You are on page 1of 5

Mechanical properties of solid polymers

Constitutive modelling of long and short term behaviour

CIP-DATA LIBRARY TECHNISCHE UNIVERSITEIT EINDHOVEN


Klompen, Edwin T.J.
Mechanical properties of solid polymers : constitutive modelling of long
and short term behaviour / by Edwin T.J. Klompen. - Eindhoven :
Technische Universiteit Eindhoven, 2005.
Proefschrift. - ISBN 90-386-2806-4
NUR 971
Trefwoorden: glasachtige polymeren / constitutieve modellering /
post-yield deformation / strain softening / strain hardening /
moleculaire transities / fysische veroudering / lange duur falen
Subject headings: polymer glasses / constitutive modelling /
post-yield deformation / strain softening / strain hardening /
molecular transitions / physical ageing / long-term failure

Reproduction: University Press Facilities, Eindhoven, The Netherlands.


Cover design: Jan-Willem Luiten (JWL-Producties).
Cover illustration: surface representing the intrinsic response of a glassy
polymer at different strain rates; the red line is the response to a constant
rate of deformation, while the blue line is the response to a constant stress.

1.3 M OLECULAR

BACKGROUND

thermorheological complex, and the magnitude of the effect is mainly determined by


the relative position of the transitions on the time scale. Due to the shape change,
experimental data of a viscoelastic function at different temperatures are generally
no longer superimposable, like in Figure 1.4(a), by pure horizontal shifting and,
therefore, a direct application of the principle of time-temperature superposition is
no longer possible.
It was already observed by Schapery [16] and Nakayasu et al. [27], that semicrystalline materials also show signs of thermorheological complex behaviour. In
these materials the crystalline phase gives rise to a relaxation mechanism at temperatures above the glass transition temperature (in contrast to the sub-Tg secondary
transitions). The strength and magnitude of these transitions depends on the degree of crystallinity and the size of the lamellae, and therefore is influenced by the
thermal history (i.e. processing). Due to the composite nature of semi-crystalline
materials, changes in the crystalline phase lead to changes in the amorphous mechanism as well. Any secondary relaxations are generally not affected by the presence
of a crystalline phase. A detailed survey regarding relaxation processes in crystalline
polymers is provided by Boyd [28, 29].

Stress activated mobility: nonlinear flow


It was shown that the characteristic, time-dependent, material behaviour, generally
observed for polymer materials, is caused by molecular transitions, and that these
are activated by temperature. Application of stress has a similar effect as temperature, increasing mobility with increasing load. In contrast to temperature, however,
stress preferentially promotes mobility in the direction of the load applied. This is the
basis of the Eyring flow theory, which is schematically illustrated in Figure 1.5(a). An
initially symmetric potential energy barrier with magnitude U is biased by a load
. The magnitude of this bias is determined by the parameter V , the (segmental)
activation volume. This volume is characteristic for a particular flow process associated with a relaxation mechanism. It has been shown that this theory adequately
describes both strain rate and temperature dependence of polymers [30].
As an example, the tensile yield stress of polycarbonate at different temperatures as
function of strain rate, shown in Figure 1.5(b), is known to be governed by a single
relaxation mechanism in the range of temperatures and strain rates of interest. A
linear dependence of yield stress as a function of log strain rate can be observed over
a large interval of strain rates and temperatures, which are described well using
the Eyring theory (solid lines). Other flow theories, such as those of Argon [19], or
Robertson [31], can be expected to describe the data equally well since, on a fitting
level, they do not differ from the Eyring theory.

1 I NTRODUCTION

/T 102 (Kg.mm2.K1)

2.5

21.5 C
40 C

60 C
80 C

1.5

100 C
120 C

140 C

Potential energy

0.5
6

Log

Direction of applied stress

(b)

(a)

Figure 1.5: (a) Schematic representation of the Eyring flow theory. (b) Tensile yield stress
versus strain rate at different temperatures for polycarbonate, symbols indicate
experimental data and solid lines are fits using the Eyring-theory. Adapted from
Bauwens-Crowet et al. [30].

(/T) 10

(dynes cm

deg )

It is known that, in the linear viscoelastic range, the time dependent behaviour can
include contributions of secondary relaxation processes. It was shown by several
authors that this also holds for the tensile yield stress of these materials [30, 32].

2 C
30
40
50

60
70
80
90

0
6

Log

Figure 1.6: Tensile yield stress versus strain rate at different temperatures for polymethylmethacrylate, symbols are experimental data and solid lines are fits using the
Ree-Eyring theory. Adapted from Roetling [32].

As an example, Figure 1.6 shows the yield stress of PMMA as function of strain
rate for a large range of temperatures. With increasing strain rate the yield stress
shows a change in slope for all temperatures, the transition shifting to higher rates
with increasing temperature. The additional contribution at the higher strain rates
(i.e. shorter times, or a reduced time scale) is generally attributed to the secondary
relaxation process.

1.3 M OLECULAR

BACKGROUND

To describe this yield behaviour, a Ree-Eyring modification of the Eyring flow theory
is used. The modification consists of placing two Eyring flow processes in parallel.
This approach was also shown to be applicable for semi-crystalline materials, such
as e.g. isotactic polypropylene [33,34], while for some other materials an extra, third,
process was necessary [35, 36].

Influence of history: physical ageing and mechanical rejuvenation

Volume

log(Compliance)

Temperature and stress both increase the molecular mobility and, consequently, accelerate the time scale at which the material deforms. From this it might be concluded
that when both are kept constant, the mobility and its resulting rate of deformation
do not change. It was, however, shown by Struik that for many polymers this is not
correct [37]. With increasing time elapsed after a (thermal) quench from above Tg , he
observed a decrease in molecular mobility: the response to linear creep experiments
at constant temperature shifts to longer loading times, see Figure 1.7(a).

Equilibrium
melt

Nonequilibrium
glass

Ageing
Equilibrium
glass
Tg

log(Time)

(a)

Temperature

(b)

Figure 1.7: (a) Schematic representation of the influence of ageing time te on the linear creep
compliance. (b) Schematic illustration of physical ageing: volume as a function of
temperature.

This effect is generally referred to as physical ageing and can be explained in


terms of the presence of a non-equilibrium thermodynamic state. When cooling a
polymer melt that is in equilibrium, the mobility of the molecules decreases with
decreasing temperature, thus increasing the time required to attain equilibrium.
At a certain temperature, which depends on the cooling rate, the time required
exceeds the experimental time available and the melt becomes a polymer glass.
The temperature at which this occurs depends on the cooling rate and is termed
the glass-transition temperature Tg . The material is now in a non-equilibrium state,

You might also like