Sei sulla pagina 1di 18

Materials Research Express

PAPER Related content


- A quadratic piezoelectric multi-layer shell
Finite element analysis of composite laminates element for FE analysis of smart laminated
composite plates induced by MFC
subjected to low-velocity impact based on multiple actuators
Soheil Gohari, Shokrollah Sharifi, Rouzbeh
Abadi et al.
failure criteria
- Mixed-dimensional modeling of
delamination in rare earth-barium-copper-
To cite this article: Z W Wang et al 2018 Mater. Res. Express 5 065320 oxide coated conductors composed of
laminated high-aspect-ratio thin films
Peifeng Gao, Wan-Kan Chan, Xingzhe
Wang et al.

- Multiscale modeling of PVDF matrix


View the article online for updates and enhancements. carbon fiber composites
Michael Greminger and Ghazaleh
Haghiashtiani

This content was downloaded from IP address 148.197.84.103 on 03/08/2018 at 08:16


Mater. Res. Express 5 (2018) 065320 https://doi.org/10.1088/2053-1591/aacca3

PAPER

Finite element analysis of composite laminates subjected to low-


RECEIVED
23 April 2018
velocity impact based on multiple failure criteria
REVISED
22 May 2018
ACCEPTED FOR PUBLICATION
Z W Wang1,2, J P Zhao1,2 and X Zhang1,2
14 June 2018 1
School of Mechanical and Power Engineering, Nanjing Tech University, Nanjing 211816, People’s Republic of China
2
PUBLISHED Jiangsu Key Lab of Design and Manufacture of Extreme Pressure Equipment, Nanjing 211816, People’s Republic of China
27 June 2018
E-mail: jpzhao_njtech@163.com

Keywords: composite laminates, low-velocity impact, failure criteria, analytical approximation, computational cost

Abstract
Progressive damage models based on continuum damage mechanics are used in combination with
cohesive elements to explore the effect of three different failure criteria including Chang-Chang,
Hashin and Puck criteria, on the structural response and the failure mechanisms of composite
laminates subjected to low-velocity impact. Three different failure criteria and damage evolution
laws based on equivalent strain are used for intralaminar damage models, and the delamination is
simulated by the bilinear cohesive model based on quadratic criteria. A new numerical
optimization method combining analytical approximation and Golden section Search has been
applied in Puck criteria to search the fracture plane. Numerical analysis is performed on two
composite laminates specimens with different materials, layups and impact energy to study the
impact force-time, force-displacement and absorbed energy, computational cost, as well as the
damage evolution behaviors of fiber, matrix and delamination. The numerical results with three
different failure criteria show acceptable accord with available experimental data, which validate
the accuracy of the proposed damage model. Moreover, this research can be helpful to select
appropriate failure criteria in the progressive failure analysis of composite laminates under low
velocity impact.

1. Introduction

Composite laminates composed of fiber-reinforced plies are widely used in aerospace, automotive, blade of
wind power generation and marine because of high strength/stiffness-to-weight ratio, and excellent resistance
to fatigue and corrosion. However, composite laminates are susceptible to impact, which may cause barely
visible impact damage (BVID), therefore, it is important to predict low-velocity impact responses of structures
by developing theoretical and numerical methods [1, 2].
The composite material under impact load often appear multi-damage coupling failure, besides, the
damage initiation and evolution process are complex, the development of failure criteria for fiber-reinforced
composite, have been the focus for many years. Tsai-Wu [3] introduced a method that based on the
maximum stress with seven parameters interacting with each other, which is still considered as the simplest
way to predict the initiation of damage. Hill [4, 5] proposed the Tsai–Hill criterion which is derived from a
yield criterion for orthotropic ductile materials, and extends the classical von Mises yield condition to
orthotropic materials. Chang-Chang criteria have been put forward by Chang F K [6], which was applied to
laminated composites containing stress concentrations, however, ignore the out-of-plane shear deformation
effects. Hashin [7, 8] who believe that the composite failure was caused by the stresses acting on the inclined
fracture plane, proposed failure criteria for plane stress based on the Mohr Coulomb failure theory, though
he didn’t try to get the inclined fracture plane because of the calculation cost. Puck and Schürmann [9]
extend the model of Hashin, and take the fracture plane into consideration, the theory was ranked very highly

© 2018 IOP Publishing Ltd


Mater. Res. Express 5 (2018) 065320 Z W Wang et al

in the world wide failure exercise(WWFE) [10, 11]. LaRC04 criteria [12] take consideration of non-linear
matrix shear behavior, which was developed by the NASA after the WWFE. So many failure criteria have
been used to predict the damage onset of fiber-reinforced composite, three representative criteria of them
were adopted in this paper, aim to compare their effect on composite laminates subjected to low-velocity
impact.
Currently, many experimental and numerical analysis have been performed to explore the response of
composite laminates under low-velocity impact. Hyung Yun Choi et al [13] and Meo et al [14] adopted the
Chang-Chang criteria to predict composite failure. R C Batra et al [15] used Hashin criteria and exponential
stiffness degradation mode to predict the damage mode of composite laminates at low-velocity impact, and the
simulated force-time curves had little error with experimental results. Y Shi et al [16], D Feng et al [17] and B B
Liao et al [18] used simplified Puck criteria. N LI et al [19] adopted Puck criteria with a degradation strategy
depending on the orientation of the inter-fiber failure (IFF) plane, Golden section Search method was applied to
search the fracture plane. Markus Kober et al [20] compared three failure criteria (Puck, LaRC04, Tsai–Hill)
theoretically only, and ignore the actual application. P F Liu et al [21] compared matrix compression failure of
Chang-Chang, Hashin and Puck criteria, and the fracture plane angle was set to the same value in the Puck
criteria. However, comprehensive comparison between three respective failure criteria on the efficiency and
accuracy is not operated until now.
For Puck criteria, It’s necessary to determine the orientation of the potential fracture plane. Wiegand et al
[22] put forward a methods combining Golden section Search (GSS) and curve interpolation technique, and the
problem of the numerical search converging on a local maxima was not solved. F J Schirmaier et al [23] proposed
an approach called Selective Range Golden section Search (SRGSS), in order to get reliable results, 18 periodic
supporting points plus 6 supporting points per local maxima need to be calculated, it will lead a high
computational cost. D M Thomson et al [24] applied analytical method to narrow the search range, however,
ignored the circumstance that minimum value of stress tractions may also attract the global fracture planes. In
this work, a new efficient method is proposed based on the research of D M Thomson et al [24].
Intralaminar damage and interlaminar delamination of composite laminates under low-velocity impact are
studied in this paper. Three representative criteria (Chang-Chang, Hashin, and Puck criteria) are compared with
each other. In this way, the characteristics, similarities, differences and possible weak points of the criteria
become obvious. Chang-Chang and Hashin criteria adopt strain-based failure criteria. In order to reduce
computational cost, Puck criteria apply a new numerical optimization method to search fracture plane angle.
The delamination failure is simulated by the bilinear cohesive model proposed by Camanho and Davila [25].
Finally, two composite specimens subjected to low-velocity impact are used to demonstrate the proposed
numerical scheme.

2. Intralaminar damage model

2.1. Damaged constitutive of composites


In this paper, the composite plastic deformation and the strain rate effect are neglected. The damage variable is
an effective method to simulate the damage process of composite material. The constitutive of orthotropic single
damaged composite plate can be expressed as:

si, j = C i, j (d) · ei, j (1)

where si, j and ei, j are the effective Cauchy stress and Cauchy strain, respectively, while Ci, j (d) is damaged
stiffness matrix, which varies with damage variables d. The damaged stiffness matrix was given by W Guo
et al [26]
⎡ dC11 dC12 dC13 ⎤
⎢ ⎥
⎢ dC 21 dC 22 dC 23 ⎥
⎢ dC31 dC32 dC33 ⎥
C i , j ( d) = ⎢ ⎥ (2)
dC 44
⎢ ⎥
⎢ dC55 ⎥
⎢⎣ dC 66 ⎥⎦

2
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

dC11 = (1 - d f ) E1(1 - n 23 n 32) G


dC12 = (1 - d f )(1 - d m) E 2(n12 + n13 n 32) G
dC13 = (1 - d f )(1 - d m) E3(n13 + n12 n 23) G
dC 21 = (1 - d f )(1 - d m) E1(n 21 + n 31 n 23) G
dC 22 = (1 - d f )(1 - d m) E 2(1 - n13 n 31) G
dC 23 = (1 - d f )(1 - d m) E3(n 23 + n 21 n13) G
dC31 = (1 - d f )(1 - d m) E1(n 31 + n 21 n 32) G
dC32 = (1 - d f )(1 - d m) E2 (n 32 + n12 n 31) G
dC33 = (1 - d f )(1 - d m) E3(1 - n12 n 21) G
dC 44 = (1 - d f )(1 - smt d mt )(1 - smc d mc ) G12 G
dC55 = (1 - d f )(1 - smt d mt )(1 - smc d mc ) G 23 G
dC 66 = (1 - d f )(1 - smt d mt )(1 - smc d mc ) G 31 G
G = 1 (1 - n 12
2
- n 223 - n 13
2
- 2n12 n 23 n13) (3)
in which df and dm can be expressed as
d f = 1 - (1 - d fc )(1 - d ft )
d m = 1 - (1 - d mc )(1 - d mt ) (4)

E1, E2, E3 are the Young’ s moduli, where ‘1’ denotes the fiber direction and ‘2’ and ‘3’ denote the transverse
directions. G12, G13, G23 are the shear moduli and νij(i, j=1, 2, 3) are the Poisson’s ratios. d f and dm represent
damage degree of fiber and matrix, respectively, ‘t’ and ‘c’ in the damage variables d ft , d fc , dmt , and dmc denote
the tension and compression, considering the convergence problem of finite element computation, the
maximum value of damage variables are set as 0.99. smt=0.9 and smc=0.5 are shear stiffness damage
coefficients due to matrix tension and compression [27].

2.2. Intralaminar damage models


Chang-Chang, Hashin and Puck criteria have been widely used to predict the damage of fiber-reinforced
composite. The intralaminar damage model includes fiber tension, fiber compression, matrix tension and
matrix compression. Considering that the low-speed impact damage of the composite is mainly based on the
matrix failure, the same fiber failure criteria is chosen in this paper.

2.2.1. Fiber tension and compression

(a) Fiber tension (e11  0):


⎧ ⎛ ⎞2
⎪ f T = ⎜ e11 ⎟ - 1  0
⎪ 1 ⎜ T ⎟
⎨ ⎝ e0,1 ⎠ (5)
⎪ T XT
⎪ e0,1 =
⎩ E1

(b) Fiber compression (e11 < 0):

⎧ ⎛ ⎞2
⎪ f C = ⎜ e11 ⎟ - 1  0
⎪ 1 ⎜ C ⎟
⎨ ⎝ e0,1 ⎠ (6)
⎪ C XC
⎪ e0,1 =
⎩ E1

where f1T and f1C represents the levels of fiber tension and compression failure, eT0,1 and eC0,1 are the initial failure
strains, XT and XC are the fiber tensile and compressive strengths, respectively.
The damage evolution laws of fiber tension and compression can be seen in figure 1 and expressed as

eTf ,1(C) ⎛ eT (C) ⎞


d ft ( fc ) = ⎜⎜1 - 0,1 ⎟⎟ (7)
eTf ,1(C) - eT0,1(C) ⎝ e11 ⎠

where eTf ,1 and eCf ,1 are the fiber critical failure strains when the damage variables reach complete failure.

3
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 1. Fiber failure model.

The characteristic length L is introduced to eliminate the mesh sensitivity problem, and the critical failure
strains can be calculated by the fiber fracture energy Gft ( fc ), which can be expressed as

G ft ( fc )
eTf ,1(C) = (8)
XT (C) L

2.2.2. Matrix tension and compression


① Chang-Chang criteria [6, 14]

(a) Matrix tension (e22  0)

⎧ ⎛ ⎞2 ⎛ ⎞2
⎪ f T = ⎜ e22 ⎟ + ⎜ g12 ⎟ - 1  0
⎪ 2 ⎜ ⎟
⎨ ⎝ e0,2 ⎠
T
⎝ e0,12 ⎠
(9)
⎪ T Y T S
⎪ e0,2 = , e0,12 = 12
⎩ E2 2G12

(b) Matrix compression (e22 < 0)

⎧ ⎛ ⎞2 ⎛ ⎞2 eC e
⎪ f C = ⎜ e22 ⎟ + ⎜ g12 ⎟ + 0,2 22 - e22 - 1  0
⎪ 2 ⎝ 2e0,12 ⎠ ⎝ e0,12 ⎠ 4e 20,12 eC0,2
⎨ (10)
⎪ C YC
⎪ e0,2 =
⎩ E2

where f 2T and f 2C represents the levels of matrix tension and compression failure, YT and YC are the matrix
tensile and compressive strengths, respectively, S12 is the in-plane shear strength. eT0,2 and eC0,2 are the matrix
initial failure strains, e0,12 is the initial shear failure strain.
② Hashin criteria [7, 8, 28]

(a) Matrix tension (e22 + e33  0)

⎧ ⎛ ⎞2 ⎛ ⎞2 ⎛ ⎞2 ⎛ ⎞2
⎪ f T = ⎜ e22 + e33 ⎟ + ⎜ 1 ⎟ ⎛⎜g 2 - E2 · E3 e e ⎞⎟ + ⎜ g12 ⎟ + ⎜ g13 ⎟ - 1  0
⎪ 2 ⎜ ⎟ 23 22 33
⎨ ⎝ e0,2 ⎠
T
⎝ e0,23 ⎠ ⎝ G 23 2 ⎠ ⎝ e0,12 ⎠ ⎝ e0,13 ⎠
(11)
⎪ T YT S S S
⎪ e0,2 = , e0,12 = 12 , e0,13 = 13 , e0,23 = 23
⎩ E2 2G12 2G13 2G 23

(b) Matrix compression (e22 + e33 < 0)

4
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 2. Stress state and fracture surface.

⎧ ⎛ ⎞ ⎛⎛ ⎞2 ⎞
⎪ f C = ⎛⎜ E2 e22 + E3 e33 ⎞⎟ + ⎜ e22 + e33 ⎟ ⎜⎜ E2 e0,2 ⎟ - 1⎟
2 C

⎪ 2 ⎜ ⎟ ⎜ ⎜ ⎟ ⎟
⎝ 2G12 e0,12 ⎠ ⎝ e0,2 ⎠ ⎝⎝ 2G12 e0,12 ⎠
C


⎪ ⎛ 1 ⎞⎛

2
E 2 · E3 ⎞ ⎛ g12 ⎞ 2 ⎛ g13 ⎞2 (12)
⎪ + ⎜ ⎟ ⎜ g 2
23 - e22 e33 ⎟ + ⎜ ⎟ + ⎜ ⎟ -10
⎝ e0,23 ⎠ ⎝ G 23 2 ⎠ ⎝ e0,12 ⎠ ⎝ e0,13 ⎠

⎪ C YC
⎪ e0,2 =
⎩ E2

where e0,13 and e0,23 are the initial shear failure strain., S13 and S23 are the out-of-plane shear strength.
③ Puck criteria [9]
According to the experimental evidences, Puck and Schürmann formulated failure criteria for IFF assuming
that the fracture is mainly provoked by the stress components acting on the so-called fracture plane [9], which
can be seen in figure 2. Providing the fracture plane angle as θ, the inter-fiber failure is a function of the fracture
plane angle, and is related to the normal stress and the shear stress on the fracture plane.
In figure 2, the coordinate systems 1-2-3 and l-n-t are the global laminate coordinate system and the fracture
plane coordinate system, respectively. The action plane-related stresses and strains are only a function of the
fracture plane angle and derived from the effective stresses and strains as follows:

⎧ sn (q ) = s22 cos2 q + s33 sin2 q + 2t23 cos q sin q



⎨ snt (q ) = (s22 - s33) cos q sin q + t23 (cos2 q - sin2 q )

⎩ snl (q ) = t31 sin q + t12 cos q
⎧ en (q ) = e22 cos2 q + e33 sin2 q + g23 cos q sin q

⎨ ent (q ) = 2(e22 - e33) cos q sin q + g23 (cos2 q - sin2 q ) (13)

⎩ enl (q ) = g31 sin q + g12 cos q

According to the fracture plane’s normal stress (sn), the IFF can be divided into two cases

(a) Inter-fiber tension (sn (q )  0)


1
⎡⎛ p^t y ⎞
2
⎛ s ( q ) ⎞2 ⎛ s ( q ) ⎞2 ⎤ 2 p^t y
1
f 2 (q ) = ⎜⎜ T - A ⎟⎟ s n2 (q ) + ⎜ nt A ⎟ + ⎜ n1 ⎟ ⎥ + A sn (q )
T ⎢ (14)
⎢⎝ Y R^ y ⎠ ⎝ R^^ ⎠ ⎝ S12 ⎠ ⎥⎦ R^ y

(b) Inter-fiber compression (sn (q ) < 0)

1
⎡⎛ p c ⎞2 ⎛ s ( q ) ⎞2 ⎛ s ( q ) ⎞2 ⎤ 2 p^c y
f 2C (q ) = ⎢⎜⎜ A sn (q ) ⎟⎟ + ⎜ nt A ⎟ + ⎜ n1 ⎟ ⎥ + A sn (q )
^y
(15)
⎢⎝ R^ y ⎠ ⎝ R^^ ⎠ ⎝ S12 ⎠ ⎥⎦ R^ y

5
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Table 1. Recommended inclination


parameters [9].
t c
p^ p^ t (c )
p^^

GFRP 0.30 0.25 0.20 to 0.25


CFRP 0.35 0.30 0.25 to 0.30

with
⎧ A YC
⎪ R^^ =
⎪ 2(1 + p^^ c
)
⎪ t (c )
⎪ p^ y t
p^^(c ) pt (c )
⎨ = cos 2 y + ^ sin2 y (16)
⎪ R^Ay R^^A
S12

⎪ cos2 y = s nt
2
s2
⎪ , sin2 y = 2 n1 2
⎩ s nt + s n1
2 2
s nt + s n1
A t c
where R^^ is fracture strength, p^^ , p^^ , p^t , p^c  are the inclination parameters, the recommended values were
given in table 1.
In Puck criterion, the stress exposure is acquired by three tractions, sn, snl and snt , which act on the potential
fracture plane and can be obtained by equation (13). Since the failure criterion is a function of the angle, θ,
through three tractions. Using the knowledge of matrix failure function characteristics, a new fracture angle
search algorithm based on the research of D M Thomson et al [24], the Analytical Approximation Golden
section Search (AAGSS), is developed. Firstly, the expressions of three tractions need to be simplified into a
single sinusoidal function of θ, which can be achieved by using basic trigonometric relations and superposition
of sinusoidal waves of the same frequency, as shown in equation (17)
A cos (q + a) + B cos (q + b ) = [A cos (a) + B cos (b )] + [A sin (a) + B sin (b )]
⎧ ⎡ A sin (a) + B sin (b ) ⎤ ⎫
· cos ⎨q + tan-1 ⎢ ⎥⎬ (17)
⎩ ⎣ A cos (a) + B cos (b ) ⎦ ⎭

where A and B are amplitude of waves, α and β are offset phases. The resulting expressions of three tractions are
given below [24]
⎧ ⎛ ⎡ ⎤⎞
⎪ sn (q ) = (s22 + s33) + ⎛⎜ s22 - s33 ⎞⎟ + t 223 cos ⎜2q + tan-1 ⎢ - 2t23 ⎥ ⎟
2

⎪ 2 ⎝ 2 2 ⎠ ⎝ ⎣ s22 - s33 ⎦ ⎠

⎪ ⎛ ⎞2 ⎛ ⎡ ⎤⎞
⎨ snt (q ) = ⎜ s22 - s33 ⎟ + t 223 cos ⎜2q + tan-1 ⎢ ∣s22 - s33∣ ⎥ ⎟ (18)
⎪ ⎝ 2 2 ⎠ ⎝ ⎣ 2t23 ⎦ ⎠

⎪ s (q ) = t 2 + t 2 cos ⎛⎜q + tan-1 ⎡ t13 ⎤ ⎞⎟
⎪ nl 13 12 ⎢ ⎥
⎩ ⎝ ⎣ t12 ⎦ ⎠

For arbitrary loading conditions, the orientation of the fracture plane can be determined by the IFF function,
which achieved the maximum value within the interval (−90°, 90°). The Puck IFF criterion is a combination of
three sinusoidal, each of the tractions, sn, snl and snt , promotes their preferred fracture plane and will attract the
resulting global fracture plane. Therefore, when one of ∣sn∣max , ∣snl∣max and ∣snt ∣max takes the dominant position
with respect to the others, the resulting fracture plane will tend to the respective value of q1 to q6, which are
shown in table 2. However, in general case, the final outcome is decided by the interaction of three tractions and
the angle of potential fracture plane will lie in between these values.
Briefly, the paper considers these six angles, q1 to q6, as initial candidates for the resulting global maximum
and take the plane with the highest two exposures as starting point for Golden section Search. In this way, the
cost of the search is greatly reduced compared with traditional GSS.
An illustrative example for arbitrary stress state (s22 = 8, s33 = 2, s12 = 3, s23 = -4 and s13 = 4 ) is
adopted to compare the accuracy and efficiency of AAGSS with GSS. The vertical lines in figure 3(a) indicate the
preferred fracture planes for each traction. After the analytical approximation, the search interval is narrowed to
(q1, q3), and the numerical search quickly converges in the second step. Obviously, the AAGSS can significantly
reduce the computational cost compared with traditional GSS.
Even with the AAGSS, the fracture plane search is still repeated at every time step and integration point.
Another way proposed by D M Thomson et al [24] is applied in this paper. Actually, it is unnecessary to calculate

6
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 3. (a) Analytical approximation of AAGSS (b) Comparison of numerical efficiency and accuracy.

Table 2. Extreme value of three tractions.

Maxima Location value

tan-1 ⎡⎣ s ⎤
1 2t23 (s22 + s33) s22 s33 2
sn (s22 + s33) > 0 q1 = ⎦ ∣sn∣max = + ( - ) + t 223
2 22 - s33 2 2 2

(s22 + s33) < 0 q1 = (p + tan ⎡⎣


1
2
-1 2t23 ⎤
s22 - s33 ⎦ ), ( 2t23
s22 - s33
<0) ∣sn∣max =
(s22 + s33)
2
- ( s22
2
-
s33 2
2 ) + t 223

or q = (-p + tan -1 ⎡ 2t23 ⎤


1
1
2
⎡ ⎤ t
⎣ s22 - s33 ⎦ )
snl q2 = -tan-1 ⎣ t13 ⎦ ∣snl∣max = t13
2
+ t12
2
12

q3,4 =  2 tan-1 ⎡⎣ 222t 33 ⎤⎦


1 s -s
snt
23

q5,6 = 2 p  tan-1 ⎡⎣ 222t 33 ⎤⎦ ,


( )( )
s -s s22 - s33 s22 s33 2
1
23 2t23
<0 ∣snt ∣max = ( 2
- 2 ) + t 223

-1 ⎡ s22 - s33 ⎤
or q5,6 =
1
2 (-p  tan ⎣ 2t23 ⎦ )

the integration points whose stress states do not satisfy the requirements of damage initiation, since a criterion
has been defined to determine the upper bound, or worst case scenario, for the given stress state by using the
known maxima for the three tractions obtained in table 2, as shown in equation (19)
1
⎡⎛ p^t y ⎞
2
⎛ ∣s ∣max ⎞2 ⎤ p^t y
⎢ 1 ⎛ ∣s ∣max ⎞2 ⎥ 2
fup = ⎜⎜ T - A ⎟⎟ (∣sn∣max )2 + ⎜ nl A ⎟ + ⎜ nt ⎟ + A ∣sn∣max (19)
⎢⎝ Y R^ y ⎠ ⎝ R^^ ⎠ ⎝ S12 ⎠ ⎥⎦ R^ y

where fup is the up bound. If fup of one integration point is less than one, the f2 will also be less than one,
meaning that the possibility of failure at this point can be immediately excluded without the search of fracture
plane.
This paper adopts the damage evolution laws for matrix based on equivalent strain, the tensile stress-strain
curve is shown in figure 4. The damage evolution law is written as

⎧ eTf ,(eqC) ⎛ eT (C) ⎞


⎪d ⎜1 - 0, eq ⎟
mt (mc ) = T (C ) ⎜ (mc) ⎟

⎪ e f , eq - eT0,(eqC ) ⎝ emt
eq ⎠
⎨ (20)
⎪ T (C) s eq
mt (mc )
eeq
mt (mc )
⎪ s 0, eq = T (C) 0.5 , eT0,(eqC ) = T (C) 0.5

⎩ ( f2 ) ( f2 )

where eT0, eq and eC0, eq are the matrix tension and compression initial failure equivalent strain. emt eq and e eq are
mc

equivalent strain, s eq and s eq are equivalent stress, for the three different failure criteria, the equivalent strain
mt mc

and stress are shown in table 3, which can be expressed in terms of the effective strain and stress components
[29]. eTf , eq and eCf , eq are the matrix tension and compression critical failure equivalent strain, respectively, for the
Chang-Chang and Hashin, eTf , eq and eCf , eq can be expressed as

7
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 4. Matrix failure model.

Table 3. Equivalent strain and stress for three failure criteria.

eeq seq
ás22ñáe22ñ + t12 g12
Chang-Chang emt
eq = áe22ñ2 + g 12
2
s eq
mt
=
emt
eq
á-s22ñá-e22ñ + t12 g12
emc
eq = á-e22ñ2 + g 12
2
s eq
mc
= emc
eq
ás22ñáe22ñ + ás33ñáe33ñ + t12 g12 + t13 g13 + t23 g23
Hashin emt
eq = áe22ñ2 + áe33ñ2 + g 12
2
+ g 223 + g 13
2
s eq
mt
=
emt
eq
á-s22ñá-e22ñ + á-s33ñá-e33ñ + t12 g12 + t13 g13 + t23 g23
emc
eq = á-e22ñ2 + á-e33ñ2 + g 12
2
+ g 223 + g 13
2
s eq
mc
= emc
eq

Puck emt
eq = en2 + enl
2
+ ent
2
s mt
eq = s n2 + s nl
2
+ s nt
2

emc
eq = enl
2
+ ent
2
s mc
eq = s nl
2
+ s nt
2

2Gmt (mc )
eTf ,(eqC) = (21)
sT0,(eqC) L
where Gmt (mc ) are the matrix fracture energy, for Puck criteria, it can be expressed as [30]
⎡ 0 -0.5
2 s b
2
⎛ s 0 b ⎞2 ⎛ s 0 b ⎞2 ⎤
eTf ,(eqC) = ⎢ nkn + ⎜ n1 n1 ⎟ + ⎜ nt nt ⎟ ⎥ (22)
L ⎢⎣ G2c ⎝ G12c ⎠ ⎝ G 23c ⎠ ⎥⎦

with
⎧ 0 sn snl snt
⎪ sn = ( f T (C))0.5 , snl = ( f T (C))0.5 , snt = ( f T (C))0.5
0 0
⎪ 2 2 2
⎨ (23)
⎪ b = áenñ , b = enl , b = ent


n
eeq
nl
eeq
nt
eeq

where G2kc is the critical fracture energy release rates corresponding to transverse tension or compression, G12c
and G23c are the in-plane shear critical fracture energy release rates, si0 (i = n, nl , nt ) is the stress component at
matrix failure initiation, b i0 (i = n, nl , nt ) denotes mix-mode radio, the symbol áñ indicates the McCauley
Operator, defined áxñ = max (0, x).

3. Interlaminar delamination damage models

Camanho and Davila [25] proposed quadratic failure criteria to predict damage initiation at the interface of
adjacent composite layers, which ignore the influence of normal compressive stress, and can be expressed as

⎛ ásnñ ⎞2 ⎛ st ⎞2 ⎛ ss ⎞2
Fcoh = ⎜ ⎟ +⎜ ⎟ +⎜ ⎟ (24)
⎝ N ⎠ ⎝T ⎠ ⎝S⎠

where si (i = n, s, t ) is the traction stress vector in the normal n and shear directions s and t, respectively, N, S
and T are the interface strength, while Fcoh is the level of interface layer damage. Once Fcoh = 1 is satisfied, the
material stiffness is gradually degraded corresponding with a damage variable dcoh, which value range from zero
when damage initiates to one when complete delamination occurs.

8
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 5. Bilinear cohesive model.

For a linear softening process, the damage dcoh for delamination evolution can be defined by
dmf (dm
max
- dm
0
)
dcoh = (25)
dm
max
(dmf - d m
0
)

where d mmax
denotes the maximum equivalent displacement in the load process, which can be expressed as
dm max = max (d m max
, dm), where dm = dn2 + d2s + dt2 , di (i = n, s, t ) is the traction displacement vector. d m
0

and dmf are the displacement at the initial failure and complete failure, respectively, and they can be determined
by the Benzeggagh-Kenane Law [31], which can be written as
⎧ 1 + b2

⎪ d n0 d s0 , dn > 0
dm0
=⎨ (d s0)2 + (bd n0)2


⎩ (d s ) + (d t ) , dn < 0
0 2 0 2


⎪ 2 ⎡ ⎛ b 2 ⎞h ⎤
⎪ 0 ⎢ G IC + ( G IIC - G IC ⎜
) ⎟ ⎥ , dn > 0
dmf = ⎨ Kd m ⎣ ⎝1 + b2 ⎠ ⎦ (26)


⎩ (d s ) + (dt ) , dn < 0
f 2 f 2

where h is the parameter defined in the B-K law [31], β denotes the mixed-mode radio (b = (d2s + dt2) dn2 ),
G IC and G IIC are the mode-I and mode-II interlaminar fracture toughnesses, respectively. The typical
interlaminar bilinear cohesive model is shown in figure 5.

4. Numerical results and discussion

In this paper, two specimens are adopted to compare the effect of three failure criteria on dynamic response of
composite laminates subjected to low-velocity impact. The finite element package ABAQUS/Explicit is
employed to establish the finite element models, and the damage initiation and evolution of intralaminar
damage presented in the previous section are implemented by a user-material subroutine VUMAT with
FORTRAN.
In the finite element model, the eight-node solid elements (C3D8R) with reduced integration are chosen for
the laminate, impactor and supporting plate, and the eight-node cohesive elements (COH3D8) are embedded
between adjacent composite layers. Considering the real situation, the failed elements are allowed to remain in
the model. The steel spherical impactor which is fixed in all directions except impact direction, and the
supporting plate which is fixed in all directions to represent the experimental constraint conditions, are regarded
as rigid body. The impactor is imposed the appropriate velocity at the initial contact with composite laminate to
simulate the drop-weight impact, the penalty method is used to deal with the contact between laminate and
impactor, and the friction contact coefficient is set to 0.3. Moreover, hourglass control and distortion control are
adopted in this research. The finite element model is shown in figure 6

4.1. [03/45/−45]s laminate


D Feng and F Aymerich [17] operated drop-weight impact tests on [03/45/−45]s laminate for various impact
energies ranging between 1 J and 8 J. HS300/ET223 is adopted for this laminated plate and material parameters
are listed in table 4. The geometry size of the rectangular specimen is 65 mm×87.5 mm×3.2 mm, which is
put on a supporting plate with 45 mm×67.5 mm rectangular opening, the thickness of single layer is 0.25 mm.
A 2.34 kg steel spherical impactor with a diameter of 12.5 mm is used in the drop-weight impact simulation. The

9
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 6. Finite element model for impact analysis.

Figure 7. Experimental and numerical results at 2 J impact energy (a) Impact force-time curves (b) Impact force-displacement curves.

Table 4. Material parameters for HS300/ET223 [17].

Single layer properties E1=122 GPa; E2=E3=6.2 GPa; u12 = u13=0.35; u23=0.5; G12=G13=4.4 GPa;
G23=3.7 GPa;
XT=1850 Mpa; XC=1470 MPa; YT=29 MPa; YC=140 MPa;
S12=S13=S23=65 Mpa;
Gft = 92 N mm-1; Gfc = 80 N mm-1; Gmt = 0.52 N mm-1; Gmc = 1.61 N mm-1
k
G2C = 0.52 N mm-1; G12C = G 23C = 0.92 N mm-1;
Interface properties G IC = 0.52 N mm-1; G IIC = G IIIC = 0.92 N mm-1;
KN = 120 GPa mm-1; KN = KT = 43 GPa mm-1;
N = 30 MPa; S = T = 80 MPa

impact energy which is employed in this research are 2 J, 4 J, and 8 J. The physical thickness of cohesive element
is set to 0.001 mm. From the drop-weight impact experiment by D Feng and F Aymerich [17], the impact contact
time is about 4.2 ms and a 5 ms calculation time period is set in the explicit FEA.
Figures 7–9 show the impact force-time curves and impact force-displacement curves using three failure
criteria at the impact energy of 2 J, 4 J, and 8 J, respectively. In general, the FEA results are in good agreement
with the experimental data. When the impactor gets the lowest location at about t=2 ms, the impact force gets
the maximum value and the velocity of impactor drops to zero. And then the impactor begins to rebound until
the complete separation with composite laminate at about t=4.2 ms, the predicted time of peak force time and

10
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 8. Experimental and numerical results at 4 J impact energy (a) Impact force-time curves (b) Impact force-displacement curves.

Figure 9. Experimental and numerical results at 8 J impact energy (a) Impact force-time curves (b) Impact force-displacement curves.

separation time for Chang-Chang, Hashin and Puck criteria are 2.3 ms and 4.75 ms, 2.3 ms and 4.5 ms and 2 ms
and 4.1 ms, respectively. Puck criteria lead higher consistent with experimental data.
Through experiments, D Feng and F Aymerich [17] found that the fiber tension damage occurs when the
impact energy is larger than 6 J, it also can be seen in experimental impact force-time curve that a sudden drop
occurs at the peak force, and only Puck criteria succeeds to capture the phenomenon. From experimental and
simulation results, it can be concluded that fiber tension damage is the main reason for the loss of composite
laminates bearing capacity.
Figures 7(b), 8(b), and 9(b) show the impact force-displacement curves, the macroscopic stiffness of the
material decreased due to the composite material damage, which leads that the impact force of loading stage is
greater than unloading stage at the same displacement. By comparison, the maximum displacement obtained by
Puck criteria leads to minimum errors, which are 5.6% at 2 J, 0.7% at 4 J, and 4.7% at 8 J.
Figure 10 shows the peak impact force and absorbed energy at different energy (2 J, 4 J and 8 J). Compared
with the experimental data, the errors for peak impact force using Chang-Chang, Hashin and Puck criteria are
6.5%, 4.7% and 1.5% (at 2 J impact energy), 12.3%, 4.1%, 8.5% (at 4 J impact energy) and 7.3%, 10.9% and
4.4% (at 8 J impact energy). The absorbed energy can be calculated by initial kinetic energy reducing the final
kinetic energy, it can be clearly seen that the absorbed energy rises with the increase of impact energy, because
higher impact energy leading to a larger damage area, which will dissipate more energy. By comparison, three
failure criteria lead to basically consistent results, and Chang-Chang criteria leads to slightly larger errors.
Figure 11 shows the comparison of computational cost at the impact energy of 8 J. It’s obvious that the Puck
criteria with AAGSS has reduced the computational cost significantly compared with GSS, however, it is still
higher than the Chang-Chang criteria. It’s can be seen that the initial contact lead to a highest cost, and a sudden-
rise occurs at approximate same time by using different failure criteria when the damage initiates. The cost raises
with the increase of damage degree. A significant rise appears at 40% of simulation progress for Chang-Chang,

11
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 10. Experimental and numerical results at different impact energy (a) Peak impact force (b) Absorbed energy.

Figure 11. The comparison of computational cost.

Figure 12. Cross-sectional view of matrix damage at about t=2.4 ms under the energy of 4 J.

12
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 13. Matrix tension and matrix compression at the energy of 8 J (a) Chang-Chang criteria, (b) Hashin criteria, (c) Puck criteria.

Table 5. AS4/PEEK material parameters [34, 35].

Single layer properties E1=93.7 GPa; E2=E3=7.45 GPa; u12 = u13 = u23=0.201; G12=G13=G23=3.97 GPa;
XT=2945 Mpa; XC=1650 MPa; YT=54 MPa; YC=240 MPa;
S12=S13=S23=89.9 Mpa;
Gft = 91.6 N mm-1; Gfc = 79.9 N mm-1; Gmt = 0.22 N mm-1; Gmc = 0.76 N mm-1
G2kC = 0.52 N mm-1; G12C = G 23C = 0.97 N mm-1
Interface properties G IC = 0.52 N mm-1; G IIC = G IIIC = 0.97 N mm-1;
KN = 120 GPa mm-1; KN = KT = 43 GPa mm-1;
N = 30 MPa; S = T = 80 MPa

Puck with GSS and Puck with AAGSS, respectively, where the peak force reaches the maximum value and the
impactor near the lowest point. Then the computational cost drops until reaching the termination time. When
the impactor begins to rebound, the stress states of less and less integration points satisfy the condition given in
equation (19) and needn’t search the fracture plane, which leads that the computational cost between Chang-
Chang criteria and Puck with AAGSS tends to be consistent gradually.

13
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 14. Experimental and numerical results at 2 J impact energy (a) Impact force-time curves (b) Impact force-displacement
curves.

Figure 15. Experimental and numerical results at 5 J impact energy (a) Impact force-time curves (b) Impact force-displacement
curves.

Figure 16. Experimental and numerical results at 9.2 J impact energy (a) Impact force-time curves (b) Impact force-displacement
curves.

Figure 12 shows the cross-sectional images of matrix damage under the energy of 4 J at about t=2 ms, the
matrix tension damage appears initially at the bottom of composite, and evolves from bottom to the top. On the
contrary, the matrix compression damage appears initially at the top and evolves to the bottom, which is similar
to the conclusion Faggiani et al [32] and Kim et al [33] got. The delamination can be clearly observed in

14
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Figure 17. Experimental and numerical results at different impact energy (a) Peak impact force (b) Absorbed energy.

Figure 18. Delamination damage of 90°/0° interface under the energy of 5 J (a) Experimental (b) Chang-Chang criteria (c) Hashin
criteria (d) Puck criteria.

figure 12(c). Figure 13 shows the matrix damage of each layer under the impact energy of 8 J. The most severe
matrix tension and matrix compression damage regions are at the bottom layer and up layer, respectively. The
damage morphology of the matrix presents groundnut shape evolving along the direction of fiber, which is
consistent with the simulation result with B B Liao et al [18] and P F LIU et al [21]. However, the matrix damage
evolution direction of several top layer was perpendicular to the direction of fiber, because of buckling of top
plate. By comparison, the matrix compression damage predicted by Chang-Chang and Hashin criteria is more
severe than Puck criteria, P F LIU et al [21] got the similar conclusion.

4.2. [03/903]s laminate


[03/903]s laminate is employed for further comparison of three criteria. The composite material is AS4/PEEK,
and the material parameters are listed in table 5. The geometry size of the rectangular specimen is
65 mm×87.5 mm×2 mm, which is put on a supporting plate with 45 mm×67.5 mm rectangular opening.
The mass and diameter of the impactor are 2.28 kg and 12.5 mm, and three impact energies: 2 J, 5 J, and 9.2 J are
used. The time period is set as 7 ms according to the experiment carried on by F Aymerich and C Pani [34].
Figures 14–16 shows the impact force-time curves and impact force-displacement curves using three failure
criteria under the impact energy of 2 J, 5 J, and 9.2 J. The impactor reaches the lowest location at t=2.5 ms,
complete separation at t=5.5 ms, by comparison, Puck criteria lead to higher prediction accuracy than Chang-
Chang and Hashin criteria. Moreover, a phenomenon can be observed that a sudden drop occurs after the peak
impact force reaching the maximum value under the impact energy of 9.2 J, and fiber damage appears at the
same time, which Puck criteria succeeds to capture. Figure 17 shows the peak impact force and absorbed energy

15
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

at different energy (2 J, 5 J and 9.2 J), under the impact energy of 9.2 J, Puck criteria lead to minimum error, up to
30% though.
It can be observed that severe delamination damage occurs on the 90°/0° interface with the characteristic of
two lobed shape and evolves along the fiber direction of lower layer by using three failure criteria, and in
accordance with the experimental results by F Aymerich and F Dore [35]. Figure 18 shows the delamination
damage of 90°/0° interface under the energy of 5 J, Puck criteria denotes the most accurate prediction result.
From the simulation of two specimens above, Puck criteria which have been ranked very highly in WWF
[10, 11], leads more accurate results than the other two failure criteria compared with the experimental data
obtained by D Feng and F Aymerich [17], F Aymerich and C Pani [34] and F Aymerich and F Dore [35].
However, Puck criteria also have some disadvantages. (1) A new fracture angle search algorithm, AAGSS, have
been proposed in this paper to reduce the computational cost and a certain improvement has achieved.
However, it is still a serious problem for Puck criteria. (2) Many parameters are difficult to obtain and
approximate values can only be used, such as the fracture strength and inclination factors. (3) Compared with
the other criteria, Puck criteria have more complicated expression. Although the results predicted by Chang-
Chang and Hashin criteria are worse than Puck criteria, the errors are still acceptable, and they own the simple
expression and easier obtained material parameters at the same time, this is why the two criteria still get the
interest of many researchers.

5. Conclusions

In this paper, the effects of three respective failure criteria on the responses of composite laminates subject to
low-velocity impact are compared by operating FEA on [03/45/−45]s laminate and [03/903]s laminate. Damage
models with Chang-Chang, Hashin and Puck criteria and damage evolution laws based on equivalent strain are
implemented by ABAQUS-VUMAT, delamination is simulated by the bilinear cohesive model based on
quadratic criteria. Besides, a new fracture angle search algorithm (AAGSS) is developed to reduce the
computational cost of Puck criteria. From the simulation on the two different composite laminates, four main
conclusions can be attained:

(1) The numerical analysis results obtained by using three failure criteria have good agreement with
experimental results [17, 34, 35]. Puck criteria owns the minimum errors, however, consumes more time
on computation. Puck take the fracture plane and restraining effect of transverse compression on shear
failure into consideration, which may account for it.
(2) The matrix damage and delamination are the main damage forms when the composite laminates suffer the
low-velocity impact. Matrix tension initiates on the bottom layer and evolves to the top layer, with the
characteristic of groundnut shape elongated along the fiber direction, in the contrast, matrix compression
evolves from bottom to top. In addition, the matrix compression predicted by Chang-Chang and Hashin is
more severe than Puck criteria. Delamination grows with the characteristic of two lobed shape evolving
along the fiber direction of lower layer.
(3) The peak force and absorbed energy promote rapidly with the increase of impact energy. Moreover, when
the impact energy reaches a certain value, the fiber tension may occur, which is the main reason accounting
for the sudden drop in the bearing capacity of composite laminate. The peak force and absorbed energy
promote rapidly with the increase of impact energy. However, only Puck failure criteria capture this
phenomenon successfully.
(4) The AAGSS proposed for Puck criteria improve the computational efficiency significantly. The
computational cost raises with the increase of damage elements, and drops when the impactor begins to
rebound. The Puck criteria with AAGSS leads a consistent computational cost with Chang-Chang criteria at
the stage of rebound.

Obviously Puck criteria owns the most accurate prediction results, and the problem of computational cost
for Puck criteria have been improved at a certain degree with the application of AAGSS. Although difficult
obtained material parameters and complicate expressions may also restrict its application, there is no doubt that
Puck criteria is the first choice for response prediction of composite laminates subjected to low-velocity impact.
Moreover, Chang-Chang and Hashin criteria could still be suitable choice for researchers because of the
acceptable simulation results and simple expressions.

16
Mater. Res. Express 5 (2018) 065320 Z W Wang et al

Acknowledgments

The authors wish to acknowledge the financial support of this research by the National Key Research and
Development Program of China (2017YFC0805601& 2017YFC0805605).

ORCID iDs

J P Zhao https://orcid.org/0000-0002-7712-1412

References
[1] Abrate S 1998 Impact on Composite Structures (Cambridge: Cambridge University Press)
[2] Richardson M O W and Wisheart M J 1996 Review of low-velocity impact properties of composite materials Composites Part A: Applied
Science & Manufacturing 27 1123–31
[3] Tsai S W and Wu E M 1971 A general theory of strength for anisotropic materials J. Compos. Mater. 5 58–80
[4] Hill R 1948 A theory of yielding and plastic flow of anisotropic materials Proc. R Soc. Lond. Ser. A, Math Phys. Sci. 193 281–97
[5] Hill R 1998 The Mathematical Theory of Plasticity new ed (Oxford: Clarendon)
[6] Chang F K and Chang K Y 1987 A progressive damage model for laminated composites containing stress concentrations J. Compos.
Mater. 21 834–55
[7] Hashin Z and Rotem A 1973 A fatigue failure criterion for fiber reinforced materials J. Compos. Mater. 7 448–64
[8] Hashin Z 1980 Failure criteria for unidirectional fiber composites Journal of Applied Mechanics 47 329–34
[9] Puck A and Schürmann H 2002 Failure analysis of FRP laminates by means of physically based phenomenological models Composites
Science & Technology 62 1633–62
[10] Hinton M J and Soden P D 1998 Predicting failure in composite laminates: the background to the exercise Composites Science &
Technology 58 1001–10
[11] Soden P D, Hinton M J and Kaddour A S 1998 A comparison of the predictive capabilities of current failure theories for composite
laminates Composites Science & Technology 58 1225–54
[12] Pinho S T 2005 Failure models and criteria for FRP under in-plane or three-dimensional stress states including shear non-linearity
Nasa Technical Memorandum NASA/TM-2005-213530 NASA
[13] Choi H Y, Wu H Y T and Chang F K 1991 A new approach toward understanding damage mechanisms and mechanics of laminated
composites due to low-velocity impact: part II—analysis J. Compos. Mater. 25 1012–38
[14] Meo M et al 2003 Numerical simulations of low-velocity impact on an aircraft sandwich panel Compos. Struct. 62 353–60
[15] Batra R C, Gopinath G and Zheng J Q 2012 Damage and failure in low energy impact of fiber-reinforced polymeric composite
laminates Compos. Struct. 94 540–7
[16] Shi Y, Swait T and Soutis C 2012 Modelling damage evolution in composite laminates subjected to low velocity impact Compos. Struct.
94 2902–13
[17] Feng D and Aymerich F 2014 Finite element modelling of damage induced by low-velocity impact on composite laminates Compos.
Struct. 108 161–71
[18] Liao B B and Liu P F 2016 Finite element analysis of dynamic progressive failure of plastic composite laminates under low velocity
impact Compos. Struct. 159 567–78
[19] Li N, Chen P H and Ye Q 2017 A damage mechanics model for low-velocity impact damage analysis of composite laminates
Aeronautical Journal -New Series 121 1–18
[20] Kober M and Kühhorn A 2012 Comparison of different failure criteria for fiber-reinforced plastics in terms of fracture curves for
arbitrary stress combinations Composites Science & Technology 72 1941–51
[21] Liu P F et al 2016 Finite element analysis of dynamic progressive failure of carbon fiber composite laminates under low velocity impact
Compos. Struct. 149 408–22
[22] Wiegand J, Petrinic N and Elliott B 2008 An algorithm for determination of the fracture angle for the three-dimensional Puck matrix
failure criterion for UD composites Composites Science & Technology 68 2511–7
[23] Schirmaier F J et al 2014 A new efficient and reliable algorithm to determine the fracture angle for Puck’s 3D matrix failure criterion for
UD composites Composites Science & Technology 100 19–25
[24] Thomson D M et al 2017 Experimental and numerical study of strain-rate effects on the IFF fracture angle using a new efficient
implementation of Puck’s criterion Compos. Struct. 181
[25] Camanho P P, Davila C G and De Moura M F 2003 Numerical simulation of mixed-mode progressive delamination in composite
materials J. Compos. Mater. 37 1415–38
[26] Guo W, Xue P and Yang J 2013 Nonlinear progressive damage model for composite laminates used for low-velocity impact Applied
Mathematics and Mechanics 34 1145–54
[27] Pederson J 2008 Finite element analysis of carbon fiber composite ripping using ABAQUS All Thesis 512 Clemson University, USA
[28] Yu G C et al 2015 Low velocity impact of carbon fiber aluminum laminates Compos. Struct. 119 757–66
[29] Riccio A et al 2017 Intra-laminar progressive failure analysis of composite laminates with a large notch damage Engineering Failure
Analysis 73 97–112
[30] Li N and Chen P 2017 Failure prediction of T-stiffened composite panels subjected to compression after edge impact Compos. Struct. 162 210–26
[31] Benzeggagh M L and Kenane M 1996 Measurement of mixed-mode delamination fracture toughness of unidirectional glass/epoxy
composites with mixed-mode bending apparatus Composites Science & Technology 56 439–49
[32] Faggiani A and Falzon B G 2010 Predicting low-velocity impact damage on a stiffened composite panel Composites Part A: Applied
Science and Manufacturing 41 737–49
[33] Kim E H et al 2013 Composite damage model based on continuum damage mechanics and low velocity impact analysis of composite
plates Compos. Struct. 95 123–34
[34] Aymerich F, Pani C and Priolo P 2007 Damage response of stitched cross-ply laminates under impact loadings Eng. Fract. Mech. 74 500–14
[35] Aymerich F, Dore F and Priolo P 2008 Prediction of impact-induced delamination in cross-ply composite laminates using cohesive
interface elements Composites Science & Technology 68 2383–90

17

Potrebbero piacerti anche