You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233978569

On the Viscoelastic Poisson’s ratio of Amorphous Polymers

Article  in  Journal of Rheology · September 2010


DOI: 10.1122/1.3473811

CITATIONS READS

31 116

3 authors:

Luigi Grassia Alberto D'Amore


Università degli Studi della Campania "Luigi Vanvitelli University of Campania Luigi Vanvitelli (formerly The Second Uni…
56 PUBLICATIONS   481 CITATIONS    117 PUBLICATIONS   1,308 CITATIONS   

SEE PROFILE SEE PROFILE

Sindee L Simon
Texas Tech University
187 PUBLICATIONS   3,818 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Glass Transition of TNT View project

National Science Foundation grants, DMR 0308762 and 0606500 View project

All content following this page was uploaded by Luigi Grassia on 28 June 2016.

The user has requested enhancement of the downloaded file.


On the viscoelastic Poisson’s ratio in amorphous
polymers

Luigi Grassiaa) and Alberto D’Amore

Department of Aerospace and Mechanical Engineering, The Second University of


Naples, Via Roma 19, 81031 Aversa (CE), Italy

Sindee L. Simon

Department of Chemical Engineering, Texas Tech University,


Lubbock, Texas 79409-3121

(Received 25 January 2010; final revision received 18 May 2010;


published 24 August 2010兲

Synopsis
Poisson’s ratio is defined as the ratio of the lateral contraction to the elongation in the infinitesimal
uniaxial extension of a homogeneous isotropic body. In a viscoelastic material, Poisson’s ratio is a
function of time 共or frequency兲. In this paper, the time-dependence of the Poisson’s ratio is
analytically evaluated from the bulk and shear responses using the relations between the
viscoelastic functions in the Laplace domain. It has been found that, in the region of ␣-relaxation,
Poisson’s ratio may be a nonmonotonic function of time, with a weak minimum at short times,
when the shear response is broader than bulk response such that the ratio ␶G / ␶K is much larger than
1, or a monotonically increasing function of time if the shear and bulk responses share similar
timescales and relaxation time distributions. The latter case is verified using experimental data
from the literature for a cross-linked polymer, whereas the former case is verified for two linear
polymers. © 2010 The Society of Rheology. 关DOI: 10.1122/1.3473811兴

I. INTRODUCTION
We reserve the term viscoelastic Poisson ratio for the lateral contraction ratio mea-
sured in an infinitesimally small uniaxial deformation of a viscoelastic material in re-
sponse to a strain, ␧1共t兲 = ␧0 · h共t兲, as a unit step function, h共t兲, of time. Thus, ␯共t兲
= −␧2共t兲 / ␧0, where ␧0 is the imposed constant strain and ␧2共t兲 is the time-dependent
lateral contraction in the transverse direction 关Tschoegl et al. 共2002兲兴. The experimental
characterization of viscoelastic Poisson’s ratio is probably as difficult as the determina-
tion of the bulk functions 关Tschoegl et al. 共2002兲; O’Brien et al. 共2007兲兴 as evidenced by
the amount of reliable literature data. In fact, several studies present Poisson’s ratio data,
but the data are limited to narrow temperature and time ranges or large strains 关Bogy et
al. 共1979兲; Yoe and Takemori 共1982兲; Giovagnoni 共1994兲; Delin et al., 共1995兲; Tsou et al.
共1995兲兴. Measurements of Poisson’s ratio at room temperature on a relatively wide time-
scale were performed by Tsou et al. 共1995兲 on cellulose acetate and polycarbonate films

a兲
Electronic mail: luigi.grassia@unina2.it

© 2010 by The Society of Rheology, Inc.


J. Rheol. 54共5兲, 1009-1022 September/October 共2010兲 0148-6055/2010/54共5兲/1009/14/$30.00 1009
1010 GRASSIA, D’AMORE, and SIMON

by means of independent experiments in bending and tension. The curve of Poisson’s


ratio vs time for polycarbonate was essentially flat, whereas that for cellulose acetate was
roughly parabolic with a definite maximum around 103 s. O’Brien et al. 共2007兲 reported
on a series of Moiré interferometry experiments used to simultaneously measure the
creep compliance and time-dependent Poisson’s ratio of epoxy resins in the vicinity of
their glass transition temperatures 共Tgs兲. The results appear among those more reliable for
the accuracy of the measurement techniques at elevated temperatures when materials fall
within the rubbery state. The data showed that the viscoelastic Poisson’s ratio is a mono-
tonically increasing function of time. Interestingly, a monotonic scaling function between
Poisson’s ratio and the shear relaxation modulus was also reported by plotting Poisson’s
ratio versus the shear relaxation modulus determined independently.
Calculation of Poisson’s ratio from the shear and tensile viscoelastic functions requires
high precision in the input data, owing to the nature of the interrelation equations
关Tschoegl et al. 共2002兲; Lakes 共1992兲; Lakes and Wineman 共2006兲兴. In fact, in polymers
as well as in other materials, the viscoelastic properties depend on temperature, aging
time after preparation, as well as time/frequency. Therefore the input viscoelastic func-
tions should be measured using the same specimen, in the same environment, at the same
time, and with high accuracy and precision, as mentioned by Tschoegl et al. 共2002兲. Kono
共1960兲 reported the first simultaneous measurements of the dynamic bulk and shear
relaxation responses by using the transverse and longitudinal wave propagation tech-
niques. These data, re-analyzed by Mott et al. 共2008兲, show that the bulk response occurs
before the shear one and that Poisson’s ratios for both polystyrene and poly共methyl
methacrylate兲 show a minimum at low temperature, which in the framework of time-
temperature superposition corresponds to having a minimum at short time. A nonmono-
tonic Poisson ratio showing a minimum at short time was also observed by Koppelmann
共1958兲 for poly共methyl methacrylate兲 at room temperature. The data reported by Tsou et
al. 共1995兲, O’Brien et al. 共2007兲, Koppelmann 共1958兲, and Mott et al. 共2008兲 seem
contradictory at first glance, but they are both consistent with the theory of linear vis-
coelasticity. In fact, Lakes 共1992兲 and Lakes and Wineman 共2006兲 in their discussion of
the matter concluded that “the viscoelastic Poisson’s ratio need not increase with time,
and it need not be monotonic with time.” Finally, we mention that the viscoelastic Pois-
son ratio is obtainable under stress relaxation experiments 关Tschoegl et al. 共2002兲兴, so that
one might consider comparing its placement with the relaxation moduli. However, in
tensile stress relaxation experiments, the axial behavior is a relaxation process, whereas
the transverse strain is a retardation process. Therefore, the characteristic times of the
viscoelastic Poisson ratio are neither the relaxation nor the retardation times; rather they
are called delay times.

II. ANALYSIS: A NUMERICAL EXAMPLE


Calculation of the Poisson ratio can be performed by means of the algebraic relation-
ships between the viscoelastic functions in the Laplace domain 关Ferry 共1980兲; Chris-
tensen 共1982兲兴. One such correlation has the following form:

1 3K共s兲 − 2G共s兲
␯共s兲 = , 共1兲
s 6K共s兲 + 2G共s兲
where ␯共s兲, K共s兲, and G共s兲 are the Laplace transform of the Poisson ratio, the bulk
relaxation modulus, and the shear relaxation modulus, respectively. Thus, the time-
dependence of the viscoelastic Poisson ratio is obtainable if the bulk and the shear
relaxation moduli are known.
POISSON’S RATIO IN AMORPHOUS POLYMERS 1011

In what follows, we analyze Eq. 共1兲 in order to examine the relevant parameters
determining the shape of Poisson’s ratio and its timescale placement with respect to the
shear and bulk moduli. Let us first assume the simplest case in which the shear the bulk
moduli, G共t兲 and K共t兲, respectively, relax exponentially, each with a single relaxation
time:

G共t兲 = G⬁ + 共Go − G⬁兲exp − 冋 册 t


␶G
, 共2兲

K共t兲 = K⬁ + 共Ko − K⬁兲exp −冋 册 t


␶K
, 共3兲

where the subscripts ⬁ and 0 indicate the relaxed and the unrelaxed moduli, respectively,
and ␶G and ␶K are the shear and the bulk relaxation times, respectively. In the Laplace
domain, Eqs. 共2兲 and 共3兲 become
G⬁ 共Go − G⬁兲
G共s兲 = + , 共4兲
s s + 1/␶G

K⬁ 共Ko − K⬁兲
K共s兲 = + . 共5兲
s s + 1/␶K
Then, the Laplace transform of Poisson’s ratio can be obtained from Eq. 共1兲 as follows:
cs␶G关3aK0共1 + s␶G兲 − 2G0共1 + as␶G兲兴 + b关3aK0共1 + s␶G兲 − 2c共G0 + aG0s␶G兲兴
␯共s兲 = ,
2s兵cs␶G关G0 + 3aK0 + a共G0 + 3K0兲s␶G兴 + b关3aK0共1 + s␶G兲 + c共G0 + aG0s␶G兲兴其
共6兲
where a = G0 / G⬁, b = ␶G / ␶K, and c = K0 / K⬁. Transforming Eq. 共6兲 back into the time
domain, the time-dependence of Poisson’s ratio can be obtained allowing us to manage it
in analytical form. However, the explicit expression of Poisson’s ratio as function of time
is not reported here for the sake of simplicity as its whole expression is very large.
Based on the works by Mott et al. 共2008兲 and Koppelmann 共1958兲, we recall that
Poisson’s ratio can be a nonmonotonic function of time, showing a minimum at short
time if lim共d␯ / dt兲 ⬍ 0, according to the fact that its limiting values are such that
t→0
lim␯共t兲 = 3K0 − 2G0 / 2共3K0 + 2G0兲 ⬍ lim ␯共t兲 = 3K0 / c − 2G0 / a / 2共3K0 / c + 2G0 / a兲. This last
t→0 t→⬁
condition is true as long as a = G0 / G⬁ ⬎ c = K0 / K⬁. Hence, the nonmonotonic behavior of
Poisson’s ratio at short times can be checked by evaluating the sign of its derivative at
short times, namely, lim共d␯ / dt兲. In the short-time limit, the derivative of Poisson’s ratio
t→0
is given by the following:
d␯ 9G0K0兵a关b共− 1 + c兲 − c兴 + c其
lim = . 共7兲
t→0 dt 2ac共G0 + 3K0兲2␶G
The contour plot of lim共d␯ / dt兲 is reported in Fig. 1 for the case of G0 = 1 GPa, K0
t→0
= 4 GPa, ␶G = 1 s, and K0 / K⬁ = 2. It is clear from the figure that the limiting Poisson ratio
derivative is negative 共i.e., lim共d␯ / dt兲 ⬍ 0兲 when ␶G / ␶K Ⰷ 1, whereas it is positive 共i.e.,
t→0
lim共d␯ / dt兲 ⬎ 0兲 when ␶G / ␶K Ⰶ 1, as long as log关G0 / G⬁兴 ⬎ 1 共a condition that is always
t→0
1012 GRASSIA, D’AMORE, and SIMON

LogΤG ΤK 
20
0 5.0
1.0
0.20
1 0
0.06
2 0.095
0.105

3
0 1 2 3 4 5
LogG0 G 

FIG. 1. Contour plot of lim共d␯ / dt兲 for the case of G0 = 1 GPa, K0 = 4 GPa, ␶G = 1 s, and K0 / K⬁ = 2. The legend
t→0
on the right highlights the values of the limiting short-time derivative of Poisson ratio.

verified for amorphous polymers in the region of ␣-relaxation兲. Hence, Poisson’s ratio
shows a local minimum if the bulk relaxation modulus evolves faster than the shear
relaxation modulus.
For the sake of clarity, in Fig. 2, Poisson’s ratio is plotted as a function of time for
different values of the ratio ␶G / ␶K highlighting the effect on ␯共t兲 of the relative positions
on the time axis of the shear and bulk responses. The curves are obtained for the same
conditions displayed in Fig. 1 共i.e., G0 = 1 GPa, K0 = 4 GPa, ␶G = 1 s, K0 / K⬁ = 2, and
log关G0 / G⬁兴 = 4兲. From Fig. 2, it should be noted that Poisson’s ratio goes from a mono-
tonic to a nonmonotonic behavior, showing a deeper minimum at short times as the
timescale of the shear response translates toward longer times with respect to the bulk
response.
The effect of the bulk relaxation strength, K0 / K⬁, on the shape of the Poisson ratio
response is illustrated in Fig. 3 where the contour of lim共d␯ / dt兲 = 0 is plotted for different
t→0
values of K0 / K⬁. The contours do not depend on G0, K0, and ␶G. From Fig. 3, one can

0.50
ΤG ΤK 102
ΤG ΤK 10
0.45 ΤG ΤK 100.5
ΤG ΤK 102
0.40
Ν

0.35

0.30
104 0.001 0.01 0.1 1 10 100
ts

FIG. 2. Time dependence of the viscoelastic Poisson ratio calculated as a function of the relative position on
the time axis of the shear and bulk responses as indicated by the values of ␶G / ␶K.
POISSON’S RATIO IN AMORPHOUS POLYMERS 1013

2
K0 K 1.5
K0 K 2.0
K0 K 3.0
lim Ν t 0 K0 K 10
1 t0

LogΤG ΤK 
0

1 lim Ν t  0
t0

2
0 1 2 3 4 5
LogG0 G 
FIG. 3. Contours of lim共d␯ / dt兲 = 0 for different values of K0 / K⬁. The contours do not depend on G0, K0, and
t→0
␶ G.

observe that the value of ␶G / ␶K above which lim共d␯ / dt兲 is negative 共which corresponds
t→0
to a nonmonotonic time-dependence of Poisson’s ratio兲 increases as the ratio K0 / K⬁
decreases.
The ratio ␶G / ␶K that corresponds to lim共d␯ / dt兲 = 0 can be evaluated from Eq. 共7兲:
t→0

冏 冏
␶G
␶K lim 共d␯/dt兲=0
=
−c+a·c
a共− 1 + c兲
, 共8兲
t→0

where c = K0 / K⬁ and a = G0 / G⬁. From Eq. 共8兲, it comes out that ␶G / ␶K 兩 lim 共d␯/dt兲=0 does not
t→0
depend strongly on a = G0 / G⬁ when log关G0 / G⬁兴 ⬎ 1. To better clarify, we report in Fig.
4 the ratio ␶G / ␶K as function of K0 / K⬁ for lim共d␯ / dt兲 = 0 and log关G0 / G⬁兴 ⬎ 1. It should
t→0

1000

100
lim Ν t 0
t0
ΤG ΤK

10

1 lim Ν t  0
t0

0.1
0 1 2 3 4 5
K0 K

FIG. 4. ␶G / ␶K vs K0 / K⬁ evaluated at lim共d␯ / dt兲 = 0 and log关G0 / G⬁兴 ⬎ 1.


t→0
1014 GRASSIA, D’AMORE, and SIMON

be noted that as the bulk relaxation strength decreases, the value of ␶G / ␶K above which
lim共d␯ / dt兲 is negative increases.
t→0
From this simplified analysis, it is clear that, in the framework of ␣-relaxation where
G0 / G⬁ Ⰷ 1, the viscoelastic Poisson ratio is a nonmonotonic function of time with a
minimum at short times when the ratio ␶G / ␶K is much larger than 1, or a monotonically
increasing function of time if the shear and bulk responses share similar timescales. The
latter case is verified for cross-linked polymers while the former case is verified for linear
polymers, as pointed out in Sec. III.
We note that our simplified model also predicts that ␯共t兲 shows a maximum when
共G0 / G⬁兲 ⬍ 共K0 / K⬁兲共␶K / ␶G兲 and ␶G / ␶K ⬍ 1. This condition is not observed in the frame-
work of ␣-relaxation since G0 / G⬁ is large. However, in the region of the ␤-relaxation,
the condition may be met and Poisson’s ratio may show a maximum since one expects
that the ratio G0 / G⬁ to be smaller for the ␤-relaxation given that the relaxation is usually
not as strong as the ␣-relaxation 共where G0 now refers to the glassy modulus at times
shorter than the ␤-relaxation and G⬁ refers to the glassy modulus at times longer than the
␤-relaxation兲.
The presence of a maximum in the data of Tsou et al. 共1995兲 for cellulose acetate is
dubious and is not supported by the present analysis. Indeed, Tsou et al. 共1995兲 have
calculated Poisson’s ratio at room temperature from the normalized tensile relaxation
modulus, EN, and the normalized bending relaxation modulus, DN, both evaluated experi-
mentally. In the limit of linear viscoelastic behavior, EN and DN should be very close each
other. This implies that small errors in the determination of EN and DN can give large
errors in the calculation of ␯共t兲. Furthermore, from the work of Tsou et al. 共1995兲, a
strong ␤-relaxation does not seem to be present for cellulose acetate, and consequently it
seems that their result is due to inaccuracies in EN and DN. As pointed out in the Intro-
duction, the importance of using precise data to obtain ␯共t兲 by interconversion has been
thoroughly discussed in the literature 关Tschoegl et al. 共2002兲; Lakes 共1992兲; Lakes and
Wineman 共2006兲兴.

III. EVIDENCE COMING FROM LITERATURE DATA


In this section, we compare the results of the simplified analysis reported above with
the features of the viscoelastic Poisson ratio for two linear polymers, 共polystyrene and
polycarbonate兲 and for a thermosetting polycyanurate having two different crosslink den-
sities. Poisson’s ratio is evaluated from experimental data for the bulk and shear re-
sponses using the relations between the viscoelastic functions in the Laplace domain. In
particular, for polycarbonate we will calculate Poisson’s ratio from the shear relaxation
modulus and the bulk compliance, whereas for polystyrene, it is obtained from the bulk
relaxation modulus and the shear compliance and for polycyanurate, it is calculated from
the bulk and the shear relaxation moduli.
The experimental data for polycarbonate, polystyrene, and polycyanurate, in terms of
bulk and shear responses, are taken from the recent papers by Grassia and D’Amore
共2009a, 2009b兲, by Meng and Simon 共2007兲, by Meng et al. 共2009兲, by Li and Simon
共2007兲, by Guo and Simon 共2009兲, and by Guo and Simon 共2010兲. The shear relaxation
data are fitted using a double Kohlraush–Williams–Watts 共KWW兲 function as
2
G共t兲 = G⬁ + 共G0 − G⬁兲 兺 wi exp −
i=1
冋冉 冊册t
␶i
␤i
, w1 + w2 = 1, 共9兲

whereas the bulk response is fitted using a simple KWW function as follows:
POISSON’S RATIO IN AMORPHOUS POLYMERS 1015

1.000
0.45 0.500
Polycarbonate
0.40 0.100

Bt  GPa1
Gt

Gt  GPa
0.050
Bt
0.35
0.010
0.30 0.005

0.25 0.001

104 0.1 100 105 108


ts

FIG. 5. Polycarbonate: Experimental data for the shear relaxation modulus 共symbols兲 and PVT-extracted bulk
compliance 共dashed line兲 at a reference temperature of T = Tg, as measured at 0.1 K/min, from the works by
Grassia and D’Amore 共2009a, 2009b兲. The solid line indicates the fit of Eq. 共12兲.

M共t兲 = M ⬁ + 共M 0 − M ⬁兲exp − 冋 冉 冊册 t


, 共10兲

where the symbol M indicates either the bulk modulus K or the bulk compliance B.
The shear relaxation modulus G and the bulk compliance B for polycarbonate are
shown in Fig. 5 at a reference temperature of 137 ° C that coincides with Tg, as measured
at 0.1 K/min. The master curve is taken from the papers by Grassia and D’Amore 共2009a,
2009b兲 and the details of the procedure utilized to obtain G and B are reported in the
same papers 共the master curve was obtained using the time-temperature superposition
principle; the experiments were performed in the temperature range 135– 165 ° C and in
the time window 0.1– 103 s兲. Here we only point out that the shear relaxation modulus
was measured experimentally, whereas the bulk compliance was obtained directly in form
of a simple KWW equation by fitting the pressure-volume-temperature 共PVT兲 data to the
Kovacs–Aklonis–Hutcheson–Ramos model of structural recovery 关Kovacs et al. 共1979兲;
Grassia and D’Amore 共2006兲兴. The parameters used in the fits of Eqs. 共9兲 and 共10兲 to the
polycarbonate data are reported in Tables I and II, respectively. From knowledge of G
and B, Poisson’s ratio for polycarbonate can be numerically derived using the relation
between these viscoelastic functions in the Laplace domain, i.e.,
3 − 2B共s兲G共s兲s2
␯共s兲 = , 共11兲
6s + 2B共s兲G共s兲s3
where ␯共s兲, B共s兲, and G共s兲 are, respectively, the Laplace transforms of ␯, B, and G.

TABLE I. Parameters of Eq. 共9兲 utilized to fit the shear relaxation modulus.

Polycyanurate Polycyanurate
Polycarbonate xM = 0.1 x M = 0.3

G0 共GPa兲 1.00 0.862 0.850


G⬁ 3.75⫻ 10−4 1.23⫻ 10−3 4.45⫻ 10−5
w2 0.936 0.957 0.927
␶1 共s兲 0.491⫻ 102 18.9 3.17⫻ 10−2
␶2 共s兲 3.06⫻ 102 43.3 35.7
␤1 0.167 0.180 0.112
␤2 0.438 0.359 0.353
1016 GRASSIA, D’AMORE, and SIMON

TABLE II. Parameters of Eq. 共10兲 utilized to fit the bulk response.

Polycyanurate Polycyanurate
Polycarbonate Polystyrene xM = 0.1 x M = 0.3

M0 0.208 GPa−1 3.07 GPa 5.05 GPa 5.63 GPa


M⬁ 0.484 GPa−1 1.76 GPa 2.07 GPa 2.38 GPa
␶ 共s兲 230 1.35⫻ 103 1.10⫻ 103 0.890⫻ 103
␤ 0.442 0.255 0.217 0.210

Concerning polystyrene, the shear compliance J and the bulk modulus K are reported
in Fig. 6, at a reference temperature of 93.8 ° C that coincides with Tg, as measured at 0.1
K/min. The shear creep compliance, J, is calculated by integrating the retardation spectra
reported in Fig. 11 of the paper by Meng and Simon 共2007兲, whereas the master curve
regarding the bulk modulus is directly taken from the paper by Meng et al. 共2009兲, where
the details of the procedure utilized to obtain K from pressure relaxation measurements
are reported. We mention that the bulk modulus was measured at two elevated pressures
共namely, at 42 and 76 MPa兲; the master curve was obtained using the time-temperature
superposition principle; the pressure relaxation experiments were performed in the tem-
perature ranges 85.5 ° C – 120.5 ° C and 89 ° C – 131 ° C at 42 MPa and 76 MPa, respec-
tively; the data were obtained in the time window 102.5 – 105 s. In order to obtain the bulk
modulus at ambient pressure, it is assumed that the magnitude of dispersion of the bulk
modulus 共namely, K0 − K⬁兲 does not change with pressure, according to the findings
reported by Meng et al. 共2009兲. Consequently, the bulk modulus at ambient pressure is
obtained by simply shifting downward the original data such that K⬁ is consistent with
that obtained from Tait equation 关Tait 共1888兲兴 at the conditions of interest 共P
= 0.1 MPa and T = 93.8 ° C兲 using the Tait parameters from the paper by Meng et al.
共2009兲. Furthermore, the parameters used to fit Eq. 共10兲 to the polystyrene bulk modulus
response are reported in Table II. From knowledge of J and K, Poisson’s ratio for poly-
styrene can be numerically derived by using the following relation in the Laplace do-
main:

3.0 105
Polystyrene
2.8
104
Jt  GPa1

2.6 Kt
Kt  GPa

Jt 1000
2.4
2.2 100
2.0 10
1.8
1
0.001 1 1000 106 109 1012 1015
ts

FIG. 6. Polystyrene: Experimental data for the shear creep compliance 共dashed line兲 and for the bulk relaxation
modulus 共symbols兲 at a reference temperature of T = Tg, as measured at 0.1 K/min, from the works by Meng and
Simon 共2007兲 and by Meng et al. 共2009兲, respectively. The solid line indicates the fit of Eq. 共11兲.
POISSON’S RATIO IN AMORPHOUS POLYMERS 1017

5.0 1.000
Polycyanurate 0.500
4.5
x M  0.1
4.0
0.100

Kt  GPa

Gt  GPa
Kt
3.5 Gt 0.050
3.0
0.010
2.5
0.005
2.0
1.5 0.001
0.01 1 100 104 106 108
ts

FIG. 7. Polycyanurate, x M = 0.1: Experimental data for the shear and bulk relaxation moduli 共symbols兲 at a
reference temperature of T = Tg, as measured at 0.1 K/min, from the works by Li and Simon 共2007兲, by Guo and
Simon 共2009兲, and by Guo and Simon 共2010兲. The solid lines display the fit of Eqs. 共11兲 and 共12兲.

3
− 1 + J共s兲K共s兲s2
2
␯共s兲 = , 共12兲
s + 3J共s兲K共s兲s3
where ␯共s兲, K共s兲, and J共s兲 are, respectively, the Laplace transforms of ␯, K, and J.
The shear relaxation modulus, G, and the bulk modulus, B, for two polycyanurates are
shown in Figs. 7 and 8 at the respective Tgs of the two materials measured at 0.1K/min.
The master curves, taken from the works by Li and Simon 共2007兲, by Guo and Simon
共2009兲, and by Guo and Simon 共2010兲, are reported for two different crosslink densities,
as indicated by x M = 0.1 and 0.3, where x M is the mole fraction of monofunctional mono-
mer used in the material preparation; the sample with x M = 0.1 has a higher cross-link
density and a corresponding higher Tg value. The details of the procedure, utilized to
obtain G and K, are reported in the same papers 关Li and Simon 共2007兲; Guo and Simon
共2009兲; Guo and Simon 共2010兲兴. We mention that while the shear response was measured
at ambient pressure 共temperature range Tg − 20 ° C − Tg + 20 ° C; time window:
0.1– 103 s兲, the bulk modulus was measured at higher pressure 共namely, at the average
pressure of 62 MPa; x M = 0.1, temperature range: 136– 180 ° C; x M = 0.3, temperature
range: 91– 140 ° C; time window: 102.5 – 105 s兲. The bulk response at ambient pressure is
obtained by the same procedure utilized for polystyrene, i.e., by simply shifting down-

5.5 1
5.0 Polycyanurate
x M  0.3 0.1
4.5
Gt  GPa
Kt  GPa

Kt
4.0 Gt 0.01
3.5
0.001
3.0
2.5 104
2.0
0.01 1 100 104 106 108 1010
ts

FIG. 8. Polycyanurate, x M = 0.3: Experimental data for the shear and bulk relaxation moduli 共symbols兲 at a
reference temperature of T = Tg, as measured at 0.1 K/min from the works by Li and Simon 共2007兲, by Guo and
Simon 共2009兲, and by Guo and Simon 共2010兲. The solid lines display the fit of Eqs. 共11兲 and 共12兲.
1018 GRASSIA, D’AMORE, and SIMON

0.50 PC 0.405
PS 0.404
0.48 0.403
PCy xM 0.1
PCy xM 0.3 0.402

Poisson's Ratio
0.46 0.401 PC
0.400
0.44 0.399
103 101 10
0.42
0.40 0.358
0.356 PS
0.38
0.354
0.36 0.352

107 104 0.1 100 105 108 107 104 101

t s

FIG. 9. Poisson’s ratio for polycarbonate 共PC: dot-dashed line兲, polystyrene 共PS: dashed line兲, and polycyanu-
rates 共PCy xM = 0.1: dot dot-dashed line and PCy x M = 0.3: solid line兲. All curves are for a reference temperature
of T = Tg, as measured at 0.1 K/min. The insets highlight the minima at short times for the two linear polymers
共polycarbonate and polystyrene兲. Here Poisson’s ratio, calculated by shifting the bulk response, is displayed as
thin lines.

ward the bulk response at elevated pressure such that K⬁ is consistent with the Tait
equation, the parameters of which are reported in the work by Guo and Simon 共2010兲.
The parameters used to fit Eqs. 共9兲 and 共10兲 to the shear and bulk responses for the
polycyanurates are reported in Tables I and II. From the knowledge of G and K, Poisson’s
ratio for the polycynurates can be numerically evaluated using the following relation in
the Laplace domain:
1 3K共s兲 − 2G共s兲
␯共s兲 = . 共13兲
s 6K共s兲 + 2G共s兲
Transforming Eqs. 共11兲–共13兲 back into the time domain, Poisson’s ratio for the four
materials of interest at their respective Tgs can be calculated. In particular, the Laplace
transform and the inverse Laplace transform have been calculated numerically using the
software MATHEMATICA®. In particular, the Laplace transform of the KWW function has
been obtained by calculating the cubic spline that interpolate the KWW function and then
performing the Laplace transformations. This ensures more accurate results than replac-
ing the KWW with Prony’s series.
The results of these calculations are reported in Fig. 9, where it can be observed that
␯共t兲 increases from a short-time value to a liquidlike value near 0.50. The time scale over
which ␯共t兲 evolves appears to be broader for the linear polymers than for the cross-linked
materials presumably because the latter have narrower relaxation time distributions. At
the shortest times, Poisson’s ratio value is approximately 0.36 for polystyrene, 0.40 for
polycarbonate, and 0.42 and 0.43 for the two polycyanurates. A similar result was ob-
tained in the work by O’Brien et al. 共2007兲, where the short-time limit of Poisson’s ratio
for an epoxy material near Tg was found to be approximately 0.40. Perhaps of more
interest in the context of the present work, for polycarbonate and polystyrene, Poisson’s
ratio is a nonmonotonic function of time, whereas for both of the polycyanurate materi-
als, Poisson’s ratio is a monotonically increasing function of time. In particular, for the
linear polymers 共polycarbonate and polystyrene兲 at short times, Poisson’s ratio shows a
weak minimum, as reported by Koppelmann 共1958兲 and Mott et al. 共2008兲.
A sensitivity analysis has been performed in order to further assess our result that
Poisson’s ratio is a nonmonotonic function of time for linear polymers. In particular, we
have done again the calculations of Poisson’s ratio for polystyrene and polycarbonate,
shifting the bulk function by half a decade on the time axis, both forward and backward.
A half-decade shift is considerably large than the experimental error in constructing the
POISSON’S RATIO IN AMORPHOUS POLYMERS 1019

1 1

0.01 0.01

Ht  GPa

Ht  GPa
104 104
Polycarbonate Polystyrene
106 106

108 108
104 0.1 100 105 108 105 0.1 1000 107 1011 1015

ts ts

0.1 0.1
Ht  GPa

Ht  GPa
0.001 0.001
Polycyanurate Polycyanurate
105 x M  0.1 105 x M  0.3

107 107
105 0.01 10 104 107 105 0.01 10 104 107
ts ts

FIG. 10. Bulk 共dashed line兲 and shear 共solid line兲 relaxation spectra a reference temperature of T = Tg, as
measured at 0.1 K/min.

bulk modulus master curves and hence this constitutes a severe and conservative test of
our results. From these calculations, it follows that Poisson’s ratio continues to show a
weak minimum at short times as displayed in the inset of Fig. 9, where Poisson’s ratio
calculated from the shifted bulk response is displayed as thin lines. In addition to the
effect of errors incurred by errors in the placement of the bulk modulus master curve on
the time axis, the effect of our assumption that the magnitude of the bulk dispersion is
independent of pressure on the results is examined. As was previously shown, the value
of ␶G / ␶K above, which a minimum in Poisson’s ratio is observed, increases with decreas-
ing K0 / K⬁. Our assumption, if anything, underestimates the value of K0 / K⬁ since 共K0
− K⬁兲 at ambient pressure cannot be less than at higher pressure. Thus, our results based
on this assumption are conservative with respect to the presence of the minimum.
The nonmonotonic behavior of Poisson’s ratio for polycarbonate and polystyrene can
be explained taking into account the relative position of the shear and bulk relaxation
moduli on the time axis. From Sec. II, it was shown that Poisson’s ratio shows a mini-
mum at short time when ␶G / ␶K is large. In order to determine if this is indeed the origin
of the minima observed for the linear polymers, the relaxation time spectra, H共t兲, for
polycarbonate, polystyrene, and the polycyanurates are reported in Fig. 10. The spectra
are evaluated with the second method proposed by Schwarzl and Staverman 共1953兲 关i.e.,
H共␶兲 = −共dG共t兲 / d ln t兲 + 共d2G共t兲 / d共ln t兲2兲 兩t=2␶兴. To minimize the calculation error during
differentiation, we use the KWW fits rather than the raw data to calculate the first and
second derivatives required. From Fig. 10, it is clear that for the linear polymers 共i.e.,
polycarbonate and polystyrene兲, the shear response is much broader than the bulk re-
sponse such that ␶G / ␶K Ⰷ 1, whereas for the cross-linked polycyanurates, the shear and
the bulk response share similar relaxation timescales such that ␶G / ␶K is on the order of
unity; hence, the results of the calculation reported in Figs. 9 and 10 are consistent with
the prediction obtained from the simple analysis performed in Sec. II.
There is quite a bit of evidence now that for linear polymers the viscoelastic shear
response occurs at longer time scales relative to the bulk response in the vicinity of the
␣-relaxation 关Ferry 共1980兲; Grassia and D’Amore 共2009a, 2009b兲; Meng and Simon
1020 GRASSIA, D’AMORE, and SIMON

共2007兲; Bero and Plazek 共1991兲兴, not necessarily because the two arise from different
molecular mechanisms but because the shear response is broader since it includes long-
time chain modes that are not available to the bulk response. This is an important point
because as cross-linking occurs, for example, in an epoxy or polycyanurate material, one
begins to lose the long-time chain modes and the relaxation spectra narrows. Therefore,
in a thermosetting polymer, especially if one assumes that the bulk and shear relaxation
spectra are similar at short times, the two relaxation spectra will become more alike as
cross-link density increases. Furthermore, as the relaxation spectra become more similar
with cross-linking and long-time chain mechanisms are lost in the shear response, ␶G / ␶K
will decrease. For K0 / K⬁ ⬇ 1.5– 2.0, ␶G / ␶K must be less than 3 or 2, respectively, for a
monotonic response as pointed out in the simplified model reported in Sec. II 共see also
Fig. 4兲. This is possible if the material has high enough cross-link density.
Our results that Poisson’s ratio is nonmonotonic function of time for linear polymers
and a monotonically increasing function for cross-linked materials is consistent with the
previous work of O’Brien et al. 共2007兲 on epoxy resins. It is also consistent with the
analysis of Kono’s data 关Kono 共1960兲兴 by Mott et al. 共2008兲 for a polystyrene and a
poly共methyl methacrylate兲. Both the works by O’Brien et al. 共2007兲 and by Kono 共1960兲
were performed in the vicinity of the ␣-relaxation process, where our analyses, both
analytical and experimental, are pertinent. Our approach is also not inconsistent with the
work in the region of the ␤-transition on poly共methyl methyacrylate兲 关Koppelmann
共1958兲兴. Koppelmann’s data were obtained at room temperature, i.e., in the vicinity of the
␤-relaxation for poly共methyl methacrylate兲, and show that the Poisson’s ratio undergoes
a minimum, while its relaxed and unrelaxed values are similar 共i.e., about 0.3兲. In the
framework of our modeling approach, these two features indicate that in the region of
␤-relaxation G0 / G⬁ is on the order of K0 / K⬁ and that ␶G / ␶K ⬎ 1, respectively 共inciden-
tally Koppelamann’s data also indicate that G0 / G⬁ in the ␤-relaxation region is about
1.5–2.0兲. On the other hand, the results of Tsou et al. 共1995兲 on cellulose acetate do not
agree with our predictions, since their results showed a maximum in the time-dependent
Poisson ratio, which appears to be unreasonable according to our analysis. Furthermore,
the results of Tsou et al. 共1995兲 on polycarbonate, which showed a flat 共time-
independent兲 Poisson ratio is not inconsistent with our results given that the measure-
ments were made at room temperature deep in the glassy state, although there is a
␤-relaxation in polycarbonate at this temperature, albeit considerably weaker than the
␤-relaxation in poly共methyl methacrylate兲 关Flory and McKenna 共2005兲兴.
A final point of discussion concerns our prediction that the time-dependent Poisson
ratio of thermosetting polymers in the vicinity of the ␣-relaxation should be a monotoni-
cally increasing function because the time scales and distribution of the relaxation times
for the shear and bulk responses become more alike as long-time chain modes 共which
seem to be operative in the shear response but not in the bulk response兲 关Grassia and
D’Amore 共2009a, 2009b兲; Meng and Simon 共2007兲; Guo and Simon 共2009兲; Bero and
Plazek 共1991兲兴 are progressively lost due to cross-linking. For low molecular weight
glass-forming materials, chain modes are also not present and hence, one might expect
that the shear and bulk relaxation responses should have similar time scales and relax-
ation time distributions. Hence, for the case of low molecular weight glass-forming
molecules, as well as for the case of inorganic network-forming glasses, we also predict
that the time-dependent Poisson ratio should be a monotonically increasing function in
the vicinity of the ␣-relaxation.
POISSON’S RATIO IN AMORPHOUS POLYMERS 1021

IV. CONCLUSIONS
The strength and the relative placement of viscoelastic functions depend on the par-
ticular material, so that no conclusions regarding ␯共t兲 can be drawn a priori. In linear
amorphous polymers, the shear response in the vicinity of the ␣-relaxation is broader and
contains long-time mechanisms that are unavailable to the bulk response such that
␶G / ␶K Ⰷ 1, and, according to our analysis, nonmonotonic Poisson’s ratio behavior may be
expected for this class of materials. This is shown to be the case for a polycarbonate and
a polystyrene material using data from the literature. However, in the case of tightly
cross-linked polymeric networks, the long-time chain modes are not present and hence,
the shear and bulk response share similar timescales, such that ␶G / ␶K is on the order of
unity. Consequently, Poisson’s ratio for cross-linked polymeric networks is expected to be
a monotonically increasing function of time, and we showed this to be the case for two
polycyanurate materials. Our results are consistent with the time-dependent Poisson ratio
reported in the literature for epoxies 关O’Brien et al. 共2007兲兴, as well as for polystyrene
and poly共methyl methacrylate兲 关Mott et al. 共2008兲兴, both in the vicinity of the
␣-relaxation, and they are not inconsistent with work in the region of the ␤-transition on
poly共methyl methyacrylate兲 关Koppelmann 共1958兲兴 and polycarbonate 关Tsou et al.
共1995兲兴. However, previous findings of a maximum in the Poisson’s ratio for cellulose
acetate 关Tsou et al. 共1995兲兴 appear to be inconsistent with our analysis. Based on our
results, we further suggest that low molecular weight glass-forming materials and inor-
ganic network glasses will also display a Poisson ratio that is a monotonically increasing
function of time.

ACKNOWLEDGMENTS
S.L.S. gratefully acknowledges funding from the National Science Foundation Divi-
sion of Materials Research Grant No. DMR-0606500.

References
Bero, C. A., and D. J. Plazek, “Volume-dependent rate processes in an epoxy resin,” J. Polym. Sci., Part B:
Polym. Phys. 29, 39–47 共1991兲.
Bogy, D. B., N. Bugdayci, and F. E. Talke, “Experimental determination of creep functions for thin orthotropic
polymer films,” IBM J. Res. Dev. 23, 450–458 共1979兲.
Christensen, R. M., Theory of Viscoelasticity: An Introduction, 2nd ed. 共Academic, New York, 1982兲.
Delin, M., R. Rychwalski, and J. Kubat, “Volume changes during stress relaxation in polyethylene,” Rheol. Acta
34, 182–195 共1995兲.
Ferry, J. D., Viscoelastic Properties of Polymers, 3rd ed. 共Wiley, New York, 1980兲.
Flory, A. L., and G. B. McKenna, “Chemical structure—Normal force relationships in polymer glasses,”
Polymer 46, 5211–5217 共2005兲.
Giovagnoni, M., “On the direct measurements of the dynamic Poisson’s ratio,” Mech. Mater. 17, 33–46 共1994兲.
Grassia, L., and A. D’Amore, “Constitutive law describing the phenomenology of sub-yield mechanically
stimulated glasses,” Phys. Rev. E 74, 021504 共2006兲.
Grassia, L., and A. D’Amore, “The relative placement of linear viscoelastic functions in amorphous glassy
polymers,” J. Rheol. 53, 339–356 共2009a兲.
Grassia, L., and A. D’Amore, “On the interplay between viscoelasticity and structural relaxation in glassy
amorphous polymers,” J. Polym. Sci., Part B: Polym. Phys. 47, 724–739 共2009b兲.
Guo, J., and S. L. Simon, “Effect of crosslink density on the pressure relaxation response of polycyanurate
networks,” J. Polym. Sci., Part B: Polym. Phys. 47, 2477–2486 共2009兲.
1022 GRASSIA, D’AMORE, and SIMON

Guo, J., and S. L. Simon, J. Polym. Sci., Part B: Polym. Phys. 共to be published兲.
Kono, R. J., “The dynamic bulk viscosity of polystyrene and polymethyl methacrylate,” J. Phys. Soc. Jpn. 15,
718–725 共1960兲.
Koppelmann, J., “Über die Bestimmung des dynamischen Elastizitätsmoduls und des dynamischen Schub-
moduls im Frequenzbereich von 10−5 bis 10−1 Hz,” Rheol. Acta 1, 20–28 共1958兲.
Kovacs, A. J., J. J. Aklonis, J. M. Hutchinson, and A. R. Ramos, “Isobaric volume and enthalpy recovery of
Glasses. II. A transparent multiparameter theory,” J. Polym. Sci., Polym. Phys. Ed. 17, 1097 共1979兲.
Lakes, R. S., “The time dependent Poisson’s ratio of viscoelastic materials can increase or decrease,” Cell.
Polym. 11, 466–469 共1992兲.
Lakes, R. S., and A. Wineman, “On Poisson’s ratio in linearly viscoelastic solids,” J. Elast. 85, 45–63 共2006兲.
Li, Q. X., and S. L. Simon, “Viscoelastic shear response and network structure in polycyanurates,” Macromol-
ecules 40, 2246–2256 共2007兲.
Meng, Y., and S. L. Simon, “Pressure relaxation of polystyrene and its comparison to the shear response,” J.
Polym. Sci., Part B: Polym. Phys. 45, 3375–3385 共2007兲.
Meng, Y., P. Bernazzani, P. A. O’Connell, G. B. McKenna, and S. L. Simon, “A new pressurizable dilatometer
for measuring the time-dependent bulk modulus and pressure-volume-temperature properties of polymeric
materials,” Rev. Sci. Instrum. 80, 053903 共2009兲.
Mott, P. H., J. R. Dorgan, and C. M. Roland, “The bulk modulus and Poisson’s ratio of “incompressible”
materials,” J. Sound Vib. 312, 572–575 共2008兲.
O’Brien, D. J., N. R. Sottos, and S. R. White, “Cure-dependent viscoelastic Poisson’s ratio of epoxy,” Exp.
Mech. 47, 237 共2007兲.
Schwarzl, F., and A. J. Staverman, “Higher approximation methods for the relaxation spectrum from static and
dynamic measurements of viscoelastic materials,” Appl. Sci. Res., Sect. A 4, 127 共1953兲.
Tait, P. G., The Voyage of H. M. S. Challenger 共Her Majesty’s Stationary Office, London, 1888兲, Vol. 2.
Tschoegl, N. W., W. G. Knauss, and I. Emri, “Poisson’s ratio in linear viscoelasticity—A critical review,” Mech.
Time-Depend. Mater. 6, 3–51 共2002兲.
Tsou, A. H., J. Greener, and G. D. Smith, “Stress relaxation of polymer films in bending,” Polymer 36,
949–954 共1995兲.
Yoe, A. F., and M. T. Takemori, “Dynamic bulk and shear relaxation in glassy polymers. I. Experimental and
techniques on PMMA,” J. Appl. Polym. Sci. 20, 205–224 共1982兲.

View publication stats

You might also like