You are on page 1of 4

Materials Letters 63 (2009) 877880

Contents lists available at ScienceDirect

Materials Letters
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / m a t l e t

A hyper-viscoelastic constitutive model for polyurea


Chunyu Li , Jim Lua
Global Engineering and Materials Inc, Princeton, NJ 08540, USA

a r t i c l e

i n f o

Article history:
Received 18 November 2008
Accepted 13 January 2009
Available online 26 January 2009
Keywords:
Hyper-viscoelasticity
Constitutive modeling
Polyurea
Rate-dependent material
High strain rate

a b s t r a c t
This letter presents a new constitutive model for polyurea by superposing the hyperelastic and viscoelastic
behaviors of polyurea. The Ogden model is used for the hyperelastic part and its parameters are determined
from curve tting of quasi-static test data. A nonlinear viscoelastic model is employed for describing the
viscoelastic behavior and its relaxation time is obtained based on the test data of shear relaxation modulus. A
special form of Zapas kernel for the damping function is found to be very effective to capture the viscoelastic
behavior of polyurea subjected to wide ranges of strain rate. Both the versatility and accuracy of the model
are examined via virtual testing.
2009 Elsevier B.V. All rights reserved.

1. Introduction
Polyurea is a product from the chemical reaction between an
isocyanate and an amine. It has been widely used in the coating
industry, because of its extensive benets over existing epoxy
adhesives and rubber linings in terms of impact, abrasion and
corrosion resistance. With the development of polyurea spray coatings technology, specically formulated polyurea can be directly and
efciently sprayed on the surface of structural components to enhance
the mechanical strength and durability of civil and military structures.
Previous experimental studies [1] have revealed that polyurea
exhibits elastic and nearly incompressible behavior to volumetric
deformations and its stressstrain behavior depends on strain rate,
temperature and pressure [2,3]. Different from the rubbery behavior
under low strain rates, polyurea displays a distinct leathery behavior [4]
at high strain rates. Polyurea can be used in a wide range of temperature
(from 50 C to 150 C) and has shown a high heat resistance. However,
its shear modulus decreases signicantly with increasing temperature.
The glass transition temperature Tg for polyurea is roughly 50 C [1].
Given the stiffening behavior of polyurea material with both increasing
strain and strain rate, it has been used either as a protection coating on a
metallic structure or an inserted layer between the outer facesheet and
the foam core in a blast-tolerant sandwich structure.
Both material certication and performance evaluation of polyurea
coated structural components under hostile environment by tests will
be prohibitively expensive and time consuming. Development of a high
delity constitutive model of polyurea is imperative to perform an
optimal design of its coated components subjected to a combined
dynamic and thermal loading. Currently, there are only two models have
Corresponding author.
E-mail address: chunyulee@hotmail.com (C. Li).
0167-577X/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.matlet.2009.01.055

been developed specically for polyurea although some similar models


have been proposed for other hyperelastic or viscoelastic materials [5,6].
One is an experimentally-based linear viscoelastic constitutive model
proposed by Amirkhizi et al. [2]. It incorporates the classical Williams
LandelFerry (WLF) time-temperature transformation and pressure
sensitivity, in addition to a thermodynamic energy dissipation mechanism. The model can reproduce experimental results for conned
polyurea tests but has limited capability in simulating the unconned
test data. The other is a more complex model proposed by Elsayed [7]
which consists of an elastoplastic network acting in parallel with several
viscoelastic networks. Quasi-incompressible Ogden-type potentials
were used and the number of model parameters is a function of the
number of active Ogden terms and relaxation mechanisms, thus a
procedure based on genetic algorithms was employed to calibrate model
parameters based on existing experimental data. Although this model is
relatively successful in reproducing experimental data, it seems to be too
complicated for engineering practice.
A constitutive model for polyurea should cover a wide range of strain
rates from static to high speed impact and capture both hyperelastic and
viscoelastic material behavior. A rational approach based on a combination
of hyperelasticity theory and viscoelasticity theory is needed in deriving
the constitutive model. Also, for its practical applications, the model
should be comparatively simple to be implemented in a general purpose
nite element code while its parameters can be determined experimentally from the standard tests. To meet the model requirements, this letter
presents a new hyper-viscoelastic constitutive model for polyurea.
2. Hyperelastic model for low strain rates
Given the compliant rubber-like behavior of polyurea from quasistatic tests, a hyperelastic constitutive model is selected to characterize
its rubbery stressstrain behavior at low strain rates. The constitutive

878

C. Li, J. Lua / Materials Letters 63 (2009) 877880

listed in Table 1. Fig. 1 shows that the curve-t results are in good
agreement with the experimental data.

Table 1
Fitting parameters

i(MPa)
i
Gi(MPa)
i(ms)

40.9183
7.6257
15.8879
301.61

16.6532
3.9473
28.3111
0.13146

34.7554
3.4994
38.1695
1.0556 10 3

27.7561
3.9430
44.0718
66.9582
86.1898
4.3933 10 5 2.4659 10 6 1.6715 10 7

law for an isotropic hyperelastic material is decomposed into volumetric


and deviatoric parts. The deviatoric part is dened by an equation
relating the strain energy density of the material to the three invariants
of the strain tensor. By assuming the incompressible behavior of the
rubber-like material, the strain energy density W becomes a function of
the rst two invariants, i.e., W =W(I1, I2), where I1 =tr(B), I2 = [I21 tr(B2)]/
2 and B is the left Cauchy-Green deformation tensor. Here, we only
consider the uniaxial loading case given the availability of experimental
data. For a specimen under uniaxial loading, the principal stretches can
be expressed as 1 =u, 2 =3 =u 1/2, where u stands for the stretch in
the loading direction and u = 1 +u with u denotes the engineering
strain in the loading direction. The left CauchyGreen deformation
tensor B is then B = diag[2u, u 1, u 1] (the right CauchyGreen
deformation tensor C = B). There are several specic forms of the strain
energy density, such as by Arruda-Boyce [8], Mooney-Rivlin [9] and NeoHookean [10]. Here we select the Ogden model [11], given by
n

W=
i=1


2 i  i
1 + 2 i + 3 i 3 :
2i

i=1

2 i  i i =2 
;
u u
i

Having established the rate-independent hyperelastic model for


polyurea under static loading conditions, we are now focusing on the
rate-dependent characteristic of polyurea. A viscoelastic constitutive
law is a good choice for capturing its rate-dependent behavior. For
viscoelastic materials, the stress state depends on the strain and strain
rate histories. Here, we assume the material is isotropic and subjected
to small strains under an isothermal condition. But different from a
linear viscoelastic model used by Amirkhizi et al. [2] an assumption of
nonlinear viscoelasticity is made. The general constitutive relationship can be expressed as [12]:
t

v = pv + Ft R

e11

where
represents the engineering stress. The hydrostatic pressure
ph representing the volumetric part can be obtained by posing the
boundary condition associated with unconned tests h22 = h33 = 0, i.e.,

fCgFT t :

where v is the Cauchy stress tensor, pv is the pressure in the viscoelastic


material, a Kronecker tensor, R a matrix functional that describes the
effect of strain history on stress, and F represents the deformation
gradient (F = diag[u, u 1/2, u 1/2] for uniaxial loading). There have been
numerous approximations to represent the functional for viscoelastic
solids and uids [12,13]. Since the K-BKZ integral form [14] and its
variants are well-accepted, we adopt the form shown below
R

Thus, the Cauchy (true) stress h11 derived from the Ogden model is
h11 = ph + u e11 =

3. Nonlinear viscoelastic model for high strain rates

0 0
:
t
/ I1 ; I2 mt E d:
fC g =

where (I1',I2') is a damping function with I1',I2' as the rst two


invariants of strain tensor C, and the strain rate is dened by
: 1 : T
:
F F + FTF ;
E=
2

To determine the parameters i and i, the experimental data from


Sarva et al. [4] for polyurea under very low strain rate (0.0016 s 1) is
used for the nonlinear least square tting. The tted parameters are

and m(t) is a relaxation function.


The relaxation function is generally represented by a series of simple
exponentials, or so-called Prony series, with its individually dened
relaxation time. In the literature, different number of exponential terms
(e.g.1, 2 or 5), thus different relaxation times were selected driven by the
required delity of the curve-t model [1517]. While the use of
combination of relaxation times and other undetermined parameters
can provide a good t of the experimental data of stressstrain curves, it
has no physical foundation and rational for a specic combination. In
fact, the relaxation times depend on the microstructure of a material.
Motion within the microstructure determines the relaxation time of the
material and its activation depends on temperatures and corresponding
loading rates. For example, at higher temperatures, the molecular
relaxation processes are more easily and frequently activated due to the

Fig. 1. Fitting the Ogden hyperelastic model.

Fig. 2. Fitting of the master curve to nd relaxation times.

ph =
i=1

2 i  i =2 i =4 
:
u u
i

The nal form of the Cauchy stress in the loading direction is thus
written as
n

h11 =

i=1


2 i  i
i =2
u 2
+ u i =4 :
u
i

C. Li, J. Lua / Materials Letters 63 (2009) 877880

879

= and the pressure pv into


Substitution of the strain rate E11
u u
Eq. (12) yields the stress for the viscoelastic polyurea in the loading
direction:

v11 =

 0 i
:
1 1 t 2 h
u 0 u A1 + A2 I2 3 mt u d
2
h
 0 i
:
+ 2u 0t u A1 + A2 I2 3 mt u d:

14

Adding up Eqs. (4) and (14) together, we have the hyperviscoelastic stress for polyurea 11 = h11 + v11. In this total stressdeformation formulation, only two parameters A1 and A2 are
undetermined and they can be obtained by curve-tting of the
experimental data associated with a uniaxial high strain rate testing.
Fig. 3. Comparison of experimental data and model predictions.
4. Results and discussions

higher thermal energy. In this study, therefore, we propose to determine


the relaxation function based on the shear modulus relaxation rst
before computing other curve-tting parameters.
Knauss [1] conducted relaxation tests on polyurea at various
temperatures. The experimental data is shown in Fig. 2, in which the
data has been shifted according to the time-temperature shift
principle. The applicability of the shift process is assured by the
rather smooth master relaxation curve. The standard WLF shift
function has been given as [1]
aT = 10ATTref =B + TTref ;

with A = 8.86, B = 171 C and Tref = 0 C. The master curve can be


approximated by a Prony series with a number of terms:
n

Gref t = G + Gi et=i :

i=1

Fig. 2 shows our nonlinear regression to the master curve using a


6-term Prony series. The tting parameters are summarized in Table 1.
The value of the goodness of t R2, which is computed from the sum of
the squares of the distances of the points from the best-t curve
determined by the nonlinear regression and then normalized to the
sum of the square of the distances of the points from a horizontal line
through the mean of all Gref values, is 0.9998. This is much close to 1.0
and indicates that the model ts the data better.
Because G is not related to time, the relaxation function m(t ) in
Eq. (6) can then be assumed as
6

mt = Gi e t=i ;

In this study, the data for the unconned uniaxial compression test at strain rate of
2250 s 1 from Sarva et al. [4] is used to determine the parameters A1 and A2. The
parameters determined from the tting are A1 = 0.1435 and A2 = 0.1098. Fig. 3 displays
the comparison between the test data and our prediction from the present model. The
symbols triangle, square, pentagon, hexagon and circle represent, respectively, the test
data under the strain rate 0.0016/s, 80/s, 800/s, 2250/s and 6500/s. The color lines
represent predictions of our present model with gray, cyan, wine, blue and red stand for
the result under the strain rate 0.0016/s, 80/s, 800/s, 2250/s and 6500/s, respectively. It
can be seen that the prediction of the present model is in a very good agreement with
the experimental data for a large range of strain rates. In addition, the delity of the
present model is also evaluated via its comparison with prediction from other existing
models, as shown in Fig. 4. Clearly, higher accuracy has been achieved from the present
model in simulating the polyurea constitutive behavior than the model given by
Amirkhizi et al. [2] As shown in Fig. 4, our model captures the early response more
accurately than the model developed by Elsayed [7]. Given the simplicity of the present
model, it can be used as a strongly competitive alternative to Elsayed's model.

5. Conclusions
This letter presents a new constitutive model for polyurea by
including both its hyperelastic and viscoelastic behavior. The Ogden
model is used for the hyperelastic part and its parameters are curve
tted using the quasi-static test data. A nonlinear viscoelastic model is
proposed to characterize its viscoelastic part and the relaxation time is
determined based on the test data of the shear relaxation modulus. A
special form of Zapas kernel for the damping function is found to be
very effective in prediction of the viscoelastic behavior of polyurea
subjected to a wide range of strain rates. It is expected that this new
constitutive model can be implemented in general purpose nite
element code as a user-dened material model to assess and design of
a protective structure subjected to extreme dynamic loading.

10

i=1

for isothermal deformations. After some strenuous trials, the damping


function in Eq. (6) is taken as a special form of Zapas kernel [18],
/ = A1 =A1 + I3;

11

where I = A2I1' + (1 A2)I2' and A1 and A2 are undetermined parameters.


Substituting Eq. (11) into Eq. (6) and then Eq. (5) and neglecting the
effect of deformation history prior to loading on the stress, we obtain
n h
 0 i
o
:
v = pv + F  0t A1 + A2 I2 3 mt E d  FT :

12

Considering
:
: the free :boundary condition of the transverse stress

and E22 = 2 2 = 12 2
u u (the stretching rate u is equal to the
engineering strain rate 11), the pressure pv can be obtained as
h
 0 i
:
1
t 2 A1 + A2 I2 3 mt u d:
pv = 1
2 u 0 u

13

Fig. 4. Comparison of the present model with existing models.

880

C. Li, J. Lua / Materials Letters 63 (2009) 877880

Acknowledgement
This work is supported by the Ofce of Naval Research under
Contract N00014-08-C-0614 with Dr. Roshdy Barsoum as the Program
Manager.
References
[1] W.G. Knauss, Final report to the Ofce of Naval Research. (ONR Grant No. N0001403-1-0539), CIT, Pasadena, CA.
[2] Amirkhizi AV, Isaacs J, Mcgee J, Nemat-Nasser S. Philos Mag 2006;86:584766.
[3] Yi J, Boyce MC, Lee GF, Balizer E. Polymer 2006;47:31929.
[4] Sarva SS, Deschanel S, Boyce MC, Chen W. Polymer 2007;48:220813.
[5] Qi HJ, Boyce MC. Mech Mater 2005;37:81739.

[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]

Reese S, Govindjee S. Int J Solids Struct 1998;35:345582.


T.M. Elsayed, PhD dissertation, California Institute of Technology (2008).
Arruda E, Boyce MC. J Mech Phys Solids 1992;41:389.
Mooney M. J Appl Phys 1940;11(582).
Treloar LRG. Proc Phys Soc 1948;60:135.
Ogden RW. Proc R Soc Lond 1972;A326:56584.
Lockett FJ. Nonlinear viscoelastic solids. New York: Academic Press; 1972.
Carreau PJ, De Kee DCR, Chhabra RP. Rheology of polymeric systems: principles and
applications. New York: Hanser Publishers; 1997.
Tanner RI. J Rheology 1988;32:673.
Osaki K, Ohta S, Fukuda M, Kulata M. J Polym Sci 1976;14(1701).
Wagner MH. Rheol Acta 1979;18(33).
Yang LM, Shim VPW, Lim CT. Int J Impact Eng 2000;24:545.
Zapas LJ. J Res Natl Bur Stand 1966;70A(525).

You might also like