Sunteți pe pagina 1din 18

International Journal of Solids and Structures 44 (2007) 7615–7632

www.elsevier.com/locate/ijsolstr

A computational micromechanics constitutive model for


the unloading behavior of paper
M.K. Ramasubramanian *, Yun Wang
Department of Mechanical and Aerospace Engineering, North Carolina State University, 3211 Broughton Hall,
Raleigh, NC 27695-7910, USA

Received 1 September 2006; received in revised form 8 March 2007


Available online 10 May 2007

Abstract

In this study, a computational micromechanics material model for the unloading behavior of paper and other nonwo-
ven materials is presented. The asymptotic fiber and bond (AFB) model for paper elastic–plastic behavior [Sinha, S.K.,
Perkins, R.W., 1995. Micromechanics constitutive model for use in finite element analysis, In: Proceedings of the 1995,
Joint ASME Applied Mechanics and Materials Summer Meeting, Los Angeles, CA, USA, Jun 28–30, 1995] has been
extended to model the unloading process through a computational algorithm and implemented using the UMAT subrou-
tine in ABAQUS finite element code. For every unloading increment, the material model assumes elastic unloading with a
slope equal to the initial elastic modulus. The Jacobian matrix of the constitutive model is updated at every unloading
increment by applying the incremental form of AFB model for a planar element with an elastic fiber and bond condition.
A uniaxial tensile and a biaxial Mullen burst loading–unloading experiments were carried out for a paperboard sample and
simulated using the model. The stress–strain curve and residual strain for the uniaxial loading were in good agreement with
experimental results. The finite element model of the burst test with the AFB unloading material model predicted the gen-
eral shape of the pressure versus deflection curve. However, the model over predicted the residual deflection by more than
50%. The loading portion of the pressure–deflection curve had a significant offset from experimental curves, and the non-
linearity in the unloading curve towards the end was not predicted. The discrepancies with experimental results are attrib-
uted to the burst test itself, model parameter estimation inadequacies, boundary conditions used in the FEA, and
neglecting time-dependant effects. Nevertheless, the model can be useful in parametric studies relating microstructure to
unloading behavior in structural problems.
 2007 Elsevier Ltd. All rights reserved.

Keywords: Micromechanics; Paper; Nonwovens; Constitutive model; Plasticity; Unloading; Cyclic loading

1. Introduction

Paper is an orthotropic material, made up of self-bonding short cellulosic fibers preferentially oriented in
the machine direction (MD), and exhibiting a distinct microstructure consisting of fibers, fiber fragments, and

*
Corresponding author. Tel.: +1 919 515 5262; fax: +1 919 515 7968.
E-mail address: rammk@ncsu.edu (M.K. Ramasubramanian).

0020-7683/$ - see front matter  2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijsolstr.2007.05.002
7616 M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632

voids. Due to the long slender nature of the fibers and the papermaking process of forming the web on a por-
ous fabric from a very dilute fiber suspension, fibers generally conform to the plane of the paper, forming
essentially a two-dimensional layered structure. On a larger size scale (millimeters), paper structure exhibits
variation in mass density due to the non-uniform nature of dispersion during the papermaking process. These
non-uniformities affect the strength and appearance of paper. Paper exhibits a strong nonlinear elastic–plastic
behavior under monotonic uniaxial loading (Niskanen and Karenlampi, 1998). Upon unloading from plastic
regime, the paper exhibits permanent deformation and time-dependent recovery, and upon reloading, exhibits
a hysteresis (Bronkhorst, 2003).

1.1. Elastic behavior

The elastic behavior and strength of paper have been studied extensively in the past through a combination
of experimental, semi-empirical, and analytical approaches (Niskanen and Karenlampi, 1998). Constitutive
models to predict the elastic behavior of paper in terms of its constituents have been developed by several
researchers working at different size scales starting from hydrogen bonds to a macroscopic sheet continuum
(Cox, 1952; Corte and Kallmes, 1962; Van den Akker, 1970; Page et al., 1979; Perkins, 1986, 1990; Batten
and Nissan, 1987; Schulgasser and Page, 1988; Schulgasser and Grunseit, 1990; Suhling, 1990; Lu et al.,
1996; Ostoja-Starzewski and Stahl, 2001; Xie et al., 2002). While all the referenced work above are 2-D planar
or 2-D laminated composite or mosaic models, a 3-D network model was presented by Heyden and Gustafs-
son (2001). This model for paper takes a cubic unit cell, with dimensions 1.2–4.8 times the fiber length, and
places fibers in them randomly in 3-D. The fiber crossings were considered bonded when the proximity of
fibers were within a prescribed tolerance. Fibers and bonds were assumed to be linear elastic. The bonds were
assumed to exhibit a stick–slip behavior. The fibers were modeled as linear elastic Bernoulli beams of constant
curvature in a finite element algorithm to simulate uniaxial loading in all three directions. The results from
such a simulation were the elastic parameters of the network, and strength related parameters such as fracture
energy. The model was used to simulate fiber networks containing 173 fibers for out-of-plane density to in-
plane density ratios of 1.0–0.2. A ratio of 1 will represent fiber fluff and smaller ratios will tend towards
two-dimensional densified paper structure. Only elastic moduli alone were calculated for all density ratios.
The stress–strain curve was simulated only for ratios of 1.0 and 0.8. Below 0.8, the degrees of freedom
increased rapidly, presumably making it computationally expensive. No experimental results were presented
to validate the model. Thus, the model is useful for understanding fiber fluff and very low density paper behav-
ior and not applicable for a majority of paper and paperboard microstructures. Using this approach, it would
be almost impossible to model realistic macroscopic structural problems involving paper materials. Similar
approaches to defining a 3-D unit cell accurately with statistics and geometry have been presented in the area
of textiles and nonwovens (Komori and Makishima, 1977; Komori and Itoh, 1991). In all cases, three-dimen-
sional network models very quickly become computationally unmanageable. Hence, these models have only
been applied in elastic uniaxial loading cases.

1.2. Inelastic behavior

While extensive work has been carried out on the elastic behavior of paper, inelastic behavior of paper has
received relatively less attention. Treating paper as a homogeneous continuum, Johnson and Urbanik (1984)
formulated a nonlinear elastic plate theory and defined a strain energy density function W. The form of the
strain energy density function was not known. The stress–strain relationship for the material was obtained in a
hyperbolic tangent functional form through curve fitting uniaxial experimental data. The hyperbolic tangent
function was then integrated to obtain the strain energy density function W. Once the strain energy density
function is obtained, stress–strain relations in other directions were obtained through the relationship
oW
rij ¼ oe ij
. Suhling (1990) used a hyperelastic model in a similar framework to predict the nonlinear stress–strain
curve. In this approach, although the nonlinear stress–strain curve is correctly predicted, the curve is elastic. It
cannot predict the unloading path correctly. Castro and Ostoja-Starzewski (2003) extended the hyperbolic
tangent model to fit both uniaxial and biaxial test data. A three-dimensional, anisotropic constitutive model
has been presented to model the in-plane elastic–plastic deformation of paper and paperboard (Xia et al.,
M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632 7617

2002), which uses a continuum plasticity model with an evolving hardening parameter to predict the aniso-
tropic elastic–plastic behavior of paper. Makela and Ostlund (2003) presented a continuum orthotropic elas-
tic–plastic material model for paper applying the isotropic plastic equivalent (IPE) (Karafillis and Boyce,
1993) concept. By equating the IPE-material and the orthotropic material energy dissipation rates, the incre-
mental plastic strain tensor for the IPE-material and that of the anisotropic material were also related by the
IPE transformation tensor. Strain hardening behavior was modeled with a modified form of Ramberg–
Osgood relation and the expressions for plastic strain increments were derived. The plane-stress constitutive
model was implemented using the user-defined material (UMAT) subroutine in ABAQUS/Standard finite ele-
ment code. Uniaxial tests were simulated and compared with experiments. Unloading behavior was not con-
sidered. A continuum model that combines nonlinear elasticity and bounding surface plasticity, and
implemented using the UMAT subroutine in ABAQUS model, for modeling through-thickness elastic–plastic
behavior of paper was presented by Stenberg (2003). In this model, elastic plastic unloading and reloading was
considered in compression, out-of-plane tension was treated as elastic.
Inelastic behavior of paper has been modeled through a micromechanical analysis by Ramasubramanian
(1987), Ramasubramanian and Perkins (1988), and experimentally studied by Furukawa et al. (1991). A fiber
(inclusion phase) with portions of homogenized crossing fibers (matrix phase) for shear lag load transfer was
defined as the representative volume element (RVE). Network geometry, fiber length and orientation distribu-
tions, sheet density, and fiber and fiber-to-fiber bond constitutive behavior were taken into account. Fiber
behavior can be different in tension and compression thereby the possibility of fiber buckling under compres-
sion is included. Monotonic increase in load was simulated. Stress–strain curves and lateral contraction ratios
were predicted. Unloading was not considered. The model is computationally intensive in that for every sheet
strain value, the fiber axial strain distribution and the fiber-to-fiber bond shear stress distribution in every
direction has to be computed and the portion of the fibers and bonds that are elastic and plastic have to
be computed.
To reduce the number of computations without significant compromise in quality, Sinha (1994) proposed
an asymptotic fiber and bond model to handle the analysis of RVE used by Ramasubramanian and Perkins
(1988). Instead of determining the lengths of portions of every fiber and bond that is elastic or plastic at every
orientation and strain level, the asymptotic model assumes that when the maximum strain in a fiber or bond
reach its critical stress value, the corresponding fiber or bond yields completely along its entire length. Thus,
the end effects and shear lag mechanism are not addressed. The simplified model reduced considerably the
amount of numerical calculation required without compromising the quality of the results (Sinha, 1994).
An incremental form of the asymptotic model was implemented in as the material model using the UMAT
procedure in ABAQUS finite element code and monotonic uniaxial tensile loading was studied (Sinha and
Perkins, 1995).
A two-dimensional stochastic computational network model to simulate uniaxial tension tests was pre-
sented by Bronkhorst (2003). A two-dimensional stochastically generated network consisting of fibers in a
10 mm · 10 mm area was studied. The fiber-to-fiber bond was assumed rigid and the fibers behavior was mod-
eled with continuum plasticity models. A finite element model of the structure was created in ABAQUS finite
element code, modeling fibers as isotropic elastic–plastic beams, and loaded in uniaxial tension. The stress–
strain curve generated was compared with experiments. One cyclic tensile test result was presented and com-
pared with experiment. The discrepancy in later stages of unloading was attributed to time-dependent effects
not accounted for in the model. While this model could be used to study larger problems, the RVE of the
structure is a macroscopic sheet (10 mm · 10 mm, for example). Thus, local effects of fiber-to-fiber bond yield-
ing and fiber failure could not be taken into account. Reducing the size of RVE may be one option, but it
might increase the computational requirements as reported in other stochastic network model approaches
(Heyden and Gustafsson, 2001).

1.3. Unloading and cyclic loading behavior

While there has been good progress in recent years in modeling the inelastic behavior of paper, very little
has been done to model the unloading and cyclic loading behaviors (Ramasubramanian and Wang, 1999). As
discussed in previous section, Bronkhorst (2003) showed a result from cyclic loading simulation based on
7618 M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632

stochastic network model with rigid bonds and elastic plastic fibers. Kang et al. (2006) presented a numerical
approach for uniaxial cyclic deformation of discontinuously reinforced (particulate and short fiber) metal-
matrix composites. The fibers were treated as an elastic material and the matrix was treated as an elasto-plastic
material. Infinitesimal elasto-plasticity with isotropic elasticity, along with kinematic hardening and an evolu-
tion rule for back stress was used. The constitutive model was revised in an incremental form and placed in the
UMAT subroutine in ABAQUS. Cyclic deformation of metal-matrix composites was simulated through a sin-
gle fiber and matrix model and verified with experiments. The effects of weak interfacial bonding, fiber orien-
tation, thermal residual stress and other microstructural features, and stress relaxation were not taken into
account. Nevertheless, the model offers a good framework to develop a computational constitutional model
for unloading and cyclic loading behavior of short fiber composites such as paper. In another study, (Doghri
and Tinel, 2006), an incremental micromechanics constitutive model was proposed for rate independent inelas-
tic problems in short fiber reinforced composites. The incremental form of model was implemented through
the UMAT procedure in ABAQUS and unloading cases were simulated, and compared with experiments.

2. Objectives

In this paper, we present a Computational methodology for simulating the unloading behavior of paper in
a micromechanics framework already established for the simulation of inelastic behavior of paper (Ramasubr-
amanian and Perkins, 1988; Sinha, 1994; Sinha and Perkins, 1995). We present the incremental form of the
asymptotic fiber and bond model and the unloading algorithm implemented into ABAQUS finite element code
using the UMAT subroutine. We report results from simulation of a uniaxial and a biaxial loading–unloading
problem, and compare results with tensile tests and biaxial tests on a Mullen burst tester, respectively.

3. Micromechanics model

3.1. Full plasticity model

The representative volume element (RVE) or the mesoelement for the model consists of a fiber and a por-
tion of all crossing fibers attached to it as shown in Fig. 1 (Ramasubramanian and Perkins, 1988). The fiber-
to-fiber bonds are distributed on the fiber and essentially transfers load to the reference fiber in the RVE
through a shear lag mechanism. The effect of fiber-to-fiber bonds is smeared or homogenized along the fiber
like a surrounding deformable matrix. A bilinear elastic–plastic material model is used to model the constitu-
tive behavior of the fiber and then bond, shown in Figs. 2 and 3 (Ramasubramanian, 1987; Sinha, 1994). In
this two-slope model, the elastic modulus of fiber is the same in tension and compression. In order to address
localized failure, such as fiber buckling, the compressive strength and tangent modulus (second slope) of the
fiber cell wall material could be much smaller than their respective tensile counterparts.
The constitutive model of fiber in tension would then be given by

θ
Fiber Machine
Direction

RVE
Element
Boundary

Bonds

Fig. 1. Representative Volume Element (RVE) for the micromechanics model.


M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632 7619

σf

σ pt E2t

Ef
ε pc
ε pt εf

E2c
σ pc

Fig. 2. Fiber constitutive model.

τb G2

Gb
γ
γb

Fig. 3. Fiber-to-fiber bond constitutive model.

rf ¼ Ef ef ; rf < rpt ð1Þ


rf ¼ rpt þ E2t ðef  ept Þ; rf P rpt ð2Þ

and in compression
rf ¼ Ef ef ; jrf j < jrpc j ð3Þ
rf ¼ rpc þ E2c ðef  epc Þ; jrf j P jrpt j ð4Þ

where rf = fiber stress, ef = fiber strain, Ef = fiber elastic modulus, E2t = fiber tangent modulus (second slope)
in tension, E2c = fiber tangent modulus (second slope) in compression, rpt = fiber yield stress in tension,
rpc = fiber yield stress in compression, ept = fiber yield strain in tension, and epc fiber yield stress in
compression.
The fiber-to-fiber bond, the constitutive behavior can be described by
sb ¼ G b cb ; jsb j < jsp j ð5Þ
sb ¼ sp þ G2 ðc  cb Þ; jsb j P jsp j ð6Þ

where, sb = bond shear stress, cb = bond shear strain, Gb = bond shear elastic modulus, G2 = bond tangent
modulus (second slope), sp = bond yield stress, and cp = bond yield strain.
One half of the RVE is shown in Fig. 4. The length Lf represents the portion of the fiber that is elastic, Lb
represents portion of the bond that is plastic, and L is the half length of RVE.
The location of these boundaries were computed in a full plasticity model (Ramasubramanian, 1987) for
the following three possibilities, namely, elastic fiber and bonds, elastic fiber–plastic bond, and plastic fiber
and bond for an RVE. Once, these boundaries were determined, the total work function was numerically
7620 M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632

Lf
Lb
L

Fig. 4. Portion of a RVE showing location of elastic–plastic interfaces from the center of the element.

computed. Determining the location of elastic–plastic boundaries for every macroscopic applied strain and
fiber orientation made this computationally intensive and limited the scope of this approach.

3.2. Asymptotic fiber and bond model

In the asymptotic model, the average fiber stress was computed once a critical stress was reached in the mid-
dle of the fiber, thus homogenizing the fiber stress in the RVE. The average stress is computed by
Z
1 L
f ¼
r rf dn ð7Þ
L 0

Four critical strains values are used to determine the fiber and bond conditions, namely, ecb, ecf, ebf, and efb.
Starting with elastic fiber and elastic bond, ecb is associated with the initiation of bond yielding before fiber
yielding, and ecf is associated with the initiation of fiber yielding before bond yielding. The strain ebf is asso-
ciated with the initiation of fiber yielding after the bond has already yielded, and efb is associated with the ini-
tiation of bond yielding after the fiber has already yielded. If ebf is greater than efb, then bond would yield first
and vice versa. In this paper, we only consider the case of bond yielding while the fiber remains elastic.
From equilibrium considerations of an axial section of the RVE, the fiber stress expressions can be derived
for each case using the full plasticity model (Ramasubramanian and Perkins, 1988). The expressions can be
then simplified using the asymptotic approximation for this work (Sinha, 1994) and the results are shown
in Eqs. (8) and (9).
Case of bond yielding before fiber (ecf > ecb)
Average fiber stress for elastic fiber and elastic bond is given by
 
tanh aL
 f ¼ E f es 1 
r ð8Þ
aL

Average fiber stress for elastic fiber and plastic bond is given by
   
tanh aL tanh ab L
f ¼ Ef ecb 1 
r þ Ef ðes  ecb Þ 1  ð9Þ
aL ab L

where es is the sheet strain in the direction of the RVE. In terms of the applied macroscopic strains, and the
orientation of the RVE, the sheet strain is given by
es ¼ ex cos2 h þ ey sin2 h þ 2exy cos h sin h ð10Þ

and
Gb we
a2 ¼ ð11Þ
E f Af t b
G2 we
a2b ¼ ð12Þ
E f Af t b
M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632 7621

where we is the width of the RVE, Af is the cross section area of the fiber, and tb is the thickness of the bond
region in the RVE. The critical strain expression for the case of elastic fiber and plastic bond in terms of bond
yield stress sp is (Sinha, 1994):
 
sp we L 1
ecb ¼ ð13Þ
Ef Af aL tanh aL
The sheet stress rx and ry are then obtained from the integral sum of the appropriate average fiber stress com-
ponent for fibers in all orientations,
Z p=2
q
rx ¼ s f ðhÞ cos2 hf ðhÞdh
r ð14Þ
qf p=2
Z p=2
q
ry ¼ s f ðhÞ sin2 hf ðhÞdh
r ð15Þ
qf p=2
where f(h) is the von Mises function used to describe the fiber orientation distribution, qs is the sheet density
and qf is the density of the fiber cell wall material (Ramasubramanian and Perkins, 1988).

3.3. Incremental form of the asymptotic fiber and bond model

In the incremental form (Sinha and Perkins, 1995), the incremental sheet strain in the direction of the RVE
in terms of applied incremental two-dimensional macroscopic strains field is given by
Des ¼ Dex cos2 h þ Dey sin2 h þ 2Dexy cos h sin h ð16Þ
The total axial strain in the direction of RVE is given by the sum of the incremental strains
Xn
es ¼ ðDes Þi ð17Þ
i¼1

The elemental Drij is assumed to be sufficiently small such that it is still linearly related to Deij by the instan-
taneous slope of the incremental stress–strain curve. The fiber stresses in the incremental form can be derived
in a fashion similar to the full asymptotic approach. For the case of elastic fiber and elastic bond (cf. Eq. (8)),
the average incremental stress is given by (Sinha, 1994)
 
tanh aL
hDrf i ¼ Ef ðDes Þ 1  ð18Þ
aL
The incremental sheet stresses are obtained by
Z
qs p=2
Drx ¼ hDrf ðhÞi cos2 hf ðhÞdh ð19Þ
qf p=2
Z
qs p=2
Dry ¼ hDrf ðhÞi sin2 hf ðhÞdh ð20Þ
qf p=2
The total sheet stress is a sum of all the previous sheet stress increments
X
rx ¼ ðDrx Þ ð21Þ
X
ry ¼ ðDry Þ ð22Þ

4. Computational material model

The incremental asymptotic fiber and bond model was incorporated into ABAQUS/Standard for the solu-
tion of problems with complex geometries and multi-axial loading, through the UMAT subroutine (ABA-
QUS, 1998). This subroutine will be called at all material calculation points of elements and the subroutine
7622 M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632

must update the stresses and solution-dependent state variables to their values at the end of the increment. The
user-defined
 material
 model is incorporated by providing complete expressions for the terms in the Jacobian
matrix oDrij =oDeij . The Jacobian matrix represents the incremental change in material properties for an ele-
ment and is determined by applying the incremental micromechanics model for a planar element in our case.
The constitutive relationships based on the asymptotic fiber and bond model in its incremental form can be
written in a matrix form as shown in Eq. (23).
8 9 2 38 9
< Drx >
> = k 11 k 12 k 13 >< Dex >=
6 7
Dry ¼ 4 k 21 k 22 k 23 5 Dey ð23Þ
>
: >
; >
: >
;
Drxy k 31 k 32 k 33 Dexy
where
Z 1 Z p=2
q
k 11 ¼ s F cos4 hf ðhÞf ðkÞdh dk ð24Þ
qf 0 p=2
Z 1 Z p=2
q
k 12 ¼ s F sin2 h cos2 hf ðhÞf ðkÞdh dk ð25Þ
qf 0 p=2
Z 1 Z p=2
qs
k 13 ¼ F sin h cos3 hf ðhÞf ðkÞdh dk ð26Þ
qf 0 p=2
Z 1 Z p=2
qs
k 22 ¼ F sin4 hf ðhÞf ðkÞdh dk ð27Þ
qf 0 p=2
Z 1 Z p=2
qs
k 23 ¼ F sin3 h cos hf ðhÞf ðkÞdh dk ð28Þ
qf 0 p=2

k 33 ¼ k 12 ð29Þ
The F functions are derived from micromechanics which are essentially the fiber stresses in the RVE for
different loading cases (Sinha, 1994). In our study, we have considered the fiber to be elastic and bond to
be elastic–plastic. The two relevant cases are given below.
Fiber and bond elastic (es 6 ecb)
 
tanh aL
F ¼ Ef 1  ð30Þ
aL
Fiber elastic and bond plastic (es > ecb)
 
tanh ab L
F ¼ Ef 1  ð31Þ
ab L

4.1. General unloading algorithm

When a fiber is unloading, es would have an opposite sense to that of Des. Both fiber and bond would
behave elastically. The F function that corresponds to elastic fiber and bond condition (Eq. (30)) is used in
the computation of the Jacobian matrix and the corresponding change in rij. The algorithm for unloading cal-
culations is shown in Fig. 5. For every time step, each iteration of the Finite Element equilibrium conditions,
and at each integration (Gauss) point of the mesh, ABAQUS calls UMAT as shown in Fig. 5. Consider now
an undeformed sheet subjected to tensile loading in the machine direction. The values of the model parameters
are defined in the input file for use in the UMAT subroutine. In this work, ecb is greater than ecf in tension and
compression. Therefore, bond would yield first when the critical strain ecb is reached. The fiber orientation
distribution, f(h), is discretized into 30 increments varying from p/2 to p/2. For a given loading condition
on the sheet, the algorithm begins with h = p/2 and calculates the corresponding es (h). The critical strains
are calculated. Suppose es(h) is loaded in tension (sgn(es) = sgn(Des) and Des > 0), beyond the yield point ecb.
M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632 7623

Call UMAT Calculate εs(θ) and Δεs(θ )


Subroutine Orientation dependent model parameters Ef, E2, G b,
G2, a, ab, Critical strain ε cb

Read Input Parameters ε x , ε y , ε xy , σ x , σ y , and σ xy for


the kth time step; Δε x , Δε y , and Δε xy for the (k+1)th NO
Sign of σ f ( θ) =
time step Sign of Δε f(θ) ?

Unloading
Define model parameters Process:
E f and E 2 . Gb and G2 YES Calculate F
function using the
τ p ,ε pt ,ε pc , ρ s , ρ f , κ
Loading process expression for
elastic fiber and
bond

NO
I=1 I=
I+1= 31?

θ ( I ) = -π / 2 + ( I -1) (π / 30)
YES

Determine kij of Jacobian Matrix


using Simpson’s 1/3 rule.

Calculate Δσ ij

σ ijk +1 = σ ijk + Δσ ij

Return to
ABAQUS

Fig. 5. Incremental form of asymptotic model UMAT subroutine implementation flow chart.

Bond would yield and the total strain of es could be broken into two components, i.e. the elastic and plastic
strain component. The es,el is recovered when the sheet is unloaded. The kij in the Jacobian matrix are obtained
by evaluating the integral of the F functions numerically using the Simpson’s 1/3 rule at every material calcu-
lation point. The incremental stress is then calculated and the total stress value is returned to ABAQUS. The
algorithm is capable of addressing cyclic loading. During cyclic loading, the critical strains have to be modified
accordingly so that the ensuing plastic deformation would occur at the correct stress limit. Moisture and tem-
perature effects can be easily added to the incremental constitutive relationships (Sinha, 1994).

4.2. Cyclic loading

Consider a fiber with an orientation h, to the MD and loaded in tension until the bond in the RVE has
yielded. When it is unloaded, the fiber strain would have a nonzero plastic component, es,pl. This nonzero
value of the fiber plastic strain component is added to the critical strains in compression. After complete
unloading, if the fiber is subjected to compression, the updated critical strains allow the RVE to yield at
the correct stress limit in compression.
7624 M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632

The elastic and plastic strain components of es are also stored for decision making in the algorithm and
updating the critical strains for subsequent loading cycles. In this study, only the bond plastic critical strain
ecb for the RVE is considered. It is updated as follows.
ecb ¼ ecb;initial þ es;pl ð32Þ
For a fiber subjected to multiple cycles of loading and unloading process, the critical strains are updated by the
same equation, except the ecb,initial. is replaced by the previous value of ecb. When subjected to compression
after unloading, the original critical strains are adjusted by adding the cumulative absolute value of the plastic
strains, eoffset (cf. Eq. (33)). In tension, the offset value is subtracted from the original critical strains.
Xn  
i 
eoffset ¼ es;pl  ð33Þ
i¼1

At any instance of unloading, the F function for elastic fiber and elastic bond is used to calculate the material
contribution to the Jacobian matrix.

5. Results

5.1. Uniaxial loading–unloading

The nonlinearity of the sheet stress–strain behavior is due to the nonlinear behavior of the fibers and fiber-
to-fiber bonds, and could be modeled by either allowing the fiber or the bond to yield first. Since, in this study
the later case is used to demonstrate the method, the model parameters are reduced to j, Ef, Gb, sp, and G2.
The model parameters were determined by a curve fitting procedure using uniaxial stress–strain data for the
material, in this case, a bleached softwood kraft board paper cup stock material (density = 720 kg/m3, thick-
ness = 0.48 mm). Tests were conducted on an InstronTM tensile tester with rectangular samples,
25.4 mm · 177.8 mm, at a displacement rate of 25.4 mm/min. In order to determine the orientation parameter
j, and the lateral contraction ratio m, the samples were first relieved from drying induced stresses by exposing
to high humidity (85% RH, 23 C) for 24 h and later conditioning the samples in a TAPPI standard testing
conditions (50% RH, 23 ± 1.0 C) for 24 h. The ratio of EMD/ECD of these samples measured through uniaxial
tests is then a function of j and the Poisson’s ratio m, and the relationship is given by (Wang, 2000)
R p=2  2 
EMD p=2
cos h  mMD sin2 h cos2 hf ðhÞdh
¼ R p=2  2  ð34Þ
ECD sin h  mCD cos2 h sin2 hf ðhÞdh
p=2

The orientation distribution f(h) is a function of the orientation parameter j Perkins, 1986). A family of EMD/
ECD versus j (0.0 6 j 6 1.0) curves were generated for a range of discrete values for m 0.06m61.0) using Eq.
(34). For the experimentally measured value of EMD/ECD (2.06), the range of possible values for j were deter-
mined from the plots and the average value was chosen, in this case, 0.6. Once the value of j is determined, the
value of lateral contraction ratio is determined by interpolation (Wang, 2000). In the elastic region, the elastic
modulus is a function of fiber and bond elastic properties. In a sensitivity study by Sinha (1994), Ef was found
to have a strong influence the sheet elastic modulus, while Gb had negligible effect. Meanwhile, the sheet yield
stress was dependent on sp and the slope of the stress–strain curve in the plastic region was influenced by G2.
The values for these parameters were determined using the procedure discussed by Wang (2000) through curve
fitting uniaxial tensile stress–strain curves, once the elastic parameters were determined. The hardening param-
eters (due to drying restraint effects HEf0, HGb0,Hsp0, HG20) could be obtained in similar manner using the
experimental stress–strain data from samples that without relieving drying stresses. Table 1 shows the list of
parameters and their values chosen through this procedure for the simulation.
Experiments were carried out by first loading the sample into the plastic regime and unloading subsequently
at a rate of 25.4 mm/min. The displacements at zero force are difficult to obtain in a tensile tester because the
samples tend to buckle when nearing zero force. Therefore, the displacement at zero force was approximated
by using the displacement when the applied load is close to zero.
M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632 7625

Table 1
Experimentally determined and estimated material parameters for UMAT input
Parameter Values
j 0.6
l 0 (MD)
axM. ayD 0.045, 0
exD, evD 0,0
Ef0 18 GPa
HEf0 0
Gbo 15 MPa
HGbo 0
sp0 6.5 MPa
Hsp0 2.0
Gb2 10 kPa
HG20 70

5.2. Asymptotic fiber and bond model calculations

The uniaxial loading–unloading problem can be simulated directly using the asymptotic fiber and bond
model. Since we assume that unloading is elastic, with a slope equal to the initial elastic modulus, the compu-
tation of residual stress is quite straight forward and can be computed as follows.
For a uniaxially strained sheet in the x-direction, Eqs. (8), (10) and (14) can be combined to determine the
sheet stress as:
Z  
qs p=2   tanhðaLÞ
rx ¼ E f ex cos2 h  m sin2 h 1 cos2 hf ðhÞdh ð35Þ
qf p=2 aL

where m is Poisson’s ratio and the sheet elastic modulus is given by


Z  
q p=2  2
 tanhðaLÞ
Ex ¼ s Ef 2
cos h  m sin h 1 cos2 hf ðhÞdh ð36Þ
qf p=2 aL

where Ex is the elastic modulus of the sheet. The asymptotic fiber and bond model described in Section 3.2 is
used to generate the uniaxial tensile loading portion of the stress–strain curve. Upon unloading the sheet from
the nonlinear portion of the stress–strain curve, the sheet unloads elastically (elastic fibers and bonds). Then,
the elastic strain recovered by the sheet at the end of unloading would be given by the sheet stress at the point
of unloading divided by the sheet elastic modulus, given by
R p=2
f ðhÞ cos2 hf ðhÞdh
r p=2
ex;el ¼ R p=2  
ð37Þ
Ef p=2 cos2 h  m sin2 h 1  tanhðaLÞ
aL
cos2 hf ðhÞdh

The integrals in Eqs. (35)–(37) are evaluated numerically using multiple applications of Simpson’s 1/3 rule.
Once the elastic recovered strain is obtained, the non-recoverable plastic strain, epl, is given by
epl ¼ etotal  eel ð38Þ

MD and CD tensile elastic–plastic unloading experiments described above for paper cup stock material
were simulated using the asymptotic fiber and bond model. The results are shown with the experimental curves
in Figs. 6 and 7. The deviation between the predicted plastic residual strain and the experimental values varies
7626 M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632

3.5E+07
model
experimental
3.0E+07

2.5E+07
MD Stress, Pa

2.0E+07

1.5E+07

1.0E+07

5.0E+06

0.005 0.01 0.015


MD Strain
Fig. 6. MD unloading curve and residual strain from model and experiment. The unloading was done at a target strain value of
0.01726 m/m.

from 5% to 13% for the MD curves and between 1% and 2% for the CD curves. The measured plastic strains at
zero stress are all higher than the model predictions as the experiments did not take the sample down to zero
stress. The asymptotic model predicts uniaxial unloading effects satisfactorily.

5.3. Finite element simulation of uniaxial unloading

The unloading algorithm implemented within UMAT in ABAQUSTM was used to simulate the uniaxial ten-
sile test. The sample was modeled using a 4-node bilinear plane stress element (CPS4). Each node of this ele-
ment has two degrees of freedom, i.e. ux and uy. The nodes on the left boundary were constrained in the two
degrees of freedom while the right boundary was displaced (cf. Fig. 8) by the amount of the average MD dis-
placement obtained in experiments, i.e. 3.120 · 103 m.
For unloading, the right boundary was displaced in the opposite direction until the average experimental
displacement at unloading was reached, i.e. 1.752 · 103 m. The results from the center element on the right
boundary were used to obtain the stress–strain curves. A plot of the 100-node model solution and comparison
with the closed form asymptotic fiber bond model solution are shown in Fig. 9. Results show excellent agree-
ment with the asymptotic fiber and bond calculation results, validating the finite element implementation of
the incremental asymptotic fiber bond model.

5.4. Biaxial loading–unloading

In order to subject the paper sample to a state of biaxial loading, we used the standard Mullen burst tester.
In this test, a paper sample is gripped between two circular disks (like two flat washers) and a hydraulic piston
moves a fluid against a neoprene diaphragm contacting the circular unsupported sample area in the disk
thereby increasing the pressure until the sample fails. The maximum pressure is recorded as the burst strength.
The standard Mullen tester, shown in Fig. 10, was instrumented to measure the sample central deflection as
a function of applied pressure. A schematic of our test setup is shown in Fig. 11. The measurement of central
deflection of the sample presents a difficulty due to limited space available between the sample and the clamp.
A 12.7 mm · 12.7 mm metal foil target was placed in the center on the sample with a double sided adhesive
M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632 7627

2E+07

model
1.8E+07 experimental

1.6E+07

1.4E+07
CD Stress, Pa

1.2E+07

1E+07

8E+06

6E+06

4E+06

2E+06
0.01 0.02 0.03
CD Strain
Fig. 7. CD unloading curve and residual strain from model and experiment. The unloading was done at a target strain value of
2.5 · 102 m/m.

Fig. 8. The tensile test sample was modeled with 100 4-node bilinear plane stress elements in ABAQUSTM.

tape. A capacitance sensor (HPB-500 Button Probe, by Capacitec), in the form of a circular disk, measuring
12.7 mm in diameter, was placed above the target. The sensor measures change in capacitance caused by a
change in distance between the sensor and the metal target. The output signal is converted to voltage by
the amplifier and measured by a data acquisition system (CIO-DAS1400 DAQ board from Computer Boards,
Inc., LabTech NotebookTM data acquisition software, and a PC). The hydraulic pressure is measured with
using an in-line pressure transducer in the hydraulic circuit and interfaced to the same data acquisition system.
The output voltage of the pressure transducer varies linearly with the applied pressure. In the experimental
work, the sample was clamped down in the burst tester. The 3.0-cm diameter sample area was subjected to
the applied pressure and the displacement data was collected.

5.5. Finite element simulation of mullen burst loading–unloading test

The test area was modeled with 96 shear deformable shell elements in ABAQUS. The innermost ring was
made up of 6-node triangular thin shell elements (STRI65), while the rest were 8-node doubly curved thin shell
elements (S8R5). The boundary of the model was constrained in the displacement degrees of freedom only, i.e.
ux,uy, and uz. Although the sample was clamped down during the test, the rotational degrees of freedom were
7628 M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632

Fig. 9. Asymptotic fiber bond model and ABAQUS finite element implementation comparison, MD orientation.

Fig. 10. Standard mullen burst tester and view of the clamp area.

not restrained in the finite element analysis to simulate correctly the typical central failure location in a burst
test. The shell elements used in ABAQUS, STRI65 and S8R5, require shear elastic moduli values to calculate
the transverse shear stiffness. Since the asymptotic fiber and bond model does not predict out-of-plane shear
properties, experimental values obtained for the samples using ultrasonic measurements (Baum et al., 1981).
For the coated paper cup stock used in this work, we determined G23 = 0.109 GPa and G13 = 0.119 GPa. The
corresponding transverse stiffness are given by
5 5
k ts11 ¼ G13 t; k ts22 ¼ G23 t; k ts12 ¼ 0:0 ð39Þ
6 6
where t is the average sample thickness (0.48 mm). Loading unloading experiments and simulations were car-
ried out at four different maximum pressure values, namely, 0.49, 0.51, 0.54, and 0.72 MPa. Results are shown
in Fig. 12.
M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632 7629

Fig. 11. Test setup for measuring pressure–deflection curve with a Mullen tester.

Fig. 12. ABAQUS FEA model with incremental form of asymptotic fiber bond material model compared with mullen burst test.
7630 M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632

From Fig. 12, it can bee seen that the agreement between FEA simulation and the experimental results
are not very good, although a general shape of a nonlinearly increasing pressure versus deflection behavior
and a residual deflection upon unloading are evident. However, several important observations can be made
to explain the discrepancies. Historically, Mullen burst test is not the most reproducible test. It was selected
because it offers a quick and convenient way to conduct a reasonable approximation to a standard biaxial
test, and it simulates some end use applications for paperboard. The hydraulic loading as indicated by the
pressure of the fluid is only an indirect way to measure the force applied to the sample. The differences in
measured pressure of 0.49, 0.51, and 0.55 MPa are not significant in a burst tester, and should be considered
essentially the same test. While the first two experimental curves (0.49 and 0.51 MPa) are sufficiently close in
terms of peak central deflection (about 2.3 mm), and residual deflection after unloading (about 1 mm), the
third test in the same group (Peak pressure = 0.55 MPa), exhibited an anomalous behavior; the peak deflec-
tion was 2.6 mm and the residual deflection was 1.8 mm, exceeding that of the higher pressure test at
0.72 MPa. The highest pressure test at 0.72 MPa measured a peak deflection of 3 mm and residual deflection
of 1.75 mm. The model results, on the other hand, were consistent with one another; the residual deflections
were 1.65, 1.7, 1.8, and 2.4 mm, for pressures 0.49, 0.51, 0.54, and 0.72 MPa peak pressures, respectively.
Thus, the burst test itself contributes significantly to the discrepancy. The model parameters were deter-
mined by curve fitting uniaxial tensile test data. Inaccuracies in estimating parameters, particularly, the
bond shear modulus and bond hardening slope that the inelastic portion is sensitive to (Sinha, 1994), will
contribute to the discrepancy. The boundary condition used in simulating the Burst test (rotation uncon-
strained) to simulate correctly the failure location (high stress) near the center of the sample and not at
the boundary, makes the sample more compliant in the model and produce a pressure deflection curve that
lags the experimental curve as observed. The unloading part of the curve is highly nonlinear towards the
end. This feature is not captured in the model due to not including time-dependent effects. The experimen-
tally obtained pressure–displacement curves show a step increase before 0.1 MPa. The specific cause of this
step increase is unclear. In spite of these experimental artifacts, and model parameter estimation inaccura-
cies, the incremental computational approach has shown promising results in the one-dimensional and two-
dimensional unloading applications. The model can be used in parametric studies relating microstructural
parameters to elastic–plastic response in unloading in applications such as creasing, bending and scoring
of paperboard.

6. Summary and conclusions

The asymptotic fiber and bond model for the elastic–plastic behavior of paper has been extended to
include the unloading behavior and cyclic loading. The incremental form of the asymptotic fiber and
bond model and the unloading algorithm have been implemented in ABAQUS finite element code
through the UMAT subroutine. Uniaxial loading–unloading tests were carried out with a paper cup stock
material (density 720 kg/m3, t = 0.48 mm.). The asymptotic fiber–bond model was used to predict the
unloading behavior, specifically, the stress–strain curve and residual plastic strain after unloading. The
finite element code with the incremental form of the asymptotic model was also used to predict the
unloading curve and the residual strain under uniaxial loading. The discrepancy between the experimental
measurements and asymptotic model prediction was between 5% and 13% in MD tests and between 1%
and 2% in CD tests. The finite element solution agrees well with the numerical solution of the asymptotic
model. Mullen burst tester was used to obtain the pressure versus central deflection in a biaxial loading–
unloading test for the cup stock paper material. The finite element analysis with the incremental form of
the asymptotic fiber and bond material model shows some promise with the methodology, but also shows
significant discrepancies with experimental results. The discrepancies are attributed to poor reproducibility
of the Mullen burst test, inaccuracies in the model parameter estimation, and boundary conditions in
FEA simulation, and not accounting for time-dependent effects in the model. However, the model results
themselves were consistent among each other when tested at different maximum pressures. In spite of
this, the finite element implementation of the incremental form of the asymptotic fiber and bond model,
together with the unloading algorithm, can be used for modeling complex geometries subjected to loading
and unloading.
M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632 7631

Acknowledgements

This research was supported in part by unrestricted a gift from James River Corporation (now Georgia-
Pacific Corporation). The authors thank Professor Richard Perkins of Syracuse University for insightful dis-
cussions and suggestions throughout the course of this study, and his valued friendship.

References

ABAQUSTM, 1998. User-defined mechanical material behavior, Section 12.8.1, 25.2.30, Version 5.8, Hibbitt, Karlsson & Sorensen,
Pawtucket, RI, USA.
Batten, G.L., Nissan, A.H., 1987. Unified theory of the mechanical properties of paper and other H-bond-dominated solids 3. Tappi
Journal 70 (11), 137–140.
Baum, G.A., Brenna, D.C., Habeger, C.C., 1981. Orthotropic elastic constants of paper. Tappi 64 (8), 97–101.
Bronkhorst, C.A., 2003. Modelling paper as a two-dimensional elastic–plastic stochastic network. Int. J. Solids Struct. 40, 5441–5454.
Castro, J., Ostoja-Starzewski, M., 2003. Elasto-plasticity of paper. Int. J. Plast. 19, 2083–2098.
Corte, H. Kallmes, O. J. 1962. Statistical geometry of a fibrous network. In: Bolam, F. (Ed.), Formation and structure of Paper, vol. 1,
Tech. Sect. Brit. Paper and Board Makers Association, London, pp. 13.
Cox, H.L., 1952. The elasticity and strength of paper and other orthotropic fibrous materials. Brit. J. Appl. Phys. 3, 72–79.
Doghri, I., Tinel, L., 2006. Micromechanics of inelastic composites with misaligned inclusions: Numerical treatment of orientation.
Comput. Methods Appl. Mech. Eng. 195, 1387–1406.
Furukawa, I., Mark, R.E., Crosby, C.M., Perkins, R.W., 1991. Inelastic behavior of machine-made paper related to its structural changes.
Japan Tappi J. 45 (5), 582–590.
Heyden, S., Gustafsson, P.J., 2001. Stress–strain performance of paper and fluff by network modeling, In: Proc. 12th Fundamental
Research Symposium, Oxford, September 2001, pp. 1385–1401.
Johnson, M.W., Urbanik, T.J., 1984. A nonlinear theory for plastic plates with application to characterizing paper properties. J. Appl.
Mech. 51 (3), 146–152.
Kang, G., Gou, S., Deng, C., 2006. Numerical simulation for uniaxial cyclic deformation of discontinuously reinforced metal matrix
composites. Mater. Sci. Eng. A 426, 66–76.
Karafillis, A.P., Boyce, M.C., 1993. A general anisotropic yield criterion using bounds and a transformation weighting tensor. J. Mech.
Phys. Solids 41 (12), 1859–1882.
Komori, T., Itoh, M., 1991. Theory of the general deformation of fiber assemblies. Text. Res. J 61 (10), 588–594.
Komori, T., Makishima, K., 1977. Numbers of fiber-to-fiber contacts in general fiber assemblies. Text. Res. J. 47 (1), 13–17.
Lu, Wentao, Carlsson, Leif A., de Ruvo, Alf, 1996. Micro-model of paper. Tappi J. 79 (2), 197–205.
Makela, P., Ostlund, S., 2003. Orthotropic elastic–plastic material model for paper materials. Int. J. Solids Struct. 40, 5599–5620.
Niskanen, K., Karenlampi, P., 1998. In-plane tensile properties, In: Paper Physics, Book 16, Papermaking Science and Technology, A
series of 19 Books, Niskanen, K. (Ed.), published by Fapet Oy, PO Box 146, FIN-00171, Helsinki, Finland, 1998, pp. 138–191.
Page, D.H., Seth, R.S., De Grace, J., 1979. Elastic modulus of paper (1). controlling mechanisms. Tappi J. 62 (9), 99–102.
Perkins, R.W., 1986. Fiber Networks (Models for Predicting Mechanical Behavior), Encyclopedia of Materials Science and Engineering,
Bever, M.B. (Ed.)-in-Chief, Pergamon Press, 1986, pp. 1712–1719.
Perkins, R.W., 1990. Micromechanics Models for Predicting the Elastic and Strength Behavior of Paper Materials, Materials Interactions
Relevant to Pulp, Paper, & Wood Ind. (Caulfield, Passaretti, & Sobczynski, ed.)/Materials Res. Soc. Symp. (San Francisco) Proc. vol.
197, April 18–20, 1990, pp. 99–118.
Ramasubramanian, M.K. 1987. Computer simulation of the uniaxial stress–strain behavior of ribbonlike fiber Nonwovens, Syracuse
Univ. Ph.D. Thesis, 1987, 203p
Ramasubramanian, M.K., Perkins, R.W., 1988. Computer simulation of the uniaxial elastic–plastic behavior of paper. J. Eng. Mater.
Technol. 110 (2), 117–123.
Ramasubramanian, M. K., Wang, Y.Y., 1999. Constitutive models for paper and other ribbon-like nonwovens – a literature review. In:
Perkins, R. (Ed.) Mechanics of Cellulosic Materials, AMD-vol. 231, MD-vol. 85, Am. Soc. Mech. Eng., New York, pp. 31–42.
Schulgasser, K., Page, D.H., 1988. The influence of transverse us properties on the in-plane elastic behavior of paper. Composites Sci.
Technol. 32, 279–292.
Schulgasser, K., Grunseit, Z., 1990. The influence of micro-formation on the inplane elastic behavior of paper. J. Mater. Sci. 25 (5), 2433–
2440.
Sinha, S.K., 1994. Viscoplastic Micromechanics and Continuum Models of Fibrous Cellulosic Materials. Syracuse University, 322 p.
Sinha, S.K., Perkins, R.W., 1995. Micromechanics constitutive model for use in finite element analysis, In: Proc. 1995, Joint ASME Appl.
Mech. Mater. Summer Meeting, Los Angeles, CA, USA, Jun 28–30, 1995.
Ostoja-Starzewski, M., Stahl, D.C., 2001. Random fiber networks and special elastic orthotropy of paper. J. Elasticity 60, 131–149, 2000.
Stenberg, N., 2003. A model for the through-thickness elastic–plastic behaviour of paper. Int. J. Solids Struct. 40, 7483–7498.
Suhling, J.C., 1990. Continuum Models for the Mechanical Response of Paper and Paper Composites: Past, Present, and Future,
Materials Interactions Relevant to Pulp, Paper, & Wood Ind. (Caulfield, Passaretti, & Sobczynski, ed.)/Materials Res. Soc. Symp. (San
Francisco), Proc. vol. 197, April 18–20, 1990, pp. 245–255.
7632 M.K. Ramasubramanian, Y. Wang / International Journal of Solids and Structures 44 (2007) 7615–7632

Van den Akker, J.A., 1970. Structure and textile characteristics of papers. Tappi 53 (3), 388–400.
Wang, Y.Y. 2000. Constitutive Modeling of the Unloading Behavior of Paper Material Using the Asymptotic Fiber and Bond Model,
Ph.D. Dissertation, Mechanical and Aerospace Engineering, NC State University, NC, USA.
Xia, Q.S., Boyce, M.C., Parks, D.M., 2002. A constitutive model for the anisotropic elastic–plastic deformation of paper and paperboard.
Int. J. Solids Struct. 39 (15), 4053–4071.
Xie, Z., Gilliksson, M., Hagglund, R., 2002. Determining the elastic constants of paper with optimization methods. Inverse Problems Eng.
10 (5), 393–411.

S-ar putea să vă placă și