You are on page 1of 145

This article was downloaded by: [Princeton University]

On: 03 June 2013, At: 06:30


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number:
1072954 Registered office: Mortimer House, 37-41 Mortimer Street,
London W1T 3JH, UK

Journal of Macromolecular
Science, Part C
Publication details, including instructions for
authors and subscription information:
http://www.tandfonline.com/loi/lmsc19

Mechanical Properties of
Polymers: The Influence
of Molecular Weight
and Molecular Weight
Distribution
a b
John R. Martin , Julian F. Johnson & Anthony
c d
R. Cooper
a
Institute of Materials Science
b
Institute of Materials Science and Department
of Chemistry, University of Connecticut, Storrs,
Connecticut, 06268
c
Chevron Research Company, Richmond,
California, 94802
d
Department of Chemistry and Institute of
Materials Science, University of Connecticut,
Storrs, Connecticut, 06268
Published online: 06 Dec 2006.

To cite this article: John R. Martin , Julian F. Johnson & Anthony R. Cooper
(1972): Mechanical Properties of Polymers: The Influence of Molecular Weight and
Molecular Weight Distribution, Journal of Macromolecular Science, Part C, 8:1,
57-199
To link to this article: http://dx.doi.org/10.1080/15321797208068169

PLEASE SCROLL DOWN FOR ARTICLE


Full terms and conditions of use: http://www.tandfonline.com/page/
terms-and-conditions

This article may be used for research, teaching, and private study
purposes. Any substantial or systematic reproduction, redistribution,
reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make
any representation that the contents will be complete or accurate or
up to date. The accuracy of any instructions, formulae, and drug doses
should be independently verified with primary sources. The publisher
shall not be liable for any loss, actions, claims, proceedings, demand, or
costs or damages whatsoever or howsoever caused arising directly or
indirectly in connection with or arising out of the use of this material.
Downloaded by [Princeton University] at 06:30 03 June 2013
J. MACROMOL. %I.-REVS. MACROMOL. CHEM., C8(1), 57-199 (1972)

Mechanical Properties of Polymers:


The Influence of Molecular Weight
and Molecular Weight Distribution
Downloaded by [Princeton University] at 06:30 03 June 2013

JOHN R. MARTIN
Institute of Materials Science
JULIAN F. JOHNSON
Institute of Materials Science and Department of Chemistry
University of Connecticut
S t o w s , Connecticut 06268
ANTHONY R. COOPER*
Chevron Research Company
Richmond, California 94802

I. INTRODUCTION............................ 58
11. INDIRECT EFFECTS . . . . . . . . . . . . . . . . . . . . . . . . 59
111. STRESS-STRAIN PROPERTIES .................. 64
A. General Considerations . . . . . . . . . . . . . . . . . . . . . 64
B. Tensile Strength, Tenacity . . . . . . . . . . . . . . . . . . . 66
C. Elongation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
D. Yield Point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
E. Brittle Point . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
F. Modulus.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

*Current address: Department of Chemistry and Institute of Materials


Science, University of Connecticut, S t o r r s , Connecticut 06268.

57
Copyright 0 1972 by Marcel Dekker, Inc. NO PARTof this work may be reproduced or utilized in any
form or by any means, electronic or mechanical, including xerography, photocopying, microfdm, and
recording, or by any information storage and retrieval system, without the written permission of the
publisher.
58 MARTIN. JOHNSON. AND COOPER

IV . IMPACT STRENGTH ......................... 129


V . STRESS CRACKING AND CRAZING . . . . . . . . . ...... 136
A . Introduction . . . . . . . . . . . . . . . . . . . . . . . . .
.... 136
B . Amorphous Polymers ...................... 138
C . Crystalline Polymers ...................... 141
VI . FATIGUE AND ENDURANCE . . . . . . . . . . . . . . . . . . . 151
VII . DYNAMIC STRESS-STRAIN PROPERTIES . . . . . . . . . . 155
A . General Considerations . . . . . . . . . . . . . . . . . . . . . 155
B . Stress Relaxation . . . . . . . . . . . . . . . . . . . . . . . . . 156
C . Creep . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
D. Vibration Measurements . . . . . . . . . . . . . . . . . . . . .
Downloaded by [Princeton University] at 06:30 03 June 2013

168
E . Internal Losses . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
F . Ultrasonic Properties . . . . . . . . . . . . . . . . . . . . . . 174
VIII . GLASS TRANSITION TEMPERATURE . . . . . . . . . . . . . 175
IX . MELTING TEMPERATURE .................... 183
X . MISCELLANEOUS PROPERTIES . . . . . . . . . . . . . . . . . 184
A . Hardness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
B. T e a r Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
C . Burst Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
D . Heat Distortion and Softening Temperatures . . . . . . . 186
E . Wear and Abrasion . . . . . . . . . . . . . . . . . . . . . . . . 186
ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . 188
REFERENCES ............................. 189

I . INTRODUCTION
The molecular weight and molecular weight distribution of a
polymer can markedly affect its mechanical properties . While this
has been recognized for a long time. quantitative studies have been
relatively difficult to perform for a number of reasons . A major
problem has been the characterization of the complete molecular
weight distribution curve by fractionation methods because fraction
collection and the determination of the various molecular weight
averages of polymers is time consuming . Additionally. measurement
of the number-average molecular weight. Hn. by classical osmometry
procedure is often unreliable .
The precision of some mechanical tests is low s o a substantial
MECHANICAL PROPERTIES OF POLYMERS 59

amount of sample is required for replicate testing. This is unfor-


tunate because the preparation of even small amounts of a fractionated
polymer which has a narrow and defined molecular weight distribu-
tion is tedious.
Recent advances have considerably improved the situation. For
example, gel permeation chromatography now provides a rapid
precise method for measuring molecular weight distributions. Cali-
bration with polymer fractions of known molecular weight is required
but, once calibrated, the determinations a r e very rapid. Availability
of recording dynamic osmometers has materially improved both the
speed and accuracy of a n measurements. Thus two of the difficulties
Downloaded by [Princeton University] at 06:30 03 June 2013

described above no longer exist. Preparation of sufficient amounts


of narrow fractions is still difficult but recent improvements in
techniques of polymer fractionation have reduced the effort required
to the point where it is feasible to prepare such fractions.
There appears to be no current comprehensive review on the
effect of molecular weight and molecular weight distribution on the
mechanical properties of polymers. In Compiling this paper an
attempt has been made to include most of the pertinent mechanical
properties and transitions of solid polymers. Ftheology of melts and
solutions is excluded. The subjects of gel permeation chromatog-
raphy and polymer fractionation a r e covered in various reviews
[38,60,139]and only topics of specific application to solid-state
properties a r e included here.
It is necessary to carefully distinguish between molecular weight
and molecular weight distribution. The presence of high and low
molecular weight “tails” in fractions which have unknown o r broad
distributions can often mask the effect of molecular weight on physi-
cal properties. This is one of several considerations which have
made questionable the significance and reliability of some of the
literature data reported here. The evaluations, while often severe,
a r e not intended a s a criticism of previous research. In most cases
they a r i s e because of presently available techniques which did not
exist when the cited research was carried out,

II. INDIRECT EFFECTS


Solid-state properties can be correlated with a number of parame-
t e r s : molecular weight (MW) molecular weight distribution (MWD),
branching, orientation, crystallinity, crystal structure, processing
conditions, etc. These variables a r e not independent s o a change in
either MW o r MWD can influence material properties through its
effect on one of the other parameters. For example, when uniaxially
60 MARTIN, JOHNSON, AND COOPER

oriented films of atactic polymethyl methacrylate (PMMA), poly-


styrene, polyvinyl acetate, and poly(2,2’-octamethylene-5,5’bibenzi-
midazole) were tested, it was found that molecular weight did not
directly affect tensile strength at the higher molecular weights
[164-1661. It did control orientation however. A s a result, the struc-
ture and properties of the oriented products were indirectly deter-
mined by the MW of the samples. MWD also influenced orientation
characteristics a s shown in Ref. 166 and Figs. 1 and 2.
A- STRETCHING TEMPERATURE 16O0C(SAMPLES OF FRACTIONATED
POLYMERS)
8- STRETCHING TEMPERATURE I7O0C( SAMPLES WITH BROAD
Downloaded by [Princeton University] at 06:30 03 June 2013

DISTRIBUTION OF MOLECULAR WEIGHT)

0 I I I I I I
0 0.5 I .o 1.5 20 25 30
MV x lo6

Fig. 1. Maximum degree of stretching, Am. of PMMA films as a function of


the molecular weight, Ev [166].

In one of the more comprehensive studies of MW effects, William-


son [298]fractionated several pounds of two high-density polyethyl-
enes by coacervation. Intrinsic viscosity measurements on subfrac-
tions from some of the samples indicated that the fractions had a
MECHANICAL PROPERTIES OF POLYMERS 61

C- STRETCHING TEMPERATURE = 160% (SAMPLES OF FRACTIONATED


POLYMERS)

0- STRETCHING TEMPERATURE- 17OoC( SAMPLES WITH BROAD


DISTRIBUTION OF MOLECULAR WEIGHT)

*Oo0r
I
I-
W
z
w
Lz
I-
v)
Downloaded by [Princeton University] at 06:30 03 June 2013

0 I I I I
0 05 10 15 2 .o 2.5 30
M, x lo6
Fig. 2. Maximum tensile strength of P M M A films a s a function of molecular
weight 1166).

Q (= aw/’Mn) of 1.1 t o 3.1 as compared to Q = 15 i n the original


materials. (Gw is the weight-average molecular weight.)
When molded under identical conditions, the density (or crystal-
line content) of the fractions decreased as the molecular weight was
increased (Fig. 3). Densities of the fractions were lower than the
corresponding values measured on the parent polymers. These
trends were attributed to crystallization kinetics. They imply that
performance in properties which are primarily controlled by the
degree of crystallinity can be reduced if a high M W , narrow distri-
bution material is used. Many investigators have reached s i m i l a r
conclusions with respect t o the effect of molecular weight on c r y s -
tallization kinetics [Chapters 2 and 7 of Ref. 23, and Refs. 51, 176,
181,183,184,216,279] and percent crystallinity [Chapters 2 and 7 of
Ref. 23, and Refs. 52,76,85,102,154,181,183,187,203,233,255,274,300,
3031. Numerous r e p o r t s [52,76,154,184,187,203] indicate that low
MW material can a l s o reduce crystallinity because of the increased
concentration of chain ends (Figs. 4 and 5).
Changes in crystal structure and morphology due t o variations in
62 MARTIN, JOHNSON, AND COOPER

I I I 1 -

0.96 - -

4
$ 0.9s - -
Downloaded by [Princeton University] at 06:30 03 June 2013

*
c
u)
z
W
0

0.94 - -
Y

0.93 - 0
-
I I 1 I
0 2 4 6 8

Flg. 3. Effect of intrinsic viscosity on density of polyethylenes [298]:


( A ) Phillips polyethylenes, ( A ) Ziegler polyethylenes, (0)fractions from
Phillips polyethylenes, (0)fractions from Ziegler polyethylenes.

MW can be quite substantial [6,7,13,Chapter 2 of 23,183,194,213,2381.


One example, which relates M W to spherulite diameter and brittle
temperature, is given in Table 1. Ross [238] investigated the effects
of MW, MWD, extrusion temperature, orientation, crystal structure,
and draw ratio on polypropylene filaments. He found that the crys-
talline state of undrawn fibers could be correlated with extrusion
temperature and M W D ; fibers produced from broad distribution
resins had a monoclinic structure while the narrow distribution
resins gave paracrystalline fibers. Upon drawing, all of the fibers
assumed the monoclinic structure, The study also showed that the
degree of orientation was lower for resins with low a w . When com-
MECHANICAL PROPERTIES OF POLYMERS 63

L I N E A R PARAFFINS
Downloaded by [Princeton University] at 06:30 03 June 2013

Weight - Average M o l e c u l a r Weight

FIg. 4. Dependence of the crystalline fraction of linear polyethylene on molecu-


l a r weight and thermal history (1871.

*k0.940-
m
5 0.930- o Commercial
0
0.920. \Experi menta I
I,
10 I o4
:jZiegler PES
I o5 I06
MOLECULAR WE I GHT
Fig. 6. Density vs. molecular weight of high density polyethylene fractions
[154].

pared at constant draw ratio however, IWn had little effect on modu-
lus and tenacity except a t the lower molecular weights (Fig. 6).
From the above discussion it is clear that, in some cases, MW
and M W D can affect physical properties through their effects on other
material parameters. In other cases the relationships are either
small o r tenuous. Unfortunately it is difficult to determine the sig-
64 MARTIN, JOHNSON, AND COOPER

Table 1
Effect of MW on Spherulite Structure of
Polyethylene and Polypropylene [ 13 I

Brittle Spheruli te
Polymer MW range temp ( O F ) diam ( p )

Polyethylene 280,000 <-80 6-10


550,000 <-80 3-8
1,000,000 <-80 1
Polypropylene 200,000 125 3-15
Downloaded by [Princeton University] at 06:30 03 June 2013

300,000 80 2- 6
500,000 60 1- I

nificance of these secondary effects because of the coupling between


parameters. Whenever possible, these effects, whether due to MW
or some other variable, have been noted. However, an exhaustive
compilation of indirect effects would be impractical s o this report
is largely limited to direct relationships between MW o r MWD and
physical properties.

In. STRESS-STRAIN PROPERTIES


A. General Considerations
As mentioned above, the specific effects of MW o r MWD can be
determined only if all other variables a r e held constant or allowed
for quantitatively. The simplest systems a r e films of amorphous
polymers cast from solution. This eliminates orientation effects
which may be introduced during compression molding and definitely
a r e present with injection molding. Crystallinity causes further
complications in that this will change a s the MW is changed to study
the MWD effect. The stress-strain properties also depend on the
speed of testing, loading mode, environment, and the temperature
relative to the glass transition temperature. Discussion of these
effects and the related theories a r e included in several recent books
[9,17,234,237,245].
The question of which MW average should correlate with the
physical and mechanical properties of polymers has not been settled.
Flory [81] developed an equation based on experimental work by
Sookne and Harris [258] on fractionated cellulose acetate and his own
work on butyl rubber vulcanizates [82]:
Property = A B
--
zn
MECHANICAL PROPERTIES OF POLYMERS 65

The tensile strengths depended explicitly upon Mn, and blends


followed weight average summation of the component properties.
Flory showed that these two statements were compatible to a first
approximation. The similar behavior of cellulose acetate and vul-
canized butyl rubber was considered to be coincidental. The de-
pendence of tensile strength of tutyl rubber upon l / R n at constant
cross-linking could be explained, but the linear relationship for
cellulose acetate could not.
There is no theoretical basis for expecting polymer properties to
depend upon a particular MW average. Nor is there any reason to
expect the same relationships to apply to amorphous, semicrystalline,
Downloaded by [Princeton University] at 06:30 03 June 2013

90
80
DR 5

-6.0 DR 4
* 5.0
0 DR 3
a4.0
z
E30, DR 2
2.0,
- -I- . ,
1.0. UNDRAWN

3 mi

aopoo 1e0,ooo -Mw 210,000 240,000

Fig. 6. Tenacity and modulus a s a function of molecular weight of f i b e r s from


Polymer D a t various d r a w ratios [238].
66 MARTIN, JOHNSON, AND COOPER

and cross-linked polymers. When one attempts to interpret a prop-


erty change in terms of distribution effects, the problem becomes
much more complex because, as shown in Tables 2 and 3 , there can
be many reasons for a relationship between a particular property
and MWD. If one assumes that a property depends solely upon G n
up to a particular Gn,gin, then Table 2 shows the M W D dependence
for various values of Mn and GW. A similar tabulation would apply

Table 2
Effect of MWD upon a Property Solely Dependent on E n
Downloaded by [Princeton University] at 06:30 03 June 2013

Property-MWD
Case MW data dependence
-
Mn constant 5 E n , m i n
-
Mw constant or varying None
- -
Mn varying 5 Mn,min
-
Mw constant Narrow MWD
better proper-
ties
MWD effect
variable
None

to properties which were functions of MW except that, in Case 2, a


broader MWD would lead to better properties.
In some publications it is suggested that properties depend on
both G n and aw,with each having a minimum value above which
maximum values a r e obtained. This hypothesis leads to the four
possible combinations outlined in Table 3.
B. Tensile Strength, Tenacity
1. Amorphous Polymers below Tg. The effect of ii?n and GWon
t h e stress-strain properties of solvent cast polystyrene films was
examined by Merz et al. [196]. Maximum film strength was obtained
for all samples which simultaneously had GW> 1.5 x lo5 and G n
> 2 X lo4. The detailed shape of the MWD curve was unimportant in
samples which met both of these criteria. Below Mn = 2 X lo4, tensile
strength was increased by narrowing the MWD at constant Gw be-
cause tensile properties increase with G n in this region. These
MECHANICAL PROPERTIES OF POLYMERS 67

Table 3
Effect of MWD upon a Property Dependent on Enand m,
Case MW data Property-MWD dependence
-
1 Mn
- 5 mn
- minimum
M, 2 M, minimum None
-
2 Mn 2 Mn minimum a. At constant an-broad MWD
- better properties
M, < KWminimum b. At constant mw-none
c. Vary both E n and mw, proper-
Downloaded by [Princeton University] at 06:30 03 June 2013

ties depend on & but MWD


effect variable
-
3 Mn < E n minimum a. Constant En-none
-
Mw 2 mw
minimum b. Constant a r n a r r o w MWD
better properties
c. Vary both ii?n and EWproper-
ties depend on a n but MWD
effect variable
-
4 Mn < ii?n minimum a . Constant Mn-broad MWD better
properties
Til, < XT, minimum b. Constant mw-narrow MWD
better properties
c. Vary
- both an and a w - h i g h e r
M’s favor better properties,
MWD effect variable

results a r e valid only for nonoriented, strain-free polystyrene. An


earlier study [195] by the same group showed that the tensile strength
of solvent cast films was also affected by the choice of solvent. Poly-
styrene film cast from methyl ethyl ketone had higher tensile
strengths and lower creep rate than those cast from benzene solution.
The conclusion was that some small amount of tenaciously held ben-
zene was responsible for the poorer properties. It was suggested
[lll] that this could also be explained on an entanglement basis;
films cast from the thermodynamically poorer solvent yielded films
with a greater degree of molecular entanglement.
McCormick et al. [ 1891 used anionic, isothermal, and thermally
prepared polystyrenes to examine a number of physical properties.
The tensile samples were prepared by compression and injection
68 MARTIN, JOHNSON, AND COOPER

molding and characterized by ultracentrifuge sedimentation velocity


data. They all had G n above the lower limit reported previously [ 1961
for constant property levels in unoriented polystyrene. Several
-
samples, however (their anionic Samples S4 S7 and isothermal
Sample Bl), had GWbelow the lower limit set for constant property
levels, 1.5 x lo5. The data presented for tensile strength and elonga-
tion a r e consistent with constant property levels for compression-
molded samples having simultaneously G n > 2 x lo4 and ii?, > 1.5
X lo5. In Fig. 7A the original plot of the authors is shown for tensile
strength as a function of Gn. For compression-molded specimens
the authors suggest that a MWD dependence can be seen by the two
Downloaded by [Princeton University] at 06:30 03 June 2013

dashed lines. If one accepts the results of Merz et al. that are cited
above 11961, it is possible to reinterpret this data as shown in Fig. 7B.
Most of the points can be represented within experimental e r r o r by
-
Line A at a tensile strength of 5000 psi. Samples B1 and S4 S7 fall
below this line solely because of their low Gw. A plot of tensile
strength against fiw (Fig. 7C) shows that all of the points except B1
lie on two lines regardless of their MWD. Point B1 is the only datum
with both a broad MWD and an @W below the limiting value for con-
stant property levels s o under this method of interpretation, there
would be insufficient data to examine MWD effects.
Thomas and Hagan [266] investigated the effect of MWD on the
processing and mechanical properties of polystyrene. Two samples
which had the same MW (1.9 x lo5) were compared; one had a Q of
2.6, the other 1.06. With samples taken from injection-molded sheets
the narrow distribution material was shown t o have consistently
higher tensile strengths than the broad distribution material. The
broad distribution material also exhibited greater anisotropy at all
of the molding temperatures. This was attributed to its greater melt
elasticity. As shown in other sections, elongation at rupture and
tensile creep properties were also improved in the narrow distribu-
tion sample. Much of these data are also included in an earlier paper
[ 1091.
A series of physical property measurements were made on anion-
ically polymerized polystyrene (& < 1.1) of varying molecular weights
[299]. Both linear and branched (tetrachain startype molecule) mate-
rials were included in this study as well as one broad distribution
commercial polystyrene. As in the studies noted above, the tensile
strength of the linear fractions showed no marked MW effects once
the MW exceeded 1.6 X lo5. The minimum MW was somewhat greater
for the tetrachain polymers, however, because a reduction in Mv
(viscosity-average molecular weight) from 4.3 x lo5 to 2.04 x lo5
decreased the tensile strength from 6100-7300 to 4900 psi. The
MECHANICAL PROPERTIES OF POLYMERS 69

latter value brackets the tensile strength of the broad distribution


linear polymer (Hv = 2.29 x lo5).
In contrast to the results obtained in Refs. 189, 196, and 299,
Bondurant [24] found that the tensile strength of polystyrene fractions
does not become constant until GW= 3.5 x lo5 (Fig. 8). A s e r i e s of
blends was also prepared such that the peak in each MWD curve
occurred a t D P (degree of polymerization) = 3460. Despite the fact
that these blends had different normal and skewed distributions, the
tensile strength a t this D P m U was unaffected by the shape of the
distribution. The results of tests on another s e t of blends is included
Downloaded by [Princeton University] at 06:30 03 June 2013

in Tables 4 and 5 and plotted in Fig. 9. The MV of the blends was


calculated using the assumption that the component fractions were
perfectly homogeneous. These data show an MWD effect which quali-
tatively agrees with that reported by McCormick et al. [189]. How-
ever, as noted above, the MW range in these blends is well above the
minimum limit required for constant properties [ 189,196,2991.
Although most of the evidence outlined above was obtained in tests
performed on polystyrene, the results appear to be qualitatively
applicable to most amorphous polymers. F o r example, when Kishi
and Kamata [ 1551 irradiated PMMA specimens t o obtain samples of
different molecular weights, they found that tensile strength was
almost independent of MW a t higher MW. However, when the degree
of polymerization fell below 1500 (%w < 1.5 x lo5),tensile properties
were severely diminished.
Berry [22]measured tensile strength and fracture surface energy
in PMMA specimens whose nv varied from 9.8 x lo4 to 6 X lo6. He
showed that fracture surface energy is inversely proportional to MW.
Extrapolation of the data showed that zero surface energy, and hence
zero tensile strength, would be expected when the MW was reduced
to 2.5 x lo4. Actual measurements of tensile strength showed little
MW dependence for MV > 9.8 X lo4. The effect of MW on the flexural
properties of a polycarbonate, poly-[2,2-prapane-bis(4-phenyl c a r -
bonate)], has also been studied [ 1001. Irradiation produced chain
scission without cross-linking, and viscometry was used to charac-
terize molecular weights. The results for the MV dependence of
tensile strength and flexural results are shown in Figs. 10 and 11.
Tensile strength appears to be virtually constant for fiv > 1.5 X lo4
but it decreases to zero a tzv- 8 x lo3. Flexural strength follows
a similar dependence. A reasonable correlation between these
parameters and l/zv was obtained in the brittle region (mv = 8 x lo3
to 1.5 x lo4). In the ductile region, these properties were independent
of MW. MWD of these polymers was not discussed.
A number of physical properties were reported for poly(norborn-
70 MARTIN, JOHNSON, AND COOPER

ANIONIC
m ISOTHERMAL
A THERMAL
Downloaded by [Princeton University] at 06:30 03 June 2013

(a) on 10-5

Fig. 7A. A plot of tensile strength of compression-molded polystyrenes


against number-average molecular weight [l89].

6-
I !IA
-
r
I-
w 4-
z
Wm
Eb -
L n x
W V ,

Ln
z 2-
W
I-

- 81 I

0 I I I I
MECHANICAL PROPERTIES OF POLYMERS 71

5- - m D
A

.A
0
x
-
4-
I
c
0
z ANIONIC
2 3- ISOTHERMAL
G A THERMAL
Downloaded by [Princeton University] at 06:30 03 June 2013

W
2
ul
z 2-
W
c

l -

I I 4 1 1 I I 1
0 ;
I0 20 30 40 50

Flg. 7C. A plot of tensile strength of compression-molded polystyrenes


against weight average molecular weight [189].
ene) in Ref. 93. Tensile strength increased when intrinsic viscosity
[77] and the ratio of trans to cis unsaturation were increased simul-
taneously. The data which were presented did not separate the effects

Fig. 8. Tensile strength vs. molecular weight of fractions, polystyrene [24 I .


72 MARTIN, JOHNSON, AND COOPER

Table 4
Tensile Strength Variation of Special Blends
of Polystyrene Prepared from Two Fractions [24]

Fraction Fraction - Tensile strength Number of


5 (%I 12 ( % ) M~ x 1 0 - ~ ( p s i ) x lo-* samples

100 0 5.0 43 32
90 10 4.5 44 26
80 20 4.1 45 30
70 30 3.6 41 26
65 35 3.4 20 14
Downloaded by [Princeton University] at 06:30 03 June 2013

Table 5
Tensile Strength Variation of Special Blends
of Polystyrene Prepared from Two Fractions [24 ]

Fraction Fraction - Tensile strength Number of


1 (%I 12 ( % ) M~ x ( p s i ) x 10-2 samples

100 0 14.6 46 34
80 20 11.8 48 32
70 30 10.3 46 37
60 40 8.9 27 28

of structural isomerism from those due to variations in [q].


2. Amorphous Polymers above Tg. Flory [81,82] derived a rela-
tionship between the tensile strength and the MW of a butyl rubber
vulcanizate at constant cross-link density:

Experimental verification of this relationship is shown in Figs. 12


and 13 using fractionated samples with GVranging from 1.7 x lo4 to
1.08 x lo6. The tensile strength of both fractions and blends was
related to d n . The only exception was in blends which contained a
low MW component because the low MW materials could not be ap-
preciably incorporated into the cross-linked network.
Using a standard t i r e tread recipe, Yank0 [301]ran a similar
study on 25/75 SBR fractions (En = 1.24 X lo4 to 1.65 x lo6). He found
that only the six fractions with an initial G n > 6.5 x l(r (unmilled)
were vulcanizable. Tensile strength increased with G n and leveled
MECHANICAL PROPERTIES OF POLYMERS 73

X SHARP FRACTIONS
0 BLENDS O F FRACTIONS 12 AND 5
A BLENDS OF FRACTIONS I2 AND I

50 -
40 -
0
030
\
-
-
Downloaded by [Princeton University] at 06:30 03 June 2013

ui 20
I
10 -

0 I I I
0 I .o 2.0 3.0
M, x

Fig. 9. Tensile strength vs. number-average molecular weights, polystyrene


blends [ 24 1.
RADIATION DOSE, R (megarads)
200 150 100 75 50 40 30 2 0 10 0
I 1

.-
1

d-+-W-+
I-
tn
W

W
I-

0
10000 12500 15000 17500 20000

Fig. 10. Tensile strength vs. viscometric-average molecular weight for


polycarbonate bars [loo].
74 MARTIN, JOHNSON, AND COOPER

RADIATION DOSE, R ( megads)

10 8z
-
2
7.1c
Downloaded by [Princeton University] at 06:30 03 June 2013

Q)

2.5

0
0 1000 10000 11000 20000 21000
MV

Fig. 11. Flexural properties vs. viscometric-average molecdar weight for


polycarbonate bars 1100 I.

off at fin = 3 x l o 6 (unmilled). This trend was also observed in un-


filled natural rubber [ 1381 and SBR vulcanizates [ 161 but the effect
was reduced when the cure system was adjusted to give a constant
equilibrium modulus [16]. When Yank0 varied sulfur levels to obtain
optimum cure states, tensile strength increased with Gn and did not
level off until G n > 5 x los (unmilled). It was noted that the five
fractions with the highest G n values all had tensile strengths greater
than that observed in the unfractionated polymer. Johnson came to a
similar conclusion in another SBR tire-tread study [138]. A statisti-
cal study of SBR production samples has shown that MWD variations
can affect tensile strength [160] although an earlier report by the
same group did not indicate systematic changes due to M W D [141].
Cerny [46] studied the effects of MW on the processing and me-
chanical properties of a 29/71 SBR (Kralex). Generally, process
variables were quite dependent on MW but the tensile strength of the
vulcanizates also showed a marked dependence on MW for MW
< -1.5 x lo5 (Fig. 14). At high cross-link densities the tensile
MECHANICAL PROPERTIES OF POLYMERS 75

o 30-MIN. CURE
Downloaded by [Princeton University] at 06:30 03 June 2013

60-MIN. CURE

M / 100,000
Fig. 12. Tensile strength of pure-gum vulcanizates of low-unsaturation frac-
tions vs. molecular weight [82].

strength was unaffected by MW over a range of 6.4 x 104 to 2.46 x 105.


However, major differences were apparent when the cross-link den-
sity, u, was less than -2.5 x 10" mole cm3 (Fig. 15).
Temperature performance was a major parameter in a study [254]
which determined the effect of MW on the tensile strength and elonga-
tion of filled and unfilled polyisoprene elastomers. MW effects were
small a t 20°C. However, a t 100°C the tensile strength of the filled
specimens increased by a factor of 7-9 when nv
of the initial poly-
mer was raised from 3.5 x lo5 to 3.0 x loe (Table 6). Similar results
were obtained on the unfilled specimens. Interpretation of this work
is difficult because the stereoregularity of the material varied with
MW, and curing conditions were not discussed.
76 MARTIN, JOHNSON, AND COOPER

5000 I I 1 I
Y I 1 1

-
4000 -
c5

.-
2
Y

3000 -
I
Downloaded by [Princeton University] at 06:30 03 June 2013

I-
c3
z
-
w
0:
2000 -
w
1
u)
-
z
w
I- 1000-

105(M + M,, I
Fig. 13. Tensile strength of low-unsaturation fractions and mixtures after
pure-gum cure vs. 1 / ( M + Mc) o r l / ( m n + Me) [82]: (0)fractions,
(a) mixtures of fractions, (a) unfractionated Polymer 11-A.

9. Slightly Cryetalline Polymers. The results discussed in this


section a r e based on polymers which are amorphous but which may
have some small degree of long-range molecular order. Typical
materials include cellulose derivatives where the substituent groups
have disrupted most of the cellulose structure and paracrystalline
fibers made from an acrylonitrile copolymer, Few of the papers
discussed sample morphology, SO the “slightly crystalline” assign-
ment should be regarded as somewhat arbitrary.
A number of papers which were published in the 1930s discussed
the effects of MW and MWD on tensile strength. For example, Doug-
las and co-workers [56,62]found a substantial decrease in the tensile
MECHANICAL PROPERTIES OF POLYMERS 77

sop00
N
-

I
I-
z
(3
w
e
a z
30 - 300 Pa
l-
(3
Downloaded by [Princeton University] at 06:30 03 June 2013

w z
d 9w
v)
z
:20 - 200

10 100 I I

Fig. 14. Tensile strength (Curve 1) and elongation (Curve 2) vs. molecular
weight of Kralex [46].

strength of vinyl chloride-acetate copolymers when the molecular


weight fell below lo4. Spurlin [261] used fractionated cellulose nitrate
to show that both low MW and broad M W D adversely affect tensile
strength. He compared his data to results obtained by earlier investi-
gators and concluded that low fractionation efficiency and insufficient
sample replication were major problems in most M W D studies. In
one of the more widely cited early studies, Sookne and Harris [258]
fractionated cellulose acetate and characterized the fractions by
intrinsic viscosity and dynamic osmometry . The tensile strength,
elongation, and fold endurance of the fractions and some blends were
then evaluated. Tensile strengths of the fractions increased with DPn
and intrinsic viscosity over the whole range studied. Blends were
then prepared, and the results are shown in Fig. 16. The author’s
interpretation was that the blends followed the number-average D P
rather than the weight-average. However, the data appear to be con-
sistent with the behavior that would be expected when both B n and
Downloaded by [Princeton University] at 06:30 03 June 2013

IOL I I I 1 1
0 I 2 3 4 S
v x i04mo~-cm-3
Fig. 15. Tensile strength of Kralex vs. cross-link density at various molecu-
lar weights [46]. Molecular weights: (0) = 6.4 x lo4, (0) = 1.03 x lo5,
( A ) = 1.75 x lo5, (a)= 2.46 x lo5.

UWa r e below their respective limits for maximum properties. At


constant aW, upper graph, a narrow MWD favors better properties
because a n is approaching its limit for maximum properties. Poly-
disperse samples should fall below the fraction curve a s occurs on
the graph. Similarly, a t constant an,Jower graph, broad distribu-
tions give better properties because Mw is approaching its limit. A
few blends lie below the line, but the majority lie above it; further-
more, the points farthest above the line have the widest MWD. An-
other cellulose acetate study [257] indicated that the tensile strength
loss in wet or knotted fibers is increased when the average MW is
decreased o r when the content of low MW fractions is increased.
Scherer and Rouse [249] prepared 12 fractions of cellulose nitrate
@pw = 30 to 550) and examined mechanical properties as a function
of MW and the shape of the MWD. Tests conducted on solvent cast
films indicated that tensile strength was independent of MW and MWD
when DPw > 100-120. Below this level, tensile strength was a func-
tion of both parameters. In order to correlate physical properties
MECHANICAL PROPERTIES OF POLYMERS 79

with both MW and MWD, an empirical shape factor was defined:

mmax
Shape factor =
log H
~

where

av width of peak
log H = log
max height of peak 1
Increasing the MW o r narrowing the MWD gives higher shape factors
Downloaded by [Princeton University] at 06:30 03 June 2013

and better properties. The tensile strength vs. shape factor curve
was found to level off at SF > 30. A similar series of tests on ethyl
cellulose was also reported [248]. The results showed tensile strength
increasing with D P before leveling off at mw-300. The authors
noted that the mechanical properties of the fractions were better
than those of the original sample except for one fraction with a very
low DP. In addition, the presence of 17-25% of material with D P <
100 was also reported to improve the physical properties of blends.
This was assumed to be due to some plasticizing effect.
The MW and MWD dependence of PVC polymers has been rather
widely studied. For example, four PVC samples which had identical
values of [q]but different M W D were compared by Kaminska [145].
He found that tensile strength increased as the MWD was narrowed.
Furthermore, this trend was linear when MWD, expressed in terms
of Scherer’s differential shape [249], was plotted against tensile
strength.
An attempt to modify the properties of PVC by adding vinyl chlo-
ride telomers (DP = 5.5-6.0) was made [246]. Small amounts of the
telomer (20-25%) increased tensile strength by 10-20%. This is not
unexpected because small amounts of conventional plasticizers
(5-10%) can greatly increase the crystallization rate in PVC (page
254 of Ref. 204). Further increases in the telomer concentration
lowered tensile strength.
Pezzin and Zinelli [221] fractionated 3000g of PVC and charac-
terized the fractions by light scattering and osmometry. They found
that [q] of the unplasticized PVC decreased -20% during the milling
and molding operations, s o most of the testing was carried out on
samples which were plasticized with 60 parts of dibutyl phthalate per
100 parts PVC to decrease degradation. A s shown in Fig. 17, the
tensile strength of compression-molded, plasticized PVC increased
-
strongly with f i w and showed no signs of leveling off below GW 2
x lo5. Fractions, blends, and whole polymers all fell on the same
Downloaded by [Princeton University] at 06:30 03 June 2013

Table 6
Comparison of P r o p e r t i e s of SKI-3 and SKI-L Vulcanizates
Unfilled and Reinforced with 30 P a r t s of Channel Black [254]

Characteristics of r u b b e r P r o p e r t i e s of vulcanizates
Tensile strength Elongation a t Tension set
Plasticity Content Glass-
(GOST) of cis-1.4 transition ( kgf/cm2) break (%) (%)
[qlb M W X 10- 415-53) groups ( % ) temp ("C) a t 20°C a t 100°C a t 20°C a t 100°C a t 20°C at 100°C

Unfilled Vulcanizates
SKI- L
2.24 365 0.71 75.0 - 321 32 1210 630 14 8
4.11 835 0.50 75.0 - 326 49 1100 570 8 4
5.77 1325 0.38 81.5 - 3 20 65 1070 720 12 2
8.09 2120 0.17 93.5 - 306 167 1000 1160 10 8
8.98 2440 0.28 92.5 - 292 153 1090 1240 10 12
3.37 645 0.45 97.0 - 235 41 1100 645 8 6
10.3 2950 0.14 92.5 - 303 145 960 1140 8 14
10.3= 2950 0.14 92.5 - 240 32 870 550 6 4
Downloaded by [Princeton University] at 06:30 03 June 2013

SKI-3
3.55 - 0.38 97.0 - 308 108 840 975 8 15
6.0 - 0.4 96.0 - 316 250 850 1095 9 16
Filled Vulcanizates I

SKI-L
2.26 350 0.68 86.0 -63 309 20 1030 360 12 4
2.92 520 0.66 88.0 -64 316 31 1015 450 12 4
6.01 1400 0.3 92.5 -64 320 47 1100 520 8 2
7.51 1950 0.17 93.5 -67 306 167 1000 1160 10 .8
7.64 1960 - 88.0 -65 282 191 800 900 20 62
10.3 2950 0.14 92.5 -69 323 146 760 920 30 48
SKI-3
3.55 - 0.38 97.0 -68 371 201 810 840 28 20
6.0 - 0.4 96.0 -73 378 230 810 840 30 30

'The rubber was mixed with the ingredients at the solution stage, dried in a vacuum at 60°C, and vulcanized.
bToluene, 30°C.
82 MARTIN, JOHNSON, AND COOPER
Downloaded by [Princeton University] at 06:30 03 June 2013

INTRINSIC VISCOSITY

“,leoo- 0 0
1

d
s.
/
0
I
L 000
z
W 502+204
0
!0:
- 502+127
0
Q) 0 502+43
9 502+92
- 400
f j
Q)
@ 502+l60
0 331+127
2 d STARTING
W
I- MATERIAL
1 1 I
0‘ 200 400 600
NUMBER-AVERAQE DP
Fig. 16. Tensile strength of blends of cellulose acetate a s a function of intrin-
sic viscosity (upper graph) and mn(lower graph) [258].
curve so, within the limits of the test, MWD effects were not observed.
The Q of these samples was less than 3.0 except for two blends where
Q =7.0 and 13.6.
Unlike Pezzin’s results on plasticized PVC, Wallach [291] found
that the tensile strength of unoriented polyimide films levels off at
high MW (Gn 4 x l(r and & 1.5 X lo5). With some scatter in
N N

the data, tensile strength was found to be linear when plotted against
l/Gn. These results are shown in Figs. 18-21. A small M W D effect
is apparent when comparisons are made at constant [TJ] (Fig. 20).
Crystallinity of the films ranged from 2 to 5%.
MECHANICAL PROPERTIES OF POLYMERS 83
Downloaded by [Princeton University] at 06:30 03 June 2013

L
D 150 i

Fig. 17. Tensile strength, uB, vs. Ti?w of PVC fractions ( O ) , blends ( x ) , and
whole polymers (0)(2211.

A study of fractionated polyvinyl acetate [247] using cast films


also showed an asymptotic approach to maximum values at higher
MW for tensile strength. Blending these fractions allowed the evalu-
ation of a shape factor, dependent upon the maximum DPw of the
distribution curve, the height of the maximum, and a skew factor. The
mechanical properties related to this shape factor were tensile
strength, fold endurance, and elongation. Changing the three variables
but keeping the shape factor constant resulted in identical properties.
The larger the shape factor, the better were the physical properties.
Others have found that increases in the MW of polyvinyl acetate
substantially increase the tensile strength of oriented films [ 14,1641.
84 MARTIN, JOHNSON, AND COOPER

- 200 - B
A TENSILE STRENGTH ELONGATION
-
-z
(D
n
24 -
0 0
E 0
t-
Downloaded by [Princeton University] at 06:30 03 June 2013

I" IMPACT STRENTH

Fig. 18. Polyimide mechanical properties vs. precursor weight-average


molecular weight [291].

This is an indirect effect because only minor changes were observed


in unoriented samples.
In a sense, tensile strength measurements a r e a special case of
creep rupture tests. Takaku and Kishi [265] noted this in their ten-
sile creep measurement of acrylonitrile methylacrylate copolymers
(91/9 composition by weight) which ranged from EW= 7.9 x lo4 to
1.52 x lo5, When tensile strength was plotted against log time, all
samples had a similar shape. However, the shift factor was a func-
tion of MW so, at a given temperature, tensile strength increased
with MW if the measurements were made in either the transition or
rubbery regions.
4. Highly Crystalline Polymers. In an early study [45], Carothers
reported that the tensile strength of polyester fractions made from
MECHANICAL PROPERTIES OF POLYMERS 85

IB ELONGATION

P
In
a
t
In
Downloaded by [Princeton University] at 06:30 03 June 2013

20,000 Mn 60,000

D IMPACT
STRENGTH
t

IZ'
5

1,
6:
2

I-

Flg. 19. Polyimide mechanical properties vs. p r ecursor number-average


molecular weight [291].

w-hydroxydecanoic acid increased sharply when the MW was raised


from 9,300 to 16,900. This trend was also noted in tests conducted
on whole polymers of low-density polyethylene [Chapter 4 of Ref. 23,
and Refs. 199,235,253,2591.A check of other factors revealed that
crystallinity had no effect on the tensile strength while long-chain

-
branching caused a reduction in the property. In fractionagd samples,
tensile strength increased with MW before leveling off at Mw 3 x
a6 MARTIN, JOHNSON, AND COOPER

A TENSILE STRENGTH
1 200
0 E LO NGAT I0,N
3.9
Downloaded by [Princeton University] at 06:30 03 June 2013

c TEAR STRENGTH D IMPACT STRENGTH

.-- 24 -
E

I- 4-
0
W d
I- 4
2-

I 2 3 4

Fig. 20. Polyimide mechanical properties vs. precursor intrinsic viscosity


[291].

lo5 [253]. Narrowing the MWD was also reported to increase the
tensile strength when the measurements were made at constant melt
index [57].
Lawton [ 1701 measured a variety of mechanical properties on
seven commercial polyethylenes as a function of y-radiation dose,
temperature, and initial MW. Irradiation at low levels gave the same
effect as increasing MW (higher tensile strength and elongation).
However, at high dose levels, the increased cross-linking made ini-
tial MW unimportant. These tests included measurements above the
crystalline melting point, where, at constant dose level, tensile
strength increased with MW, Limited data on two sulfur vulcanized
ethylene-butadiene copolymers also indicate that tensile strength
MECHANICAL PROPERTIES OF POLYMERS 87
Downloaded by [Princeton University] at 06:30 03 June 2013

0.2 0.4 0.6 0.0 I


I/Mn x
Flg. 21. Polyimide tensile strength vs. reciprocal of number-average molecu-
lar weight [291].

and elongation increase with initial MW at room temperature and


above the melting point [ 121. The MW effect was less important after
thermal aging.
In the Williamson study described in Section 11, the ultimate tensile
strength of unfractionated, high-density polyethylene increased sharply
with [Q] [298]. Increasing density apparently also raised the tensile
strength (Fig. 22). Since fractionation resulted in further improve-
ment, one can conclude that either narrowing the MWD o r reducing
the density caused the property changes. This latter possibility
conflicts with the behavior observed on the whole polymers but it
does conform with the results obtained on the higher MW fractions
where samples with lower crystallinity had higher tensile strength
at a given [Q]. Other effects could also be important. F o r example,
fractionation efficiency was relatively low in the high MW samples.
I
I 0.932 I I

.-
a
Y)
6000 - A (0.945
0.932

\ I
I
5000
z
w
0:
I-
v)

- 0.959 '
w 3506 I I \ 0979')) I
Downloaded by [Princeton University] at 06:30 03 June 2013

2
v)
z
w
I-
w
4000
P''
0.941
r l
; r
j

\s3j " I
5 ' I
3000 0.9571.4 1
5
3

0 2 4 6 8
I N T R I N S I C VISCOSITY, d t / g

Fig. 22. Relationships between ultimate tensile strength and intrinsic viscos-
ity. Symbols as in Fig. 3; numerals on curves a r e densities [298].

Figure 22 indicates that, at high [q],tensile strength decreases a s


MW increased. This may be a MW effect but changes in morphology
and increases in the MWD and residual stresses at the higher MW
were probably responsible because most other studies, such a s Ref.
146, do not show this trend (Table 7). The data in Table 7 also indi-
cate that tensile strength over a broad range of temperatures is
improved when the MWD is reduced. No MWD effect was noted [283]
when comparisons were based on GWor zv.
Broadening the MWD
improved tensile strength a t constant Gn however.
A M W D dependence was also demonstrated [273]for high-density
polyethylene. As in Ref. 298, fractions exhibited higher tensile
strengths than whole polymers which had the same [q]. Tensile
strength increased with MW for both fractions and whole polymers
but the rate of change was much greater in the fractions.
MECHANICAL PROPERTIES OF POLYMERS 89

In contrast to the above results, MWD effects were not substantial


[36,113] where comparisons were made at constant relative specific
viscosity and constant melt index, respectively. However, tensile
strength was found to increase by 50-75% when melt index was de-
creased from 1.0 to 0.10 [113]. Similar trends have been observed
in Refs. 148 and 283.
Still other workers [ 1541 found that, for high-density polyethylene
fractions, tensile strength increased with GVup to 82,000 and then
decreased slightly (Table 8). Limited results from tests on bimodally
distributed polyethylenes indicate that tensile strength is improved
when MW is raised and MWD is reduced. The addition of low MW
processing waxes impairs tensile strength [ 148,2171 although the
Downloaded by [Princeton University] at 06:30 03 June 2013

change was considered to be small for additive levels < 3% [217].


Extraction of low MW atactic material with n-heptane was re-
ported to increase the tensile strength of polypropylene in one study
1801. However, the opposite result was obtained in other investiga-
tions “72,2521. Little dependence on atactic content was noted on
injection- and compression-molded samples [255]. At constant [TI,
higher tensile strength has been associated with narrow MWD on
compression-molded films [244]. This conclusion was reached in a
comparison of polypropylene fractions with blends which had a Q of
2 and 10 (Fig. 23). The figure also shows a substantial decrease in
tensile strength when [q] is less than -1.5 dl/g (decalin, 135°C).
Kamide and co-workers [144] measured the effect of MW and MWD
on the mechanical properties of undrawn isotactic polypropylene
filaments. Crystallinity and orientation were held constant while the
properties were measured on fractions, blends, and whole polymers.
Tensile strength and energy to break increased with MW. Unlike the
results in Ref. 244, however, MWD effects were small.
In oriented specimens of isotactic polypropylene, a reduction in
the MWD reportedly gives higher tenacity at the maximum draw
ratio [80]. Increasing a n [202] and & [39] also improves this
property.
The data in Fig. 6 indicate that tenacity is dependent on Mw a t
constant draw ratio. However, this trend is largely due to orienta-
tion. A s shown in Figs. 24 and 25, tenacity is highly dependent on
molecular and crystallite orientation [256]. MW and MWD were
found to be less important while the per cent crystallinity was shown
to be insignificant,
Ross [238] observed that it was possible to obtain polypropylene
fibers of different MW and MWD by varying the extrusion tempera-
ture to change the amount of degradation. He concluded that tenacity
increases as the MWD is narrowed if draw ratio and GWa r e held
Downloaded by [Princeton University] at 06:30 03 June 2013

W
0

Table 7
Relationships Between Stress-Strain Properties, MW Parameters, and Temperature
in High-Density Polyethylene [146]

~~~~ ~ ~~ ~

CHS C=C
- - - - Percent of Melt per 100 per1000
Fraction [TI' MV MW Mn Mw fraction with
= index Density carbon carbon
No. (dl/g) x x x Mn mol wt > lo6 (g/lO min) (g/cc) atoms atoms

1 1.2 50 60 16 3.7 0 50 0.956 0.3 0.7


2 2.2 114 113 61 1.8 0 0.08 0.954 0.05 0.1
3 1.85 90 97 18 5.4 0 6.6 0.951 0.3 0.6
4 1.95 97 110 19 5.8 0 4.2 0.951 0.3 0.6
5 3.5 210 218 96 2.3 0 0.01 0.950 0.05 0.1
6 2.8 153 154 13 12 3 3.2 0.950 0.35 0.7
7 2.85 160 190 19 10 6 1.6 0.950 0.15 0.6
8 3.15 185 255 - - 7 1.5 0.955 0.2 0.7
9 3.6 220 260 16 16 11 0.6 0.950 0.2 0.65
10 ,15
- -1500 1500 220 7 45 Nonflowing 0.937 <0.1 0.05
D
z
0
Downloaded by [Princeton University] at 06:30 03 June 2013

Yield strength (kg/cm*) Tensile strength (kg/cm2) Ultimate elongation (%)


Frac- Frac- Frac-
Temperature ("C) Temperature ("C) Temperature ("C)
tion tion tion
NO, -40 -20 0 20 40 60 80 100 NO. 4 0 -20 0 20 40 60 80 100 NO. -40 -20 0 20 40 60 80 100

1 - - - 292 205 127 85 48 1 473 388 353 142 109 79 61 33 1 <5 <I0 <I5 <90 1 140b 230Ob 25Wb 980b
270 1120 570 41 0
2 420 380 336 276 200 151 104 62 2 443 410 363 336 230 100 52 52 2 470 510 580 840 1900 3300 2800 970
3 442 369 312 247 174 127 75 46 3 - - 192 189 153 125 72 44 3 40 50 140 550 1150 2200 1950 1280
4 447 376 310 243 171 113 63 5 1 4 - - - 180 189 110 69 46 4 40 40 110 640 1400 2350 2110 1320
5 410 360 310 266 193 152 102 60 5 580 500 486 365 388 220 78 52 5 500 500 610 690 960 2100 2900 1400
6 445 377 326 258 186 122 81 47 6 - - - 183 151 155 102 53 6 30 35 75 780 870 1700 1970 1750
7 430 382 315 240 164 120 75 SO 7 - 230 195 180 161 185 114 65 7 70 90 95 660 950 1600 1790 1660
8 422 378 327 275 178 136 86 48 8 - 240 - 199 160 172 108 57 8 80 85 190 710 980 1750 2000 2050
9 424 380 327 257 188 126 78 SO 9 300 243 230 206 252 215 141 69 9 100 120 230 790 1150 1540 1730 21 70
10 340 299 265 228 168 133 89 56 10 740 740 634 497 400 378 313 219 10 340 420 440 420 540 650 820 1300

aa-Chloronaphthalene, 135°C.
bElongation calculation based on the drawn portion of the sample (less than 25 mm).
Downloaded by [Princeton University] at 06:30 03 June 2013

Table 8
Physical Properties of High-Density Polyethylene Fractions [154 ]

Elongation
Frac- Tensile (%I Impact 25°C
tion Mv X Density strength ( p s i ) (yield/ Melt strength ( p s i ) Modulus
No. '7SP lo-' (g/cc ) (yieldhreak) break) index in. of notch (psi1

1 0.019
2 0.027
3 0.034
4 0.056 1.5 0.969 - /1283 -/1.8 - - 235,000
5 0.064 1.7 0.966 843/1225 2/2 - - 245,000
7 0.101 3.0 0.956 4009/2194 10/287 22 0.65 170,000
8 0.130 4.0 0.953 3766/2533 11/2000 2.8 2.64 155,000
9 0.173 5.6 0.951 3672/4199 l0/2000 1.25 13.71 140,000
10 0.244 8.2 0.950 3411/5198 14/815 0.25 16.07 98,000
11 0.412 14.5 0.949 3244 /4241 17/350 0.60 16.08 95,000
12 0.641 23.0 0.949 3131 /4496 15/296 Too hard No break 100,000
MECHANICAL PROPERTIES OF POLYMERS 93

4t
Downloaded by [Princeton University] at 06:30 03 June 2013

I
c
0
z
w 2-
a
c
rn
0

'L
00
I
I
2 3
I

I N T R I N S I C VISCOSITY ( 7 ) )
1
4

Fig. 23. Relation between intrinsic viscosity and tensile strength of


polypropylene fractions and mixtures [244 I: ( 0 )fraction, ( 0 )mixture
Mw/Mn = 2, ( 0 )mixture Mw/Mn = 10.

constant. When crystal structure and M W D were held constant,


tensile strength increased with MW.
Relationships between melt index and the tensile strength of
ethylene copolymers [84]and propylene copolymers [ 133,1881 are
also available. As illustrated in Fig. 26, the higher M W polypropylene
copolymer is stronger than the lower MW material over a wide range
of densities.
Relatively few studies have been published concerning MW effects
on the tensile strength of nonolefinic polymers, For example, little
specific data is available on polytetrafluoroethylene, PTFE, although
one review paper noted that increases in the MW of PTFE can raise
the tensile strength by 25% [267]. There is also an indirect effect
because MW influences crystallization rates.
The tensile strength of polyacrylonitrile fibers was related to
both MW and crystallite orientation [156]. Comparisons made at
94 MARTIN, JOHNSON, AND COOPER

I2 -

II -

10 -

9-
Downloaded by [Princeton University] at 06:30 03 June 2013

8-

7-

6 -

a
z 5 -
w
c
4-

3 c
0 1 1
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.e 0.9 1.0

MOLECULAR ORIENTATION
Fig. 24. Relationship between molecular orientation and tenacity of polypropyl-
ene monofilaments of different molecular weights (2561.
MECHANICAL PROPERTIES OF POLYMERS 95

10 -

9-

-4
C
8-

7-
-
\
0
6-
Downloaded by [Princeton University] at 06:30 03 June 2013

k I-
0
U
2 4-
W
I-
3-

2-

I-

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0


CRYSTALLITE ORIENTATION FACTOR
Fig. 25. Relationship between Herman’s Crystallite Orientation Factor and
the tenacity of polypropylene monofilaments 12561.
constant draw ratio did not always show the molecular weight de-
pendence because, at constant draw ratio, crystallite orientation is
a function of MW.
Nagano [200]measured the tensile strength and elongation of poly-
vinyl alcohol films cast from a water solution. The five samples
used in this study had bimodal MWD (Fig. 27). As shown in Fig. 28,
tensile strength increased linearly with shape factor. This empirical
quantity was originally proposed by Scheres [249] and is defined in
Section 111-B-3.Basically, an increase in MW o r a decrease in MWD
raises the shape factor. In contrast to Nagano’s results on unoriented
films, the tenacity of polyvinyl alcohol fibers was not strongly affected
by MW characteristics when measurements were made at constant
draw ratio [263].
The mechanical properties of several commercial and experi-
mental poly(oxymethy1ene) diacetates were measured in Ref. 106.
The data indicate that a broad MWD impairs tensile strength because
96 MARTIN, JOHNSON, AND COOPER

DENSITY ( g / c c )

Flg. 26. Density of copolymers vs. the tensile strength a t the fail point for
two levels of average melt flow 11331.
Downloaded by [Princeton University] at 06:30 03 June 2013

the sample with Q = 15.1 was less than half as strong a s the narrower
distribution samples (Table 9). The authors also correlated this
property with Gn. Flex strength was said to increase with Ewbut
an examination of the data in Table 9 shows that OWhad no substantial
effect on this property over the range of values tested. The one
sample with an extremely broad M W D did exhibit a sharp decrease
in flex strength however. One might also have interpreted this as an
effect of E n .

3
I.o
z
0
I-
2 0 000 2 00 3000 0 1000 2000 3000 0 1000 2000 3000

1
G
a DP DP DP

;k
i"h
3.0 No.5 6p=1305

I.o

I ..
0 1000 2000 3000 0 1000 2000 3000
DP DP
Fig. 27. Differential molecular weight distributions of polyvinyl alcohol films
[ZOO].
MECHANICAL PROPERTIES OF POLYMERS 97

-
Al
E
8-

-
\
:
0

7 -
I
I-
W
Z
Downloaded by [Princeton University] at 06:30 03 June 2013

W
a 6-
c
u)

W
4
5-
2
W
c

4-

200 300 400


SHAPE FACTOR

Fig. 28. Tensile strength vs. shape factor of polyvinyl alcohol films [ 2 0 0 ] .

Relationships between crystalline content, MW and mechanical


properties of 610 and 66 nylons have been investigated 12621. The
crystallinity of the compression-molded films (Gn = 12,200 and
23,000 f o r the 610 material) was varied by thermal treatments. A s
shown in Fig. 29, the tensile strength of 610 nylon is a function of
per cent crystallinity and independent of MW over this range of G n .
Similar results were obtained on the 66 samples.
A number of other polyamide studies concluded that MW does
affect tensile strength. For example, Hamilton and Epstein [110]
investigated a terpolymer system of nylons 66/610/piperazine 10.
The tensile strength of this flexible nylon increased linearly from
5000 to 9000 psi when Gv was raised from 4,600 to 16,200. Pipera-
zine 10 is a secondary amine so it has no hydrogens available for
bonding. This, along with the fact that the components of the terpoly-
98 MARTIN, JOHNSON, AND COOPER

Table 9
Relation Between Mechanical Properties and
MW Parameters in Poly(oxymethy1ene) Diacetates [lo61

1.39 2.12 2.17 2.40 2.41


5.77 10.35 10.70 15.87 14.26
3.82 9.20 7.10 2.30 0.94
1.51 1.13 1.50 6.9 15.1
Tensile strength 664 750 710 570 267
( kp/cm )
Elongation (%) 12 25 15 6 3
Flexural strength 1162 1100 1262 1275 7 04
Downloaded by [Princeton University] at 06:30 03 June 2013

(kp/cm2)
Flexural modulus 22000 22000 27000 34000 50000
(kp/cm 2,
Ball p r e s s u r e hardness 1894 1950 1960 2324 2284
HB 2/5 after 60 sec
( kp/cm )
Impact strength, notched 8 15 4.3 1.2 1.3
bar (cm kp/cm2)
Shock resistance
(cm kp/cm2)
+ 20°C > 90 > 90 90 40 11
-4 0°C > 90 > 90 > 90 - -

a Tetrachloroethane/phenol, 3/1 by weight plus 2% a-pinene; 90°C.


b Calculated from fractionation data.

mer were present in almost equal proportions (35/30/35), reduced


the per cent crystallinity and resulted in an unusually flexible poly-
amide.
Prevorsek [226] found that the tenacity of fully drawn nylon 6 and
polyethylene terphthalate fibers increased monotonically with MW.
However, when the fibers were thermally treated to alter the degree
of chain folding, maximums were observed in the tenacity-MW curves.
The results were interpreted in terms of the morphology of these
highly oriented fibers. Zhurkov and Abasov [ 3041 noted that the
strength of both oriented and unoriented Capron fibers increased
with MW. They correlated their results and the data of several other
papers in terms of a fluctuation mechanism of chemical bond rupture.
At constant time and temperature, the tenacity of both the oriented
and unoriented fibers was inversely proportional to MW. This work
is extensively critiqued by Bartenev and Zuyev in Chapter 4 of their
book [ 171.
MECHANICAL PROPERTIES OF POLYMERS 99

SAMPLE A DRY ( M n = 12 200)


SAMPLE A C O N D I T I O N E ~TO 5001~ R.H.
Downloaded by [Princeton University] at 06:30 03 June 2013

SAMPLE A CONDITIONED TO 1 0 0 % R . H .
SAMPLE B DRY (&ln=23,000)
FRESHLY MOLDED SPECIMEN DRY

I I I I I
10 20 30 40
PER CENT CRYSTALLINITY
Fig. 29. Tensile strength vs. crystallinity of 610 nylon (2621. The crys-
tallinity axis was rescaled in a later paper [262a].
Cumberbirch and Harland [55] related the tenacity of regenerated
cellulose to D P and orientation. Filaments of cellulose acetate frac-
tions (DP = 100-800) were spun, washed, stretched in boiling water,
saponified, and again washed. When tested, it was found that D P had
little effect on the tenacity of unoriented filaments except at the
lower MW (Fig. 30). However, this property did increase significantly
with D P in the oriented filaments. Part of the increase was attributed
to the fact that MW influences orientation. When tenacity was plotted
against D P at constant orientation the curves leveled off above D P
400 (Fig. 31). Limited tests on cellulose acetate filaments gave re-
-
sults which were similar to those obtained on the cellulose specimens.
Schieber [250], Tachikawa [264], KrPssig [161], and Krassig and
Kitchen [162] also studied the tensile properties of cellulose fibers.
In the latter two papers it was observed that tensile strength does
not drop off linearly with l/DP. To account for this, a model was
proposed where tensile strength was expressed in terms of DP, DP,,
orientation, and crystallinity. DP, is the effective length of the mor-
phological moieties and is expressed in terms of glucose units. Fig-
ure 32 illustrates how DP, was determined. At constant DP and
crystallinity the tensile strength increased and elongation decreased
with increasing orientation for both wet and dry samples. After re-
lating tensile strength to orientation and per cent crystallinity,
100 MARTIN, JOHNSON, AND COOPER

-
3.00

0
t
\
c?
>-200-
t ' / / 100% 1
0
a
Z
w
Downloaded by [Princeton University] at 06:30 03 June 2013

I-
>
a
a

I "0°-
" t
Effective
Y

01 1 I I
Stretch

1
I
0 200 400 600 800

Fig. 30. Relation between DP and dry tenacity at various levels of effective
stretch [55].

hydrolysis experiments were performed to alter the DP. Since this


also affected orientation and crystallinity, the data had to be cor-
rected using the relations previously developed. Figure 33 illustrates
the experimental (dark circles) and corrected (outlined circles)
results. The dashed lines parallel to the initial slope allowed the
determination of DP, at any DP. Figure 34 shows the plot from which
a relationship for dry tensile strength was obtained, This relation-
ship was used to calculate DP, for commercial rayon samples. This
work has thus accounted for the tensile strength of cellulose fibers
in terms of MW. The M W were determined from a viscosity rela-
tionship which was based on osmotic pressure measurements. No
corrections were made for variations in MWD.
A number of other studies have related the physical properties of
cellulose and its derivatives to MW and MWD. Mark [185]and Wake-
ham [290] summarize much of the early work. Krissig and Kitchen
[162] also tabulate many of the studies made before 1960.
MECHANICAL PROPERTIES OF POLYMERS 101
Downloaded by [Princeton University] at 06:30 03 June 2013

I I I I
0 200 400 600 800
-DP-
Fig. 31. Relation between DP and tenacity at constant birefringence (551.

TS Opt.
at DP=-

L
b c f (L.0.)

\. -.---_
1000,

IOOO/DPL

Fig. 32. Theoretical relationship between lOOO/DP and tensile strength. The
slope h of the solid straight line i s assumed to be related to the perfection of
order and the degree of orientation, i.e., the over-all s t a t e of lateral o r d e r
11621.
102 MARTIN, JOHNSON. AND COOPER

1s T i r e Rayon
Downloaded by [Princeton University] at 06:30 03 June 2013

Fig. 33. Relationship between tensile strength of conditioned yarns and


l O O O / D P . The D P was changed by acid hydrolysis [162]: ( 0 )Observed values
of TScond, (0) TScond corrected for the improvement in fine structure during
hydrolysis.

.
C Elongation
1. Amorphous Polymers below Tg. Many of the investigations
tabulated in Section 111-B included measurement of both tensile
strength and elongation. In evaluating the results obtained on amor-
phous polymers below Tg, it is evident that most of the tensile
strength conclusions also apply to the elongation data. F o r example,
Merz et al. [ 1861 found that certain minimum levels of both E n and
3~existed in polystyrene. Above these minima, MWD had no effect
on either tensile strength o r elongation. McCormick et al. [ 1891 also
reported the same trends in these two properties (compare Figs. 7A
and 35A), but he concluded that MWD effects were significant even
-
when E n 1 to 1.5 x lo5. Using Merz’s minimum values of G n and
Ew,one can reinterpret this elongation data as illustrated in Fig.
35B. The basis for the reinterpretation is explained in Section III-
B-1. Both approaches appear to give a reasonable correlation with
the MW parameters.
Thomas and co-workers [109,266] showed that, at constant EW
(1.9 x the rupture elongation of injection-molded polystyrene
sheets is increased by narrowing the MWD. For example, tensile
specimens made from the narrow distribution material (Q = 1.06)
went through a yield point and failed at 5.9% elongation while the
MECHANICAL PROPERTIES OF POLYMERS 103
Downloaded by [Princeton University] at 06:30 03 June 2013

0.4 0.8 I .2 1.6 C r l x f ~ X 1 0 0 0(I/DPL-I/DP)

1000
Flg. 34. Relationship between the p a r a m e t e r C r I X f; X (DP
the observed tensile strength of a conditioned yarn from the hydrolysis
-d1000
and

s e r i e s . CrI i s the crystallinity index and fr is the orientation factor. F ,


Fortisan; M, Meryl; HM, high wet modulus; T r , t i r e yarn; C, cotton; Tx,
textile rayon [I62 1.

broad distribution sample (Q = 2.6) exhibited brittle failure at 3.4%


elongation. The improved stress-strain properties of the narrow
distribution material were attributed to the formation of tiny c r a z e s
in the material when it was strained.
Elongation-MW data for glassy polymers other than polystyrene
a r e rather sparse. However , the similarity between tensile strength
behavior and elongation characteristics i s probably common to all
glassy polymers. The similarity was shown for irradiated PMMA
11551 and irradiated polycarbonate [loo] (Fig. 11). Neither of these
studies examined MWD effects. Elongation in poly(norbornene) in-
creases as [q]and the ratio of trans to c i s unsaturation a r e increased
[93]. The relative importance of these two variables was not exam-
ined in this paper.
2. Amorphous Polymers above Tg. Many elastomeric materials
crystallize at high strains. Flory, in his work on butyl rubber vul-
104 MARTIN, JOHNSON, AND COOPER

3-
0 ANIONIC
I ISOTHERMAL
A THERMAL
i-
z
w
u
fi2-
L

2
.
0
c - _---
- "r --
4 /

/
z 0
a-!,'
Downloaded by [Princeton University] at 06:30 03 June 2013

S I -
/ 0, sA, 0
W / /
/
/ /4 6
/
,4ss
q'
/
el /
/
Is4 I I 1 I
OO 0.5 1.0 I. 5 2.0 2.5

Fig. 35A. A plot of elongation of compression molded polystyrenes against


number-average molecular weight [lSS].

canizates [82], found that most samples had to be stretched to high


elongations before crystallization set in. Rupture occurred at slightly
higher strains. High MW samples crystallized at relatively low
elongations because the polymer chains in these samples were readily
oriented. Thus crystallization tended to reduce breaking strain in
the high MW samples. Since elongation was depressed by the high
concentration of terminal chains in the low MW fractions, the net
result was that a maximum occurred in the elongation vs. MW data.
Similar conclusions can be drawn from data on filled and unfilled
polyisoprene vulcanizates [254]. When samples of similar stereo-
regularity a r e compared at 20°C (Table 6), elongation is seen to
decrease slightly with MW. It also decreases a s stereoregularity
increases because higher elongations a r e required to crystallize the
polymers with low stereoregularity . At 100"C elongation clearly
increases with MW. Presumably crystallization rates a r e much
smaller a t this temperature s o the effect of chain ends is controlling.
Elongation of unfilled SBR vulcanizates [46] was found to be quite
dependent on MW for MW less than 1.5 x lo5 (Fig. 14). It also ap-
peared to decrease somewhat in the highest MW sample even though
MECHANICAL PROPERTIES OF POLYMERS 105

-
3.
0 ANIONIC
ISOTHERMAL
A THERMAL
c
t
W
:2
W
a
Downloaded by [Princeton University] at 06:30 03 June 2013

0 0.5 1.0 1.5 2.0 2.5


(b) fin 10-5

Fig. 35B. A plot of elongation of compression molded polystyrenes against


number-average molecular weight [ l a g ] .

SBR does not crystallize to any great extent. Using a standard tire
tread recipe, Yanko [301] found that elongation of SBR fractions was
virtually independent of MW for G n > 1 X lo5 (unmilled). Batches
with Gn < 6.5 X 10‘ could not be cured with this particular recipe,
In two other studies of SBR properties, elongation was found to be
independent of M W D [141,160].
Although polyethylene is normally considered to be crystalline,
its properties have been measured above the crystalline melting
point [170]. This study showed that elongation of radiation cross-
linked specimens exhibited a maximum when plotted against MW if
measurements were made on the rubbery material at constant dose
level.
3. Slightly Crystalline Polymers. A s in Section 111-B-3, the ma-
terials discussed below a r e largely amorphous, but the actual test
samples may have had some small degree of long-range order. Few
of the papers characterized sample morphology, s o the term “slightly
crystalline” is rather arbitrary.
An early study on cellulose acetate [258] indicated that the ulti-
mate elongation of both fractions and blends leveled off when DPn >
106 MARTIN, JOHNSON, AND COOPER

200 (Figs. 36 and 37). Sokolova and Rogovin 112571 observed relatively
small changes in the elongation of cellulose acetate fibers when the
D P of fractions was raised from 160 t o 305 o r when low MW material
was added to a whole polymer. Scherer and co-workers found that
Dwmin was approximately 200 in elongation tests of ethyl cellulose
[2481 and approximately 100 for cellulose nitrate [249]. However,
the experimental data on blended samples could not be directly re-
lated to D P because, in regions near and below DPw,min, the shape
of the MWD was also important. Rough correlations were obtained
when the data was plotted against shape factor (discussed in Section
111-B-3). The shape factor approach was also used by Scherer and
Downloaded by [Princeton University] at 06:30 03 June 2013

Chinai [247] to characterize the physical properties of polyvinyl


acetate fractions and blends. Properties generally increased with
shape factor. However, when elongation of these cast films was
plotted against Dw, a maximum was found at mw
= 1000.

0 - 1 2
I
INTRINSIC VISCOSITY

5 L A
200 400 600
NUMBER-AVERAOE DP

Flg. 36. Ultimate elongation of cellulose acetate fractions a s a function of


intrinsic viscosity (upper graph) and DPn (lower graph) 12581.
MECHANICAL PROPERTIES OF POLYMERS 107

0 2.75+ .23
Downloaded by [Princeton University] at 06:30 03 June 2013

I 2 3
I NTRlNSlC V I s c o s ~
~y

$30. U

z
0
I-
a
(3
2 20 ~ --

s
W
@ 502+204
W 502t127
c 0 502+43
r 10
e 502i92
502+100
5
3 @ 331+127
0 STARTING
0- 0
n
MATERIAL
i

200
- 400
NU M 8 E R AVE RAGE DP
600

Fig. 37. Ultimate elongation of blends of cellulose acetate fractions a s a


function of intrinsic viscosity (upper graph) and r P n (lower graph) [258].

Hosoda used PVC (DP = 800-1500)t o study the effect of tempera-


t u r e on maximum elongation and the crack resistance at fixed elonga-
tion [125]. He found that the higher M W samples exhibited higher
elongations and higher cracking temperatures. Pezzin [221]a l s o found
that the elongation of plasticized PVC increased with MW (Fig. 38).
Although there was no clear MWD effect, it may be significant that
almost all of the specimens with Q 2 2.2 fell above the curve in
108 MARTIN, JOHNSON, AND COOPER

Fig. 38 while samples with Q 5 2.1 lie below the curve. Weight-
average MW a r e relatively insensitive t o the presence of low molecu-
lar weight components. Thus when comparisons are made between
polymers of the s a m e Gw, the broad distribution samples will con-
tain much higher proportions of low MW material. A s shown by
Savel’ev et al. [246], the addition of PVC telomers to PVC increases
the ultimate elongation through a plasticizing action. Low MW com-
ponents in the blends, whole polymers, and fractions shown in Fig.

500
X 13.6 1.78
Downloaded by [Princeton University] at 06:30 03 June 2013

a/


400

2.2 /
X
7.0 X /
xYl.70 0 2.0
300 t.e/
0
2.2 I
-
zf 8-
I
m 0
Lu
0 2.2 1
200
0

8
01.75

8
0
/ 0 2.00
/
I00 0

50 100 150 200


M, IO-~
Fig. 38. Elongation of plasticized P V C vs. mw.N u m e r a l s a r e values of Q
[ 219,2211.
MECHANICAL PROPERTIES OF POLYMERS 109

38 would have a similar effect. The presence of such material would


increase both Q and elongation without grossly affecting EweThis
could explain the possible MWD effect noted above as well a s some
of the scatter in the data.
Wallach's data on polyimide films [291]of low crystallinity, 2-5%,
shows trends similar to that found on the plasticized PVC samples
because, a s shown in Figs. 18-20, elongation had not leveled off even
when hw= 2 x lo5. The MWD effect is also evident in Fig. 20.
4. Highly Crystalline Polymers. The ultimate elongation of un-
fractionated low-density polyethylenes has been shown [259]to depend
simultaneously on density and h v ; it increased with increasing 6iv
Downloaded by [Princeton University] at 06:30 03 June 2013

- constant density and decreased with increasing density at constant


at
Mv. By suitable combinations of mv and density, constant elongation
could be obtained over a considerable range of these variables. The
dependence of elongation on MW has also been shown in Refs. 148,
170, 199, and 235 for unfractionated, low-density polyethylene. Bi-
axial results a r e reported by Hopkins et al. [123].
In a study of fractionated high-pressure polyethylenes by Kamide
et al. [143],the elongation of fractions was never higher than that of
the whole polymer, and often considerably lower. The density of the
fractions was not reported and the scatter in the data points makes
interpretation difficult.
F o r high-density polyethylenes of similar relative specific vis-
cosity and crystallinity, a narrow MWD resulted in higher elongation
[36]. This was confirmed by Tung's work [273]where the elongation
of fractions was found to increase rapidly with MW (Table 10). At
constant MW, elongation decreased a s the density increased. An
increase in elongation with increasing Mv was also exhibited by
whole polymers in this study but the increase was much slower. F o r
samples of similar Ev, high elongation was favored by narrow MWD.
Karasev et al. [146]examined the tensile behavior of 10 poly-
ethylene fractions (a, = 6 x lo4 to 1.5 X lo5) with similar densities,
0.95. The range of test temperatures extended from -40 to +lOO°C.
The results of this study showed that the presence of low MW mate-
rial substantially decreases elongation at low temperatures. From
+ 20 to + 80"C MW changes had just the opposite effect; elongation
increased when MW was raised. A t- 100°C the high MW samples had
the highest elongations. Density, Mn, Gv, Mw, melt index, and sev-
eral molecular structure parameters were reported for these sam-
ples. However, the overwhelming effects of temperature make it
impossible to relate elongation to any particular MW average except
over narrow temperature intervals. MWD effects are somewhat
Downloaded by [Princeton University] at 06:30 03 June 2013

a
3

Table 10
Tensile Properties of High-Density Polyethylene Fractions [273]

Yield Ultimate Total Tensile


[rll strength strength elongation modulus
Fraction (dl/g) MW Nominal MW Density (Psi) (psi 1 (8) (psi x

1-1 3.10 165,000 110.000 0.936 2,470 5,310 985 -


1-2 1.55 63,000 43,000 0.942 2,730 3,660 1,630 1.24
1-3 1.01 34,000 26,000 0.949 3,060 1,680 210 1.67
1-4 0.84 27,000 20,000 0.953 3,660 - 17 1.71
2-1 2.47 118,000 76,000 0.954 3,600 5,000 973 1.72
2- 2 2.64 130,000 84,000 0.952 3,430 3,280 680 1.60
2-3 1.92 85,000 57,000 0.956 3,560 2,450 107 1.78 z
2-4 1.60 65,000 45,000 0.957 3,530 2,780 67 1.94
2-5 1.37 52,000 37,000 0.959 3,170 - 13 2.03
2-6 1.01 34,000 26,000 0.963 3,650 - 12 2.04
MECHANICAL PROPERTIES OF POLYMERS 111

clearer because the samples with narrow distributions had unusually


high elongations, particularly near the extremes of the -40 t o + 100" C
temperature range. Some of the data from this study is given in
Table 7.
In contrast to some of the above reports, elongation, at Gv>
5.6 x 104, has been shown [ 1541 to decrease with GVat fairly similar
densities, Table 8. This trend was also noted in limited tests on
bimodal narrow distribution polyethylenes [21] and in a study of
medium-pressure polyethylene [3]. The data in the former study
show little sensitivity to changes in MWD. These conflicts are at
least partially reconciled by Williamson et al. [298] where elongation
Downloaded by [Princeton University] at 06:30 03 June 2013

of high-density polyethylenes increased rapidly with [a] until a maxi-


mum was reached (Fig. 39). Further increases in [77] caused a
decrease in the elongation of both fractions and whole polymers. The
effects of MWD were substantial only a t the lower MW where narrow-
ing the MWD increased elongation. The decrease in elongation at
high MW has also been observed in Ref. 283. MWD effects were
noticeable only when comparisons were based on f i n ; broader MWD
samples gave higher elongations. Low MW additives apparently have
a variable effect on this parameter [217].
Kamide and co-workers [ 1441 measured the effects of MW and
MWD on the mechanical properties of undrawn, isotactic polypro-
pylene filaments. Crystallinity and orientation were held constant
while the properties were measured on fractions, blends, and whole
polymers. Elongation exhibited a maximum a t GV= 4-5 x 104 and
was independent of MWD. In oriented fibers, tenacity and elongation
both increased with G n at maximum draw ratio [202].
Ultimate elongation of injection-molded polypropylene specimens
increased a t low MW while compression-molded samples showed the
opposite trend [ 1691. Others have found that compression-molded
polypropylene exhibits either a sigmoidal relationship between
elongation and MW [255], o r else goes through a maximum [133]. The
elongation of injection-molded specimens apparently goes through a
maximum when plotted against MW [255]. Increases in the amount
of low MW atactic material reportedly increase elongation because
of a plasticizing effect "721.
Although there was some scatter in the data, Sato and Ishizuka
[244] showed that the elongation of isotactic polypropylene films
made from low MW material is greatly reduced when the MWD is
broad. Once [77] (decalin, 135") was increased above 1.5 dl/g however,
the MW and MWD effects were not detectable.
Nagano [200] measured the tensile strength and elongation of
polyvinyl alcohol films cast from a water solution. The five samples
112 MARTIN, JOHNSON, AND COOPER

1400

1200
7---
8 1000

L
Downloaded by [Princeton University] at 06:30 03 June 2013

G
I-
a
(3
800
s
W

W
l-
a
2 600
+-I
3

400

200 I
I
1 8
INTRINSIC VISCOSITY, dl/g

Fig. 39. Relationship between ultimate elongation and intrinsic viscosity. Sym-
bols a s in Fig. 3 12981.
used in this study had bimodal distributions (Fig. 27). Although
elongation of the five samples ranged from 6.8 to 13.23/0,this prop-
erty could not be correlated with MW, MWD, o r Scherer’s shape
factor. Tensile strength was a linear function of this last parameter.
In another polyvinyl alcohol study the elongation of fibers was found
to be only weakly dependent on M W characteristics when measure-
ments were made at constant draw ratio [263].
MECHANICAL PROPERTIES OF POLYMERS 113

In one study of poly(oxymethy1ene) diacetates [ 1061, elongation


was correlated with E n . A s Gn was varied from 9.4 x los to 9.2 x
109, the elongation increased from 3 to 25%. The data, Table 9, are
not completely separable from MWD effects because the samples
with low values of E n also had broad distributions. Others [174]
were unable to correlate the elongation properties of polyoxymethyl-
enes with Gn. This was thought to be due to variations in sample
fabrication.
Elongation to break of filaments made from regenerated cellulose
-
fractions increases with DP up to D P 400 [55]. Further increases
in MW did not affect the elongation of unoriented filaments but small
Downloaded by [Princeton University] at 06:30 03 June 2013

decreases were observed in the oriented specimens (Fig. 40). Elon-


gation of Toramonen process viscose reportedly increases with D P
-
to D P 800 [264]. More complete characterization of this material
is included in Ref. 162. No correlation between elongation and D P of
bleached sulfate chestnut pulp was found by Raway [228] over a D P
200
L

Effective

- 150- . S t r e tc h

I
E
Z
P
.
v)
z
W
I-
X
too-
W
zY
a
W
a
m
. 7.'
* 0

>
8 50-

> 100%
0
^c=
200%

I I I I
200 400 600 000 1000
-DP-

Flg. 40. Relation between FP and dry breaking extension at various levels of
effective stretch 1551.
114 MARTIN, JOHNSON, AND COOPER

range of 650-950. Relationships between D P and the wet and dry


elongation of various cellulose fibers a r e presented in Ref. 162 in
terms of morphological factors, orientation, and tensile strength.
Thomas [267] reports that increases in the MW of polytetrafluo-
roethylene can decrease the ultimate elongation by as much as 20%.
Specific data were not given in this review.
D. Yield Point
In polymeric materials the type of failure, whether brittle o r
ductile, depends on temperature, strain rate, sample configuration,
and mode of loading as well as various material parameters. Yield
Downloaded by [Princeton University] at 06:30 03 June 2013

strength normally increases with strain rate, crystallinity, and


degree of cross-linking. Increases in temperature and plasticizer
levels usually reduce the yield strength while MW is reported to
have only indirect effects [284]. The relative importance of these
and other factors is discussed in various reviews [9,17,204,284,286,
2871.
Wyman [299] reported an MWD effect on the flexural yield strength
of polystyrene, When tested according to ASTM D-790 (73"F, 1/8 X
1 x 4 in. samples), the yield strength of the fractions leveled off at
-
MW above 1.6 x lo5. However, the values measured on these frac-
tions (11,000-13,600 psi) were significantly higher than that recorded
on a broad distribution polystyrene (9,750 psi).
Other studies which relate the yield strength of amorphous poly-
mers to MW are few but some work has been accomplished on poly-
carbonate [loll. Data from this study a r e included in Figs. 41 and
42. In the high strength region of these plots, the failures were
ductile so the stresses and strains a r e yield values. Brittle behavior
was observed in the low-strength area. Yield strength was found to
be a weak but linear function of MW as it slowly decreased toward
the transition zone. Using the data from this work, the authors pro-
posed that an equivalence exists between strain rate and MW because
stress-MW data obtained at various strain rates could be shifted to
form a composite curve. Stress vs. log (strain rate) and strain vs.
log (strain rate) curves were treated in a similar manner. As a
result of these superpositions, the authors concluded that a tenfold
change in strain rate was approximately equivalent to a change of
1000 in Bv.
Several authors [57,253,259] have shown that the yield strength of
polyethylene is a function of density and, to a first approximation, is
independent of MW. Conversely, the yield strength of the low-density
material has been shown (page 95 of Ref, 23) to increase with MW in
the range 1.6 x lo4 to 2.9 x 104. A similar trend was noted in Refs.
Downloaded by [Princeton University] at 06:30 03 June 2013

5
D
I
D
zD
P
-U
n
B
rn
n
-I
rn
v)

MOLECULAR WEIGHT x Id’


Fig. 41. Flexural yield strength (or brittle strength) v s . log strain rate and molecular weight (1011.
Downloaded by [Princeton University] at 06:30 03 June 2013

116
MARTIN, JOHNSON, AND COOPER
MECHANICAL PROPERTIES OF POLYMERS 117

3 and 199. Low MW components tend to depress the yield point of


whole polymers [148,217,283].
The relation between density and yield strength of high-density
polyethylene [273]is shown in Fig. 43, the solid circles representing
5,000
.-
u)

I
Q
- 4,000
I-
$ 3,000
W
n
Downloaded by [Princeton University] at 06:30 03 June 2013

I-
v, 2,000
0
-J
W
< 1,000

0
0-91 092 0.93 0.94 0.95 0.96 0.97
DENSITY

Fig. 43. Effect of density on yield strength [273].

the fractions and the open circles the whole polymers. The dotted
line superimposed on this plot is that determined [259]for low-
density polyethylenes. In a study of fractions of high-density poly-
ethylene [154],it was found that below 15,000 MW the samples were
too brittle to make moldings. At 17,000 MW, tensile yield (Table 8)
was low; but at 30,000 and above, the yield strength decreased with
increasing MW as did the density. Again, the effect of MWD was
seen in that tensile yield was lower for the fractions than the whole
polymers.
Williamson et al. [298]confirmed these trends and found that the
yield strength of high-density polyethylene increases linearly with
density. MW and MWD correlations were also noted because the
yield strength increased a s the MW was decreased and the MWD was
broadened. These trends appear t o be second-order effects because
they a l s o change the degree of crystallinity (compare Figs. 44 and
3). At high values of [q],where the per cent crystallinity is almost
constant, the yield strength is almost independent of [q]. In other
polyethylene studies, minor decreases in yield strength resulted
when the MW of whole polymers was increased [113], while Belov et
al. [21]found yield strength to increase with decreasing MW'D in
bimodal specimens. There is some indication [146]that reducing the
118 MARTIN, JOHNSON, AND COOPER

.-a
0

I
I-
Downloaded by [Princeton University] at 06:30 03 June 2013

3500
W
a
c
In
P
-1
w*
3000

2500

L I I I
2 4 6
I N T R I N S I C V I S C O S I T Y dl/9

Flg. 44. Relationship between yield strength and intrinsic viscosity. Legend
as in Fig. 3 [298].

MWD increases the yield strength a t elevated temperatures (Table


7). MW effects were not substantial.
Several investigators have linked the yield strength of polypro-
pylene t o over-all crystallinity [252,255,293]. The percentage of
atactic polymer is also a significant factor because this material
substantially lowers the yield strength of both compression-molded
[169,227,252,255] and injection-molded samples [255]. In various
studies yield strength has been shown t o increase slightly with MW
[144,244], decrease slightly with MW [90,231], and to be independent
of GW[227] over various MW ranges. The tensile yield of injection-
molded specimens has been reported to increase with MW [255] but
the opposite result was found by Lapshin et al. [169]. MWD appar-
ently has little effect on yield strength because as shown in Fig. 45,
MECHANICAL PROPERTIES OF POLYMERS 119

0
Downloaded by [Princeton University] at 06:30 03 June 2013

1.14 1.12 1.10 1.08


SPECIFIC VOLUME
Fig. 45. Relation between modulus, yield strength, and crystallinity [ 2931:
(0, x)measurements on fractions, ( 0 , 0 ) m e a s u r e m e n t s on whole polymers.

fractions and whole polymers both fall on the same curve when this
property is plotted against per cent crystallinity. This was also
shown by Kamide et al. [ 1441 where measurements were made on
undrawn isotactic polypropylene filaments at constant crystallinity
and orientation, but a slight increase in the yield strength of films
was noted [244] when the MWD was reduced.
Measurement on nylon 66 and 610 showed a good correlation
between yield strength and crystallinity [262] while increases in the
per cent crystallinity of polyethylene terphthalate from -2 to -40%
give relatively small increases in yield strength except at tempera-
tures above the glass transition [61]. No MW dependence was appar-
ent in either of these studies. The yield point of ethylene copolymers
was also independent of MW [84] as expressed in terms of melt flow
index (Fig. 46).
E. Brittle Point
A s defined by Vincent [287], brittle fracture is a break which
occurs at a low elongation and at a s t r e s s below the yield stress.
Unlike yield strength, brittle strength decreases substantially as the
MW is reduced. Therefore, at a given temperature and strain rate,
low MW materials are more likely to be brittle than their higher
MW homologs. Relations between the brittle-yield transition, tem-
perature, strain rate, and material parameters are reviewed in
Refs. 9, 17, 240, 284, and 287.
The effect of MW on the low-temperature brittle strength of
several polymers is illustrated in Fig. 47. These flexural tests
[284] were carried out at -196"C, but even at this temperature the
120 MARTIN, JOHNSON, AND COOPER

7000 - C 0 POLY M ERS


A - ETHYLENE/PROPYLENE

.- 6000 - x - E THYLENE/BUTENE- I
o - ETHYLENE/HEXENE-I
Y)
0 0 - ETHYLENE/OCTENE- I
' 5000-
I-
5
4000-

D - X -
A
-* -
Downloaded by [Princeton University] at 06:30 03 June 2013

3000-
n "
0

I 10 100 1000 10,000


FLOW INDEX - MG/MINUTE
FIg. 46. Yield point a s a function of flow Index [84].

higher MW (to -loE)polyethylenes were not brittle. The values of


Kin on the plot were estimated from a melt flow index relation.
Brittle/ductile behavior was also demonstrated on polycarbonate
[loll. At 20°C and a strain rate of 1 in./in, min, brittle behavior
was observed when fell below -14,000 (Fig. 41). A correlation
between the brittle strength of PMMA and g n is included in Ref. 285.
Houwink (page 218 of Ref. 126) gathered data on the glass-transi-
tion and brittle temperatures of polyisobutylene and showed that
they are closely associated at MW > lo4 (Fig. 48). This correspon-
dence between the second-order transition and brittle temperature
should not be generalized because, as shown in Refs. 9 and 284),
there a r e numerous exceptions to this behavior.
Low-temperature brittle properties of polyethylenes a r e im-
proved as the melt index is decreased (Chapter 7 of Ref. 23,50,57,
105,148,233,235,286,294).Most of the studies showed that further im-
provements could be obtained by lowering the per cent crystallinity.
Avoiding the use of low MW waxes as processing aids is also bene-
ficial [ 1481. In oriented films a reduction in the MWD improves
brittle characteristics [57]. Data from two of these papers a r e
included in Fig. 49-51. One of the main points in Ref. 286 was a
MECHANICAL PROPERTIES OF POLYMERS 121

-
Mn
106 105 50,000 25,000 20,000 15,000 10,000
Y; 3 5 1
Downloaded by [Princeton University] at 06:30 03 June 2013

,I,\Y-
'S
5

1 2 3 4 5 6 7 8 9 1
0

Fig. 47. Effect of number-average molecular weight, Mn, on brittle strength


for ( a ) polyethylene, ( b ) polymethyl methacrylate, ( c ) polystyrene [284 1.

rough correlation between ductility and the ratio uB/uy, where cr, is
the brittle strength a s measured a t a low temperature and u, is the
yield strength at ambient conditions. This ratio, along with ductility,
is increased by reducing the density and increasing the MW. Since
tough, ductile materials a r e desired in most applications, uB/uy was
proposed a s a screening parameter for evaluating new materials.
At room temperature the brittle/ductile transition in polyethylene
fractions reportedly occurs above av
= 56,000 but below a v = 2 x
lo5 [183] although the data of Kenyon et al. [154] indicate a lower
value (WV = 30,000) a s shown in Table 8. Raspopov et al. [227] found
that the transition in an isotactic polypropylene is centered about
aw = 30,000. At constant per cent crystallinity, the limits of the
brittle region shifted to slightly lower MW when the temperature was
increased. When measurements are based on flex life data at con-
stant crystallinity, the transition apparently takes place over a fairly
small molecular weight range [293]. The relationship between crys-
tallinity and MW (inherent viscosity) obtained in this study is given
in Fig. 52. Brittleness in polypropylene has also been related to
122 MARTIN, JOHNSON, AND COOPER

-20

-
0
0
-40

W
a
3
l-
a
Downloaded by [Princeton University] at 06:30 03 June 2013

a
W
a
L -60
W
I-

-a o

I I I
I I I I I I

I o3 lo4 10s lo6

MOLECULAR WEIGHT

Flg. 48. Effect of molecular weight on second-order transition point, T, and


brittle point, Tb. of polyisobutylene [126].

orientation [65], stereoregularity [72], and spherulite size [252]. The


fact that the low MW, highly stereoregular samples form large
spherulites apparently increases their tendency to be brittle. In
studies conducted on polybutene-1, the brittle behavior of low MW
samples was attributed to less extensive interlamellar reinforcement
in the spherulites [108].
The brittle temperature of polyoxymethylenes was found to be
closely related to Z n over the range of 25-80 x lo3 [174]. Brittle
temperature dropped rapidly at first but leveled off a t the higher MW.
F. Modulus
Low-strain properties of glassy polymers are usually not functions
of MW or MWD except at very low MW. Thus it is not surprising
that the tensile moduli of polystyrene [ 180,189,1961,polydichloro-
MECHANICAL PROPERTIES OF POLYMERS 123
Downloaded by [Princeton University] at 06:30 03 June 2013

I
DO
MELT FLOW INDEX

Fig. 49. The effect of melt flow index on the flexural strength a t -180°C
(brittle strength) of ethylene polymers for three types of fabrication condi-
tions 12861: ( A ) compression molded and slowly cooled, ( B ) compression
molded and rapidly cooled, ( C ) injection molded and tested near the gate.

styrene [180], and PMMA [155] a r e reported to be independent of MW.


Flexural modulus measurements on polystyrene [ 189,2991 and poly-
carbonate [ 1001 led to similar conclusions although a later test on
polycarbonate [ 1011 indicated that the modulus decreased slowly a s
ii?, was increased over a wide range of strain rates. Berry reported
[22] that the tensile modulus of PMMA increases linearly with l o g 3 v .
Several of the tests noted above [189,196,299] also examined MWD
effects. These tests, all conducted on polystyrene, showed that MWD
has no major influence on modulus.
Martynyuk et al. [ 1861 measured the effects of temperature, MW,
and MWD on the isothermal compressibility of polystyrene. Less
extensive data on poly-ry -methylstyrene was also presented. The
results of this investigation (Fig. 53) indicate that the isothermal
compressibility (the inverse of isothermal bulk modulus) is indepen-
dent of MW until the glass-transition temperature is approached. In
this region the sigmoidal curves do reflect differences in MW. MW
effects a r e also present above the glass transition because com-
pressibility over the 200-400 kg/cm2 range increased with tempera-
124 MARTIN, JOHNSON, AND COOPER
Downloaded by [Princeton University] at 06:30 03 June 2013

DENSITY AT t 2OoC

Fig. SO. The effect of density (at +20°C) on the flexural strength at -180°C
(brittle strength) for four ethylene polymers. Melt flow indices: ( A ) 1.5,
( B ) 5, ( C ) 6.5, (D)20 [ Z S S ] .

ture f o r the high MW samples but not for the telomers (Fig. 53B).
As a result, the low MW specimens were more compressible than
their higher MW homologs in the transition region but less compres-
sible at higher temperatures. Limited data indicated that MWD has
little effect on compressibility. When this property was measured
on two samples which had the same M W (1.41 x lo5) but different
distributions, the points all fell on the same curve (Fig. 53C). The
MWD of the two samples was not given.
Flory [82] showed that plots of 300% modulus (tension at 300%
elongation) vs. 1/Mn were linear for butyl rubber specimens with the
same degree of unsaturation (Fig. 54). Since a whole polymer had
the same modulus as a fraction with the same k n , it appears that
MWD effects were small. This conclusion was also reached [141,
1601 for SBR. In another study on SBR, Yank0 [301] used fractionated

-
material in a tire-tread compound to show that the 300% modulus
increases with En.The curve leveled off at E n 4 x lo5 (unmilled).
When comparisons were made at constant cross-link density, how-
ever, the 300% modulus of SBR was not appreciably affected by MW
[46]. Measurements made on radiation cross-linked polyethylene
Downloaded by [Princeton University] at 06:30 03 June 2013

-200 -160 -120 - 80 -40 0 40 80 120


TEMPERATURE ( O C )

Flg. 51. Breaking energy v s . temperature for 1 mm polyethylene films in a


mv
biaxial test [105]: ( N ) = 5 x lo4, narrow MWD, density = 0.960, crystal-
linity = 7 4 % ; ( M ) & = 2.75 X lo5, broad MWD, density = 0.945, crystallinity
= 63% ( H ) Zv > lo6, broad MWD, density = 0.932, crystallinity = 53%.

1
0 1 2 3 4 5
IN HE R E N T V I S C O S I T Y

Flg. 62. Brittleness in relation to crystallinity and inherent viscosity [293].


above its crystalline melting point indicated that modulus increases
with initial MW when comparisons a r e made at constant dose level
[170]. This is at least partially due to the differences in cross-link
densities because the number of cross-links/cc required to form a
network decreases as MW increases.
126 MARTIN, JOHNSON, AND COOPER

75

50

25

75

lD
50
- U
Downloaded by [Princeton University] at 06:30 03 June 2013

I-

bp
-I
25

0
75

50

25

0
20 60 100 140 180 220
TEMPERATURE ( " C )

Fig. 53. Temperature-compressability relationship of polymers over a pres-


s u r e range of 200-400 kg/cm2 [186]. P a r t A: Curve I is polystyrene;
(0)Ev = 6.139 x lo5, (0) & = 3.75 x 10'. Curve I1 i s poly-a-methyl styrene;
(0) mv= 6.068 X lo5, (a) Ti?, = 5.48 x l o 4 . P a r t B: Polystyrene; (0) zv =
6.139 X lo5, ( e )MV= 2.39 X lo4, ( 0 )Z n = 7 . 2 X lo3, ( 0 ) mn = l o 3 . P a r t C:
Polystyrene; (0)EV= 1.41 x lo5, narrow MWD; (0)m w = 1.41 x lo5, broad
MWD.

Early work on vinyl chloride/vinyl acetate copolymers indicated


that the modulus increased sharply with M W up to 6000-8000 [56,62].
Further increases due to MW were relatively small. As expected,
broadening the MWD of PVC by adding vinyl chloride telomers
(DP = 5.5-6.0) lowered the 100% modulus [246]. In a study of plasti-
cized PVC fractions and mixtures [221], Youngs modulus was found
to be constant for aw
> 5 x lo4. It decreased rapidly when Gw was
reduced below this value.
The modulus of various polyethylenes [259] is plotted as a function
of density and log viscosity in Fig. 55. Since the modulus isopleths
are parallel to the viscosity axis, this parameter is apparently inde-
MECHANICAL PROPERTIES OF POLYMERS 127
Downloaded by [Princeton University] at 06:30 03 June 2013

Fig. 54. Tension (“modulus”) at 300% elongation vs. reciprocal molecular


weight M ( o r X n ) 1821: (0)low-unsaturation fractions, (a)unfractionated
Polymer XX-A, ( A ) high-unsaturation fractions from Polymer IV, ( 0 )high-
unsaturation fractions from Polymer V.

pendent of MW. Early workers proposed that short-chain branching


determined crystallinity [31,236]. Since chain branching decreased
as MW increased in a given polymerization system, it was suggested
[235,241] that modulus was indirectly dependent on MW. However,
numerous workers have shown that modulus is independent of MW
[199] and either linearly related to density over a small range [57,
2731 o r else slightly curved [191] over an extended range of density.
In tests conducted over a range of - 125 to + 80” C, Carey and co-
workers (41,431 found that Young’s modulus increases with MW
above - 50”C . Per cent crystallinity was apparently not held constant
in this work. Since crystalline content, o r density, normally de-
creases with increases in MW, one would have expected Young’s
modulus to decrease with MW instead of increasing. The discrepancy
may be due to variations in the short-chain branching of these com-
mercial materials as explained above.
128 MARTIN, JOHNSON, AND COOPER

I I I f

0.94 *
80

).
0.93 - 7 4 -
5's
1
58
4 r 40
I- 2
.
9 34 30
5 0.92 * '30
z 20
W
* 0.91 *
..I4
14
16 %;-I6
16
0.90 -
Downloaded by [Princeton University] at 06:30 03 June 2013

0.89

Broadening the MWD increased the tensile modulus [36] of poly-


ethylene when comparisons were made at a constant solution vis-
cosity. This was presumably due to the fact that broadening the
MWD at constant aw
increases the per cent crystallinity of a mate-
rial [157]. Orientation introduced during the injection molding also
influenced sample stiffness.
Wijga [293] and Fil'bert [80]found that the modulus of unoriented
polypropylene increases with per cent crystallinity. Since the same
relationship held for both fractions and whole polymers (Fig. 45),
M W D effects a r e apparently unimportant. Tensile modulus of com-
pression-molded samples is also independent of M W although a
maximum in the modulus vs. MW data was noted for injection-molded
specimens [255]. The influence of orientation on modulus is also
evident in Fig. 6 where the draw ratio of polypropylene fibers is
clearly much more important than ZW.
The moduli of 66 and 610 nylons were related to per cent crystal-
linity and moisture content [262]. Per cent crystallinity is also the
controlling parameter in determining the modulus of polyethylene
terphthalate [61]. No MW effect was found in either of these studies.
Wet and dry stiffness of viscose fibers was expressed in terms of
DP, the size of the morphological units, and orientation [ 1611.
In contrast to the results obtained on other crystalline polymers,
Grohn and Friedrich [ 1061 reported that the flex modulus of poly-
(oxymethylene) diacetate increases with $k However, one could
also use this data to show that, at constant Mw, flex modulus in-
creases a s Q (MWD) is increased or a s S n is decreased (Table 9).
These changes both tend to increase the per cent crystallinity, s o an
MECHANICAL PROPERTIES OF POLYMERS 129

increase in modulus as these parameters a r e varied would be ex-


pected from the results observed on other crystalline polymers. The
per cent crystallinity of the samples used in Ref. 106 was not re-
ported.

IV. IMPACT STRENGTH


The impact strength of a material is a measure of its ability to
resist breakage under high-velocity loading conditions. As the veloc-
ities and modes of loading vary widely, both in practical applications
and in impact tests, there is often no more than a qualitative agree-
Downloaded by [Princeton University] at 06:30 03 June 2013

ment between test results. In some cases even this qualitative agree-
ment is poor because test bars and molded products often have
different orientation characteristics, particularly near the surfaces
[ 1241. In spite of these correlation difficulties, most investigators
have concluded that MW is a very significant parameter in determin-
ing impact strength. In fact, a s shown below, an order of magnitude
increase in impact strength can be obtained by increasing MW.
The significance of MWD is not so clear, although some workers
have reported that the use of fractionated polymer increases impact
strength. The basic problem in determining MWD effects is that
most impact studies have been conducted on unfractionated o r poorly
fractionated materials. This limitation has been necessary because
impact studies require a relatively large number of samples in order
to obtain statistically valid results. This factor has effectively ex-
cluded the use of narrow fractions because the preparation and
characterization of large quantities of fractionated polymer is diffi-
cult.
The use of high MW polymer with a narrow MW distribution also
maximizes orientation. This factor must be considered when inter-
preting impact data. F o r example, McCormick et al. [189] found that
both wide and narrow distribution polystyrene samples fell on the
same curve when tensile impact strength was plotted against Mw.
Their results showed that the impact strength increased by a factor
of 10 when Gw was increased from 1 x lo5 to 4 x lo5. As the samples
used in this work were injection molded, part of the change was un-
doubtedly due to sample anisotropy. However, the effect of MW is
undoubtedly real because Maigeldinov et al. also found that an in-
crease in the MW of polystyrene from GV= 5 x lo4 to 15 x lo4 gave
an order of magnitude change in impact strength [180]. In both of
these studies impact strength appeared to approach an asymptotic
limit at high MW, Gw = 5 x lo5 and ifv = 1.5 x lo5, respectively. A
limit of MW = 2 x lo5 was reported [291] for polyimide films with
2-5% crystallinity (Figs. 18-20).
130 MARTIN, JOHNSON, AND COOPER

Savel'ev [246] studied the effect of bimodal MWD by adding poly-


vinyl chloride telomers to a polyvinyl chloride polymer. A small
amount of the telomer, up to lo%, increased the notch impact strength
slightly because the additive improved processing characteristics
and apparently reduced residual stresses in the material. However,
additional amounts of the telomer substantially reduced the impact
strength. Early work [56,62] indicates that the impact properties of
vinyl chloride-acetate copolymers a r e enhanced when the MW is
raised.
Haward et al. [113] used polyethylenes of various melt indices to
measure the effect of blending on physical properties. They found
that when two materials of widely different melt indices were blended,
Downloaded by [Princeton University] at 06:30 03 June 2013

the resulting mixture had an impact strength which was similar to


that measured on an unblended sample of the same melt index. This
result showed that, within the scope of the investigation, MWD did
not greatly affect impact strength because the blending operation
increases the width of the MWD. It should be pointed out that com-
mercial polyethylenes were used in this work rather than narrow
fractions, hence the conclusions may not apply t o narrow distribution
materials. A definite MW effect was found, however, a s the Izod
impact strength decreased by a factor of six when the melt index was
increased from 0.01 to 1.0.
Notched Izod impact strength has been measured on compression-
molded polyethylene fractions [298]. The results showed that impact
strength increases with MW a t constant density. It also appeared to
increase when the MWD was reduced because the impact strength of
fractions was substantially higher than that measured on whole poly-
mers (Fig. 56). However, these results do not conclusively show an
MWD effect because, a t constant [ q ] ,the per cent crystallinity of the
fractions was lower than that of the whole polymers. Other investi-
gators have shown similar MW [50,103,154,273] and MWD [36,103,
105,154,2731 effects in high-density polyethylene (Figs. 57 and 58 and
Table 7). Since the fractions (solid circles) in Fig. 57 were taken
from polymers of two different densities, one might conclude that
small variations in the per cent crystallinity have little effect on
impact strength. Relationships between impact strength, zw, thermal
history, and the spherulite size of linear polyethylene are given in
Ref. 213.
In unfractionated low-density polyethylene, Charpy impact strength
is a function of both melt index and density as shown in Fig. 59 [294].
However, when these data were plotted as a function of an, the den-
sity effects were insignificant (Fig. 60). Analogous results were
obtained in an environmental stress crack resistance test (Figs. 61
and 62).
MECHANICAL PROPERTIES OF POLYMERS 131

i- A- A
" &

-
Downloaded by [Princeton University] at 06:30 03 June 2013

MW x 10-3
Fig, 57. Effect of molecular weight on tensile impact strength [273].

Another study on polyethylene [191]found that the tensile impact


strength could not be related to density, o r density and MW. The
third variable, MWD, was investigated; and on the basis of varying
one characteristic a t a time, the toughness was found to increase
with decreasing density, melt index, and MWD. This was confirmed
by other workers [281]who related the increased strength of the
narrow MWD a t constant inherent viscosity to orientation during
flow of these injection-molded specimens. It was shown that frozen-
132 MARTIN, JOHNSON, AND COOPER

I
0

COLUMN
Downloaded by [Princeton University] at 06:30 03 June 2013

0 0.5 1.0 1.5 2.0 2.5 3.0 35 4.0 4.5


MELT INDEX
Fig. 58. Impact-flow relations for high-density polyethylene [1541.

\o

DENSITY
0- .917
A- -924
0- .929

MELT INDEX

Fig. 69. Influence of melt index and density on impact resistance of low-den-
sity polyethylenes [294 I.
MECHANICAL PROPERTIES OF POLYMERS 133

5-

DENSITY

4 - 0 -- ,917
A
0 - ,924
,929
7
0 -
23
I I
i
s .
2
Y
Downloaded by [Princeton University] at 06:30 03 June 2013

3
3
2-
0
W

1 I I 1 I
4 5 6 7

IMPACT R E SlST ANCE


Fig. 60. Impact resistance-molecular weight relationship for low-density
polyethylenes [294 1.

in orientation was strongly dependent on both MW and MWD. Similar


results were found for high-density polyethylenes [36,154,273].
Energy-to-break, deformation-at-break, and breaking strength
of three linear polyethylenes were determined in a falling ball test
[ 1051. The results of this biaxial test indicated that performance in
these categories is significantly improved over a temperature range
of -180 to +120°C when the MW is increased. The data, Figs. 51 and
63, also show that the low-temperature limit on impact strength is
extended when MW is increased. Part of these improvements a r e
undoubtedly due to the large differences in the per cent crystallinity
of the three samples.
Van Schooten et al. 12521observed the effects of MW, MWD, per cent
crystallinity, crystal structure, and stereoregularity on the impact
strength of polypropylene fractions. When other factors were held
constant, the data indicated that samples made from high MW, narrow
MWD material gave the best impact strength. Increases in the per
cent crystallinity and the size of the spherulites reduced impact
strength. This is actually an indirect MW effect because the lowest
134 MARTIN, JOHNSON, AND COOPER

\O

\
DENSITY
0,.917
&-. 9 2 4
QL* ,929

\<\
lot
1

L ' O\ 0
=.
s1.0-
Downloaded by [Princeton University] at 06:30 03 June 2013

,
a .
0 '
r :

i
0.1
0.I -
I
1.0
\a\

MELT
"\

INDEX
1-
10
\ A

rl
100

Fig. 61. Influence of melt index and density on ESCR of low-density polyethyl-
enes [294].

MW stereoregular fractions generally had both high crystalline con-


tent and large spherulites . Introduction of stereoregularity greatly
complicates the situation because the addition of either amorphous
atactic material o r rubbery stereoblock fractions t o isotactic poly-
propylene increases the impact strength even though it lowers the
MW of the blend [169,252,255]. Others have also related the impact
strength of polypropylene to MW [65,90,133,153] and density [65,133].
Because of the number of interrelationships among the parameters
listed above, it is difficult to cite more than qualitative trends with
respect to the impact strength of polypropylene. The MW effects
noted by Fujioka [go] were observed on two commercial polypro-
pylenes of similar density and isotactic content. Measurements were
made on a falling dart instrument which was modified s o that force
and acceleration could be monitored as a function of time at impact.
Although the higher MW sample (melt index = 3 . 1 g/10 min) had a
brittle temperature which was about 10°C below that of the compari-
son sample (M.I. = 7.9 g/10 min), the author concluded that changes
MECHANICAL PROPERTIES OF POLYMERS 135

5-
DENSITY
4- 01.917
~-.924

9
-3’
c
I
w .
z
a
Downloaded by [Princeton University] at 06:30 03 June 2013

U
J
32-
0
w
_1

t
1 . . . ....
I 1 1

0.1 I .o 10
ESC R Fw,hr

Flg. 62. ESCR-molecular weight relationship for low-density polyethylenes


[294 1.

in MW alone will not be sufficient to give polypropylene adequate


impact strength at low temperatures.
Kelly and Dunn [153] modified the conventional tensile impact test
so that the entire stress-strain curve could be recorded instead of
just rupture energy. The data, as measured on linear polyethylenes,
polypropylenes, polyacetals, and ethylene-butene- 1 copolymers,
show that rupture energy increased with MW for all four types of
polymer. The stress-strain curves indicated that the changes were
largely due to plastic deformation a s the initial elastic behavior was
relatively unaffected by the MW variations. Impact behavior differ-
ences were detectable even when specimens had similar rupture
energies as measured in the conventional tensile impact test.
Although much of this discussion has centered on polyolefins, the
MW and crystallinity effects appear to be quite general. For exam-
ple, at constant crystallinity the pneumatic impact strength of nylons
66 and 610 increased with MW [262]. At constant MW, however, high
crystallinity was found to reduce impact strength (Fig. 64). A similar
interrelationship between the per cent crystallinity and MW has been
Downloaded by [Princeton University] at 06:30 03 June 2013

-200 -160 -120 -80 - 40 0 40 80 I20

TEMPERATURE ('C)

Fig. 63. Deformation-at-break vs. temperature for 1 mm polyethylene films


in a biaxial test. Legend as in Fig. 51 [105].

noted in polyethylene terphthalate [ 611. Spherulite size and storage


conditions were also significant factors in this study. Other impact-
MW correlations include papers on propylene-ethylene block copoly-
mers [275], polyacetals [106], and regenerated cellulose film [ 1971.
The polyacetal data is included in Table 9.

V. STRESS CRACKING AND CRAZING


A. Introduction
Brittle fracture in polymers at relatively high-strain rates is
discussed in Sections 111-E and IV, This mode of failure can also be
observed in some long-term tests at low stresses. Most long-term
failures fall into one or more of the following categories:

(1) Creep rupture. Data on both brittle and ductile failures are
reviewed in Section VII-C.
(2) Thermal s t r e s s cracking. Heat alone can cause stress failure
MECHANICAL PROPERTIES OF POLYMERS 137

i
I-
Q
2
W
E 300-
l-
0
ti
g>oo -
-I
oc:
I-L
a
z
Downloaded by [Princeton University] at 06:30 03 June 2013

3
W
loo-
z
a

I I 1

PE R C E N T CRY STALL I NI TY
Flg. 64. Impact s t r e n g th vs. crystallinity: comparison of 66 and 610 nylons.
Points: 610 nylon, Samples A and B. Shaded a r e a s : 66 nylon, Samples C and
D. Lower pattern: lower MW Samples A and C. Upper pattern: higher MW
Samples B and D [262]. The crystallinity axis w a s r escal ed in a l a t e r p ap er
[262a].

by allowing internal physical changes to occur, i.e., morphological


structure. Unlike other forms of s t r e s s cracking, rupture usually
initiates within the material rather than a t a surface [ 1151.
(3) Chemical s t r e s s cracking. An example of this process is the
irreversible oxidative attack on the surface of a material while it
is being subjected to thermal s t r e s s cracking conditions.
(4) Solvent crazing and cracking. Here a swelling agent initiates
either brittle or ductile failure by weakening and plasticizing the
surface layers.
(5) Environmental s t r e s s cracking. This rather general term is
normally restricted to failure that occurs when a stressed sample
contacts a surface-active agent. Unlike solvents, the surface-active
agents appear to cause surface embrittlement [ 1311 but have no
chemical or physical effect on unstressed specimens.

Most of the material in the next two sections is drawn from the
literature on environmental and solvent stress cracking although
several examples from the other categories will also be discussed.
138 MARTIN, JOHNSON, AND COOPER

B. Amorphous Polymers
Polymeric materials often fail prematurely a t very low s t r e s s
levels. In order to determine whether MW o r MWD influenced s t r e s s
cracking in polystyrene, Rudd [239] immersed a series of narrow
distribution specimens in n-butanol and applied an initial tensile
s t r e s s of 1000 psi. A set of broad distribution samples was also
exposed t o this environment (Table 11). The results (Figs. 65-67)
Table 11
Polystyrene Samples [239]
Downloaded by [Princeton University] at 06:30 03 June 2013

Heat distortion
- temp Tav ( se c )a
Material K4 Mwmn (“C) (uo= 1000 p s i )

Anionic Polystyrene
S103 124,700 1.05 103 38.0
S105 153,500 1.04 101 148
s109 193,000 1.06 102 8,300
Slll 239,000 1.08 100 24,300
S108 267,000 1.08 105 169.500
Isothermal Polystyrene
B2 187,000 1.90 97 95.3
B3 214,000 1.94 96 1,700
I34 255,000 1.96 100 7,700
B5 285,900 1.77 100 53,900
B6 44 7,000 2.85 97 268,000

a ~ a V= time to zero s t r e s s in butanol at 26 * 1°C.

showed that MW drastically affected s t r e s s crack resistance because


the time to failure increased by almost four orders of magnitude
when the weight-average MW of the fractions was increased from
1.247 x lo5 to 2.67 x lo5. A substantial MWD effect was also observed
because, when comparisons were made at the same &J, it was found
that the narrow distribution specimens generally lasted 10- 100 times
as long as the broad distribution samples (Fig. 67). The curves in
this plot do not, of course, continue to increase indefinitely. An
MWD-solvent stress creeping relationship was also demonstrated in
a tensile creep test of polystyrene [266]. Reducing Q from 2.6 to 1.06
tripled the rupture time in oil when RW= i.9 x lo5.
In contrast to Rudd’s data, Hopfenberg [122]found that polystyrene
MECHANICAL PROPERTIES OF POLYMERS 139

I 1 1 I I
100 - I I

00 -
b”
IL
0
60 -
t-
$ 40 -
0
E
Downloaded by [Princeton University] at 06:30 03 June 2013

W
a
20 -

I I 1 I 1 I I
100 10’ 10‘ lo3 lo4 lo5 I06

TIME (SEC)

Fig. 66. S t r e s s decay curves for narrow distribution (anionic) polystyrenes


in butanol environment, uo = 1000 p s i 12391.
I I I I I I
100

00

b”
L 60
0
t-
z
W 40
0
a
w
Q
20

0
I 1 I I I I
100 10’ 10‘ lo3 lo4 lo5 lo6
T I M E (SEC)

Flg. 66. S t r e s s decay curves for broad distribution (isothermal) polystyrenes


in butanol environment, (TO = 1000 p s i 12391.
140 MARTIN, JOHNSON, AND COOPER

lo'
106
I
-
POLY STY R EN E
(BUTANOL. 1000 PSI)

lo5 -
0
w
8
c lo4-
Y
4
Downloaded by [Princeton University] at 06:30 03 June 2013

Km 103-

0
I-
w 102-
EI-

10' -
I00 -
i
/
/
I 1 I I
lo-'
0 100,000 200,000 300,000
-
Mw

FIg. 67. Dependence of time to break upon mw for polystyrene in butanol envi-
ronment, u0 = 1000 psi, (-) = narrow distribution samples, (- - ) = broad
distribution samples 12391.

films of 2.4 x lo5 MW crazed faster than those made from 1.8 x lo5
MW material. He attributed the difference to residual stresses in
the higher MW material. This was confirmed later when Bray and
Hopfenberg [27]separated MW and orientation effects by casting one
set of polystyrene films (MW = 1.16 X lo5 and 5.37 x lo5) onto mer-
cury while a control set was cast and drawn down on glass. After
drying and annealing 15 min at 105"C,the films were exposed to
n-pentane and the decrease in uncrazed core thickness was mea-
sured as a function of time and temperature. The results obtained
on the mercury-cast films indicated that low MW polystyrenes do
indeed craze faster than higher MW materials when orientation and
residual s t r e s s effects a r e absent. However, the opposite result was
found in the sample which had been cast on glass, apparently because
MECHANICAL PROPERTIES OF POLYMERS 141

the rapid draw down of the casting bar induced different orientation
and s t r e s s e s which were not removed during the short annealing
period. Sat0 [243] exposed polycarbonate specimens to CCl, in order
to measure the effect of orientation, plasticization, and MW on stress
cracking. He found that stress crack resistance increases when the
s t r e s s is applied in the direction of the orientation. It decreased
when tested in the perpendicular direction. Annealing was partially
effective in improving stress crack resistance while plasticizers
decreased the resistance. Increases in MW greatly increased the
time to failure when the MW was above 25,000.
Kambour [142] reviewed craze phenomena in glassy polymers and
Downloaded by [Princeton University] at 06:30 03 June 2013

showed that material in a craze is composed of soft, tough polymer


instead of glassy material. This structural change is due to the fact
that high localized strains cause rarefaction and void formation in
the polymer. When a crack is formed in a glassy polymer, the frac-
ture tip is preceded by a craze zone. Thus fracture is actually a
two-step process; formation of a craze and breakdown of the craze.
Although Kambour’s tests were on polycarbonate, he used Rudd’s
data to illustrate the significance of the two steps. He concluded that
craze formation is not strongly dependent on MW because the initial
portion of Rudd’s s t r e s s relaxation curves a r e similar for polysty-
rene fractions of different MW (Fig. 65). However, as noted above,
craze failure differed by orders of magnitude in these curves. Hence
craze breakdown and the consequent crack propagation is strongly
affected by MW. Berry’s [22] photomicrographs on polymethyl
methacrylate show similar results with the high MW samples exhibit-
ing greater crack toughness. They also gave more complex craze
patterns instead of forming cracks.
C. Crystalline Polymers
A s characterized in numerous reviews [8,9,50,127-131,140,201],
the environmental s t r e s s crack resistance (ESCR) of crystalline
polymers can be affected by thermal history, MW, MWD, chain
branching, crystallinity, crystal structure, orientation, surface im -
perfections, chemical nature of the medium, temperature, stress
level, modes of imposed and internal stresses, catalyst residues,
molding pressures, and storage time. Since polyethylene is unusually
sensitive to environmental s t r e s s cracking, most of the published
work in the field deals with this material. Unfortunately, quantitative
comparisons of the data obtained by different investigators are usu-
ally impossible because of the overwhelming importance of thermal
history [58,128-130,132,147,218].
142 MARTIN, JOHNSON, AND COOPER

Qualitative comparisons leave no doubt that MW can be extremely


significant in determining the ESCR of polyethylene. Increases in
MW improve ESCR whether measured in t e r m s of melt index [ll,
Chapter 5 of 23,64,69,9 1,92,113,117-119,127-130,146-148,167,168,
190,201,218,260,282,294,302], solution viscosity [40,41,91,92,94,128,146,
193,2981, a n [294], o r Williams plasticity [59,71]. Kenyon et al. [154]
and Williamson et al. [298]used fractionated polyethylene in their
studies while most of the other reports are based on commercial
whole polymers. Typical results from several of these papers are
given in Figs. 61, 62, and 68. The MW effects tend t o be less impor-
tant in high-density resins [64,92,167,168,190].
Downloaded by [Princeton University] at 06:30 03 June 2013

Density correlations a r e more complex than the MW relationships.


When comparisons a r e made at constant strain, ESCR decreases a s
density increases [59,92,127,128,130,201,294].This is not unexpected
because, a t constant strain, the stiffer high-density resins a r e sub-
jected to much higher stresses. In tests conducted a t a high constant
stress, the resistance to stress cracking increases with density if
comparisons are made at similar melt indices [69,119,128,130,135,
16’7,1901.At lower levels of constant stress this dependency is re-
versed because, a t low stresses, the superior creep resistance of
the high-densitv resins is less important 1118.119.1681. When E n
[294],[~][298],o r RSV (relative specific viscosity) [94]a r e used to
characterize the MW of whole polymers, the apparent changes due
to variations in density become insignificant, at least over limited
ranges of density. This effect is apparently related to MWD a s well
a s MW because the curves of ESCR vs. [v]of polyethylene fractions
taken from 0.945 and 0.960 material did not superimpose [298]. Plots
of ESCR vs. melt index and KIn of low-density commercial polyethyl-
enes are presented in Figs. 61 and 62.
Since rapidly cooled specimens are less crystalline than annealed
samples, one would expect that quenching a sample would improve its
ESCR in a constant strain test if excessive internal s t r e s s e s could
be avoided. Actually, the MW-crystallinity-thermal history relation-
ship is much more complex as shown by Howard and Gilroy [132],
Table 12. The expected effect of crystallinity on the ESCR of “as
molded” samples is observed for short cooling periods. However,
when the cooling cycle is extended, the decreases in the melt index,
possibly due to entanglements o r cross-linking, becomes the con-
trolling factor After annealing, samples with the lowest melt indices
had the highest ESCR despite the fact that they also had the highest
per cent crystallinity. The thermal conditioning generally decreased
ESCR even though MW, as characterized by melt index, and crystal-
linity, a s determined by density measurements, were almost un-
MECHANICAL PROPERTIES OF POLYMERS 143
Downloaded by [Princeton University] at 06:30 03 June 2013

71s P
Fig. 68. Effect of molecular weight on environmental stress crack resistance
[154].

changed in some of the specimens. Thus a model which describes


the ESCR of polyethylene in terms of over-all density o r crystallinity
is insufficient; details of the morphological structure are required.
In an early paper, Hopkins et al. [123] proposed that stress-crack-
ing agents a r e adsorbed onto the polymer surface. By exerting a
film-spreading pressure at the apex of any flaw, the fluid would
accelerate crack propagation. Indeed, surfaces which a r e molded or
Downloaded by [Princeton University] at 06:30 03 June 2013

Table 12
Changes in Properties of a Typical Polyethylene on Moldinga [132]

Stress-crack r e s i s t a n c e
M e l t indexb Density F50 ( h r ) *
Cooling time
(min) A s molded Conditioned' A s molded Conditioned' A s molded Conditioned'

Nonee 0.28 0.9338


Shock-cooled 0.27 0.27 0.9322 0.9329 18 1
3 0.26 0.26 0.9336 0.9339 12 2.5
10 0.26 0.25 0.9346 0.9347 6 3
30 0.25 0.24 0.9350 0.9350 7.5 3.8
140 0.23 0.23 0.9357 0.9357 9 5
300 0.9358 0.9358 11 5.8
1200 0.18 0.17 0.9360 0.9359 23 15

aHeld a t 165'C for 3 min p r i o r to cooling to room temperature a s shown.


bASTM D 1238, Condition E.
'Conditioned f o r 18 h r at 70°C after molding.
dASTM D 1693 using 10% aqueous Igepal CO-630 a t 50°C. Boiling water conditioning omitted. Fso ( t i m e to 50%
failure) determined graphically.
eOriginal pellets a s supplied.
MECHANICAL PROPERTIES OF POLYMERS 145

fire polished a r e much less susceptible to s t r e s s cracking than


freshly cut surfaces which have numerous flaws [40,127,128,131].
However, elimination of surface flaws will not prevent stress crack-
ing because Haas and MacRae found that even low-level biaxial
stresses create microscopic cracks o r voids a t spherulite centers,
and, to a lesser extent, at spherulite boundaries [107,108]. Neither
environmental nor solvent s t r e s s crack agents appeared to increase
the rate of crack formation but they did influence the propagation
rate. When a surface-active agent was applied to stressed films, it
entered void areas at the spherulite centers and spread out along the
radial cracks which formed a s the lamallae separated. Since this
Downloaded by [Princeton University] at 06:30 03 June 2013

liquid, Igepal CO-630, penetrated the spherulite before spreading, it


was suggested that initiation of macroscopic cracks may occur below
the surface of the specimen under certain circumstances. These
tests were conducted on linear polyethylene and polybutene-1. In the
low MW polyethylene sample, cracks also propagated from the mate-
rial which was left between spherulites. However, once such a crack
came to the boundary joining two spherulites, it tended to change
direction and enter one of them, presumably because the imposed
biaxial s t r e s s spreads the lamallae. Raising the MW reduced brittle
fracture because, a s MW is increased, the number of interlamaller
Utie"molecules also increases. The tendency for cracks to propa-
gate radially in spherulites was also noted by Hittmair and Ullman
[119].
Microscopic examination of polyethylene samples exposed t o dis-
.
tilled water while under constant load showed that Ufissures o r . .
microzones of cold drawing" form in the region beyond a razor cut
[135]. These cold-drawn areas were observed at both spherulite
boundaries and within spherulites . Gradually they grew, coalesced,
and caused macroscopic cold drawing at the cut. Similar fissures
were formed in samples immersed in a more active environment,
isopropanol. However, the microscopic cold drawn regions did not
expand in this environment because rupture occurred in the boundary
region between drawn and undrawn material. Thus crack growth was
accelerated and the failure mode was brittle instead of ductile. The
susceptibility of the drawn-undrawn transition region was confirmed
on a tensile sample which had been stretched in air so that part of
the specimen was cold drawn. When stressed and exposed to isopro-
panol, macroscopic cracks appeared almost immediately at the
shoulder between the drawn and undrawn regions. Others have re-
ported similar results [ 50,115,1321. This observation is important
because it implies that stress crack-tests must be designed s o that
some localized portion of the sample is near its yield point. A test
146 MARTIN, JOHNSON, AND COOPER

in which all of the material is above o r below the yield point would
give misleading ESCR data.
Although MW clearly affects ESCR, the significance of MWD is
uncertain. Zabusky [302] showed that the shape of the MWD can
affect both processing variables and ESCR. When two polyethylene
resins of the same melt index were compared in a modified Bell Lab
test, ASTM D1693-59T, the specimens with a bimodal MWD lasted
three times as long as the commercial materials. In a second com-
parison, the exposure time required for 50% failure was increased
4-8-fold when the bimodal product was substituted for conventional
products in a detergent bottle stress cracking test.
Downloaded by [Princeton University] at 06:30 03 June 2013

Although one recent study [146] linked ESCR to the proportion of


high M W material, a number of workers [35,119,127-131,1681 have
reported that extraction of low MW material improves the ESCR of
polyethylene, Table 13. A s expected, addition of low MW waxes to low-
density resins impairs this property [148]. Veselovskaya et al. [282]
obtained just the opposite result, Tables 14 and 15. They concluded
that small amounts (0.6-13.5%) of low MW (-500-700) waxes act a s

-
plasticizers. When Isaksen et al. [135] investigated this effect by
adding linear hydrocarbons (MW 500, 1500, 2000) to a moderately
sharp fraction of high-density polyethylene, only minor changes in
s t r e s s crack behavior were observed even though some of the blends
were quite brittle. Additive levels ranged from 1 to 10%. It might be
significant that the fraction used in this test was a fairly low MW
material (Gn = 20,000, f i v = 33,000, melt index = 40) so its ESCR
would be rather low even without the additives. However, other
studies have also shown either no difference between the ESCR of
fractions and a whole polymer [154], Fig. 68, or ambiguous results
[298]. In the latter paper, whole polymers with densities of 0.96 and
0.945 had similar values of ESCR when compared at constant [17] in
the Bell test, ASTM D1693-60T. Fractionation impaired the ESCR
of the 0.96 material but improved performance in most of the speci-
mens taken from the 0.945 polymer.
Narrowing the MWD at constant melt index reportedly improves
the s t r e s s crack resistance of whole polymers [57,130,260]. This
trend was also observed in annealed specimens at similar values of
GWby Herman and Biesenberger [116]. However, when they exam-
ined quenched samples they found that ESCR is degraded when M W D
is reduced. The authors cited evidence that narrow MWD materials
have a more uniform crystallite structure and that high and low MW
"tails" a r e pushed out of the growing lamallae. In annealed speci-
mens the reject material forms secondary lamallae in the intercrys-
tallite regions These regions have unusually high concentrations of
Downloaded by [Princeton University] at 06:30 03 June 2013

Table 13
Effect of Variation in Chloroform Extractable Fraction [130 J n
i
n
Biaxial stress-strain n
properties
Melt indexa Inherent Density CHCI a extract Cracking time' Stress at Stress at
(g/1O min) viscosityb (25OC ) (YOby wt) F50 ( h r ) rupture ( p s i ) rupture ('56)

1.0 1.042 0.9209 11.8 4 13,400 385


1.0 1.040 0.9201 10.6 25 12,800 375
1.0 1.066 0.9226 9.7 > 1000 16,600 420
1.0 1.063 0.9213 9.3 > 1000 17,900 450
1.0 1.052 0.9211 7.9 > 1000 14,800 4 10

aASTM Designation: D 1238, Condition E.


bAt 0.5 g/dl in xylene at 85°C.
'Bent strip test using Igepal at 50°C.
148 MARTIN, JOHNSON, AND COOPER

Table 1 4
Resistance of Polyethylene to Cracking a s a
Function of the Percentage of Low M W
Material [ 282 ]

Melt flow Low MW Resistance to


index material cracking,
15 (%) F50 ( h r )

6.75 1.15 10
6.0 9.8 15
Downloaded by [Princeton University] at 06:30 03 June 2013

5.7 0.86 12
5.7 13.5 > 500
4.79 2.9 35
4.64 7.7 60
4.12 11.4 400
3.36 1.45 15
3.4 2.6 25
3.6 6.4 170
3.78 9.25 >1,000
2.04 0.76 28
2.45 1.75 28
2.82 3.35 46
2.45 5.8 28
2.4 7.9 >400
1.65 0.64 88
1.64 1.12 130
1.15 2.70 115
1.35 3.04 295
1.39 3.36 700
1.03 4.05 >1.000

MW material, voids, amorphous material, and internal stresses s o


they a r e very susceptible to s t r e s s cracking. Narrowing the MWD
reduces the amount of low MW reject material, refines the crystal-
lite structure, and lowers the void content so the observed improve-
ments in the ESCR of annealed samples when the MWD is reduced is
not unexpected. Similar arguments a r e presented t o explain the
opposite trend which was found in the quenched samples. Although
their work was not directly applied to ESCR, Keith and co-workers
[151,152] have extensively investigated M W fractionation during
crystal growth and the nature of intercrystalline linkages. These
MECHANICAL PROPERTIES OF POLYMERS 149

Table 15
Influence of Removal of Low M W
Fractions on the Properties of Polyethylene (2821

Low M W material Melt flow index Cracking resistance


(%I 15 F50 ( h r )
Refore After Before After Before After
extraction extraction extraction extraction extraction extraction

13.5 - 5.7 2.64 500 60


11.4 0.92 4.12 2.13 400 150
Downloaded by [Princeton University] at 06:30 03 June 2013

9.25 1.85 3.78 2.28 1,000 60


2.96 0.56 1.39 0.9 1,000 130
0.77 0.23 0.58 0.5 1,000 230

concepts offer tenable explanations for a number of observed phe-


nomena including the effect of annealing on ESCR, the tendency
towards radial crack propagation in spherulites, and the sensitivity
of ESCR to molding cycle variations, storage time, and low MW
components.
Dunkel and Westlund [65] were unable to induce s t r e s s cracking
in polypropylene because its s t r e s s crack resistance is much greater
than that of polyethylene. However, Forsman et al. [86] did observe
stress cracking in a constant load test of commercial polypropylenes
(melt index = 0.13-15: 83-96% isotactic) which were immersed in
water. Since water is a relatively inactive s t r e s s cracking agent for
polyolefins, the failures were presumably caused by some combina-
tion of environmental, thermal, and chemical s t r e s s cracking as
well as creep. Typical results for one of the test materials is given
in Fig. 69. No MW effects were evident when 100 h r rupture s t r e s s
was plotted against melt index. However, MW became quite important
in the longer term, lower s t r e s s , tests, Figs. 70 and 7 1 . Other factors
considered in this study were melt index-temperature interactions,
stereoregularity, crystallinity, orientation, elastomer blending, and
the applicability of time-temperature superposition.
Although most of the data presented in Sections V-B and V-C per-
tain to either environmental o r solvent s t r e s s cracking, contributing
effects from other s t r e s s cracking modes are usually present. Howard
[ 1281 discussed this point and concluded that material parameters
which affect ESCR in crystalline polyolefins also control other types
of s t r e s s cracking although their relative importance may be different.
150 MARTIN, JOHNSON, AND COOPER

I
..
I I 1

\ ?0002
- O V o F
-\
-.
( 250c)

140.6

\.-
._
0 O----OO
5
\

lo
a
Ln -8 x
0

\(49"c) c-
? Y 2 O O F m
rl+
Q
u)
W
Lz
$I

'. '.
' m \ ~ 0 F ( 6 6 0 C ~ ~
Downloaded by [Princeton University] at 06:30 03 June 2013

m 1000- 70 3
\
0 DUCTILE FAILURE \
BRITTLE FAILURE

OL I I I I

For example, in constant strain tests, the resistance to thermal s t r e s s


cracking increases with MW [44,114,134] and decreases with increas-
ing density [44,114,115]. The latter parameter is particularly impor-
tant if the sample is held at constant strain [114]. Constant s t r e s s o r
creep rupture tests usually indicate that MW is the more important
variable, as discussed in Section VII-C.
Although literature data was scanty, Howard [ 1281 suggests that
s t r e s s cracking induced by oxidation will be affected by the same
material parameters which control ESCR because the two modes of
failure differ only in the initiation step. Hinton and Keller [117]
showed that this conclusion is not valid for severe oxidation environ-
ments. When high-density polyethylene was exposed t o fuming nitric
acid at 60°C and constant s t r e s s , failure time was almost independent
of MW. Surface crystalline texture, thermal history, orientation, and
stress level were important, however. The authors noted that oxida-
tive cracking is a chain scission process s o it should be virtually
independent of IvlW. ESCR was considered to be progressive chain
disentanglement. Comparative melt index vs. rupture time curves for
polyethylene in nitric acid and in a surface-active environment a r e
presented in Fig. 72.
MECHANICAL PROPERTIES OF POLYMERS 151

.-a2000
\ 40.6

a
.
) N
u) E
u)
$
Downloaded by [Princeton University] at 06:30 03 June 2013

W m
a
I- UT
u)
v)
a W
I a
0 +
u)
0
0
IOOC ‘0.3

0 I I
I 1.0 10.0 I( .o
MELT FLOW RATE (23OoC)

Fig. 70. Effect of molecular weight on 1000 h r s t r e s s rupture [86].

VI. FATIGUE AND ENDURANCE


In early papers the fold resistance of cellulose nitrate [249,261],
ethyl cellulose [248], cellulose acetate [258], polyvinyl acetate [247],
and vinyl chloride-acetate copolymers [62] reportedly increased with
MW to some characteristic value before leveling off. The samples
used in these studies were at least roughly fractionated. Later work
by Morrow [I981 on two polystyrene fractions (Gw= 1.6 X lo5 and
8.6 x lo5) and unfractionated material also indicated that fatigue life
increases with MW a t a given stress level. When comparisons were
made between fractions and blends, most investigators concluded that
broad o r bimodal distributions usually decrease the fold resistance of
152 MARTIN, JOHNSON, AND COOPER

.;200( 40.6
a
u)-
u)
(Y
W E
a
4
Downloaded by [Princeton University] at 06:30 03 June 2013

I-
m s
a
I $
W
0 a
0
I-
9 m
0
loo( 'Oa3

I 1.0 10.0
MELT FLOW RATE (23OOC)
Fig. 71. Effect of molecular weight on 10,000 h r s t r e s s rupture [SS].

low MW specimens [247-249,2611.However, Sookne and Harris [258]


indicated that data from fractions and blends were similar when com-
parisons were based on the number-average MW.
The flex life of elastomers is often a function of environmental
conditions and heat buildup [17,204]. Bueche [34]discussed the latter
factor and showed that the lower flex life and higher rate of heat
generation in SBR, relative to natural rubber, is due t o the presence
of larger amounts of low MW material. Yank0 [301],and later Bar-
tenev et al. [16], came to essentially the same conclusions because
when the cure state of SBR compounds was adjusted to give a con-
stant 300% modulus, the flex life of SBR fractions dropped for a n <
lo5 (unmilled). Heat buildup, as measured by Goodrich hysteresis at
MECHANICAL P R O P E R T I E S OF P O L Y M E R S 153

6 I I I
1
Downloaded by [Princeton University] at 06:30 03 June 2013

"
0 5 10 IS 20
Tim. to failure in h o u r s

Fig. 72. Dependence of time to failure upon melt flow index (M.F.I.) for
polyethylene subjected to s t r e s s cracking in either fuming nitric acid o r
Teepol [ 117 1.

lOO"C,also increased a s the MW was decreased. If the proportion


of cross-linking agents is not adjusted to obtain constant cross-link
density, the fatigue life at constant dynamic force of a 30/70 SBR is
reduced for MW < 5 x lo5, Fig. 73 [16]. Fatigue life increases at
low MW when the strain amplitude is held constant because the lower
cross-link density of these samples allows plastic deformation to
occur. In a statistical study of SBR production samples, cycles to
failure in a DeMattia tear test was correlated with MWD [160]. No
correlations had been found in an earlier study [141].
Although details of the test procedure were not given, Haward et
al. [ 1131 showed that the fold endurance of polyethylene increases
rapidly with decreases in the melt index As in most flex tests, there
is substantial scatter in the data but the relation between melt index
and fold endurance was approximately linear on a log-log scale. Wil-
liamson et al. [298] used a Kohler-Molin flex tester to measure the
fold endurance of polyethylene. Although actual test data were not
presented, they reported that both fractions and whole polymers fell
on the same curve when log (fold endurance) was plotted against [q].
Since the fractions were less crystalline than the original polymers,
the authors concluded that neither MWD nor per cent crystallinity
were significant variables within the limits of the test and that only
1 MARTIN JOHNSON AND COOPER9
Downloaded by [Princeton University] at 06:30 03 June 2013

MOLECULAR WEIGHT x

Flg. 73. Effect of molecular weight of the fraction on the dynamic endurance
of vulcanizates with equal contents of combined sulfur (2.05%) and different
values of equilibrium modulus (17.2, 16.2, 15.2, 8.0, 3.6 kg/cm2). Curve 1:
constant strain; Curve 2: constant force (161.

MW was important. Wijga 12931 reported somewhat different results


when he tested polypropylene on a Kohler-Molin flex tester because,
a s shown in Fig. 52, both the MW and the per cent crystallinity affect
flex life. When compared at a constant per cent crystallinity, the
transition between the brittle and flexible regions occurs over a
surprisingly narrow MW range.
In tests conducted on polytetrafluoroethylene (Fig. 74), Thomas
[267] reported that either increasing the MW or decreasing the per
cent crystallinity can extend flex life by several orders of magnitude.
The flex life of poly(oxymethy1ene)diacetate [ 1061, regenerated cellu-
lose film [197], and bleached sulfate chestnut pulp [228] a r e also
reported to increase with MW. An MWD effect was noted by Grohn
and Friedrich [lo61 because, at constant [q],the flex life of broad
distribution samples was shorter than the flex life of fractions. This
trend was attributed to the lower MW tail of the distribution curve
although indirect effects due to differences in crystal structure o r
crystalline content may also have been contributing factors.
MECHANICAL PROPERTIES OF POLYMERS 155

VII. DYNAMIC STRESS-STRAIN PROPERTIES


A. General Considerations
The organization of Section VII differs from the format used in
Section I11 because it is subdivided according to test method rather
than basic material properties. This change was made because the
amount and type of data presented in most of the papers a r e insuf-
ficient to allow quantitative property comparisons when different
test methods were employed.
Few measurements are truly static s o the decision to review
some dynamic properties such a s impact strength and fatigue life
Downloaded by [Princeton University] at 06:30 03 June 2013

separately was rather arbitrary. Aside from the material on creep


rupture, this section is limited to studies which pertain to low strain
moduli (or compliances) sound velocity and their associated losses.
In summary, most of the work cited here indicates that neither

IC'01 , I /

>
k
z 60
J
-I
a
I-
v)
>
LI
0 40
8 F L E X LIFE VALUES

I
20

INON STANDARD TEST


45-CYCLES/ MIN.
180" BEND
I
2.30 2.26 2.22 2.18 2.14

(STANDARD SPECIFIC GRAVITY)


MOLE CU LAR WEIGHT -
Fig. 74. Flex life vs. % crystallinity and molecular weight [267].
156 MARTIN, JOHNSON, AND COOPER

MW nor MWD have a substantial effect on the modulus and losses of


amorphous polymers in the glassy region. The location of the transi-
tion zone and its shape can be influenced by changes in these param-
eters at moderately high MW according to a few creep studies which
were made on narrow distribution materials. However, stress re-
laxation and vibration tests have generally indicated that this zone
is independent of both MW and MWD except a t very low MW. MW
and MWD effects are quite significant in the rubbery region. Losses
normally increased and modulus decreased when low MW o r broad
distribution materials were tested.
Per cent crystallinity appears to be the controlling parameter in
determining modulus a s a function of time for crystalline materials
Downloaded by [Princeton University] at 06:30 03 June 2013

although long term creep studies indicate that MW is also important.


Rupture time under load was greatly enhanced by increases in MW.
Since internal losses in amorphous materials a r e much higher than
those measured on crystalline polymers, it is not surprising that
loss characteristics are primarily determined by per cent crystal-
linity. In some cases MW can affect losses indirectly because of its
influence on crystallization kinetics.
B. Stress Relaxation
A number of investigators have examined the effect of MW on
stress relaxation. The results have generally indicated that the re-
laxation modulus, Er(t), is independent of MW in the glassy and
transition regions. These conclusions were reached in studies of
polystyrene [Chapter 2 of Ref. 269, 2701, PMMA [192], polyisobu-
tylene [30,271], polyvinyl acetate [211], and plasticized PVC [112].
Although fractions were used in some of these studies, none of them
included measurements below @ V = 8 x lo4. Figure 75 presents data
from one of these papers. Curve A in this plot was considered to be
“slightly in doubt” for Er(t) > lo* dynes/cma because of sample
brittleness. The curve also shows that the rubbery region is affected
by MW because an increase in MW raises the rubbery plateau and
extends it to longer times and higher temperatures. Similar behavior
has been observed by others [1,10,30,179,192,211,271,272,and Chap-
t e r 2 and Appendix J of Ref. 2691.
Ninomiya [206,212] studied the effect of blending on the s t r e s s
relaxation of polyvinyl acetate in the rubbery region. Composite
s t r e s s relaxation curves were prepared from data generated on six
fractions (& = 8.5 x lo4 to 7.6 x lo5). Then, assuming that the re-
laxation spectra were linearly additive, new curves were calculated
for six blends. When these curves were compared with experimental
data, it was shown that interactions occurred, particularly when the
MECHANICAL PROPERTIES OF POLYMERS 157

10

-I
N
9

w
z
-*
0 7

-
c
c
Downloaded by [Princeton University] at 06:30 03 June 2013

3=
W
0
-I5

-I 0 I 2 3 4 5 6 7 8 9 10
LOG TIME ( S E C O N D S )

Fig. 75. Master curves for five monodisperse polystyrene samples at 100°C.
A is S102, MW = 8 X lo4; B is 5103, m w = 1.25 x lo5; C is S109, Ew= 1.93 x
lo5; D is S111, mw
= 2.39 X lo5; E is S108, Rw = 2.67 X l o 5 [270].

two fractions in the blend had large differences in MW. This MWD
effect was related to a n of the components through a defined shift
factor. A later paper [207]showed that the blending laws developed
for polyvinyl acetate could also be applied to polyisobutylene and
polystyrene. Tobolsky 12721 has also investigated MWD effects in
the rubbery flow region. H i s data on anionic and thermal polystyrene
-
showed that samples with Q 2 had steady-state shear compliances
which were roughly three times as great a s those measured on
“monodisperse” samples.
In measurements made on undrawn, isotactic polypropylene fila-
ments, the relaxation modulus at a given time reportedly went
through a maximum at f i v = 8 x 104 to 13 x lo4 [144]. For example,
when av
= 6 x lo4 and 20 x lo4, the 1 and 10 min readings were
approximately 40% lower than values recorded in the intermediate
range. MWD had no apparent influence on this effect.
C . Creep
Creep is usually considered to be the long-term deformation of a
material while under constant load. Figures 76 and 77 illustrate
158 MARTIN, JOHNSON, AND COOPER

RECOVERY
Downloaded by [Princeton University] at 06:30 03 June 2013

__---- _----
TIME (SECONDS)
u u
Fig. 76. Creep of a four element model, t=-+ - (1 - d7) +-Ot, with
E l E2 73
El = 5 x l o 9 dynes/cm2, E2 = l o 9 dynes/cm2, T = q 2 / E 2 , q 2 = 5 X lo9 poises,
q 3 = 5 x l o i i poises, and u = 10' dynes/cm2 [204].

typical creep curves for amorphous polymers in t e r m s of strain and


compliance [204,225]. No isothermal test can encompass the entire
time span shown in Fig. 77 so time-temperature superposition pro-
cedures a r e used to generate the complete curve at a given tempera-
ture. Details of this method and a comprehensive treatment of creep
phenomena are presented by F e r r y [77]. Nielsen [204] and Rimrott
and Schwaighofer [234] also review many practical aspects of creep
and creep rupture.
In an extensive study, Plazek and O'Rourke [225] measured tor-
sional creep and creep recovery on anionically polymerized polysty-
rene. Parts of this paper are included in Refs. 222, 223, and 225a; the
extremely wide range instrumentation is also described in Ref. 223.
Their initial observation was that time -temperature superposition
did not give a single response curve for the reduced creep compli-
ance, Jp, at temperatures near the glass transition, Fig. 78. However,
this procedure was successful in reducing the recoverable compli-
ance data, Fig. 79. The poor fit of the Jp data was attributed to a
difference in the temperature dependence of the viscoelastic and
viscous deformation mechanisms. Other discrepancies were noted
in the rubbery o r entanglement plateau region. For example, when
the MW of these narrow distribution samples was greater than the
MECHANICAL PROPERTIES OF POLYMERS 159

MW between entanglements but less than Mv = 5 x lo6, the observed


entanglement plateau occurred at a reduced recoverable compliance
which was above that found in samples with Ev > 5 x lo6, Fig. 80.
This behavior was consistent with several theoretical models but
incompatible with the commonly accepted FLW extension of the
Rouse-Bueche dilute solution theories [79]. Stress relaxation tests
on a series of narrow distribution polystyrenes by Tobolsky and
co-workers have also indicated deficiencies in the FLW model [l,
2701.
For comparative purposes the curves in Fig. 80 were shifted
along the time axis s o that their transition regions superimposed
Downloaded by [Princeton University] at 06:30 03 June 2013

Log Time
@ Glassy response @ Transient network response

@ Primary transition @ Steady slate complianco

@ Entanglement p l a t e a u 8 Vircour deformotion

Fig. 77. Semiquantitative schematic representation of the expected behavior


of the creep compliance, J ( t ) , for a linear amorphous polymer of moderate
molecular weight ( M > Me). Jg is the time-independent glassy compliance;
Jd is the contribution of the delayed deformation to the steady-state compli-
ance, Je. @ ( t )i s the normalized retarded compliance function and 1) is the
viscosity [225 I.
Downloaded by [Princeton University] at 06:30 03 June 2013

-4

A
Q)
144.9OC 0
133.8"
-: 125.0"

-6 _--

z
D
P
-I
z
b
I
5
D
z
U
c)
0
FA. 78. The reduced c r e e p compliance, J p ( t ) =
J(t)Tp/Topo, cm2/dyne, of polystyrene Sample A-25 2rn
(Mv = 4.69 x lo4); Q = 1.04,) plotted logarithmically against the logarithm of a time s c a l e reduced with W
shift factors calculated f r o m viscosity values, vp = vTp/T0pO, dyne s e c / c m 2 . Reference temperature,
To, is 1OO"C. The long-dashed line represents the viscous contribution to the compliance, t / v (100").
The reduced recoverable compliance for 100°C i s represented by the short-dashed line [222,225].
MECHANICAL PROPERTIES OF POLYMERS 161

I I
I I 1 I I 1 I I I I I

-6

-7
n
e
-
5
c
c
I
.
0
Downloaded by [Princeton University] at 06:30 03 June 2013

7
0 - 8
0
J

-9

-10

1 I I I I I I I I I I I 1
-I 0 I 2 3 4 5 6 7 8 9 C
I
Log t/a,

Flg. 79, The logarithm of the reduced recoverable compliance, J p ( t ) - t/up,


of A-25 plotted against the logarithm of the empirically reduced time scale.
Temperatures and sample a r e identified in Fig. 78 [222,225].

upon the A-19 data (av = 5.92 x lo5,Q = 1.05,). A check on these and
one other reduced curve in Ref. 225 showed that shifts of approxi-
mately 1, 0.25, and 0.7 decades were required t o obtain superposition
for samples of ZV= 4.69 x lo4, 9.4 x lo4, and 1.89 x lo5, respectively.
The erratic relationship between MW and shift magnitude in this
series may have been caused by the lack of a significant high MW
tail in the third sample because this material had an unusually nar-
row MWD (Q = 1.01J. The other two specimens (Q < 1.08 and Q =
1.04,) probably had distributions which were closer to that of A-19.
The apparent effects of MW and MWD upon the location of the
transition region contradicts the numerous s t r e s s relaxation and
vibration r e p r t s cited in Sections VII-B and VII-D. However, as
162 MARTIN, JOHNSON, AND COOPER

-5 I
1 I I I 1 1 I I I I I I I
Downloaded by [Princeton University] at 06:30 03 June 2013

10 12 14
Lag t / a T ( A19 TIME SCALE)

Flg. 80. Logarithmic comparison plot of reduced recoverable compliance


curves a s a function of reduced time. All curves shifted to superimpose with
response of A-19 in transition region. A-61[3]: z v = 1.64 x l o 4 , &is is a
fraction taken from a sample whose Q was l es s than 1.08. A-25: Mv = 4 . 6 9 x
l o 4 , Q = 1 . 0 4 , . L-5: & = 1.22 X lo5, Q = 1.055LL-2: mv = 1 . 8 9 x l o 5 , Q =
1.015, A-19: Z v = 5.92 X l o 5 , Q = 1.Oij6, A-16: Mv = 8 X 10'. Q not reported
f o r this anionically polymerized sample 1225 1.

discussed later, other creep studies have shown similar trends. It


is conceivable that shifts at moderately high MW are observable
only when sharp fractions a r e used.
Figure 80 also indicates that low MW specimens deviate from
their higher MW homologs in both the glassy and steady-state pla-
teaus, The difference in the glassy region probably reflects a broad-
ening of the dispersion a t low MW while the poor fit in the steady-
state compliance was attributed to a decrease in the number of
active viscoelastic elements at lower temperatures. Because of this
latter change, the classical temperature correction for entropy
effects was shown to be inadequate for low MW fractions.
MECHANICAL PROPERTIES OF P O L Y M E R S 163

The effect of MWD upon recoverable compliance was also demon-


strated in this study. Reduced curves for a mixture made up of a
low MW, narrow M W D polystyrene (98 wt %) and a high MW, narrow
MWD polystyrene (2 wt %) a r e given in Fig. 81. Even though the MW
of the main component in this blend is only slightly greater than the
MW of sample A-61[3] of Fig. 80, it is clear that the recoverable
compliance of this specimen is much higher than that measured on
any of the samples in Fig. 80. Others have observed similar effects
in polystyrene (2081 and polyisobutylene [ 1711 blends. The response
of the blend in Fig. 81 was also nonlinear. The anomalies were ex-
plained by noting that the higher M W component in the blend governed
Downloaded by [Princeton University] at 06:30 03 June 2013

recoverable compliance while the shearing field was largely con-


-3

-4

n
9
\
c
I

z -5
v
Q
7
U
cn
3

-6

I I 1 I I I 1 I I
-7 I

i
0 I 2 3 4
Log t
Fig. 81. Logarithm of apparent recoverable compliance vs. log t f o r mixture
of 98.0% A-58 (av = 2 x l o 4 , Q < 1.08) and 2.0% A-16 (ZV= 8 X lo5, narrow
MWD). Maximum s t r e s s in sample a t 138.2"C from largest compliance t o
lowest. 240: ( 0 )1000, 2300, and 6200 dyne/cm*, A t 129.3"C: ( 0 )1000, ( 0 )
3800 dyne/cm2. At 119.4"C: ( 0 )3800, ( 0 )14,600. At 109.4"C ( 0 )14,600 [225,
225al.
164 MARTIN, JOHNSON, AND COOPER

trolled by the lower viscosity, low MW “solvent.” Under the high


shear stresses the longer molecules ‘assumed grossly distorted
configurations, hence the nonlinear response. Notice that, unlike the
A-61[3] data, the reduced curves in Fig. 81 can be shifted into a
composite master curve with a single, well-defined steady-state
compliance. Sample A-61[3] is the middle cut taken from an anioni-
cally polymerized polystyrene (Q c 1.08) which was fractionated.
The original material, A-61, despite its narrow MWD, exhibited an
extremely high, nonlinear recoverable compliance similar to that
found in the blend. A s stated above, these phenomena were not ex-
hibited b y the fraction s o one would suspect that differences in the
Downloaded by [Princeton University] at 06:30 03 June 2013

amount of high MW material caused the changes. However, gel


permeation chromatography and ultracentrifuge tests of A-61 failed
to show any trace of high MW components s o the amount of this
material, if present, must have been extremely small. The reduced
A-61 data, like that of A-61[3], exhibited a steady-state compliance
which was temperature dependent. Since only 2% of a high MW mate-
rial was needed to give the high level, temperature-independent
steady-state compliance shown in Fig. 81, one would conclude that
even minor amounts of high MW impurities in low MW fractions can
completely mask the MW effects and will make low MW fractions
appear to be amenable to the standard reduction procedures.
Thomas and Hagen [266] conducted creep tests on polystyrene at
room temperature using a tensile load of 5000 psi. Two materials
which had the same aw (1.9 x lo5) were compared: one had a Q of
2.6, the other 1.06. The total strain in the broad distribution samples
exceeded that of the narrow distribution material over most of the
test. In addition, the average rupture time of the latter specimens
was twice a s long as that measured on the broad distribution samples.
A similar test conducted in oil showed a threefold difference in rup-
ture time for these materials.
Other workers have found that the creep compliance of polystyrene
[32] and fractionated PMMA [33] is insensitive to MW in the glassy

-
region. However, at higher temperatures and/or longer times, creep
and creep rate increases a s MW (Gw o r ii?v 105-106)decreases
[32,33,205]. Recoverable compliance was reported to be independent
of both MW and MWD over the range investigated [32].
In an early creep study Flory [82] attributed creep of cross-linked
polyisobutylene to diffusion of the distorted chain ends. His model
predicted that a derived creep relaxation constant would increase
linearly with 1/Mn. This relationship was approximately followed in
measurements made on fractionated samples but effects due to the
length of the terminal chains and rearrangements of the network
MECHANICAL PROPERTIES OF POLYMERS 165

chains were also noticeable. At a given stress, creep generally de-


creased when fin was increased. Similar results were obtained on
vulcanized SBR and polyisoprene samples when the long time por-
tions of these creep curves were evaluated [34]. Van Holde and
Williams [ 1211 reported that the steady-state compliance of uncured
polyisobutylene was not affected by Gn. MWD did appear to be im-
portant, however, because narrowing the distribution reduced the
elastic compliance. Both the creep curve and a defined recovery
function were sigmoidal when plotted against log (t/at). The recovery
curve shifted to longer times at higher MW while broadening the
MWD extended the transition area over a wider time scale. Leader-
Downloaded by [Princeton University] at 06:30 03 June 2013

man et al. [172] found that the steady-state compliance of polyiso-


butylene was affected by both fiw and MWD.
The viscoelastic properties of six polydimethyl siloxanes (Gw =
0.41 x lo6 to 4.9 x lo6) in the plateau and terminal zones were in:
vestigated using both creep and vibration techniques [ 2241. Although
these noncrystalline polymers were not cross-linked and completely
soluble in solvents, the highest MW samples exhibited no viscous
flow and complete recovery at low stresses. Yielding and consequent
flow did occur at higher stresses. This behavior, indicative of a
quasi-permanent network was correlated in terms of Andrade creep.
Ninomiya and co-workers [209,21Olmeasured tensile and shear
creep on polyvtnyl acetate fractions (Mv = 5.5 x lo3 to 7.8 x lo5) and
blends over a temperature range of 19 to 143°C. It was found that
the shift factors used in the data reduction were functions of MW.
However, all points fell on the same curve when different reference
temperatures were used. These temperatures were characterized
a s iso-free volume states. Although the transition region measure-
ments for most fractions and blends were superimposed when the
reduction procedure was based on these iso-free volume states,
analysis of the rubbery plateau proved to be much more complex.
A s expected, entanglements extended this region in the higher MW
samples.
Oyanagi and Ferry [215] studied creep under a shearing load for
both plasticized and unplasticized polyvinyl acetate. Two fractions
(MW = 5.34 x lo4, Q = 1.04; MW = 2.3 x lo5, Q = 1.10) were used in
this work. A comparison of the 40°C master curves for the unplas-
ticized materials shows that increasing MW over this range moves
the transition to longer times. The shift amounted to approximately
half a decade near the glassy region. Data included from a previous
investigation indicated that further increases in MW would give
relatively little additional shift. A second notable feature of the
master curves is that the transition region is much less pronounced
166 MARTIN, JOHNSON, AND COOPER

in the lower MW fraction. A s a result, the curves for the two frac-
tions were roughly three decades apart in the rubbery region. Simi-
lar behavior was noted in the nine systems plasticized with diethyl
phthalate although the shifts due to MW were generally smaller.
Creep, creep rupture, and creep recovery of an acrylonitrile-
methylacrylate copolymer (91:9 composition by weight) have also
been examined as functions of temperature and MW [265]. Below
~o'c, creep of the lowest MW sample (Gv = 7.9 x 104) was only
slightly greater than that measured on the other two samples (Gv =
1.16 x lo5 and 1.52 x lo5). However, above this temperature, larger
differences were apparent both in creep and creep recovery. MW
Downloaded by [Princeton University] at 06:30 03 June 2013

effects were also observed in the creep rupture tests because, at a


given temperature, increasing the MW shifted the transition region
of the compliance-at -break curves t o longer times.
Pezzin [221]concluded that the creep compliance of plasticized
P V C was independent of MW for Gw > 4 X lo4. However, he did find
a small MWD effect because broadening the distribution tended to
increase n when the strain-time relationship was expressed in terms
of the Nutting equation: E = ktn.
Creep rupture data at 60°C on high-density polyethylenes indicate
that lowering the melt index substantially increases time-to-failure,
Fig. 82 [168]. Most of the improvement is due to a shift in the brittle-
ductile transition. In this particular study, the test was terminated
and the failure considered to be ductile if the strain exceeded 0.20.
A s expected, reductions in s t r e s s level extend the failure time when

1.0 10 to* I 03 lo4 to5


TIME TO FAILURE IN HOURS

FIg. 82. S t r e s s r u p t u r e of 0.95 density polyethylene a t 60°C in water. The


w a t e r was r e p o r t e d t o have little o r no effect on t h e s e d a t a [168].
MECHANICAL PROPERTIES OF POLYMERS 167

the applied s t r e s s is high enough to cause ductile failure. However,


failure time in the brittle region is relatively insensitive to s t r e s s
level. The effect of per cent crystallinity or density was also con-
sidered in this work, Fig. 83. Although a comparison of Curves 2
and 3 in this figure implies that the lower density resin has a greater
creep life at a given melt index and s t r e s s level, the data a r e not
really comparable because the high-density measurements were
stopped a t 0.20 strain while the low-density tests were carried out
to rupture. Rupture elongation in the latter materials was on the
order of 200%, well beyond the allowable deformation in most appli-
cations.
Downloaded by [Princeton University] at 06:30 03 June 2013

C 0.920 DENSITY

I 0.950850 DsiI
I
II
, 0 2 1
10 I02 I 03 I o4 lo5
TIME TO FAILURE IN HOURS
Fig. 83. Melt index vs. s t r e s s r u p t u r e life a t 60°C in w a t e r [168].

Carey and co-workers presented similar results in earlier papers


[42,44]. The amount or rate of creep was considered to be controlled
by per cent crystallinity while fracture properties were related to
MW. They also noted that the relatively high tensile strengths which
characterize the high-density resin will not be observed in long-term
service unless a low melt index resin is used. With high melt index,
low MW resins, the brittle-ductile transition occurs at relatively
short times. Because of this factor, they found [44] that the long-
term strength of these materials at low stresses was no greater
than that exhibited by low-density polyethylene.
In Ref. 98, Gohn and Cummings updated an earlier paper [99]
which discussed the creep characteristics of low-density polyethyl-
ene. Their tests, which were conducted at various s t r e s s levels for
168 MARTIN, JOHNSON, AND COOPER

periods of 1, 8, and 11 years, show that short-term creep data


(1,000-10,000 hr) cannot be simple extrapolated to longer times.
Even at 40,000 hr, constant creep rates were not observed except at
very low stresses. Typical curves from this test program are in-
cluded in Fig. 84. A s shown in Fig. 85, increases in MW reduced
creep in these compression-molded samples. Creep resistance of
high-density polyethylene as a function of RSV over a 3-1/2 year
period was reported in Refs. 94 and 232. Figure 86 includes some
of these data.
Creep characteristics of two propylene -ethylene block copolymers
indicate that increases in MW will increase rupture strength at a
Downloaded by [Princeton University] at 06:30 03 June 2013

given time [275]. However, this paper also cited unpublished data
which apparently shows that, for low stress levels, low MW mate-
rials can exhibit lower creep rates than their high MW homologs.
D. Vibration Measurements
Measurements from 24 to 2400 Hz on poly-n-butyl methacrylate
( a w = 3.7 x lo5 and 3.05 x lo6)[48] and polyethyl methacrylate (aw=
1.64 x lo5 and 1.73 x lo6) [78] in the rubbery region indicated that
J' and J" were essentially independent of MW at these values of aw.
A s might be expected, increases in MW lengthen the rubbery plateau
region of polydimethylsiloxane [224] but have little effect on the
location of the midpoint in the glass transition of polyisobutylene
and polystyrene [251]. High frequency (10 KHz to 78 MHz) measure-
ments on commercial polydimethylsiloxanes have shown that .the
transition in this material is also insensitive to MW for MW > -lo4
[15]. In vibrating reed and torsional pendulum tests [93], the loss
modulus of poly(norbornene) decreased by a factor of six when [77]
was increased from 3.5 to 9.0 dl/g (benzene, 30°C). This change
appears to be the result of a simultaneous shift in Tg caused by
variations in the ratio of trans-to-cis unsaturation rather than a MW
effect. However, the paper is not clear on this point.
Read 12301 measured the dynamic shear modulus and damping of
five commercial polyethylene oxides (Gw = 4 X lo3 to 5 x los) at
-1 Hz and over a temperature range of -190°C to ambient tempera-
ture. Connor et al. [52] made dynamic mechanical (-1 Hz), dielec-
tric, and NMR measurements on this material (MW = 2 x lo2 to
2.8 x los) from -190 to +6O"C. Both of these studies concluded that
G' increases slightly with MW in the glassy region. Major effects
were found in the G" and log decrement curves, however, because
the maximum in these curves shifted to much higher temperatures
at intermediate MW, Fig. 87. This anomalous behavior was attrib-
uted to restricted chain mobility in the amorphous regions because
Downloaded by [Princeton University] at 06:30 03 June 2013

MECHANICAL PROPERTIES OF POLYMERS


169
170 MARTIN, JOHNSON, AND COOPER
Downloaded by [Princeton University] at 06:30 03 June 2013

0
I02 2 4 6 0 103 2 4 6 8 ,04 2 4 6 0 ,05
ELAPSED TIME, hr

Fig. 85. Effect of melt index on c r e e p of compression molded polyethylene


sheet. Test temperature: 85°F. Stress: 400 psi. Nominal thickness: 0.125 in.
[98 I.

samples in this MW range also had a much higher crystalline content.


The shape and location of the G' , G", and log decrement curves in
the transition zone was largely controlled by the per cent crystal-
linity. This in turn, is a function of MW and thermal history.
The elastic modulus and loss factor of polypropylene fractions
was investigated from -30 to +50°C over a time scale of 20-120sec-'
[252]. Both of these parameters were related to the per cent crys-
tallinity with MW having no noticeable effect. Chu et al. 1491 reached
similar conclusions in their study of fractionated polypropylene
fibers as Young's modulus and tan 6 were primarily affected by
crystallinity, thermal history, and drawing conditions.
E. Internal Losses
Although the moduli and associated losses of amorphous polymers
a r e virtually independent of MW in the glassy state, both the loss
modulus and the loss tangent, tan 6 = G " / G o r E"/E' o r J"/J', a r e
affected by MW and MWD in the rubbery region. HGgberg et al. [120]
showed this dependence with two anionically polymerized samples of
polystyrene (Mw = 1.71 x lo5 and 2.86 X lo5). When tan 6 and G were
plotted against log wat, these functions went through a minimum in
the rubbery region. The minima of the lower MW material were
somewhat sharper and occurred at higher values of frequency, G"
and tan 6. With a broad distribution sample, tan 6min deepened and
MECHANICAL PROPERTIES OF POLYMERS 171
Downloaded by [Princeton University] at 06:30 03 June 2013

TIME (hrr)
Fig. 86. Time-to-break curves for tensile b a r s of low-pressure polyethylene
a t 20, 50, and 80°C. In 80°C water bath: 0 RSV = 2.5-3, RSV = 4-5. 50°C
water bath: A RSV = 2.5-3, A RSV = 4-5. 20°C in air: 0 RSV = 2.5-3, 0 RSV =
4-5 [94].

shifted to a higher frequency. G ’ decreased in magnitude and did not


exhibit a minimum in this material. Theoretical relations were de-
veloped to link MW, tan b i n , and the MW between entanglements.
However, the equations were found to be valid only for narrow dis-
tribution samples. In a similar study Cox e t al. [54lrelated the
minimum in the damping vs. temperature curve to Mn. However, as
in Ref. 120, the equation failed to describe the behavior of a broad
distribution (bimodal) polystyrene. The complex relationship between
tan 6min, MW, and MWD has also been demonstrated on polyvinyl
acetate [214].
Data reduced to 25°C f o r polydimethylsiloxane in the plateau and
172 MARTIN, JOHNSON, AND COOPER
Downloaded by [Princeton University] at 06:30 03 June 2013

D -180 - 140 - 100 -60 -20 20 , D


TEMPERATURE , C
Fig. 87. Variation of G', G", and the log decrement, A , with temperature for
Polyox FC118 ( n v = 2.8 X 10') ( O ) , Polyox FC2075 (% = 8.4 X lo5). ( O ) ,
Polyox 2464 (% = 2.8 Xil"
o 5 ) ( A ) , Carbowax 20M (mv= 3 X l o 4 ) ( 0 ) .Carbo-
wax 4000 ( T n = 1000) ( A ) [52,230].

terminal zones indicated that tan 6 goes through a maximum at a


frequency of -3 x Hz and a minimum at a somewhat lower fre-
quency of -1.5 x 10" [224]. The position of the peaks appeared to be
independent of MW, but decreasing MW raised the level of the curves,
MECHANICAL PROPERTIES OF POLYMERS 173

Fig. 88. As discussed in Section VII-C, the higher MW samples in


this series exhibited quasi-permanent network characteristics.
Most applications for rubbery polymers require cross-linked
materials. Hence, from a practical standpoint, comparison of loss
characteristics a s a function of the original MW of vulcanized elas-
tomers would be of interest. In an early study Johnson [138] showed
Downloaded by [Princeton University] at 06:30 03 June 2013

Fig. 88. Loss tangent values at 25°C vs. logarithm of reduced frequency.
Short-dashed lines, direct f r ee oscillation torsion pendulum measurements;
long-dashed lines, from forced oscillation measurements. (0)Calculated
from measured r e al (J') and imaginary (J") components of the complex
compliance and the viscosities. ( 0 )Computed from J( t ) using relationships
from the literature. ( x ) Using J" values obtained from graphical integration.
) Calculated from the Andrade cr eep model. Long dash on ordinate shows
( * a

low frequency limiting value of 0.577 12241.

that a compounded, high M W SBR fraction had a higher dynamic


modulus and lower loss parameters than the whole polymer, the gel
portion, o r a bimodal mixture of high and low MW material. Limited
service tests on tires indicated that this rough fraction also had
slightly better tread wear than either the whole polymer o r the gel
fraction. The bimodal material was inferior to all three. Both the
174 MARTIN, JOHNSON, AND COOPER

high MW fraction and the bimodal mixture presented severe process-


ing difficulties; one was extremely tough and viscous while the other
had a low viscosity. Bueche [34] derived an expression in which the
heat generation of flexed rubber was proportional to the slope of the
long-time portion of a creep curve. His data on SBR as well as syn-
thetic and natural rubber confirmed this trend, at least qualitatively.
Since the slopes were increased by the presence of low MW material,
the author concluded that low MW components cause the excessive
heat buildup which is observed in some SBR tire formulations. Yanko
[301] came to similar conclusions in his Schopper rebound tests on
vulcanizates compounded from SBR fractions and a whole polymer.
Downloaded by [Princeton University] at 06:30 03 June 2013

The fractions with high values of G n (> 1.9 x lo5 unmilled) had re-
bounds of 58-62%. This is close to typical values measured on
natural rubber. However, even when the cure state was adjusted to
give the same 300% modulus, the unfractionated specimen ( a n -
9.6 x lo4 unmilled) had a rebound of only 38-40%. Low MW fractions
were in this range o r lower. Goodrich hysteresis tests at 100°C gave
similar results for the fractions. However, the whole polymer was
indistinguishable from the higher MW fractions in this test. Resil-
ience of SBR vulcanizates at constant cross-link density has also
been shown to increase with MW [46]. At high cross-link densities
o r at MW > lo5, the MW effects were insignificant.
A secondary loss mechanism a t low reduced frequencies has been
observed in cross-linked polybutadienes [ 1781. The peak was related
to network strand ends because increases in the initial MW and
cross -link density both reduced its magnitude.
Damping in crystalline polymers can usually be correlated with
per cent crystallinity because most of the losses occur in amorphous
regions. This was shown for polypropylene [49,252] and polyethylene
oxide [52,230]. MW effects were also important in the latter material
because the crystallization rate and the per cent crystallinity of this
polyether is particularly sensitive to MW. The location of the maxi-
mum in the damping curve of polyethylene oxide shifted to higher
temperatures when the crystalline content w a s increased. This was
presumably due to restrictions on chain mobility a t the higher crys-
tallinities .
F. Ultrasonic Properties
Many material properties can be measured nondestructively by
using ultrasonic techniques. An example of t hi s is the experiments
conducted on fractionated samples of polystyrene, polymethyl methac-
rylate, and a 50/50 copolymer of styrene and methyl methacrylate
[ 1631. Longitudinal (dilitational) sound velocity, v1, and transverse
MECHANICAL PROPERTIES OF POLYMERS 175

sound velocity, vt of the fractions were measured at 1.67 MHz. From


the smoothed data it was then possible to calculate Young’s modulus
E’, shear modulus G’, bulk modulus K’, and Poisson’s ratio p ’ . These
parameters are slightly dependent upon molecular weight as shown
in Table 16.
Butta and Giusti [37] measured vl and damping on polyisobutylene
at a frequency of -lo4 Hz. Commercial polymers with MW from
6 x lo3 to 3 x lo5 were used and the t e s t temperatures ranged from
-179 to 0°C. As shown in Fig. 89, v1 is temperature dependent and
increases slightly with increases in MW in the glassy and transition
regions for MW < 2.5 X lo5. The lower MW (6 x lo3 to 15 x lo3)
Downloaded by [Princeton University] at 06:30 03 June 2013

samples showed slightly different behavior as the authors detected


two distinct transitions, Fig. 90. The transition temperatures in-
creased with MW. Although the t e r m “glass transition temperature”
has been assigned to the break point in the v1 curve, it s h a l d be
recognized that this curve would shift along the temperature axis if
the test frequency was substantially changed [136].
A conclusion that v1 is only slightly dependent upon MW above
-
MW lo4 has been reached in studies of other low MW polymers
[96,292].

VIII. GLASS TRANSITION TEMPERATURE


When an amorphous polymer is cooled from the rubbery t o glassy
state, it passes through a transition frequently referred to as a
second-order transition. This transition, Tg, occurs when chain
mobility o r segmental jumping frequency is reduced t o such an extent
that the main chain groups become essentially frozen in place. At
least 20 methods have been devised to measure the glass transition
temperature [ 1731 but the results of these different methods often
show poor agreement. Much of the difficulty is caused by the fact
that the location of T g depends on the time scale of the experiment.
Variations in the p e r cent crystallinity, MW, impurities, orientation,
and stereoregularity also account for many of the inconsistencies.
Eisenberg [66] has discussed the multidimensional nature of the
glass transition and defined analogous values of Pg, Cg, and Mg
where these term s r e f e r to the location of the glass transition as a
function of pressure, diluent concentration, and MW.
From both theoretical relationships [70,89] and experimental
results [ 5,18,19,87-89,220,2681 the glass transition-MW relationship
can be expressed in the form:
K
Tg = Tg m - z
Downloaded by [Princeton University] at 06:30 03 June 2013

Table 16
Ultrasonic Mechanical Properties of Fractionated Amorphous Polymers [163]

Longitudinal Transverse Shear Young's Bulk Poisson's


- sound velocity sound velocity modulus modulus modulus ratio
[Sl M~ x lo5 (m/sec) (m/sec) (kg/mm2) (kg/mm2) (kg/mm2) (m/sec)

Polystyrene
1.092 2.512 2370 1200 154 407 376.6 0.32
1.026 2.286 2365 1175 148 393 381.2 0.33
0.780 1.557 2345 1120 134 362 402.6 0.35
0.700 1.365 2340 1100 130 352 414.5 0.36
0.369 0.540 2325 1030 113 315 422.9 0.37

Polymethyl Methacrylate
4.79 12.33 2745 1340 226 606 631.1 0.34
4.00 10.50 2710 1310 206 553 576.2 0.34
Downloaded by [Princeton University] at 06:30 03 June 2013

3.92 10.21 2705 1305 205 54 9 571.7 0.34


3.80 9.886 2700 1300 203 54 5 567.2 0.34
3.33 8.128 2670 1270 194 5 24 581.7 0.35
3.30 8.035 2665 1265 192 519 577.2 0.35
2.58 5.808 2625 1225 183 496 590.0 0.36
2.09 4.416 2600 1200 173 471 560.8 0.36

Polystyrene/PMMA 50: 50 Copolymer


1.84 2750 1370 231 614 601.8 0.33
1.60 2720 1350 2 24 596 584.6 0.33
1.36 2710 1345 223 592 580.2 0.33
1.14 2690 1335 219 583 571.6 0.33
1.10 2685 1330 218 579 567.4 0.33
0.98 2680 1320 214 574 598.1 0.34
0.88 2665 1315 213 566 554.7 0.33
0.72 2650 1310 211 562 550.5 0.33
0.40 2630 1295 206 54 8 537.7 0.33
178 MARTIN, JOHNSON, AND COOPER
Downloaded by [Princeton University] at 06:30 03 June 2013

-180 -160 -140 -120 -100 -80 -60 -40 -20


TEMPERATURE ( O C )

Fig. 89. Sound velocity and internal dissipation as functions of temperature


[371.

An alternate relationship has also been used [89,155,278]:

In these equations Tg.mis the extrapolated glass transition tempera-


ture of a polymer of infinite MW, M is the MW, and K and C are
empirically determined constants. More complex models such as
those based on the Gibbs -DiMarzio statistical theory of supercooled
liquids [97] can often be fitted to a wider range of MW.
Typical values for the constants of Eqs. (1) and (2) a r e included
in Table 17 for polystyrene and PMMA. The discrepancies in the
PMMA data can be attributed to differences in stereoregularity, the
MECHANICAL PROPERTIES OF POLYMERS 179

-
3200 VISTANEX S
M = 9000 - 10,000
3000 -

2800.

2600 -
Downloaded by [Princeton University] at 06:30 03 June 2013

2400 *

2200.

2000 *

- IooooQ-l
h

z
Y

>
=: =: "

I
1
-180 -140 -100 -60 -20

Fig. 90. Sound velocity and internal dissipation of Vistanex S a s a function of


temperature [37 I

indirect methods used to determine MW, and possibly chain branch-


ing in the irradiated samples. The reason for the differences in the
polystyrene constants is less obvious but it can be traced to the
method used to analyze the data in Ref. 278. In this work the author
postulated that polystyrene can be treated as a binary system of end
groups and internal chain groups where the end groups act as plas-
ticizers. A relationship was derived which showed that a plot of
1/Tg vs. the mole fraction of end groups, Ne, should be linear. With
some scatter, this was true, Fig. 91. The derivation assumed that
the free volume of all end groups is the same-a reasonable assump-
tion at high MW but questionable in the telomers. Unfortunately, in
a plot of l/Tg vs.Ne, 75% of the curve is determined by telomers of
DP 5 8 (Ne 2 0.25); data on the materials with D P > 8 is condensed
into the remaining 25%. Thus the effect of any deviations which
Downloaded by [Princeton University] at 06:30 03 June 2013

Table 17
a
Glass Transition Temperature-MW Relationships for Polystyrene and PMMAa 3)
9

MW
Sample Tp" (OK)' K x 10-~ C average Test method Ref.

Polystyrene
-
Fractions 373 1.0 - Mv Dilatometry 88
Fractions 373 1.2 0.864
-
Mv Dilatometry 88
Fractions and -
mixtures 357-361 - 0.515 Mn and mv Dilatometry 278
Anionically
polymerized -
butyl end groups 371 1.02 - MV Dilatometry 5

Polymethyl Methacrylate
Whole polymer, -
atactic 387 2.1 f 0.2 - Refractive index 19 b
Mn D
n
Whole polymer, - r!
atactic - 1.6 f 0.8 - Mn Dilatometry 19 z
Whole polymer, -
irradiated 35 1 - 3.33 MV Dilatometry 155 bI
Whole polymer, z
irradiated, loo'% 2
isotactic 321 1.06 - -
MV DTA 268
D
Whole polymer, z
irradiated, 64% 0

syndiotactic, 36% - 8
atactic 399 3.4 - Mv DTA 268 8rn
n
MECHANICAL PROPERTIES OF POLYMERS 181
Downloaded by [Princeton University] at 06:30 03 June 2013

0 0.5
- Na
1.0

Flg. 91. Second-order transition temperature of p u r e polystyrene fractions


and mixtures of fractions vs. mole fraction of chain end groups. P u r e frac-
tions ( 0 ) .mixtures of fractions with D P = 8 and 39.4 ( X ) , mixtures of frac-
tions with D P = 3.3 and 928 (0),mixtures of fractions with DP = 2 and 48.1
(+), mixtures of fractions with D P = 2 and 222 ( A ) [278].

might be caused by telomer end groups is magnified in this type of


plot. When the curve is redrawn without using the values f o r D P 9 8,
then Tgm = 372°K and C = 0.717 x l o 5 . This is in reasonable a g r e e -
ment with the data of Fox and Loshaek [89] where D P was 2 16 and
is consistent with the results of other workers [2,5,25,87,88,137,159,
196,2991. Nonlinear behavior at low MW has a l s o been found in other
polymers [19,53,220]. It is interesting to note that Evans et al. [73]
made a plot which is analogous to Fig. 91 using the melting tempera-
ture, Tm, instead of Tg. As in Fig. 91, the intercept, Tmm, and the
slope, C , of the straight line were unreasonably low when extremely
low MW products were included in the data.
In other studies T g has been determined as a function of MW f o r
182 MARTIN, JOHNSON, AND COOPER

polyvinyl acetate [159,296], poly(ethy1 acrylate) [295], poly cr-methyl-


styrene [53], PMMA [26 and Table 171, sodium phosphate polymers
[67,681, polyisobuty lene [25 1,277,2971, PVC [220,22 1,3031, atactic
polypropylene [ 1491, polyacrylonitrile [ 18,20,150], polyethylene
terephthalate [ 1581, polydimethylsiloxane [300], polyethylene oxide
[52,76,230], polypropylene oxide [4,75], and polybutylene oxide [75].
Test methods included dilatometry [4,5,19,67,68,87-89,137,155,158,
196,277,278,2991, refractive index [ 18-20,26,295,296], DTA [53,149,
150,268,3031, DSC [220,221,300], dielectric loss [52,159], mechanical
loss [52,75,76,230,251], thermomechanical [303], and NMR [52,76].
When indirect effects were not present, most of the above studies
Downloaded by [Princeton University] at 06:30 03 June 2013

found that Tg was substantially reduced if the MW of the samples fell


below -3 x 104. However, once Gv o r Gw exceeded -lo5, Tg was
generally within 2-4°C of the assumed asymptotic limit. Thus varia-
tions in MWD would be expected to have no effect unless the varia-
tions were due to changes in low MW content. Several investigators
have examined the possibility of using monomers or telomers to
plasticize rigid plastics [2,246,251,276,278], but conventional plasti-
cizers a r e normally considered to be more efficient.
Complications often arise in measurements made on semicrys-
talline materials because the crystallites reduce chain mobility in
the adjacent amorphous regions. For example, an amorphous (p =
1.3390 g/cc) polyethylene terephthalate sample was reported to have
a Tg of 67"C, but when the density was increased to 1.395 g/cc the
transition occurred a t 80°C [158]. The behavior of polyethylene oxide
is also interesting because a maximum in plots of Tg vs. MW is
-
observed at MW 5 x 109 to 30 X lo3 [52,76,230]. This peak, Fig. 92,
is caused by the fact that samples in this intermediate MW range
attain higher crystallinities than other specimens under normal
fabrication conditions. Attempts to quench them in liquid nitrogen
resulted in lower values of Tg but the data were erratic [76]. This
observation confirms the belief that smooth Tg-MW curves would be
observed in these polyethers if the indirect effect due to crystallinity
was eliminated.
Faucher [75] obtained quite different results in his tests of poly-
propylene oxide and polybutylene oxide. Although the MW of these
two materials were varied from 1.34 x lo2 to 4 x 105 and 5 x 102 to
1.7 x lo5, respectively, all of the samples had a glass transition
temperature of -60 * 2°C. A torsional pendulum operating a t 1-3 Hz
was used in these tests. The absence of MW effects was tentatively
ascribed to an assumption that the free volume of end groups and
chain groups was roughly the same for these materials.
MECHANICAL PROPERTIES OF POLYMERS 183
Downloaded by [Princeton University] at 06:30 03 June 2013

0 BROADLINE NMR

I o2 I'
0 lo6 loe
MOLECULAR WEIGHT

Fig. 92. Glass transition temperature vs. molecular weight for polyethylene
oxide [76].

M. MELTING TEMPERATURE
A s noted in Section I1 and in various reviews [Chapters 2 and 8 of
Ref. 181,1821, changes in MW and MWD can affect both the crystal-
lization kinetics and the degree of crystallization observed under a
given s e t of conditions. Crystalline melting temperature, T m , is
also a function of MW, but an asymptotic limit is normally approached
below D P = 200 [Chapter 2 of Ref. 1811.
In most practical applications the thermal and process variables
which control morphology will a l s o determine the T m of a material.
In fact, one early work [259] related the melting point of polyethylene
whole polymers to p e r cent crystallinity and showed no effects of
MW. However, later studies [28,29,74,85,102,183,274] on polyethylene
fractions and the longer n-paraffins have demonstrated MW effects
on the equilibrium melting temperature ( T h ) . This has been d i s -
184 MARTIN, JOHNSON, AND COOPER

cussed theoretically by Flory and V r i j [83] who predicted that the


equilibrium melting temperature of an infinitely long linear poly-
ethylene molecule would be 145.5 f 1°C. Experimentally, T h is
somewhat lower (-139°C). This melting point depression is caused
by crystallite characteristics: the lamallae thickness relative to the
extended chain length of high MW samples is small and the interfacial
free energy is large [83,102,181,183].
The melting temperatwe of polypropylene fractions increases
from 89 to 150°C when a n is raised from 870 to 1 X lo4 [203]. Fur-
ther increases had a much smaller effect as a change in f i v from
3 x 104 to 1 x lo5 only increased Tm from 168 to 175°C. In a similar
Downloaded by [Princeton University] at 06:30 03 June 2013

study [244], broad MWD polypropylene samples (Q = 10) had a slightly


lower Tm than fractions with the s a m e [771.
Rybnikar [242] investigated isothermally crystallized polypro-
pylene and concluded that Tm is not a function of spherulite size,
crystallization time, o r conditions in the melt prior to crystallization.
He also plotted Tm vs. viscosity number for 13 commercial poly-
propylenes and found that the highest MW material had a melting
point which was -10°C above the lowest MW sample. No such MW
dependence was found in another study of commercial polypropylenes
over an zw range of 2.5 x lo5 to 2.4 x los [256].
MW-melting point relationships have also been reported for sev-
e r a l polyesters [45,73,279], low MW syndiotactic PVC fractions
[303], polyethylene oxide [ 1511, and polydimethylsiloxane fractions
[300]. The silicone data showed no substantial MW (MW = 1.02 x l o 5
to 1.13 x lo6) effects on Tm although double melting peaks were ob-
served a t the lower MW. The PVC data went through a minimum a t
E n = 2000. Both of these phenomena a r e presumably related to
differences in crystal structure and per cent crystallinity. In Refs.
73 and 279, 1/Tm was reported to be proportional to l/DP a t low
MW. However, as shown by Evans et al. "731, the data deviate from
linearity a t high MW.

X. MISCELLANEOUS PROPERTIES
A. Hardness
A number of hardness tests have been applied to polymers [104,
1'75,2041. Most of these measure the ability of a material to resist
indentation o r scratching by a rigid body and a r e generally indepen-
dent of MW for MW > -104. This conclusion is based on tests con-
ducted on polystyrene [299], PMMA [26], vinyl chloride-acetate
copolymers [ 56,621, polyethylene [ 1991, polypropylene [252], and
sulfur vulcanized ethylene-butadiene copolymers [ 121. The hardness
of nylon 610 is also independent of MW (Gn = 12,200 and 23,000)
MECHANICAL PROPERTIES OF POLYMERS 185

except at low crystallinities [262]. Hardness measurements on poly-


propylene have been reported to decrease with increasing MW [255]
while tests on polyethylene showed the same trend if the MWD was
reduced [36]. These results a r e thought to be indirect effects of MW
and MWD, however, because both of the changes would also lower
the crystalline content. A s expected, the per cent crystallinity can
be related to hardness in polyethylene [36,259], polypropylene [252],
and nylon [262]. In contrast to the results cited above, the Shore C
hardness of polyethylene, at constant density, and the ball pressure
hardness of polyacetals [lo61 were found to increase with MW. The
data in the latter paper may have been influenced by changes in crys-
tallinity because the samples with high values of GWalso had broad
Downloaded by [Princeton University] at 06:30 03 June 2013

distributions (Table 9),


When the Shore A hardness of elastomers is measured at constant
cross-link density [46] or constant 300% modulus [301], the values
increase at low MW. This is due to the lower MW between cross-
links in these materials.

B. T e a r Strength
Studies on polyimide films of 2-5% crystallinity [291] have indi-
cated that the Elmendorf tear strength (ASTM D-1922-61T) increases
with MW to at least GW= 2 x lo5 without leveling off, Fig. 18. When
plotted against Gn, a limiting value appears to be reached at G n
6 x 104, Fig. 19. A s shown in Fig. 20, narrowing the MWD greatly
-
increases the tear strength when comparisons a r e made at constant
[d.
-way [228] conducted a statistical study of the physical and
chemical properties of bleached sulfate chestnut pulp. His data show
a close relationship between Elmendorf tear strength and D P because
the correlation coefficient for this set of parameters was 0.906 over
a DP range of 625 to 990.
Less extensive data on regenerated cellulose film [ 1971 and low-
density polyethylene [199] also indicate that increases in MW gener-
ally increase tear strength. Somewhat different results were obtained
in studies of cellulose nitrate [249] and ethyl cellulose [248] frac-
tions because the Elmendorf tear strength (ASTM D869-44) of these
materials leveled off when Ew exceeded 100-150. An MWD effect
was also noted when FPpwwas near o r below this limiting range.

C. Burst Strength
Correlations between burst strength and MW are rather sparse.
Data for polystyrene fractions (aw = 3 x lo5 to 1.46 x los) indicated
that burst strength was not affected by MW over this range [24]. This
-
186 MARTIN, JOHNSON, AND COOPER

result was also obtained for bleached sulfate chestnut pulp when D P
was varied from 625 to 990 [228]. However, the burst strength of
various cellulose fabrics reportedly decreased a s the D P fell [63].
Scherer and co-workers [248,249]found that the Muller burst strength
of cellulose nitrate and ethyl cellulose fractions decreased when
mW was reduced below -100 and -150, respectively. Blends of
these fractions were related to a shape factor which was defined in
terms of MW and MWD. Limited data by Belov et al. [21]showed
that two bimodal narrow distribution polyethylenes had much higher
burst strengths than a low MW comparison sample which had an
unusually broad MWD (Q = 290). Although the MW averages of the
Downloaded by [Princeton University] at 06:30 03 June 2013

two bimodal samples were considerably different, their burst


strengths were almost identical. Thus it would appear that burst
strength is not grossly affected by MW except when large amounts of
low MW material is present.
D. Heat Distortion and Softening Temperatures
A number of heat distortion and softening tests have been used a s
guidelines to fix temperature limits and to evaluate orientation char-
acteristics in polymers [204]. Usually the heat distortion tempera-
ture occurs close to Tg in amorphous materials and Tm in the highly
crystalline polymers. Since neither of these transitions is markedly
affected by MW, one would expect that the heat distortion tempera-
tures would also show little MW dependence above MW -104. Early
tests on vinyl chloride-acetate copolymers [ 56,621 confirmed this
belief. The softening temperature of polyethylene can either decrease
[Chapter 7 of Ref. 231 o r increase [113,235,259]slightly with MW. For
example, a reduction in the melt index from 5 to 0.01 increased the
Vicat softening point from 120 to 127°C [113]. When annealed poly-
styrene fractions were tested according to ASTM D648 [299],the 66
psi heat distortion temperature was increased from 203 to 215°F
when the MW was raised from 8.1 x lo4 to 5.39 X lo5. A broad dis-
tribution sample also fell within this range but two polystyrene
tetrachain fractions (Gv = 2.04 X lo5 and 4.3 x lo5) had values of
219 and 220°F. Merz et al. [196]presented similar data on polysty-
rene fractions a s did Rudd [239].Data from this latter paper show
small increases in the heat distortion temperature when the MW is
raised and the MWD is reduced, Table 11. No MW effects were noted
in tests conducted on polystyrene and polydichlorostyrene whole
polymers [ 1801.
E. Wear and Abrasion
Friction and wear a r e interrelated in that friction is invariably a
part of any abrasive process. Normally, sliding friction is consid-
MECHANICAL PROPERTIES OF POLYMERS 187

ered to be a combination of two mechanisms: an adhesion term which


involves a shearing of the junctions which are formed at regions of
intimate contact and a ploughing term where the harder material
scratches the mating surface. Since this process involves the s u r -
faces of two bodies, relative material properties a r e more important
than absolute properties [204]. Most workers have correlated sliding
friction in terms of hardness, modulus, shear strength, softening
temperature, molecular adhesion, and surface energy but, in general,
agreement between different instruments and investigators is quite
poor because of the complexity of the process. When one considers
rolling friction, viscoelastic properties assume major importance
Downloaded by [Princeton University] at 06:30 03 June 2013

[ 177,2041.
Aside from behavior near t h e transition temperatures, most of
the properties listed above'are relatively insensitive to MW. Hence,
abrasion should not be grossly affected by changes in friction char-
acteristics if the MW or MWD of a material is varied. At low s u r -
face velocities the adhesion term may be an exception however. A
number of papers [229,280,288,289] have shown that adhesion between
some mating surfaces increases as MW is decreased if measure-
ments are made at constant time and temperature. This trend is due
to t h e effect of MW on the diffusion rate of chain segments.
Although friction is a part of any abrasion process, other factors
can be equally important. In various applications wear o r abrasion
has been expressed in terms of resistance to tearing and cutting,
hardness, stress-strain relations, chemical and thermal stability,
position relative to the glass transition and melting temperatures,
and viscoelastic properties such as resilience [47,95,204]. Obviously,
the relative importance of these parameters will be quite sensitive
to the particular service conditions s o good correlations between
fundamental material properties and wear resistance a r e rare.
Direct studies between wear rates and MW a r e almost nonexist-
ent. In an early work the abrasion resistance of low-density poly-
ethylene decreased with increasing MW [199]. The opposite trend
was found in limited tests on isotactic polypropylene [ 1441 while
measurement on two sulfur vulcanized ethylene -butadiene copolymers
showed no conclusive MW effect [12]. In a service test on t i r e com-
pounds, six cold SBR polymers were extended with varying amounts
of a highly aromatic oil, compounded, and extruded into an experi-
mental tread [58]. Relative wear ratings were then compared with
oil content and MW (expressed in terms of Mooney viscosity). The
results showed that increases in the MW and decreases in the oil
extension both improved the wear rating. However, the MW effect
was surprisingly small because a 25-point decrease in the Mooney
viscosity of the unextended polymer typically reduced the tread wear
188 MARTIN, JOHNSON, AND COOPER

rating of the extended material by 5% at a constant oil level. This


small but noticeable relationship between MW and the tread wear of
SBR tires was also observed by Johnson [138]. In addition, he found
that skid loss resistance in natural rubber tires was improved by
using higher MW stock.

Acknowledgments
It is a pleasure to acknowledge the courtesy of the following pub-
lishers, journals, and authors f o r permission t o reproduce the
designated figures:
Downloaded by [Princeton University] at 06:30 03 June 2013

Academic Press, publishers of Journal of Colloid Science, Figs. 88


and 91.
The Academy of Sciences of the Latvian SSR, publishers of
Mekhanika Polimerov, Figs. 1 and 2.
The Academy of Sciences of the White Russian SSR, publishers of
Doklady Akademii Kauk Belorusskoi SSR, Fig. 53.
The American Chemical Society, publishers of Journal of the
A m w i c a n Chemical Society, Fig. 55; Polymer Preprints, Figs. 18-
21;Industrial and Engineering Chemistry, Figs. 12, 13, 16, 36, 37,
and 54; and Journal of Physical Chemistry, Figs. 78 and 79.
The American Institute of Physics, publishers of Journal of
Applied Physics, Fig. 92;and Journal of Chemical P h y s i c s , Fig. 75.
The American Society for Testing Materials, publishers of ASTM
Bulletin, Figs. 84 and 85;and Materials Research and Standards,
Fig. 4; a s well a s sponsor of the 65th Annual Meeting of the ASTM,
Figs. 82 and 83.
ARTIA, publishers of Plaste Hmoty Kaucuk, Figs. 14 and 15.
Dr. C. W. Bondurant, curves from his Ph.D. thesis, Figs. 8 and 9.
Consultants Bureau, a Division of Plenum Publishing Corporation,
publishers of Colloid Journal (a translation of Kolloidnyi Zhurnal),
Fig. 73.
Elsevier Publishing Company, publishers of Elastomers and
Plastomers, Fig. 48.
C arl HanserVerlag, publishers of Kunststoffe, Figs. 51, 63, and
86.
Heywood-Temple Industrial Publications, Ltd., publishers of
Plastics (London), Figs. 49, 50, 82, and 83.
INAC, publishers of Ricerca Scientvica, Figs. 89 and 90.
Interscience Publishers, a division of John Wiley & Sons, Inc.,
publishers of Journal of Applied Polymer Science, Figs. 6, 17, 26,
38, 41, 42, and 72; Journal of Polymer Science, Figs. 5, 7A, 10, 11,
29, 32-35A, 58, 64-68 and 81.
MECHANICAL PROPERTIES OF POLYMERS 189

IPC Science and Technology, Ltd., publishers of P o l y m e r , Figs.


47 and 87.
National Bureau of Standards, publishers of Journal of Research
of the National Bureau of Standards, Figs. 16, 36, and 37.
Doctors Donald J . Plazek and Michael O’Rourke, unpublished
material presented a t the Polymer Conference Series, Wayne State
University, Figs. 77 and 80.
Tokyo Kogyo Daigaku, publishers of Sen-i Gakkaishi, Figs. 27 and
28.
The Society of Chemical Industry, publishers of Journal of Applied
Chemistry, Figs. 3, 22, 39, 44, 56, and 87;and Physical Properties
Downloaded by [Princeton University] at 06:30 03 June 2013

of P o l y m e r s , SCI Monograph No. 5, Figs. 45 and 52.


The Society of Plastics Engineers, publishers of SPE Journal,
Figs. 43, 57, and 74;SPE Tech P a p e r s , Figs. 59-62;SPE Transac-
tions, Fig. 46;and Polymer Engineeringand Science, Figs. 4 and 69-71.
The Society of Polymer Science, Japan, publishers of Kobunshi
Kagaku, Fig. 23.
The Textile Research Institute, publishers of Textile Research
Journal, Figs. 24 and 25.
The Textile Institute, publishers of the Journal of the Textile
Institute, Figs. 30, 31, and 40.
Van Nostrand Reinhold Company, publishers of Mechanical Prop -
e r t i e s of P o l y m e r s , Fig. 76.

The assistance of a number of translators, particularly Miss


Barnee M. Escott, M r . Richard F . Freyer, and Mr. Jau-Yi Jwang,
is gratefully acknowledged
This work was supported in part by the University of Connecticut
Research Foundation and in part by the National Science Foundation
through Grant G P 28613.

References

[ I ] G. Akovali, J . P o l y m . S c i . , Part A - 2 , 5 , 875 (1967).


[ 2 ] A. P. Alexandrov and J. S. Lazurkin, C . R . (Dokl.) Acad. Sci. Z’URSS,
43, 376 (1944).
(31 R. M. Aliguliev, S. A. Abasov, A. S. Dzhafarov, and V. G. Abashin, Plast.
Massy, 1969(6), 41.
[4 ] G. Allen, Techniques of Polymer Science (SCI Monograph) No. 17), SCI,
London, 1963, pp. 167-174.
[ 5 ] T. Altares, Unpublished Results cited in Refs. 223 and 225.
[ 6 ] F. R. Anderson, J. Polym. S c i . , Part C , 3, 123 (1963).
171 F. R. Anderson, J. Appl. Phys., 35, 64 (1964).
[ 8 ] R. J. Anderson and S. R. Melvin,SPE Tech. Pap., 4 , 940 (1958).
190 MARTIN, JOHNSON, AND COOPER

[ 9 ] E. H. Andrews, Fracture in P o l y m e r s , Oliver and Boyd, Edinburgh and


London, 1968.
[ l o ] R. D. Andrews and A. V. Tobolsky, J. Polym.Sci., 7 , 221 (1951).
[ l l ] J. R. Aspen, SPE J., 2 1 ( 7 ) , 643 (1965); 2 1 ( 8 ) , 745 (1965).
[12] F. W. Bailey, C. F. Leslie, D. R. Witt, and J. E. Prichard, SPE J., 2 2 ( 8 )
93 (1966).
[ 1 3 ] G . W. Bailey, J. Polym. Sci., 62, 4 1 (1962).
[14 ] E. I. Barg, D. M. Spitkovskii, and N. N. Mel’t’yeva, Dokl. Akad. Nauk
SSSR, 84, 257 (1952).
[151 A. J. Barlow, G . Harrison, and J. Lamb, P r o c . Roy. Soc., Ser. A , 282,
228 (1964).
[16] G. M. Bartenev, A. S. Novikov, and F. A. Galil-Ogliy, Kolloid. Zh., 18,
Downloaded by [Princeton University] at 06:30 03 June 2013

7 (1956).
(17 ] G. M. Bartenev and Yu. S. Zuyev, Strength and Failure of Viscoelastic
Materials, Pergamon, New York, 1968.
[ l H ] R. B. Beevers, J. Polym. Sci., Part A , 2 , 5257 (1964).
[19]R. B. Beevers and E. F. T. White, Trans. Faraday Soc., 56, 744 (1960).
[20] R. B. Beevers and E. F. T. White, Trans. FaraduySoc., 56, 1529 (1960).
[21 ] G. P. Belov. S. K. Kaltochikhina, L. I. Atanasova, V. I. Tsvetkova, and
I. M. Chirkov, Zzv. Akad. Nauk SSSR, Ser. Khim., 1966, 1275.
[ 2 2 1 J. P. Rerry, J. Polym. Sci., Part A , 2 , 4069 (1964).
[23 ] H. V. Boenig. Polyolefins : Structure and Properties, Elsevier,
Amsterdam and New York, 1966.
[24 ] C. W. Bondurant, Ph.D. Thesis, Virginia Polytechnic Institute, 1960;
University Microfilms, Order No. 62- 1150.
[25] J. Brandrup and E. H. Immergut, eds., PoZymerHandbook, Wiley (Inter-
science), New York, 1966.
[26] G. M. Brauer and W. T. Sweeney, Mod. Plast., 3 2 ( 9 ) , 138 (1955).
[27] J. C. Bray and H. B. Hopfenberg, J. Polym. S c i . , Part B , 7, 679 (1969).
[28] M. G. Broadhurst,J. R e s . Nut. Bur. Stand.,A, 88, 241 (1962).
[29] M. G . Broadhurst, J. Chem. Phys., 36, 2578 (1962).
[ 3 0 ] G. M. Brown and A. V. Tobolsky, J. Polym. Sci., 8 , 1 6 5 (1951).
[ 3 1 ] W. M. D. Bryant and R. C. Voter,J. A m e r . Chem. Soc., 7 5 , 6113 (1953).
[32] R. Buchdahl, L. E. Nielsen, and E. H. Merz, J . Polym. Sci., 6, 4 0 3 (1951).
[33] F. Bueche, J. Appl. P h y s . , 26, 738 (1955).
[34 1 F. Bueche, J . Polym. Sci., 25, 305 (1957).
(351 A. A. Buniyat-Zade and A. B. Azimova,Plast. Massy, 1970(6), 37.
[36] G. N. B. Burch, G. B. Field, F. H. McTigue, and H. M. Spurlin, SPE J.,
13, 34 (1957).
[37] E. Butta and P. Giusti, R i c . Sci., Parte 2 : S e z . A , 2 , 362 (1962).
[38]M. J. R. Cantow, ed., Polymer Fractionation, Academic, New York,
1967.
(391 V. Cappuccio, A. Coen, F. Bertinotti, and W. Conti, Chim. Ind. (Milan),
44, 463 (1962).
[401 R. H. Carey, ASTM Bull., 167, 56 (1950).
(41 ] R. H. Carey, SPE J . , 10(3), 1 6 (1954).
MECHANICAL PROPERTIES OF POLYMERS 191

1421R. H. Carey, Znd. Eng. Chem., 50, 1045 (1958).


(431 R. H. Carey, E. F. Schulz, and G. 6. Dienes, Znd. Eng. Chem., 42, 842
( 1950).
144 ] R. H. Carey, J. A. Snyder, and H. C. Vakos, Wire Wire P r o d . , 32, 998
(1957).
[45] W. H. C a r o t h e r s and F. J. Van Natta, J . A m e r . C h e m . S O C . , 55, 4714
( 1933).
[461 J. Cern?, Plust. Hmoty Kuuc., 3(4), 103 (1966).
[47] C. K. Chatten and A. Thiruvengadam, in Testing of Polymers, Vol. 3
(J. V. Schmitz and W. E. Brown, eds.), Wiley (Interscience), New York,
1967, pp. 245-263.
[48] W. C. Child and J. D. F e r r y , J . Colloid Sci., 12, 327 (1957).
Downloaded by [Princeton University] at 06:30 03 June 2013

[49] H.-D. Chu, R. Kitamaru, and W. Tsuji, Bull. Znst. Chem. R e s . , Kyoto
Univ., 43, 193 (1965).
[50] P. L. Clegg, S. T u r n e r , and P. I. Vincent, Plastics (London), 24, 31
(1959).
[51] M. Compostella, A. Coen, and F. Bertinotti, Angew. Chem., 74, 618
( 1962).
[ 5 2 ] T. M. Connor, B. E. Read, and G. Williams, J . Appl. Chem., 14, 74
(1964).
[ 5 3 ] J. M. G. Cowie and P. M. Toporowski, Eur. P o l y m . J . , 4 , 621 (1968).
[ 5 4 ] W. P. Cox, R. A. Isaksen, and E. H. Merz, J . Polym. S c i . , 44, 1 4 9 (1960).
[55] R. J . E. Cumberbirch and W. G. Harland, J . T e x t . Znst., T r a n s . , 50, 311
(1959).
[56] G. 0. C u r m e and S. D. Douglas, Znd. Eng. Chem., 28, 1123 (1936).
[57] E. T. Darden and C. F. H a m m e r , Muter. Methods, 44, 94 (1956).
[58] H. K. d e Decker, D. D. Dunnom, and C. A. McCall, Rubber Age (New
York), 101(5), 5 3 (1969).
[v] J. B. DeCoste, F. S. Malm, and V. T. Wallder, Ind. Eng. Chem.. 43, 117
(1951).
[SO] H. Determann, Gel ChrumutograPhy, Springer, New York. 1968.
[ 6 1 ] E. R. Dixon and J. B. Jackson, J. Muter. S c i . , 3. 464 (1968).
[62] S. D. Douglas and W. N. Stoops, Znd. Eng. Chem., 28, 1152 (1936).
[ 6 3 ] N . D r i s c h and L. Soep, T e x t . R e s . J., 23, 513 (1953).
1641 W. A. Dukes, B r i t . P l a s t . , 3 4 ( 3 ) , 123 (1961).
[65] W. L. Dunkel and R. A. Westlund, SPE J . , 16, 1039 (1960).
[66] A. Eisenberg, J . Phys. Chem., 67, 1333 (1963).
[67] A. Eisenberg, Fortschr. Hochpolym. Forsch., 5 , 59 (1967).
[ 6 8 ] A. Eisenberg and T. Sasada, in Physics of Non-crystalline Solids (J. A .
P r i n s , ed.), North Holland, A m s t e r d a m , E l s e v i e r , New York, 1965, pp.
99-116.
169) F. E l b e r s and F. Fischer. Kunststoffe, 5 0 ( 9 ) , 485 (1960).
[ 7 0 ] S. M. Ellerstein, J . Polym. S c i . , Part B . 2, 379 (1964).
[ 7 1 ] W. C. Ellis and J. D. Cummings, ASTM Bull ., 178, 4 7 (1951).
(721 G. A. Ermilova, I. A. Ragozina, and N. M. Leont’eva, Plust. Mussy,
1 9 6 9 ( 5 ) , 52.
192 MARTIN, JOHNSON, AND COOPER

[73]R. E. Evans, H. R. Mighton, and P. J. Flory, J. A m e r . Chem. Soc., 72,


2018 (1950).
[74]J. G. Fatou and L. Mandelkern, J. Phy s. Che m., 69, 417 (1965).
(75)J. A. Faucher, J. Polym. Sc i. , Part B , 3 , 143 (1965).
[76]J. A. Faucher, J. V. Koleske, E. R. Santee, J. Stratta and C. W. Wilson,
111, J . Appl. P hy s. , 37, 3962 (1966).
(771 J. D. F e r r y , Viscoelastic Properties of Poly mers, 2nd ed., Wiley,
New York, 1970.
[78]J. D. F e r r y , W. C. Child, R. Zand, D. M. Stern, M. L. Williams, and
R. F. Landel, J. Colloid Sci., 12, 53 (1957).
[79]J. D. Ferry, R. F. Landel, and M. L. Williams, J . Appl. Phys., 26, 359
(1955).
Downloaded by [Princeton University] at 06:30 03 June 2013

[80]D. V. Fil’bert, G. N. Mol’koua, and A. B. Pakshuer, Khim. Volokna,


1965(5), 6.
I811 P. J. Flory, J. A m e r . Chem. Soc., 67, 2048 (1945).
[82]P. J. Flory,Znd. Eng. Chem., 38, 417 (1946).
[83]P. J. Flory and A. Vrij, J. A m e r . Chem. SOC., 86, 3548 (1963).
[84]V. L. Folt, SPE T r a n s . , 2(4), 285 (1962).
[85]R. W. Ford, J. D. Ilavsky, and R. A. Scott, in Analytical Calorimetry
(R. S. P o r t e r and J. F. Johnson, eds.), Plenum, New York, 1968, pp. 41-
44.
(861J. P. Forsman, H. C. North, and T. H. Hakala, Polym. Eng. Sci., 6, 145
(1966).
[87]T. G Fox and P. J. Flory, J. Appl. Phy s. , 21, 581 (1950).
[ 8 8 ] T. G Fox and P. J. Flory, J. Poly m . Sc i. , 14, 315 (1954).
[89]T. G Fox and S. Loshaek, J. Polym. S c i . , 15, 371 (1955).
[go] K. Fujioka, J . Appl. Polym. Sc i. , 13, 1421 (1969).
[911G. E. Fulmer, Trans. S O C . Rheol., 9(2), 121 (1965).
[92]G. E. Fulmer, Trans. SOC. Rheol., 10(2), 501 (1966).
[93]I. Galperin, J. H. Carter, and P. R. Hein, J. Appl. Polym. Sci., 12, 1751
(1968).
[94]E. Gaube, Kunststoffe, 49(9), 446 (1959).
[95] F. M. Gavin, in Testing of Poly m e rs, Vol. 3 ( J . V. Schmitz and W. E.
Brown, eds.), Wiley (Interscience), New York, 1967, pp. 139-201.
[96] V. I. G6czy. Kolloid-Z., 201(2), 99 (1965).
[97] J. H. Gibbs and E. A. DiMarzi0.J. Chem. P h y s . , 28, 373 (1958).
[98] G. R. Gohn and J. D. Cummings, ASTM B u l l . , 247, 64 (1960).
1991 G. R. Gohn, J. D. Cummings, and W. C. Ellis, A m e r . SOC.T e s t . M a t e r . ,
P r o c . , 49, 1139 (1949).
[loo] J. H. Golden, B. L. Hammant. and E. A. Hazell, J. Polym. Sci., Part A ,
2, 4787 (1964).
[loll J. H. Golden, B. L. Hammant, and E. A. Hazell, J . A p p l . Polym. Sci.,
12, 557 (1968).
[lo21 M. Gopalan and J. Mandelkern, J. Phy s. Chem., 71, 3833 (1967).
MECHANICAL PROPERTIES OF POLYMERS 193

[ l o 3 1 J. M. Goppel and A. K. van d e r Vegt, P r o c . Discuss. Znt. P l a s t . Congv.,


3 r d , 177-187 (1966).
[ l o 4 1 J. J. Gouza, in Testing of P o l y m e r s , Vol. 2 (J. V. Schmitz, ed.), Wiley
(Interscience), New York, 1966, pp. 225-278.
[lo51 H. Grimminger, Kanststoffe, 57, 4 9 6 (1967).
[ l o 6 1 H. Grohn and H. Friedrich, Plaste Kaut., 14, 795 (1967).
[ l o 7 1 T. W. Haas and P. H. MacRae, SPE J . , 24, 27 (1968).
[ l o 8 1 T. W. Haas and P. H. MacRae,SPE Tech. Pap., 15, 1 6 (1969).
[lo91 R. S. Hagan, D. P. Thomas, and W. R. Schlich, P o l y m . Eng.Sci., 6,
373 (1966).
[110] C. W. Hamilton and M. M. Epstein, Mod. P l a s t . , 36(8), 150 ( 1 9 5 9 ) .
[ill] H. C. Hass and D. I. Livingston, J . Polym. Sci., 17, 135 (1955).
Downloaded by [Princeton University] at 06:30 03 June 2013

[112] N. Hata and A. V. Tobolsky, J . Appl. Polym. S c i . , 12, 2597 (1968).


[113] R. N. Haward, B. Wright, G. R. Williamson, and G . J. Thackray, J .
Polym. Sci., Part A , 2 , 2977 (1964).
[114] J. H. Heiss and V. L. Lanza, Wire Wire Prod. 33. 1182 (1958).
11151 J. H. Heiss, V. L. Lanza, and W. M. Martin, Wire Wire P r o d . , 34, 592
(1959).
[116] J. N. Herman and J. A. Biesenberger, Polym. Eng. Sci., 8, 341 (1966).
(1171 T. Hinton and A. Keller, J. Appl. Polym. S c i . , 13, 745 (1969).
[I181 P. Hittmair, H. F. Mark, and R. Ullman, J . Polym. S c i . , 33, 5 0 5 (1958).
[119] P. Hittmair and R. Ullman, J. Appl. Polym. S c i . , 6 , 1 (1962).
[120] H. Hogberg, S. E. Lovell, and J. D. Ferry, Acta Chem. Scand., 14, 1424
( 1 960).
(1211 K. E. Van Holde and J. W. Williams, J. Polym. Sci., 11, 243 (1953).
[122] H. B. Hopfenberg, Ph.D. Thesis, Dept. Chem. Eng., Massachusetts
Institute of Technology, 1964.
[123] I. L. Hopkins, W. 0. Baker, and J. B. Howard, J . A p p l . P h y s . , 2 1 , 206
(1950).
(1241 R. A. Horsley, D. J. A. Lee, and P. B. Wright, The Physical Properties
of Polymers (SCI Monogr. No. 5), SCI, London, 1959, pp. 63-79.
[125] K. Hosoda, Kogyo Kagaku Zasshi, 64, 1869 (1961).
(1261 R. Houwink, ed., Elastomers and Plastomers, Vol. 1, Elsevier.
Amsterdam and New York, 1950.
[127] J. B. Howard, SPE J., 15, 397 (1959).
(1281 J. B. Howard, in Crystalline Olefin P o l y m e r s , P a r t I1 (R. A. V. Raff
and K. W. Doak, eds.), Wiley (Interscience), New York, 1964, pp. 47-103.
[129] J. B. Howard, in Engineering Design for Plastics, (E. Baer, ed.),
Reinhold, New York, Chapman and Hall, London, 1964, pp. 742-794.
[130] J. B. Howard, SPE Trans., 4 ( 3 ) , 217 (July 1964).
[131] J. B. Howard, Polym. Eng. Sci., 5, 1 2 5 (1965).
11321 J. B. Howard and H. M. Gilroy,SPE J., 24, 68 (1968).
[133] T. Huff, C. J. Bushman, and J. V. Cavender, J . Appl. Polym. Sci., 0,
825 (1964).
194 MARTIN, JOHNSON, AND COOPER

[134] C. S. Imig, SPE Tech. P a p . , 4 , 934 (1958).


[1:15] R. A. Isaksen, S. Newman, and R. J. C l a r k , J . Appl. Polym. Sci., 7,
515 (1963).
[136] D. G. Ivey, B. A. Mrowca, and E. Guth,Rubber Chem. Technol., 23,
172 (1950).
[137] E . Jenckel and K. Ueberreiter, 2.Phys. Chem., Abt. A , 182, 361
(1938).
[138] B. L. Johnson, Znd. Eng. Chem., 40, 351 (1948).
[139] J. F. Johnson and R. S. P o r t e r , P r o p . Polym. Sci., 2, 201 (1970).
11401 D. F. Kagan and P. F. Lokshin, Plast. Massy, 1966(8), 65.
[141] M. Kalfus, J. Kopytowski. S. L e h i a k , and 2. Skupinska, Polimery, 9(2),
55 (1964).
Downloaded by [Princeton University] at 06:30 03 June 2013

(1421 R. P. Kambour, PoEym. Sci. Eng., 8, 281 (1968).


[143] K. Kamide. Y. Inamoto, and K. Ono, Sen-i Gakkaishi, 23, 269 (1967).
[144] K. Kamide. S. Mochizuki, and S. Ikeda, Sen3 Gakkaishi, 23, 578 (1967).
[145] A. Kaminska, Polimery, 9 ( 2 ) , 4 8 (1964).
[146] A. N. Karasev, I. N. Andreeva, N. M. Domareva, K. I. Kosmatykh,
M. G. Karaseva, and N. A. Domnicheva, Vysokomol. Soedin., Ser. A , 12,
1127 (1970).
11471 K. A. Kaufman, Mod. Plast., 3 6 ( 7 ) , 146 (1959).
[148] K. A. Kaufman and C. S. Imig, Mod. Plast., 3 6 ( 6 ) , 1 3 7 (1959).
[149] B. Ke, J. Polym. Sci.,Part B , 1, 167 (1963).
(1501 J. J. Keavney and E. C. Eberlin, J . Appl. Polym. S c i . , 3, 47 (1960).
[151] H. D. Keith and F. J. Padden, Jr., J . Appl. Phys., 35, 1270 (1964).
[152] H. D. Keith, F. J. Padden, Jr.. and R. G. Vadimsky, J.Polym. Sci.,
Part A-2, 4 , 267 (1966).
[153] P. P. Kelly and T. J. Dunn,Mater. Res. Stand., 3, 545 (1963).
[154] A. S. Kenyon, I. 0. Salyer, J. E. Kurz, and D. R. Brown, J . Polym.
Sci., Part C , 8 , 205 (1965).
[155] N. Kishi and K. Kamata, Oyo Butsuri, 3 0 ( 6 ) , 411 (1961).
[156] H. Kobayashi, K. Sasaguri, Y. Fujisaki, and T. Amano, J. Polym. Sci.,
Part A , 2, 313 (1964).
[157] H. Kojima and K. Yamaguchi, Kobunshi Kagaku, 19, 715 (1962).
[158] H. J. Kolb and E. F. Izard, J . AppZ. Phys., 20, 564 (1949).
[159] S. N. Kolesov, Vysokomol. Soedin, Ser. A , 9, 1860 (1967).
(1601 J. Kopytowski, M. Kalfus, 2. Skupinska. and S. Leiniak, Polimery, 10(2),
55 (1965).
[161] H. Krtissig, Chemiefasem, 17, 821 (1967).
[162] H. Krassig and W. Kitchen, J . Polym. Sci, 51, 123 (1961).
[163] J. Kriauciunas and R. Ziemys, Liet. TSR Mokslu Akad. Darb, Ser. B ,
1963(1), 8393.
[161] L. A. Laius and E. V . Kuvshinskii, Vysokomol. Soedin., 3 ( 2 ) , 215
(1961).
[165] L. A. Laius and E. V . Kuvshinskii. Mekh. Polim., 1967, 455.
[166] L. A. Laius and E. V . Kuvshinskii, Mekh. Polim., 1967, 579.
MECHANICAL PROPERTIES OF POLYMERS 195

[1671 L. L. L a n d e r , SPE J., 16,1329 (1960).


11681 L. L. L a n d e r a n d R. H. C a r e y , Plastics (London), 28 (310). 87 (1963).
[I691 V. V. Lapshin, N. A. P o s p e l o v a , a n d V. N. G r i n b l a t , P l a s t . M a s s y .
1966(2). 20.
[1701 E. J. Lawton, J. S. Balwit, and A. M. Bueche, 2nd. Eng. Chem., 46,
1703 (1954).
[1711 H. L e a d e r m a n , R. G. Smith, and R. W. J o n e s , J . Polym. S c i . . 14,47
(1954).
[I721 H. L e a d e r m a n , R. G. Smith, and L. C. W i l l i a m s , J . Polym. Sci., 36,
233 (1959).
11731 W. A. Lee and G. J. Knight, in Polymer Handbook (J. B r a n d r u p a n d
E. H. I m m e r g u t , eds.) , Wiley ( I n t e r s c i e n c e ) , New Y o r k , 1966,pp. 111-61
Downloaded by [Princeton University] at 06:30 03 June 2013

t o 111-91.
[174] W. H. Linton and H. M. Goodman, J . Appl. Polym. Sci., 1, 179 (1959).
[I751 D. I. Livingston, in Testing of Polymers, Vol. 3 ( J . V. S c h m i t z and
W. E. Brown, eds.), Wiley ( I n t e r s c i e n c e ) , New Y o r k , 1967, pp. 111-138.
[176] E. G. Lovering, J . Polym. Sci., Part A - 2 , 8 , 747 (1970).
[I771 K. C. L u d e m a and D. T a b o r , W e a r , 0 ( 5 ) , 329 (1966).
I1781 E. Maekawa, R. G. Mancke, a n d J. D. F e r r y , J . P h y s . Chem., 69, 2811
( 196 5 ) .
[179] E. Maekawa, G. Y a s u d a , K. Ninomiya, and S. Ido, Nippon Gomu
Kyokaishi, 36, 714 (1963).
[I801 I. A. Maigel’dinov, A. V. G r i g o r ’ e v a , and K. I. T s y u r . Plast. M a s s y ,
1961(3), 7.
[I81] L. Mandelkern, Crystallization of Polymers, McGraw-Hill, New Y o r k ,
1964.
[ 1 8 2 ] L. Mandelkern, A m e r . Chem. SOC., Div. Polym. Chem. Preprints,7(1),
35 (1966).
[183] L. Mandelkern, J. Polym. S c i . , Part C , 15, 129 (1966).
[184] L. Mandelkern, J. G. F a t o u , and K. Ohno, J. Polym. Sci., Part B , 6.
615 (1968).
[185] H. F. M a r k , in Cellulose and Cellulose Derivatives High P o l y m e r s ,
Vol V). ( E . Ott a n d 13. M. S p u r l i n , eds.), I n t e r s c i e n c e , New Y o r k and
London, 1943,pp. 1007-1009.
[I861 M. J. Martynyuk, Z . A. Machionis, B. V. E r o f e e v , a n d V. K.
Semenchenko, Dokl. A k a d . Nauk B e l o m s s . SSR, 7(3), 170 (1963).
[187] S. Matsuoka, Polym. Eng. Sci., 5(3), 142 (1965); S. Matsuoka and
F. H. Winslow, Muter. R e s . Stand., 5(3), 134 (1965).
[188] A. N. May, A p p l . M a t e r . R e s . , 5(2), 81 (1966).
[189] H. W. M c C o r m i c k , F. M. B r o w e r . a n d L. Kin, J . Polym. Sci., 39, 87
(1959).
[190] R. M c F e d r i e s , W. E. Brown, and F. J. M c G a r r y , SPE T r a n s . , 2 , 170
(1962).
[191]F. C. McGrew, Mod. Plast., 36, 125 (1958).
[192]J. R. McLoughlin and A. V. Tobolsky, J . Colloid Sci., 7 , 555 (1952).
196 MARTIN, JOHNSON, AND COOPER

[ l 9 3 ] F. H. McTigue, Plast. Technol., 6 ( 2 ) , 35 (1959).


11941 G. Meinel and A. Peterlin, Bull. A m e r . P h y s . Soc., 1 4 ( 3 ) , 363 (1969).
1195) E. H. Merz, L. E. Nielsen, and R. Buchdahl, J. Polym. Sci., 4, 605
(1949).
[196] E. H. Merz, L. E. Nielsen, and R. Buchdahl, Znd. Eng. Chem.,43, 1396
(1951).
[197] M. Miller and V. C. H a s k e l l , J . A p p l . Polym. Sci., 6, 627 (1961).
[198] D. R. Morrow, E. Foden, and J. A. Sauer, Bull. A m e r . Phys. Soc., 14 (3),
363 (1969).
[199] C. S. Myers, Mod. Plast. 2 1 ( 1 2 ) , 1 0 3 (1944).
12001 M. Nagano, Sen-i Gakkaishi 7, 544 (1951).
[ 2 0 l l E. I. Nalivaiko and A. G. Sirota, Plast. Massy, 1 9 6 8 ( 2 ) , 13.
Downloaded by [Princeton University] at 06:30 03 June 2013

[202) G. Natta, Ref. 22 of Ref. 238.


12031 G. Natta. I. Pasquon, A. Zambelli, and G. Gatti, Makromol. Chem., 70,
1 9 1 (1964).
[204] L. E. Nielsen, Mechanical Properties of Polymers ,Van Nostrand
Reinhold, New York, 1962.
(2051 L. E. Nielsen and R. Buchdahl, J. Colloid Sci., 5 , 282 (1950).
[206] K. Ninomiya. J. Colloid Sci., 14, 49 (1959).
(2071 K. Ninomiya, J. Colloid Sci., 17, 759 (1962).
12081 K. Ninomiya and J. D. F e r r y , J. Colloid S c i . , 18. 421 (1963).
[209] K. Ninomiya and J. D. F e r r y , J. Phy s. Chem., 67, 2292 (1963).
[210] K. Ninomiya, J. D. F e r r y , and Y. Uyanagi, J. Phys. Chem., 67. 2297
(1963).
12111 K. Ninomiya and H. Fujita, J . Colloid Sc i. , 12, 204 (1957).
[212] K. Ninomiya and M. S a k a m o t o , J . P h y s . Chem., 64, 181 (1960).
[213] S. M. Ohlberg, J. Roth, and R. A. V. Raff, J. Appl. Polym. Sci., 1 ( 1 ) ,
114 (1959).
(2141 Y . 6 a n a g i and J. D. Ferry, in Pr o c. Znt. Congr. Rheol, 4th, Part 2
(E. H. Lee, ed.), Wiley (Interscience), New York, 1965, pp. 491-503.
(2151 Y. G a n a g i and J. D. F e r r y , J. Colloid Sci., 21, 547 (1966).
[216] P. P a r r i n i and G. C o r r i e r i , Makromol. Chem., 62, 8 3 (1963).
[217] S. E. Paskvitovskaya, M. D. Medvedeva, 0. A. Makarova, A. S.
Seleznev. and V. I. Gusev, Plast. Mussy, 1 9 6 9 ( 8 ) , 46.
( 2 1 8 ) U. Pelagatti and G. Baretta, Mod. Plast. , 3 6 ( 1 0 ) , 140 (1959).
[219) G. Pezzin, G. Sanmartin, and F. Zilio-Grandi.J. Appl. Polym. Sci., 13,
1539 (1967).
[220] G. Pezzin, F. Zilio-Grandi, and P. Sanmartin, Eur. Polym. J., 6 , 1053
(1970).
(2211 G. Pezzin and G. Z i n e l l i , J . Appl. PoZym. Sci., 12, 1119 (1968).
12221 D. J. P l a z e k , J . P h y s . Chem., 69, 3480 (1965).
[223) D. J. Plazek, J . Polym. Sci., Part A - 2 , 6, 621 (1968).
(2241 D. J. Plazek. W. Dannhauser, and J. D. F e r r y , J . Colloid Sci., 16, 1 0 1
(1961).
MECHANICAL PROPERTIES OF POLYMERS 197

[2251 D. J. Plazek and V. M. O’Rourke, Rheology of Polymeric Materials-


Polymer Conference Series, M a y 22-26,1967, Wayne State University,
Detroit, Michigan, 49 pp.
I225al D. J. Plazek and V. M. O’Rourke, J. Polym. Sci., Part A-2, 9, 209 (1971).
[ 2 2 6 ] D. C. P r ev o r s e k , A m e r . Chem. S O C . ,Div. Polym. Chem. Preprints,
1 0 ( 2 ) , 1203 (1969).
[227] L. N. Raspopov, 0. N. Pirogov, N. M. Chirkov, and D. M. Lisitsyn,
Vysokomol. Soedin., 6, 1761 (1963).
12281 P. Raway,Assoc. Tech. 2nd. Papet. Bull., 1961, 117.
[ 2 2 9 ] V. G. Rayevskii, and S. S. Voyutskii, Vysokomol. Soedin., 6(1), 108
(1963).
( 2 3 0 ) B. E. Read, Polymer, 3, 529 (1962).
Downloaded by [Princeton University] at 06:30 03 June 2013

(2311 0. Redlich, A. L. Jacobson, and W. H. M c F a d d en ,J. Polym. Sci., Part


A , 1, 393 (1963).
[ 2 3 2 ) K. Richard, G. Diedrich, and E . Gaube,Rubber Plast. A g e , 39, 364
(1958).
(2331 R. B. R i c h a r d s , J . Appl. Chem., 1, 370 (1951).
[234] F. P. J. Rimrott and J. Schwaighofer, eds., Mechanics o f t h e Solid
State, Univ. Toronto Press, Toronto, 1968. p. 277.
[ 2 3 5 ] H. F. Robertson, India Rubber World, 1 2 7 ( 1 ) , 80 (1952).
(2361 M. J. Roedel, J . A m e r . Chem. SOC., 76, 6110 (1953).
[ 2 3 7 ] B. Rosen, ed., Fracture Processes in Polymeric Solids, Wiley (Inter-
science), New York, 1964.
(2381 S. E. Ross, J. Appl. Polym. Sci., 9, 2729 (1965).
(2391 J. F. Rudd, J . Polym. Sci., Part B , 1, l ( 1 9 6 3 ) .
12401 A. Rudin, in Crystalline Olefin Polymers P a r t I1 (R. A. V. Raff and
K. W. Doak, e d s . ) , Wiley (Interscience), New York, 1964, pp. 1-45.
(2411 F. M. Rugg, J. J. Smith, and L. H. Wartmen, J . Polym. Sci., 11, 1
(1953).
12.121 F. Rybnikar, Collect. Czech. Chem. Commun., 28, 320 (1963).
[ 2 4 3 ] Y. Sato, Kobunshi Kagaku, 2 2 ( 2 3 9 ) , 145 (1965).
[2.14] M. Sato and 0. Ishizuka, Kobunshi Kagaku, 2 3 ( 2 5 9 ) , 799 (1966).
[ 2 4 5 ] J. A. Sauer and A. E. Woodward, in Thermal Characterization
Techniques, Vol. 2 ( P . E. Slade and L. T. Jenkins, eds.), Dekker,
New York, 1970, pp. 107-224.
[ 2 4 6 ] A. P. Savel’ev, K. S. Minsker, L. M. Bobinova, V. L. Zvezdin, V. P.
Stesikov, and Yu. V. Ovchinnikov, Plast. Massy. 1 9 6 8 ( 9 ) , 6.
[ 2 4 7 ] J . C. S ch er er and S. N. Chinai, Mod. T e x t . Mag., 3 6 ( 1 ) , 74 (1955);
3 6 ( 2 ) , 4 7 (1955).
[ 2 4 8 ] P. C. S ch er er and R. D. McNeer, Rayon S y n . T e x t . , 3 1 ( 2 ) , 5 3 (1950);
3 1 ( 4 ) , 54 (1950).
(2491 P. C. Scherer and B. P. Rouse,Rayon Syn. T e x t . , 3 0 ( 1 1 ) , 42(1949);
3 0 ( 1 2 ) , 1 7 (1949).
[250] W. Schieber, Zellulolle Deut. Kunstseiden-Ztg., 6, 266 (1939).
(2511 K. Schmieder and K. Wolf, Kolloid-Z., 134, 149 (1953).
198 MARTIN, JOHNSON, AND COOPER

[ 2 5 2 ] J. van Schooten, H. van Hoorn, and J. Boerma, Polymer, 2 ( 2 ) , 1 6 1


(1961).
[253] L. F. Shalaeva and N. M. Domareva, P l a s t . M a s s y , 1 9 6 1 ( 9 ) , 10.
[ 254 ] V. P. Shatalov, V. P. Yudin, A. R. Samotsvetov, and E. I. Kolesnikova,
Kauch. Rezina, 2 7 , ( 2 ) , 13 (1968).
[ 2551 N. H. S h ear er , J. E. Guillet, and H. W. Coover, SPE J . , 17, 83 (1961).
[256] W. C. Sheehan, R. E. Wellman, and T. B. Cole, Text. R e s . J . , 35, 626
(1965).
[257] V. A. Sokolova and Z. A. Rogovin, Khim. Volokna, 1 9 5 9 ( 5 ) , 45.
[ 2 5 8 ] A. M. Sookne and M. H a r r i s , J . R e s . N a t . B u r . S t a n d . , 34, 467 (1945);
Ind. Eng. Chem., 37, 478 (1945).
12591 C. A. Sperati, W. A. Franta, and H. W. Starkweather, J . A m e r . Chem.
Downloaded by [Princeton University] at 06:30 03 June 2013

S o c . , 75, 6127 (1953).


[ 2 6 0 ] W. W. Spohn and H. J. F r e y , A 1 . E . E . Publ. N o . S95, 5 1 (1958).
[261] H. M. Spurlin, Ind. Eng. Chem., 30, 538 (1938).
[262] H. W. Starkweather. G. E. Moore, J. E. Hanson, T. M. Roder, and R. E.
Brooks, J . Polym. Sci., 21, 1 8 9 (1956).
[262a]H. W. Starkweather and R. E. Moynihan,J. Polym. Sci., 22, 363 (1956).
[ 2 6 3 ] H. Suyama, S. Satake, M. Ishii, and M. Kiyota, Sen-i Gakkaishi, 18. 321
(1962).
[264] S. Tachikawa, Rayon S y n . Text., 3 2 ( 3 ) , 3 1 (1951).
[ 2 6 5 ] A. Takaku and N. Kishi, Sen-i Gakkaishi, 2 2 ( 3 ) , 103 (1966).
[ 2 6 6 ] D. P. Thomas and R. S. Hagan, Polym. Eng. s c i . , 9, 164 (1969).
[ 2 6 7 ] P. E. Thomas, J. F. Lontz, C. A. Sperati, and J. L. McPherson. SPE
J . , 12, ( 6 ) . 89 (1956).
[ 2 6 8 ] E. V. Thompson, J . P o l y m . S c i . , Part A-2, 4 , 199 (1966).
[ 2691 A. V. Tobolsky, Properties and Structure of Polymers, Wiley, New York.
1960.
[ 2 7 0 ] A. V. Tobolsky, J. J. Aklonis, and G. Akovali, J . Chem. P h y s . , 42, 723
( 1965).
1271) A, V. Tobolsky and J. R. McLoughlin, J . P o l y m . Sci., 8, 543 (1952).
[ 2 7 2 ] A. V. Tobolsky, R. Schaffhauser, and R. Bohme, J . Polym. Sci., Part B ,
2, 1 0 3 (1964).
12731 L. H. Tung, S P E J . , 1 4 ( 7 ) , 25 (1958).
[274 1 L. H. Tung and S. Buckser, J . Phys. Chem., 62, 1530 (1958).
[ 275 I R. L. Tusch, Polym. Eng. S c i . , 6, 255 (1966).
[ 2761 K. Ueberreiter. Angew. Chem , , 53, 247 (1940).
12771 K. Ueberreiter, 2 . Phys. Chem., Abt. B . 45, 361 (1940).
[ 2 7 8 ] K. Ueb er r ei t e r and G . Kanig, J . Colloid Sci., 7, 569 (1952).
12791 K. Ueberreiter, G. Kanig, and A. Brenner, J . Polym. Sci., 16, 5 3
(1955).
[280] R. M. Vasenin, V. K. Gromov, V. L. Vakula, and S. S. Voyutskii, Adgez.
Polim., 1963, 52.
1281 I A. K. van d e r Vegt and W. R. Wilson, Techniques of Polymer Science
(SCI Monograph No. 17), SCI, London, 1963, pp. 108-121.
MECHANICAL PROPERTIES OF POLYMERS 199

[282] E. V. Veselovskaya, A. A. Levina, E. I. Nalivaiko, and M. D.


Pukshanskii, Plast. Massy, 1 9 6 6 ( 6 ) , 43.
[283] E. V. Veselovskaya, M. D. Pukshanskii, and L. F. Shalaeva, Plast.
M a s s y , 1969(9), 44.
12841 P. I . Vincent, Polymer. 1, 425 (1960).
12851 P. I . Vincent, Plastics (London), 27, 105 (1962).
[ 2 8 6 ] P. I. Vincent, Plastics (London), 28(307), 95 (1963).
[ 2 8 7 ] P. I. Vincent, Plastics (London), 29(315), 79 (1964).
[ 2 8 8 ] S. S. Voyutskii, Vysokomol. Soedin. Vsesoyuz. Khim. Obshchestvo i m .
D . I . Mendeleeva 1 (2). 230 (1959).
(2891 S. S. Voyutskii, V. L. Vakula, V. E. Gul’, and Y. T. Ho, in Issled. Obl.
Poverkh.Si1. S b . Dokl. Konf., 1st (B. V. Deryagin, ed.). Academy of
Downloaded by [Princeton University] at 06:30 03 June 2013

Sciences of USSR P r e s s , Moscow, 1960 (Published 1961), p p 35-45.


Translation: Research in Surface Forces, Consultants Bureau, New York,
1963, pp. 42-49.
12901 H. Wakeham, in Cellulose and Cellulose Derivatives ( E . Ott, H. M.
Spurlin, and M. W. Grafflin, eds.) , Interscience, New York and London,
1955, pp. 1329-1334.
[ 2 9 1 ] M. L. Wallach, A m e r . Chem. Soc., Div. Polym. Chem. Preprints, 8 ( 1 ) ,
656 (1967); J . Polym. Sci., Part A - 2 , 6, 953 (1968).
[ 2 9 2 ] A. Weissler, J . A m e r . Chem. Soc., 71, 93 (1949).
(2931 P. W. 0. Wijga, Physical Properties o f P o l y m e r s (SCI Monogr. No. 5).
SCI, London, 1963, pp. 35-45.
(2941 L. Wild, J. F. Woldering, and R. T. Guliana, SPE T e c h . Pap., 13, 91
(May 1967).
(2951 R. H. Wiley and G. M. Brauer, J . Polym. Sci., 3, 647 (1948).
12961 R. H. Wiley and G. M. B r a u e r , J . Polym. Sci., 11, 221, (1953).
[ 2 9 7 ] R. H. Wiley, G. M. B r a u e r , and A. R. Bennett, J . Polym. Sci., 5. 609
(1950).
[ 2 9 8 ] G. R. Williamson, B. Wright, and R. N. Haward, J . Appl. Chem., 14,
131 (1964).
[299 J D. P. Wyman, L. J. Elyash, and W. J. F r a z e r , J . Polym. Sci., 3, 681,
(1965).
(3001 M. Sh. Yagfarov and V. S. Ionkin, Vysokomol. Soedin., Part A , 10, 1613
(1968).
13011 J. A. Yanko, J . Polym. sci., 3 , 576 (1948).
[ 3 0 2 ] H. I-I. Zabusky and R. F. Heitmiller, SPE Trans., 4, 17 (1964).
(3031 V. I. Zegel’man, N. A. Okladnov, and Ye. N. Zil’berman, Vysokomol.
Soedin., Part A , 10, 1319 (1968).
(3041 S. N . Zhurkov and S. A. Abasov, F i z . Tverd. Tela. 4, 2184 (1962).

You might also like