You are on page 1of 78

Journal Pre-proof

Constitutive-damage modeling and computational implementation for simulation of


elasto-viscoplastic-damage behavior of polymeric foams over a wide range of strain
rates and temperatures

Jeong-Ho Lee, Dong-Man Ryu, Chi-Seung Lee

PII: S0749-6419(19)30541-8
DOI: https://doi.org/10.1016/j.ijplas.2020.102712
Reference: INTPLA 102712

To appear in: International Journal of Plasticity

Received Date: 29 July 2019


Revised Date: 13 February 2020
Accepted Date: 13 February 2020

Please cite this article as: Lee, J.-H., Ryu, D.-M., Lee, C.-S., Constitutive-damage modeling and
computational implementation for simulation of elasto-viscoplastic-damage behavior of polymeric foams
over a wide range of strain rates and temperatures, International Journal of Plasticity (2020), doi: https://
doi.org/10.1016/j.ijplas.2020.102712.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.


Author Statement

Jeong-Ho Lee: Conceptualization, Methodology, Programming, Validation, Formal analysis,

Investigation, Resources, Data curation, Writing – Review and editing, Visualization. Dong-

Man Ryu: Validation, Formal analysis, Investigation, Resources, Data curation, Writing –

Review and editing. Chi-Seung Lee: Conceptualization, Methodology, Validation, Formal

analysis, Investigation, Resources, Data curation, Writing – Original draft, Writing – Review

and editing, Visualization, Supervision, Project administration, Funding acquisition.


Constitutive-damage modeling and computational implementation for

simulation of elasto-viscoplastic-damage behavior of polymeric foams

over a wide range of strain rates and temperatures

Jeong-Ho Lee1, Dong-Man Ryu2,3 and Chi-Seung Lee3,4,5,*


1
Department of Mechanical Engineering, Boston University, Boston MA 02215, USA
2
Department of Mechanical Engineering, Michigan State University, East Lansing MI

48824, USA
3
Biomedical Research Institute, Pusan National University Hospital, Busan 49241,

Republic of Korea
4
Department of Convergence Medicine, School of Medicine, Pusan National University,

Busan 49241, Republic of Korea


5
Department of Biomedical Engineering, School of Medicine, Pusan National University,

Busan 49241, Republic of Korea

*Corresponding author: (Tel) +82-51-240-6867 (Email) victorich@pusan.ac.kr

Abstract

The material behavior and damage of polymeric foams (e.g., polyurethane foam and

polystyrene foam) are extremely complex under compression, depending on strain

rates and temperatures. In the present study, the elasto-viscoplastic-damage behavior

of polymeric foams over a wide range of strain rates and temperature is evaluated and

predicted using the modified Gurson-Tvergaard-Needleman-Lemaitre (GTNL) model

and the modified Khan-Huang-Liang (KHL) model. A modified GTNL model is

1
introduced to describe phenomena of decreasing void volume fraction and elastic

modulus for polymeric foams under uniaxial compression. The modified KHL model is

proposed to evaluate the nonlinear strain rate- and temperature-dependent material

behavior of polymeric foams. The full derivation of the implicit integration algorithm of

the proposed modified GTNL-KHL model is introduced, and a user-defined material

(UMAT) subroutine for commercial finite element analysis code (i.e., ABAQUS) is

established. Using the UMAT, the damage characteristics (e.g., void volume fraction

and elastic modulus changes) and the nonlinear material behavior (e.g., linear elastic,

softening/plateau and densification) of polymeric foams under uniaxial compression

with various strain rates and temperatures are successfully simulated.

Keywords: Polymeric foams; Constitutive behavior; Elastic-viscoplastic material; Finite

elements; Numerical algorithms.

2
Nomenclature

1. Characters

The specific meaning of some important characters is listed below.

Exponential function of the equivalent plastic strain

, , , , , Material parameters in the original KHL model

, , , , , , , , Material parameters in the modified KHL model

Component for the inverse of the consistent


tangent operator

Damage variable

Maximum plastic strain rate

Fourth-order Infinitesimal elasticity tensor

Fourth-order Infinitesimal elastoplastic consistent


tangent operator

Fourth-order Infinitesimal effective elasticity tensor

Elastic modulus

Elastic modulus at loading state

Elastic modulus at unloading state

Void volume fraction

Critical void volume fraction

! Void volume fraction at fracture

!
̅ Ultimate void volume fraction

#
$% Growth of existing voids

#
&' Nucleation of new voids

3
( Quantity of nucleated voids per unit volume

Initial void volume fraction

∗ Effective void volume fraction

* Shear modulus

+ Second-order identity tensor

+, Fourth-order deviatoric projection tensor

+- Fourth-order symmetric identity tensor

. /-0 Second principal invariant of the deviatoric stress


tensor

1 Bulk modulus

2 Equivalent plastic strain

3 Acceleration factor in void-growth term of


modified GTNL model

4 ,4 ,4 Material parameters in GTN model

5, 6, 3
Material parameters in the Lemaitre damage

model

- Deviatoric stress tensor

7( Standard deviation

8 Current temperature

89 Melting temperature

8: Maximum temperature

8% Reference temperature

; Damage energy release rate

4
<# Plastic multiplier

=# Infinitesimal equivalent strain rate

= > Strain for the stress-drop

=( Mean nucleating strain

?@ Infinitesimal elastic strain tensor

?@, Deviatoric strain tensor

?ABCDE
, Deviatoric trial strain tensor

?# Plastic strain rate tensor

?@F Volumetric strain tensor

GH Stress ratio for the void volume fraction, namely,


the ratio of the stresses between the porous state
with initial void volume fraction

I Poisson’s ratio

J̅ Reference density

K von Mises equivalent stress

L Cauchy stress tensor

KM Hydrostatic stress

LN Hydrostatic stress tensor

KO Yield stress

KO,PQ& Yield stress at the dense state of polymeric foam

KO,QR Yield stress obtained from uniaxial compression


test of polymeric foam

KO,SQ: Temperature-dependent yield stress

T Yield function

5
U Helmholtz free-energy per unit mass

2. Mathematical Operations

The meaning of operations is listed below.

V
/∙0 Derivative of /∙0 with respect to W
VW

Y∙Z Single contraction of vectors and tensors

Y∶Z Double contraction of tensors

Y⨂Z Tensor product of vectors and tensors

6
1. Introduction

Polymeric foams, such as polyurethane foam (PUF) and polystyrene foam (PSF), can be

easily observed in nature. They are widely adopted in the industrial field, because they

have excellent and unique material characteristics. They are lightweight, exhibit robust

specific stiffness and strength, possess superior energy absorption capacity, etc. The

polymeric foam is a representative porous material and has remarkably unique

properties as it contains voids itself. First, it has orthotropic, transversely isotropic and

isotropic properties, depending on the foaming process, additives such as glass fiber

during fabrication, etc. Second, owing to the voids, the material behavior in uniaxial

tension and compression significantly differs from each other. For example, under a

tensile load, it exhibits elastic behavior, and a sudden brittle fracture can occur after a

certain strain. However, under a compressive load, the elastic and inelastic behaviors

appear sequentially, and the softening/plateau, including a stress-drop and

densification, are observed with the inelastic behavior. Additionally, as the compressive

strain rate increases, the amount of stress-drop increases. Fig. 1 shows the schematic

uniaxial compressive stress-strain curve of the PUF and the deformed test specimens at

each strain. As shown in this figure, the linear elastic region (I), softening/plateau

region (II), and densification region (III) can be clearly identified. Third, under tensile

load, the void volume fraction increases, and the elastic modulus of the material matrix

decreases, while the void volume fraction and the elastic modulus decrease

simultaneously under the compressive load.

Experimental studies have been carried out by a lot of researchers to


7
investigate the material nonlinearity, damage, failure, and fracture phenomena of

polymer materials such as polymeric foams and amorphous/semi-crystalline polymers

(Ayoub et al., 2011; Balieu et al., 2013; Guessasma and Nouri, 2016; Hachour et al.,

2014; Khan and Farrokh, 2006; Lee et al., 2019; Marsavina and Constantinescu, 2015;

Park et al., 2016; Ponçot et al., 2013; Shafiq et al., 2015; Srivastava et al., 2010; Whisler

and Kim, 2015; Yao et al., 2018; Zaïri et al., 2008; Zaïri et al., 2011; Zaïri et al., 2005,

2007; Zhang et al., 1998; Zhang et al., 1997; Zhou et al., 2019). In their studies, various

(thermo-)mechanical tests of polymer materials, such as uniaxial/multiaxial/hydrostatic

tension/compression, shear, creep, torsion tests under low-to-high temperature and

static-to-dynamic strain rates were conducted. Variations of material characteristics,

such as nonlinear stress-strain relationship, and yield, failure and fracture phenomena,

were investigated with respect to temperature, strain rate, and stress condition.

Among others, several researchers have carried out the material tests and have

investigated the mechanical behaviors of various types of polymeric foams. Marsavina

and Constantinescu (Marsavina and Constantinescu, 2015) investigated changes of

material properties according to polymeric foam density, microstructural morphology

(open or closed cell), and foaming direction (rise or transverse) over a wide range of

tensile/compressive loading speeds and temperatures. In addition, they carried out the

fracture toughness test under static and dynamic loadings, and they identified that the

important parameters influencing the fracture toughness are foam density, orientation

and dimension of cell, loading direction and loading speed. In their study, moreover,

the damage initiation and growth of polymeric foams during loading were

quantitatively evaluated using digital image correlation (DIC) technique. Guessasma

8
and Nouri (Guessasma and Nouri, 2016) investigated the microstructural changes as

well as material nonlinearity of bio-polymeric foams under excessive uniaxial

compression using universal testing machine with X-ray micro-tomography imaging

equipment. They monitored all of the microstructural changes inside the foam from

before deformation to densification using this facility. In addition, they found that the

cell connectivity, morphology and size distributions are related to structural anisotropy

generated by compressive loading. Shafiq et al. (Shafiq et al., 2015) investigated the

yield surface of transversely isotropic Divinycell H-100 PVC foam under a multitude of

uniaxial, biaxial and triaxial strain paths using a new custom-built multi-axial testing

apparatus. In this context, the experimental results reveal that yielding in foams

exhibits not only a quadratic pressure dependence, which is widely recognized in

literature, but also a significant linear pressure dependence, which has been largely

overlooked in previous studies. A new energy-based yield criterion developed for

transversely isotropic foams is also validated using the experimental yield data. Whisler

and Kim (Whisler and Kim, 2015) investigated the mechanical characteristics of

polyurethane foams from quasi-static to dynamic compressive loads (2,600/s) at high

strain levels (>80%) using both universal testing machine and modified Hopkinson bar

facility. In their study, the experimental data indicated that the polyurethane foam is

strain rate sensitive

On the other hand, theoretical and numerical studies based on continuum and

computational mechanical approaches have been conducted by numerous scientists to

evaluate and predict the aforementioned mechanical features of polymer materials

such as amorphous/semi-crystalline polymers (Ames et al., 2009; Anand et al., 2009;

9
Ayoub et al., 2011; Ayoub et al., 2010; Ayoub et al., 2014; Baghani et al., 2012; Balieu

et al., 2013; Bouvard et al., 2013; Frank and Brockman, 2001; Hachour et al., 2014;

Khan et al., 2006; Khan and Yeakle, 2011; Launay et al., 2011; Li and Buckley, 2010;

Ponçot et al., 2013; Regrain et al., 2009; Shojaei and Li, 2013; Srivastava et al., 2010;

Voyiadjis et al., 2011, 2012; Zaïri et al., 2008; Zaïri et al., 2011; Zaïri et al., 2005, 2007),

and polymeric foams or cellular polymers (af Segerstad et al., 2008; Alkhader and Vural,

2010; Ayyagari and Vural, 2015, 2016; Cayzac et al., 2013; Chen and Ghosh, 2012; Chen

et al., 2018; Ehlers and Markert, 2003; Fahlbusch et al., 2016; Kim et al., 2017; Lee et

al., 2011; Lee et al., 2015; Lee and Lee, 2014a, b; Lee et al., 2018; Lee et al., 2016;

Machado et al., 2011; Rodas et al., 2016; Su et al., 2016; Wang and Pan, 2006; Yang and

Li, 2016). In these studies, temperature- and strain rate-dependent constitutive models

were proposed to describe the nonlinear material behavior of polymeric materials with

initial/growth voids. Additionally, the damage-coupled constitutive model was

introduced to quantitatively evaluate the variation of void volume fraction and damage

within material, and the explicit/implicit formulation procedure and computational

implementation method for the introduced constitutive-damage model was addressed.

Among these, Anand et al. (Anand et al., 2009) have presented a detailed

continuum mechanical development of a thermo-mechanically coupled elasto-

viscoplastic theory to model the strain rate and temperature dependent large-

deformation response of amorphous polymeric materials. The model was developed

based on the fully-thermomechanically coupled theory, and the rapid increase in

stresses at large deformation as the polymer chains start to lock-up was considered. In

order to allow for important energy storage mechanisms due to plastic deformation,

10
the development of an internal back-stress, and to account for Bauschinger-like

phenomena on unloading and reverse loading, they have introduced a symmetric and

unimodular tensor field.

Ames et al. (Ames et al., 2009) have conducted the large-strain compression

experiments on three representative amorphous polymeric materials such as poly

methyl methacrylate (PMMA), polycarbonate (PC), and a cyclo olefin polymer (Zeonex-

690R), in a temperature range spanning room temperature to slightly below the glass

transition temperature of each material, in a strain rate range of 10-4/s to 10-1/s, and

compressive true strains exceeding 100%. In addition, they have evaluated/predicted

the thermo-mechanically coupled strain rate and temperature dependent large

deformation mechanical response of three amorphous polymers through the

constitutive model proposed by Anand et al. (Anand et al., 2009). In their study, the

following major intrinsic features of the macroscopic stress-strain response of

amorphous polymers were successfully evaluated/predicted: (a) the strain rate and

temperature dependent yield strength; (b) the transient yield-peak and strain-

softening which occurs due to deformation-induced disordering; (c) the subsequent

rapid strain-hardening due to alignment of the polymer chains at large strains; (d) the

unloading response at large strains; and (e) the temperature rise due to plastic-

dissipation and the limited time for heat-conduction for the compression experiments

performed at strain rates ≳0.01/s.

Ayoub et al. (Ayoub et al., 2010) have carried out the mechanical test to

investigate the nonlinear time-dependent response of a high density polyethylene

(HDPE) at room temperature under constant true strain rates and stress relaxation

11
conditions. Moreover, they have proposed a physically-based inelastic model written

under finite strain formation to describe the mechanical behavior of a HDPE. In their

model, the inelastic mechanisms involved two parallel elements: a visco-hyperelastic

network resistance acting in parallel with a viscoelastic-viscoplastic intermolecular

resistance where the amorphous and crystalline phases were explicitly taken into

consideration. The semicrystalline polymer was considered as a two-phase composite,

and the influence of the crystallinity on the loading and unloading behavior was

investigated. They have applied the proposed constitutive model to a polyethylene (PE)

(Ayoub et al., 2011). Furthermore, in order to describe the large deformation time-

dependent mechanical response of rubber-like materials under cyclic loading

characterized by stress-softening, hysteresis and permanent set, they have developed a

constitutive model integrating the physics of polymer chains and their alteration

(Ayoub et al., 2014).

Balieu et al. (Balieu et al., 2013) have proposed a phenomenological finite

strain non-associated elasto-viscoplastic model coupled with damage to simulate the

behavior of a 20% mineral-filled semicrystalline polymer for a large strain rate range

under several loading conditions. In their model, the direct relation between the

logarithmic rate of the Eulerian Hencky strain tensor (work-conjugate of the Cauchy

stress) with the additively decomposition of the stretching tensor in an elastic and a

viscoplastic part was adopted.

Bouvard et al. (Bouvard et al., 2013) have presented a complete theoretical

accounting for the thermomechanical coupling within a viscoplastic model to predict

the time, temperature, and stress state dependent mechanical behavior of amorphous

12
glassy polymers. The proposed model account for (i) the material strain softening

induced by the polymer chain slippage; (ii) the material strain hardening at large strains

induced by chain stretching between entanglement points; (iii) the time, temperature,

and stress state dependence exhibited by polymers under deformation.

Zairi et al. (Zaïri et al., 2008; Zaïri et al., 2005, 2007) have proposed a modified

viscoplastic model and a damage-coupled viscoplastic model which were originally

developed by Bonder and Partom (Bodner and Partom, 1975), Frank and Brockman

(Frank and Brockman, 2001), and Gurson, Tvergaard and Needleman (Gurson, 1977;

Tvergaard, 1981; Tvergaard and Needleman, 1984) to describe the mechanical

response including strain hardening/softening and void evolution of the rubber-

toughened (or rubber-modified) polymers under various strain rate and isothermal

conditions in the regime of the infinitesimal strain theory. In order to describe the

strain hardening and softening simultaneously, they have modified the original Frank-

Brockman model by introducing the strain hardening and softening terms. In addition,

they have combined this model with the Gruson-Tvergaard-Needleman model to

calculate the void evolution inside the material under the arbitrary strain rates. After

that, Zairi et al. (Zaïri et al., 2011) have proposed a physically-based model for large

deformation stress-strain response and anisotropic damage in rubber-toughened glassy

polymers in the regime of the finite strain theory. The main features leading to a

microstructural evolution (regarding cavitation, void aspect ratio, matrix plastic

anisotropy and rubbery phase deformation) in rubber-toughened glassy polymers were

introduced in their constitutive model. The constitutive response of the glassy polymer

matrix was modeled using the hyperelastic-viscoplastic model of Boyce et al. (Boyce

13
and Arruda, 2000; Boyce et al., 1988). The deformation mechanisms of the matrix

material were accounted for by two resistances: an elastic-viscoplastic isotropic

intermolecular resistance acting in parallel with a visco-hyperelastic anisotropic

network resistance, each resistance were modified to account for damage effects by

void growth with a variation of the void aspect ratio.

Meanwhile, some of these scientists have focused on developing novel

material models such as yield functions, constitutive models and damage equations to

precisely predict the damage initiation/growth as well as the nonlinear material

behavior of polymeric foams. Ehlers and Markert (Ehlers and Markert, 2003) proposed

an efficient, large strain, multiphasic continuum model based on the theory of porous

media accounting for the porous cell structure, the moving and interacting pore-fluid

and the intrinsic viscoelastic behavior of the polymer solid matrix. Wang and Pan

(Wang and Pan, 2006) investigated the yield behavior of a closed-cell polymeric foams

under multiaxial loading, and developed a phenomenological yield function to

characterize the initial yield behavior under a full range of loading conditions. Af

Segerstad and Toll (af Segerstad et al., 2008) developed a new generic model for an

open-cell cellular solid under finite deformation with a random microstructure. The

new model they proposed is not a multi strut model but a single strut model, which

not only saves computation time but can also simplify the modeling of dissipative

mechanisms, plasticity, viscoplasticity and damage. Alkhader and Vural (Alkhader and

Vural, 2010) introduced the initial and subsequent yield surfaces for an anisotropic and

pressure-dependent 2D stochastic cellular foams under biaxial loading. Machado et al.

(Machado et al., 2011) established a single-surface yield criterion-based rate-

14
independent elasto-plastic foam constitutive model for considering the relative density

dependence effect on polymeric foams subjected to large deformations. Ayyagari and

Vural (Ayyagari and Vural, 2015, 2016) developed a new yield criterion for transversely

isotropic solid foams. Its derivation is based on the central hypothesis that the yielding

in foams is driven by the total strain energy density rather than a completely

phenomenological approach. Fahlbusch et al. (Fahlbusch et al., 2016) proposed an

analytical and a numerical model for calculation of the effective failure surface of

closed-cell foams. Su et al. (Su et al., 2016) established the elongated

tetrakaidecahedron and Voronoi models in the rise direction to investigate the

multiaxial creep behavior of approximately transversely isotropic open-cell foams

based on the geometric stretch method. Chen et al. (Chen et al., 2018) developed a

new constitutive model to estimate the effective mechanical behaviors of the

incompressible neo-Hookean materials with randomly distributed spherical voids

under finite deformations.

Despite these efforts, it has been difficult to comprehensively study the

variation of material nonlinearity, damage, and void volume fraction of polymeric

foams over a wide range of strain rates and temperatures under compression. In

particular, it is exceedingly difficult to find literature that covers the development, full

derivation, and computational implementation of the constitutive-damage model to

date.

In this study, a unified constitutive-damage model for the simulation of the

nonlinear elasto-viscoplastic material behaviors and material degradation induced by

damage and void volume fraction is proposed. To describe the strain rate- and

15
temperature-dependent material nonlinearity of polymeric foams, the Khan-Huang-

Liang (KHL) model is introduced. Because it is difficult to simulate the

softening/plateau and densification phenomena after the elastic domain of foam

materials, the original KHL model is modified. To calculate the degradation of void

volume fraction according to accumulation of compressive strain, the Gurson-

Tvergaard-Needleman (GTN) model is introduced and modified. Additionally, the

Lemaitre’s damage evolution equation is implemented with the modified GTN model to

evaluate the increase of damage (or decrease of elastic modulus) caused by the

increase of compressive strain. It is called the modified GTNL-KHL model in this study.

The full derivation of the implicit integration algorithm of the two proposed

models is addressed, and the discretized equations are implemented into a ABAQUS

user-defined material (UMAT) subroutine for simulation of strain rate- and

temperature-dependent material nonlinearity, void volume fraction and damage of

polymeric foams under uniaxial compression. Using the UMAT, a series of uniaxial

compression tests of PUF and PSF under various experimental conditions are simulated,

and the simulation results, such as compressive stress-strain behavior, void volume

fraction decrease and elastic modulus degradation, are compared to the experimental

results to validate the proposed method.

2. Constitutive Modeling

To describe the damage initiation and growth, and the strain rate- and temperature-

dependent material nonlinearities of polymeric foams, modified GTNL and KHL models

are proposed. The degradation of void volume fraction and elastic modulus under
16
uniaxial compressive loads is described using the modified GTNL model, described later,

and the strain rate- and temperature-dependent nonlinear material behavior is

represented using the modified KHL model. The specific procedure for material-

modeling is addressed below based on the infinitesimal strain theory. On the other

hand, it is postulated that the polymeric foam is an isotropic material in the present

study.

2.1 Modified GTNL Model

Gurson (Gurson, 1977), Tvergaard (Tvergaard, 1981) and Tvergaard and Needleman

(Tvergaard and Needleman, 1984) developed a constitutive model to describe the

mechanism of nucleation, growth and coalescence of internal voids in porous metals


using the effective void volume fraction, , as follows (de Souza Neto et al., 2011):

KO 34 KM
T = . /-0 − a1 + 4 ∗
− 24 ∗
cosh i jk,
3 2KO
(1)

where T is the yield function, - is the deviatoric stress tensor, . /-0 is the second

principal invariant of the deviatoric stress tensor, KO and KM represent the yield and

hydrostatic stresses, and 4 , 4 , and 4 are the material parameters. In particular,

Tvergaard and Needleman (Tvergaard and Needleman, 1984) introduced the following

function for the effective void volume fraction to account for the effects of rapid void

coalescence at failure (Li and Wang, 2008):

17
o ≤
m ̅ −
= + / − 0 < < !,
∗ !
n ! −
m ̅
(2)

l ! ! ≤

where , , ! , and ̅ s= s4 + t4 − 4 uv4 u are the void volume fraction, the


!

critical void volume fraction, the void volume fraction at fracture, and the ultimate void

volume fraction, respectively. The effective void volume fraction is obtained from both

nucleation and growth mechanism if the void volume fraction is less than the critical

value, . The coalescence mechanism becomes active when the void volume fraction

is higher than the critical value, (Malcher et al., 2014). Additionally, The rate of

change of voids in terms of growth and nucleation relying on the material behavior is

calculated using the following equations (Lee et al., 2018; Mkaddem et al., 2004):

# = $%
# + &'
# , (3)

where #
$% and &'
# represent the growth of existing voids and nucleation of new

voids, respectively. Growth of existing voids is based on the law of conservation of

mass and is expressed in terms of voids fraction:

# = /1 − 0?# : +,
$% (4)

where ?# is the infinitesimal plastic strain rate tensor, and + is the second-order

identity tensor. The nucleation of voids is given by a strain-controlled relationship:


18
#
&' = 2# , (5)

where is an exponential function of the infinitesimal equivalent plastic strain, 2.

The volume fraction of the nucleated voids is as given in the following expression:

1 2 − =(
= exp a− } ~ k,
(
7( √2y 2 7(
(6)

where ( is the quantity of nucleated voids per unit volume, 7( is the standard

deviation, and =( is the mean nucleating strain.

Meanwhile, Lemaitre and his colleagues introduced a constitutive equation for

elastoplasticity coupled with damage (Lemaitre, 1984, 1985a, b, 2012; Lemaitre and

Chaboche, 1994; Lemaitre and Desmorat, 2005). Based on the concept of effective

stress and the hypothesis of strain equivalence, Lemaitre’s model includes the

evolution of internal damage and nonlinear isotropic and kinematic hardening in the

description of the behavior of ductile metals (de Souza Neto et al., 2011).

The elasticity law and thermodynamic force conjugate to the damage variable

are given by following equations (Lemaitre, 1985a, b, 2012; Lemaitre and Chaboche,

1994):

VU
L = J̅ = /1 − 0 : ?@ ,
V?@
(7)

19
VU 1 . /-0 KM
; = J̅ = − ?@ : : ?@ = − i + j,
V 2 2*/1 − 0 21/1 − 0
(8)

where L is the Cauchy stress tensor, J̅ is the reference density (mass per unit

volume in the reference configuration), U is the Helmholtz free-energy per unit mass,

is the damage variable, is the fourth-order infinitesimal elasticity tensor, ?@ is

the infinitesimal elastic strain tensor, ; is the damage energy release rate that

corresponds to the variation of internal energy density caused by damage growth at

constant stress, and * and 1 are the shear and bulk moduli, respectively.

Based on the hypothesis of strain equivalence, the fourth-order infinitesimal

effective elasticity tensor, , is expressed as a coupling of elasticity and damage, as

follows:

= /1 − 0 @
. (9)

Additionally, damage growth is calculated using the plastic multiplier and damage

energy-release rate, as follows:

<# ; 9
# =} ~ }− ~ ,
1− 5
(10)

where <# is the plastic multiplier, and 5 and 6 are the material parameters.

In the GTN model, the evolution of the damage variable is not directly

associated with a dissipative mechanism. The damage variable, , in this case, is the

20
void volume fraction (i.e., the local fraction of volume occupied by voids), and its

evolution law follows as a direct consequence of mass conservation. On the other hand,

in Lemaitre’s update to the model, the calculated damage variable affects the elastic

behavior. The elasticity tensor is a function of the damage variable. Hence, to describe

the decreasing features of void volume fraction and the elastic modulus for PUF under

compression, we apply the modified GTN-Lemaitre (GTNL) model, as follows (Lee et al.,

2018):

. /-0 KO 34 KM
T= − a1 + 4 ∗
− 24 ∗
cosh i jk,
/1 − 0€• 3 2KO
(11)

= /1 − 0€• @
, (12)

where 3 is a material parameter that controls the effect of damage on the level of

degradation of the elasticity tensor. According to this effect, the effective elasticity

tensor in Eq. (9) is modified as Eq. (12). Additionally, the effect of the effective void

volume fraction is ignored in this study, because the compression gives rise to only

negative growth of voids without nucleation for PUF. = ∗


was adopted during the

numerical implementation of the modified GTNL model for describing the nonlinear

compressive behavior of PUF.

In the proposed modified GTNL model, the increment of the damage variable

is calculated using Eqs. (8)-(10), without changes for the Lemaitre model. An

acceleration factor, 3, is applied to the void-growth term, as addressed in Eq. (13), for

calculating the increment of the void volume fraction, because it is insufficient to

21
describe the decrease of void volume fraction under compression using only the

growth term (Lee et al., 2016):

# = 3/1 − 0?# ‚ : +.
$% (13)

2.2 Modified KHL model

Khan et al. (Khan and Huang, 1992; Khan and Liang, 1999; Khan et al., 2004) proposed

a constitutive model to calculate the dependence of material behavior on strain rate

and temperature with experimental results as follows:

ln =# =# ˆ… 89 − 8 :…
&…†
K=a + i1 − j 2 &…‡
k} ~ } ~ ,
ln =# 89 − 8%
(14)

where K is the von Mises equivalent stress, =# is the infinitesimal equivalent strain

rates, , , , , and are the material parameters, 89 , 8 and 8% are

the melting, current and reference temperatures, respectively, =# is the reference

strain rate, i.e., =# =1/s, and is the maximum plastic strain rate chosen to be 106/s

(Bodner and Partom, 1975; Bodner, 2001; Khan and Huang, 1992; Liang and Khan,

1999).

In a previous study, to address the strain rate- and temperature-dependent

behavior of PUF, the modified KHL model was proposed as follows (Lee et al., 2018):

22
ln=# =# =# /1 − 8 ∗ 0:†
&†
K = ‰a + i1 − j 2&‡ k exp Š ln } ~‹ + exp Š ln } ~‹Œ ,
ln =# =# GH
(15)

8 − 8%
8∗ = ,
8: − 8%
(16)

where , , , , , , , and are the material parameters, 8: is the

maximum temperatures, at which PUF is adopted for industrial structures. In Eq. (15),

, , and define the trend of yield stress with respect to the strain rate,

and determine the shape of stress-strain relationship, and and , control

the strain rate and temperature sensitivity. GH represents the ratio of the stresses

between the porous state with initial void volume fraction, , and the dense state

without initial void volume fraction.

In the modified GTNL model, the material was assumed to be a porous state

with initial void volume fraction, while in the original KHL model, it was postulated to

be a dense state ( = 0). In this study, the target material is porous state materials

such as polymeric foam, so it is inevitable to modify the original KHL model. Hence, this

obstacle was solved by introducing the material parameter GH in the original KHL

model as shown in Eq. (15). Under uniaxial compression, because the axial stress,

KŽR Ž9 , the equivalent stress, 4, and 3 × hydrostatic stress, 3KM , are equal, a value of

GH can be acquired by using a solution of Eq. (17):

GH + 24 cosh s u−1−4 = 0.
•• ‘•
(17)

This modified KHL model is implemented into the part of the modified GTNL model
23
dealing with the yield stress, KO , in Section 2.1 during the numerical implementation

to describe the strain-hardening of the PUF, which strongly depends on strain rate and

temperature.

3. Numerical Implementation

To adopt the introduced constitutive-damage model using the finite element method, a

full derivation for the implicit integration algorithm of the proposed model is

introduced. A computational algorithm and the related ABAQUS UMAT subroutine for

the polymeric foam is proposed.

3.1 Implicit Integration Algorithm

The full derivation for the implicit integration algorithm of the proposed model is

addressed in this section. Eq. (11) is rewritten as Eq. (18), and its associated key

equations, Eqs. (19)-(22), are addressed as follows:

. /-0 KO 34 KM /1 − 0€• . “-ABCDE ” KO 34 KM


’ = − a1 + 4 ∗
− 24 ∗
cosh i jk = − a1 + 4 ∗
− 24 ∗
cosh i jk,
/1 − 0€• 3 2KO •/1 − 0€• + 2*∆<— 3 2KO
(18)

34 KM
’ = KM − KM S% Ž9 + 1∆<4 4 ∗
KO sinh i j,
2KO
(19)

34 KM
’ = − − 3/1 − 0∆<4 4 ∗
KO sinh i j − ∆2,
&™
2KO
(20)

’ = 2 − 2&™ − ∆2, (21)

∆< ; 9
’š = − −3 } ~ }− ~ .
&™
1− 5
(22)

The main parameters of Eqs. (20)-(22), , ∆2, and ;, are written as follows:

24
1 2 − =(
›= exp a− } ~ k,
(
7( √2y 2 7(
(23)

2 2. /-ABCDE 0 1 34 KM
∆2 = ∆<œ • + a4 4 ∗K sinh i jk ž,
3 •/1 − 0€• + 2*∆<— 3 O
2KO
(24)

/1 − 0 €• ™ . “-ABCDE ” KM
; = −a + k.
2*•/1 − 0 • + 2*∆<—
€ 21/1 − 0
(25)

The full derivation of differentiating Eqs. (18)-(22) and Eqs. (24)-(25) by ∆<, KM , , 2,

and is listed in Appendix A.

The plastic strain rate can be obtained by differentiating the yield function, Eq.

(11), as follows:

VT - 1 34 KM
?# ‚ = <# = <# a + 4 4 ∗
KO sinh i j +k.
VL /1 − 0 €• 3 2KO
(26)

Based on the additive decomposition of strain into an infinitesimal strain tensor and

both elastic predictor and plastic corrector processes (or return-mapping algorithm),

the following equations can be derived:

?@ = ?@, + ?@F , (27)

∆<
?@, = ?ABCDE − ∆?, = ?ABCDE − -,

, ,
/1 − 0€•
(28)

34 KM
=ŸQ = =ŸS% Ž9 − ∆=Ÿ = =ŸS% Ž9 − ∆<4 4 ∗
KO sinh i j,
2KO
(29)

25
where ?@, and ?@F are the deviatoric and volumetric strain tensors, respectively. By

solving Eqs. (28)-(29) using - = 2*?@, and LN = 1=ŸQ +, the deviatoric and hydrostatic

stress tensors can be obtained as follows:

/1 − 0€•
-=a k -ABCDE ,
/1 − 0€• + 2*∆<
(30)

34 KM
LN = LN ABCDE − 1∆<4 4 ∗
KO sinh i j +.
2KO
(31)

Additionally, the equivalent (or accumulated) plastic strain rate in Eq. (5), the growth of

existing voids, Eq. (3), and damage growth, Eq. (10), can be rewritten as follows:

2 2 2. /-0 1 34 KM
2# = œ ?# ‚ : ?# ‚ = <# œ • + a4 4 ∗K sinh i jk ž,
3 3 /1 − 0 €• 3 O
2KO
(32)

34 KM
# = 3/1 − 0?# ‚F : + = 3/1 − 0<# 4 4 ∗
KO sinh i j +,
$%
2KO
(33)

<# 1 3. /-0 KM
9
# =3 } ~a i + jk .
1− 5 6*/1 − 0 21/1 − 0
(34)

The tangent operator for the present model is addressed as follows. The total

stress can be decomposed into deviatoric and volumetric parts:

26
2*/1 − 0€•
L = - + KM + = a k ?ABCDE + KM +.
/1 − 0€• + 2*∆< ,
(35)

Thereafter, the consistent tangent operator can be obtained by differentiating the

above equation, giving

2*/1 − 0€• 4* ∆<3 /1 − 0€• ™ 4* /1 − 0€•


dL = a k d?ABCDE −‰ Œ ?ABCDE d −‰ Œ ?ABCDE d∆<
/1 − 0€• + 2*∆< ,
•/1 − 0€• + 2*∆<— ,
•/1 − 0€• + 2*∆<— ,
(36)
+ dKM +,

dL 2*/1 − 0€•
= =a k + − +⨂/ 8 +, + 8 ++ 8 +, + 8 +, + š 8š +, 0
d? ABCDE /1 − 0€• + 2*∆< ,

4* ∆<3 /1 − 0€•™
+‰ Œ ?ABCDE ⨂/ 8 +, + 8 ++ 8 +, + 8 +, + šš 8š +, 0
•/1 − 0€• + 2*∆<— , š š š š (37)

4* /1 − 0€•
+‰ Œ ?ABCDE ⨂/ 8 +, + 8 ++ 8 +, + 8 +, + š 8š +, 0,
•/1 − 0€• + 2*∆<— ,

where is the fourth-order infinitesimal elastoplastic consistent tangent operator,

+, s≡ +- − +⨂+u is the fourth-order deviatoric projection tensor, +- is the fourth-

order symmetric identity tensor, and is the component for the inverse of the

consistent tangent operator. The differentiation of ’ to ’š with respect to ?ABCDE


,

(i.e., 8 to 8š ) are listed below:

V’ 4* /1 − 0€•
8 = =‰ Œ ?ABCDE ,
V?ABCDE
,
•/1 − 0€• + 2*∆<— , (38)

V’
8 = = −1,
V?ABCDE
,
(39)

27
V’ V∆2 8∆<*
8 = =− =− ?ABCDE ,
V?ABCDE V?ABCDE
,
, , 2 2. /-ABCDE0 1 ∗ K sinh }34 KM ~‹ Œ •/1 − 0€• + 2*∆<—
3œ ‰ + Š4 4
(40)
3 •/1 − 0€• + 2*∆<— 3 O 2KO

V’ V∆2 8∆<*
8 = =− =− ?ABCDE ,
V?ABCDE V?ABCDE ,
, , 2 2./-ABCDE 0 1 ∗ K sinh }34 KM ~‹
3œ ‰ + Š4 4 Œ •/1 − 0€• + 2*∆<—
(41)
3 •/1 − 0€• + 2*∆<— 3 O 2KO

; 9 ; 9
V’š 3 ∆<6 s− u V; 3 ∆<6 s− u 2*•/1 − 0€•™ —
8š = =− 5 = 5
€• + 2*∆<— ?, .
ABCDE
V?ABCDE ;/1 − 0 V?ABCDE ;/1 − 0 •/1 − 0
(42)
, ,

3.2 Computerization Procedure

The proposed constitutive-damage model is implemented into an ABAQUS UMAT

based on the introduced implicit integration algorithm. The UMAT allows researchers

to define material characteristics such as elasticity, inelasticity, and damage at each

integration point of an element (Lee et al., 2011; Lee et al., 2015; Lee and Lee, 2014b;

Lee et al., 2017; Lee et al., 2018; Lee et al., 2016). The computational algorithm for

numerical implementation of the proposed model is shown in Fig. 2. The simulation is

carried out in four stages. First, the initial analysis condition (i.e., initial material

properties, material parameters, geometry, element and node information) are

imported from the ABAQUS *.INP file. Second, the increment of stress and strain

tensor and solution-dependent state variables (SDV) are calculated in the UMAT. Third,

if the proposed yield function is satisfied, the decrease of void volume fraction and

degeneration of elastic modulus are calculated using Newton-Raphson method. Fourth,

the material Jacobian matrix, V¥L⁄V¥?, is calculated, and stress, strain, and SDVs are

updated at the end of the analysis iteration procedure. After UMAT calculation, the

tangent stiffness matrix of finite elements is calculated with the ABAQUS solver. The

increment and iteration are thereafter continued, and the analysis results are stored in

28
the ABAQUS *.ODB file until the analysis step is finished.

4. Numerical Simulation Procedure and Results

Based on the proposed constitutive-damage model, implicit integration algorithm, and

ABAQUS UMAT, the material nonlinearities of linear elastic, softening/plateau, and

densification and damage features, such as the decrease of void volume fraction and

degradation of elastic modulus of PUF and PSF under uniaxial compression with

various strain rates and temperatures, are simulated.

4.1 Experimental Overview in Literature

The nonlinear material behaviors and damage characteristics of polymeric foams were

experimentally investigated by a few researchers (Marsavina and Constantinescu, 2015;

Zhang et al., 1998; Zhang et al., 1997). In their studies, the material and mechanical

features were observed from three viewpoints: strain rate- and temperature-

dependent material nonlinearity; decrease of void volume fraction; and degradation of

elastic modulus. Each item is summarized below.

4.1.1 Strain rate- and temperature-dependent material nonlinearity

Marsavina and Constantinescu experimentally investigated the nonlinear strain rate-

and temperature-dependent material behavior of PUF (Marsavina and Constantinescu,

2015). In their research, a series of uniaxial compression tests of closed-cell PUF with a

density of 200 kg/m3 were carried out under diverse experimental conditions,

29
including a strain rate of 0.003-500/s and the temperature of 353, 296, and 213 K. The

nonlinear stress-strain behaviors were quantitatively investigated as shown in Fig. 3.

Zhang et al. carried out a series of experiments for several polymeric foams,

such as closed-cell PSF having a density of 16 kg/m3 and open-cell PUF with a density

of 68 kg/m3, evaluating uniaxial compression, hydrostatic compression, uniaxial

tension and simple shear tests under various strain rates and temperatures (Zhang et

al., 1998; Zhang et al., 1997). Additionally, the relationship between stress and strain,

according to strain rate (0.0016-88/s) and temperature (253-353 K) was analyzed.

In both studies, the experimental results represented that the elastic modulus,

yield stress, and the stress and strain ranges of the stress-drop increased as the strain

rate increased and the temperature decreased. Additionally, the three stages of

material nonlinearity for polymeric foams were identified in all experimental results. It

was confirmed that the stress-drop phenomenon was noticeable when the density of

the polymeric foams and the strain rate were high. In particular, it was confirmed that

the stress amplitude and the strain range of stress-drop increased as the strain rate

increased and the temperature decreased.

4.1.2 Decrease of void volume fraction

Fig. 4 shows the scanning electron microscopy (SEM) image of PUF at 500- and 300-μm

scales, respectively, before compression (Lee and Lee, 2014a). As shown in this figure,

there are many connected pores, and the space between each is filled with air.

Additionally, Fig. 5 represents the SEM images of PUF before and after compression

(Lee and Lee, 2014a). As shown, three PUF stages can be observed under uniaxial

30
compression: the pores of PUF endure the external loads, retaining a porous shape;

the pores cannot resist the external loads, and the foam is crushed gradually; and the

foam is squeezed excessively, acting as a rigid medium. These three stages are clearly

represented in the stress-strain behavior as a linear elastic region, a softening/plateau

region, and a densification region, respectively, as illustrated in Fig. 1.

As shown in Figs. 4 and 5, the void volume fraction of PUF decreased under

compression as the size of the internal voids (or cells) decreased with loading. In this

regard, Di Prima et al. investigated the change of cell sizes under compressive strain

for polymeric foam using micro-computed tomography (CT) scanning (Di Prima et al.,

2007). Fig. 6 shows the microstructural response of polymeric foam to deformation at

398 K. Fig. 6a shows the relationship among normalized average cell size, compressive

strain, and stress, whereas Fig. 6b represents the micro-CT snapshot of the foam

during deformation from 0 to 80% at 10% strain intervals. As shown in Fig. 6, the

average cell size drops less than 10% in the first 10% of strain. Subsequently, it rapidly

decreases until ~45% strain. After this point, the decrease in the average cell size

begins to slow. The snapshot of the foam structure shows that the buckling of the

main cells did not occur, on average, until the sample was compressed to 20% strain

and that the densification started near 40% strain. Each of the snapshots shown

correspond to a volume of 2×2×0.8 mm (Di Prima et al., 2007).

4.1.3 Degradation of elastic modulus

When the applied compressive load exceeds the material yield stress, the cell structure

of PUF collapses. Owing to this phenomenon, the PUF is permanently deformed and

31
does not recover its original shape. Accordingly, the collapse can be considered as

internal damage of the material. Generally, this damage reduces the elastic modulus of

the material (Lemaitre, 2012; Lemaitre and Chaboche, 1994; Lemaitre and Desmorat,

2005), and the degradation of elastic modulus for PUF, owing to the collapse of cell

structure, should be identified.

To identify the relationship between cyclic compressive load and elastic

modulus degradation, Hou et al. carried out quasi-static cyclic compression tests for

both neat and fiber-reinforced PUFs, investigating the changes of the elastic modulus

and stress-strain curves according to pre-strain amounts and numbers of cyclic loads

(Hou et al., 2014). Fig. 7 shows the engineering stress-strain curves of neat and fiber-

reinforced PUFs under loading (first cycle), unloading and reloading (second cycle),

respectively. In this experiment, three amounts of pre-strain (0.15, 0.50, and 0.80)

were adopted, and the degradation of elastic modulus between the compressive

loading and reloading cycles were measured as 0.521 (0.15 pre-strain), 0.242 (0.50 pre-

strain), and 0.139 (0.80 pre-strain), using the following equation:

Degradation = , (43)

where and are the elastic moduli at loading and reloading states,

respectively.

4.2 Finite Element Models and Analysis Conditions

32
To evaluate the changes of material characteristics of PUFs under compressive loads,

the strain rate- and temperature-dependent material behaviors, the decrease of void

volume fraction and the elastic modulus were computationally simulated based on the

proposed model with ABAQUS UMAT.

Fig. 8 shows the finite element (FE) model and its loading and boundary

conditions for a polymeric foam-test specimen. The element type was a reduced

integrated brick element (i.e., C3D8R element in ABAQUS). One element (1x1x1

element) or eight elements (2x2x2 elements) are sufficient to simulate the material

behavior using constitutive model. Hence, the number of elements and the element

sizes were set as 8 and 1x1x1 mm3, respectively. Before simulation, three surfaces of

the FE model were simply supported (i.e., UX = 0, UY = UZ ≠ 0 on the right surface, UY =

0, UX = UZ ≠ 0 on the leU surface, and UZ = 0, UX = UY ≠ 0 on the lower surface). The

upper surface was strain rate-controlled at the same rates of experiments in Section

4.1.1.

4.3 Identification of Material Parameters

In this study, most of the material parameters for the proposed constitutive-damage

model were identified by the deterministic method, namely, the analytical estimation

approach (Pyrz and Zairi, 2007). To do this, the experimental data obtained from

various literatures were used to determine the values of material parameters so that

the theoretical description corresponds to experiments. The material parameters that

cannot be obtained through the deterministic approach were determined by using the

values proposed in various literatures or by the evolutionary method to simulate the

33
experimental results well. The identification procedures of material parameters for the

modified KHL model and the modified GTNL model are schematically shown in Fig. 9.

4.3.1 Identification of Material Parameters for the modified KHL model

To determine the material constants of the modified KHL model for PUF and PSF, a

series of uniaxial compressive stress-strain curve data form Marsavina et al.

(Marsavina and Constantinescu, 2015) and Zhang et al. (Zhang et al., 1998; Zhang et al.,

1997) were adopted. Here, the reference temperature, 8% is room temperature

(8% =296K). Because the strain at yield stress is relatively small at room temperature, Eq.

(15) can be approximated as Eq. (44), and , , , and can be determined:

=# =#
KO = exp Š ln } ~‹ + exp Š ln } ~‹.
=# =#
(44)

When the strain rate is 1/s at room temperature, Eq. (15) can be reduced to Eq.

(45), and and can be identified. In the present study, the interpolation is

adopted to determine the stress-strain curve at strain rate of 1/s, because there is no

experimental data at this strain rate in the literature, as shown in Fig. 3:

K− − = 2&‡ . (45)

34
Meanwhile, Eq. (15) can be addressed as Eq. (46) at room temperature, and can

be determined within a range of calculated average values without the strain rate of

1/s:

=#
oK − exp ¬ ln s=# u- °
m − m
m exp ¬ ln s =# u- m
=# ln=#
= ln ±ln i1 − j.
n 2 ‡
&
ln
(46)

m m̄
m m
l ®

The values of and in Eq. (15) can be obtained via Eqs. (47)-(48):

KO,QR
KO,PQ& = ,
GH (47)

=# =# /1 − 8 ∗ 0:†
KO,SQ: = ² exp Š ln } ~‹ + exp Š ln } ~‹³ ,
=# =# GH
(48)

where KO,PQ& is the yield stress at the dense state of the polymeric foam, KO,QR is

the yield stress obtained from uniaxial compressive test of polymeric foam, and KO,SQ:

is the temperature-dependent yield stress. When the current temperature is not

equivalent to room temperature, Eq. (15) can be reduced to Eq. (48). First, after

was set to the range of [0, 1], the minimum positive integer among the positive

integers with an error sum of squares less than 0.5 was selected as . Subsequently,

values having the minimum-error sum of squares within the range of [0, 1] under each

strain rate were determined as the values of using the determined value of .

35
Namely, was a function of ln/=#⁄=# 0. Lastly, was expressed as a function of

temperature using 8 ∗∗ = 8⁄8% , and values of at temperature other than room

temperature were identified using the experimental results obtained under the highest

strain rate.

Table 1 shows the finally identified material parameters of the modified KHL

model for polymeric foams. In here, 8: was postulated as 473 K, which is the

maximum temperature of the polymeric foams in industrial fields.

4.3.2 Identification of Material Parameters for the modified GTNL model

To identify the material parameters of the modified GTNL model, both the cell-size

change curve and elastic-modulus degradation curve, as shown in Figs. 6 and 7, were

used.

As mentioned in Section 4.1.1, the densities of the polymeric foams in the

present study were 200 kg/m3 for closed-cell PUF, 68 kg/m3 for open-cell PUF, and 16

kg/m3 for closed-cell PSF. The initial values of void volume fraction, , of the closed-

cell PUF was calculated via interpolation of literature data (Linul et al., 2013). However,

owing to the lack of information about initial values of void volume fraction for open-

cell PUF and closed-cell PSF in the literature, they were postulated as 0.9. However, 4 ,

4 , and 4 control the effects of void volume fraction and hydrostatic stress on the

material behavior. The following set of parameters has been introduced for metallic

materials by many scientists: 4 =1.5, 4 =1.0, and 4 = 4 . In the present study, 4

was determined within the range of (0, 1.5], because the polymeric foams has a much

higher initial void than do the metallic materials.

36
Parameter 3 plays a role in defining the level of the elastic modulus

degradation according to the increase of the damage variable, and accordingly, a larger

3 causes greater degradation of material performance. In the present study, this

value was identified by repetitive comparison between the degradation values

calculated through Fig. 7 and the values of /1 − 0€• in the modified GTNL model

according to the compressive deformation of the polymeric foams.

Parameter 6 is a variable related to the rate of increase in material damage.

For example, the shape of the damage-strain curve is determined according to 6, and

the rate of decrease of elastic modulus is identified. Thus, this value was determined

within the rage of <1.0 in this study.

Parameters 3 and 5 control the extent of the increment in void volume

fraction and the damage variable, calculated by Eqs. (13)-(10), respectively.

Accordingly, these variables were addressed as functions of strain rate and

temperature, as shown in following equation:

=# =#
exp ¬ ln s=# u- + exp ¬ ln s=# u-

3 and 5 = ´ µ · /1 − 8 ∗ 0 ¸ .
5.221
(49)

To maintain the consistency when expressing the effects of strain and

temperature, two parts were adopted in this equation: a normalized term relevant to

the dependence on the strain rate of Eq. (44), and a term relevant to the dependence

on the temperature in the modified KHL model.

37
First, because the normalized term becomes unity when the strain rate is the

lowest, parameter ´ was determined based on the experimental results obtained

under this strain rate at room temperature. Subsequently, parameter ´ was decided

from experimental results obtained under the highest strain rate at room temperature

with the determined ´ . Lastly, parameter ´ was determined using the experimental

results obtained under the lowest strain rate at temperatures other than room

temperature.

Table 2 lists the finally determined material parameters of the modified GTNL

model for polymeric foams.

4.3.3 Identification of Material Properties

Except for the material parameters in the proposed constitutive-damage model, the

material properties adopted in simulation are elastic modulus, , Poisson’ s ratio, I,

and strain for the stress-drop, = >, which means that the stress-drop occurs from

yield strain to this strain owing to cell collapse without a change in void volume

fraction. In this respect, Poisson’s ratio for polymeric foams was identified as 0.3, and

the elastic modulus and strain for the stress-drop were determined through the

functions of strain rate and temperature, as addressed in Eq. (50), because these two

properties strongly depend on strain rate and temperature, as shown in Fig. 3. The

coefficients for Eq. (50) are listed in Tables 3 and 4. However, three coefficients of

= > for the experimental results of Zhang et al. (Zhang et al., 1998; Zhang et al., 1997)

cannot be obtained, as shown in Table 4, because the stress-drop phenomenon could

not be observed in their studies.

38
¹# º» Ž•
and = > =W s u +W .
¹# ‡
(50)

4.4 Prediction Results and Comparison with Experimental Results

Figs. 10-14 show the comparison results of the compressive stress-strain curve, the

normalized void volume fraction, and the degradation of the elastic modulus of the

closed-cell PUF, closed-PSF, and open-cell PUF. As shown in these figures, the material

nonlinearities (e.g., linear elastic, softening/plateau, and densification under various

temperatures and strain rates) were successfully predicted. In particular, the

phenomenon was successfully simulated for which the increase of elastic modulus,

yield stress, and the stress and strain ranges of stress-drop as the strain rate increased,

and the temperature decreased.

Although there is a small mismatch between the experiment and simulation in

terms of both void volume fraction-strain and degradation of elastic modulus-strain

curves for the closed-cell PUF, as illustrated in Figs. 10-12, the decreasing tendency of

void volume fraction and damage growth over a wide range of strain rates and

temperatures was reasonably predicted.

On the other hand, the compressive stress-strain, void volume fraction-strain

and degradation of elastic modulus-strain curves were successfully predicted for the

closed-cell PSF as shown in Fig. 13. At strains above 0.5, there is a slight mismatch

between the experiment and simulation in compressive stress-strain curve for the

open-cell PUF as shown in Fig. 14. Except for this, it could be confirmed that the

simulation results of the open-cell PUF were reasonably predicted.


39
5. Concluding Remarks

The constitutive-damage modeling and its computational implementation for

simulation of porous materials, such as polymeric foams, were proposed to

simultaneously describe the compressive stress-strain, void volume fraction-strain, and

damage-strain relationships over a wide range of strain rates and temperatures. A full

derivation of the implicit integration algorithm of the proposed model was also

introduced. Based on this computational algorithm, an ABAQUS UMAT was developed

for the application of polymeric foams. The proposed method was validated by

comparing a series of experimental results from literatures. The results and limitations

of this study are summarized below.

(1) A unified strain rate- and temperature-dependent elasto-viscoplastic-damage

model for polymeric foams was developed by modifying two existing material

models that were originally applied to metals (i.e., KHL and GTNL). To describe

the dependence of strain rate and temperature on material nonlinearity, the

modified KHL model was introduced and inserted into the yield function of the

modified GTNL model. To express the decrease phenomena of both void

volume fraction and damage growth (or decrease of elastic modulus)

simultaneously, the modified GTNL model was proposed. Through the unified

material model, it was confirmed that the micro-mechanical (i.e., void volume

fraction and damage changes) and the macro-mechanical (i.e., stress-strain

behavior) characteristics for closed- and open-cell PUFs and closed-cell PSFs

were effectively analyzed.


40
(2) The full derivation for the implicit integration algorithm of the proposed

material model and the ABAQUS UMAT-based finite element analysis

procedure for the polymeric foam were introduced. Based on the proposed

method, it might be possible to analyze the micro-mechanical and macro-

mechanical features of various polymeric foams over a wide range of strain

rates and temperatures.

(3) Because the experimental data for one type of polymeric foam, such as

compressive stress-strain, void volume fraction-strain, and damage-strain

under various strain rates and temperatures, could not be obtained from the

literature, the experimental data of a few types of foams reported in other

studies were adopted to determine the material parameters of the proposed

model. Accordingly, the authors plan to carry out a series of experiments for

one target material for identifying material parameters in a future study.

(4) The material parameters related to relative density were not taken into account

during the constitutive-damage modeling because there is no parameter for

relative density in the original GTN, Lemaitre and KHL models. However, there

are a large number of material parameters in the constitutive-damage model

proposed in this study, which depend on the type and density of the polymeric

foam. For example, in this study, three types of polymeric foams, namely,

closed-cell PUF (200kg/m3), closed-cell PSF (16kg/m3) and open-cell PUF

(68kg/m3) were selected, and their material parameters are totally different

each other as shown in Tables 1-4. In other words, although the parameters for

relative density were not directly introduced, it can be seen that various

41
material parameters differ depending on the density. Therefore, it is expected

that the relationship between material parameters and density can be

expressed by equation through the series simulation of material behavior for a

large number of polymeric foams with various densities. Furthermore, it is

expected that the material parameters of the polymeric foam with any density

under arbitrary strain rate and temperature can be identified rapidly and

effectively through this equation. If this equation is combined with the

constitutive-damage model proposed in this study, it is expected to be a great

tool for predicting the material behavior of polymeric foams with a wide range

of densities.

Acknowledgements

This work was supported by a National Research Foundation (NRF) of Korea grant

funded by the Korean Government (MSIT) (NRF-2019R1F1A1062037).

Appendix

(a) Derivatives for ’

∂’ −4*/1 − 0€• . /-ABCDE 0


= ,
∂∆< •/1 − 0€• + 2*∆<—

∂’ 34 KM
=4 4 ∗
KO sinh i j,
∂KM 2KO
(A1)

∂’ 2KO V ∗ 34 KM
= a4 cosh i j−4 ∗
k,
∂ 3 V 2KO

42
2KO Š1 + 4 ∗ − 24 ∗ cosh }34 KM
~‹
∂’ ∂KO 2KO 34 KM
=− µ +4 4 ∗ sinh i j KM ·,
∂2 ∂2 3 2KO

∂’ 3 /1 − 0€• ™ •/1 − 0€• − 2*∆<—. “-ABCDE ”


= ,
∂ •/1 − 0€• + 2*∆<—

(b) Derivatives for ’

∂’ 34 KM
= 14 4 ∗
KO sinh i j,
∂∆< 2KO

∂’ 34 34 KM
= a1 + 1∆<4 4 ∗
KO cosh i jk,
∂KM 2KO 2KO

∂’ 34 KM V ∗
= 1∆<4 4 ∗
KO sinh i j ,
∂ 2KO V
(A2)

∂’ ∂KO 34 KM 34 KM 34 KM
= 1∆<4 4 ∗
asinh i j − cosh i j k,
∂2 ∂2 2KO 2KO 2KO

∂’
= 0,

(c) Derivatives for ’

∂’ 34 KM V∆2
= −3/1 − 04 4 ∗
KO sinh i j− ,
∂∆< 2KO V∆<

∂’ 34 KM 34 V∆2
= −3/1 − 0∆<4 4 ∗
KO cosh i j − ,
∂KM 2KO 2KO VKM

∂’ 34 KM V ∗
V∆2
= 1 − 3∆<4 4 KO sinh i j Š/1 − 0 − ∗
‹− ,
∂ 2KO V V
(A3)

∂’ ∂KO 34 KM 34 KM 34 KM V V∆2
= −3/1 − 0∆<4 4 ∗ asinh i j − cosh i j k − } ∆2 + ~,
∂2 ∂2 2KO 2KO 2KO V2 V2

∂’ V∆2
=− ,
∂ V

(d) Derivatives for ’

43
∂’ V∆2
=− ,
∂∆< V∆<

∂’ V∆2
=− ,
∂KM VKM

∂’ V∆2
=− ,
∂ V
(A4)

∂’ V∆2
=1− ,
∂2 V2

∂’ V∆2
=− ,
∂ V

(e) Derivatives for ’š

; 9 V;
∂’š −3 s− u ∆<6
< V∆<
= ½1 + ¾,
∂∆< /1 − 0 ;

; 9 V;
∂’š −3 ∆<6 s− 5 u VKM
= ,
∂KM ;/1 − 0

∂’š
= 0,

(A5)

∂’š
= 0,
∂2

; 9
∂’š Á3 ∆< s− uÄ 1 6 V;
< Ã
= 1−À Š + ‹,
∂ À 1 − Ã 1 − ;V
¿ Â

(f) Derivatives for ∆2

V∆2
V∆<

2 2. /-ABCDE 0 1 34 KM
=œ • + a4 4 ∗K sinh i jk ž
3 •/1 − 0 €• + 2*∆<— 3 O
2KO (A6)

8∆<*. “-ABCDE ”
− ,
2 2. /-ABCDE 0 1 34 KM
3œ ‰ + Š4 4 ∗K sinh } ~‹ Œ •/1 − 0€• + 2*∆<—
3 •/1 − 0€• + 2*∆<— 3 O 2KO

44
34 KM 34 KM 34
2∆< Š4 4 ∗K sinh } ~‹ Š4 4 ∗K cosh } ~‹
V∆2 O 2KO O 2KO 2KO
= ,
VKM
2 2. /-ABCDE 0 1 ∗ K sinh }34 KM ~‹ Œ
9œ ‰ + Š4 4
3 •/1 − 0€• + 2*∆<— 3 O 2KO

34 KM V ∗
2∆< Š4 4 ∗
KO sinh } ~‹
V∆2 2KO V
= ,
V
2 2. /-ABCDE 0
1 34 KM
9œ ‰ + Š4 4 ∗K sinh } ~‹ Œ
3 •/1 − 0€• + 2*∆<— 3 O 2KO

34 KM VKO 34 KM 34 KM 34 KM
2∆< Š4 4 ∗K sinh } ~‹ 4 4 ∗ Šsinh } ~ − cosh } ~ ‹
V∆2 O 2KO V2 2KO 2KO 2KO
= ,
V2
2 2. /-ABCDE 0 1 34 KM
9œ ‰ + Š4 4 ∗K sinh } ~‹ Œ
3 •/1 − 0€• + 2*∆<— 3 O 2KO

V∆2 4∆<3 /1 − 0€•™ . “-ABCDE ”


= ,
V
2 2. /-ABCDE 0 1 ∗ K sinh }34 KM ~‹ Œ •/1 − 0€• + 2*∆<—
3œ ‰ + Š4 4
3 •/1 − 0€• + 2*∆<— 3 O 2KO

(g) Derivatives for ;

V; 2•/1 − 0€• ™ — . “-ABCDE ”


= ,
V∆< •/1 − 0€• + 2*∆<—

V; KM
=− ,
VKM 1/1 − 0

∂;
= 0,
∂ (A7)

∂;
= 0,
∂2

/3 − 10•/1 − 0€• ™ — •/1 − 0€• + 2*∆<—


∂; . “-ABCDE ” − 3 •/1 − 0€•™ — KM
/1 − 0€•
= µ ·− .
∂ * •/1 − 0€• + 2*∆<— 1/1 − 0

References

af Segerstad, P.H., Larsson, R., Toll, S., 2008. A constitutive equation for open-cell cellular solids,

including viscoplasticity, damage and deformation induced anisotropy. International journal of

plasticity 24, 896-914.

45
Alkhader, M., Vural, M., 2010. A plasticity model for pressure-dependent anisotropic cellular

solids. International Journal of Plasticity 26, 1591-1605.

Ames, N.M., Srivastava, V., Chester, S.A., Anand, L., 2009. A thermo-mechanically coupled theory

for large deformations of amorphous polymers. Part II: Applications. International Journal of

Plasticity 25, 1495-1539.

Anand, L., Ames, N.M., Srivastava, V., Chester, S.A., 2009. A thermo-mechanically coupled theory

for large deformations of amorphous polymers. Part I: Formulation. International Journal of

Plasticity 25, 1474-1494.

Ayoub, G., Zaïri, F., Fréderix, C., Gloaguen, J.M., Naït-Abdelaziz, M., Seguela, R., Lefebvre, J.M.,

2011. Effects of crystal content on the mechanical behaviour of polyethylene under finite strains:

experiments and constitutive modelling. International Journal of Plasticity 27, 492-511.

Ayoub, G., Zaïri, F., Naït-Abdelaziz, M., Gloaguen, J.M., 2010. Modelling large deformation

behaviour under loading–unloading of semicrystalline polymers: application to a high density

polyethylene. International Journal of Plasticity 26, 329-347.

Ayoub, G., Zaïri, F., Naït-Abdelaziz, M., Gloaguen, J.M., Kridli, G., 2014. A visco-hyperelastic

damage model for cyclic stress-softening, hysteresis and permanent set in rubber using the

network alteration theory. International Journal of Plasticity 54, 19-33.

Ayyagari, R.S., Vural, M., 2015. Multiaxial yield surface of transversely isotropic foams: Part I—

Modeling. Journal of the Mechanics and Physics of Solids 74, 49-67.

Ayyagari, R.S., Vural, M., 2016. On the nature of pressure dependence in foams. International

Journal of Solids and Structures 78, 160-173.

Baghani, M., Naghdabadi, R., Arghavani, J., Sohrabpour, S., 2012. A thermodynamically-

consistent 3 D constitutive model for shape memory polymers. International Journal of Plasticity

46
35, 13-30.

Balieu, R., Lauro, F., Bennani, B., Delille, R., Matsumoto, T., Mottola, E., 2013. A fully coupled

elastoviscoplastic damage model at finite strains for mineral filled semi-crystalline polymer.

International Journal of plasticity 51, 241-270.

Bodner, S., Partom, Y., 1975. Constitutive equations for elastic-viscoplastic strain-hardening

materials. Journal of Applied Mechanics 42, 385-389.

Bodner, S.R., 2001. Unified plasticity for engineering applications. Springer Science & Business

Media.

Bouvard, J.-L., Francis, D.K., Tschopp, M.A., Marin, E., Bammann, D., Horstemeyer, M., 2013. An

internal state variable material model for predicting the time, thermomechanical, and stress state

dependence of amorphous glassy polymers under large deformation. International Journal of

Plasticity 42, 168-193.

Boyce, M.C., Arruda, E.M., 2000. Constitutive models of rubber elasticity: a review. Rubber

chemistry and technology 73, 504-523.

Boyce, M.C., Parks, D.M., Argon, A.S., 1988. Large inelastic deformation of glassy polymers. Part I:

rate dependent constitutive model. Mechanics of materials 7, 15-33.

Cayzac, H.-A., Saï, K., Laiarinandrasana, L., 2013. Damage based constitutive relationships in semi-

crystalline polymer by using multi-mechanisms model. International Journal of Plasticity 51, 47-

64.

Chen, Y., Ghosh, S., 2012. Micromechanical analysis of strain rate-dependent deformation and

failure in composite microstructures under dynamic loading conditions. International Journal of

Plasticity 32, 218-247.

Chen, Y., Guo, W., Yang, P., Zhao, J., Guo, Z., Dong, L., Zhong, Z., 2018. Constitutive modeling of

47
neo-Hookean materials with spherical voids in finite deformation. Extreme Mechanics Letters 24,

47-57.

de Souza Neto, E.A., Peric, D., Owen, D.R., 2011. Computational methods for plasticity: theory

and applications. John Wiley & Sons.

Di Prima, M., Lesniewski, M., Gall, K., McDowell, D., Sanderson, T., Campbell, D., 2007. Thermo-

mechanical behavior of epoxy shape memory polymer foams. Smart Materials and Structures 16,

2330.

Ehlers, W., Markert, B., 2003. A macroscopic finite strain model for cellular polymers.

International Journal of Plasticity 19, 961-976.

Fahlbusch, N.-C., Grenestedt, J.L., Becker, W., 2016. Effective failure behavior of an analytical and

a numerical model for closed-cell foams. International Journal of Solids and Structures 97, 417-

430.

Frank, G.J., Brockman, R.A., 2001. A viscoelastic–viscoplastic constitutive model for glassy

polymers. International Journal of Solids and Structures 38, 5149-5164.

Guessasma, S., Nouri, H., 2016. Comprehensive study of biopolymer foam compression up to

densification using X-ray micro-tomography and finite element computation. European Polymer

Journal 84, 715-733.

Gurson, A.L., 1977. Continuum Theory of Ductile Rupture by Void Nucleation and Growth: Part

I—Yield Criteria and Flow Rules for Porous Ductile Media. Journal of Engineering Materials and

Technology 99, 2-15.

Hachour, K., Zaïri, F., Nait-Abdelaziz, M., Gloaguen, J.-M., Aberkane, M., Lefebvre, J.-M., 2014.

Experiments and modeling of high-crystalline polyethylene yielding under different stress states.

International journal of plasticity 54, 1-18.

48
Hou, C., Czubernat, K., Jin, S.Y., Altenhof, W., Maeva, E., Seviaryna, I., Bandyopadhyay-Ghosh, S.,

Sain, M., Gu, R., 2014. Mechanical response of hard bio-based PU foams under cyclic quasi-static

compressive loading conditions. International Journal of Fatigue 59, 76-89.

Khan, A.S., Farrokh, B., 2006. Thermo-mechanical response of nylon 101 under uniaxial and

multi-axial loadings: Part I, Experimental results over wide ranges of temperatures and strain

rates. International Journal of Plasticity 22, 1506-1529.

Khan, A.S., Huang, S., 1992. Experimental and theoretical study of mechanical behavior of 1100

aluminum in the strain rate range 10− 5− 104s− 1. Interna^onal Journal of Plas^city 8, 397-424.

Khan, A.S., Liang, R., 1999. Behaviors of three BCC metal over a wide range of strain rates and

temperatures: experiments and modeling. International Journal of Plasticity 15, 1089-1109.

Khan, A.S., Lopez-Pamies, O., Kazmi, R., 2006. Thermo-mechanical large deformation response

and constitutive modeling of viscoelastic polymers over a wide range of strain rates and

temperatures. International Journal of Plasticity 22, 581-601.

Khan, A.S., Suh, Y.S., Kazmi, R., 2004. Quasi-static and dynamic loading responses and constitutive

modeling of titanium alloys. International Journal of Plasticity 20, 2233-2248.

Khan, F., Yeakle, C., 2011. Experimental investigation and modeling of non-monotonic creep

behavior in polymers. International Journal of Plasticity 27, 512-521.

Kim, T.-R., Shin, J.K., Goh, T.S., Kim, H.-S., Lee, J.S., Lee, C.-S., 2017. Modeling of elasto-viscoplastic

behavior for polyurethane foam under various strain rates and temperatures. Composite

Structures 180, 686-695.

Launay, A., Maitournam, M., Marco, Y., Raoult, I., Szmytka, F., 2011. Cyclic behaviour of short glass

fibre reinforced polyamide: Experimental study and constitutive equations. International Journal

of Plasticity 27, 1267-1293.

49
Lee, C.-S., Chun, M.-S., Kim, M.-H., Lee, J.-M., 2011. Numerical evaluation for debonding failure

phenomenon of adhesively bonded joints at cryogenic temperatures. Composites Science and

Technology 71, 1921-1929.

Lee, C.-S., Kim, M.-S., Park, S.-B., Kim, J.-H., Bang, C.-S., Lee, J.-M., 2015. A temperature- and

strain-rate-dependent isotropic elasto-viscoplastic model for glass-fiber-reinforced polyurethane

foam. Materials & Design 84, 163-172.

Lee, C.-S., Lee, J.-M., 2014a. Anisotropic elasto-viscoplastic damage model for glass-fiber-

reinforced polyurethane foam. Journal of Composite Materials 48, 3367-3380.

Lee, C.-S., Lee, J.-M., 2014b. Failure analysis of reinforced polyurethane foam-based LNG

insulation structure using damage-coupled finite element analysis. Composite Structures 107,

231-245.

Lee, C.-S., Lee, J.-M., Youn, B., Kim, H.-S., Shin, J.K., Goh, T.S., Lee, J.S., 2017. A new constitutive

model for simulation of softening, plateau, and densification phenomena for trabecular bone

under compression. Journal of the mechanical behavior of biomedical materials 65, 213-223.

Lee, E.S., Goh, T.S., Lee, J.S., Heo, J.-Y., Kim, G.-B., Lee, C.-S., 2019. Experimental investigation of

macroscopic material nonlinear behavior and microscopic void volume fraction change for

porous materials under uniaxial compression. Composites Part B: Engineering 163, 130-138.

Lee, J.-H., Kim, S.-K., Park, S., Park, K.H., Lee, J.-M., 2018. Unified constitutive model with

consideration for effects of porosity and its application to polyurethane foam. Composites Part B:

Engineering 138, 87-100.

Lee, J.H., Kim, S.K., Park, S.B., Bang, C.S., Lee, J.M., 2016. Application of Gurson Model for

Evaluation of Density-Dependent Mechanical Behavior of Polyurethane Foam: Comparative

Study on Explicit and Implicit Method. Macromolecular Materials and Engineering 301, 694-706.

50
Lemaitre, J., 1984. How to use damage mechanics. Nuclear engineering and design 80, 233-245.

Lemaitre, J., 1985a. A continuous damage mechanics model for ductile fracture. Transactions of

the ASME. Journal of Engineering Materials and Technology 107, 83-89.

Lemaitre, J., 1985b. Coupled elasto-plasticity and damage constitutive equations. Computer

methods in applied mechanics and engineering 51, 31-49.

Lemaitre, J., 2012. A course on damage mechanics. Springer Science & Business Media.

Lemaitre, J., Chaboche, J.-L., 1994. Mechanics of solid materials. Cambridge university press.

Lemaitre, J., Desmorat, R., 2005. Engineering damage mechanics: ductile, creep, fatigue and

brittle failures. Springer Science & Business Media.

Li, H., Buckley, C., 2010. Necking in glassy polymers: Effects of intrinsic anisotropy and structural

evolution kinetics in their viscoplastic flow. International Journal of Plasticity 26, 1726-1745.

Li, S., Wang, G., 2008. Introduction to micromechanics and nanomechanics. World Scientific

Publishing Company.

Liang, R., Khan, A.S., 1999. A critical review of experimental results and constitutive models for

BCC and FCC metals over a wide range of strain rates and temperatures. International Journal of

Plasticity 15, 963-980.

Linul, E., Marsavina, L., Voiconi, T., Sadowski, T., 2013. Study of factors influencing the mechanical

properties of polyurethane foams under dynamic compression, Journal of Physics: Conference

Series. IOP Publishing, p. 012002.

Machado, G., Alves, M., Rossi, R., Silva Jr, C., 2011. Numerical modeling of large strain behavior of

polymeric crushable foams. Applied Mathematical Modelling 35, 1271-1281.

Malcher, L., Pires, F.A., De Sá, J.C., 2014. An extended GTN model for ductile fracture under high

and low stress triaxiality. International Journal of Plasticity 54, 193-228.

51
Marsavina, L., Constantinescu, D.M., 2015. Failure and damage in cellular materials, Failure and

Damage Analysis of Advanced Materials. Springer, pp. 119-190.

Mkaddem, A., Hambli, R., Potiron, A., 2004. Comparison between Gurson and Lemaitre damage

models in wiping die bending processes. The International Journal of Advanced Manufacturing

Technology 23, 451-461.

Park, S.-B., Lee, C.-S., Choi, S.-W., Kim, J.-H., Bang, C.-S., Lee, J.-M., 2016. Polymeric foams for

cryogenic temperature application: Temperature range for non-recovery and brittle-fracture of

microstructure. Composite Structures 136, 258-269.

Ponçot, M., Addiego, F., Dahoun, A., 2013. True intrinsic mechanical behaviour of semi-crystalline

and amorphous polymers: influences of volume deformation and cavities shape. International

Journal of Plasticity 40, 126-139.

Pyrz, M., Zairi, F., 2007. Identification of viscoplastic parameters of phenomenological constitutive

equations for polymers by deterministic and evolutionary approach. Modelling and Simulation in

Materials Science and Engineering 15, 85.

Regrain, C., Laiarinandrasana, L., Toillon, S., Saï, K., 2009. Multi-mechanism models for semi-

crystalline polymer: Constitutive relations and finite element implementation. International

Journal of Plasticity 25, 1253-1279.

Rodas, C.O., Zaïri, F., Naït-Abdelaziz, M., Charrier, P., 2016. A thermo-visco-hyperelastic model for

the heat build-up during low-cycle fatigue of filled rubbers: formulation, implementation and

experimental verification. International Journal of Plasticity 79, 217-236.

Shafiq, M., Ayyagari, R.S., Ehaab, M., Vural, M., 2015. Multiaxial yield surface of transversely

isotropic foams: Part II—Experimental. Journal of the Mechanics and Physics of Solids 76, 224-

236.

52
Shojaei, A., Li, G., 2013. Viscoplasticity analysis of semicrystalline polymers: a multiscale approach

within micromechanics framework. International Journal of Plasticity 42, 31-49.

Srivastava, V., Chester, S.A., Ames, N.M., Anand, L., 2010. A thermo-mechanically-coupled large-

deformation theory for amorphous polymers in a temperature range which spans their glass

transition. International Journal of Plasticity 26, 1138-1182.

Su, B., Zhou, Z., Shu, X., Wang, Z., Li, Z., Zhao, L., 2016. Multiaxial creep of transversely isotropic

foams. Materials Science and Engineering: A 658, 289-295.

Tvergaard, V., 1981. Influence of voids on shear band instabilities under plane strain conditions.

International Journal of Fracture 17, 389-407.

Tvergaard, V., Needleman, A., 1984. Analysis of the cup-cone fracture in a round tensile bar. Acta

metallurgica 32, 157-169.

Voyiadjis, G.Z., Shojaei, A., Li, G., 2011. A thermodynamic consistent damage and healing model

for self healing materials. International Journal of Plasticity 27, 1025-1044.

Voyiadjis, G.Z., Shojaei, A., Li, G., 2012. A generalized coupled viscoplastic–viscodamage–

viscohealing theory for glassy polymers. International Journal of Plasticity 28, 21-45.

Wang, D.-A., Pan, J., 2006. A non-quadratic yield function for polymeric foams. International

journal of plasticity 22, 434-458.

Whisler, D., Kim, H., 2015. Experimental and simulated high strain dynamic loading of

polyurethane foam. Polymer Testing 41, 219-230.

Yang, Q., Li, G., 2016. Temperature and rate dependent thermomechanical modeling of shape

memory polymers with physics based phase evolution law. International Journal of Plasticity 80,

168-186.

Yao, Y., Luo, Y., Xu, Y., Wang, B., Li, J., Deng, H., Lu, H., 2018. Fabrication and characterization of

53
auxetic shape memory composite foams. Composites Part B: Engineering 152, 1-7.

Zaïri, F., Naït-Abdelaziz, M., Gloaguen, J.-M., Lefebvre, J.-M., 2008. Modelling of the elasto-

viscoplastic damage behaviour of glassy polymers. International Journal of Plasticity 24, 945-965.

Zaïri, F., Naït-Abdelaziz, M., Gloaguen, J.M., Lefebvre, J.M., 2011. A physically-based constitutive

model for anisotropic damage in rubber-toughened glassy polymers during finite deformation.

International Journal of Plasticity 27, 25-51.

Zaïri, F., Naït-Abdelaziz, M., Woznica, K., Gloaguen, J.-M., 2005. Constitutive equations for the

viscoplastic-damage behaviour of a rubber-modified polymer. European Journal of Mechanics-

A/Solids 24, 169-182.

Zaïri, F., Naït-Abdelaziz, M., Woznica, K., Gloaguen, J.-M., 2007. Elasto-viscoplastic constitutive

equations for the description of glassy polymers behavior at constant strain rate. Journal of

Engineering Materials and Technology 129, 29-35.

Zhang, J., Kikuchi, N., Li, V., Yee, A., Nusholtz, G., 1998. Constitutive modeling of polymeric foam

material subjected to dynamic crash loading. International journal of impact engineering 21, 369-

386.

Zhang, J., Lin, Z., Wong, A., Kikuchi, N., Li, V., Yee, A., Nusholtz, G., 1997. Constitutive modeling

and material characterization of polymeric foams. Journal of engineering materials and

technology 119, 284-291.

Zhou, D., Xiong, Y., Yuan, H., Luo, G., Zhang, J., Shen, Q., Zhang, L., 2019. Synthesis and

compressive behaviors of PMMA microporous foam with multi-layer cell structure. Composites

Part B: Engineering 165, 272-278.

List of Figures
54
Fig. 1 Schematic uniaxial compressive stress-strain curves of the PUF and the deformed

test specimens at each strain

Fig. 2 Computational algorithm of ABAQUS UMAT

Fig. 3 Compressive stress-strain relationship of PUF according to various strain rates

under (a) 213 K, (b) 296 K, and (c) 353 K (Kim et al., 2017; Marsavina and

Constantinescu, 2015)

Fig. 4 SEM magnification view of RPUF at (a) 500 μm and (b) 300 μm scales (Lee and

Lee, 2014a)

Fig. 5 SEM photographs of (a) initial and (b) after compression of RPUF (Lee and Lee,

2014a)

Fig. 6 Microstructural response of polymeric foam to deformation at 398 K: (a)

relationship among normalized average cell size, compressive strain and stress, and (b)

micro-CT snapshot of the foam during deformation from 0 to 80% at 10% strain

intervals (read left-to-right, top-to-bottom) (Di Prima et al., 2007)

Fig. 7 Degradation of elastic modulus for (a) neat and (b) reinforced PUF (Hou et al.,

2014; Lee et al., 2018)

Fig. 8 Finite element model and its loading/boundary conditions for simulation

Fig. 9 Identification procedures of material parameters for the modified KHL model and

the modified GTNL model

Fig. 10 Comparison results of compressive stress-strain curve, normalized void volume

fraction, and degradation of elastic modulus of closed-cell PUF between simulations

and experiments at 0.003/s, 0.075/s, and 0.7/s under 213, 29, and 353 K

Fig. 11 Comparison of compressive stress-strain curve, normalized void volume fraction,


55
and degradation of elastic modulus of closed-cell PUF between simulations and

experiments at 8.3/s, 27.8/s, and 55.6/s under 213, 296, and 353 K

Fig. 12 Comparison of compressive stress-strain curve, normalized void volume fraction,

and degradation of elastic modulus of closed-cell PUF between simulations and

experiments at 83.3/s, 250/s, and 500/s under 213, 296, and 353 K

Fig. 13 Comparison of compressive stress-strain curve, normalized void volume fraction

and degradation of elastic modulus of closed-cell PSF between simulations and

experiments at 0.0016/s, 0.08/s, and 4.6/s under 296 K

Fig. 14 Comparison of compressive stress-strain curve, normalized void volume fraction

and degradation of elastic modulus of open-cell PUF between simulations and

experiments at 0.0016/s, 0.08/s, and 4.6/s under 296 K

List of Tables

Table 1 Material parameters of the modified KHL model

Table 2 Material parameters of the modified GTNL model

Table 3 Coefficient of equations expressing the elastic modulus

Table 4 Coefficient of equations expressing the strain for stress-drop

56
Table 1 Material parameters of the modified KHL model

Parameter Closed-cell PUF Open-cell PUF Closed-cell PSF

6.316 0.00945 0.0002283

0.03175 -0.02534 -0.2938

0.01191 0.005872 0.09827

0.978 0.2354 0.03656

1.844E+4 0.9439 19.86

7.563 3.134 3.969

(K) 296.15 296.15 296.15

(K) 473.15 473.15 473.15

0.296 0.4699 0.4699

1.0E+6 1.0E+6 1.0E+6


. !
1.654 ∗∗ .
− 1.755 -1 3.923 − 8.229

# 25 1 1

# 0.049 exp (−0.033ln + + 0.001exp (0.103ln + 0 0

1
Table 2 Material parameters of the modified GTNL model

Parameter Closed-cell PUF Open-cell PUF Closed-cell PSF

- 0.829 0.9 0.9

. 9 13 13

.
. !/ . / /
1 0.03139 + 0.1145 0.1153 − 0.009378

0 0.6 0.3 0.001

1 0.82 0.91 0.92

1 1 1 1

1 1 1 1

. exp 3 ln 4 56 + exp 3A ln 4 56 . !!; . ; /


39 2
5.221
8 91 − ∗ : .!
1.563 + 47.230 −8.822 + 50.870

<
.
exp 3 ln 4 56 + exp 3A ln 4 56
0.02 2 8 91 − ∗: . 1 1
5.221

2
Table 3 Coefficient of equations expressing the elastic modulus

Coefficient Elastic modulus, =

(Unit) Closed-cell PUF Open-cell PUF Closed-cell PSF

> (MPa) −0.0495 + 20.01 0.2099 -0.05088

> (-) 0.4437 0.05733 -0.3854

> (MPa)
. /
99.001 × 10 : + 62.6 0.03161 2.829

Note: = 1K.

3
Table 4 Coefficient of equations expressing the strain for the stress-drop

Coefficient Strain for the stress-drop, BCDE

(Unit) Closed-cell PUF Open-cell PUF Closed-cell PSF

> (mm/mm) 94.958 × 10 / : − 0.004 - -

> (-)
/.F;
93.551 × 10 ;:
+ 0.215 - -

> (mm/mm)
.
9−1.504 × 10 : + 0.096 - -

Note: = 1K.

4
Fig. 1 Schematic uniaxial compressive stress-strain curves of the PUF and the deformed

test specimens at each strain


Fig. 2 Computational algorithm of ABAQUS UMAT
(a) (b)

(c)

Fig. 3 Compressive stress-strain relationship of PUF according to various strain rates

under (a) 213 K, (b) 296 K, and (c) 353 K (Kim et al., 2017; Marsavina and

Constantinescu, 2015)
(a) (b)

Fig. 4 SEM magnification view of RPUF at (a) 500 μm and (b) 300 μm scales (Lee

and Lee, 2014)


Loading direction

(a) (b)

Fig. 5 SEM photographs of (a) initial and (b) after compression of RPUF (Lee and Lee,

2014)
(a) (b)

Fig. 6 Microstructural response of polymeric foam to deformation at 398 K: (a)

relationship among normalized average cell size, compressive strain and stress, and (b)

micro-CT snapshot of the foam during deformation from 0 to 80% at 10% strain

intervals (read left-to-right, top-to-bottom) (Di Prima et al., 2007)


(a) (b)

Fig. 7 Degradation of elastic modulus for (a) neat and (b) reinforced PUF (Hou et al.,

2014; Lee et al., 2018)


Fig. 8 Finite element model and its loading/boundary conditions for simulation
Fig. 9 Identification procedures of material parameters for the modified KHL model and

the modified GTNL model


Fig. 10 Comparison results of compressive stress-strain curve, normalized void volume

fraction, and degradation of elastic modulus of closed-cell PUF between simulations

and experiments at 0.003/s, 0.075/s, and 0.7/s under 213, 29, and 353 K
Fig. 11 Comparison of compressive stress-strain curve, normalized void volume fraction,

and degradation of elastic modulus of closed-cell PUF between simulations and

experiments at 8.3/s, 27.8/s, and 55.6/s under 213, 296, and 353 K
Fig. 12 Comparison of compressive stress-strain curve, normalized void volume fraction,

and degradation of elastic modulus of closed-cell PUF between simulations and

experiments at 83.3/s, 250/s, and 500/s under 213, 296, and 353 K
Fig. 13 Comparison of compressive stress-strain curve, normalized void volume fraction

and degradation of elastic modulus of closed-cell PSF between simulations and

experiments at 0.0016/s, 0.08/s, and 4.6/s under 296 K


Fig. 14 Comparison of compressive stress-strain curve, normalized void volume fraction

and degradation of elastic modulus of open-cell PUF between simulations and

experiments at 0.0016/s, 0.08/s, and 4.6/s under 296 K

Di Prima, M., Lesniewski, M., Gall, K., McDowell, D., Sanderson, T., Campbell, D., 2007.
Thermo-mechanical behavior of epoxy shape memory polymer foams. Smart Materials and
Structures 16, 2330.
Hou, C., Czubernat, K., Jin, S.Y., Altenhof, W., Maeva, E., Seviaryna, I., Bandyopadhyay-Ghosh,
S., Sain, M., Gu, R., 2014. Mechanical response of hard bio-based PU foams under cyclic quasi-
static compressive loading conditions. International Journal of Fatigue 59, 76-89.
Kim, T.-R., Shin, J.K., Goh, T.S., Kim, H.-S., Lee, J.S., Lee, C.-S., 2017. Modeling of elasto-
viscoplastic behavior for polyurethane foam under various strain rates and temperatures.
Composite Structures 180, 686-695.
Lee, C.-S., Lee, J.-M., 2014. Anisotropic elasto-viscoplastic damage model for glass-fiber-
reinforced polyurethane foam. Journal of Composite Materials 48, 3367-3380.
Lee, J.-H., Kim, S.-K., Park, S., Park, K.H., Lee, J.-M., 2018. Unified constitutive model with
consideration for effects of porosity and its application to polyurethane foam. Composites Part B:
Engineering 138, 87-100.
Marsavina, L., Constantinescu, D.M., 2015. Failure and damage in cellular materials, Failure and
Damage Analysis of Advanced Materials. Springer, pp. 119-190.
Highlights

Unified elasto-viscoplastic-damage model for polymeric foams was developed.

Stress-strain curve, void volume fraction change and damage growth were evaluated.

Full derivation for implicit integration algorithm of present model was introduced.

ABAQUS user-defined subroutine (UMAT) for polymeric foams were developed.

Proposed method was validated by comparing a series of experimental results.


Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

You might also like