
Rosenzweig, N ., 262, 263, 266,
Qayyum, M. M., 243, 324 326
Quigley, J. M., 559, 561, 565 Rubin, I. I . , 435, 459, 473
682 Author Index

Rudd, J. F . , 2, 131 Schmidt, L. R ., 125, 136, 247,


Rudd, T. F . , 243, 324 325
Rusch, K. C ., 299, 301, 302, Schmidt, G. P ., 348, 375
305, 328 Schott, H ., 2, 98, 99, 129
Russell, D. P ., 264, 286, 287, Schowalter, W. R. , 160, 223
326, 327 Schreiber, H. P ., 87, 134
Ryabchikov, I. D ., 340, 373 Schultz, J. M ., 228, 322
Ryabov, A. V ., 14, 79, 132 Schurmann, E ., 598
Ryan, M. E ., 61, 134, 335-340, Schurz, J., 159, 222
345-349, 351-353, 357, 366, Schweizer, R ., 565
370, 372-375, 410-412, 433, Scott, A. J . , 441, 446, 475
444, 455, 458, 476, 478 Scott, J. R. , 533, 563
Segalman, D ., 193, 225
Seiden, R . , 259, 326
Semjonow, V ., 9, 12, 15, 17-20,
Sacher, E ., 410, 433, 444, 476 132
Sabsai, O. Y ., 439, 474 Serino, J. F ., 444, 475
SaffeU, J. R . , 242, 244, 245, Sestak, J . , 340, 373
246, 267, 268, 324 Sezna, J. A . , 443, 475
Saibel, E ., 268, 275, 277, 278, Shabab, Y ., 444, 475
287, 327, 527, 563 Shaffer, B . W ., 527, 563
Saini, D. R ., 85, 134 Shen, S. F . , 41, 43-45, 50, 51,
Sakai, T ., 358, 375 53, 56-59, 126, 127, 133, 136,
Sakovtseva, M. B . , 2, 10, 12, 161, 181, 198, 222-225, 235,
13, 130, 132 323, 447, 455, 477, 478, 488,
Salinas, A ., 559, 565 489, 562, 634, 650, 668, 669
Salovey, R ., 462, 479 Shenoy, A . V ., 2, 85, 130, 134
Samuels, R. J ., 230, 231, 323 Shibayama, K ., 356, 375
Sandilands, G. J., 259, 265, Shih, P. T. K. , 444, 476
269, 286, 325 Shishkin, N ., 22, 133
Santamaria, A ., 101, 105, 135 Shusman, L. , 1, 129
Santi, G. C . , 235, 323 Sidi, S ., 343-345, 371, 373, 374
Sastry, V. M. K ., 448, 477 Sieder, E. N . , 572
Satava, V ., 340, 373 Siegmann, A . , 262-266, 269, 286,
Sato, T ., 243, 260, 324 288, 297, 326, 327
Sauer, J. A ., 353, 375 Silva-Nieto, R. J., 345, 374, 492,
Saunders, D. W., 441, 446, 475 496, 535, 537, 562, 564
Saunders, T. F ., 444, 475 Simmons, E. L ., 340, 373
Schauer, K. R ., 598 Simon, R. H. M ., 2, 13, 131
Schechter, R. S ., 448, 477 Simpson, G. M., 261, 326
Schlanger, H. P ., 459, 460, 479 Singh, K. J . , 577-580, 598
Schlichting, H., 46, 47, 133, Siskovic, N ., 2, 92, 96, 98, 100,
488, 562 129
Schmidt, L ., 258, 264, 269, 278, Skelton, J. R ., 406, 433
285, 287, 288, 325, 326 Smith, D. H. , 438, 474
Author Index 683

Smith, G. R ., 554, 564 Sutherland, I. E ., 610, 668


Smith, J. B ., 448, 477 Sylvester, N. D ., 160, 223, 455,
Smith, K. L ., 264, 266, 285, 287, 477
326
Smith, L. C . , 340, 372
Smith, R. W ., 396, 432
Smith, T . , 367, 375
Sneller, J. A ., 360, 362, 375 Tadmor, Z ., 2, 44, 53, 124, 126,
Snowdown, J. C ., 380, 421, 431 130, 133, 136, 254, 264, 325,
So, P ., 263, 265, 267, 269, 326 438, 447, 474, 476, 477, 525,
Socha, L. S ., 44, 45, 56-59, 133, 563, 638, 668, 669
447, 477 Taggart, D. G ., 556-558, 565
Sourour, S ., 333-335, 341-343, Tajima, Y. A ., 345, 349, 374, 375
345, 347, 353-356, 372, 375, Takao, J ., 159, 222
410, 411, 412, 433, 444, 476, Takeuchi, N ., 187, 224
513, 562 Takserman-Krozer, R ., 108, 135
Southwart, D. W., 462, 479 Tan, V ., 239, 247, 248, 254, 265,
Sparrow, E. M ., 160, 223 324, 327, 344, 371, 374
Spelta, O ., 340, 373 Tanaka, H ., 358, 375,440,475
Spencer, R. S ., 1, 2, 61, 62, Tanner, J . , 340, 373
64, 69, 129, 131, 133, 247, Tanner, R. I ., 179, 182, 187,
254, 264, 324 189, 193, 223-225, 439, 474,
Springer, G. S ., 543, 564 534, 535, 563
SpruieU, J. E ., 2, 79, 130, 131, Tant, M. R ., 299, 300, 328
231, 323 Tayler, A. B ., 66, 125, 126, 134,
Steidler, F. E ., 108, 135 136
Stein, R. S . , 243, 260, 324 Taylor, L. J. , 340, 373
Sternberg, E ., 276, 277, 327, Tee, T. T . , 264, 326
527, 563 Te Nijenhuis, K . , 92-95, 103, 134
Stevens, J. R ., 308, 309, 328 Thakar, A. , 335, 372
Stevenson, J. F ., 23, 36, 133, Thakkar, B. S ., 262, 264, 326
513-519, 562 Theocaris, P. S ., 243, 260, 324
Stewart, W. E. , 126, 136 Thienel, P ., 20, 85, 133
Stigale, F. H ., 231, 323 Thomas, D. G ., 358, 375
Stratton, R. A ., 2, 13, 18, Thomas, D. P ., 2, 101, 131
131, 132 Thomas, K ., 264, 285, 286, 326
Struik, L. C. E ., 264, 266, 277, Thompson, B. W ., 48, 133
278, 280, 298, 299, 312, 326, Thorton, E. A ., 50, 133
328, 525, 563 Throne, J. L . , 458, 478
Suehiro, S ., 243, 260, 324 Titomanlio, T ., 235, 323
Suetsugu, Y ., 440, 442, 475 Tiu, C. , 448, 477
Suganuma, A ., 2, 129 Tobolsky, A. V ., 90, 134, 280,
Sukanek, P. C ., 108, 135 281, 308, 327, 328
Sulzhenko, L. L ., 159, 222 Toki, S ., 439, 441, 474
Sundstrom, D. W., 345, 374 Tokita, N ., 392, 399, 432, 433
684 Author Index

Tolstukhina, F. S ., 439, 474 Van der Vijgh, R ., 247, 248, 265,


Tomlinson, J. N ., 353, 375 325
Tool, A. Q . , 277, 327 Van Donselaar, R ., 108, 135
Toor, H. L ., 1, 108, 129, 135, Van Krevelen, D. W., 18, 21, 79,
247, 264, 324, 447, 476 82, 84, 132, 134
Tordella, J. P ., 101, 105, 107, Van Leeuwen, T r. J ., 247, 248,
135 265, 325
Toyoda, H., 448, 477 Van Rijn, C. F. H . , 90, 134
Tsai, B. C . , 462, 479 Van Wijngaarden, H ., 23, 125,
Treloar, L. R. G ., 233, 323, 126, 133
381, 431 Vargin, S ., 340, 373
Treutting, R. G ., 243, 260, 262, Varner, A. M., 447, 448, 477
263, 265, 286, 289, 324 Vigness, I. , 257, 325
Tseng, H. C . , 100, 101, 135 Vila, G. R ., 393, 432
Tucker, C. L ., 492, 493, 495, Villamizar, C. A ., 99, 103, 135,
496, 529, 544-552, 554, 555, 247, 325
558, 559, 562, 563, 564, 565 Vinogradov, G. V ., 2, 8, 9, 12,
Tung, L. H ., 87, 134 14, 100, 131, 132, 140, 142-144,
Turnbull, D ., 299, 328 151, 160, 222, 223, 233, 254,
Turnbull, N ., 392, 399, 432, 433 323, 325, 439, 441, 443, 455,
474, 475, 478
Viryayuthakorn, M., 187, 192, 194,
224, 439, 455, 474, 478
Vishnuakov, I. I . , 243, 260, 324
Ueberreiter, K ., 353, 375 Voelcker, H. B . , 616, 668
Ueda, K ., 608, 668 Volkov, V. Z., 160, 223, 455,
Ueda, S ., 1, 107, 129 478
Upadhyay, R. K ., 37, 133, 144, Von Turkovich, R ., 559, 560,
147, 181, 187, 193, 198-219, 565
222, 224, 225, 235, 254, 323, Vrentas, J. S ., 179, 185-187,
448, 455, 477, 478, 636, 668 223, 224
Ushirokawa, M ., 243, 260, 324 Vyvoda, J. C ., 231, 232, 323
Utracki, L. A ., 2, 92, 98, 99,
103, 107, 130, 131
W

Wagner, M. H ;, 2, 92-95, 107, 130


Wales, J. L. S ., 2, 90, 91, 93, 96-
Vachagin, K. D ., 37, 133, 448, 99, 102, 103, 131, 134, 135, 233,
477 237, 239, 243, 247, 248, 252,
Valenza, A., 2, 98, 130, 162, 254, 265, 294, 323, 324
171, 223, 455, 478 WaU, L. A ., 340, 372
Van Dam, A . , 613, 668 Walters, K ., 179, 187, 191, 223,
Van Dam, J . , 108, 135 224, 225, 395, 396, 432, 533-
Van der Vegt, A. K . , 2, 84, 131 535, 563
Author Index 685

Wang, K. K ., 23, 41, 44, 45, Wüliams, S. G ., 23, 133, 447,


56-59, 100, 101, 133, 135, 476
161, 223, 447, 477, 598, 609, Williams, H. L. , 345, 374
622, 625, 634, 649, 650, 654, Williams, J. G ., 438, 447
657, 660, 661, 668, 669 Williams, M. C ., 459, 479
Wang, V. W., 622, 634, 666, Wüliams, M. L ., 18, 81-85, 132,
668, 669 301, 328, 381, 383, 393, 431
Wang, W. P ., 654, 657, 660, Williams, R. W ., 179, 223
661, 669 Windle, A. H ., 242, 244-246, 267,
Wang, Y. L ., 160, 223 268, 324
Watson, S. W., 340, 373 Wineman, A. S ., 439, 474
Webster, M. F ., 191, 224, 225 Winer, W. O ., 108, 135
Weissberg, H. L ., 160, 222 Winkel, E ., 264, 269, 278, 285,
Weitsman, Y ., 282, 327 288, 326
Wendlandt, W. W., 340, 373 Winslow, A . M., 635, 636, 668
Wesseling, P ., 23, 125, 126, Winter, H. H . , 108, 135
133 Wise, R. W., 403, 406, 433
Westover, R. F ., 2, 19, 20, 23, Wissler, E. H., 448, 477
117, 120, 129, 132 Woo, T. C ., 268, 276-278, 280-
Wheelans, M. A ., 423, 433, 435- 282, 327, 527, 563
437, 444, 464, 473 Woods, W. C . , 464, 479
Wheeler, J. A . , 448, 477 Wortberg, J., 20, 85, 133
White, E. F. T ., 383, 431 Wu, P. C. , 1, 23, 129, 447, 476
White, J. L ., 2, 79, 92, 107, 125, Wubken, G ., 20, 85, 133, 232,
126, 129, 130, 131, 134, 136, 264, 323, 328
153, 167, 222, 223, 228, 231, Wulf, K ,, 159, 222
234, 235, 239, 241, 247, 248, Wust, C. 242, 246, 268, 298,
250-252, 256, 264, 265, 285, 302, 324
286, 294, 322, 323, 325, 327,
358, 366, 375, 392, 432, 439,
443, 447, 448, 455, 474-478,
559, 565
White, J. R ., 243, 245, 259, Yamada, M ., 2, 19, 21, 83, 130,
264, 265, 269, 286, 324, 167, 223
325-327 Yamamoto, O ., 353, 375
White, R. P ., 345, 347, 374 Yang, P. H . , 462, 479
Whitehead, J. C. 139, 140, 141 Yannas, I. V ., 243, 260, 282,
222 324, 327
Whorlow, R. W., 441, 475 Yanyo, L. C ., 379-383, 385-388,
Wüey, J. P . , 348, 375 412, 413, 415, 416, 431
Wilks, C. E ., 2, 131 Yap, C. Y ., 264, 326
Willard, P. E ., 344, 374 Yeh, P . W. , 189, 190, 192, 193,
Wükes, G. L ., 231, 299, 300, 224, 455, 478
328 Yoo, H. J . , 455, 478
Wilkes, C. E ., 396, 432 Yoo, H. Y ., 140, 142, 222
685 Author Index

Ziegel, K. D ., 423, 433


Zienkiewicz, O. C ., 50, 133, 634,
668
Zakharenko, N. V ., 439, 474
Zosel, A. , 22, 133
Zamodits, H ., 393, 432, 438, Zucker, W., 660, 669
447, 474 Zurn, R. M ., 380, 381, 389, 412-
Zentner, M. M., 535-538, 564 417, 419, 420, 431
Subject Index

Air quenching 246


Air-shot temperature 50
Abrupt contraction 138, 207, 218, Amino resins 362
220 Analytical solutions 181
Abrupt expansion 138, 207, 209, Analyzer 229
218, 220 Anisotropic mechanical behavior
Accelerator 393 264
Acetal 23 Anisotropic thermal conductivity
Acrylonitrile - butadiene - styrene of SMC 507
(A B S) 2, 6-10, 13, 14, 19, 23, Anisotropic thermal expansion 525
644, 648 Anisotropic viscosity 530, 531
Activation energy 336 Anisotropy of physical properties
Activation energy of cure reaction 461, 463
412 Annealed moldings 287
Activation energy of viscosity Annealing 246, 265
82-85 Annular cross-section 574
Activation energy of viscous flow Antivibration mounting, 421, 424
415 Apparent shear rate 3-5, 96, 108-
Actual volume 299 117, 119-124
Addition polymerization 506, 511 Apparent viscosity 3-5, 111
Advancing melt front 234, 235, APT 659, 660, 662
247 Arrhenius equation 393
Aggarwala-Saibel theory 275, 278, Arrhenius temperature depen­
287 dence 2, 8, 25, 81
Aging 298 Artificial intelligence 667
Aging in liquid nitrogen 286 ASCII (code) 615

687
688 Subject Index

Asymmetric Cooling 288 [Birefringence in moldings]


Automatically programmed tools effect of packing pressure 254,
(A PT ) 659, 660, 662 255, 257
Auxiliary modulus function 277, effect of processing conditions
280 248
Average velocities - in compres­ maximum 247-249, 254-256
sion molding 487, 489 minimum 247, 251
Axisymmetric flow 138, 176, 178 residual 265
Axisymmetric geometry 162 secondary maximum 256, 257
Blends 70-80, 86
BMWD (broad mol. wt. distrib.)
11-14, 83, 89, 90, 92, 94, 95,
105-108, 110, 116, 118
Boolean operation 619, 659, 660,
Baffles 585 664
Bagley correction 162, 164 Bosses 137, 220
Bagley plot 19, 96, 122, 123 Boundary element method 649-657,
Balanced pressure 424 597
Balanced runner system 429 Boundary layer 160
Bandwidth 53 Boundary representation 618,
Bartenev theory 273 619, 653, 654, 660, 662
Bifurcation phenomenon 193 Brabender torque rheometer 344
Biot number 266, 575 Bulk molding compounds 363
Birefringence 96, 97, 139, 144, Bulk velocity 3, 4, 25, 26, 29,
228, 233 40, 57, 75, 76, 122, 125, 127,
along centerline 148, 207, 209 128
in converging channel 145
buildup 149
distribution 206
effect of annealing 245
effect of bath temperature 246
flow-induced 242, 246 CAD 609, 612, 616, 620, 643,
frozen-in 242, 246, 254 659
gap wise distribution 229, 240 CAD/CAE/CAM 620, 621, 662-668
gap wise variation 145 CAD/CAM 609-611, 621, 651, 653,
in polycarbonate 244 657, 659, 660, 664, 666
in polymethylmethacrylate 244 CAE 1, 663
maximum 151, 212, 216 Calorimetry 444
prediction 206 CAM 609, 659
relaxation 245, 250 Capillary rheometer 3, 4, 13, 36,
thermal 242, 245, 246 71, 96, 107, 111, 121, 124,
Birefringence in moldings 345
effect of cavity shape 249 pressurized 18, 19, 117-120
effect of cavity thickness 251 Sieglaff McKelvey 3
effect of flow rate 249, 250 Capillary rheometry 152, 164, 395,
effect of melt temperature 249 441, 445
Subject Index 689

Carbon black-filled Compensator 229


isoprene rubber 440 Compliance (steady shear) 8, 79,
polystyrene 441, 442 86-93, 95
Carreau model 104-110, 183 Composite mechanics 389
Carreau number 184 Compressibility (variable density)
Cathode ray tube (C R T) 611, 612, 62-78
617 Compression mold filling - gener­
Cavity 447 alized Hele-Shaw model 488-498
center-gated disk 28, 29, 36, lubricated squeezing flow model
39, 41 497, 504
circular tube 28, 29, 36, 39, 40, order of magnitude analysis 486,
41, 111, 116, 122 488
end-gated strip 23, 29, 31, 36 thick parts 528-529
39, 41, 50, 57, 64-78, 125, 126 thin cavities 485-504
filling 1, 22-61, 67, 70, 72, 77 Compression molding 421
multi-gated 41-44, 622, 625, 627, Compressive stress 266
630, 631 Computer aided design 597
three-dimensional (thin) 610, Computer graphics 609-620, 628-
634, 637, 642, 643, 667 634, 641-668
variable thickness 56-61, 68, Computer simulation 363
635, 639, 642, 643 Conduction 569
with insert 48, 54-58, 60, 61 Conduction shape factors 580,
Cavity-filling simulation 218, 470 582, 590
Centerline velocity 192, 205 Cone-and-plate 96, 97
overshoot 192 Cone and plate geometry 345
Characteristic relaxation time 277, Conservation of energy 24, 28,
280 37, 44, 45, 51, 67, 113, 114,
Characteristic shear rate 206 639, 650, 651, 653
Characteristic time 183, 198 Conservation of mass 23, 24, 45,
Chemical analysis 331 53, 62-78, 635-639
Chemorheology 374, 421 Conservation of momentum 24, 44,
Circular arc runner 453 62, 67, 76
Circular die 161 Consistency index 350
Circular polariscope 259 Constitutive equation 178
Cicular tube 28, 29, 36, 39-41, Constitutive equation - curing
111, 116, 122 material 527
Clamp force 447, 608, 642, 648, 667 Constrained quenching 245, 246
Coefficient of thermal expansion Constructive solid geometry (CSG)
273, 276, 278, 307 618, 660
Cold-feed extruder 438 Contact micro-radiography 554
Cold thermal layers 25, 30, 32, 49, Continuity equation for fiber or­
50, 125, 126 ientation 545
Color-shaded graphics 610, 613, Contraction
620, 641-643, 648, 649, 654, angle 151, 216
655, 663 flow 144, 151, 218
Compensational method 229 ratio 151, 176, 212, 216
690 Subject Index

Contraction in mold 254, 265 [Curing - and heat transfer]


Control volume 53, 635-640 unsaturated polyesters 506,
Convection 570 511, 512
Convective heat transfer 278 Curing material - mechanical prop­
Converging funnel 147 erties 527
Coolant temperature 26 Curing kinetics 444, 457, 458
Coolants 586 Curing reaction 504
Cooling of plastics 575 Curing stresses 459, 461, 472
Cooling lines 581 Curing temperature 457
Cooling stage 22, 66, 67, 70, 78 Curing time vs. mold temperature
Cooling system 581, 607, 608, 620, 522
649-659, 664, 665 Cycle time 621, 649
Correction due to normal stress
162
Correction factor 157
Couette coefficient 160
Couette correction 152
Coupled-flow-path method 36-44, Damkohler number 517, 518,
57 519
Cox-Merz relation 87 Data bank (viscosity) 2, 3, 6,
CPU 612, 615 648
Crazing 286 Data base 610, 665-667
Crazing by bending 287 Deborah number 198-202
Crazing in solvent 287 in axisymmetric flow 188
Creep 262, 268, 282 in planar flow 188
Creep (shear) 90 Defects in plastic parts 593
Creeping flow 160 Degradation 621, 622, 628
Critical molecular weight 394 Degree of cure 412, 423, 458,
Critical shear stress for fiber 462, 463, 512
damage 559 Degree of orientation 399
Critical stress 96, 107 Degree of scorch 423
Cross model 2, 79, 104-110 Delivery system 607, 608, 621,
Crosslinks 381 635, 641
Crosslink density 356, 409 Density (mass) 24, 25, 28, 31,
Crosslinking agents 403 43-45, 49, 50, 62-78, 111,
Crosslinked sample 286 114, 115, 118, 125, 242, 296,
CRT 611, 612, 617, 660 297, 302, 642, 646
Crystallinity (semi) 14, 15, 19, Density in moldings 312
66, 83, 228 effect of processing 313
CSC 618, 660 Density in
Cure, degree of 512 inorganic glass 298
Cure cycle 412 polyethylene 298
Cure rate 457 polymethylmethacrylate 303,
Curing - and heat transfer 516 307, 310
dimensional analysis 516-517 polyphenylene oxide 297
profiles 518, 519, 520 polystyrene 298, 302, 310
Subject Index 691

Density in quenched samples Effect of fillers 357


effect of bath temperature 302 Effective extra length 157
effect of initial temperature 303 Effective pressure drop 157
effect of sample thickness 302 Effective temperature 302
Density modelling* 296, 309 Ejector pins (plate) 608, 657-659
Density relaxation 304, 305 Elastic compliance 142
Design of cooling system 587 Elastic effect 139
Desulfuration 407 Elastic modulus 384
Diallyl ortho-phthalate 362 Elastic recoil 159
Diallyl isophthalate 362 Elasticity 62, 67, 78
Dichroic ratio 231 Elastomer-filled composite 400
Die entry region 153 Elastomer-metal composite 421
Die pak 598 Electrical conductivity 444
Die swell 396 Ellipsoidal axis ratio 541, 542
Dielectric heating 484 Ellipsoidal particle - orientation
Dielectric method 444 540-542
Differential scanning calorimetry Elongational flow 239, 254
332, 410, 511 Elongational viscosity 161
Differential thermal analysis 332 Energy equation 24, 28, 37, 44,
Diffraction apparatus for mea­ 45, 51, 67, 113, 114
suring fiber orientation 555 Energy of vaporization 394
Diffusivity 299, 307 Engineering material design 380
Dilatometry 332 Entrance length 184, 186
Dimensional nonuniformity 137 Entrance pressure loss or drop
Direct-view storage tube (DVST) 161, 162
612 Entry flow effects 152, 161
Disk cavity 28, 29, 36, 41 Epoxy resins 333, 362
Displacement method 186 Equation of
Display processing unit (DPU) confirmity 176
612, 615 momentum 178
Divergence theorem 51 Equation of state 62, 69, 297
Diverging flow 209, 218 Equipment 359
Doolittle free-volume theory 383 Equilibrium thickness of SMC 535
Dynamic measurements 348 Equilibrium volume 306
Dynamic mechanical properties 332 Equivalent circular runner 448
Dynamic modulus 94-97, 383 Estimating kinetic parameters 340
Dynamic modulus of SMC 536, 538 Ethylene-propylene-diene ter-
Dynamic viscosity 87, 94, 96, 97, polymer 463
348, 396 Execution time (computer) 53
Exit flow effect 152, 162
Exit pressure 96, 162, 171, 175
Expansion angle 216
Expansion flow 144, 151
EBCDIC (code) 615 Expansion in mold 265
Eccentric rotating disks 345 Experimental techniques 228-233
EDM 609, 610 Extensional stresses 264, 265
692 Subject Index

Extent of reaction 339 Finger measure 197


Extinction angle 139, 233 Finger strain 191
Extra equivalent length 99, 100, Finite-difference mesh 28, 31, 51,
106 73, 74, 116, 117, 639
Extra length 165, 168 Finite-difference method (FDM)
Extra pressure loss 96-105, 160, 1, 26, 28, 30, 37, 41-44, 50,
161, 165, 168, 220, 221, 453, 51, 53, 70, 71, 108, 111, 113-
454, 463 115, 121, 123, 179, 198, 279,
Extrudate swell 443 307, 634
Extrusion of rubber compound Finite-element mesh 641, 643, 644,
467 647
Extrusion pressure 396 Finite-element method (FEM) 41-
Eyring’s theory 393 44, 50-53, 56, 57, 71, 127,
128, 278, 597, 634-649, 663,
666
First normal stress difference
140, 159, 203-205, 233, 235,
Falling sphere 345 238-240, 264, 265, 348
Family mold 625 First-order chemical reaction
Fatigue crack in moldings 287 412, 415
Fiber compaction 535 First-order rate equation 300,
Fiber damage 559, 560 305, 307
Fiber orientation 539-559 Fitting 443
continuity equation 545 Flow activation energy 348
Fol gar-Tucker model 545-547, Flow front: see melt front
549 Flow-induced anisotropy 462
interaction coefficient 545, 546 Flow path 450, 625, 627-630,
Jeffrey’s equation 540-542, 552, 632-634
553 Flow patterns 153
measurement 554-556 Flow rate 151
normalization condition 545 Flow rate of coolant 587
particle tracing method 549-552 Flow residual stresses 265, 294
planar extensional flow 541 Flow rheology 392
probability distribution function Flow stress development 221
544-546 Flow stresses 264, 285, 294
simple shear flow 541-544 Flow visualization 247
tensor description 558, 559 Force balance 24, 44, 45, 62, 67,
two-body interactions 543 76, 122
Fictitious temperature 273 Folgar-Tucker model for fiber
Fill time 240, 247, 621, 623, 628- orientation 545-547, 549
630, 632, 633, 648 Forced convection 571
Filled polymers 168 Fountain effect 254, 264
Filler dispersion 396 Fountain flow 490, 509, 510
Filler loading 384 Fountain-flow region 26, 50, 124-
Filling stage 1, 23-61, 67, 70, 72, 128
77, 364 Fourier spectrum 421
Subject Index 693

Fourier transform infrared 228, Glass transition temperature


230 18-22, 77, 78, 82-85, 232,
Fraunhoffer diffraction for mea­ 265, 299, 302, 305, 350, 354,
suring fiber orientation 555 380
Free energy 393 in curing systems 512
Free-radical copolymerization 511 Glassy state 380
Free-radical initiator 511 Graetz number 111
Free quenching 245
Free volume 299, 301, 302, 306
Free volume model 393, 394 H
Frequency 87, 95
Friction factor 590 Hagenbach coefficient 160
Fringe method 230 Hardness 463
Frozen-in orientation 228, 249, Heat balance 650, 651
459 Heat flow through molds 580
Frozen surface layer 251 Heat flux 650, 651, 654, 655
Heat of reaction 341
Heat pipes 581
Heat shrinkage 228
Heat shrinkage technique 232
Galerkin procedure 51, 636 Heat transfer - and curing 504,
Gapwise average velocities 487, 489 516
Gap wise velocity 447 dimensional analysis, with
Gate 41, 43, 44, 48-50, 52, 55, curing 516, 517
58, 59, 220, 447, 469, 607, in lubricated squeezing flow 507
608, 621, 624, 626-629, 632, order of magnitude analysis in
643, 647, 657, 663 mold filling 506
Gate pressure 447 Heat transfer coefficient 266, 279,
Gate size 247, 443 284, 307, 571, 651
Gauss elimination 39 Heat transfer in molds 568
Gaussian network 381 Heat transfer rate 652, 653
Gel point 445 Hele-Shaw flow 46, 47, 447, 470,
Gelation 506 634
Generalized Hele-Shaw flow model Hermans orientation function 231,
487, 488, 509, 533 232
governing equation 489 Hermans orientation factor 558
isothermal Newtonian fluid 490 High-impact polystyrene (HIPS)
limitations 490, 491 6-10, 13, 14, 80, 82, 83, 86
mold closing force 538 Hold pressure (hydraulic) 65
Generalized Newtonian fluid 144, Hold time 65
159, 179, 206 Hole drilling method 257, 258
Geometry of shape factors 582 Hookean shear modulus 161
Qass-fiber filled 6-10, 82, 163, Hot-cone molding 360
167, 168 Hot core region 22, 25, 26, 32,
Glass transition region 380, 383, 125, 126
390, 400 Hot-feed extruder 438
694 Subject Index

Hot runner 621 Interaction matrix method 381


Hybrid equation 389 Interaction coefficient for fiber
Hybrid model 384 orientation 545, 546
Hydraulic radius 37, 163, 452, Interactive computer graphics
622, 624, 636, 642, 643 (IC G ) 610-620, 628-634, 641-
Hydrodynamic model 389, 390 668
Isobar 53-55, 58, 60
Isochromatic patterns 229, 259,
I 260
Isoparametric transformation
ICG 610-620, 628-634, 641-668 634
Ideal rubber 382 Isothermal case 33, 35, 37, 54,
Impermeable boundary 48, 52, 53, 57, 59, 60, 62, 72, 73, 75,
57, 70 122, 126
Indenbom theory 274, 277 Isothermal cure condition 413
Index of polydispersity 163 Isothermal modeling 622, 628,
In dicat rix equation 230 633, 643
Induction time for thermoset cure Isotherms 15
515 Iterative techniques 179, 187
Inelastic fluids 159, 221
Inelastic model 239, 240
Inertial effects 25, 45, 62, 67, 76,
139
Infrared dichroism 231
Je° (steady shear compliance) 8,
Infrared spectroscopy 331
79, 86-93, 95
Inhibitor 511, 512
Jaumann tensor derivative 197
Initial temperature 242, 246, 284
Jeffrey's equation 540-542, 552,
effect on thermal stresses 284
553
Injection machines for rubber
Jetting 247
437
John son-Segalman fluid 195
ram type 437
Joystick 614
in-line screw type 437
Juncture region 137, 218
out-line nonreciprocating-screw
type 437
Injection molding 137, 168
Injection pressure 247, 447
Injection time 427
In-mold coating 482 Kamal-Sourour kinetic equation
Insert 48, 54-58, 60, 61, 608, 643, 513
657 Keyboard 611, 615, 642, 650
Instron capillary rheometer 441 Kinetic characterization 330
Integral type Maxwell model 191 Kinetic energy correction 162
Integrated continuity equation Kinetic models 333, 458
487 Kinetic model, of cure 411, 424
Interaction in fiber orientation Kinetic parameters, unsaturated
543, 545 polyester SMC 514
Subject Index 695

Master curve 138, 174, 220, 221


Master plots (viscosity) 8, 11-15,
89-92, 107, 108, 119
Laminar flow 572
Mathematical modeling 363
Laminated elastomer composite
Maximum of reaction rate 337
424
Maxwell model 189, 191, 193, 235
Laplace equation 47, 53, 54, 56
Maxwellian-type equation 455
Layer-removal method 257, 260-265
MCAP 598
Lay-flat approximation 448, 451,
Measurement of fiber orientation
458
554-556
LCD 611
Mechanical property control -
Lead glass fibers 554
composites 482
Lee-Roberts-Woo theory 278
Mechanical spectrometer 445
Length of circuits 587
Mechanical spectrum 384
Leonov model 173, 195, 206, 220,
Melt elasticity 248
455
Melt fracture 8, 96, 101-103, 107,
Leveque thermal layer 126
396
Light pen 611, 613, 615, 660
Melt front 26, 30, 31, 33, 37, 39,
Linear shape function 50, 127,
40, 43, 44, 48-50, 52-54, 56-
636
58, 70, 72, 124-127, 220, 447,
Liquid crystal display (LCD) 611
627-629, 632, 633, 637-639,
Liquid flow 379
641-644, 646, 648
Local mesh distortion 493
Melt instability 92, 96
Local power-law index 109-111
Melt temperature 247, 251, 286,
Location of cooling lines 587
287
Lodgers rubberlike elastic liquid
Melting point 232, 265
191
Memory effects 78, 220
Loss modulus 90, 93-95, 97
Memory integral 192
Lost-core hollow-part injection
Mesh stretching 493, 494
360
Microprocessor 610
Low-angle X-ray diffraction 228,
Microscopic mechanical energy
231
balance 161
Low-profile resins 482
Mixed method 187
Lubricated squeezing flow 487,
Mixing cup (bulk) temperature
497-502, 535
41, 50
boundary conditions 503
Mold
heat transfer 507
assembly 616, 617, 620, 657-659,
665
base 607, 608, 617, 658, 659,
M
663, 665
multi-cavity 621, 622, 624, 625,
Machining damage 263 632
Macroscopic momentum balance 161 MOLD COOL 598
Marrucci model 235 Mold design 429
Mass conservation 23, 24, 45, 53, Mold design - thermal 520
62-78 Mold flow analysis 429
696 Subject Index

Mold parallelism 504, 505 Nonisothermal cure 415


Mold surface temperature varia­ Nonisothermal effect 220, 628,
tion 521, 522, 523, 524 629, 631, 634, 636
Moldability 444-446 Nonisothermal flow 221, 239
Molding simulation 424 Nonisothermal stress relaxation
Molecular mobility 298, 299, 393 235, 236
Molecular orientation 525 Nonlinear behavior during quench­
Mouse 614 ing 281
Mooney viscosity 441 Nonlinear stability analysis 189
Mooney viscometer 445 Nonlinearity 282
Monsanto processability tester Non-Newtonian behavior 350
441, 466 Nonuniform hardening and resid­
Monsanto rheometer 459 ual stresses 527
Morphology 228 Normal-stress difference 91-93,
Multi-gated cavity 41-44, 622, 96, 97, 102
625, 627, 630, 631 Normal stresses 248, 264, 285
Mn (number-averaged mol. w t.) Normalization condition for fiber
8, 15 orientation 545, 559
Mvv (weight-averaged mol. w t.) Nozzle 50, 446, 469
8, 15, 85-87, 90-92 Numerical iteration 30, 31, 39,
Mw/Mn 8, 14, 15, 78-81, 89, 90, 49, 52, 70, 116, 628, 630, 631,
93-97 637
Muki-Sternberg theory 276 Numerical relaxation 630, 637, 641
Multiparameter model 301 Numerical quadrature 30, 31
Numerically controlled (N C ) ma­
chine tools 609, 610, 659-664,
N 666
Nusselt number 572
Natural boundary condition 53, Nylon 2, 6-10, 13, 15, 107, 108,
127 112, 118
Natural convection 571 Nylon-6 moldings 287
Natural frequency 421
Natural rubber 413, 463
NC 609, 610, 659-664, 666
Network structure 381
Newtonian fluid 47, 54, 55, 62-64, Obstacles 137
67, 73, 126, 127, 137, 160, 180, Oldroyd model 191
398, 643 Oldroyd-B fluid 193
Newton-Raphson procedure 37, 39 One-dimensional flow 447, 448
Nitrile-butadiene rubber 463 On-line parallel-plate viscometer
NMWD (narrow mol. wt. distrib.) 345
6-10, 13, 14, 82, 83, 86, 89-91, Operating window 445
105, 107-109 Optimum parameters 592
Non-circular cooling lines 574 Order of magnitude analysis - com­
Noncircular duct 448, 453 pression mold filling 486, 488
Nonisothermal cavity filling 285 heat transfer in mold filling 506
Subject Index 697

Orientation 228, 622, 664 Polybutadiene 140, 142


Orientation distribution function Polarizer 229
544-546 Polycarbonate (P C ) 2, 6-10, 11-
Orientation function 231 15, 18, 19, 21, 22, 83, 108,
Orientation in molded parts 246 113, 118, 119, 163, 165, 167,
Orientation modeling 233 242
Orientation parameters 556-558 POLYCOOL 598
Orientation tensors 558, 559 Polyethylene
Oscillating disk rheometer 445 high density (HOPE) 2, 6-10,
Oscillatory deformation 282 13, 15, 19, 23, 79, 86, 87,
Oscillatory shear 384 97, 102, 103, 106, 107, 162,
Overshoot in centerline velocity 171, 286
192, 195 low density (LDPE) 2, 6-10, 13,
15, 19, 23, 97, 99, 103, 106,
107, 162, 175
blends 162
Polyisoprene 140, 142
Packing 366 Polyisobutylene (P IB ) 9, 206
Packing factor 399 Polymer-filler interaction 390, 417
Packing pressure 65, 66, 70, 73, Polymethyl methacrylate (PMMA)
74, 76, 77, 248 2, 6-10, 13-15, 18, 19, 21, 22,
Packing stage 62, 64-67, 70, 72, 79, 83, 84, 107, 108, 114, 118,
73, 75-77, 455 122, 242
Parallel plate 345 Polyphenylene oxide moldings 288
Parallel squeezing 529 Polypropylene (PP) 2, 6-10, 13,
Particle tracing method - fiber 15, 19, 20, 23, 41-43, 56, 61,
orientation 549-552 65, 66, 68, 78, 79, 83, 85, 89,
Part weight 622 90, 93, 96, 97, 100, 104, 106-
Partine line (plane) 608, 657, 666 108, 115, 117-120, 163
Peclet number 25, 111, 125, 126 glass-fiber filled 286
Perfect network 382 residual stresses 286
Phan Thien-Tanner model 193, Polystyrene (PS) 2, 3, 6-15, 18-
195 22, 31, 32, 56, 64, 65, 68, 70,
Phenolic systems 335, 362 73, 76-83, 86, 89, 93, 94, 97-
Photoelasticity 257, 259, 260 101, 105-108,117, 119, 120,
Photoelastic stress patterns 254 140, 142, 163, 165, 167, 168,
Photon-correlation spectroscopy 171, 173, 237, 239, 242, 247,
308 280, 294, 452, 454
Pixel 611, 613, 615, 642, 643 blends 78-80, 86
Planar flow 138, 176, 178 BMWD 11-14, 83, 89, 90, 92, 94,
Planar geometry 162 95, 105-108, 110, 116, 118
Planar two-dimensional flow 139 glass-fiber-filled 6-10, 82
Plasticized 6-10, 13, 15 HIPS 6-10, 13, 14, 80, 82, 83,
Platen 607 86
Poiseuille flow 127 NMWD 6-10, 13, 14, 82, 83, 86,
Poisson^s ratio 273 89-91, 105, 107-109
698 Subject Index

[Polystyrene (P S )] Primitives 448, 450


in Aroclor 140, 141 circular tube 448
moldings 287, 291-293, 295, 456, end-gated strip 448
457 center-gated strip 448
Polysidfone moldings 287 Principal stress difference 233
Polyvinyl butyral (P V B ) 9 Probability distribution function
Polyvinyl chloride (PVC ) 2, 6-10, fiber orientation 544-546
19, 23 Process control 609, 610, 664,
plasticized 6-10, 13, 15 665
unplasticized (rigid) 6-10, 13, Programmed function keyboard
15, 108, 117, 118, 120, 121 (PFK) 615
Porosity 459
Post-cure 404
Post-filling stage 1, 22, 36, 61-
78
Potential barrier 393 Quadratic shape function 50, 51,
Potential flow 48 127
Power-law fluid 161, 181, 184, Quadrilateral elements 634
234, 235, 395, 622, 633 Quenching of
squeezing flow 533 polycarbonate 265-268
Power-law index 2, 4-8, 11-14, polymethylmethacrylate 265-270,
16-19, 25, 27-29, 31, 38-41, 278-284
47, 54, 55, 58, 59, 75, 78, polyphenylene oxide 266
79, 88, 89, 91-93, 97, 102- polysulfone 266
106, 109, 111, 118-122, 124, polystyrene 267-270, 278-284
125, 161, 184, 350, 627 Quenching temperature
Prandtl number 25, 280, 572 effect on residual stresses 284
Precision molding 608, 657
Predictor-corrector procedure
44, 53 R
Preferential flow 510
Pressure Rabino wit sch correction 4, 5, 12,
gradient 25-27, 29, 31, 40, 44- 13, 164
47, 53, 55, 59-61, 69, 70, 74- Radiation 574
77, 180, 237, 239, 623, 628 Radiographic tracers for fiber or­
loss at juncture 8, 41, 58, 71, ientation 554
73, 96-105, 622, 647 Random scan 611
pack 65, 66, 70, 73, 74, 76, 77, Raster graphics 611-613, 615,
622 642, 654
traces 42, 58, 59, 64-68, 72, 73, Rate constants 335
76, 77, 644-646 Rate equations 300
transducers 56-59, 65, 66, 72, Rate-of-deformation tensor 179
73, 75, 76, 96, 644, 646 Rate theory 393
Pressure drop or loss 139, 144, Reaction kinetics - measurement
151, 152, 218, 239, 423, 589 511
Pressure singularity 181 thermoset cure 513, 514, 515
Subject Index 699

Recoverable strain 8 Reynolds number 25, 160, 572


Reduced compressive stress 266 Rheo-optical method 139, 472
temperature 266 Rheological characterization 344
Recirculation region 199, 204 Rheological properties 377
Reference temperature 301, 306, Rheology - reinforced thermo-
307 sets 528
Refractive index 230, 296 Rheometrics mechanical spectrom­
Refresh buffer 613 eter 345
Refresh display 612 Ribs 137, 220
Relaxation 238-240, 249, 262 Rigid silicone molding compounds
Relaxation distribution function 363
395 Robot 608, 610
Relaxation function 394 Rotational rheometry 441
Relaxation modulus, 280, 463 Rotating viscometer 19
Relaxation of residual stresses Roughness of pipes 591
263 Rubber 24
Relaxation spectrum 250, 377, Rubber like behavior 402
389 Rubber-metal composite 428
Relaxation time 7, 10, 173, 198, Rubbery flow 380
298, 306-308, 380, 395, 415, Rubbery state 380, 381
461 Runner 28, 33, 37, 39-43, 50, 66,
Relaxation of normal stress 140 607, 608, 621-633, 643, 647,
Reservoir 56, 58, 66, 73, 120 648, 657, 663, 664
Residual birefringence 242, 265, artificially balanced 623-626,
267 629, 630, 632, 633, 643
Residual cure 343 naturally balanced 623, 626,
Residual flow stresses 265 633
Residual stresses 459, 484, 525, Runnerless injection/compression
597, 649, 657, 664 360
effect of cavity thickness 291
effect of processing 285, 288,
289
effect of holding pressure 288
effect of injection rate 293 Scorch time 419
effect of injection pressure 286, Scorching 444-446
288 Scott equation 533-534
in moldings 285-288, 295 Screw extruder 446
in thermosets 523-528 Second normal-stress difference
modeling 273 234, 237, 264, 265
parallel to flow 289-291 Secondary flow 145
perpendicular to flow 289-291 Second-order fluid 189
Resins 359 Semicircular runner 453
Retardation 229 Shape factor (thermal) 652-654
Reversal of flow direction 144, Shape functions 50, 51, 127
151 Shear-induced crystallization
Reverse fountain flow 510 396
700 Subject Index

Shear modulus 282, 384 Specific heat 24, 25, 28, 31,
Shear rate 1, 2, 4, 5, 8, 9, 11- 43-45, 49, 50, 111, 114,
17, 24-31, 34, 40, 41, 43- 115, 118, 125, 341, 642, 646
47, 49, 50, 55, 62, 71, 72, Spectroscopic methods 332, 444
76, 79, 87, 104, 105, 108- Spencer-Gilmore equation of
110, 236 state 62, 69
Shear rate non dimensional 236 Spiral flow 344
Shear stress 2, 13, 16, 17, 24, Sprue 38, 41, 42, 66, 446, 469,
45, 72, 74, 79, 88, 91-93, 607, 608, 618, 619, 621, 647,
96, 97, 102, 103, 127, 140, 648
202, 233-235, 238, 239, 242, Square grid patterns 267
246, 264, 285 Squeezing flow 345
Shear stress maximum 238 Squeezing flow - power-law fluid
Sheet molding compounds 363 533
Shift factor 301 stress overshoot 535
Shock isolators 381 viscoelastic fluids 533-535
Short fiber composites - mechan­ with stratified viscosity 528,
ical property control 482 530
Short shot 57, 58, 427, 628, 643, Stanton diagram 591
644, 646 State of cure 459
Shot size 608, 667 Steady natural convection 280
Shrinkage 242, 264, 296, 455-457, Steady-shear viscosity 395
471, 579, 608 Stefan-Boltzmann constant 574
modeling 297 Storage modulus 90, 93-97, 348
multilayer model 297 Storage (computer) requirements
Silicone elastomer 396 54
Simplex method 2, 71, 111, 112, Storage tube display 612
122, 123 Stored elastic energy 159, 162
SIMUCOOL 598 Strain-optical coefficient 268
Simulation of post-filling stage Strain-optical rule 268, 277
471 Stream function 54, 55
Single-parameter model 301 Stream wise -integration procedure
Sink marks 482 187, 198
Size of cooling lines 587 Stress analysis 663, 666
Slip effect 441 Stress concentration 248
Slip velocity 465, 468 Stress component perpendicular
Slit die 161, 171 to plane 267
Slit rheometer 96, 441 Stess distribution in converging
SMC - equilibrium thickness 535 dies 140
kinetic parameters 514 Stress-optical coefficient 233, 234,
preheating 484 237, 242, 243, 260, 265, 294
Sodium-butadiene rubber 443 effect of memory 260
Solidified layer 239 of polycarbonate 243
Solid modeling 610, 616, 618-620, of polymethylmethacrylate 243
653, 654, 659-661, 663-666 of polystyrene 237, 242, 265,
Sol-gel analysis 331 294
Subject Index 701

Stress-optical laws 139, 206, 233, Temperature profiles during


260, 265, 294 curing 518, 519, 520
Stress relaxation 90, 250, 268, Tensile modulus 463
287 Tensile stress 265, 463
Stress relaxation after quench­ Tensor of refractive index 228
ing 281 Theoretical minimum cooling time
Stress relaxation method 257-259 575
Strip cavity 23, 29, 31, 36, 39, Thermal birefringence 242, 245,
41, 50, 57, 64-78, 125, 126 246, 294
Structural relaxation 277 Thermal boundary layer 25, 30,
Styrene-acrylonitrile (SAN ) 2, 32, 50, 125, 126
6-10, 13, 14, 19, 23 Thermal conduction 640, 651
Styrene-butadiene rubber 462, Thermal conductivity 24, 25, 28,
463 30, 31, 43, 44, 49, 50, 111,
Successive over relaxation 53 114, 115, 118, 125, 266, 280,
Successive substitution 191 307, 350, 353, 640, 642, 646,
Successive underrelaxation 39, 651
51-53, 71 Thermal convection 234, 239, 254,
Sudden axisymmetric contraction 640
186 Thermal degradation 506, 519,
Surface layer 251, 252, 254 520
Surface strain 267 Thermal design of molds 520
Surface stress 267 Thermal diffusivity 25, 49, 65,
Surface tension 127, 233 111, 125, 126, 350, 353
Surface waviness 482 Thermal layer 239
Thermal properties of materials
571
Thermal properties of water 586
Thermal shrinkage 278
Tablet 613, 614 Thermal stress 264, 285, 294, 459
Temperature calculation 278, 280
air-shot 50, 646 distribution 266
coolant 26, 646, 650, 652, 653 effect of bath temperature 268
glass-transition 18-22, 77, 78, effect of initial temperature 268
82, 85 effect of thickness 268
inlet 25, 30, 32, 35, 41, 42, 48, in inorganic glass 273
50, 57, 65, 72, 76, 115, 120- relaxation 245, 270-272
123, 125, 632, 642, 646, 650, Thermodynamic fluctuation theory
652, 653 394
mixing cup (bulk) 41, 50, 628, Thermoelastic theory 266
630, 642, 648 Thermoset cure - reaction kinet­
wall 26, 30, 32, 35, 41, 42, 45, ics 513, 514, 515
57, 65, 72, 76, 115, 632, 642, Thermo sets 329
646, 650 Thick molding compounds 363
Temperature distribution in a Thickening reaction 511, 537
mold 447, 521, 522, 523, 524 Thixotropy 439
702 Subject Index

Three-dimensional mold geometry Vector display 615


610, 637, 642, 643, 651-657 Vector-refresh unit 612
Tool-path verification 610, 660- Velocity 24-26, 31-33, 44-46, 48,
663, 666 49, 53-55, 62, 67-69, 128
Total pressure drop 180 Velocity centerline 192
Total stress tensor 179 Velocity profile
Transfer molding 421 in axisymmetric flow 180, 181
Transient conduction 234, 235, in planar flow 180, 181
239 Velocity, gap wise distribution
Transient flow 573 236
Transient stresses 277 Vent 608
Transmissibility 421 Vibration isolator 381
Trapezoidal runner 453 Viscoelastic fluid 144, 160, 186,
Tree structure 448 221
Triangular elements 50, 127, 634- Viscoelastic fluids - squeezing
637, 639, 642-644, 648 flow 533-535
Tridiagonal elimination 30, 70, Viscoelastic model 235, 239, 240
115, 279 Viscoelastic modeling in molding
Trifringence 228, 230 239
Tubular elements 635, 637, 639, Viscoelastic simulation 221
642-644, 648 Viscosity 1-29, 31, 34, 41, 44-47,
Turbulent flow 573 50, 55, 70-72, 74, 77, 79, 96,
Two-dimensional flow 176, 206, 97, 102-104, 108-110, 112,
220, 221, 447, 448 116, 121, 127, 167, 173, 299,
Type of cooling lines 587 332, 383, 642, 644, 646, 648
Carreau model 104-110
Cross model 2, 79, 104-110
U data bank 2, 3, 6
five-constant model 17, 18, 35,
Unit cell 231 36, 71-73, 75, 76, 113, 118,
Unplasticized 6-10, 13, 15, 108, 120, 124, 644, 646
117, 118, 120, 121 four-constant model 2, 4, 11,
Unsaturated polyesters 334, 363 31, 33-35, 61, 81, 82, 91
Unsaturated polyester - kinetic master plots 8, 11-15, 89-92,
parameters 514 107, 108, 119
curing 506, 511, 512 Newtonian limit 2, 4, 6-19, 21,
Upper-convected Maxwell model 47, 54, 62, 73, 74, 85-87, 93,
189, 201 94, 96, 97, 102-109, 113, 119,
Upwind numerical procedure 28, 127
51, 640 power-law limit (model) 2, 5-7,
9, 11, 13, 16, 25, 27-29, 31-
35, 38-41, 47, 55, 56, 58, 59,
V 61, 73-75, 79, 86, 107, 109,
119, 121, 122, 124
Variable density 62-78 pressure dependence 1, 15-25,
Variable-thickness cavity 56-61, 68 36, 71, 73, 74, 110-125
Subject Index 703

[Viscosity] W
seven constant model 18, 19,
22, 35, 36, 71, 72, 76, 78, Wall shear stress 4, 5, 8, 35, 40,
113 60, 61, 75, 97-107, 119
shear-thinning effect 2-15, 47, Wall temperature 26, 30, 32, 35,
55, 71, 75, 79, 87, 88, 102- 41, 42, 45, 57, 65, 72, 76,
110, 128 115
temperature dependence 1, 2, Warpage 264, 459, 594, 649, 657
6-8, 13, 16-22, 24, 25, 27, Warpage in molding 288
34, 36, 71, 80-86, 110, 111, effect of cooling 280
122, 126, 128 effect of geometry 288
Viscosity-elastomer-filler inter­ Weissberg^s equation 160
action 415 Weissenberg-Rabinowitsch correc­
Viscosity model 401 tion 4, 5, 12, 13
Viscosity reduction 167 Weissenberg number 188, 198
Viscous dissipation 159 Weissenberg rheogoniometer 345
Viscous effect 139 Weldline 41, 43, 44, 48, 49, 52,
Viscous heating 13, 14, 25, 28, 57, 58, 447, 448, 622, 625,
30, 32, 33, 35, 36, 44, 51, 627, 630, 631, 643-648
71, 72, 107-121, 234, 396, Wide angle x -ray diffraction 228,
423, 621, 622, 640 231
Volumetric flow rate 3, 4, 23, 27, Wire-frame representation 610,
31-38, 40-44, 48-50, 52, 57, 616, 617, 641, 642, 653, 654,
58, 67, 69, 72, 76, 122, 625, 659, 660, 663
627, 628, 630, 632, 636, 642, WLF equation 18, 20, 36, 81-84,
646, 647 301, 349, 383, 384
Volume fraction 390, 399
Volume relaxation 300
Volumetric shrinkage 302
effect of processing 314
prediction 316 Yield stress 439
Vulcanization 403 Young^s modulus 273, 278
Vulcanization in the mold 459 in molding 261
Vulcanization stage 455 gapwise distribution 262

You might also like