Vous êtes sur la page 1sur 27

3D analysis of stress transfer in the

micromechanics of ber reinforced composites


by using an eigen-function expansion method
Z.J. Wu
1
, J.Q. Ye*, J.G. Cabrera
School of Civil Engineering, The University of Leeds, Leeds, LS2 9JT, UK
Received 11 February 1999; received in revised form 19 July 1999
Abstract
This paper presents an exact solution for an inhomogeneous, transversely isotropic,
elastic circular cylinder subjected to axisymmetric force and displacement boundary
conditions. The solution is obtained on the basis of an eigen-function expansion method
and can satisfy all the boundary conditions prescribed on the curved and end surfaces of
the cylinder. It can be used directly in the micromechanical analysis of ber reinforced
composites to investigate the typical Representative Volume Element (RVE). The element
consists of a combined circular cylinder composed of a solid inner circular cylinder of
transversely isotropic ber and a concentric outer circular cylinder of isotropic matrix
material. Using this solution, all the stress and displacement components of both the inner
ber and the outer matrix, and hence the stress transfer in the interface between the ber
and matrix, are expressed analytically. The numerical results presented show that stress
concentration occurs near the ends of the cylinder where external forces are applied. # 2000
Elsevier Science Ltd. All rights reserved.
Keywords: A. Microstructures; Stress transfer; B. Fiber-reinforced composite materials; C. Analytic
functions
0022-5096/00/$ - see front matter # 2000 Elsevier Science Ltd. All rights reserved.
PII: S0022- 5096( 99) 00058- 7
Journal of the Mechanics and Physics of Solids
48 (2000) 10371063
www.elsevier.com/locate/jmps
1
On leave from Tsinghua University, Beijing 1000084, China
* Corresponding author. Fax: +44-0113-2332265.
E-mail address: j.ye@leeds.ac.uk (J.Q. Ye).
1. Introduction
The research on stress transfer at the interface of two dierent materials
occupies a prominent position in the micromechanics of ber reinforced
composites because it is essential to understand how and to what extent the
interface properties inuence the mechanical performance and fracture behavior of
composites. At present, the theories about the strength and toughness of ber
reinforced composites are predominantly based on various assumptions on how
stresses are transferred between bers and matrix. To investigate the stress
transfer, a combined cylinder composed of an inner ber and a surrounding
matrix is normally used as a RVE of a ber reinforced composite. In the past few
decades, signicant work has been done in order to achieve a better understanding
of the stress transfer mechanism. Some investigators (Aveston and Kelly, 1973;
Kelly, 1970) assumed that a constant shear stress acted at the bermatrix
interface. This constant shear stress was often interpreted as the ow stress of the
matrix (especially in metalmatrix composites) or as the friction stress at the
interface. By considering the equilibrium of a ber isolated from matrix, the
average ber stress obtained varies linearly along the ber and is proportional to
the shear stress. Clearly, this solution is highly approximate and only the global
equilibrium in the ber direction can be satised.
By considering the combination of material's constitutive behavior, Cox (1952)
proposed a shear-lag model which is the most widely used in the study of the
micromechanics of stress transfer across the bermatrix interface. Practically this
model is based on the assumption used in the classical theory of strength of
materials (Timoshenko, 1957). Although several improvements on this model have
been introduced thereafter (Gao et al., 1988; Piggott, 1980; Rosen, 1964), there is
still no universal measure to assess these models and corresponding results.
Researches on stress transfer have also been carried out for ber pull- or push-
out (Hsueh, 1992; Keran and Parthasarathy, 1991; Kim and Mai, 1998), ber
fragmentation (Curtin, 1991) and rod pullout in concrete. These researches
provided important information for predicting interfacial debonding and friction
sliding. For example, based on the assumption of shear stress distribution at the
interface, McCartney (1989) gave a rened approximate solution on stress transfer
in ber pull-out. However, it has been recognized that the whole picture of
interface stress transfer has not yet been well understood (Tsai and Kim, 1996)
though many experimental studies (Atkinson et al., 1982; Cook et al., 1989) and
theoretical analyses (Freund, 1992; Hutchinson and Jensen, 1990; Povirk and
Needleman, 1993) have been done.
Recognizing the signicance of stress concentration at the ber broken ends, the
researchers in this area have also encountered the stress transfer problem when
they wanted to perform an elegant analysis. Many investigators have therefore
being inclined to employ numerical methods, particular the FE analysis by which
the eects of specic end conditions as well as the interface properties can be
properly evaluated (Case and Reifsnider, 1996; Marotzke, 1994).
Based on the above discussions, it is obvious that stress transfer is still a
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1038
challenging topic for mathematicians and engineers. Using the RVE model to
study stress transfer involves solutions of 3D boundary value problems in
cylindrical coordinates. Some of these problems are relatively simple. For
example, there are some successful results including the potential function solution
for determination of stresses in load decay or diusion in innite or semi-innite
isotropic domains (Dollar and Steif, 1988; Knowles and Horgan, 1969; Muki and
Sternberg, 1969). Some others, however, are more complicated, particularly, when
stress transfer in a nite domain (e.g., nite length combined cylinder) is
considered. In fact, even for homogeneous isotropic materials, determining the
stresses within an elastic circular cylinder subjected to prescribed forces or
displacements at its surface, is one of the classical problems of the mathematical
theory of elasticity which has never been solved completely (Lur'e, 1964). About a
century ago, Chree (1889) and Filon (1902) obtained solutions for isotropic
materials by the method of series expansion. However, these solutions cannot
achieve a complete satisfaction of boundary conditions on all surfaces even for
axisymmetric problems. In this aspect, two research papers deserve mention. One
was contributed by Flu gge and Kelkar (1968). Based on a similar eigenvalue
method which was used by Filon (1902), Flu gge and Kelkar analyzed an isotropic
nite cylinder by superimposing the solutions of an innitely long cylinder and a
semi-innite one (in Flu gge's original paper, they were respectively called the rst
and the second fundamental problem). By careful observation, we nd that the
two solutions corresponding to the two fundamental problems can be represented
by a single solution if the eigenvalues are allowed to be complex. Another
important contribution was given by Smith and Spencer (1970). Interfacial
tractions of combined isotropic composite cylinders with nite length were
investigated in that paper. By using a dierent method of eigenfunction expansion
and superimposing a classical Lame solution to the expansion, they obtained some
results for the cylinders. The method used to nd the complex eigenvalues,
however, was very complicated and had the shortcoming that eigenvalues could be
missed even if major computational penalty of using small intervals of the trial
values was accepted. It was reported in that paper that they could not prove if the
eigenfunctions they found numerically were complete.
In the present paper, a more general solution to the stress transfer in a
transversely isotropic cylinder is given. To the authors' best knowledge, this
solution is not available in the literature. The governing equations for the cylinder
are represented by using 3D state space approach of elasticity. The solution is
then sought analytically. Applying this solution to investigate the stress transfer in
a RVE of a ber-reinforced composite, it is found that all the boundary
conditions prescribed at the curved and end surfaces of the cylinder can be
satised.
The arrangement of subsequent sections of this paper is as follows. The
composite model for unidirectional ber reinforced composites and the
corresponding basic equations are outlined in section 2. The governing equation
of the problem is also given in this section, with more details being provided in
Appendix A. In section 3 general solutions for isotropic and transversely isotropic
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1039
axisymmetric cylinders are given respectively. The application of these solutions to
a RVE, i.e., a composite combined cylinder of nite length, is then studied in
section 4. After introducing all the boundary conditions of the cylinder, the
unknown coecients appearing in the solutions are determined uniquely by a set
of linear algebra equations. Numerical results in section 5 demonstrate the
variations of stress and displacement components within the RVE. Shear stress at
the interface between ber and matrix of the cylinder elements changes sharply
along the ber axis within a length of a quarter of the ber radius. Finally, section
6 presents the conclusions of this investigation.
2. Model and basic equations
2.1. Model assumption
An appropriate analytical model (or RVE) selection is the basis of the research
on micromechanics of ber reinforced composites. A RVE must reect the main
properties of a practical composite material. To select a proper model, one has to
know the distribution of the reinforcement (ber) in the matrix. Based on the
successful experiences gained by the researchers in this area and the experimental
results obtained by Wu (1991), a hexagonal array ber distribution (Fig. 1) is
considered to be a reasonable model for unidirectional ber reinforced composites.
To simplify the theoretical analysis, the following has been assumed.
1. The ber is elastic and transversely isotropic.
2. The matrix is made of a linear elastic isotropic material.
3. Having the same ber volume fraction V
f
, the cylinder model (Fig. 2) is
equivalent to the hexagonal prism model (Fig. 1).
Assumption (3) suggests that the nal RVE to be analyzed is a coaxial combined
cylinder composed of an inner circular ber cylinder, surrounded by a concentric
circular cylinder of matrix material. The same circular cylindrical RVE was also
Fig. 1. Hexagonal array of bers in a composite.
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1040
used by Smith and Spencer (1970) and some other investigators. The problem to
be solved now is to nd the stresses (or displacements) in the composite cylinder,
and, in particular, at the interface, when specied displacements or stresses are
applied to its surface including its ends.
2.2. Basic equations
Generally, the basic equations of three-dimensional elasticity in the absence of
body forces can be written as follows:
s
ij, j
= 0, (1)
s
ij
= c
ijkl
e
kl
, (2)
e
ij
=
1
2
(u
j, i
u
i, j
), (3)
where s
ij
, e
ij
and u
i
are the components, respectively, of stress, strain and
displacement, and c
ijkl
is the elastic stiness tensor. Considering the model
assumptions introduced above, Eqs. (1)(3) can be expressed explicitly below in
cylindrical polar coordinates (r, y, z ) for transversely isotropic material
@s
r
@r

1
r
@t
ry
@y

s
r
s
y
r

@t
rz
@z
= 0
Fig. 2. RVE of a ber reinforced composite.
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1041
@t
ry
@r

2t
ry
r

1
r
@s
y
@y

@t
yz
@z
= 0
@t
rz
@r

t
rz
r

1
r
@t
yz
@y

@s
z
@z
= 0, (4)
_

_
s
r
s
y
s
z
t
yz
t
rz
t
ry
_

_
=
_
_
_
_
_
_
_
_
_
_
c
11
c
12
c
13
0 0 0
c
12
c
11
c
13
0 0 0
c
13
c
13
c
33
0 0 0
0 0 0 c
44
0 0
0 0 0 0 c
44
0
0 0 0 0 0
c
11
c
12
2
_

_
_

_
e
r
e
y
e
z
g
yz
g
rz
g
ry
_

_
, (5)
and
e
r
=
@u
@r
, g
ry
=
1
r
@u
@y

@v
@r

v
r
e
y
=
u
r

1
r
@v
@y
, g
yz
=
1
r
@w
@y

@v
@z
e
z
=
@w
@z
, g
rz
=
@u
@z

@w
@r
, (6)
where u, v and w are, respectively, the displacements in r, y and z directions. For
axial symmetric (about the z-axis) case, the non-zero stress and displacement
components are s
r
, s
y
, s
z
, t
rz
, u and w, respectively. In order to derive the nal
governing equation of the symmetric problem systematically, a state space
approach is adopted based on Eqs. (4)(6) (see Appendix A)
@
@r
_

_
u
w
s
r
t
rz
_

_
=
_
_
_
_
_
_
_
_
_
_
_
_
_

c
5
r
c
6
a c
7
0
a 0 0
1
G
rz
c
1
r
2
c
2
r
a
c
4
r
a

c
2
r
a c
3
a
2
c
6
a
1
r
_

_
_

_
u
w
s
r
t
rz
_

_
(7)
and the other two stress components can be found subsequently from
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1042
_
s
y
s
z
_
=
_
_
_
_
c
1
r
c
2
a c
5
0
c
2
r
c
3
a c
6
0
_

_
_

_
u
w
s
r
t
rz
_

_
: (8)
In the above equations a=@/@z denotes a dierential operator. The constants c
k
(k
= 1, 2, . . . 7) are c
1
= c
11
c
2
12
=c
11
, c
2
=c
13
c
13
c
12
/c
11
, c
3
= c
33
c
2
13
=c
11
,
c
4
=c
12
/c
11
1, c
5
=c
12
/c
11
, c
6
=c
13
/c
11
, c
7
=1/c
11
and G
rz
=c
44
. If the material is
isotropic, these constants are c
1
=c
3
=E/(1n
2
), c
2
=n E/(1n
2
), c
4
=(2n1)/
(1n ), c
5
=c
6
=n/(1n ), c
7
=(1+n )(12n )/[E(1n )], where E is Young's modulus
and n the Poisson's ratio.
3. Solutions of the governing equation
There are many existing methods to solve a state equation which is frequently
used in modern automatic control systems (Fairman, 1998). Most of these
methods are based on numerical solutions. In this study, analytical solutions are
sought to solve the state Eq. (7) (Wu, 1991).
Using algebraic manipulation of operators, an equivalent high order governing
equation with respect to any of the variables u, w, s
r
and t
rz
can be obtained from
the state Eq. (7). For example, the governing equation about w, the displacement
in the z-axis direction, is
(V
2
x
1
a
2
)(V
2
x
2
a
2
)w = 0 (9)
where
V
2
=
1
r
d
dr
_
r
d
dr
_
is Laplace's operator in the axisymmetric polar coordinate system. x
1
and x
2
are
calculated from
x
1
x
2
=
1
2
_
_
_
_
c
33
G
rz

G
rz
c
11

(c
13
G
rz
)
2
c
11
G
rz
_
3

_
c
33
G
rz

G
rz
c
11

(c
13
G
rz
)
2
c
11
G
rz
_
2

4c
33
c
11

_
_

_
: (10)
Assuming that there exists an eigenvalue solution for Eq. (9), it is readily veried
by the method of separation of variables that the solution has the following form
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1043
w
k
= (b
1
sh

a
k
_
z b
2
ch

a
k
_
z)[D
1
J
0
(

a
1k
_
r) D
2
Y
0
(

a
1k
_
r)
D
3
J
0
(

a
2k
_
r) D
4
Y
0
(

a
2k
_
r)]: (11)
Similarly, the governing equation for the displacement u can be obtained and the
corresponding solution is
u
k
= (b
1

a
k
_
ch

a
k
_
z b
2

a
k
_
sh

a
k
_
z)[F
1
J
1
(

a
1k
_
r) F
2
Y
1
(

a
1k
_
r)
F
3
J
1
(

a
2k
_
r) F
4
Y
1
(

a
2k
_
r)], (12)
in which the J
i
, and Y
i
(i = 0, 1) are, respectively, the Bessel functions of the rst
and the second kinds of order i. Integral constants b
1
, b
2
, D
j
, F
j
( j = 1, 2, 3, 4)
and the eigenvalue a
k
are to be determined. The unknown constants have the
following relation
D
j
=
(G
rz
c
11
b
j
)
(c
13
G
rz
)

b
j
_

a
k
_
F
j
, (13)
where
b
j
=
_
x
1
( j = 1, 2)
x
2
( j = 3, 4)
(x
1
,= x
2
) (14)
and
a
ik
= x
i
a
k
(i = 1, 2): (15)
Specially, corresponding to zero eigenvalue, the solutions can be obtained by
considering the limit case of Eq. (9) or by solving Eq. (7) independently. The
following two linearly independent solutions are respectively found
w
0
= (b
10
z)(D
30
D
40
ln r), (16)
u
0
= D
10
r D
20
1
r

c
13
G
rz
2c
11
D
40
r ln r, (17)
where b
10
and D
j0
( j = 1, 2, 3, 4) are also integral constants. Combining (11) and
(12) with (16) and (17), respectively, the eigenvalue solutions which are
appropriate to the boundary value problem considered are
w =

o
k=1
w
k
w
0
(18)
and
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1044
u =

o
k=1
u
k
u
0
: (19)
The solutions (18) and (19) are more general than those given by Smith and
Spencer (1970) because both symmetric and antisymmetric displacements about z
= 0 are included here. Moreover, the present solutions are for transversely
isotropic materials from which Smith and Spencer's solutions for isotropic
materials can be obtained as a special case.
For isotropic materials, solutions (18) and (19) cannot be used because x
1
=x
2
.
The following displacements can be used for this case
w
k
= (b
1
sh

a
k
_
z b
2
ch

a
k
_
z)[D
1
J
0
(

a
k
_
r) D
2
Y
0
(

a
k
_
r)
D
3

a
k
_
r tJ
1
(

a
k
r
_
) D
4

a
k
r
_
Y
1
(

a
k
_
r)],
(20)
u
k
= (b
1

a
k
_
ch

a
k
_
z b
2

a
k
_
sh

a
k
_
z)[F
1
J
1

a
k
_
r) F
2
Y
1
(

a
k
_
r)
F
3

a
k
_
r J
0
(

a
k
_
r) F
4

a
k
_
r Y
0
(

a
k
_
r)],
(21)
where
D
1
=

a
k
_
[F
1
4(1 n)F
3
], D
3
=

a
k
_
F
3
D
2
=

a
k
_
[F
2
4(1 n)F
4
], D
4
=

a
k
_
F
4
: (22)
For zero eigenvalue solutions, it can be proved that the solution forms (16) and
(17) can be used for both isotropic and transversely isotropic materials. Hence,
after substituting isotropic elastic constants into (16) and (17), zero eigenvalue
solutions for isotropic materials are
w
0
= (b
10
z)(D
30
D
40
ln r), (23)
u
0
= D
10
r D
20
1
r

1
4(1 n)
D
40
r ln r: (24)
It is noted that the classical Lame solution is a special case of the present zero-
eigenvalue solution.
4. Boundary conditions
From solutions (11) and (12) or (20) and (21), it is found that for a given k
there are six independent parameters which are in terms of the unknown
coecients, b
1
, b
2
, D
j
or F
j
( j = 1, 2, 3, 4) and the eigenvalue a
k
. The unknown
coecients should have been identied by using k as a subscript in the previous
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1045
sections. For convenience, this subscript was omitted but is used from now as one
of the subscripts. Moreover, in order to introduce boundary conditions and
distinguish displacement and stress components in dierent materials, f and m are
used as superscripts to indicate ber and matrix, respectively.
The problem considered here is the model shown in Fig. 3. The combined
cylinder with a radius of r
m
and a length of 2 l, is subjected to traction P on the
ber ends. The transversely isotropic ber occupies the region 0 R r < r
f
. The
matrix occupies the region r
r
< r < r
m
and has Young's modulus E and Poisson's
ratio n. The two ends of the matrix cylinder are, respectively, constrained by two
xed frictionless rigid plates. The outer curved surface of the matrix can be
specied by either force boundary conditions (Case 1) or mixed conditions (Case
2). The RVE model can be used to simulate the stress transfer in some pull-out
problems. Because of symmetry in the z-direction, the origin of the z coordinate is
Fig. 3. Combined composite cylinder of RVE subjected to external force P acting on both ends of the
bre.
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1046
laid on the symmetric plane (Fig. 3). Two dierent types of boundary conditions
are studied below.
4.1. Case 1
The cylinder has a stress-free outer surface. The boundary conditions of this
case are
Continuity conditions at the interface:
w
f
[
r=r
f
w
m
[
r=r
f
= 0 (0Rz < l ) (25)
u
f
[
r=r
f
u
m
[
r=r
f
= 0 (0Rz < l ) (26)
t
f
rz
[
r=r
f
t
m
rz
[
r=r
f
= 0 (0Rz < l ) (27)
s
f
r
[
r=r
f
s
m
r
[
r=r
f
= 0 (0Rz < l ): (28)
Surface conditions:
t
m
rz
[
r=r
m
= 0 (0Rz < l ) (29)
s
m
r
[
r=r
m
= 0 (0Rz < l ): (30)
End conditions:
w
m
[
z=0
= 0, t
m
rz
[
z=0
= 0 (r
f
< r < r
m
) (31)
w
m
[
z=l
= 0, t
m
rz
[
z=l
= 0 (r
f
< r < r
m
) (32)
w
f
[
z=0
= 0, t
f
rz
[
z=0
= 0 (0Rr < r
f
) (33)
s
f
z
[
z=l
= p(r), t
f
rz
[
z=l
= 0 (0Rr < r
f
): (34)
Using (18) and (19), it is easy to nd all the stresses from (5) and (6). Before
substituting these displacements and stresses into boundary conditions (25)(34), it
is noticed that the coecients D
f
2
, D
f
4
(or F
f
2
, F
f
4
) and D
f
20
, D
f
40
in the solutions
(11), (12) and (16), (17) reect the singularities in the case of a solid ber cylinder.
To remove these singularities of displacements and stresses at r = 0, the singular
terms must vanish, i.e.,
D
f
2
= D
f
4
= D
f
20
= D
f
40
= 0: (35)
For the matrix, (31) and (32) lead to
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1047
b
m
2
= D
m
30
= D
m
40
= 0, a
k
=
_
kp
l
_
2
: (36)
For the ber, (33) gives
b
f
2
= b
f
10
= 0: (37)
To satisfy the rst end condition in (34), the applied external stress p(r ) must be
known in advance. In practical engineering, however, it is impossible to specify
the exact distribution of the force in the ber end of a RVE. In a realistic case
such as pull-out or push-out tests, only the resultant P acting on the ber end can
be measured. This condition can, therefore, be approximately replaced by
2
_
r
f
0
s
f
z
[
z=l
r dr = r
2
f
p
0
, (38)
where p
0
=P/(pr
2
f
) is the nominal average stress applied at the ber end. It has to
be mentioned here that by using (38) the distribution of p(r ) at z=l can only be
specied after a convergent solution of the problem has been found. In the
numerical example shown later, the calculated real distribution of p(r ) at z=l is
shown in Fig. 10. In other words, using (34) to achieve the same results as using
(38), the applied stress p(r ) at z=l must be prescribed exactly in the form shown
in Fig. 10. However, from the discussions in section 5, if the load is applied
suciently far from z=l (see Fig. 3), its distribution on the loaded end can be in
any form as long as the resultant is P and the resulting p(r) at z=l will be the
same as shown in Fig. 10. As mentioned before, in a real engineering case, e.g., a
pull-out test, it is hardly possible to prescribe a stress at z=l and the external
forces are normally applied at the far end of the ber. This case can always be
represented by the RVE shown in Fig. 3. In this respect, the solutions obtained
based on (38) will provide stress distributions within the RVE no matter how the
tensile force is applied at z > l. The second condition of (34), i.e., t
f
rz
[
z=l
= 0, is
found to be satised automatically.
The four continuity conditions at interface (25)(28) and the two surface
conditions (29), (30), along with (38), form a set of linear algebraic equations
from which the remaining coecients in the solutions of the present case can be
determined. To this end, the displacement solutions of the ber and the matrix are
rewritten below after considering (35)(37). They are, respectively
w
f
=

o
k=1
sh

a
k
_
z[D
f
1k
J
0
(

a
1k
_
r) D
f
3k
J
0
(

a
2k
_
r)] D
f
30
z, (39)
u
f
=

o
k=1
ch

a
k
_
z

a
k
_
[F
f
1k
J
1
(

a
1k
_
r) F
f
3k
J
1
(

a
2k
_
r)] D
f
10
r (40)
and
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1048
w
f
=

o
k=1
sh

a
k
_
z[D
m
1k
J
0
(

a
k
_
r) D
m
2k
Y
0
(

a
k
_
r) D
m
3k

a
k
_
rJ
1
(

a
k
_
r)
D
m
4k

a
k
_
rY
1
(

a
k
_
r)], (41)
u
m
=

o
k=1
ch

a
k
_
z

a
k
_
[F
m
1k
J
1
(

a
k
_
r) F
m
2k
Y
1
(

a
k
_
r) F
m
3k

a
k
_
rJ
0
(

a
k
_
r)
F
m
4k

a
k
_
rY
0
(

a
k
_
r)] D
m
10
r
D
m
20
r
: (42)
Substituting (39)(42) into (6) and then (5) gives all the stresses in terms of the
eigenvalues and the unknown constants. For solutions corresponding to each
eigenvalue, the continuity conditions at interface and the surface conditions at
outer cylinder must be satised.
By comparing the coecients in the corresponding expansion for zero
eigenvalue solution, continuity conditions at interface (26), (28) and surface
condition (30) give
D
f
10
= D
m
10

D
m
20
r
2
f
, (43)
(c
11
c
12
)D
f
10
c
13
D
f
30
=
E
(1 n)(1 2n)
_
D
m
10
(1 2n)
D
m
20
r
2
f
_
, (44)
D
m
10
(1 2n)
D
m
20
r
2
m
= 0: (45)
From (43)(45), D
m
10
, D
m
20
and D
f
10
can be represented in terms of D
f
30
. They are,
respectively
D
m
10
=
(1 2n)V
f
1 (1 2n)V
f
ZV
m

c
13
c
11
c
12
D
f
30
, (46)
D
m
20
=
r
2
f
1 (1 2n)V
f
ZV
m

c
13
c
11
c
12
D
f
30
, (47)
D
f
10
=
1 (1 2n)V
f
1 (1 2n)V
f
ZV
m

c
13
c
11
c
12
D
f
30
, (48)
in which V
m
=1V
f
; V
f
(=r
2
f
/r
2
m
) is the volume fraction of the ber reinforced
composite; Z = 2G=(c
11
c
12
) and G is the shear modulus of matrix.
Corresponding to an arbitrary non-zero eigenvalue a
k
, boundary conditions
(25)(30) lead to a set of six independent linear equations in terms of the
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1049
unknown coecients, i.e.,
[A
k
B
k
]
_
F
k
D
f
30
_
= {0], (49)
where, B
k
= [2 (1)
k1
=a
k
0 0 0 0 0]
T
, F
k
= [F
f
1k
F
f
3k
F
m
1k
F
m
2k
F
m
3k
F
m
4k
]
T
and {0} is a (6 1) zero matrix. A
k
is a (6 6) constant matrix, the
elements of which are shown in Appendix B. Superscript T denotes the transpose
of a matrix. These six equations are arranged in the same order as that of the
boundary conditions listed in (25)(30).
During the formation of (49), the displacements and stresses corresponding to
zero eigenvalue solutions have been appropriately expanded into series expansions.
For example, to have the rst equation in (49), w
f
of (39) is expressed by
w
f
=

o
k=1
sh

a
k
_
z[D
f
1k
J
0
(

a
1k
_
r) D
f
3k
J
0
(

a
2k
_
r)
2D
f
30

a
k
_ (1)
k1
]
(0Rz < l )
(50)
and compared with (41) to satisfy the continuity condition at the interface (25).
Following a similar procedure, the other equations in (49) can be found by
imposing relevant boundary conditions in (26)(30).
Eqs. (49) and (38) form a linear algebraic system from which the coecients F
k
and D
f
30
are solved. The solution of the boundary value problem can then be
constructed. Detailed numerical calculations are given in section 5.
4.2. Case 2
The outer surface of the combined cylinder is free from shear stress and has no
displacement in the r direction. In this case, all the boundary conditions are the
same as those specied in Case 1 except Eq. (30). It will be replaced by
u
m
[
r=r
m
= 0 (0Rz < l ): (51)
The solution to this case can be found by following the similar procedure as
shown in Case 1 except that the treatment of the outer surface condition is
dierent. The following displacement condition is used to replace Eq. (45) of
Case 1
D
m
10

D
m
20
r
2
m
= 0: (52)
From (42), (43) and (52), the correlations between D
m
10
, D
m
20
, D
f
10
and D
f
30
for this
case are
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1050
D
m
10
=
V
f
Z V
m

ZV
f
1 2n

c
13
c
11
c
12
D
f
30
, (53)
D
m
20
=
r
2
f
Z V
m

ZV
f
1 2n

c
13
c
11
c
12
D
f
30
, (54)
D
f
10
=
V
m
Z V
m

ZV
f
1 2n

c
13
c
11
c
12
D
f
30
: (55)
As a result of Eq. (51), the elements in the last row of matrix A
k
in Eq. (49) are
also changed (Appendix B).
To obtain a satisfactory solution, a suitably large number of the eigenvalue
expansions is required. If taking the rst n terms (k = 1, 2, . . . n) in the solutions
(18) and (19), the algebraic equations, i.e., Eqs. (49) and (38) can be represented
in the following matrix form
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
A
1
. . . B
1
.
.
.
A
2
B
2
.
.
.
.
.
.
A
k
B
k
.
.
.
.
.
.
.
.
.
. . . A
n
B
n
H
1
H
2
. . . H
k
. . . H
n
A
6n1, 6n1
_

_
_

_
F
1
F
2
.
.
.
F
k
.
.
.
F
n
D
f
30
_

_
=
_

_
0
0
.
.
.
0
.
.
.
0
p
0
_

_
, (56)
in which H
k
is a (1 6) matrix (see Appendix B). Both H
k
and A
6n + 1,6n + 1
are
formed from Eq. (38), whereas
A
6n1, 6n1
= c
33

1 (1 2n)V
f
1 (1 2n)V
f
ZV
m

2c
2
13
c
11
c
12
(57)
for the stress-free case (Case 1) and
A
6n1, 6n1
= c
33

V
m
Z V
m

ZV
f
1 2n

2c
2
13
c
11
c
12
(58)
for Case 2. The 6n+1 unknown coecients F
k
= [F
f
1k
F
f
3k
F
m
1k
F
m
2k
F
m
3k
F
m
4k
]
T
(k = 1, 2, . . . n) and D
f
30
can then be determined by solving the 6n +
1 linear equations (56).
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1051
5. Numerical calculations
As explained in the Introduction, stress transfer at the interface between ber
and matrix in a composite is the main interest of this study. Numerical
investigations are carried out in this section to show stress and displacement
distribution around interface of a ber reinforced composite element. For
brevity, only the results for a typical composite reinforced by an transversely
isotropic ber are presented here. The material constants used are for a
carbon ber reinforced epoxy matrix composite (T300/ST350) and are shown in
Table 1.
The geometric parameters of the RVE cylinder are shown in Fig.3. In the
numerical calculations, the ratio of r
m
/r
f
is taken to be 2.5, which corresponds to
a ber volume fraction of 0.16. The half length of the cylinder, l, is of 10r
f
for all
calculations except for the results shown in Fig. 11 in which the normal stress in
the ber at z = 0 against the length of the cylinder is plotted. In the calculations,
the rst n (n = 34) terms in the eigenfunction expansions have been used to
achieve an accuracy of about 1% for the maximum displacements. As a matter of
fact, only the rst nine terms are required to achieve an accuracy of about 3% for
displacements. To have the same accuracy for stresses, more terms have to be
included. It is also worthwhile to mention that due to the use of the truncated
series expansion, the singularities of the interfacial shear stress that would have
appeared at the interface near the ends of the cylinder have been smoothed out.
However, the results in Fig. 4 show that the peak shear stress occurs near the end
of the cylinder. It is observed that as more terms are added to the expansion, the
peak shear stress increases and its location moves further towards the end of the
cylinder. In Figs. (4)(11), the numerical results for Cases 1 and 2 are presented,
respectively, by solid and dash lines.
Fig. 4 shows the interface (r=r
f
) shear stress along the z-direction. Only the
variation of shear stress in the vicinity of the loaded end is given because all the
displacements and stresses including the shear stress change markedly in this
region. It can be seen from the gure that the location where the maximum shear
stress occurs is within a distance of about 0.2r
f
from the loaded end. Beyond this
Table 1
The material properties of T300/ST350
Materials E
r
(GPa) E
y
(GPa) E
z
(GPa) G
rz
(GPa) n
ry
n
rz
Fiber 20 20 210 23 0.25 0.025
Matrix 2.7 2.7 2.7 1.0 0.35 0.35
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1052
Fig. 4. The variation of shear stress t
rz
at interface (r=r
f
).
Fig. 5. The variation of shear stress t
rz
vs r at the cross section (lz )=0.2r
f
.
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1053
Fig. 6. The normal stress s
r
distribution at interface (r=r
f
).
Fig. 7. The variation of stress s
y
vs r at the cross section (lz )=0.2r
f
.
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1054
Fig. 8. The variation of the displacement w vs r at the cross section (lz )=0.2r
f
.
Fig. 9. The variation of the displacement u vs r at the end of the cylinder (z=l ).
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1055
Fig. 10. The normal stress s
z
distribution at the end of the cylinder (z=l ).
Fig. 11. The variation of the bre normal stress s
z
[
z=0
vs the length l of the cylinder at interface
(r=r
f
).
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1056
distance, the interfacial shear stress decays rapidly. At about a half of the ber
radius from the end, the shear stress is less than one-seventh of the one at 0.2r
f
.
The variation of the shear stress vs r at the cross section (lz = 0.2r
f
) is shown
in Fig. 5. It is noticed that the shear stress increases steeply when r approaches r
f
.
Within a quarter of the ber radius from the interface in r-direction, the shear
stresses in both the ber and the matrix change their signs (from negative to
positive). There is no substantial dierence in shear stress distribution between
Case 1 and Case 2.
Figs. 6 and 7 are, respectively, the normal stress distributions of s
r
at interface
and s
y
at the cross section where (lz )=0.2r
f
. The maximum s
r
occurs at the
loaded end. It is larger in Case 1 than in Case 2 (Fig. 6). From Fig. 7, it can be
seen that s
y
within the ber is always positive but changes to negative after
crossing the interface. The matrix is always in compression for Case 2, while for
Case 1, the matrix is in tension again at a distance between 0.5r
f
and r
f
from
the interface. The maximum values of both s
r
and s
y
in these cases are much
smaller compared with the shear stresses shown in Fig. 4 or Fig. 5. Obviously,
the shear stresses are the main concern in the research of interface stress
transfer.
Fig. 8 shows the variation of displacement w against r at the cross section (lz =
0.2r
f
). Within the ber, w almost remains unchanged with a value of 0.75p
0
l/c
33
for Case 1 and 0.71p
0
l/c
33
for Case 2. Only a slight decline is observed at the
interface. However, it drops rapidly in the matrix and approaches zero as r 4r
m
.
The curves shown in Fig. 9 are the variations of displacement u vs r at the
loaded end of the cylinder. Obviously, u of Case 1 is larger than that of Case 2.
The maximum values of u, which occur in the matrix within a distance of a
quarter of the ber radius away from the interface, are, respectively, 0.26p
0
l/c
33
for Case 1 and 0.20p
0
l/c
33
for Case 2. The displacement u at the outer surface
(r=r
m
) is about 0.23p
0
l/c
33
in Case 1.
The distribution of the normal stress s
z
at the end of the cylinder (z=l ) is
given in Fig. 10. It can be seen that for both cases the normal stresses are almost
constant in the central area (r/r
f
< 1) of the ber's cross-section and change
dramatically only when they approach the interface (r/r
f
=1). This observation
suggests that if a uniform stress is applied away from z=l (see Fig. 3), the
resulting p(r ) at z=l would be the same as shown. On the basis of St Venant
principle, the applied uniform stress may be replaced by any equivalent stress
system without aecting signicantly the distribution of p(r ) at z=l. This means
that the solutions presented in this paper can be used for relevant engineering
problems such as ber pull out, etc. as long as the external loads are applied
at the ber ends which are far enough away from the xed ends of the
cylinder.
It can also be seen from the gure that the magnitudes and the variations of s
z
in both cases are almost the same. Again the maximum s
z
occurs at the interface.
The average value of s
z
in the ber is p
0
.
To identify the eciency of the stress transfer, the normal stress s
z
of the ber
(r=r
f
) at the symmetric plane (z = 0) vs the length of the cylinder are plotted in
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1057
Fig. 11. Since smaller ber stress s
z
at the plane suggests that more loads (or
stresses) are transferred to the matrix, it is evident that cylinder length has a
considerable inuence on the eciency of the stress transfer. In the present study,
Case 2 is more ecient in stress transfer than Case 1.
6. Conclusions
The present investigation gives a more general solution to the stress transfer in
the analysis of micromechanics of ber reinforced composites with a particular
application to transversely isotropic nite combined-cylinder (RVE). The
governing equations for the cylinder were derived from 3D state space equations
of elasticity. The solutions were then sought systematically and can satisfy all the
boundary conditions prescribed at the surfaces of the cylinder. The present
solutions give a set of eigenvalue expansions and can be used to solve problems
involving either real or complex eigenvalues. For the non-self equilibrium
boundary value problems, the classical Lame solution used by other investigators
is only a special case of the zero eigen solution in the present expansions. It is
expected that more general boundary value problems can be solved by using this
solution.
Numerical results demonstrate that all stress and displacement components vary
markedly near the ends of the cylinder. The interfacial shear stress, which is the
most important component in the stress transfer, reaches its maximum within a
distance of a quarter of the ber radius from the end. It has also been shown how
the length of the cylinders aects the eciency of the stress transfer.
The solution provides an accurate tool to deal with axisymmetric stress
problems in the micromechanics of ber reinforced composites, by which a
continuous displacement and transverse stress eld across an interface can always
be obtained. As a result, the stress distribution obtained is more reliable,
particularly, at the ends of a combined composite cylinder where stress
concentration occurs. Using simple shape functions in 3D FE analysis, however,
would be unlikely to fully represent the displacements and stresses that would
occur particularly at the edges of the cylinder. Without the ability to handle
stress singularities or discontinuities at the ends, as it is in most traditional FE
analysis, signicant penalties could be faced in transitioning to a ne mesh at the
edges.
The solutions and the results presented may also be used to assess various
numerical and simplied theoretical models used at present in the research of stress
transfer.
Acknowledgements
The authors would like to acknowledge the support of EPSRC (Grant No. GR/
L91450) which has allowed this project to be carried out. The rst author is also
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1058
grateful for the support from Cao-Guang-Biao Advanced Science and Technology
Foundation of Tsinghua University.
Appendix A
State equation of transversely isotropic axisymmetric problem in cylindrical polar
coordinates
In a cylindrical polar coordinate system, the basic Eqs. (4), (5) and (6) of
elasticity are, respectively, reduced to
@s
r
@r

s
r
s
y
r

@t
rz
@z
= 0
@t
rz
@r

t
rz
r

@s
z
@z
= 0, (A1)
_

_
s
r
s
y
s
z
t
rz
_

_
=
_
_
_
_
c
11
c
12
c
13
0
c
12
c
11
c
13
0
c
13
c
13
c
33
0
0 0 0 c
44
_

_
_

_
e
r
e
y
e
z
g
rz
_

_
, (A2)
and
e
r
=
@u
@r
, e
y
=
u
r
, e
z
=
@w
@z
, g
rz
=
@u
@z

@w
@r
: (A3)
From the rst equation of (A2) and the strain-displacement relations (A3), one
has
s
r
= c
11
@u
@r
c
12
u
r
c
13
@w
@z
or
@u
@r
=
c
12
c
11
u
r

c
13
c
11
@w
@z

1
c
11
s
r
: (A4)
Similarly, the last equation in (A2) leads
@w
@r
=
@u
@z

1
c
44
t
rz
: (A5)
Substituting the expressions of s
r
and s
y
in (A2) into the rst equilibrium
equation of (A1), the following equation can be found
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1059
@s
r
@r
=
1
r
[(c
12
c
11
)e
r
(c
11
c
12
)e
y
]
@t
rz
@z
,
e
r
and e
y
are given in (A3). After eliminating e
r
and e
y
from (A3) and @u/@r from
Eq. (A4), the above equilibrium equation becomes
@s
r
@r
=
_
c
11

c
2
12
c
11
_
u
r
2

_
c
13

c
12
c
13
c
11
_
1
r
@w
@z

_
c
12
c
11
1
_
s
r
r

@t
rz
@z
: (A6)
From the second equation of (A1) and following the same procedure used to
derive (A6), one has
@t
rz
@r
=
_
c
13

c
12
c
13
c
11
_
1
r
@u
@z

_
c
33

c
2
13
c
11
_
@
2
w
@z
2

c
13
c
11
@s
r
@z

t
rz
r
: (A7)
Denoting dierential operators @/@z by a and @
2
/@z
2
by a
2
and letting
c
1
= c
11
c
2
12
=c
11
, c
2
=c
13
c
13
c
12
/c
11
, c
3
= c
33
c
2
13
=c
11
, c
4
=c
12
/c
11
1, c
5
=c
12
/
c
11
, c
6
=c
13
/c
11
, c
7
=1/c
11
and G
rz
=c
44
, the nal state Eq. (7) can be formed from
(A4)(A7).
Appendix B
Elements of matrices A
k
and H
k
in Eqs. (49) and (56)
Let
k
f1
=

a
1k
_
r
f
, k
f2
=

a
2k
_
r
f
, k
f
=

a
k
_
r
f
, k
m
=

a
k
_
r
m
and
A
11
=
G
rz
c
11
x
1
(c
13
G
rz
)

x
1
_ , A
12
=
G
rz
c
11
x
2
(c
13
G
rz
)

x
2
_ , A
31
= G
rz
c
13
c
11
x
1
c
13
G
rz
,
A
32
= G
rz
c
13
c
11
x
2
c
13
G
rz
, A
41
=
_
c
11
c
13
G
rz
c
11
x
1
(c
13
G
rz
)x
1
_

x
1
_
,
A
42
=
_
c
11
c
13
G
rz
c
11
x
2
(c
13
GG
rz
)x
2
_

x
2
_
, A
40
= c
12
c
11
,
A
71
=
_
c
13
c
33
G
rz
c
11
x
1
(c
13
G
rz
)x
1
_

x
1
_
, A
72
=
_
c
13
c
33
G
rz
c
11
x
2
(c
13
G
rz
)x
2
_

x
2
_
,
then
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1060
A
k
=
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
A
11
J
0
(k
f1
) A
12
J
0
(k
f2
) J
0
(k
f
) Y
0
(k
f
) 4(1 n)J
0
(k
f
) k
f
J
1
(k
f
) 4(1 n)Y
0
(k
f
) k
f
Y
1
(k
f
)
J
1
(k
f1
) J
1
(k
f2
) J
1
(k
f
) Y
1
(k
f
) k
f
J
0
(k
f
) k
f
Y
0
(k
f
)
A
31
2G
J
1
(k
f1
)
A
32
2G
J
1
(k
f2
) J
1
(k
f
) Y
1
(k
f
) k
f
J
0
(k
f
) 2(1 n)J
1
(k
f
) k
f
Y
0
(k
f
) 2(1 n)Y
1
(k
f
)
A
41
2G
J
0
(k
f1
)
A
40
2G
J
1
(k
f1
)
k
f
A
42
2G
J
0
(k
f2
)
A
40
2G
J
1
(k
f2
)
k
f
J
1
(k
f
)
k
f
J
0
(k
f
)
Y
1
(k
f
)
k
f
Y
0
(k
f
) k
f
J
1
(k
f
) (1 2n)J
0
(k
f
) k
f
Y
1
(k
f
) (1 2n)Y
0
(k
f
)
0 0 J
1
(k
m
) Y
1
(k
m
) k
m
J
0
(k
m
) 2(1 n)J
1
(k
m
) k
m
Y
0
(k
m
) 2(1 n)Y
1
(k
m
)
0 0 J
0
(k
m
)
J
1
(k
m
)
k
m
Y
0
(k
m
)
Y
1
(k
m
)
k
m
(1 2n)J
0
(k
m
) k
m
J
1
(k
m
) (1 2n)Y
0
(k
m
) k
m
Y
1
(k
m
)
_

_
,
H
k
=
_
2A
71
a
k
J
1
(k
f1
)
k
f1
(1)
k
2A
72
a
k
J
1
(k
f2
)
k
f2
(1)
k
0 0 0 0
_
:
In Case 2, the last row of matrix A
k
is replaced by [0 0 J
1
(k
m
) Y
1
(k
m
) k
m
J
0
(k
m
) k
m
Y
0
(k
m
)].
Z
.
J
.
W
u
e
t
a
l
.
/
J
.
M
e
c
h
.
P
h
y
s
.
S
o
l
i
d
s
4
8
(
2
0
0
0
)
1
0
3
7

1
0
6
3
1
0
6
1
References
Atkinson, C., Avila, J., Betz, E., Smelser, R.E., 1982. The rod pull out problem, theory and
experiment. J. Mech. Phys. Solids 30, 97120.
Aveston, J., Kelly, A., 1973. Theory of multiple fracture of brous composites. J. Mater. Sci. 8, 352
362.
Case, S.C., Reifsnider, K.L., 1996. Micromechanical analysis of ber fracture in unidirectional
composite materials. Int. J. Solids Structures 33 (26), 37953812.
Chree, C., 1889. The equations of an isotropic elastic solid in polar and cylindrical coordinates, their
solution and application. Camb. Phil. Soc. Trans. 14, 250369.
Cook, R.F., Thouless, M.D., Clarke, D.R., Kroll, M.C., 1989. Stick-slip during ber pull-out. Scripta
Metall. 23, 17251730.
Cox, H.L., 1952. The elasticity and strength of paper and other brous materials. Brit. J. Appl. Phys.
3, 7279.
Curtin, W.A., 1991. Exact theory of ber fragmentation in a single lament composite. J. Mater. Sci.
26, 52395253.
Dollar, A., Steif, P.S., 1988. Load transfer in composites with a Coulomb friction interface. Int. J.
Solids Structures 24 (8), 789803.
Fairman, F.W., 1998. Linear Control Theory, The State Space Approach. John Wiley, England.
Filon, L.N.G., 1902. On the elastic equilibrium of circular cylinder under certain practical systems of
load. Proc. R. Soc. Lond. A198, 147233.
Flu gge, W., Kelkar, V.S., 1968. The problem of an elastic circular cylinder. Int. J. Solids Structures 4,
397420.
Freund, L.B., 1992. The axial force needed to overcome frictional resistance on a circular ber as it
slides through a hole in an elastic material. Eur. J. Mech, A/Solids 11 (1), 119.
Gao, Y.C., Mai, Y.W., Cotterell, B., 1988. Fracture of ber-reinforced materials. J. Appl. Math. Phys.
(ZAMP) 39, 550572.
Hsueh, C.H., 1992. Interfacial debonding and ber pull-out stresses of ber-reinforced composites. VII:
improved analyses for bonded interfaces. Mater. Sci. Eng. A154, 125132.
Hutchinson, J.W., Jensen, H.M., 1990. Models of ber debonding and pullout in brittle composites
with friction. Mech. of Mater. 9 (2), 139163.
Kelly, A., 1970. Interface eects and the work of fracture of a brous composite. Proc. R. Soc. Lond.
A319, 95116.
Keran, R.J., Parthasarathy, T.A., 1991. Theoretical analysis of the ber pullout and pushout tests. J.
Am. Ceram. Soc. 74, 15851596.
Kim, J.K., Mai, Y.W., 1998. Engineered Interfaces in Fiber Reinforced Composites. Elsevier,
Oxford.
Knowles, J.K., Horgan, C.O., 1969. On the exponential decay of stresses in circular elastic cylinders
subject to axisymmetric self-equilibrated end loads. Int. J. Solids Structures 5, 3350.
Lur'e, A.I., 1964. Three-Dimensional Problems of the Theory of Elasticity. Interscience Publications
Moscow. 1955. English transl. by Mcvean, D. B.
Marotzke, C., 1994. The elastic stress eld arising in the single ber pull-out test. Composites Sci.
Technol. 50, 393405.
McCartney, L.N., 1989. New theoretical model of stress transfer between bre and matrix in a
uniaxially bre-reinforced composite. Proc. R. Soc. Lond. A425, 215244.
Muki, R., Sternberg, E., 1969. On the diusion of an axial load from an innite cylindrical bar
embedded in an elastic medium. Int. J. Solids Structures 5, 587605.
Piggott, M.R., 1980. Load Bearing Fiber Composites. Pergamon Press, Oxford.
Povirk, G.L., Needleman, A., 1993. Finite element simulations of ber pull-out. J. Engng. Mater.
Technol. 115, 286291.
Rosen, B.W., 1964. Tensile failure of brous composites. AIAA J. 2, 19851991.
Smith, G.E., Spencer, A.J.M., 1970. Interfacial tractions in a bre-reinforced elastic composite material.
J. Mech. Phys. Solids 18, 81100.
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1062
Timoshenko, S., 1957. Strength of Materials, Part II Advanced Theory and Problems. Van Nostrand,
New Jersey.
Tsai, K.H., Kim, K.S., 1996. The micromechanics of ber pull-out. J. Mech. Phys. Solids 44 (7), 1147
1177.
Wu, Z.J., 1991. An investigation on the interface of ber reinforced composite theory and
experiment. Ph.D. Dissertation of Shanghai Jiao Tong University, Shanghai.
Z.J. Wu et al. / J. Mech. Phys. Solids 48 (2000) 10371063 1063

Vous aimerez peut-être aussi