Você está na página 1de 14

Mechanics of Composite Materials, Vol. 48, No. 4, September, 2012 (Russian Original Vol. 48, No.

4, July-August, 2012)

FAILURE AND DEFORMATION ANALYSES


OFSMARTLAMINATED COMPOSITES

Z. Hasan1 and A. Muliana2*

Keywords: smart laminated composites, failure analysis, deformation analysis, piezoelectric actuators,
active fiber composite, microfiber composite, lead zirconate titanate
The present study focuses on the failure analysis and shape control of smart composite laminates under coupled
thermal (hygro), electric, and mechanical stimuli. A linear thermo(hygro)electroelastic constitutive model
for transversely isotropic materials is used for each ply in the composite laminate and for the piezoelectric
materials that are integrated with laminates of the composite. Piezoelectric materials, such as lead zirconate
titanate, and piezoelectric fiber composites, such as an active fiber composite or a microfiber composite, are
considered as actuators for controlling unwanted bending deformations to avoid failure in such composite
laminates. Due to the high stress concentrations at the interfaces between an active layer and the host
structure, which may cause debonding, embedded actuators in which the active material is placed as part of
the plies to form geometrically continuous plies are considered in order to minimize the stress concentration
while improving the actuation capability. The first-ply failure and the ultimate laminate failure criteria of
composite laminates are used to predict the failure stress and mode of the smart composite laminates, where
commonly known macroscopic failure criteria, such as the TsaiHill, TsaiWu, and maximum stress criteria,
are employed for each lamina. Piezoelectric materials can be used to prevent the failure from hygrothermal
and mechanical loadings by applying an electric voltage in order to counteract laminate deformations. Based
on the deformation and failure analyzes of smart composite laminates having various stacking sequences,
fiber and matrix constituents, and piezoelectric materials, we could estimate the overall properties and
failure envelopes of the laminates, which is useful in the preliminary design of smart composite structures.

The Boeing Company, Mesa, Arizona, 85215


Department of Mechanical Engineering, Texas A&M University
*
Corresponding author; e-mail: amuliana@tamu.edu
1
2

Russian translation published in Mekhanika Kompozitnykh Materialov, Vol. 48, No. 4, pp. 571-590 , July-August,
2012. Original article submitted July 27, 2011; revision submitted April 3, 2012.
0191-5665/12/4804-0391 2012 Springer Science+Business Media, Inc.

391

Introduction
Composites are multiphase materials obtained through the combination of different engineered materials in order to
obtain properties that the individual components by themselves cannot reach. In new generations, the driving force for technological changes has led to a new family of engineered materials and structures exhibiting multifunctional capabilities which
are naturally seen in biological systems, leading to a new era of smart materials [1]. The structures with surface-mounted or
embedded sensors and actuators that are able to sense and adapt to external stimuli are referred to as smart structures [2]. One
of the main motivations behind the vast attention given to smart materials and is that they can be used to monitor the integrity
of a structure, to enable a structure to change its shape, or to control its vibration [3].
Many researchers can be recognized for their contributions to the shape control of structures by using piezoelectric
materials. Lee and Moon [4-6] used piezoelectric layers in order to excite a specific structural mode for controlling vibrations
caused by external disturbances. Koconis et al. [7] analyzed changes in the shapes of fiber-reinforced composite beams, plates,
and shells by using embedded piezoelectric actuators. Tzou et al. [8] studied the distributed structural control of continua of
elastic shells by using spatially distributed modal piezoelectric actuators and some generic distributed feedback algorithms
with spatial feedback functions. Agrawal et al. [9] used the finite-difference method to analyze and control the deflection of a
smart plate with piezoelectric actuators embedded in elastic plates by estimating the optimal actuation voltages. Several other
solutions related to smart structures can be found in papers [10, 11], which are focused on obtaining elasticity solutions for
smart composite laminates and multilayered piezoelectric plates and shells under mechanical loadings. Investigations into
the hygrothermal effect in smart composites are currently limited. Raja et al. [12] studied the actuation and sensing behavior
of flat and curved piezohygrothermoelastic plates with active control through a finite-element procedure involving a coupled
piezoelectric field with hygrothermal strains. Mahato and Maiti [13] examined the aeroelastic performance of smart composite
plates under aerodynamic loads, including a hygrothermal environment. The presence of hygrothermal effects weakens the
composite structure, but the structure can regain its stiffness in a hygrothermal environment when the active layer is activated
with the help of applied voltage, which can be used to enhance the flutter boundary. Controlling the thermal deformation in
composite structures by using composite piezoelectric actuators has also been studied by Dano and Julliere [14].
The failure behavior of traditional laminated composite materials has also been studied extensively; a description of
the failure criteria developed in the past decades can be found in [15, 16]. Reddy and Pandey [17] developed a finite-element
computational procedure based on the first-ply failure analysis of laminated composite plates subjected to in-plane and/or
bending loads by considering the first-order shear deformation theory and a tensor-polynomial failure criterion. Mayes and
Hansen [18] used a constituent stress-based failure criterion to construct a nonlinear progressive failure algorithm for investigating the failure of composite laminates. Takeda et al. [19] focused on understanding the deformation and progressive failure
behavior of glass/epoxy plain-weave-fabric-reinforced laminates subjected to uniaxial tension at cryogenic temperatures.
Pipes et al. [20] studied the hygrothermal response of laminated composite systems. A plate element was used to analyze the
lamina stresses resulting from hygrothermal and mechanical loadings. While the failure of traditional laminated composites
subjected to thermomechanical loadings have been studied extensively, investigations into the deformation and failure of smart
laminated composites with a thermo (hygro)electromechanical effect are currently lacking.
The present study focuses on analyzing the failure and deformation of smart laminated composites containing different types of piezoelectric materials, such as lead zirconate titanate (PZT), an active fiber composite (AFC), or a microfiber
composite (MFC). The purpose is to use these piezoelectric materials for controlling deformations in composite laminates
subject to coupled hygrothermal and mechanical loadings in order to delay their failure. In most practical applications, the
active materials are patched to host structures, and their debonding can be a problem, especially when brittle PZTs are employed. We will examine the possibity of using flexible AFCs and MFCs and integrating the active materials with the host
structures in order to overcome the discontinuities of stresses and of applying an electric field through the active layers in order
to minimize unwanted deformations.

392

1. Constitutive Relations for a Lamina and Smart Laminated Composites


Let us consider a constitutive model for a lamina with hygrothermal and electroelastic coupling effects. We assume
for the lamina a linear response that satisfies the superposition and proportionality conditions between all field variables. The
general constitutive relation for it can be derived from the Gibbs potential with temperature, moisture, stress, and electric field
as independent variables1. In the case of nonlinear response, these relations can be extended by including higher-order terms
in the Gibbs potential [21]. The linear constitutive model is given in the form

T , ,c
ij = Sijkl
kl + d kTij, ,c k + ij ,c, T + ij ,T , C ,

T , ,c
Di = dijk
jk + ijT , ,c j + pi ,c, T + qi ,T , C , (1)

where eij and sij are the components of strain and stress tensors, respectively; T and C are the temperature and moisture content
of the material, respectively; Di is the component of the dielectric displacement vector; i is the component of the electric field
vector. The material properties and b are the thermal and moisture expansion tensors, respectively; d is the piezoelectric
strain coefficient tensor; S is the compliance tensor; p is the pyroelectric vector; is the permittivity tensor; q is the measure
of changes in the electric displacement per unit moisture content. The superscripts point to the constant field variables at which
the material properties are characterized. Up to now, there are no data for moisture-caused changes in the electric displacement,
therefore, we take q to be zero. The above-mentioned constitutive model is used for active materials, i.e., AFCs, MFCs, and
piezoelectric wafers, integrated with laminated composites. For a lamina without the electromechanical coupling effect, such
as carbon/epoxy and glass/epoxy laminae, the piezoelectric and pyroelectric quantities in Eqs. (1) are zero.
For convenience, in formulating the effective response of smart laminated composites, the field variables may be
written in the engineering notation, where the pairs of indices ii = 11, 22, and 33 are replaced with single indices m = 1, 2,
and 3, respectively, and the mixed pairs of indices (which represent the shear components of strain and stress tensors) ij = 23
or 32, 13 or 31, and 12 or 21 are written as m = 4, 5, and 6, respectively. Then, the linear piezoelectric constitutive equations
are expressed as

m = Smn n + dimi + m T + m C ,

Di = dim m + ij j + pi T ; i, j = 1, 2, 3; m, n = 1, 2,..., 6.

In many structural applications, composite materials consist of thin laminates loaded predominantly in their plane,
with all stress components in the out-of-plane direction being zero. It is assumed that each laminate is transversely isotropic
and can be described by the linear hygrothermoelectroelastic constitutive model, with index 1 denoting the longitudinal fiber
direction. This results in the following hygrothermoelectroelastic constitutive equation2 for a kth layer, polarized across its
thickness (in the out-of-plane direction), in the local coordinate system:
1
Q11

2 = Q12
6
0
k

Q12
Q22
0

0
0
Q66 k

0
D1
0

0
D2 = 0
D3
e31 e32
k

1 1
0 0 e31 1
1




2 2 T 2 C 0 0 e32 2 ,
0
0 0 0
0
k 3 k
k
6 k

0
0
0 k

0
1 11 0

0
2 + 0 22
6 0
0 33 k

1 p1

2 + p2 T ,
3 p3
k
k

It is assumed that the moisture concentration has a similar effect as the expansion/contraction due to temperature
changes. In practice, moisture concentration can cause swelling in the constituents, affecting the thermoelectroelastic properties
of the composite, and its effect is often unrecoverable.
2
This equation is for an active material with the poling axis in the x3-direction. When AFC or MFC that is polarized
in the in-plane direction is considered, the above equation will have different components of the piezoelectric constants eim .
1

393

where Qim are components of the in-plane stiffness, eim are components of the piezoelectric constants ( e = dC and C = S 1 ),
and ij are the dielectric constants, which are measured at a constant or zero strain of a kth lamina in its material coordinate
system. The coefficients Q( k )im can be expressed in terms of elastic constants of the kth layer as follows:

Q11 =

E1
,
1 21 12

Q12 =

12 E2
,
1 21 12

Q22 =

E2
,
1 2112

Q66 = G12 ,

where E1, E2, G12, and n12 are the in-plane elastic moduli, shear modulus, and Poissons ratio, respectively. Generally, each
lamina in a smart laminated composites is oriented at a certain angle to the global coordinate system. Therefore, a tensor
transformation is necessary for each lamina in order to express the constitutive relation in the global coordinate system. We
will use an overbar to denote the field variables and material characteristics in the global coordinate system. Detailed tensor
transformations can be found in [22] and [23].
As we are interested in analyzing the response of smart laminated composites comprising several layers, it is necessary to obtain constitutive relations for them, too. We deal with composites able to change their shape owing to the presence
of piezoelectric materials as actuators; thus, we are interested in thin and compliant laminated composites. We use the classical lamination theory (CLT) for the thin-plate (KirchhoffLove plate) theory to obtain an effective constitutive relation for the
smart laminated composites. Let u0, n0, and w0 be midplane displacements in the x1-, x2-, and x3-directions, respectively, where
x1 and x2 are the in-plane axes, and x3 is the out-of-plane axis of the plate. It is assumed that the layers are perfectly bonded
together. Further, we will restrict the formulation to a linear response and small strains and displacements. Consider a laminate
n

made from n plies of thickness tk, so that the total thickness of the laminate is h = tk . The position of the bottom surface
k =1

of a kth ply along the thickness axis x3 is hk. According to the CLT, the effective in-plane axial and shear forces, bending moments, and the in-plane torque caused by displacement gradients and electric fields are

N1 A11
N = A
2 21
N 6 A16

M 1 B11
M = B
2 21
M 6 B16

A12
A22
A26

B12
B22
B26

u0

xx
B
A16 1
v0
11

A26
B21
x
2

B
A66
u v 16
0 + 0
x2 x1
u0

xx
D
B16 1
v0
11

B26
D21
x
2

D
B66
u v 16
0 + 0
x2 x1

B12
B22
B26

D12
D22
D26

2 w0

2
AHT AP
x1
B16 2
1 1

w
HT p

0
A
A ,
B26
x 2 2 2

p
2
HT
B66
2 A6 A6
2 w0
x1x2
2 w0

2
B HT B p
x1
D16 2
1 1

w
HT p

0
B
B ,
D26
x 2 2 2

p
2
HT
D66
2 B6 B6
2 w0
x1x2

where N1 and N2 are the normal forces per unit length, N6 is the in-plane shear force per unit length, M1 and M2 are the bending moments per unit length, and M6 is the twisting moments per unit length. A, B, and D are the extensional, axial-bending
coupling, and bending stiffness matrices, respectively. They are given as

Aij = Qij (hk hk 1 ), i = 1, 2, 6; j = 1, 2, 6,



k =1


394

Bij =

1 2
Qij (h k h2k 1 ), i = 1, 2, 6; j = 1, 2, 6,
2 k =1 k

Dij =

1 n 3
Qij (h k h3k 1 ), i = 1, 2, 6; j = 1, 2, 6.
3 k =1 k

Assuming that the electric fields vary linearly within each layer, the hygrothermal and electro-mechanical constants
are defined as [24, 25]
1 n

AiHT = Qijk kj (T1k + T2k )hk + Qijk kj (C1k + C2k )hk ,
2 k =1 j =1, 2,6

BiHT =

1 n
Qijk kj [T1k (hk + 3zk ) + T2k (2hk + 3zk )]hk + Qijk kj [C1k (hk + 3zk ) + C2k (2hk + 3zk )]hk ,
6 k =1 j =1, 2,6
Aip =

Bip =

1 Na
Qijk d3k j (23k )hk ,
2 k =1 j =1, 2,6

1 Na
Qijk d3k j [3k (hk + 3zk ) + 3k (2hk + 3zk )]hk ,
6 k =1 j =1, 2,6

where na is the number of actuating layers. The electric field 3 in the through-thickness direction of the smart layer is defined as

3k =

Vk
,
tk

where Vk is the electric potential applied across a kth layer, and tk is the thickness of the layer. The superscript p means
piezoelectric. These equations determine the stiffness of the piezoelectric layer present in the lay-up. Here, we should note
that the application direction of the electric field varies with polarization direction of the piezoelectric material. The
KirchhoffLove plate theory is applicable only to moderately thin composite plates; for a very thin plate, e.g., with length/
thickness > 60, the von Krmn plate theory should be used.

2. Failure Analysis of Smart Laminated Composites


The following section presents a failure analysis of composite laminates subject to hygrothermal and electric stimuli.
The first-ply failure (FPF) and ultimate laminate failure (ULF) criteria of composite laminates are used in order to predict the
failure stress and mode of a composite laminate under uniaxial and biaxial loadings. We focus on analyzing the response of
thin laminated composites, for which the effect of transverse shear deformations on the overall performance of composite is
less significant. Various commonly known macroscopic failure criteria, such as the TsaiHill, TsaiWu, and maximum stress
criteria, are considered for each lamina.
The failure of a laminate may be defined either as the initial failure or the ultimate failure, depending on the degree
of conservatism applied. In the FPF, a laminate is considered failed when a first layer (or group of layers) has failed. This is
established by conducting a stress analysis of the laminate for the given loading condition, by determining the stress state in
each individual layer and by assessing the strength of each layer according to the failure criterion selected. This approach is
conservative, but it can be used with low safety factors. In the ULF, a laminate is considered failed when its maximum load
level is reached or all plies (laminas) have failed. The determination of the ULF requires an iterative procedure taking into
account the development of damage in various plies. In the analysis, initially the first-ply failure is determined and the damaged ply is discounted; then the stresses are recalculated for the undamaged plies and checked against the failure criterion to
verify that the undamaged laminae will not fail immediately under the increased stresses according to the FPF criterion used.
In this analysis, the strengths of the previously failed lamina (with a reduced or totally discounted stiffness) are assumed to

395

TABLE 1. Strength of Some Actuators


aterial

F1t

AFC
MFC
PZT-5A
AS4-3601

23.9
30
40**
2275

F2t

F1c

12
14.9
20
57.2

26.9
26.9
500***
1723
*

F2c

F6

13.4
13.4
249.5
227.5

16.3
18
22
75.8

[26], **[27], and ***[28]; the rest values are assumed

TABLE 2. Material Properties of the Actuators and Composite Beam Used


Property
Youngs modulus, GPa
E1
E2
E3
Poissons raito
12
13
23
Shear modulus, GPa
G12
G13
G23
Piezoelectric coefficients, pm/V
d31
d32
d33
d15
d24
Thickness t, m

Carbon
epoxy(AS4-3601)

MFC (60%)

AFC (60%)

PZT-5H

PZT-5A

147
10.3
10.3

30.0
15.5
15.5

35
10.41
10.41

61
61
48

61
61
53.2

0.27
0.27
0.54

0.35
0.4
0.4

0.35
0.38
0.38

0.31
0.31
0.31

0.384
0.4
0.4

7
7
3.7

5.7
10.7
10.7

4.4
4.96
4.96

23.3
19.1
19.1

22.5
21
21

198
198
418
-

260
260
540
-

274
274
593
741
741

171
171
374
584
584

0.0003

0.0003

0.0003

0.0003

0.000127/ply

be fictitiously very high to avoid the indication of a repeated failure in the same plies. The procedure is continued until the
ULF criterion is met.
This study focuses on the failure analyses of smart composite laminates with carbon fiber/polymer laminae (AS43601) with various stacking sequences and active materials. The strengths of the active materials and the AS4-3601 fiber
composite used in the analysis are shown in Table 1; the assumed values are based on comparing the strength ratios with those
of fiber-reinforced composite materials. The properties of the carbon laminae and piezoelectric materials used in this study
are given in Table 2. The effect of applying a 1-MV/m electric field to the active part of an [actuator/90/60/60]s composite
lay-up at T = 38C in a uniaxial loading is shown in Table 3. The macroscopic failure analysis of each lamina reported in
this manuscript was based on the maximum stress criterion3. Other macroscopic failure criteria for each lamina, such as the
TsaiHill and TsaiWu criteria, have also been considered [29]. It was observed that the applied electric field slightly reduced
According to the maximum stress theory, failure occurs when at least one stress component along one of the principal
material axes exceeds the corresponding strength in that direction.
3

396

TABLE 3. FPF Stress under a Uniaxial Loading, MPa


Actuator

=0

= 1 V/m

= 1 V/m, T = 38C

mode AFC/31
mode AFC/33
PZT

13.2
13.2
16.2

12.5
14.6
14.6

3.2
5.2
7.5

TABLE 4. ULF Stress; n = 1, [AFC/0/90/45/45]s


Loading mode

Ply, deg

Stress in plies, MPa

Failure mode

n = 1 (tension/tension)

AFC
0
90
454
454

44.6
268
286.1
276.4
276.4

Longitudinal tensile stress


Transverse tensile stress
he same
he same
he same

n = 1 (tension/compression)

AFC
0
90
45
45

30.7
1204.5
350.2
337.8
337.8

Longitudinal tensile stress


Transverse tensile stress
he same
Shear stress
he same

TABLE 5. FPF Stress under a Biaxial Loading, MPa


Actuator

=0

= 1 V/m

= 1 V/m, T = 38C

AFC/mode 31
AFC/mode 33
PZT

13.2
13.2
16.2

12.5
14.6
16

3.8
5.2
6.2

the failure stress when considering the AFC (mode 31), but it increased the failure stress in the case of mode 33. This is due
to the positive piezoelectric strain coefficient in the 33-direction, whereas in mode 31 it was negative. Mode 33 had a more
significant effect than mode 31 due to the greater piezoelectric coefficient in the case of mode 33 (see Table 2). The temperature
variation also decreased the failure load, but to a lesser degree than the variation in the electric field.
The composite laminates were also studied under a biaxial loading. For this purpose, we introduced the factor
n = 2 1 the ratio of the stress applied in the transverse direction to that in the axial one. Following the same procedure
performed previously in the cases of FPF and ULF, with the maximum stress criterion for each lamina, the failure of
[AFC/0/90/45/45]s laminates under biaxial loadings were investigated. The results obtained are given in Tables 4 and 5.
Failure also started in the active layer due to its low strength both in the longitudinal and transverse direction compared with
that of the other carbon/epoxy layers. In addition, the stacking sequence also had a significant effect, which was caused by the
stress distribution between different plies. The effect of applying a 1-MV/m electric field to the active part of the [actuator/90/60/60]s composite and the temperature difference T = 38C is shown in Table 5. This study demonstrates that the
incorporation of piezoelectric materials changes the overall failure load of smart laminated composites insignificantly, making
it possible to develop various smart laminated materials.
In order to investigate the failure in smart laminated composites under more complex loading conditions, the commercial finite-elements software ABAQUS was used. The effect of thermal stress on the failure of a smart composite plate was
also considered in this study. The plate was cantilevered and subjected to a temperature difference of 100C. A finite-element
397

1
85 mm

20 mm

2
0.8 mm

250 mm

Fig. 1. Geometry of a composite plate.


TABLE 6. Predicted Finite-Element Stresses of the Composite Plate under a 100C Temperature Difference
Ply, deg

F1t/F1c

F2t/F2c

F6

0
45
45
90
90
-45
45
0
FC

44.1
1.82
29.7
9.3
14.8
23.7
19.1
27.1
47.09

58.5
27.1
53.2
30.9
49.4
53.6
27.2
27.4
24.6

0.15
0.48
5.02
0.97
.917
4.66
1.07
0.32
0.54

2275/227
2275/227
2275/227
2275/227
2275/227
2275/227
2275/227
2275/227
24.6/28.9

57/75
57/75
57/75
57/75
57/75
57/75
57/75
57/75
13/15.8

1723
1723
1723
1723
1723
1723
1723
1723
17.23

model was used for a [0/45/45/90]s laminate with two MFC actuators embedded in the composite structure (Fig. 1). The
characteristics of the materials used are given in Table 2. Table 6 shows the calculated in-plane stresses in composite layers
and in the actuator; by comparing them with the corresponding strength properties of each material, it can be found that the
failure first occurs in the MFC actuators because of their low strength.
4. Shape Control of Composite Laminates by using Piezoelectric Materials
The effect of different types of actuators on the bending deformation of a composite laminate was also examined. The
actuators considered and their properties are indicated in Table 2. The composite laminate was made of carbon/epoxy AS4
(3501-6) with the stacking sequence [actuator/90/45/45/0]s. Active-fiber composites can be classified into two types according
to the method used in embedding the active fibers into a passive matrix. Figure 2a shows a schematic of the first type, where
PZT rods are incorporated in the matrix across the actuator thickness. The second type is shown in Fig. 2b, where the fibers
are embedded along the longitudinal direction of the actuator, and the electrodes are attached to the upper and lower surfaces
of the actuator. Both the types were investigated in the study.
An analytical elastic solution for a smart elastic cantilever beam subjected to a point load was obtained in order to
verify the results found by using the commercial finite-element software ABAQUS. The aim of presenting this solution was
to gain strong confidence in the element type and mesh size used when dealing with more complex structures. A discussion
about the analytical solutions is presented in [29]. The beam was modeled using 20-node quadratic continuum elements, with
reduced integration (C3D20R) for the elastic part. An extra degree of freedom for the electrical potential (C3D20RE) were used
to model the part exhibiting piezoelectric properties. An elastic aluminum beam of dimensions 600 60 1.5 mm, termed a
single bonded actuator, was attached to a PZT layer along its top surface. We also considered two PZT layers, termed a double
bonded actuator, fastened to the top and bottom surfaces of the elastic beam. The thickness of each PZT layer was 0.3mm. The
FE of the elastic beam model was meshed with four elements across the thickness, eight elements along the width, and 100
398

a
2

Fig. 2. Piezoelectric-fiber composite with fibers embedded across the matrix thickness (type 1) (a)
and along the matrix length (type 2) (b).

a
x1, m

w, mm
0
1
2
3
4
5
6
7

0.2

0.4

x1, m

w, mm
0
1
2
3
4
5
6
7

0.6

1
1
2
2

0.2

0.4

0.6

1
1
2
2

Fig. 3. Deflection w of a smart beam, with single (a) and double (b) bonded actuators, subjected to
40- (1, 1 ) and 80-V (2, 2 ) electric field across the thickness of PZT layers: analytic results (1, 2)
and FE calculation (2, 2 ).

30 mm

60 mm
25 mm

30 mm

300 mm

Fig. 4. Schematic of a composite cantilever beam with attached actuators.

elements along the length, but the actuator layers had two elements in the thickness direction. Figure 3 shows deflections of the
cantilever beam along its length under the action of a through-the-thickness voltage differences of 40 and 80V, corresponding
to 20.5- and 41-V/mm electric fields, respectively. As seen, the results obtained from the analytical solution and the FE model
are very close to each other, with a max error of 2% for the single bonded actuator and of 8% for the double bonded actuator.
It can be concluded that, at the same voltage, a double bonded actuator causes greater deflections than a single bonded one.
Next, the FE method was used in order to predict the response of a cantilever beam with bonded actuators subjected
to a voltage difference across the actuators. A schematic of the composite beam used is depicted in Fig. 4. Figure 5a shows
the response of the composite cantilever beam under the action of a 1.25-MV/m electric field in the cases of the first-type
399

a
x1, m

w, mm
0
1
2
3
4
5
6
7

0.1

0.2

0.3

1
2
3
4

x1, m

w, mm
0
1
2
3
4
5
6
7

0.1

0.2

0.3

1
2

Fig. 5. Predicted steady-state deflections of composite cantilever beams with type-1 (a) and type-2(b)
actuators under a 1.25-MV/m electric field: AFC (1), MFC (2), PZT-5H (3), and PZT (4).

a
x1, m

w, mm
0
1
2
3
4
5
6
7

0.1

0.2

0.3

x1, m

w, mm
0
1
2
3
4
5
6
7

0.1

0.2

0.3

Fig. 6. Predicted steady-state deflections w of [0/45/45/0] composite beams with PZT-5H (a) and
type-1 (VFC (b) embedded ( ) and patched () actuators under 1MV/m.

piezoelectric fiber actuators and both the PZT actuators mentioned previously. The maximum displacement, equal to 2.2mm
(0.733% strain), was obtained by using the actuator PZT-5H. Figure 5b show the response of the composite cantilever beam
under a 1.25-MV/m electric field applied across the different actuators. The max displacement, equal to 4.3mm (1.4% strain),
was reached by using the second-type AFC actuator. Unlike the brittle PZT crystals, AFCs and MFCs are more compliant due
to embedding piezoelectric materials in the form of fibers into a polymeric matrix. It can be concluded that the actuators can
be used to counteract the external stimuli applied to the composite beam; also, it is desirable to use AFCs and MFCs in the
cases of bending deformation, because they are more compliant than PZT crystals.
By increasing the voltage difference applied to the actuator, the axial stress in each ply will increase leading to higher
stress concentrations between the host structure and the actuator, but this may lead to delamination, which is one of the major
reasons for failure in smart structures. In such structures, an alternative method that decreases such stress concentrations is
used. During the manufacturing process, an actuator is incorporated into a layer so that it becomes part of the composite layer,
as shown in Fig. 1. This leads to an enhanced behavior of the overall structure and excludes all the stress concentrations that
existed in the previous patched design illustrated in Fig. 4. Figure 6 depicts the deflection of a composite beam with patched and
embedded actuators, both made of carbon epoxy AS4, with the same stacking sequence [0/45/45/0]. Only one active material
was used in the smart laminated composites. As seen, the deflections in the embedded design are larger both with PZT-5H and
MFC type-1 actuators. This is because geometry discontinuities between the actuator and the composite laminate are absent
and the in-plane stress is distributed in the laminate continuously.

400

TABLE 7. Energy Density in Different Types of Piezoelectric Materials for Different Actuation Modes at an Electric Field
of 1.25 MV/m
Acurator

d33

PZT-5H
PZT-5A
MFC
AFC

593
374
418
540

d31

E1

pm/V

E3

Ev, kJ/m3

Ev, kJ/m3

48
53.2
15.5
10.41

mode 33
2115.8
2371.5

mode 31
2012.5
783.84
516.8
1039.7

GPa
274
171
198
260

61
61
30
35

5. Preliminary Design and a Control Algorithm


Based on the failure analyses of smart composite laminates having various stacking sequences and piezoelectric materials, we could estimate the overall behavior of the laminates and their failure envelopes, which is useful in the preliminary
design of smart composite structures. A recommended voltage can be obtained from the analyses and applied through the
active parts of the composites to prevent their failure.
It is also possible to determine the energy density that defines the maximum energy per unit volume produced by a
piezoelectric device. For the AFC and MFC actuators with longitudinal fibers embedded in the matrix and actuated through
the longitudinal direction (Fig. 2b), the desired extension is parallel to the actuation direction. Therefore, mode 33 of the
piezoelectric material is utilized, and the volumetric energy density is given as [30]

1
E3 d33232 .
2
The energy density is an intrinsic property of a material and does not depend on its geometry. A higher value of
Ev =

1
E3 d332 indicates that the material can perform more mechanical work at the same applied electric fields. The PZT, AFC,
2
and MFC actuators are polarized in the direction parallel to the thickness direction of the piezoelectric layer, and the desired
extension is perpendicular to the poling direction. Therefore, mode 31 of the piezoelectric material is utilized, and the energy
density function is defined as [31]
1
Ev = E1d13232 .

2
The reduction in the volumetric energy density in this case is due to the fact that d13 is usually by a factor of 2 or3
smaller than d33. The reduction in the strain coefficient in the 13-direction is somewhat offset by the increase in the elastic
modulus in the 1-direction. Table 7 shows the energy densities for the different actuators used in the shape control of the
composite structures under an electric field of 1.25MV/m. It can be observed that the energy density obtained from mode 31
of the piezoelectric material had the largest value in the case of the PZT-5H actuator, while the AFC actuator provided the
highest energy density in mode 33.
In order to simulate a shape control, a laminated [0/903]s composite plate with two MFC actuators embedded in its
first ply was considered. The plate was cantilevered at one its end and subjected to a thermal loading. The main goal was to
compensate for the distortion caused by the thermal gradient applied to the composite plate. The geometry of the plate with
the MFC actuators is shown in Fig. 1. The FPF occurred in the 90-ply at 396 MPa. In order to prevent this load from being
reached, a countervoltage was applied to the actuators. The structure was subjected to a linearly increasing temperature. The
temperature change in relation to the number of analysis steps N is shown in Fig. 7a. The evolution of the voltage applied to
the MFC actuators is represented in Fig. 7b. At a relatively large deflection, an electric potential P was applied to compensate
for the thermal deformation induced by the temperature increase. Figure 7c shows the calculated end displacement we in the

401

a
32

T, oC

1500

30
28

P, V

20

1200

16

900

12

600

we, mm

26
24

300

22

20

N
0

10

10

10

Fig. 7. Changes in the temperature T (a) and electric potential P (b) applied and the end deflection
we of a smart beam with (1) and without (2) control in relation to the number of analysis steps N.

x3-direction with and without actuation. Without control, the plate end displacement reached about 18 mm, but it was only
10mm at the tenth step when the control was used.

6. Conclusion
The failure and deformation analyses of smart composite laminates undergoing coupled hygrothermal and mechanical loadings have been studied from a macromechanical point of view. Several failure criteria for overall smart laminates
(PFP and ULF) and individual laminae (the maximum stress, TsaiHill, and TsaiWu) have been considered. Different types
of piezoelectric materials, such as PZT crystals and two types of piezoelectric fiber composites, were integrated with fiberepoxy composite laminates, and the overall failure and deformation performance of these smart laminated composites were
examined. It was found that changes in the temperature and mechanical loads significantly affected the failure behavior of
the laminates, but applying an electric field through the actuators altered the failure behavior only slightly. Thus, it is possible
to develop smart laminated composites with piezoelectric materials without significantly changing their overall mechanical
response owing to the inclusion of active materials. The PZT-5H actuator had the highest actuation value in mode 31, while
the AFC one was dominant in mode 33. The AFCs and MFCs gave more compliant actuators than the brittle PZT crystals.
Furthermore, mismatches in the elastic moduli of the piezocomposites and fiber-epoxy composite laminates were relatively
small, but the stress discontinuities (concentrations) at the interface between the active materials and composite laminates (host
structures) decreased, and the debonding was prevented. It was also observed that, when embedding the active materials into
the composite laminate instead of patching them to the composite laminate, the magnitude of the stress concentration could be
reduced further owing to higher actuation values. Furthermore, the deformation and failure in the laminated composites due
to temperature (and/or moisture) changes could be minimized by applying counter-electric fields through the active layers.
The present failure and deformation studies can support a preliminary design of smart laminated composites by determining
their failure envelopes and overall mechanical response in the case of various lamina properties and stacking sequences, active
materials, and thermomechanical loading conditions.
Acknowledgement. This research is sponsored by the Air Force Office of Scientific Research (AFOSR) under the
grant FA 9550-10-1-0002

402

REFERENCES
1. M. V. Gandhi and B. S. Thompson, Smart Materials and Structures. Springer. 1992.
2. R. E. Newnham, Ferroelectric sensors and actuators: smart ceramics, In: Setter N, Colla E L. J. Ferroelectric Ceramics, 363-80 (1993).
3. B. N. Agrawal, Spacecraft vibration suppression using smart structures, Proc. of the 4th Int. Congress on Sound and
Vibration, 1996.
4. C. K. Lee, Theory of laminated piezoelectric plates for the design of distributed sensors/actuators. Pt I. Governing
equations and reciprocal relationships, J. Acoust. Soc., 87, 144-58 (1990).
5. C. K. Lee, Piezoelectric laminates: theory and experiments for distributed sensors and actuators, J. Intelligent Structural Systems, 28, 75-167 (1992).
6. C. K. Lee and F. C. Moon, Modal sensors/actuators. J. Appl. Mechanics, 57, 434-441 (1990).
7. D. B. Koconis, L. P. Kollr, and G. S. Springer, Shape control of composite plates and shells with embedded actuators, J. Composite Materials, 28, 459-482 (1994).
8. H.S. Tzou, J.P. Zhong, J.J. Hollkamp, Spatially distributed orthogonal piezoelectric shell actuators: theory and applications, J. Sound Vibration, 177, 363-378 (1994).
9. S. K. Agrawal, D. Tong, and K. Nagaraja, Modeling and shape control of piezoelectric actuator embedded elastic
plates, J. Intelligent Material System Structure, 5, 514-522, (1994).
10. N. J. Pagano, Exact solutions for rectangular bidirectional composites and sandwich plates, J. Composite Materials,
4, 20-34 (1970).
11. C. P.Wu, K. H. Chiu, and Y. M. Wang, A review of three-dimensional analytical approaches to multilayered and functionally graded piezoelectric plates and shells, J. Computation Material Continua, 8, 93-132 (2008).
12. S. Raja, D. Dwarakanathan, P. Sinha, and G. Prathap, Bending behavior of piezo-hygrothermo-elastic smart laminated
composite flat and curved plates with active control, J. Reinforced Plastics and Composites, 23, 265-290 (2004).
13. P. Mahato and D. Maiti, Aeroelastic analysis of smart composite structures in hygro-thermal environment, J. Composite Structures, 92, 1027-1038 (2010).
14. M. L. Dano and B. Julliere, Active control of thermally induced distortion in composite structures using macro fiber
composite actuators, J. Smart Materials and Structures, 16, 2315-2322 (2007).
15. R. S. Sandhu, A survey of failure theories of isotropic anisotropic materials, Technical Report, AFFDL-TR-72-7 I, 1972.
16. D. V. Hemelrijck, A. Makris, C. Ramault, E. Lamkanfi, W. Van Paepegem, and D. Lecompte, Biaxial testing of fiberreinforced composite laminates, J. Materials, Design and Applications, 222, 231-239 (2008).
17. J. N. Reddy and A. K. Pandey, A first-ply failure analysis of composite laminates, J Compos. Struct., 25, 371-393
(1986).
18. J. S. Mayes and A. C. Hansen, Composite laminate failure analysis using multicontinuum theory, J. Compos. Sci.
and Techn., 64, 379-394 (2003).
19. T. Takeda, S. Takano, Y. Shindo, and F. Narita, Deformation and progressive failure behavior of woven-fabric-reinforced
glass/epoxy composite laminates under tensile loading at cryogenic temperatures, J. Compos. Sci. and Techn., 65,
1691-1702 (2005).
20. R. B. Pipes, J. R. Vinson, and T. W. Chou, On the hygrothermal response of laminated composite systems, J. Composite Materials, 10, 129-148 (1976).
21. D. Dragan, Ferroelectric, dielectric and piezoelectric properties of ferroelectric thin films and ceramics, Reports on
Progress in Physics, 61, 1267-1324 (1998).
22. W. S. Slaughter, Linearized Theory of Elasticity, Birkhauser, Boston, 2002.
23. C. T. Herakovich, Mechanics of Fibrous Composites, John Wiley and Sons, 1998.
24. J. N. Reddy, Mechanics of laminated plates and shells: theory and analysis, CRC press, 2004.
25. J. N. Reddy, On laminated composite plates with integrated sensors and actuators, J. Engineering Structures, 21,
568-593 (1999).

403

26. A. C. E. Dent, C. R. Bowen, R. Stevens, M. G. Cain, and M. Stewart, Tensile strength of active fiber composites
prediction and measurement, J. Ferroelectrics, 368, 209-215 (2008).
27. T. Fett, D. Munz, and G. Thun, Tensile and bending strength of piezoelectric ceramics, J. Materials Sci., 18. 18991902 (1999).
28. S. Bhalla and C. Soh, Progress in structural health monitoring and non-destructive evaluation using piezo-impedence
transducers, Smart Materials and Structures: New Research, Nova Science Publishers; 2006.
29. Z. Hasan, Controlling Performance of Laminated Composites using Piezoelectric Materials, MS Thesis Mechanical
Engineering, Texas A&M University, 2010.
30. S. Bhalla and C. Soh, Progress in structural health monitoring and non-destructive evaluation using piezo-impedence
transducers, Smart Materials and Structures: New Research, Nova Science Publishers, 2006.
31. J. Donald and D. J. Leo, Engineering Analysis of Smart Material Systems, John Wiley & Sons, Inc, 2007.

404

Você também pode gostar