Sie sind auf Seite 1von 14

This article was downloaded by: [200.17.114.

136]
On: 12 February 2014, At: 04:49
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Mechanics of Advanced Materials and Structures


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/umcm20

A Model for Viscoelastic Heterogeneous Materials Based


on the Finite-Volume Theory
a a b
Romildo S. Escarpini Filho , Severino P. C. Marques & G. J. Creus
a
LCCV/Center of Technology, Federal University of Alagoas , Maceio , Brazil
b
CEMACOM-UFGRS , Porto Alegre , Brazil
Accepted author version posted online: 02 Aug 2012.Published online: 30 Jan 2014.

To cite this article: Romildo S. Escarpini Filho , Severino P. C. Marques & G. J. Creus (2014) A Model for Viscoelastic
Heterogeneous Materials Based on the Finite-Volume Theory, Mechanics of Advanced Materials and Structures, 21:5, 349-361,
DOI: 10.1080/15376494.2012.680800

To link to this article: http://dx.doi.org/10.1080/15376494.2012.680800

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Mechanics of Advanced Materials and Structures (2014) 21, 349–361
Copyright 
C Taylor & Francis Group, LLC

ISSN: 1537-6494 print / 1537-6532 online


DOI: 10.1080/15376494.2012.680800

A Model for Viscoelastic Heterogeneous Materials Based


on the Finite-Volume Theory
ROMILDO S. ESCARPINI FILHO1, SEVERINO P. C. MARQUES1, and G. J. CREUS2
1
LCCV/Center of Technology, Federal University of Alagoas, Maceio, Brazil
2
CEMACOM-UFGRS, Porto Alegre, Brazil

Received 7 December 2010; accepted 25 July 2011.


Downloaded by [200.17.114.136] at 04:49 12 February 2014

This article presents a new numerical model for the analysis of structures of heterogeneous materials with linear viscoelastic con-
stituents. The model is based on the recently developed parametric finite-volume theory that has produced a paradigm shift in the
finite-volume theory’s development. This parametric formulation is here extended to model linear viscoelastic behavior. The present
model employs the state variables approach for the computation of the time-dependent strains. Several examples, including both ho-
mogeneous and heterogeneous situations, are analyzed. Comparison between the numerical and analytical results shows the excellent
performance of the proposed model.
Keywords: viscoelasticity, heterogeneous materials, parametric formulation, finite-volume theory, state variable approach

1. Introduction behavior of the existing micro-heterogeneous composite in an


average sense.
The application of heterogeneous materials to fill the needs of Approximate constitutive relations may be obtained using
diverse industrial sectors has significantly increased in the last micromechanical models that do not explicitly account for
few years. Such materials may be tailored to present improved local microstructural details and interaction. Many of these
properties (stiffness, resilience, thermal capacity, damping, models are derivatives of the classical Eshelby inclusion prob-
etc.) not exhibited by their individual constituents. Important lem [5] and are often referred to as mean-field approaches.
examples are the functionally graded materials and advanced Such micromechanical schemes have been originally derived
fiber reinforced composites [1]. Many of these composites are to estimate the effective elastic properties and present as ad-
constituted by high performance fibers (carbon, glass, metal, vantage their analytical or semi-analytical framework, which
ceramic, etc.) embedded in a polymeric matrix. Due to the facilitates quick generation of data, albeit at the expense of
time-dependent behavior of polymers, the composites exhibit the loss of microstructure fidelity [6]. For the case of effective
an effective viscoelastic response that is affected by environ- properties of composites with linear viscoelastic phases, they
mental agents, particularly temperature, humidity, and age are usually applied together with the well-known correspon-
[2–4]. Therefore, prediction of the long-term behavior is of dence principle [7]. For instance, Laws and McLaughlin [8]
utmost importance for the design and analysis of many struc- and Brinson and Lin [9] employed this procedure using, re-
tures made of polymeric composite materials, such as aircraft spectively, the self-consistent and the Mori-Tanaka schemes
components and wind turbine blades, which require the con- to viscoelastic composites. Hashin-Shtrikman approach has
sideration of these time-dependent effects. also been used to obtain bounds for viscoelastic composites
On the other hand, composites show a heterogeneous tex- moduli [10, 11]. Despite the conceptual simplicity, the use of
ture on the microlevel, determined by the constitutive behavior these procedures based on the correspondence principle has
of the matrix material and the embedded fibers. Usually, nu- several limitations. For instance, in many cases of interest, the
merical analyses of structural elements on the macroscale as- effective moduli of the composites in the Laplace domain are
sume a homogeneous continuum with the average composite not available in closed-form and cannot be easily inverted to
properties. Therefore, the constitutive equations of the substi- the time-domain [12]. This problem has been solved through
tute homogenized material should well reflect the mechanical the use of computational techniques [13].
Traditionally, the finite element method is the numerical
technique more often employed for the analyses of compos-
Address correspondence to Severino P. C. Marques, Center ites, particularly to understand their macro-level behavior [14,
of Technology, Federal University of Alagoas, Campus A. 15]. In many cases, the problems involving the two differ-
C. Simoes, Maceio 57072-970, Brazil. E-mail: smarques@lccv. ent scales (micro and macro) receive uncoupled treatments.
ufal.br However, this procedure has as a disadvantage the inability
350 R. S. Escarpini Filho et al.

to merge the micromechanical details directly into the macro- by using the correspondence principle. The comparison be-
level analysis. To have a more complete understanding, for tween the numerical and analytical results shows an excellent
instance, on the viscoelastic behavior of a composite, the in- performance of the proposed model.
fluence of the heterogeneous nature at the microscale on the The article is organized as follows. In Section 2, the para-
macroscale response must be examined in details. metric viscoelastic formulation of the finite volume theory
Lately, computational tools other than finite elements have is described for the cases of plane stress and plane strain.
been proposed to model heterogeneities and their effects un- Section 3 describes the procedures employed to evaluate the
der a variety of thermo-mechanical conditions [16–18]. An viscoelastic strain using the state variables formulation. Sec-
attractive alternative technique is the recently developed para- tion 4 presents the results obtained for several examples and
metric formulation of the finite-volume theory [19], which comparisons of them with analytical and numerical solutions.
incorporated a parametric mapping capability into the FV-
DAM (Finite-Volume Direct Averaging Method) [18]. In the
standard version of FVDAM, the material microstructure 2. Parametric Viscoelastic Formulation of the
is discretized into a grid of rectangular subvolumes. Due
Finite-Volume Theory–Plane Stress and Plane Strain
to the rectangular geometry of the subvolumes, the model-
Problems
ing of microstructures and geometries with curved features
using that standard version, in general, requires extensive
2.1. Local Incremental Equilibrium Equation
discretization in order to reduce the artificial stress concen-
Downloaded by [200.17.114.136] at 04:49 12 February 2014

trations at the boundaries and material interfaces. However, In the parametric finite-volume theory, the whole domain
these limitations have been overcome by the parametric for- occupied by the heterogeneous material is discretized into
mulation of the finite-volume theory, which enables the use quadrilateral subvolumes [19]. The formulation is based on the
of quadrilateral subvolumes in approximating the heteroge- mapping of a reference square subvolume onto each quadri-
neous microstructures, allowing a more efficient modeling of lateral subvolume, as shown in Figure 1. The mapping of the
microstructural details with curvilinear geometry [19]. This point (␩, ␰) in the reference subvolume to the corresponding
capability of the parametric formulation for modeling curved point (x, y) in the actual subvolume is given by:
boundaries and interfaces in material with heterogeneous mi-
crostructures has been already demonstrated in several recent x (␩, ␰) = N1 (␩, ␰) x1 + N2 (␩, ␰) x2 + N3 (␩, ␰) x3
publications [19–24]. + N4 (␩, ␰) x4 ,
This parametric formulation of the finite-volume theory, y (␩, ␰) = N1 (␩, ␰) y1 + N2 (␩, ␰) y2 + N3 (␩, ␰) y3
initially formulated for thermo-mechanical analysis of linear
+ N4 (␩, ␰) y4 , (1)
elastic composite materials [19], has been extended to include
plastic effects [22] and also used as framework for the de-
velopment of an interesting homogenization technique ap- where Ni are the shape functions given by:
plied to materials with periodic heterogeneous microstructures
1 1
[23, 24]. N1 (␩, ␰) = (1−␩) (1 − ␰) , N2 (␩, ␰) = (1+␩) (1 − ␰) ,
A systematic comparison of the convergence characteris- 4 4
tics and processing times of the parametric formulation of 1 1
N3 (␩, ␰) = (1+␩) (1 + ␰) , N4 (␩, ␰) = (1 − ␩) (1 + ␰) .
the finite-volume theory and the finite-element method is ad- 4 4
dressed in [21]. To have a meaningful comparison of the execu- (2)
tion times consumed by the two methods in the analyses, the
study was developed using similar conditions of computing The displacement field at time t is approximated using a
environmental, machine, matrix assemblage and solver, etc. second-order expansion in the reference square subvolume
For the problems analyzed in this comparative investigation, coordinates as follows:
the parametric formulation of finite-volume theory showed
1
an excellent performance with respect to the convergence and, u 1 (t) = U1(00)
t
+ ␩U1(10)
t
+ ␰U1(01)
t
+ (3␩2 − 1)U1(20)
t
particularly, in processing times. Another detailed and direct 2
comparative study on convergence of the parametric FVDAM 1
+ (3␰ 2 − 1)U1(02)
t
,
and finite–element predictions is found in [24]. 2
Considering the previous and promising results provided
by the parametric formulation of the finite-volume theory, the (-1,1) (1,1) (x3,y3)
authors developed the present novel numerical model for the F3 (x2,y2)
analysis of heterogeneous materials with linear viscoelastic F4
constituents. The model has been derived on the parametric y (x4,y4)
F2
finite-volume theory framework and employs a state variables F1
(-1,-1) (1,-1) (x1,y1)
formulation for the computation of the viscoelastic strains x
[25].
Several examples, including both homogenous and hetero- Fig. 1. Mapping of the reference square subvolume in the ␩ − ␰
geneous materials, are analyzed and their results are compared plane onto a quadrilateral subvolume in the x − y plane of the
with analytical solutions found in the literature or are obtained actual microstructure.
Model for Materials with Viscoelastic Constituents 351
⎧ ⎫
1 ⎪ ∂u 1 ⎪ 
u 2 (t) = t
U2(00) + ␩U2(10)
t
+ ␰U2(01)
t
+ (3␩2 − 1)U2(20)
t ⎪
⎪ ⎪
⎪ 
2 ⎪
⎪ ∂␩ ⎪
⎪ 

⎪ ⎪
⎪ 
1 ⎪
⎪  ⎪
⎪  ⎡ ⎤
+ (3␰ 2 − 1)U2(02)
t
, (3) ⎪
⎪ ∂u 1 ⎪
⎪  1 0 0 0 0 0 0 0
2 ⎪
⎪ ⎪
⎪ 
⎨ ∂␰ ⎬ ⎢0 1 0 ∓3 0 0 ⎥
 ⎢ 0 0 ⎥
  =⎢ ⎥
where Uit(kl) are the polynomial coefficients. Similarly, the in- ⎪
⎪ ∂u 2 ⎪
⎪  ⎣0 0 0 0 1 0 0 0 ⎦

⎪ ⎪
⎪ 
cremental displacement field, corresponding to time interval ⎪
⎪ ∂␩ ⎪
⎪  0 0 0 0 0 1 0 ∓3

⎪ ⎪
⎪ 
[t, t + t] is given by the following second-order approxima- ⎪
⎪  ⎪
⎪ 

⎪ ∂u 2 ⎪
⎪ 
tions: ⎪
⎩ ⎭

∂␰ k=1,3
⎧ ⎫
1 ⎪ U1(10) ⎪
u 1 = U1(00) + ␩U1(10) + ␰U1(01) + (3␩2 − 1)U1(20) ⎪
⎪ ⎪

2 ⎪
⎪ U1(01) ⎪


⎪ ⎪

1 ⎪
⎪ U ⎪

+ (3␰ 2 − 1)U1(02) , ⎪
⎪ 1(20) ⎪

2 ⎪
⎨ U ⎪

1 1(02)
u 2 = U2(00) + ␩U2(10) + ␰U2(01) + (3␩2 − 1)U2(20) × . (6)

⎪ U2(10) ⎪

2 ⎪
⎪ ⎪


⎪ ⎪
1
+ (3␰ 2 − 1)U2(02) . (4) ⎪ U2(01) ⎪
⎪ ⎪

⎪ ⎪
Downloaded by [200.17.114.136] at 04:49 12 February 2014

2 ⎪
⎪ U2(20) ⎪


⎩ ⎪

U2(02)
The surface-averaged partial derivatives of the incremental
displacement components on each Fk (k = 1, 2, 3, 4) of the
quadrilateral subvolume (Figure 1) are defined in the form:
The surface-averaged partial derivatives of the incremen-
   tal displacement field with respect to the Cartesian coordi-
∂u i  1 1 ∂u i
= d␩ , nates are related to those derivatives given in Eq. (6) by the
∂␩ k=1,3 2 −1 ∂␩ expressions:
  
∂u i  1 1 ∂u i
= d␩,
∂␰ k=1,3 2 −1 ∂␰ ⎧ ⎫ ⎧ ⎫
   ⎪ ∂u 1 ⎪ ⎪ ∂u 1 ⎪ 
∂u i  1 1 ∂u i ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ 
= d␰, ⎪
⎪ ∂x ⎪
⎪ ⎪
⎪ ∂␩ ⎪
⎪ 
∂␩ k=2,4 2 −1 ∂␩ ⎪

⎪ 







⎪ 





   ⎪
⎪ ∂u 1 ⎪
⎪ ⎪
⎪ ∂u 1 ⎪
⎪ 

⎪ ⎪
⎪  ⎪
⎪ ⎪
⎪ 
∂u i  1 1 ∂u i ⎪
⎨ ∂y ⎪
⎬ ⎪ ⎪
= d␰. (5)  Jˆ 0 ⎨ ∂␰ ⎬

∂␰ k=2,4 2 −1 ∂␰   =   ,

⎪ ∂u 2 ⎪
⎪ 0 ⎪ ∂u 2
Jˆ ⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ 

⎪ ∂x ⎪
⎪ ⎪
⎪ ∂␩ ⎪

Considering Eqs. (4) and (5), the following matrix relations ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ 

⎪  ⎪
⎪ ⎪
⎪  ⎪

can be readily derived: ⎪
⎪ ∂u 2 ⎪
⎪ ⎪
⎪ ∂u 2 ⎪
⎪ 

⎩ ⎪
⎭ ⎪
⎩ ⎪

∂y  ∂␰ 
⎧ ⎫ k=2,4 k=2,4
⎪ ∂u 1 ⎪  ⎧ ⎫ ⎧ ⎫

⎪ ⎪
⎪   

⎪ ∂␩ ⎪
⎪  ⎪
⎪ ∂u 1 ⎪
⎪ ⎪
⎪ ∂u 1 ⎪
⎪ 

⎪ ⎪
⎪  ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ 

⎪  ⎪
⎪  ⎡ ⎤ ⎪
⎪ ∂x ⎪
⎪ ⎪
⎪ ∂␩ ⎪
⎪ 

⎪ ∂u 1 ⎪
⎪  1 0 ±3 0 0 0 0 0 ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ 

⎪ ⎪
⎪  ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ 
⎨ ∂␰ ⎬

⎢0
⎢ 1 0 0 0 0 0 0⎥⎥ ⎪ ∂u 1



⎪
⎪  ⎪ ∂u 1
⎪


⎪

  =⎢ ⎥ ⎨ ∂y ⎬ Jˆ 0 ⎨ ∂␰ ⎬
⎪ ∂u 2 ⎪  ⎣0 0 0 0 1 0 ±3 0⎦  

⎪ ⎪
⎪    =   . (7)

⎪ ⎪
⎪  ⎪
⎪ ∂u 2 ⎪
⎪ 0 ⎪ ∂u 2
Jˆ ⎪ ⎪


⎪ ∂␩ ⎪
⎪  0 0 0 0 0 1 0 0 ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ 

⎪  ⎪
⎪  ⎪
⎪ ∂x ⎪
⎪ ⎪
⎪ ∂␩ ⎪


⎪ ∂u 2 ⎪
⎪  ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ 

⎪ ⎪
⎪ ⎪  ⎪ ⎪  ⎪
⎩ ⎭ ⎪

⎪ ∂u 2 ⎪

⎪ ⎪

⎪ ∂u 2 ⎪

⎪ 
∂␰ ⎪
⎩ ⎪
⎭ ⎪
⎩ ⎪

k=2,4
⎧ ⎫ ∂y  ∂␰ 
⎪ U1(10) ⎪ k=1,3 k=1,3

⎪ ⎪


⎪ U1(01) ⎪


⎪ ⎪


⎪ U ⎪


⎪ 1(20) ⎪


⎨ U ⎪
⎬ where Jˆ stands for the inverse of the volume-averaged Jacobian
×
1(02)
, J¯ defined by:
⎪ U2(10) ⎪
⎪ ⎪

⎪ ⎪
⎪ U2(01) ⎪
⎪ ⎪


⎪ ⎪


⎪ ⎪
⎪  

⎪ U2(20) ⎪
⎪ 1 +1 +1
⎩ ⎭ J¯ = Jd␩d␰, (8)
U2(02) 4 −1 −1
352 R. S. Escarpini Filho et al.
 1
being 1
u 1 F4 =  u 1 |k=4 = u 1 d␰ = U1(00) − U1(10)
⎡ ⎤ 2 −1
∂x ∂y + U1(20) ,
⎢ ∂␩ ∂␩ ⎥  1
⎢ ⎥ 1
J=⎢ ⎥. (9) u 2 F1 =  u 2 |k=1 = u 2 d␩ = U2(00) − U2(01)
⎣ ∂x ∂y ⎦ 2 −1
∂␰ ∂␰ + U2(02) ,

1 1
u 2 F2 =  u 2 |k=2 = u 2 d␰ = U2(00) + U2(10)
Using the strain-displacement relations, the surface-averaged 2 −1
incremental strains on the faces of the quadrilateral subvolume + U2(20) ,
are given by: 
1 1
u 2 F3 =  u 2 |k=3 = u 2 d␩ = U2(00) + U2(01)
⎧ ⎫ 2 −1

⎪ ∂u 1



⎪
⎪ + U2(02) ,

⎪ ∂x ⎪
⎪  

⎪ ⎪
⎪ 1 1

⎪  ⎪
⎪  u 2 F4 =  u 2 |k=4 = u 2 d␰ = U2(00) − U2(10)
⎧ ⎫ ⎪
⎪ ∂u 1 ⎪

 ⎪ ⎪  2 −1
⎨ ε11
⎪ ⎪
⎬ ⎪
⎨ ∂y ⎪
⎬
  + U2(20) , (12)
ε22  = Ē   ,
Downloaded by [200.17.114.136] at 04:49 12 February 2014


⎩ ε ⎭
⎪ ⎪
⎪ ∂u 2 ⎪
⎪ 

⎪ ⎪

12 k=2,4 ⎪
⎪ ∂x ⎪
⎪ 

⎪ ⎪
⎪ Using Eqs. (12), it can be shown that the coefficients of the in-

⎪  ⎪
⎪ 
⎪ ∂u 2
⎪ ⎪
⎪ cremental displacement expansions are related to the surface-

⎩ ⎭

∂y averaged increments of displacement components as follows
⎧ ⎫k=2,4 (i = 1, 2):
⎪ ∂u 1 ⎪ 

⎪ ⎪
⎪ 

⎪ ∂x ⎪
⎪ 

⎪ ⎪
⎪ 

⎪  ⎪
⎪  1
⎧ ⎫ ⎪ ⎪ 
 ⎪ ∂u 1
⎪ ⎪
⎪  Ui (10) = (u i F2 − u i F4 ),
⎨ ε11
⎪ ⎪
⎬ ⎪
⎨ ∂y ⎪
⎬ 2
  1
ε22  = Ē   , (10) Ui (01) = (u i F3 − u i F1 ),

⎩ ε ⎭
⎪ ⎪
⎪ ∂u 2 ⎪
⎪  2

⎪ ⎪
⎪ 
12 k=1,3 ⎪
⎪ ∂x ⎪
⎪  1

⎪ ⎪
⎪  Ui (20) = (u i F2 + u i F4 ) − Ui (00) ,

⎪  ⎪
⎪  2

⎪ ∂u 2 ⎪
⎪ 

⎩ ⎭
⎪ 1
∂y k=1,3
Ui (02) = (u i F1 + u i F3 ) − Ui (00) . (13)
2

being
For an isotropic linear viscoelastic material occupying the sub-
⎡ ⎤ volume, the increments of stress are related to increments of
1 0 0 0 strain, corresponding to time interval [t, t + t], through the
⎢0 0 0 1⎥ expression:
Ē = ⎢

⎥.
⎦ (11)
1 1
0 0
2 2 ⎧ ⎫ ⎛⎧ ⎫ ⎧ T ⎫ ⎧ V ⎫⎞
⎪ ε11 ⎪
⎨ ␴11 ⎪
⎪ ⎬ ⎜

⎨ ε11 ⎪⎬ ⎪⎨ ε11 ⎪
⎬ ⎪ ⎨ ⎪
⎬⎟
Through Eqs. (4), the following relations for the surface- ␴22 = C̄ ⎜
⎝⎪ ε 22 − ε T − ε V ⎟
22 ⎪⎠ . (14)

⎩ ␴ ⎪ ⎭ ⎩ ε ⎪ ⎭ ⎪⎩ 22 ⎪ ⎭ ⎪ ⎪ ⎪
averaged increments of displacement components on each face ⎩ ⎭
12 12 0 ε V 12
of the subvolume can be found:
 1
1 Here the indices T and V are used to indicate the increments
u 1 F1 =  u 1 |k=1 = u 1 d␩ = U1(00) − U1(01) of thermal and viscoelastic strains, respectively, and C̄ is the
2 −1
+ U1(02) , linear elastic constitutive matrix for plane stress and plane
 1 strain:
1
u 1 F2 =  u 1 |k=2 = u 1 d␰ = U1(00) + U1(10)
2 −1 ⎡ ⎤
+ U1(20) , C̄11 C̄12 0
 ⎢ ⎥
1 1 C̄ = ⎣ C̄12 C̄22 0 ⎦. (15)
u 1 F3 =  u 1 |k=3 = u 1 d␩ = U1(00) + U1(01)
2 −1
0 0 C̄33
+ U1(02) ,
Model for Materials with Viscoelastic Constituents 353

The surface-averaged incremental stress-strain relation for the where b1 and b2 are the increments in the volume forces
subvolume faces can be expressed in the form: in the time interval [t, t + t]. Using Eq. (14) and the rela-
tions strain-displacement in incremental form, the following
⎧ ⎫
 expressions are derived for the partial derivatives appearing in
⎨ ␴11 ⎪
⎪ ⎬
 Eqs. (19):
␴22 
⎩ ␴ ⎭
⎪ ⎪
12
⎛⎧k=2,4 ⎫ ⎧ ⎫ ⎧ V ⎫⎞ ∂␴11
⎪ ε T ⎪ ⎪ ε ⎪  = 3C̄ 11 [(Ĵ11 )2 U1(20) + (Ĵ12 )2 U1(02) ]
⎪ ε11 ⎪ ⎪ ⎪ ⎪ ⎬⎟
⎪ ∂x
⎜⎨ ⎬ ⎨  11 ⎬ ⎨  11

= C̄ ⎝ ε22 − ε22 T − ε22 V ⎟ , + 3C̄ 12 (Ĵ11 Ĵ21 U2(20) + Ĵ12 Ĵ22 U2(02) )
⎪ ⎪ ⎠
⎩ ε ⎭ ⎪⎪



⎭ ⎩

⎪  ⎪
⎭ 
⎪ ∂T ∂ε11
V
∂ε22
V
12 0 ε12
V
− 1 − C̄ 11 − C̄ 12 ,
⎧ ⎫ k=2,4
∂x ∂x ∂x

⎨ ␴11 ⎪
⎪ ⎬ ∂␴22
 = 3C̄ 21 (Ĵ11 Ĵ21 U1(20) + Ĵ12 Ĵ22 U1(02) )
␴22  ∂y
⎩ ␴ ⎪
⎪ ⎭
12 + 3C̄ 22 [(Ĵ21 )2 U2(20) + (Ĵ22 )2 U2(02) ]
⎛⎧k=1,3 ⎫ ⎧ ⎫ ⎧ V ⎫⎞
 ∂T ∂ε11 ∂ε22
⎪ ε11 ⎪ ε11
V V
⎪ ε11 ⎪ ⎪ T

⎪ ⎪ ⎪
⎪ 
⎜⎨ ⎬ ⎨  T ⎬ ⎨  V ⎬⎟ − 2
∂y
− C̄ 21
∂y
− C̄ 22
∂y
,
= C̄ ⎜
⎝⎪ ε22 − ε22 − ε22 ⎟ . (16)
Downloaded by [200.17.114.136] at 04:49 12 February 2014

⎪ ⎠
⎩ ε ⎭ ⎪⎪


⎭ ⎪
⎪ ⎪


⎪ 
⎭ ∂␴12
12 0 ε12
V  = 3C̄ 33 [Ĵ11 Ĵ21 U1(20) + Ĵ12 Ĵ22 U1(02) + (Ĵ11 )2 U2(20)
k=1,3 ∂x
∂ε12 V
+ (Ĵ12 )2 U2(02) ] − C̄ 33 ,
Using the Cauchy equation, the surface-averaged increments ∂x
of tractions on the faces of the subvolume are given as: ∂␴12
= 3C̄ 33 [(Ĵ21 )2 U1(20) + (Ĵ22 )2 U1(02) + Ĵ11 Ĵ21 U2(20)
⎛ ⎧ ⎫⎞ ∂y
   ⎪ ␴11 ⎪  ∂ε12 V
t1  ⎜ n1 0 n2 ⎨ ⎬ 
⎟ + Ĵ12 Ĵ22 U2(02) ] − C̄ 33 , (20)
=⎝ ␴22 ⎠ , ∂y
t 
2 k=1,2,3,4
0 n2 n1 ⎪
⎩ ␴ ⎭ 

12 k=1,2,3,4
(17) where 1 = (C̄ 11 + C̄ 12 )␣, 2 = (C̄ 12 + C̄ 22 )␣, T is the tem-
perature increment, and ␣ stands for the thermal expansion
where n 1 and n 2 are the components of the unit normal vector coefficient.
to each face Fk in the directions x and y, respectively. Through Substituting Eqs. (20) into Eqs. (19), the following relation
Eqs. (6), (7), (10), (16), and (17), the following relation for the for the zeroth order coefficients is obtained:
surface-averaged increments of traction components on the
⎧ ⎫
subvolume faces can be found: ⎪ u 1 F2 + u 1 F4 ⎪
  ⎪
⎪ ⎪

U1(00) ⎨ u + u 1 F3 ⎬
⎧ ⎫ ⎧ ⎫ 1 F1
⎪ t U1(10) ⎪ = Φ−1 Θ + Φ−1 Δ␺, (21)
⎪ 1 F1 ⎪
⎪ ⎪



⎪ ⎪
⎪ U2(00) ⎪
⎪ u + u 2 F4 ⎪


⎪ t2 F1 ⎪
⎪ ⎪
⎪ U1(01) ⎪ ⎪ ⎛⎧ ⎫ ⎧ ⎫⎞ ⎪

2 F2



⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ u 2 F1 + u 2 F3

⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ Δ␧TF1 ⎪
⎪ ⎪
⎪ Δ␧V F1 ⎪


⎪ t1 F2 ⎪
⎪ ⎪
⎪ U 1(20) ⎪
⎪ ⎜⎪⎪ ⎪ ⎪
⎪ ⎪ ⎪
⎪⎟
⎨ t ⎪
⎪ ⎬ ⎪
⎨ U ⎪
⎬ ⎜⎪⎨ Δ␧T ⎪ ⎬ ⎪ ⎨ Δ␧V ⎪ ⎬⎟
2 F2 1(02) ⎜ F2 F2 ⎟
= Ā − DC ⎜ + ⎟, where

⎪ t1 F3 ⎪
⎪ ⎪
⎪ U2(10) ⎪ ⎪ ⎜⎪⎪ Δ␧TF3 ⎪
⎪ ⎪
⎪ Δ␧V ⎪⎟
⎪ ⎪ ⎪ ⎪ ⎝ ⎪ ⎪ ⎪ F3 ⎪
⎪ ⎠

⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪
⎪ ⎪

⎪ t2 F3 ⎪




⎪ U2(01) ⎪ ⎪


⎩ T ⎭

⎩ V ⎭
⎪ ⎧ ⎫

⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ Δ␧ Δ␧ ⎪ ∂T ∂ε11
V
∂ε22
V


⎪ t ⎪ ⎪ ⎪
⎪ U2(20) ⎪ ⎪
F4 F4

⎪ b −  − C̄ − C̄ ⎪

⎩ 1 F4 ⎪
⎪ ⎭ ⎪
⎩ ⎪
⎭ ⎪

1 1
∂x
11
∂x
12
∂x ⎪

t2 F4 U2(02) ⎪
⎪ ⎪


⎪ ⎪


⎪ ∂ε12 V


(18) ⎪ − C̄ 33
⎪ ⎪

where Ā = DC E B A. The matrices appearing in the defini- 1 ⎨ ∂x ⎬
tion of Ā are presented explicitly in Appendix A. In Eq. (18), Δ␺ = ,
3⎪
⎪ ∂T ∂ε11
V
∂ε22
V ⎪

Δ␧TF k and Δ␧VFk represent the vectors whose components are ⎪
⎪ ⎪


⎪ b2 − 2 − C̄ 21 − C̄ 22 ⎪

the surface-averaged increments of thermal and viscoelastic ⎪
⎪ ∂y ∂y ∂y ⎪


⎪ ⎪

strains on Fk , respectively. The incremental equilibrium equa- ⎪
⎪ ∂ε12V ⎪


⎩ − C̄ 33 ⎪

tions can be written in the form: ∂y
(22)
∂␴11 ∂␴12 and the matrices Φ and Θ are presented in the appendix.
+ + b1 = 0,
∂x ∂y Introducing Eq. (21) into Eqs. (13), the first- and second-
∂␴21 ∂␴22 order coefficients of the displacement increments are related to
+ + b2 = 0, (19)
∂x ∂y the surface-averaged increments of displacement components
354 R. S. Escarpini Filho et al.

as follows: thermal and viscoelastic effects. The vectors Δt̄ and Δt̄ o are
⎧ ⎫ ⎧ ⎫ known at the beginning of each time incremental step. Due
⎪ U1(10) ⎪ ⎪ u 1 F1 ⎪ to the interfacial traction continuity conditions on the com-

⎪ ⎪
⎪ ⎪
⎪ ⎪


⎪ U1(01) ⎪
⎪ ⎪
⎪ u 2 F1 ⎪
⎪ mon interface of adjacent subvolumes, Δt̄ is sparse, with its

⎪ ⎪
⎪ ⎪
⎪ ⎪

⎪ U
⎪ ⎪
⎪ ⎪
⎪ u ⎪
⎪ nonzero terms being the increments of surface-averaged trac-

⎪ 1(20) ⎪
⎪ ⎪
⎪ 1 ⎪


⎨ ⎪
⎬ ⎪

F2
⎪ tions on the discretized boundary regions.
U1(02) u 2 F2 ⎬
= B̄ − NΦ−1 Δ␺, (23)

⎪ U ⎪
⎪ ⎪
⎪ u ⎪
F3 ⎪


2(10)
⎪ ⎪ 1

⎪ U2(01) ⎪
⎪ ⎪





u 2 F3 ⎪
⎪ 3. Evaluation of the Viscoelastic Strains

⎪ ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪


⎪ U2(20) ⎪
⎪ ⎪
⎪ u 1 F4 ⎪

⎩ ⎭ ⎩ ⎭ 3.1. Linear Viscoelastic Constitutive Relation
U2(02) u 2 F4
In the present model, the time-dependent strain components
where B̄ = P − NΦ−1 ΘM and M, N, and P are given in the are determined incrementally using the state variable approach
appendix. [3, 25] that avoids the need to carry on the whole past history
Substituting Eq. (23) into Eq. (18), the incremental equi- of the deformation.
librium equation of the subvolume is obtained as follows: The most general representation for the constitutive rela-
tion of a linear non-aging viscoelastic material is given by:
⎧ ⎫ ⎧ ⎫
Downloaded by [200.17.114.136] at 04:49 12 February 2014

⎪ t1 F1 ⎪ ⎪ u 1 F1 ⎪ 

⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ t
∂ Di j (t − ␶ )

⎪ t2 F1 ⎪
⎪ ⎪
⎪ u 2 F1 ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪ εi (t) = Di j (0) ␴ j (t) − ␴ j (␶ ) d␶ , (27)

⎪ t ⎪
⎪ ⎪
⎪ u ⎪
⎪ ∂␶

⎪ 1 ⎪
⎪ ⎪
⎪ 1 ⎪

0

⎨ t ⎬
F2
⎪ ⎪
⎨ u ⎪
F2

2 F2 2 F2 where Di j (t) are the material creep functions. Each one of
=K − Δt 0 , (24)
⎪ ⎪
⎪ t1 F3 ⎪ ⎪ u 1 F3 ⎪
⎪ ⎪ these functions can be approximated as well as needed by a

⎪ ⎪ ⎪ ⎪
⎪ t2 ⎪
⎪ ⎪





u 2 F3 ⎪
⎪ Dirichlet-Prony series:

⎪ F3 ⎪
⎪ ⎪
⎪ ⎪


⎪ ⎪
⎪ ⎪
⎪ ⎪


⎪ t1 ⎪
F4 ⎪ ⎪
⎪ u 1 F4 ⎪
⎪  # $
⎩ ⎭ ⎩ ⎭ "
n
t−␶
t2 F4 u 2 F4 Di j (t − ␶ ) = Di0j +
p
Di j 1 − exp − p , (28)
p=1
␪i j
where K = ĀB̄ is the incremental stiffness matrix of the sub-
volume and Δt 0 is an incremental initial traction vector given p p
where Di0j , Di j , and ␪i j are material constants. For the uniaxial
by: case, Eq. (28) corresponds to the creep function of a general-
! ized Kelvin model with n + 1 springs of constants E p = 1/D p
Δt 0 = ĀNΦ−1 Δ␺ + DC Δ␧T + Δ␧V . (25) and dashpots of constants ␩p = ␪ p .D p ( p = 0, 1 . . . .n), be-
p
ing ␩0 = 0 (Figure 2). The parameters ␪i j are the retardation
In this last equation, the vectors Δ␧T and Δ␧V contain times.
the surface-averaged increments of thermal and viscoelastic Substituting Eq. (28) into Eq. (27), the following expression
strains on the four faces of the subvolume, respectively. is derived:

"
n "
m
2.2. Global Incremental Equilibrium Equation εi (t) = Di j (0) ␴ j (t) +
p
␸i j (t), (29)
The global incremental stiffness matrix KG is obtained us- p=1 j =1
ing the same procedures employed in [19]. The local stiffness
matrices are assembled into a global system of equations by with the state variables defined by:
applying continuity conditions for the increments of surface- # $
averaged interfacial tractions and displacements, followed by p 
p Di j t
t−␶
the specified boundary conditions. The system of equations re- ␸i j (t) = p exp − p ␴ j (␶ ) d␶ , (30)
␪i j 0 ␪i j
sulting from this approach relates the increments of unknown
interfacial and boundary surface-averaged displacements to
the increments of surface-averaged tractions, as follows: where no summation is implied over the repeated indices i and
j . In Eq. (29), m stands for the dimension of the stress vector.
K G ΔŪ = Δt̄ + Δt̄ o . (26)

In Eq. (26), the vector ΔŪ contains the increments


of unknown interfacial and boundary surface-averaged dis-
placements, Δt̄ is the vector of the net increments of
surface-averaged tractions along the internal interfaces and
the discretized boundary, and Δt̄ o indicates a global incre-
mental loading vector with information on body forces and Fig. 2. Generalized Kelvin model.
Model for Materials with Viscoelastic Constituents 355

Thus, the total elastic and viscoelastic strain components at 3.2. Numerical Evaluation of the State Variables
time t are given, respectively, by:
Considering a small time interval, t, the following equation
for the state variables defined by Eq. (30) can be written:
εie (t) = Di j (0) ␴ j (t) ,
"n " m
p  # $
εiV (t) =
p
␸i j (t) . (31) p Di j t+t
t + t − ␶
␸i j (t + t) = p exp − p ␴ j (␶ ) d␶ .
p=1 j =1 ␪i j 0 ␪i j
(36)
For an isotropic material and considering the volumetric and
distortional effects separately, the constitutive viscoelastic re-
Assuming ␴ j (␶ ) as constant along the interval t and equal to
lations can be written in the form:
␴ j (t), it can be easily derived from Eq. (36) the expression that
p
 t allows obtaining the variables ␸i j at time t + t as functions
∂ DK (t − ␶ )
εo (t) = DK (0) ␴m (t) − ␴m (␶ ) d␶ , of their values at time t:
∂␶
 t0
# $
∂ DG (t − ␶ )
ei (t) = DG (0) si (t) − si (␶ ) d␶ , (32) p p t
0 ∂␶ ␸i j (t + t) = ␸i j (t) exp − p
␪i j
 # $
Downloaded by [200.17.114.136] at 04:49 12 February 2014

where εo , ei , ␴m , and si stand for the volumetric strain, devi- p t


ator strain components, hydrostatic stress, and deviator stress + Di j 1 − exp − p ␴ j (t) . (37)
␪i j
components, respectively. DK and DG are the material creep
functions corresponding to the volumetric and distortional be-
haviors, respectively. DK (0) = 1/K and DG (0) = 1/2G being This simple procedure is adequate as long as the material
that K and G are the elastic values of bulk and shear moduli, relaxation times are not too short. More elaborate procedures
respectively. may be found in [26, 27].
p
The creep functions in (32) can be approximated by a In this incremental scheme, ␸i j (0) = 0 and in the beginning
Dirichlet-Prony series as: of the step t + t the values of the quantities appearing on
the right side of Eq. (37) are all already known.
"
n % &  Through Eq. (31) and Eq. (34), the viscoelastic strain vector
t−␶ p at time t + t can be written in the form:
DK (t − ␶ ) = + D0K 1 − exp − p DK ,
␪K
p=1
 % & ⎧ V ⎫ ⎛⎡ p p p ⎤⎧ ⎫
" ⎪ ε11 (t + t) ⎪

n S11 S12 S13 ⎪ 1 ⎪
p t−␶ ⎨ ⎪
⎬ " ⎨ ⎬
DG (t − ␶ ) = DG0 + DG 1 − exp − p , (33) n
⎜⎢ p p ⎥
␪G ε22
V
(t + t) = ⎜⎢ S S
p
S ⎥ 1
⎪ ⎪ ⎝⎣ 21 23 ⎦

⎩ ⎪
p=1 22

⎩ V ⎪
⎭ p=1 p p p ⎭
ε12 (t + t) S31 S32 S33 1
p p p p ⎡ p p ⎤⎧ ⎫⎞
where D0K , DK , DG0 , DG , ␪K , and ␪G are material constants. p
R11 R12 R13 ⎪ ␴11 (t) ⎪
⎢ p ⎨ ⎬⎟
Through the same reasoning used above, the elastic and vis- p ⎥
+⎢ ⎥ ␴22 (t) ⎟ , (38)
p
R
⎣ 21 R R23 ⎦
coelastic parts of the volumetric and deviator strains can be 22

⎩ ⎪⎠

p p p
given by: R31 R32 R33 ␴12 (t)

"
n
εoe (t) = DK (0) ␴m (t) , εoV (t) =
p
␸K (t), or, in compact form,
p=1
"
n "
n

eie (t) = DG (0) si (t) , eiV (t) =


p
␸Gi (t). (34) ␧V (t + t) = [S p 1 + R p ␴(t)] , (39)
p=1 p=1

p p
In Eqs. (34), ␸K and ␸Gi indicate the state variables corre- where
sponding to the volumetric and distortional behaviors, which # $  # $
are defined by: t t
p p p p
Si j = ␸i j (t) exp − p , Ri j = Di j 1 − exp − p .
p  % & ␪i j ␪i j
p DK t
t−␶
␸K (t) = p exp − p ␴m (␶ ) d␶ , (40)
␪K 0 ␪K
p  t % &
DG t−␶
p
␸Gi (t) = p exp − p si (␶ ) d␶ . (35) Hence, the vector Δ␧V of surface-averaged increments of
␪G 0 ␪G viscoelastic strain on the faces Fk of the subvolume is
356 R. S. Escarpini Filho et al.

given by:
⎧ ⎫ ⎛⎡ ⎤⎧ ⎫

⎪ Δ␧V
F1 ⎪

p
SF1 0 0 0⎪ 1⎪

⎪ ⎪
⎪ ⎪ ⎪
⎪ ⎪

⎨ Δ␧V ⎪⎬ n ⎜⎢ ⎥⎪ ⎪
SF2 0 0 ⎥ ⎨ 1 ⎬
p
" ⎜⎢ 0
F2
= ⎜⎢ ⎥
⎪ ⎪ ⎜⎢ p ⎥⎪ ⎪

⎪ Δ␧V

F3 ⎪ p=1 ⎝⎣ 0 0 SF3 0 ⎦ ⎪ ⎪ 1⎪⎪

⎪ ⎪
⎪ ⎩ ⎪
⎪ ⎭
⎩ ⎭ p
Δ␧V 0 0 0 SF4 1
⎫⎞ ⎧ V ⎫
F4
⎡ p ⎤⎧ ␧F1 ⎪
R 0 0 0 ⎪ ␴F1 ⎪ ⎪
⎪ ⎪
⎪ ⎪ ⎪ ⎪
⎢ ⎥⎪⎪ ⎪
⎪ ⎟ ⎪
⎪ ⎪
⎪ Fig. 3. Viscoelastic block in a rigid die.
⎢ 0 Rp 0 0 ⎥ ⎨ ␴F2 ⎬⎟ ⎨ ␧V ⎬
+⎢ ⎥ ⎟− F2
⎢ ⎥ ⎟ , (41)
⎣ 0 0 Rp 0 ⎦ ⎪⎪ ␴F3 ⎪
⎪ ⎪⎠ ⎪
⎪ ⎪
⎪ ␧V ⎪
F3 ⎪


⎩ ⎪
⎭ ⎪
⎪ ⎪
0 Rp ␴F4 ⎩ V ⎪ ⎭
0 0 ␧F4 friction effect between block and container is neglected. The
value of the normal stress in the y-direction is the unknown.
where ␴Fk , ␧V
p The viscoelastic behavior was modeled by a Maxwell el-
Fk , and SFk indicate the surface-averaged stress
vector, surface-averaged viscoelastic strain vector, and surface- ement with a spring of constant G = 3, 846.154 MPa and a
averaged matrix Sp , on the face Fk and at time t, respectively. dashpot with ␩G = 400 MPa.s. The adopted elastic bulk mod-
Downloaded by [200.17.114.136] at 04:49 12 February 2014

When the material creep functions corresponding to the ulus was K = 8333.333 MPa. As the formulation employs a
volumetric and distortional behaviors are considered, the hy- generalized Kelvin model, the Maxwell model was simulated
drostatic and deviator viscoelastic strains at time t + t, re- by a standard model with a very small spring constant for the
sulting from Eqs. (34) and (35), are given by Kelvin element.
The analytical solution for this example is given by [28]:
"
n
εoV (t + t) = [Sp + Rp ␴m (t)] ,   
p=1 6G 3Kt
⎧ V ⎫ ␴y (t) = −␴o 1 − exp − ,
⎪ e11 (t + t) ⎪ 3K + 4G (3K + 4G) ␪

⎨ ⎪

e22 (t + t)
V

⎪ ⎪

⎩ V ⎭ where ␪ = ␩G /G. Figure 4 presents both the numerical and an-
e12 (t + t)
⎛ ⎧ ⎫ ⎡ p ⎤⎧ ⎫⎞ alytical results for the horizontal compression stress as func-
⎪1⎪ RG 0 0 ⎪ s11 (t) ⎪
"n
⎜ p⎨ ⎬ ⎢ ⎥ ⎨ ⎬⎟ tion of time t. It is observed that the proposed formulation
= ⎜S ⎢ p
0 ⎥ ⎟
⎝ G ⎪1⎪ + ⎣ 0 ⎦ ⎪ 22 ⎪⎠ , (42)
RG s (t) provides results nearly identical with the exact analytical so-
p=1 ⎩ ⎭ p ⎩ ⎭ lution. The absolute maximum errors were about 1.75% for a
1 0 0 RG s12 (t) time increment t = 0.02 s and 0.08% for t = 0.001s. In this
example, a uniform mesh discretization of 5 × 5 subvolumes
where was adopted.
% & % &
p t p t
S =
p
(t) exp − p ,
␸K R =
p
DK1 − exp − p ,
␪ ␪
% K &  % K &
p p t p p t
SG = ␸G (t) exp − p , RG = DG 1 − exp − p .
␪G ␪G
(43)

The vectors Δ␧oV and Δe V of surface-averaged increments


of hydrostatic and deviator viscoelastic strains, respectively,
on the faces Fk of the subvolume, are determined using a
procedure similar to that employed to obtain Δ␧V in Eq. (41).

4. Numerical Examples

4.1. Viscoelastic Block in a Rigid Die


A viscoelastic block confined inside an infinitely rigid con-
tainer is subjected to an instantaneously applied and constant
vertical pressure ␴o = 10 MPa, as shown in Figure 3. The block
is made of an isotropic and homogeneous material with elastic Fig. 4. Variation of the horizontal compression stress of the block
behavior in dilatation and viscoelastic behavior in shear. The with the time.
Model for Materials with Viscoelastic Constituents 357

Fig. 5. Heterogeneous layered panel.

4.2. Creep of a Heterogeneous Layered Panel


This problem consists of a compressed panel composed by
parallel layers made of two different materials and distributed Fig. 6. Variation of the axial displacements in the panel’s right
side by side as shown in Figure 5. The panel presents a
Downloaded by [200.17.114.136] at 04:49 12 February 2014

end in function of the time.


length L = 100 cm, height B = 50 cm, and thickness e = 1 cm.
The layers in gray have a width b = 4 cm and are con-
stituted by a linear elastic material with Young’s modulus The fiber material has Young’s modulus E = 15,000 MPa and
E = 1,500 MPa and Poisson ratio ␯ = 0.0. The other lay- Poisson ratio ␯ = 0.0, whereas the viscoelastic material behav-
ers, in white, have a width a = 6 cm and are made of a lin- ior is represented by a standard linear model with constants
ear viscoelastic material exhibiting volumetric and distor- E1 = 2 × 104 MPa, E2 = 104 MPa, and ␩2 = 106 MPa.s. Per-
tional behaviors governed by standard models with the follow- fect bonding is assumed between the fiber and the matrix.
ing constants: (K1 = 666,667 MPa, K2 = 333,333 MPa, ␩K = The subvolume mesh adopted in the numerical analysis is pre-
1666,667 MPa.s), and (G 1 = 1000 MPa,G 2 = 500 MPa, ␩G = sented in Figure 7.
2500 MPa.s), respectively. The compression loading is ␴0 = The relaxation curves obtained for the surface-averaged ax-
10 MPa. ial stresses on the end faces 1 and 2 are shown in Figure 7,
Using the correspondence principle, the following expres- and corresponding to the first loading condition (i.e., the in-
sion is derived for the axial displacement in the right end of stantaneous imposed displacement) are illustrated in Figure 8.
the panel at time t: This figure also shows the variation of the mean values of the
 % % && axial normal stress ␴¯ on the right end cross section along
␴0 b 1 1 t the time. The time interval used in the numerical analysis
u(t) = Ne + Nv a + 1 − exp − ␴m
E 3K 3K2 ␪k was t = 0.1 s. The analytical expressions used to obtain the
 % 1 % &&  surface-averaged axial normal stresses on faces 1 and 2 were,
1 1 t
+ + 1 − exp − s , respectively,
2G 1 2G 2 ␪G

where Ne and Nv stand for the number of elastic and viscoelas- ␴e (t) = Eε0 and
% & % % &&
tic layers, respectively. ␴m and s indicate the hydrostatic and t E1 E2 t
normal deviator stress components in the axial direction. ␴v (t) = E1 exp − + 1 − exp − ε0 ,
␪ E1 + E2 ␪
Figure 6 shows the numerical and analytical results of the
axial displacements in the panel’s right end as function of time.
The subvolume mesh employed is shown in Figure 5. It is seen being ␪ = ␩2 /E2 and εo = u o /L. The analytical values of the
that the values found by the present formulation present an mean normal axial stress was defined by:
excellent agreement with those corresponding to the analytical
solution. ␴e (t) h e + ␴v (t) (B − h e )
␴¯ (t) = .
B
4.3. Viscoelastic Bar with an Axial Elastic Reinforcement
In this example, a linear viscoelastic bar, reinforced by a cen-
tered linear elastic fiber and fixed in its left end (Figure 7),
is subjected to two cases of loading: (a) a constant hori-
zontal displacement u 0 = 0.5 cm, which is imposed instan-
taneously on the right free end of the bar, and (b) a uni-
form and instantaneous cooling of T = −153◦ C. The bar
has a length L = 100 cm, total height B = 21.5 cm, and thick-
ness e = 1 cm. The fiber transversal dimension is h e = 1.5 cm. Fig. 7. Viscoelastic bar reinforced by an elastic fiber.
358 R. S. Escarpini Filho et al.

Fig. 8. Relaxation curves for the axial normal stress due to the
imposed displacement u 0 = 0.5cm.
Downloaded by [200.17.114.136] at 04:49 12 February 2014

As observed in Figure 8, the numerical results are in excellent Fig. 10. Axial normal stress distributions over the cross sections
agreement with the analytical solution. The maximum error defined by x = 0.5L and x = 0.7L.
was of 0.046%.
The results obtained for the second loading condition (i.e.,
composite is square with sides of 100 mm. Plane stress and
the instantaneous cooling) are illustrated in Figures 9 and 10.
plane strain conditions were considered. Due to the sym-
For generation of these results, the following thermal expan-
metry of the problem, only one-quarter of the domain was
sion coefficients were adopted for the elastic and viscoelastic
modeled.
material: ␣e = 2 × 10−5 /◦ C and ␣v = 16 × 10−5 /◦ C, respec-
tively. Figure 9 shows the variation of the axial displacement
in the center of the right cross section of the reinforced bar (a)
and non-reinforced bar. Figure 10 illustrates the axial nor-
mal stress distributions over the cross sections defined by the
coordinates x = 0.5L and x = 0.7L.

4.4. Elastic Inclusion Embedded in a Viscoelastic Matrix


Figure 11a shows a problem that consists of an elastic cir-
cular inclusion embedded in a viscoelastic matrix subjected
to a uniform tensile loading given by ␴o = 100 MPa. The
(b)

Fig. 9. Variation of the axial displacements in the panel’s right Fig. 11. Composite with an elastic circular inclusion and mesh
end in function of the time for the instantaneous cooling. discretizations.
Model for Materials with Viscoelastic Constituents 359

dance with the results generated by the finite-element code for


all inclusion volume fractions.

5. Conclusions

A new numerical incremental model for the analysis of com-


posite structures made of linear viscoelastic constituents has
been developed, using the framework of the recent parametric
formulation of the finite-volume theory. Utilizing a state vari-
able approach for evaluation of the time-dependent strains,
the model demonstrated computational efficiency and a very
good potential to describe the behavior of different cases of
linear viscoelastic structures. Several examples were analyzed
Fig. 12. Comparison of the displacement curves for inclusion and comparisons of results with analytical solutions confirm
volume fraction of 30% and different meshes. the excellent performance of the proposed model. On the other
hand, considering the recognized capability of the parametric
Downloaded by [200.17.114.136] at 04:49 12 February 2014

formulation of the finite-volume theory in modeling of hetero-


The matrix was modeled using a standard model geneous solids, it is believed that the numerical tool developed
with the following mechanical properties: E1 = 9,500 MPa, can be employed to predict the viscoelastic behavior of com-
␯ = 0.3, G 2 = 2,500 MPa, ␩G = 6.81 × 106 MPa.s, K2 = plex multiphase composite structures.
1666.667 MPa, and ␩K = 4.54 × 106 MPa.s. The elastic prop-
erties of the inclusion were E = 30,250 MPa and ␯ = 0.443.
Volume fractions v f of 5, 30, and 60% were adopted for the
inclusion. The mesh discretizations used in the investigation References
are illustrated in Figure 11b. To check the performance with [1] D.D.L. Chung, Composite materials: functional materials for mod-
respect to the mesh sizes, the case corresponding to the vol- ern technologies, Springer-Verlag London Limited, London, 2004.
ume fractions of 30% was analyzed using three meshes pre- [2] D.L. Flaggs and F.W. Crossman, Analysis of the viscoelastic re-
senting distinct refinement levels. For this last case, the hori- sponse of composite laminates during hygrothermal exposure, J.
zontal displacement at the midpoint of the right end (loaded Compos. Mater., vol. 15, pp. 21–40, 1981.
side), obtained using the three different meshes, are shown in [3] S.P.C. Marques and G.J. Creus, Geometrically nonlinear finite ele-
ments analysis of viscoelastic composite materials under mechan-
Figure 12 as function of time. As observed, these results are
ical and hygrothermal loads, Comput. Struct., vol. 53, no. 2, pp.
virtually identical for the three levels of discretizations. 449–456, 1994.
Figure 13 shows the variation of that horizontal displace- [4] B.F. Oliveira and G.J. Creus, An analytical–numerical framework
ment in function of time for both plane stress and plane strain for the study of ageing in fibre reinforced polymer composites, Com-
analyses and different volume fractions. As expected, the dis- pos. Struct., vol. 65, pp. 443–457, 2004.
placement curve for the plane strain condition is located below [5] J.D. Eshelby, The determination of the elastic field of an ellipsoidal
the curve of the plane stress state. Results obtained using the inclusion and related problems, Proc. Roy. Soc. Lond., vol. A241,
pp. 376–396, 1957.
finite-element software ANSYS [29] are also shown in Fig-
[6] S. Nemat-Nasser and M. Hori, Micromechanics: Overall Properties
ure 13. As it can be observed, using the same discretization, of Heterogeneous Materials, North-Holland, Amsterdam, 1999.
the finite-volume theory solutions are in very good concor- [7] R.M. Christensen, Viscoelastic properties of heterogeneous media,
J. Mech. Phys. Solids, vol. 17, pp. 23–41, 1969.
[8] N. Laws and R. McLaughlin, Self consistent estimates for the vis-
coelastic creep compliances of composite materials, Proc. Roy. Soc.
Lond., vol. 359, pp. 251–273, 1978.
[9] L.C. Brinson and W.S. Lin, Comparison of micromechanics meth-
ods for effective properties of multiphase viscoelastic composites,
Compos. Struct., vol. 41, pp. 353–367, 1998.
[10] G. DeBotton and L. Tevet-Deree, The response of a fiber-reinforced
composite with a viscoelastic matrix phase, J. Compos. Mater., vol.
38, pp. 1255–1277, 2004.
[11] L. Gibiansky, G. Milton, and J. Berryman, On the effective vis-
coelastic moduli of two-phase media: III. Rigorous bounds on the
complex shear modulus in two dimensions, Proc. Roy. Soc. Lond.,
vol. A455, pp. 2117–2149, 1999.
[12] N. Lahellec and P. Suquet, Effective behavior of linear viscoelastic
composites: A time integration approach, Int. J. Solids Struct., vol.
44, pp. 507–529, 2007.
[13] Y. Yeong-Moo, P. Sang-Hoon, and Y. Sung-Kie, Asymptotic ho-
Fig. 13. Displacement curves for plane stress and plane strain mogenization of viscoelastic composites with periodic microstruc-
analyses. tures, Int. J. Solids Struct., vol. 35, pp. 2039–2055, 1998.
360 R. S. Escarpini Filho et al.
⎡ ⎤
[14] L.C. Brinson and W.G. Knauss, Finite element analysis of multi- Jˆ 0 0 0 0 0 0 0
phase viscoelastic solids, J. Appl. Mech., vol. 59, no. 4, pp. 730–737, ⎢ ⎥
1992. ⎢0 Jˆ 0 0 0 0 0 0⎥
⎢ ⎥
[15] E.J. Barbero, Finite Element Analysis of Composite Materials, CRC ⎢0 0 Jˆ 0 0 0 0 0⎥
⎢ ⎥
Press, Boca Raton, FL, 2007. ⎢ ⎥
[16] M. Payley and J. Aboudi, Micromechanical analysis of composites ⎢0 0 0 Jˆ 0 0 0 0⎥
B=⎢
⎢0
⎥, (A.3)
by the generalized methods of cells, Mech. Mater., vol. 14, no. 1,
⎢ 0 0 0 Jˆ 0 0 0⎥ ⎥
pp. 127–139, 1992. ⎢ ⎥
[17] J. Aboudi, M.-J. Pindera, and S.M. Arnold, Higher-order theory ⎢0 0 0 0 0 Jˆ 0 0⎥
⎢ ⎥
for functionally graded materials, Composites: Part B, vol. 30, no. ⎢ ⎥
8, pp. 772–832, 1999. ⎣0 0 0 0 0 0 Jˆ 0⎦
[18] Y. Bansal and M.-J. Pindera, Efficient reformulation of the ther- 0 0 0 0 0 0 0 Jˆ
moelastic higher-order theory for FGMs, J. Therm. Stresses, vol.
26, no. 11/12, pp. 1055–1092, 2003.
[19] M.A.A. Cavalcante, S.P.C. Marques, and M.-J. Pindera, Paramet- where
ric formulation of the finite-volume theory for functionally graded
materials–Part I: Analysis, J. Appl. Mech., vol. 74, no. 5, pp. ⎡ ⎤
935–945, 2007. 1 0 0 0 0 0 0 0
⎢0 0 ∓3 0 0 0 0 ⎥
[20] M.A.A. Cavalcante, S.P.C. Marques, and M.-J. Pindera, Paramet- ⎢ 1 ⎥
ric formulation of the finite-volume theory for functionally graded A(1,3) =⎢ ⎥,
⎣0 0 0 0 1 0 0 0 ⎦
Downloaded by [200.17.114.136] at 04:49 12 February 2014

materials–Part II: Numerical Results, J. Appl. Mech., vol. 74, no. 5,


pp. 946–957, 2007. 0 0 0 0 0 1 0 ∓3
[21] M.A.A. Cavalcante, S.P.C. Marques, and M.-J. Pindera, Computa- ⎡ ⎤
1 0 ±3 0 0 0 0 0
tional aspects of the parametric finite-volume theory for function-
⎢0 0 0⎥
ally graded materials, Comput. Mater. Sci., vol. 44, pp. 422–438, ⎢ 1 0 0 0 0 ⎥
2008. A(2,4) =⎢ ⎥, (A.4)
⎣0 0 0 0 1 0 ±3 0 ⎦
[22] M.A.A. Cavalcante and S.P.C. Marques, A parametric elastoplastic
formulation of the method of cells for composite materials. Proceed- 0 0 0 0 0 1 0 0
ings of the XXVII Iberian Latin American Congress on Computa-
tional Methods for Engineering, September 3–6, Belem-PA, Brazil,
CD, 2006 (in Portuguese). and n(k) are matrices defined in terms of the components of
[23] H. Khatam, and M.-J. Pindera, Parametric finite-volume microme- the unit normal vector to each face Fk :
chanics of periodic materials with elastoplastic phases, Int. J. Plast.,
vol. 25, no. 7, pp. 1386–1411, 2009.  
[24] M.A.A. Cavalcante, H. Khatam, and M.-J. Pindera, Homog- n1 0 n2
enization of elastic-plastic periodic materials by FVDAM and
n(k) = . (A.5)
0 n2 n1 k
FEM approaches—An assessment, Composite: Part B, DOI:
10.1016/j.compositesb.2011.03.006.
[25] G.J. Creus, Lecture Notes in Engineering: Viscoelasticity—Basic C̄, Ē, and Jˆ are given in Subsection 2.1.
Theory and Applications to Concrete Structures, Springer-Verlag,
The matrices P, M, and N appearing in the definition of the
Berlin, 1986.
[26] J.R. Masuero and G. J. Creus, Finite element analysis of viscoelastic matrix B̄ (Eq. 23) are given by:
fracture, Int. J. Fract., vol. 60, pp. 267–282, 1993.
[27] J. Sorvari and J. Hämäläinen, Time integration in linear ⎡ ⎤
0 0 1/2 0 0 0 −1/2 0
viscoelasticity—A comparative study, Mech. Time-Depend. Mater.,
⎢ −1/2 0 0 0 1/2 0 0 0 ⎥
vol. 14, pp. 307–328, 2010. ⎢ ⎥
[28] W. Flugge, Viscoelasticity, Springer-Verlag, New York, 1975. ⎢ 0 0 ⎥
⎢ 0 1/2 0 0 0 1/2 ⎥
[29] ANSYS, Inc., Southpointe, 275 Technology Drive, Canonsburg, PA ⎢ ⎥
⎢ 1/2 0 0 0 1/2 0 0 0 ⎥
P=⎢ ⎥,
15317, USA.
⎢ 0 0 0 1/2 0 0 0 −1/2 ⎥
⎢ ⎥
⎢ ⎥
Appendix ⎢ 0 −1/2 0 0 0 1/2 0 0 ⎥
⎢ ⎥
⎣ 0 0 0 1/2 0 0 0 1/2 ⎦
The matrices used to define the matrix Ā, appearing in Eq. 0 1/2 0 0 0 1/2 0 0
(18), are given by: (A.6)
⎡ ⎤
⎡ ⎤ ⎡ ⎤ 0 0
A(1) n(1) 0 0 0 ⎢0 0⎥
⎢ A(2) ⎥ ⎢ 0 ⎡ ⎤ ⎢ ⎥
⎢ ⎥ ⎢ n(2) 0 0 ⎥
⎥ 0 0 1 0 0 0 1 0 ⎢1 0⎥
⎢ ⎥
A = ⎢ (3) ⎥ , D=⎢ ⎥, (A.1) ⎢ ⎥
⎣A ⎦ ⎣ 0 0 n(3) 0 ⎦ ⎢1 0 0 0 1 0 0 0⎥ ⎢ ⎥
⎢ ⎥ 1 0
M=⎢ ⎥, N=⎢ ⎢
⎥.

A(4) 0 0 0 n(4) ⎣0 0 0 1 0 0 0 1⎦ ⎢0 0⎥
⎡ ⎤ ⎡ ⎤ ⎢ ⎥
C̄ 0 0 0 Ē 0 0 0 0 1 0 0 0 1 0 0 ⎢0 0⎥
⎢0 ⎢ ⎥
⎢ C̄ 0 0⎥

⎢0
⎢ Ē 0 0⎥
⎥ ⎣0 1⎦
C=⎢ ⎥, E=⎢ ⎥ , (A.2)
⎣0 0 C̄ 0⎦ ⎣0 0 Ē 0⎦ 0 1
0 0 0 C̄ 0 0 0 Ē (A.7)
Model for Materials with Viscoelastic Constituents 361

The matrices Φ and Θ appearing in Eq. (21) are given by: where
 
W11 + W12 W13 + W14 W11 = C̄11 (Ĵ11 )2 + C̄33 (Ĵ21 )2 , W12 = C̄11 (Ĵ12 )2 + C̄33 (Ĵ22 )2 ,
Φ= ,
W21 + W22 W23 + W24 W13 = (C̄12 + C̄33 )Ĵ11 Ĵ21 , W14 = (C̄12 + C̄33 )Ĵ12 Ĵ22 ,
 
1 W11 W12 W13 W14 W21 = (C̄33 + C̄21 )Ĵ11 Ĵ21 , W22 = (C̄33 + C̄21 )Ĵ12 Ĵ22 ,
Θ= , (A.8)
2 W21 W22 W23 W24 W23 = C̄33 (Ĵ11 )2 + C̄22 (Ĵ21 )2 , W24 = C̄33 (Ĵ12 )2 + C̄22 (Ĵ22 )2 .
(A.9)
Downloaded by [200.17.114.136] at 04:49 12 February 2014

Das könnte Ihnen auch gefallen