You are on page 1of 25

Influence of Molecular Weight and Molecular Weight

Distribution on Mechanical Properties of Polymers


RONALD W. NUNES
Institute of Materials Science
University of Connecticut

JOHN R. MARTIN
The Foxboro Company
Foxboro, MA 02035
and
JULIAN F. JOHNSON
Institute of Materials Science and
Department of Chemistry
University of Connecticut
Storrs, Connecticut 06268

This work reviews the literature published over the last ten
years on polymer mechanical properties as a function of
molecular weight and molecular weight distribution. Thermal
properties, stress-strain properties, impact, fracture, fatigue,
creep, stress relaxation and cracking and crazing are examined
for a wide variety of homopolymers and a limited number of
copolymers. In general, mechanical properties increase as the
molecular weight increases. However, above some limiting
molecular weight the mechanical property is usually unaf-
fected. Although much work has been done to describe the
effects of molecular weight on mechanical properties, little
quantitative correlation exists. The available equations to pre-
dict such properties as cracking and crazing, To, T,,4and tensile
strength from molecular characteristics are discussed in detail.
However, a more quantitative description incorporating a
wider range of mechanical properties would be more useful.
This would facilitate use of the vast amount of information
available and enable it to be applied more readily to new
polymer systems.

INTRODUCTION fair representation. From this paper the reader will be


t is well known that molecular weight (MW) and able to obtain the trends expected with regard to MW
Iaffect
molecular weight distribution (MWD) can markedly
the mechanical properties of a polymer. For many
and mechanical properties. The reader will also be-
come aware of the research being performed in this
mechanical properties and polymers this effect reaches area.
a limiting value at some relatively high MW where STRESS-STRAIN
there is no longer any appreciable change on mechani-
cal properties with increasing MW. This paper reviews General Considerations
the effect on mechanical properties for a variety of Stress-strain tests give an indication of strength and
homopolymers and copolymers. While the material in- toughness of a polymer. Since toughness is the energy a
cluded in this review is concentrated between the years material can absorb before breaking, there should be a
1972 to 1981, some previous references have been in- correlation between impact strength and toughness.
cluded when it seemed necessary to clarify or em- Brittle materials show low toughness, whereas ductile
phasize certain areas. This review paper in essence is an materials which can be cold drawn are very tough due to
update on the review paper written by Martin, et al. (1); a large elongation at the break (EB).
hence for work performed prior to 1972, one should Stress-strain curves are found to be very dependent
consult Martin's review paper. upon the type of test used; tensile and compression may
While this paper does not represent a complete re- yield different results, Fig. 1. Even the first part of the
view of all the work done between 1972 to 1981, it is a curves, which is used to determine Young's modulus,

POLYMER ENGlNEERlNG AND SCIENCE, MARCH, 1982, Vol. 22, No. 4 205
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R. Martin, and Julian F . Johnson

It has been found (3-6) that above the minimum MW


required to form a specimen, the strength and elonga-
tion increase toward a limitingvalue at high MW. Polar
polymers and those capable of hydrogen bonding reach
their maximum at a lower MW than d o non-polar poly-
mers (2).
Stress-strain properties of crystalline polymers such
as PE and PP depend on MW similar to amorphous
polymers (2, 3); however, this dependence seems to be
less apparent. This lies in the fact that crystallites help
hold the material together, much the same as chain
entanglements. Molecular weight dependence can be
further confused because the degree of crystallinity can
decrease as MW increases. Therefore, at temperatures
above T,, an increase in the MW tends to increase
toughness. The lower MW polymer tends to collect at
the boundary between spherulites, reducing the
number of “tie molecules.” Therefore, low MW poly-
mers tend to be brittle and have,low strength (7). The
type of crystal morphology and extent of spherulitic
Fig. 1 . Stress vs. strain for polystyrene comparing compression structure also may change with MW; hence, the be-
und tension (2). havior of crystalline polymers is affected by both MW
and a variable degree of crystallinity which may change
with MW.
may be quite different, resulting in higher moduli de-
termined by compression rather than tension. The following sections, titled Tensile, Yield Elonga-
In tension, some polymers can fail in a brittle man- tion, Brittle Point and Moduli describe the recent work
ner, while in compression the. same material might be- that has been done concerning the effect of M W and
have as a ductile polymer with a yield point and higher MWD on these properties.
EB. Tensile properties of brittle materials are greatly Tensile
influenced by flaws and submicroscopic cracks which
are a part of that polymer. However, cracks do not play Mark (8) discusses an overall formula which would
an important role in compression testing because the predict the tensile strength of a polymer. As shown in
stresses have a tendency to close rather than open Fig. 2 he found a simple qualitative picture of the
cracks. Therefore, compression testing tends to be relationship between strength and the degree of
characteristic of pure polymer, while tension tests are polymerization which has also been expressed quantita-
more characteristic of flawed material or material gen- tively in the form of Eq 1 (8, 9):
erally used in applications. T, = Tsa- (C/ DP,,) (1)
The temperature the test is performed is also impor-
tant in determining stress-strain curves. At tempera- where T,is the tensile strength at the degree of polym-
tures well below T , results would show an EB, and no erization for the polymer, DP,, T,, is equal to the
yield. As the temperature is increased there develops a extrapolated tensile strength for an infinite DP, C is
yield point and an increase in EB. Finally, as the mate-
rial softens EB may show a decrease. The rate of testing
also affects behavior; modulus and ultimate strength
increase whereas En generally decreases as the rate of
testing is increased.
Perhaps the largest effect on mechanical properties is
due to the MW of the polymer. At very low MW, a
polymer is a viscous liquid if the T , is below ambient
temperature. At higher M W, a polymer may become an
elastomer, having low strength and a high EB. Above an
MW of about lo5, polymers become entangled enough
to show true rubbery behavior and exhibit a greatly
increased EB.If the polymer is of very low MW and the
T, is higher than ambient temperature, the material is
brittle. Hence, molecular entanglement must take
place before the polymer can become strong enough to 100 200 300 400 500
carry a load (2). It is also true that chain ends act as OP,
imperfection in the polymer network which can ad-
versely affect the strength properties; however, they Fig. 2 . Simple, qualitutiue picture of the relationship between
have little effect on the elastic moduli. strength und DP (8).

206 POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Influence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

equal to a constant which is characteristic for a given relative humidity at temperatures ranging from 65 to
polymer and DP, is the average degree of polymeriza- 93°C. The humidity caused the samples to hydrolyze,
tion. Once Tsmand C have been determined by two hence there was a reduction in MW. When MW fell
measurements of T, at two known values of DP,, a below a critical value (HW = 33,800, M,, = 14,300), the
quantitative expression for the tensile strength of the tensile strength dropped off rapidly and the failure
polymer over the entire DP range can be arrived at. mode changed from ductile to brittle. Their data also
For poly(viny1 chloride) (PVC) Shinozaki, et al. (lo), showed an increase in tensile strength with decreasing
found that the ultimate tensile strength (UTS)increased MW above a minimum MW. They explained this
as M w increased. They state that the higher Mu.material phenomenon as an annealing effect which reduced
had a higher elongation to fracture which implied a localized stresses and increased short range order.
larger strain hardening capacity. Further, the greater Work has also been reported on a Soviet polycarbo-
the elongation that a given specimen can withstand, the nate modified with si1icon:polysiloxanecarbonate (4). It
greater the degree of orientation in that specimen be- was found that the tensile strength and EB of theqoly-
fore failure. Hence, higher M w specimens elongate mer with aviscosity average molecular weight of M , = 2
further, are more highly oriented, and therefore exhibit x lo4 to 3 x lo4 to be quite low. Raising the MW
a higher UTS. Lee and Turner (11) examined the effect increased tensile strength but only up to M u= 6 x lo4.
of crosslinking poly(methy1 methacrylate) (PMMA) and Margolies (3) and Perkins, et al. (15), found that the
found that the crosslinks had little influence on the tensile strength of high density polyethylene (HDPE)
tensile strength of networks when the primary MW was increased with MW, but only up to 1.5 x lo6, Fig. 4.
> lo5. Crosslinking did raise the strength of PMMA Increases in MW lead to an increase in the number of
with MW < lo5. They found that crosslinking caused intercrystalline tie chains that run longitudinally in
tensile strength to approach the value observed with each microfibril and laterally between microfibrils.
high MW linear polymer. They rationalized the above When stretched as in a tensile test, the tie chains re-
findings by referring to a theory which attributes the strain the microfibrils from slipping with respect to each
strength of a linear glassy polymer to entanglements other, thus increasing tensile strength (15).
that act as physical crosslinks (12). For PMMA (ll),the Wu and Black (16)found the tensile strength of PE as
density of entanglements becomes significant only a continuous filament with a draw ratio (A) of 30 at 250°C
above a MW of 10,000. As MW increases, the density of to increase with a,.
They concluded that one necessary
entanglements increases but eventually levels off above condition for a high strength PE fiber is a a,
> 2.2 x
1 x lo5. Thus, chemical crosslinking caused a marked 104.
increase in the tensile strength of low MW samples Larsen (17) found that high MW PE could tolerate
(MW = lo4to lo5),but had little effect above 1 x lo5. high filler loadings without becoming brittle. Blends
Deanin, et al. (13), in dealing with plasticized PVC, containing polymer, filler, and plasticizer required
found the UTS shown in Fig. 3 to be dependent on MW only a small amount of a very high molecular weight PE
and to correlate fairly well with all three MW averages. to maintain appreciable strength.
Some theories (11-13) ascribe the ultimate failure to Schollenberger and Dinbergs (5) examined thermo-
disentanglement of long chains, while others point to plastic polyester urethane elastomers containing 20
the more numerous ends of short chains as points of percent hard segments over a MW range of 4.8 x lo4to
weakness. Deanin’s (13) data accept both theories as 3.67 X lo5. They showed, Fig. 5 , that the tensile
factors contributing to the overall mechanism of failure. strength of this polymer series increased with increas-
Sauer (9) and Vlachopoulos, et al. (6), found that the ing Mu. up to a limiting value of M?,. = 120,000. By
tensile strength of mono-dispersed PS can be described contrast, elongation decreased slightly with increasing
using the form of E q 1 . Vlachopoulos, et al. (6), re- MW. Thus, the increased number of virtual crosslinks
ported the limiting value for M, to be between 3 X lo5 per polymer chain at higher MW apparently builds
and 4 x lo5. polymer strength while reducing polymer extensibility.
Gardner and Martin (14) aged polycarbonate (PC)
60001 , I , I I I l l 1 I I /
tensile bars for up to 18 months at 0,75, and 100 percent
.-m Effect Of MWD On Ultimote Tensile Strength
a
I

M, lo4 Molrculor wdipht


Fig. 3. Ultimate tensile strength of PVC plotted versus M,,,
M,,, Fig. 4 . Tensile yield and tensile failure of high-density polyeth-
(13). ylene vs. molecular weight (3).

POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No, 4 207
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R. Martin, and Julian F . Johnson

POLYMER K STYRENE
0 SBS-6

- 240-
N.
0

0
SBS-7
SBS- I
5 SBS-3
-6 200- AA SBS-5
ses- a
40
40
2
-
W 160-
=! c 75 mil samples w
LL
v SBS-6 ANNEALED
v) .
-
I 25 mll samples e
z POLYMER
W
I- AE 7 D EF
II
G
I
U
1 I I
I
J
I I
120-

Fig. 5 . Tensile s t r e n g t h uersus @,r f o r thermoplastic


polyurethane (5).
EXTENSION RATIO 1
Foot and Ward (18)concluded that their tensile tests Fig. 7 .Effect of composition and block size on tensile properties
caused two types of brittle failure at -160°C in (20).
poly(ethy1ene terephthalate)(PET). For the medium
and high B?,. amorphous samples, failure resulted from content does cause a significant decrease in tensile
surface initiated crazes. However, in low MW amor- strength. It was therefore concluded that the tensile
phous samples (an
= 1.2 x lo4)and all crystalline PET, strength is dependent on PS content and not the elastic
the failure is initiated by inherent flaws. This latter block length, but that it is affected by the MWD of the
process may still be governed by craze growth but, from latter.
an internal flaw, rather than from a surface flaw.
Elongation
Work has also been done on liquid hydroxy-
terminated polybutadiene (19). Tensile strength was Margolies (3), working with HDPE, found that the
found to decrease with increasing MW. This phenome- ultimate tensile elongation increased rapidly to a
non was more pronounced in the low MW samples maximum at an MW of 5 X lo5to 7.5 X lo5. However,
below 3 x lo3 as is shown on Fig. 6. further increases in MW caused a steady decrease in
Morton, et al. (2O), dealt with styrene-diene thermo- ultimate elongation, Fig. 8. Margolies gave two possible
plastic elastomers, and found the useful range of the explanations for this behavior. First, he considered the
block MW to be 1 x lo4to 2 x 104forPS and 4 x lo4to8 molecular chain length in the vicinity of MW = 5 x lo5.
x lo4 for polybutadiene. The lower limit on PS was Here the molecular chains are long enough to be
governed by the minimum chain length required to stretched and straightened and yet short enough so that
insure the formation of a heterogenous phase, while the substantial entanglement, as compared to that which
upper limit was set by the high viscosity of both blocks occurs in the ultra-high molecular weight region, does
since this factor also influences phase formation. Their not occur. In the ultra-high molecular weight region,
tensile curves, Fig. 7, indicate that both the stress value chain entanglement increases to a point where chain
and the ultimate strength are dependent only on the PS straightening is substantially inhibited prior to break,
content and not the block lengths. Increasing the MWD therefore resulting in lower elongation values.
of the center block without altering the average PS The second explanation he proposed was that in addi-
tion to increasing chain entanglements, intermolecular
forces between very high MW molecules become sig-
1 nificant (21). In low MW PE the intermolecular forces
between small molecules are weak in comparison to the
80- primary valence forces. Therefore, chain slippage is the
I I I I I I 1 1 , I I I I 1 1 1 1

60 - .-C
Ilo4
-
- a -
40 - -
-
.-c
- e
V

- .-e
20 - -a
-
I

I0

c'
I I 1 1 I I I I I I 1 1 1 1 103
0 2000 4000 6000 8000 10000
- Io5 I06 10'
Mn Moleculor weight

Fig. 6 . Tensile s t r e n g t h 6,s. a,,of hydroxyl-tertnincited F i g . 8. Ultimate elongution und tensile impact strength of high
polyDutadierre (19). density polyethylene us. molecular weight (3).

208 POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Influence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

primary factor affecting elongation. In high MW sam-

1
ples, however, it becomes easier to break primary val- A
ence bonds (chain scission) than to overcome the chain
entanglements and significant intermolecular forces. B
The net effect again is a reduction in ultimate elonga- 20 --- C
tion.
Perkins, et al. (15),examined ultra-drawn H D P E and
-
(u

found that the ultimate elongation increased with in- E


creasing MW. They stated that this was d u e to the z 16
greater number of non-extended intercrystalline tie
r
chains linking the microfibrils and resisting the separa- 0
1
tion which leads to fracture. The fibers made from Y 12
lower MW H D P E would contain fewer and shorter v,
intercrystalline tie chains resulting in less extension v,
W
before the breaking stress is reached. a
For PC with a M,, of 4.5 x lo3to 7.15 x lo3, Pitman +
v,
8
and Ward (22), found little increase in the extension to
break with increasing iGTr for tensile tests performed at
-30°C.
Hydrolysis during long term humid aging of PC 4
causes a progressive reduction in molecular weight (14).
As a result of this MW reduction, elongation drops
substantially as shown in Fig. 9. The hydrolytic effect
on elongation became apparent during the early stages
n
0
- 5 to I5 20 25
of the humidity aging period. This suggests that
S T R A I N ( '10 1
whenever the degradation mechanism is a molecular
weight reduction, toughness related properties will be F i g . 10. Nominal stress-nominal strain curces vf three polypro-
pylenes ('u,,.; A = 300,000, B = 400,000 and C = 900,000) i n
impaired before strength related properties are lost. tension ut carious pressures (23).
In pol ysiloxanecarbonate, Khukhreva, et aE. (4),
found that an increase in MW raised the value of EB.

//
Effecl Of MWO On Ultimate Elongalion
However, increases beyond @, = 6 x lo4did not bring
8
any further improvement. =- 200
The elongation to fracture of P P depends on MW cF
according to Pae, et al. (23). Figure 10 shows that at a
160
various pressures, higher elongation was obtained for II
higher MW polymer than for lower MW. 2
W

-
In plasticized PVC, Deanin, e t al. (13), show a de- $ 120
pendence of ultimate elongation on MW, Fig. 11. z
When plotted against a,,
the broad M W D blends gave 5
= 80
somewhat higher elongation than the normal commer- I 2 3 4 4 6 80 4 8 I2
cial MWD's. When they plotted elongation versus @,,. M, lo4 lo4 M, M, lo4
Fig. 11. Ultimate elongation of PVC plotted against M,,, a,,,
a,
(13).

the envelope became more narrow. The closest correla-


tion was obtained when the data were plotted against
Mr, suggesting that the high MW fraction was the con-
trolling factor and that failure depended on disentan-
glement of the longest chains.
Schollenberger and Dinbergs (5) found that with
polyurethanes there is a definite influence of low MW
components on the ultimate elongation of 25 and 75 mil
(0.6mm and 1.9 mm) samples. The results, Fig. 12, also
show a definite decrease in sample extensibility with
increasing Mw.An inflection point was observed at alL.
= 1.05 x lo5for a sample thickness of 75 mil and at a,.
= 65,000 for a 25 mil sample. They state that the
extensibility pattern is a consequence of the more ex-
tensive virtual network that develops among the
APPARENT WEIGHT AVERAGE MOLECULAR WEIGHT 116') polyurethane chains as chain length increases. They
F i g . 9. Elongation cs. MW f o r polycarhonate (14). concluded that this, rather than increased extensibility,

POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4 209
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R. Martin, and Julian F . Johnson

8 1000
z
o_ 800
-.--- 75 mil somples
25 mil samples
devitrification process would therefore be expected to
take place at similar stress levels. However the devit-
rified material will change considerably in character
from a viscous liquid at low MW below a characteristic
value M , , to a rubber-like solid at high MW. This
change from a liquid-like to a rubber-like material on
devitrification was associated with the onset of crazing
5P 200-
POLYMER
prior to fracture and a marked increase in tensile
m AB D EF I G H I J strength. They predict that these changes begin at a
0 , I

characteristic molecular weight M , , at which an entan-


glement network first becomes apparent in the corre-
sponding polymeric melt.
Foot and Ward (18) found the brittle strength of
amorphous PET at -160°C to be slightly MW depen-
dent at high and medium MW. Low MW material had a
much lower brittle strength. Their results showed two
processes which were believed to initiate the brittle
failures. Medium and high MW amorphous materials
fail from crazes that initiate at the specimen surface.
Failures in low MW amorphous material, a,,
= 12,000,
and in all crystalline samples were initiated at internal
flaws.
Grezlak and Wilkes (29) showed that block PMMA-
PE-PMMA copolymers with a MW below 2.1 x lo4
were extremely brittle in nature. This was stated to be
due to the absence of chain entanglements below this
MW.
Gardner and Martin (14)determined that a transition
from ductile to brittle failure occurred in PC at M, =
33,800, a,= 14,300. Observations from this study
indicate that PC ductile failures occur through shear
yielding, whereas brittle failures result from craze for-
mation which leads to crack nucleation and propaga-
tion. The transition from ductile to brittle failure occurs
when the craze initiation stress falls below the shear
yield stress (30).

Modulus
Strength and stiffness are closely related theoreti-
cally, but in practice, a high modulus does not necessar-
ily guarantee high strength (16). The following studies
pertain to the influence of MW on polymer modulus.
Work has been done on four LDPE homopolymers
shown in Table 1 . These were selected so effects of a,v
and an could be evaluated separately. The data show
that up to A of 29, room temperature modulus is inde-
pendent of MW. Although no large increase in A with
respect to MW was observed, Cappaccio and Ward (31)
suggested the existence of a un4ue correlation between
A and Mw, since the lower the M t r , the higher the A. This
effect on A is important because of the effect on mechan-
ical properties. Contrary to Cappaccio and Ward (31),

Table 1. Molecular Weights of Rigidex PE Samples (31)


- -
Polymer' Sample M" Mw

Rigidex 9 41) 6,060 126,600


Rigidex 50 42) 6,180 101,450
Rigidex 25 B(1) 12,950 98,800
Rigidex 140-60 B(2) 13,350 67,800
* Supplies by BP Chemicals Int. (U.K.) Ltd.

210 POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Injluence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

Wu and Black (16) measured mechanical properties on correlations between mechanical properties and M W
PE yarns using a gauge length of 25 cm rather than the and MWD were found.
conventional
- 2.5 cm and found that, for PE filaments, Kusy and Greenberg (37)studied the affect of M W of
M , rather than t h e a, was t h e more important PMMA on dynamic mechanical properties using a
molecular parameter. dynamic mechanical analyzer, DMA. The results over a
The temperature range for effective drawing de- wide range of M, showed that the modulus was rather
pends on MW and MWD. The effective drawing tem- invariant in the glassy region, while the increasing
perature range for PE shifts towards higher tempera- number of chain ends did shift the transition region to
tures for broader MWD and higher MW (32). Ultra- lower temperatures (Fig. 13). This latter decrease cor-
high modulus samples have_been _ obtained b y drawing responds to the intense spike which they observed on
HDPE with broad MWD, M,IM, = 20 and M,. z 2 x the tan 6 vs. temperature plot and represents the glass
lo5, at higher drawing temperatures. Transparent sam- transition temperature.
ples with A of 35-40 and Young's moduli of 600-650 Kbar Schollenberger, et al. (5), looked at thermoplastic
at room temperature have been obtained by drawing PE polyurethane for affects of MW on strain and stress.
at 100-105°C.Jarecki and Meier (32)concluded that the Figure 14 illustrates a full stress-strain curve showing
MW fraction in PE, forming tie molecules in the drawn the consistent pattern of a gradual steepening of the
material is responsible for the high modulus, while the curve as MW increases, Note that MW affects were
low MW fraction facilitates alignment of the chains and negligible at the low strain levels which are characteris-
retards the internal voiding (whitening) to a very high A tic of most end use applications.
during drawing at the higher temperature. Shinozaki, et al. (lo), found that the yield stress of
Barham and Keller (33) showed that maximum PVC increased and the modulus decreased when MW
modulus depended on morphology and MWD. They was increased. Tensile modulus was found to decrease
went on to state that for PE, the prerequisite for ultra- with increasing MW by about 15 percent while yield
high modulus is a high A (30X)and the presence of some stress changed only 4 percent. The modulus trend (Fig.
very low MW material. Also important was high draw- 15) was rationalized as a secondary effect, which may
ing temperature and slow crystallization. arise from a number of causes, one of which is the
Capaccio and Ward (31) found that an acceptable degree of syndiotacticity. If tacticity varies with as a
modulus could be obtained with PE of fairly broad result of polymerization conditions, then the degree of
MWD, and that much higher A were obtained after crystallinity in PVC also changes. The modulus is then
optimization of the spinning and drawing processes. the composite modulus of a twoqhase material, with
They also felt that the limit to possible improvements of only an indirect dependence on M , . Unfortunately, no
mechanical properties had not been reached, since they experimental evidence is available at this time to judge
achieved a room temperature Young's modulus of 400 the significance of this rationalization. The addition of
Kbar at a A of 20. This suggested to them that the slight increasing amounts of stabilizer to PVC resin resulted in
decrease in slope observed at a A of 14 in the modulus vs
A curve and the definite plateau in Young's modulus at
approximately 300 Kbar (34) could no longer be con-
A
sidered representative of the behavior to be expected if
both the structure of the undrawn PE and the drawing
process were fully optimized.
T h e optimum draw temperature of ultra high
modulus PP is MW independent according to Wills,
Capaccio and Ward (35). High modulus products can be
obtained over a range of M W --- 1.8 - 4.0 x lo5with the
highest modulus attained at the low end of this range.
Pae, et al. (23), showed that pressure is also an impor-
tant variable. They found that at a given pressure, the
effect of increasing MW was to decrease the modulus
and strength of PP; however, higher pressure caused a -t
significant increase in the modulus. The pressure de- A A
pendence of modulus was linear at lower pressure but
MW related deviations were observed at higher
pressure. Pae attributed this effect to a pressure in-
duced shift of the &transition at room temperature.
Results have indicated that, for high MW materials,
greater pressure is required in order to induce a shift of
this magnitude. - 50 0 50 100
Fellers and Chapman (36)examined PS for the effect Temperature ,OC
of MW as a function of strain. They found that in the Fig. 13. lnfluence of moleculur weight on dynamicmechanical
elastic region the modulus was independent of MW and properties of P M M A . 0 , 0 ,A und 0correspond to M , , of 587,000,
MWD. However, as the strain level was increased, 54,000, 6400 and 1600, respectiuely 137).

POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4 21 1


15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R. Martin, and Julian F . Johnson

I- H
quired for domain formation (38, 40). However,
changes in t h e relative proportions of PS and
polybutadiene or polyisoprene segments significantly
8000
influence both mechanical and flow properties of SBS
andSIS (41).
For SBS block copolymer films cast from methylethyl
ketone (MEK) at a constant overall M,, Kalfoglou (42)
.-u) 6000 found that the storage modulus ( E ' ) increases with the
styrene content. He also found that at constant composi-
-0.

tion, E' increases with PS block length.


v)
v)
Boussias, et al. (43), showed that for tetramethylene
4000 terephthalate-polytetrahydrofuran random block co-
I- polymers containing 10, 20 and 30 percent polyether of
v) A
M W 1 x lo3and 2 x lo3, there was a decrease in initial
modulus as the concentration of polyether was in-
creased. It has been shown by Martin and Graessley (44)
2000 and Onogi, et al. (45), that the elastic modulus (G') and
viscous modulus (G") master curves for PS could be
separated into three regions: terminal region occurring
at low frequencies, the plateau region where G' and G"
W are independent of reduced frequency (a+) and the
200 400 600 high frequency region where G' and G" are roughly the
S T R A I N (O/o) same and independent of MW. Ahuja (46) examined G'
Fig. 14. Stress-strain properties us. a,.
f o r thermoplastic and G" of a 65/35 random copolymer consisting of
polyurethane. Molecular weight increases from A(47,900) to sty-ene and n-butyl methacrylate. The G' vs. a+ for
](366,800) (5). M,. = 3.3 x lo4had a much longer terminal region than
polymers with higher a,L..
For the polymer with =a,.
3.99 X lo5 the terminal region extended two decades
whereas for 3.3 x lo4 is was roughly four decades.
Similarly Ahuja (46)found that G" for M, = 3.3 x lo4
3.0 phr was four decades long whereas the terminal region for
I
= 3.99 x lo5was only about a decade (Fig. 16). He
found that G' and G" started at increasingly lower
frequencies as MW was increased. Hence, the increase
in MW extends the rubbery plateau and reduces the
terminal region.

IMPACT
General Considerations
Impact tests are basically a high speed fracture test
which measures the energy required to break a speci-
1 I 1 J men. There are several methods of making this test.
07 08 09 1.0 II
Naninal lntrmsic Viscority
F i g . 15. Change in tensile modulus with nominal intrinsic uis-
cosity at different stabilizer contents (10).

an increase in modulus and a decrease in yield stress.


This is consistent with a model for plastic deformation
in which the stabilizer acts as a small, hard particle.
Ono (19)found for HTPB that the modulus decreased
with increasing MW. This was especially significant
with low MW HTPB < 3 x lo3.
Some work has also been reported on copolymers.
Early studies concerning styrene-butadiene-styrene
(SBS) and styrene-isoprene-styrene (SIS) triblock
polymers focused on the influence of compositional
changes in these polymers. It was concluded that M W
changes at constant composition should produce little F i g . 16. Loss modulus G"(a,w)at T , = 120°C for65130 styrene-
or no changes on stress-strain behavior (20, 38, 39) as n-butyl methacrylate random copolymer of tiarious molecular
long as the overall MW was above a critical value re- weights (46).

212 POLYMER ENGINEHUNG AND SCIENCE, MARCH, 1982, Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Influence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

Izod and Charpy impact tests consist of a pendulum


with a hammer-like weight that strikes a notched or
unnotched sample, and the energy required to break
the specimen is determined from the loss in kinetic
energy of the weight. The falling ball or dart test mea-
sures the amount of energy required to break a sheet or
plate of the material determined from the weight and
height of the substance dropped (2). There are still
other methods such as the Tensile Impact Tester (47),
and tensile stress-strain tests ASTM D1822 and D2289,
which are used to obtain impact results that are not as
widely used as the previously mentioned tests. I o5 108 I'
0
The agreement among the various impact tests is Molecutor weight
generally poor, which indicates that there are at least Fig. 17. Ultimate elongation and tensile impact strength of
two physical properties affecting impact strength. Two high-density polyethylene vs. molecular weight (3).
of these are the energy needed to initiate a crack and
energy required to propagate the crack (2). There are
also many variables which affect impact results. These
include geometry and sample size; for example, thin
samples tend to yield higher impact strengths than thick
ones (2).A notched specimen tends to yield an impact
strength which is less than the unnotched specimen (2,
48, 49). The temperature of testing also has an affect
which is to increase the impact strength as the tempera-
ture increases (48-50).
The molecular make-up also plays an important role
in the impact strength of a polymer. It has been found
that molecular orientation usually increases impact
strength if the impacting force is parallel to the orienta-
I I I I I I I l l I
tion and decreases it if perpendicular (48). The follow- I I I J I 1 1 1 1
6 I
10 10 4 x 10'
ing section deals with the influence of MW and MWD
Molecular Weight
on the impact strength. Overall, the impact strength
increases with increasing MW (3, 51-54) usually up to a Fig. 18. Izod impact of HDPE vs. molecular weight (3)
maximum value where a further increase produces no
further change (3). increases rapidly with increasing MW until a value in
excess of 20 ft-Ib/in of notch is reached. It also shows
Impact that above MW ~1 3 x lo6, impact strength begins to
Margolies (3) investigated compression molded decrease slightly with additional increases in MW.
HDPE with a MI,.of 1 x lo5to 4.5 x lo6and found that Since all failures in this region are hinged breaks, the
most physical properties reached aplateau at 1.75 x lo6 factors that affect elongation are also affecting Izod
and did not change with further increase in MW. How- impact strength.
ever, when he measured tensile impact according to Cohen (51) studied the effect of MWD on impact
ASTM D1822-68 using Type L tensile bars, he found strength using the notched Izod impact test (Fig. 19).
that as the MW of HDPE increased, the impact strength This data demonstrates that M, must be increased from
improved, reaching a maximum at a MW of 1.5 x lo6 1.50 x lo5 to 2.75 x lo5 to achieve the same level of
and then decreased gradually with increasing MW. He toughness when an end-user shifts from a broad MWD
also discussed the energy required to break a Type L to a narrow MWD grade during processing.
tensile bar. Though primarily a function of MW, it was Using a Dart Drop Impact tester, Perron (52) also
also affected by the ability of the polymer to deform or found that a broader MWD results in an increase in
elongate under sudden load. Figure 17 illustrates this impact strength. He went on to explain that this phe-
point by plotting ultimate tensile elongation and tensile nomena was a result of an increase in long chain branch-
impact as a function of MW. By noting the similarity of ing of higher MW fractions which would cause a higher
the curves in the ultra-high MW region, MW > 1.75 x degree of MW entanglements at the branch sites.
lo6,it is readily seen that factors causing the decrease in Ryan (53)compared Izod impact strengths for 2 com-
elongation with increasing MW (intermolecular forces, mercally available PC. Comparing the two graphs in
chain entanglements) are also affecting impact be- Figs. 20 and 21, the ductile-brittle transition for
havior. Izod impact was less sensitive to these elonga- supplierA (Fig. 20) occurs at lower melt flow rate (MFR)
tion effects, especially when brittle failure occurred. than supplier B (Fig. 21). For a given material, the
However, for hinged breaks, the total energy required smaller the notch radius, the lower the MFR required
again becomes a function of the ability of the polymer to for ductile behavior. When PC is annealed, the relaxa-
elongate. Figure 18 shows that Izod impact strength tion time shifts to a greater value.

POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4 213
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R . Martin, and Julian F . Johnson

I
8
1 I 1

3 a
E 4-

w-5
\

MILS AU
\ '. ---x-----
(- ---o
~

2 I 1 I 1 I I I I

MELT FLOW RATE


gm/lomin
Fig. 21. Izod impact of supplier B polycarbonate asa function of
Broad melt flow rate (53).
I
0.8
0.6
t 201
zm .-
10 MILS uw
*-
0.4 -20 - '6
MILS NOTCH RAOWS
-40- /' UN UNANNEALEO
AN ANNEALED
d'
15 MILS UN
-60 - ..
0 J I I I I J It
* *O .- 4.C
70 90 I50 250 400 -4 b d lo :2 4
: a ;s
MELT FLOW RATE
gm /!O min
M~ lo3
Fig. 19. Notched lzod impact of HDPE 11s. M,.(51). F i g . 22. lzod impact ductile brittle transition temperature "C of
supplier A polycarbonate as a function of meltpow rate (53).

H- I
t
u)
f 20
-
L
I- MILS NOTCH RAOIUS

?? 0 - UN UNANWALED
Lk AN ANNEALE017
HRS AT ( Z V C
MILS NOTCH RADIUS 22 *
UN UNANNEALED "2 -20- +0,-4-c

AN ANNEALED 17
22
H113 AT 1 2 5 % tiz
2: -40-
c
0

2 -w. I S MILS UN
z

MELT f L o I RATE
gmf 10min
-::
0

-60 I I I 1 1

Fig. 20. lzod impact of supplierA polycarbonate as afunction of


melt pow rate (53).

Ryan (53)found that the Izod impact strength of PC,


as a function of MFR, changes very little on either side strength values can have ductile-brittle transitions as
of the transition between ductile and brittle break. It is much as 35°C apart.
at the transition itself that the dramatic drop in strength Trofimovich, et al. (54), worked with isophthaloyl
occurs. Figures 22 and 23 demonstrate that the transi- chloride-m-phenylenediamine-p-phenylene-diamine
tion temperature changes continuously over the range copolymer and found that its impact strength increased
of MFR. Polycarbonate with the same Izod impact with MW.

214 POLYMER ENGINEERING AND SCIENCE, MARCH. 1982, Voi. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Influence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

CRACKING AND CRAZING range of composition. Kim, et al. (61), demonstrated


General Considerations this, when they studied the kinetics of crack propaga-
tion for PMMA. They found it to depend on one of the
It has been established that crazes are regions of
higher average MW such as iiTZand that the principal
extreme local plastic deformation with strains often
effect of adding a low MW tail was to decrease resist-
exceeding 100 percent (55).Crazes appear to the eye as a
ance to fatigue crack propagation.
hair-line, white crack; however, in reality the craze
Fellers and Huang (60) examined the molecular
contains plastically deformed highly oriented poly-
weight between entanglements, the number of chain
meric material possessing a different density and re-
entanglements per unit volume, and how this entan-
fractive index than the glassy polymer from which it was
glement density varied with MW and MWD and de-
produced (55-57). This is because a craze forms without
rived their fluctuation theory. The general concepts of
significant lateral contraction, hence the strain is dila-
their theory was specifically applied to the crazing of
tant and reduces the polymer density inside the craze by
amorphous PS, from which the following two equations
cavitation (55).Observations of the craze microstructure
were derived. Equation 2 , gives the volume necessary
indicates that the polymer chains are oriented and
to ensure finding a site for microvoid formation to occur,
drawn into fibrils 100-500 A in diameter producing
cast in the energy-related parameters of stress and tem-
voids of the same size range. When the strain hardening
of the microfibrils is not sufficient to compensate for the perature.
void content, the craze is mechanically unstable and
continues to accumulate strain until crack formation and
brittle fracture occur.
Polymer orientation has long been known to affect
craze initiation and breakdown. If the direction of the
applied stress is parallel to the orientation direction, Equation 3 shows how the number of crazes formed per
crazing will be suppressed and brittle fracture delayed. unit volume of the sample is controlled by the MW
Conversely, if the applied stress is perpendicular to the where:
molecular orientation, crazing and fracture will occur B = bulk modulus
more readily than in the isotropic polymer (55,58).For k = Boltzmann constant
stress parallel to the direction of orientation the initial M = molecular weight
voids appear to result from fibrillation, not craze initia- M, = molecular weight between physical entan-
tion (55).Farrar and Kramer (59)found that at similar glements
tensile strain levels parallel crazes were smaller and -
M; = average number of physical entanglements
more numerous than perpendicular crazes and that the per unit volume counted only after the net-
fibril volume was much greater than in perpendicular work is established
crazes in orientated PS. Attempts to explain the affect of To = equilibrium temperature
MW on crazing are presented in the following section. v = volume
Cracking and Crazing WC = stress to initiate crazing
0,
-
- crazing probability
Fellers and Huang (60) suggested a molecular ap- -
- probability of crazing that depends on MW
Urn
proach to explain the phenomena of crazing based on
the density and density distribution of physical entan- When the above equations were combined with
glements between polymer chains as a function of MW. known MW parameters, a model depicting the nature
They state that this approach permits the molecular of the amorphous state was achieved. The model has
entanglement concept and the phenomenological pa- statistical character and shows how variations of M W
rameters such as stress, and temperature associated and chain entanglement control a crazing phenomenon.
with crazing to be related through the use of a fluctua- If a network of entangled chains cannot be established,
tion theory. Results obtained for PS gave an accurate crazing will not occur. They found that as a basic net-
theoretical prediction of the volume associated with work is perfected by increasing the polymer MW, craz-
microvoid formation and showed the dependence of ing will actually be promoted up to a limit.
the number of crazes on MW. Steger and Nielson (55)used small angle x-ray scat-
The higher free volume contributed by the high con- tering, SAXS, to study the deformation process in sev-
centration of chain ends in low MW species serves to eral rubber modified styrene polymers, high impact
dilute the entanglement network and to weaken the polystyrene (HIPS). The increase in scattering intensity
resistance to cyclic loading. The role of entanglements as the specimen crazes was attributed to the formation
depends on the stability of crazes ahead of the advanc- of microvoids in the PS matrix. The relative size, shape
ing crack and, in turn, on the stability of the entangle- and concentration of the scattering elements were de-
ment network, when subjected to cyclic loading. A termined by a Porod analysis of the scattering curves.
spectrum of entanglement stabilities must, of course, Their results indicated that the void sizes are not af-
exist with the higher MW species. As the proportion of fected by M W; however, a lower concentration of voids
low MW species increases from zero the crack growth was produced in the higher Mn HIPS. They also noted a
rate rises very rapidly at first, as the entanglements are decrease in average void size with craze thickening in
diluted, and then rises more slowly at an intermediate these polymers. The fact that Steger and Nielsen (55)

POLYMER ENGINEERINGAND SCIENCE,MARCH, 1982, ~ 0 1 . 2 2 NO.


, 4 215
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R. Martin, and Julian F . Johnson

measured a smaller volume fraction of voids at the the fracture is craze controlled. This transition from a
various strain levels for a higher MW and tougher HIPS brittle glass-like fracture to craze controlled fracture
was at first somewhat bewildering since the conven- occurs between M, = 2.04 x lo4and Mu.= 5.10 x lo4,
tional reasoning was that the higher the impact strength and may be associated with the occurrence of an entan-
of HIPS, the greater the amount of crazing. They con- glement network in the polymer. At even higher MW
cluded that if the weak regions in the higher molecular well-defined crazes are observed whose fibers can ac-
weight HIPS absorb more energy in the cavitation pro- commodate large strains.
cess, or if some fraction of them are strong enough to Brady and Yeh (69), examining PS with an electron
allow shear yielding, then higher impact strength can microscope, found that both craze and shear-band
be obtained with less void formation. morphologies are not strong functions of MW. They
Results have suggested, that a minimum tensile found that regardless of MW, fibrils which formed
stress component has to be reached before crazing can within the deformation zone are always on the order of a
start (58,60,62). Fellers and Kee (58) observed that this few hundred Angstroms in diameter. However, for thin
applied stress required to initiate a craze (often called films with MW < 2 x lo4, the small number of tie
the “critical crazing stress”) is independent of MW. The molecules between fundamental structural units or
magnitude of this stress depends on deformation rate domains made it difficult for these fibers to span the
(63, 64) and temperature (62). However, the gross craze width. This suggests that below a critical MW,
structure of the craze does depend on MW. there are so few tie molecules between domains that the
Warty, et al. (65), suggested that improved fatigue polymer fails before large scale plastic deformation can
performance with increasing M, in PS is a consequence occur.
of increased craze stability resulting from greater de- In another study, Skibo, et al. (71), found for PVC,
gree of chain entanglement and a smaller proportion of that both fracture toughness and crack growth rates
chain end defects. Foden, et al. (66), found that crazes were significantly affected by MW. Fracture toughness
formed on the specimen surfaces soon after application increased with increasing MW while crack growth rates
of the load in both the low and high MW PS samples. In decreased. Crack propagation was found to be pre-
high MW samples, the density of crazing cracks was ceded by crazing and except for the lowest MW pro-
quite high but crazes appeared to stabilize and not grow ceeded in a discontinuous fashion. The craze grew in-
to a critical crack size for propagation. Foden concluded crementally with each fatigue cycle to a characteristic
that the possibility of stabilizing the craze through size, then the crack broke through the craze and the
orientation and fibril formation is not sufficiently high in process began again. The Dugdale plastic strip model
the lower MW samples and the stress concentration served quite well to describe the fracture process, with
effect of crack tips is therefore much greater. the plastic zone of the model being identifiable with the
Wellinghoff and Baer (67), found that although PS of craze. The stability of the crazes was strongly depen-
differing MW exhibited no qualitative morphological dent on MW as was shown by the fact that an increase in
differences in their nucleation and coalescence pro- MW corresponded to an increase in the number of
cesses, MW was found to exert a large effect upon the cycles needed to break a craze. In view of this evidence
cavitation process that immediately preceded fibrilla- they suggested that toughness is governed by the stabil-
tion. In 1 x lo4MW PS samples, very small amounts of ity of the crazes formed and in turn by the entanglement
high MW PS were required to induce void expansion density as has been suggested to explain static fracture
and fibril formation. Without the high MW PS rapid energies in other polymers. To explain the enhanced
crack growth bifurcated the plastic zone. This demon- sensitivity of crack growth rates as compared with static
strated that entanglements in the usual sense were not fracture toughness to MW, one may propose further
responsible for the reinforcement of the microneck. that cyclic loading disentangles the molecules spanning
The first significant void growth and fiber formation a craze and that this effect is more pronounced at lower
occurred at MW z 3.75 X lo4. MW.
Another observation was that above twice the entan- Kusy and Turner (72) state that the plastic contribu-
glement MW, crazing stress was independent of MW tion to fracture surface energy is controlled by crazing
(58, 60) and below twice the entanglement MW, crazes in the vicinity of the crack tip. Moreover, that crazing is
did not form (58, 60, 67-69). However, it was found that dependent on the presence of molecules which are
the number of crazes initiated increased as MW in- sufficiently long to form an entangled crazed network.
creased (58, 60, 70). At the very lowest MW studied, the fracture surface
Lainchbury and Bevis (68) found that, for PS, addi- energy of only several hundred erg/cm2 corresponded
tives and impurities appeared to have more effect than a to an almost purely brittle mode of fracture with little or
wide range of MWD on the crazing and fracture pro- no contribution due to plastic deformation. For PMMA
cess. They also found that at very low MW, well de- with M u> 1 x lo5,they found that fracture morphology
fined crazes did not occur in PS. They concluded there- could be described in conventional terms of mirror,
fore that fracture is not craze controlled at low MW, mist, and hackle regions. At a critical M, = 1 x lo5,
because PS has very little strength and fractures in a significant changes occur on the fracture surface of
brittle manner to give a glass-like fracture surface. At PMMA. Both the mirror region and the rib spacing in
higher MW well defined crazes are observed. Although the hackle region decrease in size, while the inter-
the craze fibers cannot accommodate very large strains ference colors, conical features and mist region largely

216 POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
lnfluence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

disappear, Kusy and Turner also found that as MW crazed ductile matter formed at the surface of higher
gradually decreased to M, = 2.6 x lo3, the fracture MW samples. The thickness of this ductile material
assumed a more glassy appearance. While a drastic increased with MW. Although the presence of ethanol
reduction of rib periodicity occurs, shatter-cone pat- vapor increased the tendency towards local ductility,
terns appear with greater frequency. Also both river the MW effects were far more dramatic.
patterns and Wallner lines are evident as the percent of Wales (79) found that crazes were initiated earlier in
featureless surface increases. higher MW PVC; however, the difference was not large
Weidman, et al. (73), investigated the effect of both considering the normal wide range in initiation times.
MW and a fracture mechanic parameter K on the length He stated that the role in craze initiation by MW was
of the stretched molecular bundles in a crazed zone and only through its influence on the localized moulding
in particular, the maximum length at the onset of crack stresses present during sample preparation. Therefore,
propagation in PMMA. They showed that the maximum a material of higher MW, higher viscosity, would tend
length increased from 1.2 ,um to 2.9 ,um and the crack to have an earlier craze initiation. This would also ex-
tip crack opening from 0.4 Prn to 1.4,urn with increasing plain the higher craze densities for higher MW PS
MW, a, greater than 1.1 x lo5but less than 8 million. found by Fellers and co-workers (58). Wales’
However the rate of increase was only large at lower Mu. hypothesis would also explain the contrary results
between 1.1 x lo5 and 2 x lo5. found by Gothan (80), whereby an injection moulded
Doll (74) found that his work dealing with PMMA grade of low MW PMMA crazed earlier and more pro-
supported the assumption that the distances between fusely than a higher MW PMMA that had been cast.
the lines or rib-like features, which occur in an approx- Howard (81)and Nalivaiko and Sirota (82)found that
imately regular manner on fracture surfaces of high the M W of PE had a significant influence on stress crack
polymers accompanying fast fractures, were dependent resistance. Their results agreed with others (83,84) that
on craze length and stress. He found that the lines in as MW increases so does stress crack resistance. Ban-
PMMA of aw = 1.63 x lo5have a larger spacing and are dyopadhyay (85) also found that resistance to environ-
formed at a higher value of strain energy release rate mental stress cracking for PE increased with increasing
than those in PMMA of = 1.1 X lo5or 1.2 X lo5.He MW and that quenched materials showed greater re-
also found that craze zone lengths which can be esti- sistance than slowly cooled ones. Figure 24 shows that
mated from the Dugdale model (75) of the plastic zone even a slight increase in the melt index, a MW increase,
at a crack tip are also greater in PMMA of a MW equal to
1.63 x lo5 (75). Such differences in strain energy re-
lease rate and length of craze zone are not, however,
sufficient to explain why no regular line markings are
found on the fracture surfaces of high MW PMMA with
aa,,. > 4.9 x 10’. The observations in this work indi- ‘OO*k
cated that the occurrence of these lines are dependent
on MW. Doll therefore concluded that it seems reason-
able to infer that this is a reflection of a molecular chain
length dependence and that there is a limiting length
above which no line markings will be formed.
Bigg, et al. (76), found that higher MW PMMA
homopolymer required a longer time for stress cracking
fluid, n-methyl formamide and methyl alcohol, to cause
cracking at the minimum residual stress level, and that
it cracked at a higher minimum threshold stress.
Pitman, Ward and Fraser (22,77)found that the craze
shapes in PC agreed reasonably well with those pre-
dicted by the Dugdale plastic zone model. There was
also an excellent agreement between the strain energy
release rate calculated from the craze shape in the plane
strain. They found that the craze dimensions were very
dependent on MW. This arose from a dependence of
both crack opening displacement and craze stress on
MW. In addition, although the energy required to pro-
duce unit volume of shear lip remained fairly constant,
the width of the shear lip was highly MW dependent.
From these results it follows that both the strain energy U f 2 3 4 5
release rate in plane strain and the overall measured
strain energy release rate are very MW dependent. M.F.I., g/lO miti
Martin and Johnson (78)working with PVC found that Fig. 24. Dependence of the resistance to stress cracking (in 20%
the fraction with the lowest MW gave brittle fracture surfactant OP-7 solution, 50°C) of PE on the melt pow index
when fatigued in nitrogen or ethanol vapor, and walls of (82).

POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4 217
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R . Martin, and Julian F . Johnson

causes a considerable reduction in the resistance to 26). Even for extraordinarily high MW material, the
stress cracking. critical strain values were unchanged. This agreed with
Buniyat-Zade and Azimova (86)investigated PE and the earlier findings of Bergen (92). However, it has been
found the cracking resistance to decrease as MWD in- found that the subsequent stages of craze and crack
creased. On that basis they concluded that low growth are highly dependent on polymer MW (93) as is
molecular fractions reduce cracking resistance. They illustrated in the relaxation curves in Fig. 27 for three
explained this by the fact that PE is a highly crystalline samples differing only in MW. The curves show that
polymer. In such polymers relatively small amounts of each sample was strained above its critical value and
low molecular fractions can act, on the completion of yielded upon exposure to the methanol crazing agent.
crystallization, as an object of competition between The lowest MW material failed totally in a few seconds,
neighboring spherulites. When the resulting stresses while the higher MW polymers retained some load
reach values e q u a l t o t h e s t r e n g t h of t h e non- bearing capability over longer times.
crystallized material which is located between the
spherulites, microcracks are formed which act as foci of FATIGUE AND ENDURANCE
breakdown. Comparable defects on the surface of a General Considerations
polymer act as centers of adsorption of active agents,
Fatigue is defined as failure or degradation of
with subsequent development cracking. These defects
mechanical properties d u e to oscillatory strain or stress.
in the internal layers of a polymer will eventually lead Fatigue life is defined as the number of cycles of oscilla-
to thermal or fatigue cracking. They also found that the tion N before fracture.
incorporation of a very small amount of C2H5branching
Fatigue is a function of stress; therefore, as one in-
in PE strongly influenced cracking resistance (Fig. 25). creases the maximum stress'per cycle, the number of
Henry (87) evaluated styrene acrylonitrile copoly-
cycles before failure decreases. Below some value of
mers according to the factors which affect craze initia-
maximum stress, the fatigue limit, the fatigue curve
tion and those which limit the rates of flaw propagation.
levels off, and the material can be subjected to a very
He considered MW, styrene acrylonitrile ratio, rubber large number of cycles without failure (2). Results
content, and rubber phase morphology as the principal
variables which affect flaw initiation and propagation.
Plastics exposed to organic media were found to fail
at stresses much less than their yield stresses. The
tendency of glassy plastics to crack or craze in chemical
environments appeared closely related to the degree of
plasticization which results in the plastics when the
6.
5-1
organic reagent was at its equilibrium level (88-90).To a ethanol
first approximation this tendency can be described by
the different solubility parameters of the polymer and
reagent. The mechanical parameter which most realisti-
cally describes the onset of craze related yielding in an
environment is the critical strain. Bergen (91)observed
that plastics which are strained below the critical value
for an environment did not yield through craze and I I I I I I /4
crack formation while those strained above the critical 6 .8 1.0 1.2 1.4 1.6 1.8 -3-5
value failed. He found no effect of MW on the critical INTRINSIC VISCOSITY IN CHLOROFORM @ 3OoC
strain (strain at which crazing first occurs) for crack and Fig. 26. Dependence of critical strain on molecular weight (in-
craze initiation in poly(styrene co-acrylonitrile), (Fig. trinsic viscosity) of polyistyrene-co-ucrylonitrile)(87).

I Applied Strain z 0.7%

W
V

( MVIJ 85,000)
I I I I I 1

7 14 0 2 4 6 8 10 2
v
W, 0
T I M E ( min 1
(L
CONTENT, '10 Fig. 27. Enuironmentul stress reluxation curues us a function of
Fig. 25. Dependence of cracking resistance of ethylene butene-1 molecular weight (intrinsic viscosity) f o r poly(styrene-co-
copolymer on the % butene-l (86). ucylonitrile) i n methanol (87).

218 POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
lnjluence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

shown in Fig. 28 show the scatter of data points one can polymers of high MW. Gent and Thomas (12)found that
expect for fatigue tests, and the spread will increase as high MW suppressed crazing in PVC. However, Kam-
stress amplitude is lowered (94). This scatter reflects the bour (99) and Fellers and Kee (58) have observed that
emphasis of surface conditions in fatigue failure which is the initiation of crazing in PS was insensitive to MW.
due to unavoidable variations in surface geometry from Kim, et al. (98), decided to work on clarifying this
specimen to specimen. distinction by studying the effect of MW on FCP in
Fatigue life generally decreases as the temperature precracked samples. The relationship used in charac-
increases and as the frequency of oscillations increases. terizing FCP in polymers is (100):
Fatigue life is often given as (95):
da
Log N = A 4- B I T
dN = AAKn (5)
(4)
where N is the fatigue life in a number of cycles, A and B where da/dN is the crack growth rate per cycle, A and n
are constants, and T is the absolute temperature. are material constants, AK is the range in stress intensity
Fatigue failure is generally due to progressive growth factor, K , at the crack tip equal to Y A d , where Y is a
of cracks. This crack growth or tearing process is found geometrical factor for the specimen shape used, (101)
in both rigid polymers and rubbers. Materials almost A u is the range in applied stress, and a is the crack
always contain flaws which develop into submicroscopic length. The relationship between da/dN and AK allows
cracks as the stress exceeds some critical value. These the comparison of da/dN at a given AK level for differ-
cracks continue to grow until they induce failure. Basi- ent MW.
cally the factors which affect cracking and crazing will Fatigue tests were performed on an MTS electrohy-
have an affect on the fatigue life. Hence, fatigue life is draulic testing machine at a cycle frequency of 10 Hz, and
increased by those factors that decrease crack propaga- the crack propagation was followed by a moveable mi-
tion. The affect of MW and MWD is discussed in the croscope (98). These FCP tests were performed on
following section. Overall the trend found is for in- PMMA samples with a, range of 4.8 X lo6to 1.0 X 10'.
creased MW to increase fatigue life up to a limiting It was found that FCP rates were dependent on MW
MW. even when fracture toughness K , changed little (98).
Results by Skibo et al. (71), were similar for PVC.
Fatigue and Endurance
Figure 29 shows the fatigue crack growth of PVC as a
In spite of the importance of fatigue behaviour in function of AK for MW between 2.25 x 106and 6 x lo4.
engineering plastics, very little work has been done The curve demonstrates a continuous shift to higher
concerning MW and its influence on fatigue life and growth rates with decreasing MW for PVC. Kim, et al.
fatigue crack propagation (FCP). The work that has been (98), drew up an exponential relationship between FCP
done has shown a considerable decrease in fatigue life rate and 1IMW. It was proposed that this specific fatigue
with decreasing MW. This has been demonstrated on effect was a consequence of the disentanglement of
PVC with MW between 5 x lo4 - 2 x 10' (96) and PS polymer networks in the craze preceding the main
with a MW range of 1.6 x 10' to 8.6 X lo5 (66, 97).
The majority of tests were performed on unnotched
fatigue specimens (98). Therefore one could not deter-
mine whether the decrease in fatigue life with decreas-
ing MW was due to an increase in the ease of crack
initiation or by a more rapid crack growth. It has been
suggested that an increase in the strength of the crazes
which precede the crack tip would contribute to the
increase in fracture stress and fatigue life of glassy

5
lo5
?a3 lo4
106 10' 02 05 I0 20

Cycles to failure AK,M P a f i


Fig. 28. Stress amplitude us. cycles to failure f o r H I P S and PS Fig. 29. Relationship betweenfatigue crack propagation in PVC
(94). and stress intensity factor range AK at 0, 6, and 13% DOP(71).

POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4 219
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R. Martin, and Julian F . Johnson

crack; hence the lower the MW, the greater the ease of with crystallites imbedded in an amorphous matrix, it i s
disentanglement. It was also proposed that the incorpo- more resistant to crack propagation (104) than is the
ration of a plasticizing comonomer n-butyl-acrylate re- amorphous PS.
sulted in complex FCP behavior. Woan, et al. (94), found the fatigue results for HIPS
Various methods are used to test fatigue life, several to be inferior to that of PS when the fatigue lifetimes
of which were either developed or improved upon by were compared on a basis of absolute stress magnitude,
Ohishi, et al. (102). Their bending fatigue tests on PC Fig. 28. They concluded that a lower matrix MW may
showed that the effect of MW was irregular. Also, blend- be a contributing factor to the observed lower fatigue
ing PE with PC decreased normal bending fatigue resistance.
strength. Filling with glass fiber was found to be advan-
tageous for testing under constant force, but disadvan- C R E E P A N D STRESS RELAXATION
tageous for constant deformation. The influence of General Considerations
molding conditions was also apparent. Creep and stress relaxation tests measure the dimen-
For torsional properties, Ohishi, et al. (102), de- sional stability of polymers, generally over a long
veloped a new testing apparatus called “MO-U”. The period of time and are therefore of great practical use.
apparatus is capable of measuring torsional fatigue Generally for elastomeric materials the apparatus is
strength under constant amplitude of deformation and quite simple; however, for more rigid materials the
static torsional properties. The results showed that ex- apparatus becomes more complex as the deformations
cellent fatigue strength was obtained with unfilled, become much smaller. Creep and stress relaxation are
high MW samples which were molded under optimum essentially the inverse of one another (58).
conditions. It was also found that specimens prepared At temperatures well below T , where polymers are
by machining were inferior to those prepared by injec- brittle, MW has a small effect on creep and stress
tion molding. This is not surprising because the surfaces relaxation. This results from very short segments of the
of molded specimens are normally smooth, oriented molecules involved in molecular motion in the glassy
and under compressive stress. These factors substan- state. Motion of large segments of the polymer chains is
tially increase durability. frozen in, and the restricted motion of small segments
A modified impact fatigue tester which applies an can take place without affecting the remainder of the
impulsive force to a specimen by dropping a weight molecule (57). The following two sections discuss the
repeatedly was used for evaluating impact fatigue influence of MW on creep and stress-relaxation at tem-
strength (102). The results showed that the PC samples peratures near and above T, where dependence upon
with lower MW have lower impact fatigue strength. MW becomes more prominent.
Also, the blendingof PE with PC, and imperfect drying
of pellets have negative effects on impact fatigue resist- Stress Relaxation
ance. This would be expected since, inadequate drying It is generally accepted that two types of relaxations
causes a reduction in MW d u e to hydrolysis of the PC govern the viscoelastic properties of concentrated so-
while in the injection molding press. Incomplete drying lutions and melts of linear polymers (106). These two
also results in a poor surface finish which reduces im- mechanisms are the intramolecular motions of seg-
pact fatigue properties. ments and the molecular motions involving the adjust-
Pitmann and Ward (103) found for PC a substantial ments and shifting of chain entanglements (107-109).
decrease in resistance to FCP as MW of the polymer The intramolecular relaxation times are independent of
decreased. Unstable fracture was found to occur at MW, whereas entanglement relaxation times increase
lower values of h K with decreasing MW. with MW (45, 110-112). The presence of two mecha-
A Tatnall-Krause fatigue instrument was used to nisms can be observed in the appearance of the plateau
demonstrate that PS fatigue life increases significantly region, as in the storage modulus versus frequency plot,
with increasing MW (65, 66, 104). Sauer, et al. (104), or in the creep compliance versus time plot. What is
found the improvement continued even beyond the observed is a minimum in the relaxation or retardation
point where T, is independent of MW, and stated that spectra (45, 110-114). These features become clearer
this improved fatigue resistance of PS could be d u e to with increasing MW (106).
higher resistance to both craze inception and craze The typical experimental procedure for measuring
breakdown. stress relaxation is to apply a strain to the specimen in as
For crystalline PE an increase in impact fatigue per- short a time as possible, then hold this strain constant
formance with increasing MW was found (104, 105). and measure the stress as it decays with time (26). Short
Sauer, et al. (104), found greater than 100 percent in- and long term continuous stress relaxation characteris-
crease in endurance limit as MW increased over the tics of polyurethane samples of varying MW were de-
range 5 x 103 to 5 x lo4. termined at varying temperatures by Schollenberger
The greater resistance of crystalline PE to fatigue and Dinbergs (5). They found for short term testing that
failure compared to amorphous PS was attributed to two polyurethane undergoes substantial stress relaxation
characteristics. First, PE has a much lower T, and con- but at a rate which decreases with increasing MW. This
sequently greater chain mobility in its amorphous is shown in Fig. 30. They suggest that this decrease in
phase. Therefore it is less sensitive to surface or volume stress relaxation rate is d u e to the increase in the
flaws. Secondly, because of its heterogeneous nature, number of crosslinks by hydrogen bonds and hard seg-

220 POLYMER ENGfNEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Injluence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

melts to MWD. His model assumed the relaxation


spectrum was progressively truncated with increasing
shear rate under steady shearing flow conditions. The
maximum allowed relaxation time, T ~ where , the spec-
trum is truncated, is assumed to be only a function of
shear rate: 7, = K / y where K 2 2. Therefore at this
in -
Mw
Polymer -
- given shear rate, the molecular species of molecular
A 47,900 weight M act as if it were of molecular weight M , .
D Il7,lOO Because the shear-rate dependence of the relaxation
(L H 289,100
0
J 366,800 spectrum was assumed to be only the effect of trunca-
8 40 tion, zero-shear (linear viscoelastic) relationships for the
melt viscosity, steady shear elastic compliance, and first
I 10 I00 loo0 10,000 normal stress difference were used to calculate these
TIME ( secl properties as a function of shear rate by including the
F i g . 30. S t r e s s relaxation us. D,(f o r thermoplastic shear-rate truncation of the relaxation spectrum (120).
polyurethanes (5). Marshall and Petrie (121) found that the enthalpy of
relaxation of atactic PS proceeded at a somewhat faster
ments, as well as the increase in chain entanglements rate, at comparable temperature intervals below the
per chain as MW increases (5). These increases would respective T,’s ( T , - T.,,),for lower MW PS. The rate of
decrease chain mobility and reduce stress relaxation. enthalpy relaxation for each M W was expressed as: (122)
A long term experiment (Fig. 31 ), shows a small stress
117 = A exp(-E,/RT) exp[C(Qs - QJ] (6)
relaxation dependency on a,.
at 30°C (5). This effect of
stress relaxation decreasing with increasing MW was where E, is the apparent activation energy of enthalpy
found to be more pronounced at 100°C and reflects the relaxation, A and C are constants for each MW, and t is
higher strains and temperature. This reduction in re- the annealing period. It was found that the slope A(QE
laxation rate decreasing with increasing MW was - QJA log t decreased somewhat with increasing MW
also found by Greco on commercial polyisobutylene up to a,. = 5 x lo4 (123). Above this MW, the depen-
(115). dence of the relaxation rate upon MW appeared to be
Murakami, et al. (116),looked at the stress relaxation negligible. This result suggested that molecular chain
of mixtures above T , for PS. They found that a mixture dimensions have a small effect on the limited segmental
containing 50 percent 1 x lo4MW and 50 percent of 5 . 0 mobility occurring in the glassy state. Consistent with
X lo5 eliminated about 40 percent of the rubbery re- this result was their observation of a slight dependence
laxations of the high MW component and decreased of relaxation rate on MWD involving low MW poly-
the maximum relaxation time by three orders of time. mers.
Onogi, etal. (117), showed that a60percent mixture o f 6 Creep
x lo5 MW PS with a 2.5 x lo4PS component reduced Wilding, et al. (123), looked at irrecoverable creep of
the fracture strain of the high MW component by 50 PE and found that the high MW grade displayed a
percent and the fracture stress by an order of magnitude critical stress below which irrecoverable creep fell to a
for low strain rates above T,. It was concluded that low negligible level. Increasing the A decreased the com-
M W components have a strong effect on long relaxation pliance at all times. The stress dependence of the creep
times and act as aplasticizer for high MW polymers (67). behavior was markedly non-linear at longer times espe-
Recently, Bersted (118-120) proposed a model that cially at low A. These authors (123) fit the creep and
would relate the viscoelastic response of PE and PS recovery behavior for drawn PE to a series combination
of a Maxwell and aVoigt element at each level of stress.
100 -
The results showed that the viscoelastic behavior of
a - these materials has two components: First a linear vis-
w
z coelastic recoverable component and second a non-
Ea: 80- linear irrecoverable component which can be modeled
as an activated creep process, the Eyring model. The
v) -
w
v)
recoverable viscoelastic behavior of PE was found not to
a:
tl be greatly affected by A or M W, apart from the increase
_I
a
60- Y in stiffness with increasing A. The irrecoverable creep
L A 47,900 A
ua: - D 117,100
component, however was affected by both A and
0 ti 289,100 MW.
k 40- J 366.800 For lower MW PE, reduction in irrecoverable creep
s with increasing A was attributed to gradual exhaustion
of the barrier sites as initial lamellar texture of the
0 01 01 10 100 100.0 1000 material was removed. High MW samples, on the other
TIME ( hrs I hand, showed a critical stress for irrecoverable creep, as
F i g . 31. Stress relaxation us. a!, f o r thermoplastic well as a much smaller activation volume for the creep
polyurethanes (5). process. Wilding, et al. (123), suggested that the ir-

POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4 22 1


15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R . Martin, and Julian F . Johnson

recoverable creep of the low MW PE was related to a Glass Transition


crystal deformation process, such as pulling out of the A generally accepted expression relating T, to MW is
chain folds by a crystal slip process. (98, 127-130):
Bhateja (124) recently examined the 100-h constant-
load uniaxial tensile creep response of two ultra-high T, = T,, - K / M , (7)
MW linear PE material and compared it to that of a derived by Fox and Flory (131). Where T, is in "K
normal molecular weight linear polyethylene, NMW observed for a polymer of a given M,,T,, is the T, of the
LPE. His data clearly showed that at all stress levels polymer when the M W approaches infinity, and K is a
examined, the creep deformation at a given time was constant whose units are O K mole gg'. This equation is
much higher in the UHMW LPE than in NMW LPE. based on the assumption that T, occurs at a constant
There are two counteracting factors which could explain fraction of free volume (132).
this effect, MW and crystallinity. Increased MW would Eyuation 7 has been restated by Bueche (133) (Ey8)
increase chain entanglements and one would expect taking into account the total increase in free volume d u e
this to reduce creep. However, he states that in view of to the number of chain ends per cm3(2p NM-'; where p
the larger creep deformation at any time t in UHMW = density, N = Avogadro's Number). Each end contri-
LPE, the lower levels of crystallinity found (UHMW buting a free volume change of 8 cm3 is related to the
LPE = 50%, NMW LPE 70%) must be overriding depression of the glass transition temperature (T,= - T,)
the MW affect. by:
The creep behavior of P E bars under uniaxial loading
has been categorized as one of two types. One type T , = T,, - ( 2 p ~ e / 4 ( 1 / ~ ) (8)
occurs for relatively high stresses, where, after some a = cq - a, where q and agare the volume coefficients
time, necking takes place as a result of plastic flow. The in the liquid and glassy states respectively.
other occurs for small stresses where the material frac- It has been noted (134, 135) that if the MW range is
tures at small deformations d u e to crack formation and large enough, the T , shows three phases of dependence
growth (125). For H D P E there is a very narrow region of on MW. One occurs at high MW where there is no
load in which one can obtain highly strained materials dependence. Phase I1 is at intermediate MW and is
(125, 126). It has been found (128) that for H D P E and characterized by moderate decreases in T, as MW is
HMWPE samples molded at pressures ranging from reduced. Phase I11 is at even lower MW (oligomeric)
100 to 500 MPa, creep strain is lower for high MW where the decrease in T, is even more pronounced with
materials. Zapas and Crissman (125) found that there decreasing MW. In general it was found (Fig. 32), that
exists a narrow region of stress for which one can obtain for polymers with a low Tgmthe chain length at which
extension ratios of up to 25 at 296°K before fracture one first achieves the limiting T, is much shorter than for
occurs in PE. The location of this region is a function of polymers with a high Tgz (135).Thus for poly(dimethy1-
MW. siloxane) the asymptotic value of T , is reached in the
GLASS TRANSITION vicinity of a degree of polymerization equal to 100 ( X =
General Considerations loo), while for poly(cx-methylstyrene) X approaches 600
Amorphous polymers and crystalline polymers with before T,, is reached. Equations to predict the three
an amorphous component are hard rigid-like glasses regions have been developed by Cowie (135). Agree-
below T,. Mechanical properties show an enormous ment with experimental data is only fair.
change in the vicinity of their T, as they go from the Couchman (136) also worked on a theory that would
glassy state to a more rubber-like state as the tempera- predict T , for homopolymer systems of arbitrary
ture passes through T,. For instance, elastic modulus polydispersity. His theory requires knowledge of four
may change by a factor of 1000 as the temperature is
raised through the glass transition region (59). 440
There are several factors which determine the Tgof a 400 -
Region I
I
polymer. Chain stiffness of course is very important, I

and as side groups which add flexibility to the polymer 360 -


are added onto the backbone of a polymer the T, will be 320 -
lowered. On the other hand if rigid groups are incorpo-
rated into the backbone or side groups, such as aromatic
-280 -
P
structures, the To will increase. For illustration: P E 240 -
(-120°C), PP(-lO"C), PS (100°C) (57).
Molecular polarity is also important: increasing the 200-
polarity of a polymer will cause its T, to increase.
Backbone symmetry also plays an important role by - --
^ ^ -- -
affecting bond rotation. Generally, symmetrical poly- ILU
t- : .<,;; . , . , , .1, , I . , , , , , , , II . . J
. . ..
mers have lower T,'s than the unsymmetrical polymers 1 10 100 1000
(57). MW and MWD also play a vital role in determin- Fig. 32. Plot of T,("K) as a function of log X , the number of
ing what the T , of a polymer will be. The next section backbone chain atomslbonds, for 0-poly(a-methylstyrene;
examines the work which has been recently carried out A-PMMA; 0 - P V C ; A-isotactic P P ; 0 - a t a c t i c PP;
in this area. 8 -poly(dimethylsiloxane) (I 35).

222 POLYMER ENGiNEFRlNG AND SCIENCE, MARCH, 1982, Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
lnfiuence of Molecular Weight and Molecular Weight Distribution on Mechanicul Properties of Polymers
I ’ 1
properties to acquire the T,: high-polymer and chain-
end transition temperatures and their corresponding
transition increments of heat capacity.
Kumler, et al. (134), measured the T , of PS using an
ESR spin-probe technique. They found that the Tg of
“monodiyerse” PS varied smoothly as a function of Mn.
At high M , , the asymptotic value of T , was 100°C. As
MW decreased, the T , fell to a value of -140°C for

-I
styrene monomer. Figure 33 graphically shows how
ESR results compare to other methods of obtaining the
T , of PS.
A study by Chee and Rudin (137)indicated that vol-
umes attributable to end groups differ significantly be-
tween anionic and thermally initiated polymers. They 30
70 80 90 I00 110 I20
found that specific volumes of end groups in anionic PS
ELUTION VOLUME, ml
liquids are larger and increase more rapidly with tem-
perature in the range of 170-230°C. Hence one might Fig. 34. GPC distribution of a polystyrene .sample reconstituted
expect that the T, values of anionic and thermally initi- w i t h f o u r low dispersity components (A, B , C , und D ) (130).
ated PS would coincide at H M W and differ sig-
nificantly at LMW, where chain end concentration are Combining E q 7 with E q 9 yields:
more important. However, Rudin and Burgin (138)
found that t h e same T , / M dependence could b e
applied to both anionic and thermally initiated PS
polymers.
For anionically prepared PS, M W = 900 to 1.8 X lo6,
Blanchard, et ul. (130), found that the experimental
results for T , agreed with E q 7, in the form of T , = 106 (10)
- 2.1 x 105/M,. However, when c o m p a r i n g i t y i t h The results calculated for the thermal PS sample
thermally prepared PS, M, = 1.35 x lo5,with M,./Mn of yielded a value of 104°C which was still higher than the
3.2, they determined a T, of 93”C, 10°C lower than the experimental result of 93°C. They concluded that fac-
value predicted from the above equation. Hypothesiz- tors over and above MWD were contributing to the
ing that the low T , from the thermal PS was d u e to broad discrepancy, such as the presence of low MW species in
dispersity, they reconstituted the sample with four low the disperse sample.
disperse components (seeFig. 34). They rewrote E q 7 to Gillham, et al. (139), using torsional braid analysis
include the effect of a highly polydisperse sample made examined the relaxation transition located above T , in
up of known weight percentages ( N l , N 2 , N J , N 4 ) of four amorphous polymers. Evidence for such a transition
low polydisperse c o q o n e n t s , each having a known M,, and its behavior have been summarized (140), and since
(M,,M 2 , M 3 , M 4 ) . M , is then related to N and M as the transformation involves change from one liquid
follows: state to another it has been designated the symbol Ti/.
They found that for PS, the relaxation Tl1 behaves like
-+-
Nl N2 +-+---
N3 N4 -
100
- (9) the glass transition in its dependence on MW. Plots of
MI M2 M3 M4 Mn Ti!vs 1/M, displayed two regions, similar to plots of T ,
vs l/@,. MW at which the relationship varied corre-
sponded to the critical MW for entanglements.
For PMMA Kim, et al. (98), found that T , increased
100 -
with increasing M, u p to a MW of 1.9 x 10’ where it
essentially remained constant giving the inverse rela-
50 - tionship described by E q 7. They found TgEto equal
111°C and K = 1.28 x lo6.
0 - For thermoplastic polyurethane elastomers. Schol-
lenberger and Dinbergs (5)found that T , increases with
-50 - increasing a,
and levels off at M,,-- 1.6 x 10‘. Like
previous workers, they related this pattern to a decline
-i in free volume with increasing MW and reduction in
-100 X
chain end concentration.
k; Prevorsek, e t al. (141), d e t e r m i n e d that
L I I I I I I I
25 3.5 4.5 5.5 poly(bispheno1-A t e r e p h t h a l a t e carbonate) and
LOG iSi, poly(bispheno1-A carbonate) (PC) results could be de-
Fig.33. Glass transition temperature of PS as a function of M , , scribed by E y 7 in agreement with Adam, et al. (128).
as determined by various methods: (0) and (a)dilatometry, (W) For poly(N-vinyl carbozole) Bergfjord, et al. (142),
DTA, (C)DSC, ( X ) E S R (134). found that T , measurements could be described by E 9

POLYMER ENGINEERING AND SCIENCE, MARCH, 1982. Vol. 22, No. 4 223
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R. Martin, and Julian F. Johnson

7. Their K value was about 20 percent larger than that microphases in styrene-isoprene diblock and triblock
reported by Griffith (143). copolymers and in styrene-ethylene oxide diblock and
Alfthan and DeRuvo (144) noted that poly(ethy1ene triblock copolymers in which the styrene blocks had
oxide) did not follow the theoretically predicted in- comparable M W. It appears therefore that the chemical
verse relationship given by E 9 7; rather, the Tgreached nature of other blocks which are attached to the glassy
a maximum at intermediate MW (4 x lo3 to 1 x lo4). microphase have little effect on Tg.The chemical nature
This anomalous behavior was attributed to the presence andlor MW of the rubbery block does influence Tg of
of a crystalline phase which restrains the chain mobility the glassy microphase only when the MW and composi-
in the amorphous regions. In terms of free volume, this tion of the block copolymer is near that at which mi-
implies a corresponding reduction on the free volume crophase separation no longer takes place. This was
of the amorphous phase (145). Such a result may be demonstrated by Toporowski and Roovers (149) where
achieved either by compressional restraints or by the they found for styrene-isoprene low MW diblock and
interaction of adhesion between crystalline regions and triblock copolymers, that Tgdecreased with decreasing
the amorphous chain segments. They went on to show size of styrene blocks. It also increased slightly as the
that longer annealing times caused an increase in Tg. size of isoprene blocks were increased holding the
They explained this dependence in terms of crystalliza- styrene content constant.
tion kinetics. Bosnyak, et aE. (150), polymerized a range of
Daniels and Collins (129) examined the Tg for PVC. hydroxy-terminated polybisphenol A terephthalate and
Their data, Fig.35,shows the variation of Tgand T, for a isophthalate blocks having a MW of 800 to 5 x lo3,
variety of PVC polymers polymerized over a broad which were then coupled with phosgene to yield alter-
temperature range, (-85 to + 125°C). From this figure it nating polyester co-polycarbonates. They found that
is evident that, across a very wide range of polymeriza- the values for Tg followed E9 7, showing a marked
tion temperature and MW, a relationship exists be- dependence on an.
tween chemical characteristics (syndiotacticity and
MW) and Tgand T,. Pezzin (146), has shown that TQfor MELTING POINT
PVC vary with II?,according to the Fox-Flory equation,
where, in their work Tgoo= 351°K and K = 8.05 x lo4. The expected overall trend for homopolymers is for
Pezzin found that above M, = 1 x lo4, the dependence lower MW polymers to melt at lower temperatures than
of Tgon I/@, decreased. Daniels, et al. (129), reported higher MW samples as given by E q 11 (151):
the Tg of PVC exhibited E 9 7 behavior for polymeriza-
tion temperatures between -15 and 70°C. Their work
does indicate that both chain length and tacticity affect
the free volume of PVC. where T , is in degrees Kelvin, T", is the melting point of
For a 20:80 acrylonitrile-methyl acrylate copolymer, a polymer with an infinite MW, Mo is the MW of the
Zabotin, et a2. (147), found that the T , was virtually monomer, R is the gas constant and A H p is the heat of
independent of MW, but the transition temperature fusion per mole of the crystalline polymer repeating
into the viscoelastic state increased linearly with in- unit.
creasing MW. Clements, et al. (152), and Perkins, et al. (15), found
Krause and Iskandar (148) examined Tg in diblock the melting point of drawn PE to increase with increas-
copolymers of styrene and dimethyl siloxane as a func- ing MW. Clements (152) also found that for A greater
tion of styrene MW. They found that, within experi- than 15 and constant MW, higher drawing tempera-
mental error, the Tgof styrene microphases was exactly tures raise T,.
the same as Tg taken from the literature for styrene Capaccio and Ward (31, 153)showed the effect that
MWD has on the melting behavior of highly drawn PE:
In Fig. 36 sample A(2) has a T, of 137°C; however,
sample B(2), with a narrower MWD and lower au.
shows a higher Tmof 139°C. They found this result to be
unexpected, and went on to explain it in terms of the
more homogeneous structure of polymer B(2). They
also suggested that the formation of extended chain
material is related to a critical MW range, i.e. a critical
value of mw,For very narrow MWD the result becomes
completely independent of the initial morphology and
A. However, with broader MWD, the phenomenon of
extended chain formation occurs only partially and only
at high A. Therefore, the smaller MWD showed a
a0I I higher T, even though it had a lower M,.
-Do -7s -w -t) 0 t) so 75 loo IS1 In polyurethane elastomers, Schollenberger and
POLYMERIZATION TEMPERATURE T 'C Dinbergs (5) found that a distinct melting endotherm
Fig. 35. Glass transition temperature and crystalline melting which consistently peaked between 67-68°C had no
point of PVC's polymerized at various temperatures (129). apparent dependence on MW.

224 POLYMER ENGINEERING AND SCIENCE, MARCH, 1982. Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Influence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

n
'2
x 1.3 -
.-
P
W

f 1.2 -
c
C
a
.-
+
sample B(2) +
-W 1.t -
A=34 0
I

2 1.0
LL
-
sample A ( 2 )
0.9 -

\
A.29
0.8 I I I I 1 I I l l I . , I

5 I I I I I I Ilp I I I

120 130 140 150 160 170 10 10 4x 1


Molecular W e i g h t
Fig. 37. Flexural stiffness of HDPE us. molecular weight (3).
Te rnpera ture ("C)
Fig. 36. Melting curves of PE samples B(2) a n d A ( 2 )after draw-
ing (31).

Predecki and Karr (154) found that the T , for


polyoxymethylene increased with both time and tem-
perature of crystallization as well as with increasing
MW.
In PVC, Daniels and Collins (129)concluded that it is
not the chain length, but rather the tacticity which
controls the melting point of the crystallites.
Mukherjea and Saha (155) found for moisture cured
polyurethanes prepared from diisocyanate and sebacic
acid-based hydroxy esters, that the lower MW esters
produced a polyurethane having a higher T , than those
of the corresponding higher MW polyester polyols of
the same glycol. They speculated that this increase in 44
I
I I
I
I
I
I
I
I
I
1 1 1 1
I l l
6
I
I
I
"
t
6
T , was due to the presence of a higher percentage of 10 10 4x 10
relatively stiff segments of aromatic groups in hydroxy Molecular Weight
ester based urethane than in polyester polyol-based F i g . 38. Hardness of HDPE us. molecular weight (3).
urethane.
MISCELLANEOUS occurs. On the other hand, Schollenberger and Din-
Stiffness bergs (5) found no systematic change in hardness with
Stiffness is one of the important mechanical property increasing polyurethane MW when measured by Shore
requirements, along with strength, for a material to be A durometer. A slight trend toward lower Shore D
useful in reinforcement. Strength and stiffness are hardness in the lower MW samples was observed.
closely related theoretically (156) however, in practice Deanin, et al. (13), showed, Fig. 39, that indentation
attainment of a high modulus does not always result in hardness of plasticized PVC was moderately depen-
proportionate increases in strength. Khukhreva, et al.
E ecl Of MWD ON Hardness
(4), found that the stiffness of silicon modified PC films
increased as a,. increased. However, Margolies (3)
I
stated that stiffness is a function of crystallinity. H e
found that as MW of linear PE increases, percent crys-
tallinity, and hence the stiffness, decreases (Fig. 37).
The stiffness continues to decrease with increasing MW
up to a MW of 1.75 x lo6. Above this point, the stiffness
plateaus and does not significantly change with either
decreasing crystallinity or increasing MW. W
IT
0
Hardness 5 30
1 2 3 42 4 6 8 0 4 8 12
Margolies (3) found that with the Rockwell type R
test, hardness of PE decreased with increasing MW up Mn lo4 M~ lo4 M~ lo4
to 1.75 X lo6, (Fig.38), beyond which no further change a,vand
F i g . 39. Hardness of PVC versus LG,~, a,(13).
POLYMER ENGINEERING AND SCIENCE. MARCH, 1982, Yo/. 22, No. 4 225
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Ronald W. Nunes, John R . Martin, and Julian F . Johnson

dent on MW. Their data suggests that the high MW Figs. 4, 8, 17, 18, 20, 21, 22, 23, 26, 27, 35, 36, 37, 38
fraction was controlling. Their data suggested that and Table I: Society of Plastics Engineers, Inc., pub-
hardness depended upon disentanglement of the lisher of Polym. Eng. Sci. and SOC. Plast. Eng. ].
longest chains rather than on segmental mobility. Fig. 16: Dr. Dietrich Steinkopff Verlag, publisher of
On the other end of the scale the softening tempera- Rheol. Acta.
ture increases in a manner proportional to MW (3). Figs. 1 , 2 9 and 33: Marcel Dekker, Inc., publisher of
J. Macromol. Sci., Phys. and “Mechanical Properties of
Tear Strength Polymers and Composites”, Vol. 2.
Figs. 24 and 25: Rubber and Plastics Research Associ-
Schollenberger and Dinbergs (5) and Ono, et al. (19), ation of Great Britain, publisher of Sou. Plast. Eng.
found that the tear strength of polymers decreased with Transl.
increasing M W. However, Schollenberger and Din- Fig. 13: Heyden and Sons, Inc., publisher of J.
bergs (5) have shown, in Fig. 40, that this change for Therm. Anal.
polyurethane is not regular. Rather there is a relatively Fig. 19: McGraw-Hill, Inc., publisher of Mod. Plast.
rapid decrease in polymer tear strength as MW in- Figs. 5, 12, 14, 30, 31 and 40: Technomic Publishing
creases for Mu.< 1 x lo5. Above this point, a plateau Co., Inc., publisher of]. Elastomers Plast.
occurs with no change in tear strength until a second Figs. 3, 11 and 39: Division of Polymer Chemistry,
inflection is reached at M,. = 2.8 x lo5.There is a slow Inc., American Chemical Society, publisher of Polym.
decrease in polymer tear strength as M, exceeds 2.8 x Prepr., Am. Chem. SOC., Div. Polym. Chem.
lo5.They suggested that the decrease in tear strength of Fig. 10: Dr. Pae, publisher of Technical Report #2,
virtually crosslinked polymer with increasing polymer Office of Naval Research, NR 356-564.
M, parallels the decrease in polymer free volume. It Fig. 34: National Research Council of Canada, pub-
may also include a factor which relates the tear strength lisher of Can. ]. Chem.
of conventional elastomers to covalent crosslinking. Fig. 28: IPC Science and Technology Press, Ltd.,
Apparently a high MW virtually crosslinked publisher of Polymer.
polyurethane, with its high degree of chain entangle-
ments and virtual crosslinks, (and thus reduced chain REFERENCES
mobility) cannot adjust by rearrangement of polymer 1. J. R. Martin, J. F. Johnson and A. R. Cooper,]. Macromol.
chains to relieve applied stresses as readily as similar Sci.-Revs. Macrornol. Chem., C8, 57 (1972).
2. L. E. Nielson, “Mechanical Properties of Polymers and
low MW polymers. Composites,” Vol. 2, Marcel Dekker Inc., New York(1974).
3. A. F. Margolies, S.P.E. ]., 27, 44 (1971).
ACKNOWLEDGMENTS 4. I. I. Khukhreua, A. S. Barkov, S . V. Dereuyagina, M. A.
Verkhotin, and V. N. Kotrelev, Soti. Plast., Engl. Transl., 1,
The authors gratefully acknowledge the following for 50 (1972).
their permission to reproduce the indicated figures. 5. C. S. Schollenberger and K. Dinbergs,]. Elast0mer.s Plast.,
Figs. 6, 7, 9 and 15: John Wiley and Sons, Inc., 11, 58 (1979).
6. J. Vlachopolous, N. Hadjis and A. E. Hamiliec, Polymer,
publishers of 1. Polym. Sci., Parts C and 1. Appl. 19, 115 (1978).
Polym. Sci. 7. P. W. 0.Wijga, “Physical Properties of Polymers,” p. 35,
Fig. 32: Pergamon Press, Ltd., publishers of Eur. Soc. Chem. Ind., Monograph #5, Macmillan, New York
Polym. J. (1959).
Fig. 2: Communication Channels, Inc., publisher of 8. H. F. Mark, Adhes. Age, 7, 35 (1979).
9. J. A. Sauer, Polymer, 19, 858 (1978).
Adhesives Age. 10. D . M. Shinozaki, J . Vlachopoulos, K. Woo, and
A. Hamielel, J. Appl. Polym. Sci., 21, 3345 (1977).
11. H. B. Lee and D. T. Turner, PoZym.Eng. Sci., 19,95(1979).

-f 500 I 12. A. N. Gent and A. G. Thomas,]. Polym. Sci. Part A-2, 10,
571 (1972).
13. R. D. Deanin, P. R. Desai, and C. R. Wilson, Polym. Prepr.,
Am. Chem. SOC., Diu. Polym. Chem., 12, 652 (1971).
14. R. J . Gardner and J. R. Martin,J. A p p l . Polym. Sci., 24,1269
(1979).
15. W. G. Perkins, N. J. Capiati and R. S. Porter, Polyrn. Eng.
Sci., 16, 200 (1976).
16. W. Wuand W. B. Black,Polym. Eng. Sci., 19,1163(1979).
% I \ 17. D . W. Larsen, S . P . E . ]., 27, 40 (1971).
18. J. S. Foot and I. M. Ward, ]. Mater. Sci., 7, 367 (1972).
19. K. Ono, H. Shimada, T. Nishimura, S. Yamashita, H .
Okamoto and Y . Minoura, ]. Appl. Polym. Sci., 21, 3223
(1977).
20. M . Morton, J. E. McGrath and P. C. Juliano,]. Polym. Sci.

POLYMER
AB D EF C H 1 J
I I
, Part C , 26,99 (1969).
21. T . Alfrey, Jr., “Mechanical Behavior of High Polymers,”
pp. 493, 494, Interscience Publishers, Inc., New York
(1948).
22. G. L. Pitman and I. M .Ward, Polymer, 20, 895 (1979).
23. K. D. Pae, B. A. Newman and T. P. Sham, Technical Report
#2, Office of Naval Research, NR 356-564, Jan., (1975).

226 POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
InfEuence of Molecular Weight and Molecular Weight Distribution on Mechanical Properties of Polymers

24. H. W. Hill, Polym. Prepr., Am. Chern. SOC., Diu. Polym. 67. S. Wellinghoff and E. Baer,J. Macromol. Sci. Phys., B-11,
Chem., 20, 173 (1979). 367 (1975).
25. S. Matsuoka and T. T. Wang, Coat. Plast. Prepr. Pap Meet, 68. D. L. G . Lainchbury and M. Bevis,J. Mater. Sci., 11,2222
Am. Chem. SOC., Diu. Org. Coat. Plast. Chem., 36, 155 (1976).
(1976). 69. T. E. Brady and G. S. Y. Yeh,J. Mater. Sci., 8,1083(1973).
26. B. Maxwell and M. Nguyen, Polym. Eng. Sci., 19, 1140 70. J. F. Rudd, Polym. Lett., 1, l(1963).
(1979). 71. M. Skibo, J. A. Manson, R. W. Hertzberg and E. A. Collins,
27. M. N. Nguyen and B. Maxwell, Polym. Eng. Sci., 20,972 J. Macromol. Sci. Phys., 14, 525 (1977).
(1980). 72. R. P. Kusy and D. T. Turner, Polymer, 18, 391 (1977).
28. F. W. Billmeyer, “Textbook of Polymer Science,” p. 221, 73. G. W. Weidmann and W. Doll, Colloid. Polym. Sci., 254,
John Wiley & Sons, New York (1962). 205 (1976).
29. J. H. Grezlak and G. L. Wilkes,J. Appl. Polym. Sci., 19,769 74. W. Doll, J . Mater Sci., 10, 935 (1975).
(1975). 75. D. S. Dugdale, J. Mech. Phys. Solids, 8, 100 (1960).
30. K. Matsushige, S. V. Radcliff and E. Baer, J . Appl. Polym. 76. D. M. Bigg, R. I. Leininger and C. S. Lee, Polymer, 22,539
Sci., 20, 1853 (1976). (1981).
31. G. Capaccio and I. M. Ward, Polym. Eng. Sci., 15, 219 77. R. A. W. Fraser and I. M. Ward, Polymer, 19, 220 (1978).
(1975). 78. J. Martin and J. F. Johnson, J . Polym. Sci., Phys. Ed., 12,
32. L. Jarecki and D. J. Meier, Polymer, 20, 1078 (1979). 1081 (1974).
33. P. J. Barham and A. Keller, J. Mater. Sci., 11,27 (1976). 79. J. L. S. Wales, Polymer, 21, 684 (1980).
34. G. Meinel, A. Peterlin and K. Sakaoku, “Analytical 80. K. V. Gotham, Plast. Polym., 40,277 (1972).
Calorimetry,” p. 135, Plenum Press, New York (1968). 81. H. R. Howard, S.P.E. J., 27,31 (1971).
35. A. J. Wills, G. Capaccio and I. M. Ward, J. Polym. Sci., 82. E. I. Nalivaiko and A. G. Sirota, Sou. Plast. Eng. Transl., 9,
Phys. Ed., 18, 493 (1980). 17 (1967).
36. J. F. Fellers and T. F. Chapman, J. Appl. Polym. Sci., 22, 83. J. B. Howard, S.P.E. J., 15, 397 (1959).
1029 (1978). 84. H. A. Stuart, Kunststoffe, 54, 618 (1964).
37. R. P. Kusy and A. R. Greenberg,J. Therm. Anal., 18, 117 85. S. Bandyopadhyay,J. Polym. Sci., Phys. Ed., 19,749(1981).
(1980). 86. A. A. Buniyat-Zade and A. B. Azimova, Sov. Plast., Eng.
38. G. Holden, E. T. Bishop and N. R. Legge, J . Polym. Sci., Transl., 6, 38 (1970).
Part C, 26, 37 (1969). 87. L. F. Henry, Polym. Eng. Sci., 14, 167 (1974).
39. M. Morton, Adu. Chem. Ser., 99,490 (1971). 88. G. A. Bernierand R. P. Kambour, Macromol., 1,393(1968).
40. R. F. Fedors, J . Polym. Sci., Part C , 26, 189 (1969). 89. R. P. Kambour, C. L. Gruner, and E. E. Romagosa, Bull.
41. Y. D. Chen and R. E. Cohen,J. Appl. Polym. Sci., 21,629 Am. Phys. SOC., 17,340 (1972).
(1977). 90. R. P. Kambour, E. E. Romagosa and C. L. Gruner, Mac-
42. N. K. Kalfoglou, J . Appl. Polym. Sci., 23, 2385 (1979). romol., 5, 335 (1972).
43. C. M. Boussias, R. H. Peters and R. H. Stil1,J. Appl. Polym. 91. R. L. Bergen, Jr., S . P . E . J., 24, 77 (1968).
Sci., 25, 869 (1980). 92. R. L. Bergen, Jr., S . P . E . J., 18, 667 (1962).
44. G. Martin and W. W. Graessley,Rheol. Acta, 16,527(1977). 93. J. P. Berry, J. Polym. Sci., 2, 4069 (1964).
45. S. Onogi, T. Masuda and K. Kitagawa, Macromolecules, 3, 94. D. Woan, M. Habibullah, J. A. Sauer, Polymer, 22, 699
109 (1970). (1981).
46. S. K. Ahuja, Rheol. Acta, 19, 299 (1980). 95. J. H. Dillon, “Advances in Colloid Science,” Vol. 3, p. 219,
47. S. Gazit, Ph.D., Dissertation, University of Connecticut, H. Mark, E. J. W. Venvey, Ed., Interscience, N.Y. (1950).
1980. 96. J. R. Martin and J. F. Johnson,]. Appl. Polym. Sci., 18,3227
48. P. I. Vincent, “Impact Tests and Service Performance of (1974).
Thermoplastics”, Plastics Inst., London (1971). 97. J. A. Sauer, E. Foden and D. R. Morrow, SPE ANTEC Tech.
49. R. A. Horsley, Appl. Polym. Sympos., 17, 117 (1971). Pap., 22, 107 (1976).
50. M. Matsuo, A. Vedaand Y. Kondo, Polym. Eng. Sci., 10,253 98. S. L. Kim, M. D. Skibo, J. A. Manson and R. W. Hertzberg,
(1970). Polym. Eng. Sci., 17, 194 (1977).
51. S. Cohen, Mod. Plast., 4, 88 (1974). 99. R. P. Kambour, J . Polym. Sci., Part D,7, l(1973).
52. P. J. Perron and P. B. Lederman, Polym. Eng. Sci., 12,340 100. P. Paris, and F. Erdogan, J. Basic Eng., 85, 528 (1963).
(1972). 101. W. F. Brown and J. E. Srawley,ASTM Spec. Tech. Puhl.,
53. J. T. Ryan, Polym. Eng. Sci., 18,264 (1978). 410, 142 (1966).
54. A. N. Trofimovich, S. I. Tverdokhleb and L. N. Fomenko, 102. F . Ohishi, S . Nakamura, D. Kogama, K. Minabe,
Chem. Abst., 80, 486116d (1974). Y. Fujisawa and Y. Tsuruga, J. Appl. Polym. Sci., 20, 79
55. T. R. Steger and L. E. Nielsen,]. Polym. Sci., Polym. Phys. (1976).
Ed., 16, 613 (1978). 103. G. Pihnan and I. M. Ward, J. Mater. Sci., 15, 635 (1980).
56. J. P. Berry, J. Polym. Sci., 50, 107 (1961). 104. J. A. Sauer, E. Foden and D. R. Morrow, Polym. Eng. Sci.,
57. R. P. Kambour, J . Polym. Sci. A., 3, 1713 (1965). 17, 246 (1977).
58. J. F. Fellers and B. F. Kee, J. Appl. Polym. Sci., 18,2355 105. S. K. Bhateja, J. K. Reike and E. H. Andrews,]. Mater. Sci.,
(1974). 14, 2103 (1979).
59. N. R. Farrar and E. J. Kramer, Polymer, 22, 691 (1981). 106. Y. Isono, T. Fujimoto, N . Takeno, H. Kjiura and
60. J. F. Fellers and D. C. Huang,]. Appl. Polym. Sci., 23,2315 M. Nagasawa, Macromolecules, 11, 888 (1978).
(1979). 107. A. V. Tobolsky, .“Properties and Structures of Polymers,”
61. S. L. Kim, J. Janiszewski, M. D. Skibo, J. A. Manson and John Wiley, New York (1960).
R. W. Hertzberg, Polym. Eng. Sci., 19, 145 (1979). 108. J. D. Ferry, “Viscoelastic Properties of Polymers,” John
62. S . S. Sternstein, L. Ongchin and A. Silvermann, Appl. Wiley, New York (1970).
Polym. Symp., 7 , 175 (1968). 109. W. W. Graessley, Ada. Polyrn. Sci., 16, 1 (1974).
63. P. Beardmore and T. L. Johnston, Phil. Mag., 23, 119 110. T. Masuda, K. Kitagawa and S. Onogi, Macromolecules, 3,
(1971). 116 (1970).
64.J. Murrayand D. Hull,]. Polym. Sci., Part A, 8,1521(1970). 111. N. N. Moto, Polym. J., 1,485 (1970).
65. S. Warty, J. A. Sauer and A. Charlesky, Eur. Polym. J., 15, 112. T. Fujimoto, N. Ozaki and M. Nagasawa, J. Polym. Sci.,
445 (1979). Part A-2, 6, 129 (1968).
66. E. Foden, J. A. Sauer and D. R. Morrow, J. A p p l . Polym. 113. W. Dannhauser, W. C. Child and J. F. Ferry,]. Colloid Sci.,
Sci., 16, 519 (1972). 13, 103 (1958).

POLYMER FNGINHRING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4 227
15482634, 1982, 4, Downloaded from https://4spepublications.onlinelibrary.wiley.com/doi/10.1002/pen.760220402 by Fudan University, Wiley Online Library on [11/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Rotitrltl W.Nirties, J o h t t €3. Murtiti, tirid Jii1iuit F. Johnson
114. J. W. Berge and P. R. Saunders, J. D. Ferry,J. Colloid Sci., 137. K. K. Chee and A. Rudin, J . Mecromol. Sci. Phys., 7,503
14, 135 (1959). (1973).
115. R. Greco, Eur. Polym. J., 14,233 (1978). 138. A. Rudin and D. Burein. Polumer. 16. 291 (1975).
“ I J , I

116. K. Murakami, K. Ono, Shuna, T. Veno and M. Matsuo, 139. J. K. Gillham, J. A. Benci and R. F . Boyer, Polym. Eng. Sci.,
Polym. J., 2, 698 (1971). 16. 357 (1976).
117. S. Onogi, T. Matsumoto and E. Kamei, Polym. J., 3, 531 140. R. ’F. Boyer, J. Polym. Sci., Part C, 14, 267 (1966).
(1972). 141. D. C. Prevorsek, Y. Keston and B. Debona, Polym. Prepr.,
118. B. H. Bersted, J. A p p l . Polym. Sci., 20, 2705 (1976). Am. Chem. S O C . , Div. Polym. Chem., 20, 187 (1979).
119. B. H. Bersted, J. Appl. Polym. Sci., 19,2167 (1975). 142. J. A. Bergfjord, R. C. Penwell and M. Stolka,]. Polym. Sci.,
120. B. H. Bersted, J . Appl. Polym. Sci., 23, 1279 (1979). Phys. Ed., 17, 711 (1979).
121. A. S. Marshall and S. E. B. Petrie, J. Appl. Phys., 46,4223 143. C. H. Griffith and V. Laeken,]. Polym. Sci., Phys. Ed., 114,
(1975). 1433 (1976).
122. S. E. B. Petrie, J. Polym. Sci. Pert A-2, 10, 1255 (1972). 144. E. Alfthan and A. DeRuvo, Polymer, 16,692 (1975).
123. M. A. Wilding and I. M. Ward, Polymer, 19,969 (1978). 145. R. F. Boyer, Rubber Chem. Technol., 36, 1303 (1963).
124. S. K. Bhateja, Polymer, 22,23 (1981). 146. G. Pezzin, Eur. Polym. J., 6, 1053 (1970).
125. L. J. Zapas and J. M. Crissman,PoZym. Prepr., Am. Chem. 147. K. P. Zabotin, V. F. Pomytkina and V. K. Panyshev, Tr.
Soc., Diu. Polym. Chem., 19, 805 (1978). Khim. Khim. Tekhnol., 1, 100 (1973).
126. K. Djurner, J. Kjbat and M. Rigdahl, Polymer, 18, 1068 148. S. Krause and M. Iskandar, “Advances in Chemistry,
(1977). Series #176; Multiphase Polymers,” S. L. Cooper, G. M.
127. G. J. Pratt, J. Meter. Sci., 10, 809 (1975). Estes, Ed., Amer. Chem. SOC.(1979).
128. G. A. Adam, J. N. Hay, I. W. Parson and R. N. Haward, 149. P. M. Toporowski and J. E. L. Roovers, 1. Polym. Sci.,
Polymer, 17, 51 (1976). Chem. Ed., 14, 2233 (1976).
129. C. A. Daniels and E. A. Collins, Polym. Eng. Sci., 19,585 150. C. P. Bosnyak, I. W. Parsons, J. N. Hay and R. N. Howard,
(1979). Polymer, 21, 1448 (1980).
130. L. P. Blanchard, J. Hesse and S. L. Malhotra, Can. J. 151. P. J. Flory, “Principles of Polymer Chemistry,” Cornell
Chem., 52, 3170 (1974). University Press, Ithaca, N.Y. (1953).
131. T. G. Fox and P. J. Flory,J. Appl. Phys., 21,581 (1950). 152. J. Clements, G. Capaccio, and I. M. Ward, J. PoZym. Sci.,
132. D. T. Turner, Polymer, 19, 789 (1978). Phys. Ed., 17,693 (1979).
133. F. Bueche, “Physical Properties of Polymers,” p. 76, In- 153. G. Capaccio and 1. M. Ward, Polymer, 15,233 (1974).
terscience Publishers, New York (1962). 154. P. Predecki and P. H. Karr, Polym. Eng. Sci., 18, l(1978).
134. P. L. Kumler, S. E. Keinath, and R. F. Boyer,J. Macromol. 155. R. N. Mukherjea and K. K. Saha,]. Appl. Polym. Sci., 25,
Sci. Phys., 13, 631 (1977). 2699 (1980).
135. J. M. G. Cowie, Eur. Polym. J., 11,297 (1975). 156. A. H. Cottrell, Proc. R. SOC. (London), Ser. A, 282(1388),2
136. P. R. Couchman, J . Muter. Sci., 15, 1680 (1980). (1964).

228 POLYMER ENGINEERING AND SCIENCE, MARCH, 1982, Vol. 22, No. 4

You might also like