Anda di halaman 1dari 12

COMPOSITES

SCIENCE AND
TECHNOLOGY
Composites Science and Technology 67 (2007) 2795–2806
www.elsevier.com/locate/compscitech

Mechanical behavior of unidirectional fiber-reinforced polymers


under transverse compression: Microscopic
mechanisms and modeling
Carlos González, Javier LLorca *

Departamento de Ciencia de Materiales, Universidad Politécnica de Madrid & Instituto Madrileño de Estudios Avanzados en Materiales
(IMDEA-Materiales) E.T.S. de Ingenieros de Caminos, 28040 Madrid, Spain

Received 1 December 2006; received in revised form 2 February 2007; accepted 2 February 2007
Available online 15 February 2007

Abstract

The mechanical behavior of polymer–matrix composites unidirectionally reinforced with carbon or glass fibers subjected to compres-
sion perpendicular to the fibers was studied using computational micromechanics. The stress–strain curve was determined by the finite
element analysis of a representative volume element of the microstructure idealized as a random dispersion of parallel fibers embedded in
the polymeric matrix. The dominant damage mechanisms experimentally observed – interface decohesion and matrix plastic deformation
– were included in the simulations, and a parametrical study was carried out to assess the influence of matrix and interface properties on
the stress–strain curve, compressive strength, ductility and the corresponding failure modes. It was found that the composite properties
under transverse compression were mainly controlled by interface strength and the matrix yield strength in uniaxial compression. Two
different fracture modes were identified, depending on whether failure was controlled by the nucleation of interface cracks or by the for-
mation of matrix shear bands. Other parameters, such as matrix friction angle, interface fracture energy or thermo-elastic residual stres-
ses, played a secondary role in the composite mechanical behavior.
 2007 Elsevier Ltd. All rights reserved.

Keywords: Computational micromechanics; Interface fracture; Transverse compression; Fiber-reinforced polymers

1. Introduction fraction, shape and spatial distribution are not available.


This is an important limitation because the experimental
Unidirectional fiber-reinforced polymers show outstand- characterization of the lamina properties in the transverse
ing specific stiffness and strength along the fiber direction direction is subjected to more uncertainties than in the lon-
and this has led to a wide range of applications as struc- gitudinal one and, in fact, experimental data are more
tural materials. Moreover, the fiber and matrix behavior scarce. In addition, the longitudinal compressive strength
follows very closely the isostrain approximation until the is severely affected by the transverse behavior [5,6], and
onset of failure and it was possible to develop analytical the development of robust failure criteria for laminates
models to accurately predict the tensile [1–3] and compres- which include the interaction between longitudinal and
sive [4,5] strength in the fiber direction. Conversely, the transverse stresses have to rely on a precise knowledge of
mechanical behavior under transverse loading cannot be the lamina behavior under transverse loading until failure.
represented by simplified isostrain or isostress approaches, Computational micromechanics is emerging as an accu-
and micromechanical models capable of predicting failure rate tool to study the mechanical behavior of composites
strength as a function of the constituent properties, volume due to the sophistication of the modeling tools and to the
ever-increasing power of digital computers. Within this
*
Corresponding author. Tel.: +34 91 336 5375; fax: +34 91 543 7845. framework, the macroscopic properties of a composite
E-mail address: jllorca@mater.upm.es (J. LLorca). lamina can be obtained by means of the numerical simula-

0266-3538/$ - see front matter  2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compscitech.2007.02.001
2796 C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806

tion of the deformation and failure of a representative vol- elastic fibers embedded in the continuous polymeric matrix.
ume element of the microstructure [7–10]. As compared The main deformation and failure mechanisms reported in
with the classic homogenization techniques, computational the literature (namely matrix nonlinear behavior and inter-
micromechanics presents two important advantages. face failure) as well as the effect of thermal residual stresses
Firstly, the influence of the geometry and spatial distribu- were taken into account in the simulations and a paramet-
tion of the phases (i.e. size, shape, clustering, connectivity, rical study was carried out to assess the influence of these
etc.) can be accurately taken into account. Secondly, the parameters on the stress–strain curve, failure strength, duc-
details of the stress and strain microfields throughout the tility and the corresponding failure modes.
microstructure are resolved, which leads to precise estima-
tions of the onset and propagation of damage, and to accu- 2. Experimental background and simulation strategy
rate predictions of the failure strength. Recent advances in
this area include the analysis of the effect of particle shape The experimental evidence shows that lamina of poly-
[11], particle clustering [12,13] and the influence of damage mer–matrix composites unidirectionally reinforced with
[14,15] on the mechanical behavior of particle-reinforced carbon or glass fibers fail under transverse compression
composites, the prediction of the mechanical behavior of along planes parallel to the fibers [20–22]. The angle a
foams and composites whose microstructure was obtained formed between the failure plane and the through-thick-
by means of X-ray computer-assisted tomography [16,17], ness (or perpendicular to the in-plane loading) direction
or the computer simulation of ‘‘virtual fracture tests’’ in is slightly above 45 and typical values reported are in
fiber-reinforced composites [18,19]. the range 50–56 [23,21,24]. Significant non-linear defor-
This strategy is applied in this investigation to analyze mation was often observed before the maximum load
the mechanical behavior of a unidirectional fiber-rein- [25,26], and this behavior was associated to the plastic
forced polymer composite subjected to transverse compres- deformation of the polymeric matrix. This is supported
sion. The composite microstructure was idealized by a by our observations on the lateral surfaces of a Hexcel
random and homogeneous dispersion of parallel, circular 8552 epoxy matrix uniaxially reinforced with 57 vol.%

Fig. 1. Scanning electron micrograph of the lateral surface of an AS4/epoxy specimen loaded under transverse compression showing bands of intense
plastic deformation in the matrix before the maximum load [24].
C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806 2797

AS4 carbon fibers loaded under transverse compression. ing the Mohr–Coulomb yield criterion, which assumes that
Bands of intense plastic deformation in the matrix, inclined yielding is induced by the shear stresses and that yield stress
at an angle of 56 with respect to the plane perpendicular to depends on the normal stress. This model has often been
the loading axis, appeared before the maximum load was used to describe plastic deformation and failure of poly-
attained (Fig. 1). Damage by interface decohesion devel- mers [27] and of polymeric matrices in composites
oped afterwards around these bands (Fig. 2a) and final [28,23,21] as it explains the asymmetry between tensile
fracture occurred by the failure of the matrix in shear, as and compressive yielding and failure in compression along
evidenced by the numerous hackles in the matrix fracture planes forming an angle of 50–56 with respect to the
surfaces (Fig. 2b). plane perpendicular to the loading axis. Fiber/matrix dec-
These results show that the strength of fiber-reinforced ohesion was introduced by means of interface elements
polymers under transverse compression is controlled by whose behavior is controlled by a cohesive crack model,
two dominant mechanisms, namely the localization of the a standard technique in the computational micromechanics
matrix plastic strain along shear bands and the develop- of composites [29–32].
ment of damage by interface decohesion. Both processes
(and their interaction) can be taken into account within 3. Computational model
the framework of computational micromechanics in which
composite behavior is analyzed by means of the finite ele- 3.1. RVE generation and discretization
ment simulation of a two-dimensional representative vol-
ume element (RVE) of the microstructure. The matrix A square RVE, which contains a random and homoge-
was represented by an isotropic, elasto-plastic solid follow- neous dispersion of circular fibers embedded in the poly-
meric matrix, was selected to determine the behavior of
the composite under transverse loading, following to Broc-
kenbrough et al. [33]. An important issue in the simulations
is the minimum size of the RVE, which should contain all
the necessary information about the statistical description
of the microstructure and its size should be large enough
so that the average properties of this volume element are
independent of its size and position within the material.
Of course, the critical RVE size depends on the phase
and interface properties and spatial distribution, and no
estimates were available for our particular problem. It is
also known that the accuracy provided by RVEs of a given
size can be improved if the results of various realizations
are averaged [34]. Thus, the compressive strength and duc-
tility for each set of matrix and interface properties was
given by the average value of the results obtained from
six different fiber distributions in a RVE which included
30 fibers. They were compared in selected cases with those
obtained with RVEs containing over 70 fibers to ensure
that the size of the RVE did not influence significantly
the model predictions.
Random and homogeneous dispersions of monosized
fibers of radius R = 5 lm were generated in square RVEs
of dimensions L0 · L0 using the modified random sequen-
tial adsortion algorithm of Segurado and LLorca [9]. It
was assumed that the microstructure of the composite
was given by a indefinite translation of the RVE along
the two coordinate axes and thus the fiber positions within
the RVE should keep this periodicity condition. Fiber cen-
ters were generated randomly and sequentially, and each
new fiber was accepted if the distance between neighboring
fiber surfaces was >0.07R to ensure an adequate discretiza-
tion of this region. In addition, the distance between the
Fig. 2. Scanning electron micrographs of an AS4/epoxy composite loaded
fiber surface and the RVE edges should be >0.1R to avoid
under transverse compression [24]. (a) Damage by interface decohesion
around the matrix shear bands. The loading axis is horizontal. (b) distorted finite elements during meshing. Fibers intersect-
Fracture surface. The presence of numerous hackles in the matrix is ing the RVE edges were split into an appropriate number
indicative of failure by shear. of parts and copied to the opposite sides of the square
2798 C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806

Table 1
Thermo-elastic constants of the fibers and the matrix [38]
Ef (GPa) mf af (106 K1) Em (GPa) mm am (106 K1)
40 0.25 10 4 0.35 50

where the integral stands for the resultant forces acting on


the edge X1 = 0 due to the traction vector ~ t. The logarith-
mic strain along the loading axis was given as
 = ln(1 + d/L0) and the corresponding true stress on the
edge was computed as the resultant force divided by the ac-
tual cross-section.
Simulations were carried out with Abaqus/Standard [35]
under plane strain conditions and within the framework of
the finite deformations theory with the initial unstressed
state as reference. Fibers were modeled as linear, thermo-
elastic and isotropic solids. The thermo-elastic constants
given in Table 1 are intermediate between those of glass
Fig. 3. Fiber distribution and finite element discretization of a represen- and C fibers in the plane perpendicular to the fiber axis.
tative volume element of the composite with 30 fibers.
The polymeric matrix was assumed to behave as an isotro-
RVE to create a periodic microstructure. New fibers were pic, thermo-elasto-plastic solid, and the thermo-elastic con-
added until the desired volume fraction of 50% was stants (typical of an epoxy matrix) are also given in Table 1.
reached. An example of the fiber distribution an RVE with Plastic deformation was governed by the Mohr–Coulumb
30 fibers is shown in Fig. 3. The RVE was automatically criterion and the total matrix strain was given by the addi-
meshed using 6-node isoparametric modified triangles tion of the thermo-elastic and plastic strain components.
(CPE6M in Abaqus Standard [35]) with integration at The Mohr–Coulomb criterion assumes that yielding takes
three Gauss points and hourglass control. Special care place when the shear stress acting on a specific plane, s,
was taken to obtain a very fine and homogeneous discreti- reaches a critical value, which depends on the normal stress
zation throughout the RVE to resolve the plastic shear r acting on that plane. This can be expressed as
bands in the matrix during deformation (Fig. 3). s ¼ c  r tan /; ð4Þ
where c and / stand, respectively, for the cohesion and the
3.2. Finite element and material models
friction angle, two materials parameters which control the
plastic behavior of the material. Physically, the cohesion c
Periodic boundary conditions were applied to the edges
represents the yield stress under pure shear while the fric-
of the RVE because the continuity between neighboring
tion angle takes into account the effect of the hydrostatic
RVEs (which deform like jigsaw puzzles) is maintained
stresses. / = 0 reduces the Mohr–Coulomb model to the
and, in addition, because the effective behavior derived
pressure-independent Tresca model while / = 90 leads to
under these conditions is always bounded by those obtained
‘‘tension cut-off’’ Rankine model. The value of both
under imposed forces or displacements [36,37]. Let X1 and
parameters for an epoxy can be assessed from its tensile
X2 stand as the Cartesian coordinate axes parallel to the
and compressive strengths, rmt and rmc, according to
RVE edges and with origin at one corner of the RVE.
The periodic boundary conditions can be expressed in terms cos / cos /
~ 1 and U
~ 2 which relate the dis- rmt ¼ 2c and rmc ¼ 2c : ð5Þ
of the displacement vectors U 1 þ sin / 1  sin /
placements between opposite edges according to
The fracture surface of a solid which follows the Mohr–
uð0; X 2 Þ  ~
~ ~1;
uðL0 ; X 2 Þ ¼ U ð1Þ Coulomb criterion and it is subjected to uniaxial compres-
~2: sion forms an angle a with plane perpendicular to the load-
uðX 1 ; 0Þ  ~
~ uðX 1 ; L0 Þ ¼ U ð2Þ
ing axis, which is related to the friction angle / by
Uniaxial compression along the X2 axis is imposed with a ¼ 45 þ /=2: ð6Þ
~ 2 ¼ ð0; dÞ and U
U ~ 1 ¼ ðu1 ; 0Þ. d stands for the imposed
displacement in the loading direction and u1 is computed Typically 50 < a < 60 in epoxy matrices [23,21,24], and
from the condition that the average stresses on the edges thus / is in the range 10–30. Once / was fixed for a given
perpendicular to the loading axis should be 0. simulation, the corresponding cohesion c was computed
Mathematically, from Eq. (5) assuming that the matrix tensile strength was
Z L 60 MPa [38]. If not indicated otherwise, the simulations pre-
~
t dX 2 ¼ 0 on X 1 ¼ 0; ð3Þ sented in this paper used / = 15 to represent the matrix
0 behavior, which corresponds to a cohesion c of 39.1 MPa.
C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806 2799

The corresponding values on the matrix tensile and com-


1
pressive strength are, respectively, 60 MPa and 101.9 MPa.
The yield surface of the Mohr–Coulomb model, written
in terms of the maximum and minimum principal stresses
(rI and rIII), is given by

F ðrI ; rIII Þ ¼ ðrI  rIII Þ þ ðrI þ rIII Þ sin /  2c cos / ¼ 0 tn , ts


N S
ð7Þ K

and it was assumed that c and / were constant and inde- 1 (1-d)K
pendent of the accumulated plastic strain. A non-associa- 1
tive flow rule was used to compute the directions of
δ
plastic flow in the stress space and the corresponding po- δ0 δ max δf
tential flow G was expressed as
2
Fig. 4. Schematic of the traction-separation law which governs the
4ð1  e2 Þ cos2 H þ ð2e  1Þ behavior of the interface elements.
G¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2ð1  e2 Þ cos H þ ð2e  1Þ 4ð1  e2 Þ cos2 H þ 5e2  4e
3  sin /
 simplicity (N = S). Once the damage begins, the stress
6 cos /
transferred through the crack is reduced depending on
ð8Þ the interface damage parameter d, which evolves from 0
in which e = (3  sin/)/(3 + sin/) and H is obtained from (in the absence of damage) to 1 (no stresses transmitted
 3 across the interface), as shown in Fig. 4. The corresponding
1 J3 traction-separation law is expressed by
H ¼ arccos ; ð9Þ
3 J2
tn ¼ ð1  dÞKdn if dn > 0;
where J2 and J3 are, respectively, the second and the third tn ¼ Kdn if dn 6 0; ð12Þ
invariants of the deviatoric stress tensor. More details
about the numerical implementation of the Mohr–Cou- ts ¼ ð1  dÞKds
lomb model can be found in [39,40]. The evolution of the damage parameter is controlled by an
The progressive interface decohesion upon loading was effective displacement, d, defined as the norm of the dis-
simulated by 4-node isoparametric linear interface elements placement jump vector across the interface as
(COH2D4 in [35]) inserted at the fiber/matrix interface. qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
The mechanical behavior of these elements was expressed d ¼ hdn i2 þ d2 ; ð13Þ
s
in terms of a traction-separation law which relates the dis-
placement jump across the interface with the traction vec- and d depends on the maximum effective displacement at
tor acting upon it. The initial response was linear in the interface attained during the loading history at each
absence of damage and, therefore, the traction-separation material integration point dmax according to
law can be written as df ðdmax  d0 Þ
d ¼ max f 0 ; ð14Þ
tn ¼ Kdn and ts ¼ Kds ð10Þ d ðd  d Þ
where tn, ts, dn and ds stand for the normal and tangential where d0 and df stand for the effective displacement at the
tractions and displacement jumps across the interface onset of damage (d = 0) and when the interface has failed
respectively. An elastic stiffness of K = 108 GPa/m was se- completely (d = 1), respectively. In this cohesive model,
lected for the interface, which was large enough to ensure the energy necessary to completely break the interface is al-
the displacement continuity at the interface and to avoid ways equal to C, the interface fracture energy, regardless of
any modification of the stress fields around the fibers in the loading path. If not indicated otherwise, the interface
the absence of damage. The linear behavior ends at the on- fracture energy in the simulations presented below was
set of damage, which is dictated by a maximum stress cri- 100 J/m2, a reasonable value for C and glass fibers embed-
terion expressed mathematically as ded in a polymeric matrix [41].
 
htn i ts
max ; ¼1 ð11Þ 4. Results
N S
in which h i stand for the Macaulay brackets, which return 4.1. Validation of the RVE size
the argument if positive and zero otherwise, to impede the
development of damage when the interface is under com- Most of the results presented in this paper were
pression, and N and S are the normal and tangential inter- obtained by the numerical simulation of RVEs containing
facial strengths which were assumed to be the equal for 30 fibers. The influence of the actual position of the fibers
2800 C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806

140 140

120 120

∞ 2c

Compressive stress (MPa)


100
Compressive stress (MPa)

100 c mc

0.5c
80 80

60 60

0.1c
40 40
0.1c

20 30 fibers 20
70 fibers
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05
Compressive strain Compressive strain
Fig. 5. Compressive stress–strain curves for six different fiber realizations Fig. 6. Influence of the interface strength N on the mechanical response
in RVEs with 30 fibers (broken line) and one fiber realization in an RVE under transverse compression. The error bars stands for the standard
with 70 fibers (solid line). The two sets of curves are representative of deviation of six simulations. The figure next to each curve stands for the
materials with very strong (=1) or very weak (=0.1c) fiber/matrix interface strength. The broken horizontal line represents the compressive
interfaces. strength of the epoxy matrix.

within the RVE on the mechanical response was analyzed mum load and remained small afterwards. The initial com-
by comparing the results obtained with six different fiber posite stiffness was not affected by the interface strength
realizations for the typical values of the matrix and fiber but the composites with low interfacial strength (N < c)
properties given previously and two sets of interface prop- departed early from the linear behavior due to the nucle-
erties corresponding to very weak (N = 0.1c) and perfect ation of interface cracks. In isolated fibers, the cracks
interfaces, respectively. The corresponding (compression) nucleated at the points equidistant from the poles and the
stress–strain curves are plotted in Fig. 5, together with equator (latitudes 45 N and 45 S), where the interfacial
those computed with an RVE which included 70 fibers. shear stress was maximum. They propagated towards the
All the simulations were practically superposed in the elas- equator and merged. The stress concentrations at the tip
tic regime; divergences arose at the onset of matrix plastic of the interface cracks induced the formation of very short
deformation and increased in the composite with weak shear bands in the matrix linking up interface cracks in
interfaces beyond the maximum load. These results are neighboring fibers, and the maximum strength was attained
in agreement with previous numerical studies, which at this point (Fig. 7a). Further deformation led to forma-
showed that the minimum size of the RVE increases with tion of interfacial voids and to the localization of the strain
the mismatch between the phase properties (e.g. at the in the matrix in shear bands whose path was dictated by the
elasto-plastic transition) and especially with the localiza- position of the voids which grew from the interface cracks
tion of the deformation due to plastic flow and/or damage (Fig. 7b).
[42,10]. Nevertheless, the dispersion among the stress– On the contrary, composites without interface decohe-
strain curves was limited and the curve obtained by aver- sion presented a linear behavior up to compressive stresses
aging the six simulations was very close to that computed very close to the strength of the epoxy matrix in uniaxial
with an RVE with 70 fibers for both sets of material compression, rmc. This linear regime was followed by a
properties. plastic response with very little hardening as the localiza-
tion of the plastic strain in the matrix led to the formation
4.2. Influence of the interface strength of shear bands which percolated the entire RVE (Fig. 8). It
is worth noting that the angle between the shear bands and
The stress–strain curve under transverse compression is the plane perpendicular to the loading axis was very close
plotted in Fig. 6 for composite materials whose interface to 45 + //2 = 52.5, the theoretical one for the matrix
strength varied from N = 0.1c to infinity. The matrix fric- alone, regardless of the actual fiber distribution, and this
tion angle was 15 (c = 39.1 MPa) and the interfacial frac- indicates that the composite strength was determined by
ture energy was 100 J/m2 in all cases. Each curve is the the propensity of the matrix to form shear bands.
average of six different realizations with an RVE with 30 The behavior of the composite with an intermediate
fibers and the error bars stand for the standard deviation interfacial strength (N = c) was initially similar to that of
of the simulations, which was negligible up to the maxi- the materials with high interfacial strength, and the pattern
C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806 2801

Fig. 8. Contour plot of the accumulated plastic strain in the matrix in the
composite without interface decohesion (N = 1) at  = 7%. The loading
axis is horizontal.

interface strength in the transverse compressive strength,


YC, is plotted in Fig. 10. Failure is controlled by the matrix
plastic deformation if N/c P 2 and the reinforcing effect of
the stiff fibers increased the composite strength approxi-
mately 10% over the matrix flow stress in compression.
The composite strength decreases rapidly with the interfa-
cial strength as the stress concentrations associated with
interface cracks favor the onset of plastic deformation in
the matrix and the nucleation of shear bands at lower stres-
ses. It is interesting to note that predictions of the microm-
echanics simulations are in good agreement with
experimental results for epoxy–matrix composites rein-
forced with either glass or carbon fibers (Fig. 10). Experi-
mental values of the matrix and composite properties
under transverse compression were obtained from [38];
Fig. 7. Contour plot of the accumulated plastic strain in the matrix in the information of the interface strength for both composite
composite with low interfacial strength (N = 0.1c). (a)  = 1.7% corre- systems was not available in this reference and the experi-
sponding to the maximum strength. (b)  = 4%. The loading axis is mental data in [43] for C/epoxy and in [44] for glass/epoxy
horizontal. Notice that the strain values in legend (b) are 10 times higher
than in (a).
were used. Thus, although the actual interface strength is
not known, it is evident that the model predictions for
of plastic deformation in the matrix at the point of maxi- the transverse compressive strength and the failure micro-
mum stress showed the incipient development of shear mechanisms (Figs. 1 and 2) support the validity of the cur-
bands oriented at 52.5 (Fig. 9a). However, final fracture rent approach to simulate the mechanical behavior of
occurred by the development of a single shear band, unidirectional PMC.
slightly misoriented with respect to the theoretical angle, The data in Fig. 10 also include the influence of the
whose orientation was dictated by the linking up of inter- interface properties in the strain at YC, which stands for
face cracks in adjacent fibers (Fig. 9b). This fracture pat- a rough approximation of the composite ductility under
tern is very similar to that observed in Fig. 2a, in which transverse compression. The ductility values presented
the matrix shear band is surrounded by interface cracks more scatter (particularly for large interface strengths, in
and points to a failure process in three steps: incipient which the stress–strain curve is very flat near YC) but they
development of shear bands in the matrix channels between clearly show the differences between interface- and matrix-
the fibers, nucleation of interface cracks, and final localiza- dominated fracture. The former occurred when N < c and
tion of the deformation in the matrix in one dominant it was characterized by a brittle behavior, while the latter
shear band. was dominant if N P 2c and led to much higher strain to
The transverse compressive strength, YC, is given by the failure (4–6%) controlled by the plastic deformation of
maximum of each curve in Fig. 6, and the overall effect of the matrix.
2802 C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806

1.2 0.1

1
0.08
epoxy/E glass
0.8
epoxy/C 0.06

Ductility
mc
0.6

YC /
0.04
0.4

0.02
0.2

0 0
0.1 1 5
N/c
Fig. 10. Influence of the interface strength (normalized by the cohesion of
the matrix c) in the transverse compressive strength, YC (normalized by
the yield strength of the matrix in compression, rmc) and in the ductility,
represented by the strain at YC. The error bars stand for the standard
deviation of the six simulations with different RVEs.

120

2c
100
c
Compressive stress (MPa)

80 0.5c

60

40
0.1c
Fig. 9. Contour plot of the accumulated plastic strain in the matrix in the
composite with intermediate interfacial strength (N = c). (a)  = 1.7% interface
corresponding to the maximum strength. (b)  = 2.5%. The loading axis fracture energy
is horizontal. Notice that the strain values in legend (b) are 10 times higher 20
100 N/m
than in (a).
10 N/m
0
0 0.01 0.02 0.03 0.04 0.05
4.3. Influence of the interface fracture energy
Compressive strain
The influence of the interface fracture energy on the Fig. 11. Influence of the interface fracture energy on the stress–strain
mechanical behavior in transverse compression is plotted curve under transverse compression for different values of the interface
in Fig. 11. Simulations were performed with the same strength.
RVE (whose behavior was very similar to the average of
six simulations with different RVEs) and three interface
fracture energies: 100 J/m2 (the baseline value), 10 J/m2 with interface fracture energies of 1000 J/m2 were practi-
and 1000 J/m2, while the interface strength was systemati- cally superposed to those with 100 J/m2 up to the maxi-
cally varied from 0.1c up to 2c. For a given value of the mum stress in all cases, even though the fracture energies
interface strength, the variations in the fracture energy differed in one order of magnitude, and the differences
modified the effective interface displacement at failure, df beyond that point were minimum. They are not plotted
(Fig. 4), leading to more brittle or more ductile behaviors. in Fig. 10 for sake of clarity. Brittle fiber/matrix interfaces
The rest of the fiber and matrix properties were those indi- (which are represented by the curves obtained with
cated in Section 3. The stress–strain curves of the materials C = 10 J/m2 in Fig. 11) did not change significantly the
with interface fracture energies of 10 and 100 J/m2 are plot- compressive strength, although the reduction in load after
ted in Fig. 11. The curves corresponding to the materials the maximum was faster as a result of the easy propagation
C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806 2803

of the cracks along particle/matrix interface. Thus, it can matrix in compression to compare the composite behavior
be concluded that the effect of the interface fracture energy on the same basis. The curves in Fig. 12a (/ = 0) are rep-
on the transverse compressive strength of fiber-reinforced resentative of a metallic matrix, which follows the Tresca
polymers is negligible, as compared with the influence of yield criterion, while those in Fig. 11b and c stand for the
the interface strength. behavior of polymeric matrices which tend to form shear
bands oriented at an angle of 45 + //2 with the plane per-
4.4. Influence of the matrix friction angle pendicular to the compression axis. In the absence of inter-
face decohesion, the matrix with / = 0 provided the
The stress–strain curves under transverse compression highest compressive strength (relative to the rmc), and
of one RVE are plotted in Fig. 12a–c for three composites YC/rmc decreased progressively with the friction angle.
with matrix friction angles of 0, 15 and 30, respectively. This behavior is the result of the trend to localize the defor-
As the matrix tensile strength was assumed to be constant mation in shear bands between the fibers, which increases
and equal to 60 MPa, changes in the friction angle modi- with the friction angle. This effect was more marked in
fied the yield strength of the matrix in compression – as presence of interface decohesion, because matrix shear
given in Eq. (5) – which increased from 60 MPa (/ = 0) bands were triggered at lower strains by the stress concen-
up to 180 MPa (/ = 30). So the stresses in Fig. 12 were trations around the interface cracks (Fig. 9a). Obviously,
normalized by the corresponding yield strength of the this mechanism is more efficient if the matrix friction angle

a 1.2 b 1.2

1 1

0.8 0.8
mc

cm

0.6 0.6
/

0.4 0.4
Interface infinity Interface infinity
strength 2c strength 2c
0.2 c 0.2 c
0.5c 0.5c
= 0º 0.1c 0.1c
= 15º
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05
Compressive strain Compressive strain

c 1.2

0.8
cm

0.6
/

0.4

Interface infinity
0.2 strength 2c
c
0.5c
= 30º 0.1c
0
0 0.01 0.02 0.03 0.04 0.05
Compressive strain
Fig. 12. Influence of the matrix friction angle on the stress–strain curve under transverse compression: (a) / = 0; (b) / = 15; (c) / = 30. The stresses are
normalized by the respective strength of the matrix in compression.
2804 C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806

is high, and thus the degradation of the composite proper- compressive strength was not affected because the
ties was faster as the interfacial strength decreased. thermo-elastic residual stresses were rapidly smoothed
out during deformation by the intense plastic deformation
4.5. Influence of the thermal residual stresses in the matrix and the interface cracks.

Residual stresses develop in PMC upon cooling at ambi- 5. Conclusions


ent temperature after curing as a result of the thermal
expansion mismatch between the matrix and fibers. As The compressive strength under transverse loading of
the thermal expansion coefficient of the epoxy matrix is fiber-reinforced polymers was studied by means of compu-
much higher than that of the fibers, tensile stress appears tational micromechanics. In this modeling strategy, the
in the matrix and compressive in the fibers, and their influ- stress–strain curve was computed by means of the finite
ence on the behavior under transverse compression can be element analysis of an RVE of the composite microstruc-
taken into account in the micromechanical model by simu- ture. The simulations showed the role played by the two
lating the composite behavior in two steps. In the first step, dominant damage mechanisms (decohesion at the inter-
the RVE was subjected to a homogeneous temperature face and shear band formation in the matrix) in control-
change of 100 C from the stress-free temperature down ling the composite strength. On the one hand, if
to ambient temperature [38]. The computational model decohesion is inhibited, failure took place by the develop-
and the fiber and matrix properties were those given in Sec- ment of shear bands in the matrix, which propagated
tion 3 but the analyses were carried out under generalized through the microstructure at an angle of ±(45 + //2)
plane strain conditions, instead of plane strain. The thick- with respect to the plane perpendicular to the compression
ness of the model (perpendicular to the X1–X2) is constant axis. The compressive strength was slightly higher than the
in plane strain simulations, and this leads to unrealistic val- matrix strength under uniaxial compression due to the
ues of the thermal residual stresses along the X3 axis. Con- additional strengthening provided by the stiff fibers. On
versely, the generalized plane strain theory assumes that the other hand, interface cracks were nucleated at very
the model lies between two parallel planes which can move low stresses in composites with weak interfaces, while
with respect to each other and can accommodate the ther- the matrix was still in the elastic regime. The stress con-
mal strain induced by the temperature change. Once the centrations at the interface crack tips nucleated plastic
residual stresses were generated, the thickness along the shear bands between neighboring cracks, and led to the
X3 axis was held constant and the RVE was deformed evolution of the cracks into large interfacial voids. Final
under uniaxial compression. The stress–strain curves with fracture occurred by the development of bands of local-
and without residual stresses of one RVE are plotted in ized deformation formed by interfacial voids linked by
Fig. 13 for different values of the interface strength. The matrix shear bands, the orientation of these bands being
matrix and fiber properties correspond to those of the controlled by the particular distribution of the fibers in
materials in Fig. 6. The non-linear deformation started at the RVE. When the interface strength was similar to the
lower strains in the presence of residual stresses, but the matrix flow stress in compression (N  c), the numerical
simulation showed that the maximum strength was mainly
120 controlled by the matrix, and coincided with the forma-
∞ tion of an incipient pattern of shear bands in the matrix,
inclined at ±(45 + //2) with respect to the plane perpen-
100 dicular to the compression axis. Final fracture took place
c
thereafter by the propagation of a dominant shear band,
Compressive stress (MPa)

80 0.5c slightly misoriented with respect to the theoretical angle,


whose path was dictated by the linking up of interface
cracks in adjacent fibers.
60
Parametrical studies showed that other factors (such as
the matrix friction angle, the interface fracture energy
0.1c
40 and the thermo-elastic residual stresses) exerted a second-
ary influence on the compressive strength of PMC under
transverse compression. The matrix was more susceptible
20
Without residual stresses to the formation of shear bands as the friction angle
With residual stresses increased, and they developed earlier, but this effect was
0 offset by the higher matrix flow stress in compression.
0 0.01 0.02 0.03 0.04 0.05 Thermal residual stress reduced the stress for the onset of
Compressive strain nonlinear deformation but they were rapidly smoothed
Fig. 13. Influence of the thermal residual stresses on the mechanical out by the intense plastic deformation in the matrix and
response under transverse compression for different values of the interface did not modify the compressive strength. Finally, changes
strength. in the interface fracture energy by two orders of magnitude
C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806 2805

did not modify significantly the compressive strength [15] LLorca J, Segurado J. Three-dimensional multiparticle cell simula-
either. tions of deformation and damage in sphere-reinforced composites.
Mater Sci Eng A 2004;365:267–74.
It is finally worth noting the potential of computational [16] Youssef S, Maire E, Gaertner R. Finite element modelling of the
micromechanics to assess the mechanical behavior of engi- actual structure of cellular materials determined by X-ray tomogra-
neering composites. By using the appropriate constitutive phy. Acta Mater 2005;53:719–30.
equations for the fiber, matrix and interfaces, this simula- [17] Borbély A, Kenesei P, Biermann H. Estimation of the
tion tool can provide a detailed picture of deformation effective properties of particle-reinforced metalmatrix composites
from microtomographic reconstructions. Acta Mater 2006;54:
and fracture mechanisms at microscopic level, including 2735–44.
the effect of all non-linear processes and of the interaction [18] González C, LLorca J. Multiscale modeling of fracture in fiber-
among them. This information can be used to develop reinforced composites. Acta Mater 2006;54:4171–81.
more accurate and reliable failure criteria at the lamina [19] González C, LLorca J. Virtual fracture testing of fiber-reinforced
level, which in turn can be used to predict the mechanical composites: a computational micromechanics approach. Eng Fract
Mech 2007;74:1126–38.
performance of laminates and composite structures. [20] Collings TA. Transverse compressive behaviour of unidirec-
tional carbon fibre reinforced plastics. Composites 1974;5:
Acknowledgments 108–16.
[21] Pinho ST, Iannucci L, Robinson P. Physically-based failure models
This investigation was supported by the Spanish Minis- and criteria for laminated fibre-reinforced composites with emphasis
on fibre-kinking. Part I: development. Compos: Part A 2006;37:
try of Education and Science through the Grant MAT 63–73.
2006-2602 and by the Comunidad de Madrid through the [22] Huang YK, Frings PH, Hennes E. Mechanical properties of Zylon/
program ESTRUMAT-CM (reference MAT/0077). epoxy composite. Compos: Part B 2002;33:109–15.
[23] Puck A, Schürmann H. Failure analysis of frp laminates by means of
physically based phenomenological models. Compos Sci Technol
References 2002;62:1633–62.
[24] Aragonés D. Fracture micromechanisms in C/epoxy composites
[1] Aveston J, Cooper GA, Kelly A. Single and multiple fracture. In: The under transverse compression, Master thesis, Universidad Politécnica
properties of fibre composites. IPC Science and Technology Press; de Madrid, 2007.
1971. p. 15–26. [25] Vogler TJ, Kyriakides S. Inelastic behavior of an AS4/PEEK
[2] Curtin WA, Takeda N. Tensile strength of fiber-reinforced compos- composite under combined transverse compression and shear. Part
ites: II. Application to polymer matrix composites. J Compos Mater I: Experiments. Int J Plast 1999;15:783–806.
1998;32:2060–81. [26] Hsiao HM, Daniel IM. Strain rate behavior of composite materials.
[3] González C, LLorca J. Micromechanical modelling of deformation Compos: Part B 1998;29:521–33.
and failure in Ti–6Al–4V/SiC composites. Acta Mater [27] Kinloch AJ, Young RJ. Fracture behavior of polymers. Elsevier
2001;49:3505–19. Applied Science; 1983.
[4] Argon AS. Fracture of composites. In: Herman H, editor. Treatise on [28] Hashin Z. Failure criteria for unidirectional fiber composites. J Appl
materials science and technology, vol. 1. Academic Press; 1972. p. Mech 1980;47:329–34.
79–114. [29] Segurado J, LLorca J. A new three-dimensional interface finite
[5] Budiansky B, Fleck NA. Compressive failure of fiber composites. J element to simulate fracture in composites. Int J Solids Struct
Mech Phys Solids 1993;41:183–211. 2004;41:2977–93.
[6] Vogler TJ, Hsu S-Y, Kyriakides S. Composite failure under combined [30] Segurado J, LLorca J. A computational micromechanics study of the
compression and shear. Int J Solids Struct 2000;37:1765–91. effect of interface decohesion on the mechanical behavior of
[7] Michel JC, Moulinec H, Suquet P. Effective properties of composite composites. Acta Mater 2005;53:4931–42.
materials with periodic microstructure: a computational approach. [31] Hashagen F, de Borst R. Numerical assessment of delamination in
Comput Meth Appl Mech Eng 1999;172:109–43. fibre metal laminates. Comput Meth Appl Mech Eng 2000;185:
[8] Lusti HR, Hine PJ, Gusev AA. Direct numerical predictions for the 141–59.
elastic and thermoelastic properties of short fibre composites. [32] Ghosh S, Ling Y, Majumdar B, Kim R. Interfacial debonding
Compos Sci Technol 2002;62:1927–34. analysis in multiple fiber reinforced composites. Mech Mater
[9] Segurado J, LLorca J. A numerical approximation to the elastic 2000;32:561–91.
properties of sphere-reinforced composites. J Mech Phys Solids [33] Brockenbrough JR, Suresh S, Wienecke HA. Deformation of
2002;50:2107–21. metal–matrix composites with continuous fibers: geometrical effects
[10] González C, Segurado J, LLorca J. Numerical simulation of elasto- of fiber distribution and shape. Acta Metall Mater 1991;39:
plastic deformation of composites: evolution of stress microfields and 735–52.
implications for homogenization models. J Mech Phys Solids [34] Hine PJ, Lusti HR, Gusev AA. Numerical simulation of the effects of
2004;52:1573–93. volume fraction, aspect ratio and fibre length distribution on the
[11] Chawla N, Sidhu RS, Ganesh VV. Three-dimensional visualization elastic and thermoelastic properties of short fibre composites.
and microstructure-based modeling of deformation in particle-rein- Compos Sci Technol 2002;62:1445–53.
forced composites. Acta Mater 2006;54:1541–8. [35] Abaqus, Users’ Manual, ABAQUS Inc; 2006.
[12] Segurado J, González C, LLorca J. A numerical investigation of the [36] Suquet P. Effective properties of nonlinear composites. In: Contin-
effect of particle clustering on the mechanical properties of compos- uum micromechanics. CISM Course and Lecture Notes; 1997, p. 197–
ites. Acta Mater 2003;51:2355–69. 264.
[13] Segurado J, LLorca J. Computational micromechanics of composites: [37] Hazanov S, Huet C. Order relationships for boundary condition
the effect of particle spatial distribution. Mech Mater 2006;38:873–83. effects in heterogeneous bodies smaller than the representative
[14] Böhm HJ, Han W, Eckschlager A. Multi-inclusion unit cell studies of volume. J Mech Phys Solids 1994;42:1995–2011.
reinforcement stresses and particle failure in discontinuously reinforced [38] Soden PD, Hinton MJ, Kaddour AS. Lamina properties, lay-up
ductile matrix composites. Comput Model Eng Sci 2004;5:5–20. configurations and loading conditions for a range of fibre-
2806 C. González, J. LLorca / Composites Science and Technology 67 (2007) 2795–2806

reinforced composite laminates. Compos Sci Technol 1998;58: [42] Khisaeva ZF, Ostoja-Starzewski M. On the size of RVE in
1011–22. finite elasticity of random composites. J Elast 2006;85:
[39] Abaqus, Theory manual, HKS Inc; 1998. 153–73.
[40] Menetrey P, Willam KJ. Triaxial failure criterion for concrete and its [43] Pitkethly MJ et al. A round robin programme on interfacial test
generalization. ACI Struct J 1995;92:311–8. methods. Compos Sci Technol 1993;48:205–14.
[41] Zhou X-F, Wagner HD, Nutt SR. Interfacial properties of polymer [44] Benzarti K, Cangemi L, Maso FD. Transverse properties of unidi-
composites measured by push-out and fragmentation tests. Compos: rectional glass/epoxy composites: influence of fibre surface treat-
Part A 2001;32:1543–51. ments. Compos: Part A 2001;32:197–206.

Anda mungkin juga menyukai