Daletal 20 JMPS
Daletal 20 JMPS
Daletal 20 JMPS
net/publication/344286291
CITATIONS READS
45 1,925
3 authors:
Kemal Acikgoz
Middle East Technical University
8 PUBLICATIONS 166 CITATIONS
SEE PROFILE
All content following this page was uploaded by Hüsnü Dal on 17 September 2020.
Paper dedicated to Professor Michael Kaliske on the occasion of his 60th birthday
Abstract
Rubberlike materials exhibit strong rate-dependent mechanical response which manifests itself in creep and relaxation tests as well as in the
hysteresis curves under cyclic loading. Unlike linear viscoelasticity, creep and relaxation response of rubber is nonlinear and amplitude-dependent.
Within this context, the contribution of this work is three fold. (i) On the experimental side, the characterization of equilibrium and non-equilibrium
responses are carried out by means of uniaxial and equibiaxial extension tests. Also performed are the creep and relaxation experiments at various
stress and strain levels. (ii) On the theoretical side, we extend the well-known eight-chain model via incorporating a simple yet instrumental tube-
constraint term composed of the second invariant into the non-affine network contribution reflecting the ground state equilibrium response. For
the non-equilibrium response, we propose a new evolution equation for the creep/relaxation behavior of rubberlike materials based on a relaxation
kinetics of a single polymer chain. The geometric non-linearity is incorporated via the finite deformation kinematics based on the multiplicative
split of the deformation gradient into elastic and viscous parts, whereas the volumetric and isochoric split of the deformation gradient is entirely
discarded. The rheology uses a nonlinear spring responsible for equilibrium elastice response in parallel to n number of Maxwell elements, leading
to a generalized Maxwell-Wiechert viscoelastic solid. (iii) The algorithmic implementation of the model features the spectral decomposition of
the respective terms and is demonstrated within the context of the finite element method. The developed model is validated by fitting both the
elastic and viscoelastic model responses with respect to the experimental data in the sense of uniaxial, (equi)biaxial extensions, and pure shear
tests. Relaxation and creep behavior of the model are thoroughly assessed. Also presented in the manuscript is the capability and the performance
of the model in the face of a relevant non-homogeneous boundary value problem. The quality of the findings earns the model vast utilization areas
from an engineering perspective.
Keywords: Finite viscoelasticity · rubber viscoelasticity · hysteresis · creep · relaxation · thermodynamical consistency · algorithmic setting ·
finite element modeling · experimental analysis
1. Introduction
Rubberlike materials exhibit strong rate-dependent response that earns rubber excellent mechanical properties
highly sought in applications suc as automotive, aerospace and soft robotics, just to name but a few. Viscoelasticity
improves the fracture toughness and energy absorption capacity rendering rubber very suitable as and shock absorbers
and vibration decouplers. The backbone of finite viscoelastictiy of rubberlike materials is an accurate description of
ground state equilibrium elastic response. In this context, the hyperelastic constitutive models for rubber can either be
represented by (i) phenomenological, or (ii) micromechanically based free energy functions. The phenomenological
models are based on invariant or principal stretch-based expressions for the free energy function that successfully
fit the experimental stress-strain curves. The invariant based models usually consist of polynomial combinations
[80, 62, 7, 89] or more sophisticated mathematical functionals [28, 76, 42, 27, 90, 30, 49, 66, 55, 40] of the first and
second invariants of the right/left Cauchy-Green deformation tensor. Various functional forms of free energy have been
proposed in terms of principal stretches e.g. [85, 4, 63, 71, 2, 5, 44]. Some material models have a mixed mathematical
representation in terms of principal stretches and invariants e.g. [32, 37]. Micromechanical models, on the other
hand, are physically-based models, condisdering the material microstructure by linking the macroscopic mechanical
∗ Corresponding author
Email address: [email protected] (Hüsnü Dal)
F : TX B0 → Tx B; F := ∇X ϕt (X) (1)
maps the unit tangent of the reference/Lagrangian configuration onto its counterpart in the current/Eulerian config-
uration. Subsequently, let TX∗ B0 and Tx∗ B denote the co-tangent spaces in the Lagrangian and Eulerian manifolds.
Weighing cof[F ] by the volume map det[F ], one can describe a normal map per unit reference volume
which maps the unit normal vector N of the reference/Lagrangian configuration onto its counterpart n in the cur-
rent/Eulerian configuration. The gradient operators ∇X [•] and ∇x [•] indicate the spatial derivative with respect to the
reference X and current x coordinates, respectively. In the sequel, let dX, dA and dV represent the infinitesimal
line, area and volume elements, in the undeformed configuration. Then, the deformation gradient F , its cofactor
cof[F ] = det[F ]F −T and the Jacobian J := det[F ] > 0 characterize the deformation of infinitesimal line, area and
volume elements
dV
dX dA
X X X
(a) (b) (c)
Figure 1: Three fundamental maps of a continuum: (a) The deformation gradient F as a mapping of an infinitesimal line element, (b) its cofactor
cof[F ] as an area map, and (c) and its determinant det[F ] as a volume map.
C = F T gF and c = F −T GF −1 (4)
as the pull back of the current metric g and push-forward of the Lagrangian metric G, respectively. The left Cauchy
Green tensor or the Finger tensor is denoted by b = c−1 . For a geometric interpretation, we refer to Fig. ??(a-b).
In order to comply with the principle of material objectivity and principle of material frame indifference, the energy
stored in an isotropic material is governed by principle stretches or invariants. Within this context, the eigenvalue
decomposition of the deformation metrics U and C are, respectively,
3
X 3
X 3
X
U := λa N a ⊗ N a , C := λ2a N a ⊗ N a and cof[C] := νa2 N a ⊗ N a , (5)
a=1 a=1 a=1
4
where
νi = J/λi such that ν1 = λ2 λ3 , ν2 = λ3 λ1 , ν3 = λ1 λ2 , (6)
are the areal stretches. Three isotropic invariants of the right Cauchy Green tensor read
I1 := tr[C], I2 := tr[cof[C]], and I3 := det[C] . (7)
The relation between the principal stretches and invariants can be established through
I1 = λ21 + λ22 + λ23 I2 = ν12 + ν22 + ν32 I3 = λ21 λ22 λ23 . (8)
In (7), we have introduced an alternative definition to the second invariant by postulating the equivalence of − tr[C 2 ] = 1
2 I12
tr[cof[C]]. Comparing (3) and (7), one can easily infer that the three isotropic invariants are associated with squares
of linear, areal and volumetric stretches of an infinitesimal cube representing a material point, see Fig. 2. To distin-
guish between volumetric and shear deformations, the deformation gradient F can be decomposed into volumetric
F vol := J 1/3 1 and unimodular F̄ := J −1/3 F parts
F = F vol F̄ . (9)
N2 N2
e3 N3 N3
ν1 a 2
0
2
ν 3a
0
0
0
λ1 a
λ1 a
e2 2
ν 2a
0
a0
a0 a0
a0 λ2 λ2
e1
a0 a0 a0
λ3 λ3
N1 N1
(a) (b) (c)
P
Figure 2: Geometric interpretation of the eigenvalue and eigenvectors of the right stretch tensor U = C 1/2 where U = 3a=1 λa N a ⊗ N a .
2 2 2 2 2 2
The invariants (a) I1 (C) = λ1 + λ2 + λ3 as sum of squares of the principal stretches, (b) I2 (C) = ν1 + ν2 + ν3 as sum of squares of principal
areal stretches where νi = J/λi and (c) I3 = J 2 = λ21 λ22 λ23 as square of volumetric stretch.
2.2. Multiplicative split of the deformation gradient and geometric mappings relative to intermediate configuration
In the sense of Kröner [45] and Lee [50], the deformation gradient can be multiplicatively decomposed into elastic
F e and viscous parts F v , i.e.
F = F eF v . (10)
The multiplicative decomposition of the deformation gradient requires the description of tangent T̄X̄ B and co-tangent
spaces T̄X̄∗ B in a fictitious intermediate configuration, respectively. Moreover, we describe a metric tensor G̃ :
T̄X̄ B → T̄X̄∗ B in the intermediate configuration which maps the unit vector in the tangent space onto its counterpart
in the co-tangent space. In the sequel, we introduce metric and deformation tensors relative to the intermediate
configuration as illustrated in Fig. 3. With the help of the intermediate configuration, one can describe the elastic right
Cauchy Green tensor and the elastic inverse left Cauchy Green tensors
C e = F eT gF e and ce = F e−T G̃F e−1 , (11)
as the pull-back of the current metric g to the intermediate configuration and the push-forward of the intermediate
metric G̃ to the current configuration, respectively. Moreover, it can be shown that the viscous metric C v = F T ce F
is the pull-back of the inverse left Cauchy Green tensor to the reference configuration.
5
2.3. Material and spatial rates of stress and deformation tensors
The material and the spatial velocity gradients are given as
Ḟ := ∇X ẋ and l = ∇x ẋ = Ḟ F −1 , (12)
respectively. By analogy with (10), one can describe the elastic and viscous parts of the spatial velocity gradient as,
The relationships between l, le and lv can be established via the insertion of (10) into (12)2 yielding
v v
l = le + F e lv F e−1 or l = le + l̃ where l̃ := F e lv F e−1 . (14)
The spatial velocity gradient l can be decomposed into covariant symmetric and covariant skew-symmetric parts
engendering the rate of deformation and the vorticity tensors, i.e.
respectively. In line with the additive decomposition of the spatial velocity gradients given in (14), the symmetric rate
of deformation tensor d can also be additively decomposed into elastic and viscous (inelastic) parts
v
d = de + d̃ . (16)
Next, we emphasize the push-forward of the covariant (•)♭ and the contravariant (•)♯ objects
By the same token, the pull-back of the covariant (•)♭ and the contravariant (•)♯ objects read
The Lie derivative of an Eulerian object is the push-forward of the material time derivative of the corresponding
˙
Lagrangian object, i.e. ϕ ∗ {ϕ ∗ (•)}. Accordingly, the objective Lie derivative of the Kirchhoff stress tensor reads
£ν τ = τ̇ − lτ − τ lT , (19)
where τ̇ is the material or the total time derivative of the Kirchhoff stress tensor.
F F
TX B0 T̄X̄ B Tx B TX B0 T̄X̄ B Tx B
Fv Fe Fv Fe
C Ce g Cv G̃ ce = be−1
(a) F −T (b) F −T
Figure 3: Push-forward and pull-back relations of the current and intermediate metric tensors. (a) Pull-back of the current metric g to the
Lagrangian C = F T gF and the intermediate C e = F eT gF e configurations. (b) Pull-back of the intermediate metric G̃ to the Lagrangian
C v = F vT G̃F v and the its push-forward to the Eulerian configuration ce = F e−T G̃F −1 .
6
Ψe
ψ1ne γ̇1
σ σ
ψnne γ̇n
F ek F vk
F = F ek F vk
Figure 4: The generalized Maxwell-Wiechert viscoelastic solid with n elements in total for k ∈ {1 . . . n} represented in a geometrically nonlinear
setting. The nonlinear spring Ψe accounts for the ground state elastic response, while the nonlinear spring ψkne and the dashpot with the associated
creep rate γ̇k in each Maxwell element reflect the non-equilibrium (viscous) response of the material considered.
The emphasis is placed upon the formulation of constitutive relations with regard to the equilibrium and non-
equilibrium response of rubberlike materials in the sense of generalized Maxwell-Wiechert model under finite defor-
mations. Besides, a new evolution equation for the creep/relaxation rate based on the relaxation kinetics of a single
polymer chain is the subject matter of this section.
Unlike the classical treatment for rubberlike materials, where volumetric and isochoric response are strictly de-
coupled, in what follows, we adopt a specific form of free energy density function
which reflects the standard viscoelastic solid, i.e. Maxwell elements are established parallel to the non-linear spring
characterizing the ground state elastic response. Hence, the non-equilibrium part of the free energy in (20) follows
from the rheological structure presented in Fig. 4 that describes the generalized Maxwell-Wiechert model with n
elements,
Xn Xn
Ψne (g, F e ) = ψkne (g, F ek ) or Ψne (C e ) = ψkne (C ek ) , (21)
k=1 k=1
where the index k indicates nothing but the Maxwell elements, i.e. k ∈ {1 . . . n} incorporated into the system.
Recalling (10), the multiplicative split of the deformation gradient for the non-equilibrium part now assumes distinct
deformation process, thereby being recast in the following form
F = F ek F vk . (22)
Accordingly, the total Kirchhoff and second Piola-Kirchhoff stress experienced by the viscoelastic material are given
by
τ := 2∂g Ψ(g, F , F e ) = τ e + τ ne and S := 2∂C Ψ(C, C e ) = S e + S ne , (23)
with the Kirchhoff stress terms τ e and τ ne characterizing the equilibrium (elastic) and the non-equilibrium (viscous)
stresses, respectively. By the same token, the second Piola-Kirchhoff stress terms S e and S ne indicate the equilibrium
(elastic) and the non-equilibrium (viscous) stresses, respectively.
7
3.2. Equilibrium part of the constitutive model: Extended eight-chain model
In view of (23), the equilibrium part of the constitutive response can be defined as follows
In the following sub-section, we will elaborate on the specific form of the constitutive response in the sense of Kirch-
hoff stress τ e .
where θ and η stand for the temperature and the entropy, respectively, while kB denotes the Boltzmann constant.
The entropy in (25) additively decomposes into free and constraint parts, i.e. ηf = kB ln pf and ηc = kB ln pc ,
respectively. Consequently, the free energy function of a single chain in a tube-like constraint environment reads
The non-Gaussian probability density function for the free chain response by Kuhn and Grün [48] reads
−1 L−1 (λr ) λ
pf (λ) = p0 exp −N λr L (λr ) + ln −1
with λr = √ , (27)
sinh L (λr ) N
where L−1 (λr ) is√the inverse Langevin function of the relative stretch λr ∈ [0, 1) normalized with respect to the
extensibility limit N characterized by the number of segments N . The probability density function for the tube-like
constraint in the sense of [60, 20] is given as
" 2 #
r0 −1
pc (ν) = p0 exp −α ν , (28)
d0
√
where r0 = N l is the mean end-to-end distance of the chain, d0 is the tube diameter, α is a shape constant and
ν −1 = (d0 /d)2 is the tube areal contraction and ν = (d/d0 )2 is the tube areal stretch. Incorporation of (27) and (28)
into (25) leads to the free energy functions corresponding to the free-chain
−1 L−1 (λr )
ψf (λ) = N kB θ λr L (λr ) + ln + ψ0 (29)
sinh L−1 (λr )
and the tube-like constraint 2
l
ψc (ν) = αkB θN ν −1 + ψ0 , (30)
d0
respectively. The key issue in the modeling of macroscopic response of a polymer network lies in the non-affine
mean-kinematic variables with the affine macro-kinematic variables. Within this context, the macroscopic free energy
functions can be additively split into
hnψf (λ̄)i = nψf (λn ) and hnψc (ν̄ −1 )i = nψc (νn ) (33)
in terms of the average kinematic quantities λn and νn , see Fig. 5. At this point, the average network stretch and the
tube areal stretch are assumed as
r r r r
2 2 2 2 2 2
2 λ1 + λ2 + λ3 2 I1 3 ν1 + ν2 + ν3 3 I2
λn = = and νn = = . (34)
3 3 3 3
Herein, the network stretch is the mean-square root average of the principal stretches λi in the sense of [1]. The mean
non-affine tube areal contraction νn of the polymer network in constraint environment is linked to the macroscopic
areal stretches in terms of (34)2 . As a result, the free-chain and the tube-constraint part of the macroscopic free energy
function for the extended eight-chain model read
e L−1 (λr )
−1
Ψf (g, F ) = µN λr L (λr ) + ln and Ψec (g, F ) = µc (νn − 1) , (35)
sinh L−1 (λr )
√
where λr = λn / N represents the relative average network stretch. The shear modulus µ = nkB θ is associated with
the free-chain response and the µc = α(l/d0 )2 nkB θ is associated with the tube constraint part.
A = νn A 0
e3
A0 = d20 N2
N3
r0
λn
r0
0
λ1 a
a0 e2
λ 2a
0
a0
0
λ3 a
e1 a0 N1
Figure 5: Idealization of the rubber network by eight-chain model and the linkage between macroscopic stretches with mean streches of the
representative chain,Prespectively. N i for i = {1, 2, 3} represent the eigenvectors of the right Cauchy-Green tensor, whereas λi are the principle
stretches, i.e. C = 3i=1 λ2i N i ⊗ N i .
where χ = κ − 23 µ and µ are the first Lamé constant and the shear modulus, respectively. The second expression in
the free energy function is the normalization term that enforces τ e = 0 at undeformed state. The stress expression for
the compressible eight-chain model reads
hχ i
τ e = 2∂g Ψe (J, I1 ) = (ln J + J(J − 1)) − µ̃ g −1 + µ̂(λn )b , (37)
2
9
where the two terms related to the shear modulus are given as
µ 3N − 1 µ 3N − λ2n
µ̃ = and µ̂(λn ) = . (38)
3 N −1 3 N − λ2n
For a Gaussian network where the number of segments approaches infinity, i.e. N → ∞, the parameters simplify to
µ̂ → µ and µ̃ → µ in (38) so that one recovers the compressible neo-Hookean model. Advancing a step forward from
(36), an extension to compressible eight-chain model can be achieved
Ψe (g, F ) = Ψev (g, F ) + Ψef (g, F ) + Ψec (g, F ) = Ψ̃ev (J) + Ψ̃ef (λr ) + Ψ̃ec (ν) , (39)
in conjunction with (35). Therein, the purely volumetric expression for the extended eight-chain model reads
χ µ 3N − 1 4
Ψev (g, F ) = Ψ̃ev (J) = [(ln J)2 + (J − 1)2 ] − ln J − µc ln J , (40)
4 3 N −1 9
in which the second and the third expressions are added in order to normalize the stresses at undeformed configuration.
Then, the stress expression for the compressible extended eight-chain model is recast into the following form,
2 µc
τ e = 2∂g Ψe (g, F ) = κ̂(J)g −1 + µ̂(λn )b + (I1 b − b2 ) , (41)
9 νn2
where
χ µ 3N − 1 4 µ 3N − λ2n
κ̂(J) = (ln J + J(J − 1)) − µ̃, µ̃ = + µc and µ̂(λn ) = . (42)
2 3 N −1 9 3 N − λ2n
The first expression in the stress equation (41) is the purely volumetric part, whereas the latter two expressions are
due to free chain response and the tube constraint, respectively. For various volumetric free energy functions and their
implementation into fnite element method in the quasi-incompressible setting, we refer to Kadapa and Hossain [36].
Therein, the individual overstress expression for each Maxwell element is given by
ne
τ̂ ne ne e ne e
k = 2∂g ψk (g, F k ) and Ŝ k = 2∂C e ψk (C k ) , (44)
according to the rheology illustrated in Fig 4. In the following sub-section, we will elucidate the individual contribu-
tions arisen from the particular selection of the free energy function ψkne (g, F ek ) in the sense of the Kirchhoff stress
τ̂ ne
k . Before giving an account of the constitutive modeling of the non-equilibrium response, we underline that the
subscript k reflecting the particular Maxwell branch in the rheological structure hereafter drops out of the relevant
equations, e.g., ψkne (g, F ek ), in order not to make the expressions too complicated. In the sequel, the effect of the tube
constraint effect on the non-equilibrium response of the rubber network is neglected. The compressible eight-chain
model for a single nonlinear elastic spring part of the Maxwell branch as follows,
χv µv 3N v − 1 β(λer )
Ψne (J e , I1e ) = ln J e + (J e − 1)2 − ln J e
+ µ v v
N λ e
r β(λe
r ) + ln . (45)
4 3 Nv − 1 sinh β(λer )
p p
Therein, λen = I1e /3 and λer = I1e /3N v denote the network and the relative network stretches for the non-
equilibrium deformations, respectively. β = L−1 (λer ) is the inverse Langevin function, µv is the shear modulus of
the non-equilibrium network, N v is the number of segments of the chains entangled around obstacles. Accordingly,
the Kirchhoff stress expression for the compressible eight-chain model is
v
χ
τ̂ ne = 2∂g ψ ne (J e , I1e ) = (ln J e + J e (J e − 1)) − µ̃v g −1 + µ̂v (λen )be , (46)
2
where the two terms related to the shear modulus read
µv 3N v − 1 µv 3N v − λe2
n
µ̃v = and µ̂v (λen ) = . (47)
3 Nv − 1 3 N v − λe2n
10
3.4. Thermodynamical consistency
The second axiom of thermodynamics restricts the proposed model by the so-called dissipation inequality
1
D := P − Ψ̇ ≥ 0 with P := S : Ċ = τ : d , (48)
2
being the stress power term. The reduced form of the dissipation inequality reads
v
Dred = τ̂ ne : d̃ ≥ 0 . (49)
The thermodynamical consistency is satisfied for the non-negative dissipation Dred . The evolution equation for the
v
viscous rate of deformation tensor d̃ must a priori satisfy the inequality (49).
r0 r r
Figure 6: Stretch and relaxation of a single chain entangled around an obstacle: (a) undeformed equilibrium state, (b) non-equilibrium deformed
state after rapid stretch, (c) deformed and fully relaxed state.
In this part, we elucidate the relaxation kinetics of a single free chain [14]. To this end, we revisit the relaxation
kinetics of a single chain entangled around an obstacle at both sides as depicted in Fig. 6. In Fig. 6(a),
√ a single chain
entangled around obstacles at both ends is shown at rest. The end-to-end distance is given by r0 = N0 ˆl where N0
is the number of segments between the two obstacles and l̂ denotes the length of a monomer unit. If the polymer
chain is instantaneously stretched, see Fig. 6(b), it elongates and attains a more ordered conformation decreasing its
entropy and increasing the free energy, while the internal energy of the chain is kept constant, see Treloar [82] (p. 25).
The stretching of the chain triggers the movement of the free ends of the chains and they retract in combination
of reptational and Brownian motion. Hence, the√length of the polymer chain portion lying between the obstacles
elongates until the most favorable state, i.e. r = N∞ l̂, is reached. Fig. 6(c) shows not only the deformed but also
the fully relaxed state in this case with the same end-to-end distance r. Then, the number of segments at the fully
relaxed state reads
r
N∞ = λ̄2 N0 where λ̄ = . (50)
r0
Recall that unlike N0 , which is a material constant, N∞ is variable dependent on the instantaneous stretch level λ̄. Let
us consider an intermediate step between rapid deformation
p as seen in Fig. 6(b) and fully relaxed state as indicated in
Fig. 6(c). We define an end-to-end distance r̄(t) = N (t)l̂ which is probabilistically the most preferred state upon
removal of the load maintaining the end-to-end distance r at time t. In a deformation driven state, λ̄ is given and can
be decomposed into elastic and inelastic (viscous) parts as follows
r r r̄(t) r r̄(t)
λ̄ = = = λe (t)λv (t) where λe (t) := and λv (t) := . (51)
r0 r̄(t) r0 r̄(t) r0
Initially at time t = 0, the end-to-end distance reads r̄(0) = r0 , leading to λe (0) = λ̄ and λv (0) = 1. However, at the
fully relaxed state at time t = ∞, the end-to-end distance reads r̄(∞) = r, leading to λe (∞) = 1 and λv (∞) = λ̄.
For the evolution of the chain length in-between the two obstacles, we adopt a generic ordinary differential equation
(ODE), i.e.
1
Ṅ (t) = [N∞ − N (t)] , (52)
τ
11
where τ designates the relaxation time. Then, the solution of the ODE reads
−t
N (t) = (N∞ − N0 ) 1 − exp + N0 . (53)
τ
Next, we substitute (53) into (51)3 and consider the rate of the viscous stretch λv (t) with respect to time. This, upon
some simple mathematical manipulations, lead to the following
p ! !
v d r̄(t) d N (t)ˆl 1 1 N∞ p
λ̇ (t) = = √ = √ p − N (t) . (54)
dt r0 dt N0 ˆl 2τ N0 N (t)
In order to comply with the large-strain kinematics in the Eulerian setting, the one-dimensional logarithmic counter-
part of the rate of deformation tensor is derived as follows,
λ̇v
ǫ̇v = = γ̇0 λe2 − 1 . (57)
λv
kτ ne k
τ ne = √ . (60)
2
p
In what follows, we also replace the λe expression for a single chain by the network stretch λen = I1e /3 of the
eight-chain model where I1e = tr C e . Finally, we end up with the following function for the effective creep rate
m
τ ne
γ̇ := γ̇0 [λe2
n − 1] . (61)
τ̂
In the above expression the relaxation time 1/2τ is replaced by the creep rate constant γ̇0 > 0 for the sake of
convenience. Note that the term m ≥ 1 controls the energy-activated relaxation and leads to viscoplastic material
response for high values. The expression τ̂ is used merely for normalization purposes and assumed to be τ̂ = 1 [MPa]
12
in the subsequent treatments. In the end, we emphasize that the nonlinear viscous flow stated in (61) is governed by
one more additional material parameter in comparison to the that of the finite linear viscoelasticity and reflects from
the continuum point of view the behavior of a single dashpot in the Maxwell element as depicted in Fig. 4. Recall that
replacing the term γ̇0 [λe2
n − 1] → γ̇0 and taking m = 1, we obtain the classical multiplicative finite viscoelasticity
in the sense of Reese and Govindjee [68]. Moreover, taking N → ∞, one obtains the simplest finite viscoelasticity
formulation in the geometrically nonlinear setting based on multiplicative decomposition of the deformation gradient
into elastic and viscous parts.
R EMARK: The proper prediction of viscous flow direction requires γ̇0 [λe2 n − 1] ≥ 0. In incompressible rubber
viscoelasticity, the condition is always satisfied since the network stretch λen is always greater than unity λe2
n ≥ 1.
However, for slightly compressible elastomers, under purely hydrostatic pressure loading, it is possible to have λe2
n ≤
1. To resolve this issue, the term [λe2 e2
n − 1] must be replaced by |λn − 1| in the evolution equation (61) for the analysis
of compressible elastomers.
In this section, the algorithmic setting of the model suitable for a finite element implementation will be discussed.
In particular, the algorithmic framework for the Eulerian stresses and moduli terms of the equilibrium and non-
equilibrium responses will be established. Furthermore, we will propose an efficient implicit update for the principle
elastic stretches of the non-equilibrium Maxwell branches, conceptually following the references [14, 68, 17].
In a time-discrete interval, i.e. ∆t = tn+1 −tn , the current stresses and the tangent expressions need to be obtained
where ∆t stands for the time-increment. The time-discrete counterpart of the continuous expression in (19) reads
where ∇x (·) represents the spatial gradient of the term considered. Therein, Calgo is the total algorithmic moduli and
must be consistently derived for the accurate and quadratically convergent nonlinear finite element analysis based on
global iterative solution algorithms, e.g., the Newton-Raphson method. Subsequently, the total algorithmic moduli
term is also additively decomposed into equilibrium and non-equilibrium parts as follows
Calgo := Ce + Cne
algo . (63)
which gives rise to the following explicit form of the equilibrium tangent moduli through the particular form of the
free energy (39),
Ce = (ŝe + p̂e)g−1 ⊗ g−1 − 2p̂eI + n̂e b ⊗ b + fˆe(I1 b − b2) ⊗ (I1 b − b2) + ĝe(b ⊗ b − Ib ). (65)
∂ 2 Ψe χ ∂ 2 Ψe χ 2
p̂e := J = [ln J + J(J − 1)] − µ̃ and ŝe := J 2 = J − ln J + 1) + µ̃ , (66)
∂J 2 2 ∂J 2 2
while the terms associated with I1 and I2 are written as
4µ 1 8µc 1 4µc 1
n̂e = , fˆe = − and ĝ e = . (67)
9N (1 − λ2r )2 81 νn5 9 νn2
13
The fourth-order tensor Ib is simply the push-forward of the fourth-order identity tensor I such that
1 ik jl 1 ik jl
Iijkl = δ δ + δ il δ jk and Ib ijkl = b b + bil bjk . (68)
2 2
In the above equation, the inverse Langevin function is computed by Padé approximation L−1 (λr ) ≈ λr (3−λ2r )/(1−
λ2r ) proposed by Cohen [10].
Both in (69) and (70) the Lie derivative of the elastic Finger tensor be is given which can be defined as follows
d
£ν be := F C v−1 F T . (71)
dt
During the elastic predictor step E (elastic trial step), d(C v−1 )/dt is equal to zero. Therefore,
It should be highlighted that the subscript denoting the current time step tn+1 hereinafter drops out of the expressions
for convenience. The merit of this proposal is to circumvent the rate of inelastic metric in computation of the consistent
tangent moduli. Afterwards, we focus on the inelastic corrector step I. Here, the spatial velocity gradient l is set to
e
zero yielding ḃ = £ν be . Inserting this result into (69) and substituting the evolution equation proposed in (58) for
v
d̃ , we obtain the following expression,
e
Inelastic corrector step I : ḃ = −2γ̇Nbe tr , (73)
for the evolution equation for the inelastic corrector step. The interested reader is referred to references [14, 17],
regarding integration of the equations [72,73] in a time-discrete setting.
in the intermediate configuration. In (75)2, the principal value of the fictitious second Piola-Kirchhoff stress reads
2
s̃a = τa /λea, tr . With the definition (75)1 at hand, the incremental rate equation can be defined as follows
∆S̃
ne
= Cne e
algo : ∆C tr where Cne
algo = 2∂C e
tr
S̃
ne
. (76)
Next, we consider (76)2 via exploiting (75)2 and obtain the moduli expression as follows
3 X
3 3
Cne cab − 2τa δab 2τa
X 2 X
algo = 2 2 N a ⊗ N a ⊗ ∂C etr (λeb, tr ) + 2 ∂C etr (N a ⊗ N a ) , (77)
a=1 b=1 λea, tr λeb, tr e
a=1 λa, tr
along with the respective partial derivatives with respect to the trial elastic right Cauchy-Green tensor,
3
1
Gab + Gba) ,
2 X
∂C etr (λeb, tr ) = N b ⊗ N b and ∂C etr (N a ⊗ N a ) = ( (78)
2(λea, tr
2
− λeb, tr )
2
b6=a
5. Experimental Analyses
In this section, we will perform uniaxial and equibiaxial extension tests on nitrile butadiene rubber (NBR, see
Table 1 for composition) in order to obtain the hyperelastic and viscoelastic characteristics of the material. In the
Table 1: NBR rubber composition.
sequel, the relaxation and creep behavior of the same NBR material will be assessed by means of uniaxial extension
tests. These experimental studies are conducted on a modular Zwick/Roell Z010 universal testing machine with two
video extensometers dedicated specifically to uniaxial and equibiaxial tests. The TestXpert III software associated with
the test machine is used in implementing experimental protocols, measurements, and control of the testing machine.
Pincer grips, especially designed for rubber specimens, are mounted between the Zwick/Roell XForceP series load cell
with 500 N nominal force capacity.
15
5.1. Uniaxial and equibiaxial extension test on NBR rubber
The specimens are cut out from the same beige colored NBR rubber sheet with 2 mm thickness. The preparation
of the uniaxial test specimen complies with ISO37-Type2 geometry [35]. In order to retrieve the hyperelastic re-
sponse of the material, single cycle loading-unloading tests are performed at various crosshead speeds, i.e {10, 1, 0.1}
[mm/min]. It is observed that at/below 1 [mm/min] constant displacement rate, no hysteresis is observed in the spec-
imen. For the quasi-static investigation, a displacement controlled free test definition with 0.1 [mm/min] constant
displacement rate is chosen and maximum stretch λ = 2 is set via the test configuration options. The result of the
experiment is documented in Fig. 7(a) in the sense of stress (first Piola-Kirchhoff stress) versus stretch values. Con-
0.8 1
0.8
0.6
P11 [MPa]
P11 [MPa]
0.6
0.4
0.4
0.2 0.2 raw data
corrected curve
0 0
1 1.2 1.4 1.6 1.8 2 1 1.1 1.2 1.3 1.4
(a) λ (b) λ
Figure 7: (a) Uniaxial and (b) equibiaxial tension test results in terms of the first Piola-Kirchhoff stress versus stretch values.
cerning the equibiaxial experiments, a custom designed equibiaxial fixture is used. The specimen geometry required is
adopted from Axel Physical Testing Services [3]. The lower plate of the custom equibiaxial test apparatus is mounted
on the Zwick/Roell XForceP series load cell having 10 [kN] nominal force capacity, while its upper plate is fixed to
the cross-head and moves together with it. A virgin specimen is marked with four black dots and placed between 16
slider guided connection points of the upper plate. Steel wires connect the upper and lower plates through pulleys
and allow pulling the sliders as a result of cross-head movement, thereby applying equal forces at the periphery of
the specimen from the connection points. For a quasi-static investigation, a 1 [mm/min] constant displacement rate
is chosen from the test definition settings for both loading and unloading during a three-cycle experiment. The max-
imum stretch value is set to λ = 1.4, whereas the point of load removal at the end of a cycle is chosen to occur at 0
[MPa]. The result of the experiment is illustrated in Fig. 7(b). Within the third cycle, where λ = 1.306, the specimen
ruptured.
1 1
λ = 1.05 λ = 1.05
0.8 λ = 1.15 0.8 λ = 1.15
P11 [MPa]
P11 [MPa]
λ = 1.45 λ = 1.45
λ = 2.20 λ = 2.20
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 100 200 300 400 500 600 0 20 40 60 80 100
(a) time [s] (b) time [s]
Figure 8: Relaxation experiments at different stretch levels, i.e. λ = 1.05, λ = 1.15, λ = 1.45, and λ = 2.2 for (a) the entire duration of the
experiments, (b) first 100 [s].
The target stress values for the creep experiments are, respectively, P11 = 0.1, P11 = 0.25, P11 = 0.5, and P11 = 1.
[MPa]. Between successive experiments, a 30 minutes waiting time is applied. The results of the creep experiments
is depicted in Fig. 9.
2.25 2.25
1.5 1.5
1.25 1.25
1 1
0 100 200 300 400 500 600 0 20 40 60 80 100
(a) time [s] (b) time [s]
Figure 9: Creep experiments at P11 = 0.1, P11 = 0.25, P11 = 0.5, and P11 = 1 in [MPa] for (a) the entire duration of the experiments, (b) first
100 [s].
The following sub-sections are concerned with some meticulous and highly erudite assessments of the proposed
extended eight-chain model in Sections 2-4 in terms of (i) the validation and the performance of hyperelastic and
viscoelastic frameworks under homogeneous uniaxial, biaxial, equibiaxial extensions and pure shear tests reported in
other studies as well as data obtained from our experiments portrayed in Section 5; (ii) the capability of the model
in describing viscoelastic mechanical response under cyclic loading by either referring to direct experimental mea-
surements conducted in Section 5 or representative geometries of relevant devices such as conical springs. To this
end, the in-house material and element subroutines are coded in the general-purpose finite element analysis program
FEAP 8.5 originally developed by Taylor [79] and is used as a simulation tool. Apart from that, the optimization tools
developed in-house for the nonlinear least squared analysis, is extensively used, see Dal et al. [15].
17
6.1. Comparison of the proposed extended eight-chain model with the eight-chain, Pucci-Saccomandi, and Carroll
models
The performance of the extended eight-chain model is compared with the eight-chain, Pucci-Saccomandi, and
Carroll models, well known 3-parameter models, in view of the classical Treloar [81] and Kawabata et al. [39]
datasets. Treloar data is used for the comparison of the uniaxial (UT), equibiaxial (ET), and pure shear (PS) responses,
whereas the Kawabata data is solely used to examine the biaxial (BT) response of the different models. In line with
Dal et al. [16, 15], the corresponding model fits are obtained with quality of fit values. In part (i) of Fig. 10, the
eight-chain model results are shown for uniaxial extension (UT) only, simultaneous (uniaxial (UT), equibiaxial (ET)
extension and pure shear (PS) tests combined), and biaxial extension (BT) tests with several initial stretch values for
λ1 . Likewise, the extended eight-chain model behavior is presented for the same tests in part (ii) of Fig. 10. The
retrieved material parameters are also listed in Table 2.
Table 2: Eight-chain and extended eight-chain model parameters fitted to the Treloar data and Kawabata (biaxial) data. The parameters in the first
row are obtained by fitting to the UT-data only. The second row parameters are obtained by simultaneous fit to the UT,ET and PS data.
Upon a closer look, the eight-chain model produces quite good results for UT and and PS under uniaxial loading
only. However, ET response of the model greatly underestimates the data points, as visualized in Fig. 10(i)(a). Though
the performance of the model is noticeably improved with slightly underestimated data points for ET in the case of
simultaneous fits, the results this time overestimate the data taken from UT and PS as observed in Fig. 10(i)(b). Such
a response is expected and has been previously observed by others which motivates the development of the extension
of the eight-chain model. Focusing on the data retrieved from Kawabata et al. [39], the predictions of the eight-chain
model agree quite well with the data for relatively small stretch values, i.e. λ1 = {1.04 . . . 1.24}, as depicted in
Fig. 10(i)(c). However, the predictions start to fall short of tracking the data for moderate stretch values and finally
exhibits an asymptotic behavior as the spectrum of the model response in terms of the first Piola-Kirchhoff stress
narrows down to a single response curve for large stretch values, see Fig. 10(i)(d).
Fig. 10(ii)(a) and Fig. 10(ii)(b) signify the substantial enhancement of the modeling quality associated with the
extended eight-chain model for both UT only and simultaneous tests. As matter of fact, the addition of the term
involving the second invariant I2 brings the modeling capability to such an extent that even a UT only tests seem
to be sufficient for the characterization of the homogeneous material behavior. Similarly, the new model are in
excellent conformity with the BT tests irrespective of the degree of the stretch, clearly displayed in Fig. 10(ii)(c) and
Fig. 10(ii)(d). In fact, the hyperelastic response of the material for small, moderate and larger stretches is excellently
traced by the model.
6 6
5 5
P11 [MPa]
P11 [MPa]
4 4
3 3
2 2
1 1
0 0
1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8
(a) (b)
λ λ
BT fit with λ1 : 1.04 − 1.24 BT fit with λ1 : 1.3 − 3.7
0.4 1.4
0.3 1.05
P22 [MPa]
P22 [MPa]
6 6
5 5
P11 [MPa]
P11 [MPa]
4 4
3 3
2 2
1 1
0 0
1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8
(a) λ (b) λ
BT fit with λ1 : 1.04 − 1.24 BT fit with λ1 : 1.3 − 3.7
0.4 1.4
0.3 1.05
P22 [MPa]
P22 [MPa]
Figure 10: (i) Eight-chain, and (ii) extended eight-chain model predictions for (a) uniaxial tension, (b) combination of uniaxial, equibiaxial, and
pure shear loadings using Treloar data (c) biaxial tension loading for λ1 : 1.04 − 1.24, (d) biaxial tension loading for λ1 : 1.3 − 3.7 using
Kawabata data. 19
(i) UT only fit Simultaneous UT, ET and PS fits
7 7
6 6
5 5
P11 [MPa]
P11 [MPa]
4 4
3 3
2 2
1 1
0 0
1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8
(a) (b)
λ λ
BT fit with λ1 : 1.04 − 1.24 BT fit with λ1 : 1.3 − 3.7
0.4 1.4
0.3 1.05
P22 [MPa]
P22 [MPa]
6 6
5 5
P11 [MPa]
P11 [MPa]
4 4
3 3
2 2
1 1
0 0
1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8
(a) λ (b) λ
BT fit with λ1 : 1.04 − 1.24 BT fit with λ1 : 1.3 − 3.7
0.4 1.4
0.3 1.05
P22 [MPa]
P22 [MPa]
Figure 11: (i) Carroll, and (ii) Pucci-Saccomandi’s model predictions for (a) uniaxial tension, (b) combination of uniaxial, equibiaxial, and pure
shear loadings using Treloar data (c) biaxial tension loading for λ1 : 1.04 − 1.24, (d) biaxial tension loading for λ1 : 1.3 − 3.7 using Kawabata
data. 20
Table 3: Ground state elastic material parameters of the extended eight-chain model simultaneously fitting the experimental data reported in
Section 5.1.
0
P11 [MPa]
0
0
UT model fit
0 UT exp
ET model fit
ET exp
0
1 1 1 1 1 2
λ
Figure 12: Extended eight-chain model fits against the data received from UT and ET experiments performed on NBR rubber.
6.3. Assessment of the viscoelastic behavior of extended eight-chain model in terms of relaxation & creep
We now assess how well the model proposed fits the experimental data. For this purpose, the model is simultane-
ously fitted against the data obtained from homogeneous tensile uniaxial tests characterizing relaxation experiments
with ultimate stretches λ = 1.05, λ = 1.15, λ = 1.45, and λ = 2.20 as seen in Fig. 8 and creep experiments with
ultimate stresses P11 = 0.1, P11 = 0.25, P11 = 0.5, and P11 = 1 [MPa], see Fig. 9. For the baseline elastic response,
the parameters identified in Table 3 is used. Firstly, we try to fit the model response against both data concurrently
by means of a single Maxwell branch. However, all the attempts invariably fall flat that we cannot achieve model
fits satisfactorily close to the experimental data. Therefore, one more Maxwell branch is added into the rheological
structure of the model and the simultaneous fitting procedure in the sense of trial and method starts again. As can
be observed in Fig. 13 and Fig. 14, both relaxation and creep response of the viscoelastic model subjected to several
target stretch and stress conditions under uniaxial extension are traced very well. The respective viscous material
parameters are reported in Table 4. Also shown in Fig. 13(b) and Fig. 14(b) are the zoom-in view of both relaxation
and creep responses during the first 100 [s], respectively. Apparently, the slight deviations from data observed in
relaxation, i.e. at λ = 1.45 and λ = 2.20, and the creep curves, i.e. at P11 = 0.5, and P11 = 1 [MPa], vanish as
the relaxation and creep processes continues to evolve in time. It needs to be highlighted that a fully incompressible
viscous free energy, i.e. χv = 0 [MPa], which appears in (45) may yield jumps in the relaxation and creep behavior
upon the descent and ascent of the curves, respectively. For this reason, a sufficiently small value for χv is hereinafter
considered throughout the examples, see Table 4. In addition, the segment number of the non-equilibrium network is
taken equal to the that of the equilibrium network, i.e. N v ≡ N , due to the lack of the physical information.
21
1.2 1.2
P11 [MPa]
λ1 = 1.45 λ1 = 1.45
λ1 = 2.20 λ1 = 2.20
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 100 200 300 400 500 600 0 20 40 60 80 100
(a) time [s] (b) time [s]
Figure 13: Model predictions against the experimental data of relaxation tests at λ1 = 1.05, λ1 = 1.15, λ1 = 1.45, and λ1 = 2.20 for a) the
entire duration of the experiment (600 [s]); b) the first 100 [s].
2.25 2.25
λ1 [-]
1.25 1.25
1 1
0 100 200 300 400 500 600 0 20 40 60 80 100
(a) time [s] (b) time [s]
Figure 14: Model predictions against the experimental data of creep tests at P11 = 0.1, P11 = 0.25, P11 = 0.5, and P11 = 1 [MPa] for a) the
entire duration of the experiment (600 [s]); b) the first 100 [s].
x
r1 h1 h2 h3 h4
r2
r3
rubber
r4
steel
(a) (b) (c)
Figure 15: (a) Schematic view of a conical spring composed of two materials, namely NBR rubber and stainless steel, (b) Discretization with
N UMEL= 5300 eight-noded mixed Q1P0 brick elements connected by N UMNP = 6200 nodes. (c) The tailored geometry involves the following
radii, i.e. r1 = 50, r2 = 64, r3 = 80, r4 = 100 and r5 = 72 [mm], while the heights are h1 = 35, h2 = 55, h3 = 75 and h4 = 125 [mm].
The domain is subjected to displacement controlled load based on Fig. 16(a).
for the essentially elastic mechanical response of the stainless steel are chosen as χ = 104 [MPa] and µ = 76923
[MPa]. Based on our experience during simulations we note that assigning an identical value for the first Lamé
parameter χ in the NBR rubber and the stainless steel seems indispensable in order to hinder the numerical stability
issues due to the mismatch in the volumetric constraint.
5 0
0 −6
T = 1 [s]
−5 −12 T = 10 [s]
−10 −18 T = 100 [s]
−15 −24
tB tD
−20 −30
0 1 2 −20 −10 0 10 20
Number of cycles uz [mm]
(a) (b)
Figure 16: (a) Conical spring undergoes a cyclic sinusoidal loading in a displacement driven context with the displacement uz (t). The corre-
sponding (b) force-displacement curves attest to the fact that the higher the loading rate (small period) is, the stiffer the mechanical response
becomes.
23
aspect of any finite element analysis, we also report in Fig. 17 the Cauchy stress distribution over the conical spring
with four snapshots taken at the instants tA , tB , tC , and tD on the cyclic loading path shown in Fig. 16(a). Aside
from the top portion of the steel which is, in fact, directly subjected to the loading, the absolute maximum stresses are
obviously concentrated in the NBR rubber components. Moreover, the larger areas of viscous overstresses are clearly
demonstrated in the rubber parts which are related to T = 1 [s].
(a) tA tB tC tD
(b) tA tB tC tD
(c) tA tB tC tD
7. Conclusion
In this article, we proposed a new constitutive framework that accounts for the finite viscoelastic mechanical
response of rubber-like materials and touched upon the theoretical, algorithmic, experimental and finally the numerical
aspects thereof. For the ground state elastic response, the eight-chain model, originally proposed by Arruda and Boyce
[1], was extended via the addition of a tube constraint part that is macroscopically linked to the the second invariant
I2 . To this end, the relation between the tube area contraction of the representative chain and the second invariant
is established. Special emphasis was given to the continuum formulation in which the multiplicative split of the
deformation gradient into volumetric and isochoric part was completely dispensed with and all the consequences
thereof drop out of the theoretical formulation. Consequently, compressible counterparts of the eight-chain and the
extended eight-chain models are derived. Such an ansatz recovers the quasi-incompressible behaviour or rubberlike
24
materials as the large κ/µ ratios. Moreover, it allows modeling of slightly compressible filled rubber and description
of relevant degradation mechanisms prior to failure. The proposed extended eight-chain model shows an excellent
fitting performance to Kawabata and Treloar data with only 3-material parameters. The computational efficiency
is considerably better than the well known accurate models in the literature. Algorithmic treatment features the
spectral decomposition of the respective terms and is demonstrated within the context of the finite element method.
Having based our modeling endeavors on micro- and macro-physical grounds, we enriched the study via experimental
observations whose results were directly compared to the model behavior thereby elucidating the performance and the
validity of the model proposed. In fact, the results are in close agreement with the creep and relaxation experiments
carried out at various amplitudes. Later, analyses pertaining to non-homogeneous tests were coducted by means of
a custom designed conical spring undergoing displacement driven cyclic extensions. The finite element simulations
yielded results that are in excellent conformity with the propensities of viscous materials. The proposed numerical
implementation is very robust and quadratically convergent.
Acknowledgment: The authors gratefully acknowledge the financial support from the Scientific Research Projects
Coordination Unit of Middle East Technical University under grant number GAP-302-2020-10203 and T ÜBITAK
Career Award 3501 under grant 315M140.
References
[1] Arruda, E. M., & Boyce, M. C. (1993). A three–dimensional constitutive model for the large stretch behavior of rubber elastic materials.
Journal of the Mechanics and Physics of Solids, 41, 389–412.
[2] Attard, M. M., & Hunt, G. W. (2004). Hyperelastic constitutive modeling under finite strain. International Journal of Solids and Structures,
41, 5327–5350.
[3] Axel Products Physical Testing Services (2019). axelproducts.com. Accessed: 2019-09-28.
[4] Ball, R. C., Doi, M., Edwards, S. F., & Warner, M. (1981). Elasticity of entangled networks. Polymer, 22, 1010–1018.
[5] Bechir, H., Chevalier, L., Chaouche, M., & Boufala, K. (2006). Hyperelastic constitutive model for rubber-like materials based on the first
seth strain measures invariant. European Journal of Mechanics-A/Solids, 25, 110–124.
[6] Bergström, J. S., & Boyce, M. C. (1998). Constitutive modeling of the large strain time–dependent behavior of elastomers. Journal of the
Mechanics and Physics of Solids, 46, 931–954.
[7] Biderman, V. (1958). Calculation of rubber parts. Rascheti na prochnost, 40.
[8] Boyce, M. C., & Arruda, E. M. (2000). Constitutive models of rubber elasticity: A review. Rubber Chemistry and Technology, 73, 504–523.
[9] Christensen, R. (1982). Theory of Viscoelasticity. (2nd ed.). Academic Press.
[10] Cohen, A. (1991). A Padé approximant to the inverse langevin function. Rheological Acta, 30, 270–273.
[11] Coleman, B., & Noll, W. (1961). Foundations of linear viscoelasticity. Quarterly of Applied Mathematics, 33, 239–249.
[12] Coleman, B. D. (1964). Thermodynamics of materials with memory. Archive for Rational Mechanics and Analysis, 17, 1–46.
[13] Cuitino, A., & Ortiz, M. (1992). A material-independent method for extending stress update algorithms from small-strain plasticity to finite
plasticity with multiplicative kinematics. Engineering Computations, 9, 437–451.
[14] Dal, H. (2011). Approaches to the modeling of inelasticity and failure of rubberlike materials: theory and numerics. Ph.D. thesis Technische
Universität Dresden.
[15] Dal, H., Açıkgöz, K., & Badienia, Y. (2021). On the performance of hyperelastic constitutive models for rubber-like materials: A review.
Applied Mechanics Reviews, . To be submitted.
[16] Dal, H., Badienia, Y., Açıkgöz, K., & Denli, F. (2019). A comparative study on hyperelastic constitutive models on rubber: State of the art
after 2006. In Constitutive Models for Rubber XI. Proceedings of the 11th European Conference on Constitutive Models for Rubber (ECCMR
XI), June 25-27, 2019, Nantes, France.
[17] Dal, H., & Kaliske, M. (2009). Bergström–Boyce model for nonlinear finite rubber viscoelasticity: Theoretical aspects and algorithmic
treatment for FE method. Computational Mechanics, 44, 809–823.
[18] Dal, H., Zopf, C., & Kaliske, M. (2018). Micro-sphere based viscoplastic constitutive model for uncured green rubber. International Journal
of Solids and Structures, 132-133, 201–217.
[19] Davidson, J. D., & Goulbourne, N. (2013). A nonaffine network model for elastomers undergoing finite deformations. Journal of the
Mechanics and Physics of Solids, 61, 1784 – 1797.
[20] Doi, M., & Edwards, S. F. (1986). The Theory of Polymer Dynamics. (1st ed.). Clarendon Press, Oxford.
[21] Eidel, B., Stumpf, F., & Schröder, J. (2013). Finite strain viscoelasticity: how to consistently couple discretizations in time and space on
quadrature-point level for full order p ≥ 2 and a considerable speed-up. Computational Mechanics, 52, 463–483.
[22] Erman, B., & Flory, P. J. (1978). Theory of elasticity of polymer networks. ii. the effect of geometric constraints on junctions. The Journal
of Chemical Physics, 68, 5363–5369.
[23] Ferry, J. D. (1980). Viscoelastic Properties of Polymers. (3rd ed.). John Wiley & Sons, Inc., New York.
[24] Flory, P. J., & Erman, B. (1982). Theory of elasticity of polymer networks. 3. Macromolecules, 15, 800–806.
[25] Flory, P. J., & Rehner, J. J. (1943). Statistical mechanics of cross–linked polymer networks: I. rubberlike elasticity. The Journal of Chemical
Physics, 11, 512–520.
[26] de Gennes, P. G. (1971). Reptation of a polymer chain in the presence of fixed obstacles. The Journal of Chemical Physics, 55, 572–579.
[27] Gent, A. N. (1996). A new constitutive relation for rubber. Rubber Chemistry and Technology, 69, 59–61.
25
[28] Gent, A. N., & Thomas, A. (1958). Forms for the stored (strain) energy function for vulcanized rubber. Journal of Polymer Science, 28,
625–628.
[29] Green, M. S., & Tobolsky, A. V. (1946). A new approach to the theory of relaxing polymeric media. The Journal of Chemical Physics, 14,
80–92.
[30] Hart-Smith, L. J. (1966). Elasticity parameters for finite deformations of rubber-like materials. Zeitschrift für angewandte Mathematik und
Physik ZAMP, 17, 608–626.
[31] Haupt, P., & Lion, A. (2002). On finite linear viscoelasticity of incompressible isotropic materials. Acta Mechanica, 159, 87–124.
[32] Heinrich, G., & Kaliske, M. (1997). Theoretical and numerical formulation of a molecular based constitutive tube–model of rubber elasticity.
Computational and Theoretical Polymer Science, 7, 227–241.
[33] Holzapfel, G. A., & Simo, J. C. (1996). A new viscoelastic constitutive model for continuous media at finite thermomechanical changes.
International Journal of Solids and Structures, 33, 3019–3034.
[34] Hong, W. (2011). Modeling viscoelastic dielectrics. Journal of the Mechanics and Physics of Solids, 59, 637–650.
[35] ISO-37:2017(E) (2017). Rubber, vulcanized or thermoplastic – Determination of tensile stress-strain properties. Standard International
Organization for Standardization Geneva, CH.
[36] Kadapa, C., & Hossain, M. (2020). A linearized consistent mixed displacement-pressure formulation for hyperelasticity. Mechanics of
Advanced Materials and Structures, (pp. 1–18). doi:https://doi.org/10.1080/15376494.2020.1762952.
[37] Kaliske, M., & Heinrich, G. (1998). An extended tube–model for rubber elasticity: Statistical–mechanical theory and finite element imple-
mentation. Rubber Chemistry and Technology, 72, 602–632.
[38] Kaliske, M., & Rothert, H. (1997). Formulation and implementation of three–dimensional viscoelasticity at small and finite strains. Compu-
tational Mechanics, 19, 228–239.
[39] Kawabata, S., Matsuda, M., Tei, K., & Kawai, H. (1981). Experimental survey of the strain energy density function of isoprene rubber
vulcanizate. Macromolecules, 14, 154–162.
[40] Khajehsaeid, H., Arghavani, J., & Naghdabadi, R. (2013). A hyperelastic constitutive model for rubber-like materials. European Journal of
Mechanics-A/Solids, 38, 144–151.
[41] Khiêm, V. N., & Itskov, M. (2016). Analytical network-averaging of the tube model:: Rubber elasticity. Journal of the Mechanics and Physics
of Solids, 95, 254–269.
[42] Kilian, H.-G., Enderle, H. F., & Unseld, K. (1986). The use of the van der waals model to elucidate universal aspects of structure-property
relationships in simply extended dry and swollen rubbers. Colloid and Polymer Science, 264, 866–876.
[43] Koprowski-Theiss, N., Johlitz, M., & Diebels, S. (2011). Characterizing the time dependence of filled epdm. Rubber Chemistry and
Technology, 84, 147–165.
[44] Korba, A. G., & Barkey, M. E. (2017). New model for hyper-elastic materials behavior with an application on natural rubber. In ASME 2017
12th International Manufacturing Science and Engineering Conference collocated with the JSME/ASME 2017 6th International Conference
on Materials and Processing (pp. V002T03A020–V002T03A020). American Society of Mechanical Engineers.
[45] Kröner, E. (1960). Allgemeine Kontinuumstheorie der Versetzungen und Eigenspannungen. Archive for Rational Mechanics and Analysis,
4, 273–334.
[46] Kuhn, W. (1934). über die gestalt fadenförmiger moleküle in lösungen. Kolloid-Zeitschrift, 68, 2–15.
[47] Kuhn, W. (1936). Beziehungen zwischen molekülgröße, statistischer molekülgestalt und elastischen eigenschaften hochpolymerer stoffe.
Kolloid-Zeitschrift, 76, 258–271.
[48] Kuhn, W., & Grün, F. (1942). Beziehungen zwischen elastischen konstanten und dehnungsdoppelbrechung hochelastischer stoffe. Kolloid-
Zeitschrift, 101, 248–271.
[49] Lambert-Diani, J., & Rey, C. (1999). New phenomenological behavior laws for rubbers and thermoplastic elastomers. European Journal of
Mechanics-A/Solids, 18, 1027–1043.
[50] Lee, E. H. (1969). Elastic-plastic deformation at finite strains. ASME. Journal for Applied Mechanics, 36, 1–6.
[51] Linder, C., Tkachuk, M., & Miehe, C. (2011). A micromechanically motivated diffusion-based transient network model and its incorporation
into finite rubber viscoelasticity. Journal of the Mechanics and Physics of Solids, 59, 2134–2156.
[52] Lion, A. (1996). A constitutive model for carbon black filled rubber. Experimental investigations and mathematical representations. Contin-
uum Mechanics and Thermodynamics, 8, 153–169.
[53] Lion, A. (1997). On the large deformation behaviour of reinforced rubber at different temperatures. Journal of the Mechanics and Physics of
Solids, 45, 1805–1834.
[54] Lodge, A. S. (1956). A network theory of flow birefringence and stress in concentrated polymer solutions. Transactions of the Faraday
Society, 52, 120–130.
[55] Lopez-Pamies, O. (2010). A new i1-based hyperelastic model for rubber elastic materials. Comptes Rendus Mecanique, 338, 3–11.
[56] Lubliner, J. (1985). A model of rubber viscoelasticity. Mechanics Research Communications, 12, 93–99.
[57] Marckmann, G., & Verron, E. (2006). Comparison of hyperelastic models for rubber-like materials. Rubber Chemistry and Technology, 12,
835–858.
[58] Marsden, J. E., & Hughes, T. J. R. (1983). Mathematical Foundations of Elasticity. Prentice-Hall, Englewood Cliffs, New Jersey.
[59] Miehe, C., & Göktepe, S. (2005). A micro–macro approach to rubber–like materials. Part II: The micro–sphere model of finite rubber
viscoelasticity. Journal of the Mechanics and Physics of Solids, 53, 2231–2258.
[60] Miehe, C., Göktepe, S., & Lulei, F. (2004). A micro–macro approach to rubber–like materials. Part I: The non–affine micro–sphere model of
rubber elasticity. Journal of the Mechanics and Physics of Solids, 52, 2617–2660.
[61] Miehe, C., & Keck, J. (2000). Superimposed finite elastic–viscoelastic–plastoelastic stress response with damage in filled rubbery polymers.
Experiments, modelling and algorithmic implementation. Journal of the Mechanics and Physics of Solids, 48, 323–365.
[62] Mooney, M. (1940). A theory of large elastic deformation. Journal of Applied physics, 11, 582–592.
[63] Ogden, R. (1972). Large deformation isotropic elasticity: on the correlation of theory and experiment for incompressible rubberlike solids.
Proceedings of the Royal Society of London A, 326, 565–584.
26
[64] Phan-Thien, N. (1978). A nonlinear network viscoelastic model. Journal of Rheology, 22, 259–283.
[65] Phan-Thien, P. (2002). Understanding Viscoelasticity. (1st ed.). Springer-Verlag Berlin-Heidelberg.
[66] Pucci, E., & Saccomandi, G. (2002). A note on the gent model for rubber-like materials. Rubber chemistry and technology, 75, 839–852.
[67] Raghunath, R., Juhre, D., & Klüppel, M. (2016). A physically motivated model for filled elastomers including strain rate and amplitude
dependency in finite viscoelasticity. International Journal of Plasticity, 78, 223–241.
[68] Reese, S., & Govindjee, S. (1998). A theory of finite viscoelasticity and numerical aspects. International Journal of Solids and Structures,
35, 3455–3482.
[69] Rivlin, R. S., & Saunders, D. W. (1951). Large elastic deformations of isotropic materials vii. experiments on the deformation of rubber.
Proceedings of the Royal Society of London A, 243, 251–288.
[70] Saxena, P., Hossain, M., & Steinmann, P. (2013). A theory of finite deformation magneto-viscoelasticity. International Journal of Solids and
Structures, 50, 3886 – 3897.
[71] Shariff, M. H. B. M. (2000). Strain energy function for filled and unfilled rubberlike material. Rubber chemistry and technology, 73, 1–18.
[72] Simo, J. C. (1987). On a fully three–dimensional finite–strain viscoelastic damage model: Formulation and computational aspects. Computer
Methods in Applied Mechanics and Engineering, 60, 153–173.
[73] Simo, J. C. (1992). Algorithms for static and dynamic multiplicative plasticity that preserve the classical return mapping schemes of the
infinitesimal theory. Computer Methods in Applied Mechanics and Engineering, 99, 61–112.
[74] Simo, J. C., & Miehe, C. (1992). Associative coupled thermoplasticity at finite strains: Formulation, numerical analysis and implementation.
ASME Journal of Applied Mechanics, 98, 41–104.
[75] Steinmann, P., Hossain, M., & Possart, G. (2012). Hyperelastic models for rubber-like materials: consistent tangent operators and suitability
for treloar’s data. Archive of Applied Mechanics, 82, 1183–1217.
[76] Swanson, S. R. (1985). A constitutive model for high elongation elastic materials. Journal of engineering materials and technology, 107,
110–114.
[77] Tanaka, F., & Edwards, S. F. (1992). Viscoelastic properties of physically crosslinked networks. I: Non–linear stationary viscoelasticity.
Journal of Non–Newtonian Fluid Mechanics, 43, 247–271.
[78] Tanaka, F., & Edwards, S. F. (1992). Viscoelastic properties of physically crosslinked networks. II: Dynamic mechanical moduli. Journal of
Non–Newtonian Fluid Mechanics, 43, 289–309.
[79] Taylor, R. L. (2017). FEAP – A Finite Element Analysis Program, Version 8.5. University of California at Berkeley, Berkeley, California.
[80] Treloar, L. R. G. (1943). The elasticity of a network of long-chain molecules-ii. Transactions of the Faraday Society, 39, 241–246.
[81] Treloar, L. R. G. (1944). Stress–strain data for vulcanised rubber under various types of deformation. Transactions of the Faraday Society,
40, 59–70.
[82] Treloar, L. R. G. (1975). The Physics of Rubber Elasticity. (3rd ed.). Clarendon Press, Oxford.
[83] Treloar, L. R. G., & Riding, G. (1979). A non–gaussian theory of rubber in biaxial strain. i.mechanical properties. Proceedings of the Royal
Society of London A, 369, 261–280.
[84] Truesdell, C., & Noll, W. (1965). The non–linear field theories of mechanics. In S. Flügge (Ed.), Encyclopedia of Physics. Springer–Verlag,
Berlin volume III/3.
[85] Valanis, K. C., & Landel, R. F. (1967). The strain-energy function of a hyperelastic material in terms of the extension ratios. Journal of
Applied Physics, 38, 2997–3002.
[86] Wall, F. T. (1942). Statistical thermodynamics of rubber. The Journal of Chemical Physics, 10, 132–.
[87] Wang, M. C., & Guth, E. (1952). Statistical theory of networks of non–gaussian flexible chains. The Journal of Chemical Physics, 20,
1144–1157.
[88] Weber, G., & Anand, L. (1990). Finite deformation constitutive equations and a time integration procedure for isotropic hyperelastic-
viscoplastic solids. Computer Methods in Applied Mechanics and Engineering, 79, 173–202.
[89] Yeoh, O. H. (1990). Characterization of elastic properties of carbon-black-filled rubber vulcanizates. Rubber chemistry and technology, 63,
792–805.
[90] Yeoh, O. H., & Fleming, P. D. (1997). A new attempt to reconcile the statistical and phenomenological theories of rubber elasticity. Journal
of Polymer Science Part B: Polymer Physics, 35, 1919–1931.
27