Home Physical Sciences A variational approach for predicting initiation of matrix cracking and induced delamination in symmetric composite laminates under in-plane loading
Article Open Access

A variational approach for predicting initiation of matrix cracking and induced delamination in symmetric composite laminates under in-plane loading

  • Seyed Siamak Taghavi Larijani and Amin Farrokhabadi EMAIL logo
Published/Copyright: October 17, 2017

Abstract

This paper discusses the formation of matrix cracking and induced delaminations in a [ϕm(2)/ψn(1)]s laminate subject to arbitrary in-plane loading using a variational approach. To this end, the effects of delaminations coming from the tips of transverse cracks in the middle sublaminates are considered to specify the critical crack density points and finally understand which modes of damage are prospects for various lay-ups in the presence of transverse cracks as well as induced delamination. After deriving a very good approximation of the principle of minimum complementary energy by considering complex equations, the stress fields are provided. In this analysis, a unit cell in the ply level of a composite laminate containing both matrix cracking and delamination is considered. Then, the values of the admissible stresses and compliance of cracked laminate are employed to evaluate the energy release rate of each mentioned damage mode. Eventually, the graphs for different lay-ups in such symmetric laminate are drawn, which indicate important points for designers in practical applications. Afterwards, the effects of thickness on the dislocation of critical crack density points are checked. It can be emphasized that the current approach opens a new insight for perceiving the composites’ structural behavior in vulnerable positions.

1 Introduction

One of the most essential discussions in composites is their innate ability to withstand damage modes. Over the past four decades, a particular focus of discussion for failure composites has been served. There are always several possible damages in laminated composites, e.g. matrix cracking [1], delamination [2], etc. Due to the utilization of composites in diverse industries, their effective parameters, properties, and relations with failing factors should be known precisely. For this purpose, researchers have tried to conduct studies about the major damage modes that are occurring in a laminate during the process of loading.

Matrix cracking and induced delamination are the most significant damage modes that often occur in composite laminates under general in-plane loading. The failure processes of composite laminates under general in-plane loading have been permanently seen in diverse situations like matrix cracking and induced delamination. For predicting matrix cracking, distinct models were suggested by researchers [3], [4], [5], [6], [7]. All of these models are based on limitations that decrease the accuracy of their calculations. Furthermore, these models are not able to estimate initiation or propagation of damages without considering continuum mechanics thermodynamic properties very well, unless using some assumptions such as changing scales in a unit cell [8], [9].

In the unit cell models, the first challenge for many researchers was the estimation of stiffness reduction. Indeed, they have sought to know how matrix cracking in a laminate could decrease the elasticity properties. For this problem, all of the investigations can be divided into three parts: variational approach [10], [11], [12], [13], shear-lag method [14], [15], [16], and stress transfer method [8], [17], [18], [19]. Even though diverse semianalytical or numerical methods can be found in some papers [20], [21], this article emphasizes the analytical methods. As a result, depending on the conditions of the problem, for instance, the lay-up or type of loading, the solutions should be suggested.

Hashin [10], [22], [23] employed a variational approach to offer an approximate stress analysis for cross-ply laminates having matrix cracking in the 90° plies. This method was complemented for the angle-ply lay-ups under general in-plane loading by Vinogradov and Hashin [24]. According to the results presented by Zhang and Minnetyan [25], it can be said that stress-based variational analysis has more rational results in relation to displacement-based variational analysis. In the latest presented studies, the stiffness reductions have been calculated for arbitrary lay-ups [11], [12]. In fact, it was attempted to conduct and expand the projects for outer-ply matrix cracking in cross-ply laminates [26].

Generally, by reviewing stress transfer and shear-lag models, some significant weaknesses can be cited, which should be considered while investigating diverse properties in a laminate. Firstly, the effects of out-plane stress are not considered; however, this stress plays a special role in the stress transfer. Aside from this, the special lay-ups and loading conditions confine results to cross-ply lay-ups or some off-axis laminates. These objections have been partially decreased in the developed model of shear-lag (McCartney method); however, complex calculations [27] persuade authors not to utilize this method. On the other hand, if special attention is paid to the results of Vinogradov and Hashin [24], it can be said that the process of investigating a laminate is also declined, so that it is decided to employ a variational approach for calculating stiffness analysis.

Due to the importance of the induced delamination mode in the final failure load, some researchers tried to suggest several related analytical or numerical models [28], [29], [30]. Nairn and Hu [31] used the variational approach to predict the initiation and growth of induced delamination by matrix cracking. As a matter of fact, the compliance of cracked laminate in cross-ply lay-ups under specific loading was used to compute the energy release rate for formation matrix cracking in the middle sublaminate and induced delamination; however, it can be seen that in the paper by Hosseini-Toudeshky et al. [8], the energy release rate was based on the stress field of cracked laminate. In other studies [32], [33], Li et al. researched the prediction of continuous matrix cracking; however, their calculations had totally the same boundary and continuity conditions that had been considered by Nairn and Hu [31], and there is no research for distinct stacking sequences.

By considering the mentioned problem, the present study provides a generalized model that is based on the unit cell method. Using Euler-Lagrange equations in the variational approach, it is attempted to evaluate the energy release rate of induced delamination as well as matrix cracking formation simultaneously for angle-ply laminates. In fact, this paper is a complementary research to Nairn and Hu’s study [31], in perspective. Nairn and Hu tried to compute the critical crack densities by using a dimensionless formulation according to the energy release rates only for cross-ply laminates; however, for angle-ply laminates, this method is impossible because of complex Euler-Lagrange equations that are supposed to be simplified. In this regard, the methods that they employed are no longer unique for this problem. Therefore, the authors attempted to simplify the approach with respect to its principles. In the following, changes that may occur to the critical size of unit cells in the laminates, according to their stacking sequences when there is an equal stress state, are revealed. Using the boundary and continuity conditions, the stress state of the cracked laminate is assessed by considering the fact that these conditions could be arbitrary in regards to the kind of loading; for example, it can be full in-plane stress or only tensile stress in X direction. In the next phase, the effects of thickness in the critical crack density dislocation are estimated. Displacement of this point depends on the size of unit cell investigated. This technique provides a precise lower bound for the stiffness matrix, so that the results of its energy release rates are the most reliable amounts for practical application.

2 Variational approach under in-plane loading

2.1 Perturbation stress field analysis

Consider an angle-ply [ϕm(2)/ψn(1)]s composite laminate that is cracked under tensional static loading. All of the cracks have equivalent distances between each other (Figure 1).

Figure 1: Geometry of a symmetric laminate containing cracks in the middle layers in a global coordinate system.
Figure 1:

Geometry of a symmetric laminate containing cracks in the middle layers in a global coordinate system.

A list of symbols and parameters is compiled in Table 1, as defined in the text and the following formulations.

Table 1:

A list of parameters and their definitons in the text.

SymbolsDefinitionSymbolsDefinition
ϕm(2)The angle of layer (2)η(x)An unknown function of x
ψn(1)The angle of layer (1)2aThe distance between the crack surfaces in a unit cell
FXXX-direction in-plane load (primary coordinate)CThe local compliance matrix
FXYIn-plane load in the XY plane (primary coordinate)UCComplementary energy due to perturbation stresses
FYYY-direction in-plane load (primary coordinate)UuComplementary energy of the uncracked body
Fxxx-Direction in-plane load (secondary coordinate)C*The effective compliance matrix of the cracked body
FxyIn-plane load in the xy plane (secondary coordinate)W(i)The strain energy density due to perturbation stress
Fyyy-Direction in-plane load (secondary coordinate)εmnp(i)The strain value of ply (i)
t2The thickness of layer (2)C̅(i)Compliance of the matrix unidirectional
t1The thickness of layer (1)ζThe ratio of x to t1
λDimensionless of quantity for thicknessρThe ratio of a to t1
hThe summation of thicknessesαThe real part of the complex root
θ(x)An unknown function of xβThe imaginary part of the complex root
ϑ(x)An unknown function of xS(ρ)The effective compliance matrix of the cracked unit cell
AtcThe created damage area in micro crackingUcbeThe strain energy before micro cracking
UcafThe strain energy after matrix crackingwThe width of laminate
d2The length of delaminated area in layer (2) – part 2d1The length of delaminated area in layer (2) – part 1
S2The compliance of outer lamina in delaminated compositeUA&CThe strain energy after delamination for outer laminas
δDimensionless number for length of delaminated areaVρδThe volume of intact region
ADThe delaminated areaωThe real and the positive root
UBThe strain energy for the intact regionGdThe energy release rate of the delaminated unit cell
GmThe energy release rate due to matrix cracking formationEAAxial Young’s modulus in fiber direction
cThe crack densityGTTransverse shear modulus
ETTransverse Young’s modulusνAAssociated axial Poisson’s ratio
νTAssociated transverse Poisson’s ratioσxx00The applied stress in the X direction
GAAxial shear modulusσyy00The applied stress in the Y direction
σxy00The applied stress in the XY planeσ1The local stress in the 1 direction
σ2The local stress in the 2 directionσ3The local stress in the 3 direction

The maximum number of layers in this analysis is four sublaminates; however, researchers are able to investigate for arbitrary lay-ups by handling their calculations following Ref. [11]. In addition, the thicknesses of these laminas can be considered arbitrary. As shown in many papers [11], [12], [24], these calculations are the same in estimating basic science; however, generally, their boundary conditions are different. The first step of calculations is assessing the stiffness degradation. For this purpose, it is needed to transfer the unit cell’s coordinate to the new one. In fact, if it is assumed to continue calculations based on the project just like Hashin did in 1985 [10], changes should be made in its primary strategy. In order to have 90° plies in the middle sublaminates, the rotation around Z axis should be [ψn(1)+90]. Indeed, this method carries out the data to the previous calculations with cross-ply sublaminates. New lay-up for this composite can be displayed in the form of [S/90]s. The term S denotes [ϕm(2)ψn(1)+90] in the general lay-up (Figure 2).

Figure 2: Cracked laminate in the rotated coordinate.
Figure 2:

Cracked laminate in the rotated coordinate.

As illustrated, under in-plane loading, which can be arbitrary, these cracks are constructed. After the rotation, it is expected that the type of forces and their amounts would be altered. Lowercase letters represent new coordinate characteristics. Next, it is shown that in some rotations depending on their angles and their quantities, occasionally they would be reversed into compressive loading. By these interpretations, we bring forward external loads for the primary coordinate as FXX, FXY, FYY and for the secondary coordinate in Fxx, Fxy, Fyy. It is definitely seen that there are always some papers that have used some different types of loadings and lay-ups [34] (which work on stiffness reduction of off-axis laminates; neither induced delamination nor matrix cracking analysis for angle-ply laminates); however, in this research, it is tried to assess general loadings and stacking sequence.

Now, it is the time to explicate the composition of stress field in a cracked laminate. In a fragment, it can be expressed as an admissible stress field by a coincidence of the stresses in the uncracked material and some special unknown perturbation stresses that are due to the presence of cracks.

(1)σ˜mn(i)=σmnu(i)+σmnp(i),

where i=1, 2 specifies the number of laminas, and u and p denote uncracked and perturbation, respectively. Proportionate with what is shown in Figure 1, h=t1+t2 and a dimensionless ratio λ=t2t1 are defined. By virtue of symmetry, it is tried to analyze only half of the laminate. In this model, as is observable in Figure 1, there are several boundary and continuity conditions that can help arrive at a solution for the perturbation stresses. The following conditions have been used in calculations:

  1. There are no stresses on the out-plane surfaces (at z=∓h, σ˜mz(2)=0).

  2. Using continuity conditions between interfaces, it can be seen that at zt1, σ˜mz(1)=σ˜mz(2).

  3. Free tractions on the crack surfaces denote that (x=±a,σ˜mx(1)=0).

  4. The behavior of an uncracked laminate with a cracked laminate just like the work done by Vinogradov and Hashin [24] can be simulated; thus, the effects of local damages can be ignored.

Using the equilibrium equations and also balancing the membrane forces, some perturbation stresses that have impacts upon the stiffness degradation are achieved. Indeed, these amounts directly influence stiffness reduction. According to the steps that have been derived by Vinogradov and Hashin [24], the following can be written.

For lamina 1 (the middle lamina):

(2)σxxp(1)=θ(x),σxyp(1)=ϑ(x),σyyp(1)=η(x),σxzp(1)=θ(x)z,σyzp(1)=ϑ(x)z,σzzp(1)=θ(x)(ht1z2)2.

For lamina 2 (the outer ply):

(3)σxxp(2)=1λθ(x),σxyp(2)=1λϑ(x),σyyp(2)=1λη(x),σxzp(2)=1λθ(x)(hz),σyzp(2)=1λϑ(x)(hz),σzzp(2)=1λθ(x)(hz)22,

where all of these stresses are the perturbation stresses in two sublaminates. The applied boundary conditions are denoted by

(4)θ(±a)=σxxu1,θ(±a)=0,ϑ(±a)=σxvu1.

Note that these implicit results of boundary conditions will be used in calculating the final answers of ordinary differential equations of (θ, ϑ, η), which are derived in the next step.

2.2 Variational formulation

Now, the principle of minimum complementary energy is going to be used for obtaining the perturbation stresses. For this purpose, the strain energy release rate must be calculated for transverse cracked laminates. It is very important to know about how the strain energy has to be derived in a model. For a loaded unit cell, the complementary energy functional can be defined as

(5)U˜Cr=12σ˜TCσ˜dV.

As it was mentioned in some papers like variational principles for generalized plane strain problems and their applications [32], the correct shape of the energy criterion should not be used in the form of strain energy; however, given that there is a linear elastic assumption, this equivalent is acceptable. Therefore, all equations that are considered have to be under this assumption. In the presence of cracks, for all elastic materials, Eq. (5) can be expressed in the following form:

(6)U˜C=Uu+UC,

where Uu and UC are the strain energies for an uncracked body and due to perturbation stresses, respectively. As mentioned in Hashin’s papers [10], [24], [35], the value of the complementary energy using admissible stresses can be derived from the principle of minimum complementary energy or, in another words, it can be written as

(7)U˜C=12σ¯TCσ¯VU˜Cr=12σ¯TCuσ¯V+12σpTCσpdV,

where the terms C and Cu are the effective compliance matrix of cracked and uncracked bodies, respectively; V is the volume of the cracked laminate; σ¯ specifies the average in-plane stresses equal to applied stresses; and σp denotes the perturbation stresses that are constructed by the presence of cracks. Actually, it is needed to minimize the right-hand side of Eq. (7) in order to derive the closest amount of the strain energy to a practical value. For this purpose, it has been suggested to minimize an integration that is produced by multiplying the perturbation energy density by other quantities, which are described in equations.

Until here, it has been attempted to put forward a model that is restricted by two cracks in the volume of a fragment with 2a length and |z|≤h. Due to the factors in Figure 1 and conditions that have been explained before, for half of a unit cell, it can be written that

(8)W(i)=12(σxxp(i)εxxp(i)+2σxyp(i)εxyp(i)+2σyzp(i)εyzp(i)+2σxzp(i)εxzp(i)+σyyp(i)εyyp(i)+σzzp(i)εzzp(i)),

where W(i) and εmnp(i) are the strain energy density and strain value of the ply (i). It is worth to mention that the strain energy density is a scalar function. The sum of integration of all sublaminates over the given volume provides the total strain energy of the body. To calculate the strain amount, or in another words (εmnp(i)=C¯(i)σmnp(i)), we have to compute the compliance of the matrix unidirectional composite. For this purpose, as it was suggested by Vinogradov and Hashin [24], it is better to rotate the coordinate as was mentioned in Section 2.1. Embedding the expressions of perturbation stresses [Eqs. (2) and (3)] and stress energy densities in the perturbation strain energy, an integral calculus with constant coefficients is derived that assists researchers in computing the value of this expression to totalize with the strain energy of the uncracked body. The general terms of this integration is expressed here:

(9)UC=aaF(x,θ,θ,θ,ϑ,ϑ,η)dx.

For solving this integration from the Euler-Lagrange method, it must be converted to dimensionless form with these changes (ζ=xt1,ρ=at1). The above equation can be illustrated as

(10)UC=t12ρρF(ζ,θ,θ,θ,ϑ,ϑ,η)dζ,

where

(11)F(ζ,θ, θ, ϑ,ϑ, η)=M1aθ2(ζ)+M2aϑ2(ζ)+M3aη2(ζ)+M12aθ(ζ)ϑ(ζ)+M13aθ(ζ)η(ζ)+M23aϑ(ζ)η(ζ)+M1b(θ(ζ))2+M2b(ϑ(ζ))2+M12bθ(ζ)ϑ(ζ)+M1cθ(ζ)θ(ζ)+M2cθ(ζ)ϑ(ζ)+M3cθ(ζ)η(ζ)+M1d(θ(ζ))2,

and the multipliers are the known amounts that are detailed in the Appendix. To compute the strain energy due to perturbation stresses, the unknown functions [θ(ζ), ϑ(ζ), η(ζ)] must be calculated from minimizing this integration. Because of three unknown functions, three Euler-Lagrange equations are needed, which are obtained from the below relations:

(12)d2dζ2(Fθ)ddζ(Fθ)+Fθ=0.
(13)d2dζ2(Fϑ)ddζ(Fϑ)+Fϑ=0.
(14)Fη=0.

Substituting the result of Eq. (14) into Eqs. (12) and (13), a new one is derived that is only a function of two variables.

(15)η=(M13a2M3aθ+M3c2M3aθ)M23a2M3a,

which helps us rewrite Eq. (11) as

(16)UC=t12ρρH(ζ,θ,θ,θ,ϑ,ϑ)dζ.

As shown in the above equation, one of the unknown functions is omitted easily (η). The other independent unknown functions (θ, ϑ) will be determined in the coupled linear differential equations when Euler-Lagrange equations are implemented for two times. Thus, it can be written as

(17)T1aθ+T2aθ+T3aθ+T4aϑ+T5aϑ=0,
(18)T4aθ+T5aθ+T6aϑ+T7aϑ=0,

where Tia are the constant coefficients that can be obtained according to the formulations based on the previous coefficients before implementing the variational approach. Depending on the boundary conditions, this coupled linear differential equation can be converted to the model that is suggested by Refs. [10], [23], [24].

2.3 Compliance of cracked unit cell

In order to solve coupled linear differential equations, there are several solutions that are suggested by commercial mathematical software. The authors prefer to utilize a more rational way, which is an analytical solution. Indeed, from the standpoint of mathematics, it is not logical that numerical methods are used instead of analytical ways when both solutions are available.

For these grounds, let us consider a matrix that is deduced from the coupled linear differential Eqs. (17) and (18):

(19)L=[T1ar4+T2ar2T4ar2+T5aT4ar2+T5aT6ar2+T7a].

As it is clear to have a specified solution for Eqs. (17) and (18), the determinate of the above relation [Eq. (19)] should be zero, and thus it can be written that

(20)|L|=0L1r6+L2r4+L3r2+L4=0,

where

(21)L1=T1aT6a,L2=T1aT7a+T2aT6a(T4a)2,
(22)L3=T2aT7a+T3aT6a2T4aT5a,L4=T3aT7a(T5a)2.

Due to the simplification of the equations, it is assumed that k=r2. Depending on the boundary conditions, it can be observed that there might be two complex roots and one real positive root.

(23)k1=ω,k2,3=α±iβ.

This is not the right place to describe more points about the process of solution. With these interpretations, it is better to refer to differential equation books like that of Simmons [36]. It has indicators to solving these kinds of problems.

The general model of a system with these roots can be noted in the form of

(24)θ(ζ)=μ1C1cosh(sζ)+C2cosh(pζ)cos(qζ)+C3sinh(pζ)sin(qζ).
(25)ϑ(ζ)=C1cosh(sζ)Δ1cosh(pζ)cos(qζ)+Δ2sinh(pζ)sin(qζ).

It is essential to note that this problem has symmetry of about (x=0). Here, multipliers of θ(ζ) and ϑ(ζ) have relations with each other. Given Eqs. (21) and (22) in calculations lead authors to express that

(26)s=ω,p,q=22α2+β2±α.
(27)Δ1=(C2μ2+C3μ3),Δ2=(C3μ2C2μ3).
(28)μ1=T5a+T4aωT3a+T2aω+T1aω2.
(29)μ2=(T5a+T4aα)(T7a+T6aα)+T4aβ2(T7a+T6aα)2+(T6a)2β2.
(30)μ3=(T4aT7aT5aT6a)β(T7a+T6aα)2+(T6a)2β2.

The insertion of the expression in Eq. (4) helps us obtain the particular solutions of Eqs. (24) and (25). At first, it is needed to compute the constants C(i) and then there are precise answers for unknown functions (θ, ϑ). This is a uniform solution with a less detailed description for calculating the stiffness degradation in comparison with Vinogradov and Hashin’s work [24]. Briefly using these functions, an equation can be obtained for compliance of the cracked body, which helps authors to evaluate a true value for strain energy in a cracked body. The expression of compliance for a cracked body can be obtained as

(31)CC0+2UCσ¯TVσ¯.

According to the obtained equation, it is noted that in the variational standpoint, the minimum amount of the perturbation strain energy in the unit cell has the principal impact on the effective compliance matrix (C). Now, in this model, which is able to estimate the stiffness of a cracked body, it can be said that these calculations would not depend on the angles of sublaminates and it can be arbitrary.

3 Energy release rate due to matrix cracking formation

The first form of damage in laminates is often matrix cracking. By referring to performed studies [6], [8], [31] and also according to the experimental results in cross-ply laminates, which are also mentioned in Ref. [31], the next matrix cracking is created in the middle layer. Among the performed studies, various methods of calculating the strain energy release rates have been suggested, which have infrastructural discrepancies. Indeed, as authors peruse these different methods carefully, they have been divided as (i) stress base [8] and (ii) stiffness base [31]. All of these methods have special advantages in their suitable positions. In this paper, it is tried to compute the strain energy release rate from the second technique. For the present state of matrix cracking, the strain energy can (before matrix cracking) be obtained from

(32)Ucbe=12σ¯TS(ρ)σ¯V.

The strain energy after matrix cracking should be obtained by

(33)Uca=14σ¯TS(ρ2)σ¯V.

It is important to mention that the parameter Atc is the created damage area and it must be evaluated from

(34)Atc=2t1w,

and the energy release rate for this damage (Gm) is finally obtained from

(35)Gm=(2UcafUcbe)Atc.

As shown in Figure 3, the initiation of matrix cracking in a unit cell is depicted. Indeed, two modes of laminate are investigated in these procedures, which should be divided after matrix cracking and before matrix cracking. It is conspicuous that after matrix cracking, the unit is separated into two segments. As a result, Eq. (33) defines an energy release rate after matrix cracking just for one segment.

Figure 3: Unit cell after matrix cracking.
Figure 3:

Unit cell after matrix cracking.

This systematic approach is often used to detect an energy release rate based on the initial distance of cracks in a unit cell. In the next step, the other mode of damage is going to be investigated.

4 Energy release rate due to induced delamination

By initiation of delamination from tips of cracks, three regions are constructed as demonstrated in Figure 4. These regions are divided into intact and delaminated regions, respectively. In the intact region, the strain energy is calculated for the laminate containing transverse cracks with length of 2(ρδ)t1. The delaminated region is also divided into two parts, which can be illustrated for outer plies and the dead zone. It is only needed to compute the strain energy for outer laminas in a special volume, which is expressed below:

Figure 4: Different region of the delaminated unit cell.
Figure 4:

Different region of the delaminated unit cell.

(36)UA&C=w(aa+d1dxt1hdzσ˜TS2σ˜+ad2adxt1hdzσ˜TS2σ˜),

where UA&C is the strain energy after delaminating of laminates for outer plies. In other words, integrating the strain energy density over a damaged area in an outer ply for sample of width w gives the total strain energy in the damaged area. It is worth to mention that the middle sublaminate is the dead zone. Indeed, the stresses of other regions are zero and the middle sections can be removed from the sample with no effect. σ˜ij and S2 are the admissible stress system and the compliance of the outer lamina, respectively. The stresses can be expressed as

(37)σ˜=[σ˜xxσ˜yyσ˜xy],

and the in-plane stresses can be obtained from

(38)σ˜xx=σxxu1ϕ(x)=(1+λ)λσxx00.
(39)σ˜yy=σyyu1ψ(x)=(1+λ)λσyy00.
(40)σ˜xy=σxyu1η(x)=(1+λ)λσxy00.

The boundary conditions in the outer plies require ϕ′(x)=0, ψ′(x)=0, η′(x)=0. Integrating these equations implies that the integrated functions are constants.

The strain energy release rate for the intact region should be obtained from

(41)UB=12σ¯TS(ρδ)σ¯Vρδ.

Note that Eq. (41) defines the strain energy for the intact region. For this purpose, it is worth to mention that when δ=0, the initiation of delamination is calculable and this value depends only on the size of unit cells. δ has to be obtained from

(42)δ=d1+d22t1,

where σ˜ij and S2 are the admissible stress system and the compliance of the outer lamina, respectively, and Vρ−δ =2(ρδ)t1w is the volume of the intact region. Therefore, the energy release rate of the delaminated unit cell can be noticed as

(43)Gd=UB+UA&CUcbeAD,

where AD is the delaminated area. All of these calculations are based on a unit cell that is demonstrated in Figure 4.

5 Results and discussion

In this section, the energy release rates for diverse laminates with different stacking sequences and different thicknesses are considered under general loading conditions. These stresses have to be converted into the local coordinates. At first, the results of these calculations are compared with Nairn and Hu’s results, and then it is tried to expand the results for variant lay-ups. Thereafter, the results of energy release rates are provided for distinct lay-ups.

5.1 Cross-ply lay-ups

Initially, it is better to introduce the same materials that had been used in Nairn and Hu’s results [31].

EA=128 GPa,GA=4 GPa,νA=0.3.ET=7.2 GPa,GT=2.4 GPA,νT=0.5,t1=t2=0.14 mm.

According to the utilized conditions by Ref. [33], the applied stresses are σxx00=100MPa,σxy00=0,σyy00=0. If the presented Eqs. (35) and (42) are thoroughly investigated, the obtained results have incredible agreements to Ref. [33] for a (02/904)s carbon fiber/epoxy as shown in Figure 5, even though this process is much easier than theirs because of the ignorance of simplified additional equations.

Figure 5: Comparison of the obtained energy release rate with Li’s results for [02/904]s carbon/epoxy laminate under σ1=100 MPa [33].
Figure 5:

Comparison of the obtained energy release rate with Li’s results for [02/904]s carbon/epoxy laminate under σ1=100 MPa [33].

Figure 5 illustrates the variation of energy release rates due to two forms of most possible damages with the increase of crack densities. It is shown that between 0 and 0.51 crack densities, the odds-on type of damage is formation of matrix cracking; however, from this critical point to the lowest distance between cracks, delamination must be observed. Generally, this critical point helps designers to know something more about the next possible damage. As shown in Figure 5, the size of the unit cell plays an essential role in the determination of the kind of damages. Moreover, this analogy proves the point that the new method of achieving energy release rate in the variational approach has the same precision that theirs have.

In the next attempts, the authors consider diverse cross-ply lay-ups of carbon fiber/epoxy, which are distinguished by their thicknesses. Likewise, the critical points are obtained easily, so that it is straightforward to estimate the effects of the number of cracked sublaminates in dislocation of critical points relative to crack densities. For this purpose, various types of fiber/epoxy cross-ply laminates are considered to report the influence of adding cracked laminas in Figures 68.

Figure 6: Crack density vs. energy release rate for [0/90]s carbon/epoxy laminate under σ1=100 MPa.
Figure 6:

Crack density vs. energy release rate for [0/90]s carbon/epoxy laminate under σ1=100 MPa.

Figure 7: Crack density vs. energy release rate for [0/902]s carbon/epoxy laminate under σ1=100 MPa.
Figure 7:

Crack density vs. energy release rate for [0/902]s carbon/epoxy laminate under σ1=100 MPa.

Figure 8: Crack density vs. energy release rate for [0/903]s carbon/epoxy laminate under σ1=100 MPa.
Figure 8:

Crack density vs. energy release rate for [0/903]s carbon/epoxy laminate under σ1=100 MPa.

In all of these calculations, the considered stresses are σxx00=100MPa,σxy00=0,σyy00=0; however, the authors try to draw some graphs that have non-zero σxy00 and σyy00. Nevertheless, Figures 911 are provided as samples with other remote loadings. In these paradigms, the stresses are σxx00=100MPa,σxy00=1MPa,σyy00=10MPa. As it is detectable in these cross-ply laminates, the increasing thicknesses of cracked sublaminates lead to regressive conditions, or, in another words, induced delamination is more prospective by declining the critical point relative to crack density.

Figure 9: Crack density vs. energy release rate for [0/90]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 9:

Crack density vs. energy release rate for [0/90]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

Figure 10: Crack density vs. energy release rate for [0/902]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 10:

Crack density vs. energy release rate for [0/902]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

Figure 11: Crack density vs. energy release rate for [0/903]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 11:

Crack density vs. energy release rate for [0/903]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

5.2 Angle-ply lay-ups

Diverse angle-ply laminates are chosen to compute their energy release rates when their stresses in local coordinates are the same as the precise amount that cross-ply laminates have. Indeed, the authors try to examine how differentiation of angles affects the energy release rate for induced delaminations and the next matrix cracking in a unit cell. According to these assessments, it is decided to provide Table 2, which introduces some essential factors about the stress and its angles for local coordinate.

Table 2:

Local and global loads of several laminates.

No.[ϕm(2)ψn(1)]s21]s (a)σ1 (MPa)σ2 (MPa)σ12 (MPa)σxx00 (MPa)σyy00 (MPa)σxy00 (MPa)
1[45/–45][180/90]1001015654−45
2[40/–40][170/90]10010148.1761.83−44.49
3[30/−30][150/90]10010133.36676.634−39.47

Relatively straightforward calculations specify the point that in order to achieve angle-ply energy release rates, the expressions have to be investigated according to their local coordinates as Hosseini-Toudeshky et al. [8] did in their micromechanical approach. The results are drawn in diverse line charts, which are shown here.

5.2.1 [45/−45n]s Angle-ply laminates

Figures 1214 show changes in the energy release rates of matrix cracking and induced delamination for [45/−45]s, [40/−452]s and [40/−453]s. There is a modest increase in the energy release rate for matrix cracking and then it plunges sharply, although this steep decrease is not the same as the decline that is seen in the induced delamination. The critical points for Figures 1214 are 0.81, 0.605 and 0.47, respectively. It should be underlined that the decreasing trend of critical points tends to be relative to the increasing number of thicknesses in the cracked laminas. In addition, according to the data shown, the amounts of energy release rate by increasing the number of cracked laminas are escalated.

Figure 12: Crack density vs. energy release rate for [45/−45]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 12:

Crack density vs. energy release rate for [45/−45]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

Figure 13: Crack density vs. energy release rate for [40/−452]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 13:

Crack density vs. energy release rate for [40/−452]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

Figure 14: Crack density vs. energy release rate for [45/−453]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 14:

Crack density vs. energy release rate for [45/−453]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

5.2.2 [40/−40n]s Angle-ply laminates

Figures 1517 depict how inclinations for the energy release rates have changed for matrix cracking and induced delamination as well as Figures 1214 illustrated. The most obvious changes, which are notable in comparison with the previous figures, are its measures for energy release rates. Indeed, whatever the stacking sequences of local coordinates ([S/90]s) is closer to [180/90]s, the energy release rates see a slower change and vice versa. It can be clearly seen that the energy release rate for the next matrix cracking in [40/−40]s starts with 20.77 J/m2; in contrast, this amount for [45/−45]s begins with 3.81 J/m2.

Figure 15: Crack density vs. energy release rate for [40/−40]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 15:

Crack density vs. energy release rate for [40/−40]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

Figure 16: Crack density vs. energy release rate for [40/−402]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 16:

Crack density vs. energy release rate for [40/−402]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

Figure 17: Crack density vs. energy release rate for [40/−403]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 17:

Crack density vs. energy release rate for [40/−403]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

Another distinguishing feature is their critical points. The variations of energy release rate versus crack density for matrix cracking and delamination intersect at 1.74 (1/mm), 0.725 (1/mm) and 0.413 (1/mm) in Figures 1618. The trend of decreasing energy release rates is virtually the same as the other stacking sequences. As can be observed, some of these graphs are totally outnumbered by increasing the number of sublaminates.

Figure 18: Crack density vs. energy release rate for [30/−30]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 18:

Crack density vs. energy release rate for [30/−30]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

5.2.3 [30/−30n]s Angle-ply laminates

In this section, the widest difference can be seen in the amounts of energy release rates that Figures 1820 demonstrate with the previous figures. As illustrated, the critical points are 1.96 (1/mm), 0.925 (1/mm) and 0.524 (1/mm), respectively, for Figures 1820.

Figure 19: Crack density vs. energy release rate for [30/−302]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 19:

Crack density vs. energy release rate for [30/−302]s carbon/epoxy laminate under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

Figure 20: Crack density vs. energy release rate for [30/−303]s carbon/epoxy laminate for under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.
Figure 20:

Crack density vs. energy release rate for [30/−303]s carbon/epoxy laminate for under σ1=100 MPa, σ12=1 MPa and σ2=10 MPa.

This opportunity has been granted to express the standpoint that local stresses and also angles, when the middle layers are 90°, are determinative. In other words, for initiation of delamination and intralaminar matrix cracking, the key point is that how the global stresses would change and how these amounts affect the energy release rates according to their crack densities.

As the Euler-Lagrange equations are derived, it can be said that variational approaches can be acceptable if the orientation of fibers would be restricted in two layers and the intralaminar cracks would occur in the middle layer.

In Figure 21, the changes in critical densities for carbon/epoxy [0/90n]s, [45/−45n]s, [40/−40n]s, [30/−30n]s and [15/−15n]s laminates are illustrated with respect to the number of middle cracked plies. The results demonstrate a rational trend according to the altering angles. In other words, whatever the number of cracked laminates are increased, the risk of delamination increases in proportion to the particular lay-up that is being investigated. According to previous studies [8], it can be proved, when the coordinates and the amounts of stresses are equal in local calculations, that critical crack density points have broadly similar amounts.

Figure 21: Number of middle cracked plies vs. critical crack densities for distinct carbon/epoxy laminates.
Figure 21:

Number of middle cracked plies vs. critical crack densities for distinct carbon/epoxy laminates.

Table 3 shows the critical energy release rates, critical crack densities, and energy release rates due to matrix cracking and induced delamination for a minute unit cell (c=0.01). Gc denotes the critical energy release rate, which represents the precise energy release rate for exceeding induced delamination in relation to matrix cracking as possible damages. Table 3 depicts that whenever stacking sequences in local coordinates get far from [0/90n]s, the energy release rates increase dramatically. Furthermore, the amounts of critical crack densities are growing rapidly. In this table, it can be seen that almost all critical crack density values are decreasing by increasing the damaged area. In other words, decreasing from 0.81 to 0.47 in [0/90n]s shows that the possibility for matrix cracking is shrank by tripling the cracked sublaminates. This trend seems set to continue for other stacking sequences.

Table 3:

The critical energy release rates, critical crack densities and energy release rates due to induced delamination and matrix cracking for a unit cell with c=0.01 (1/mm) under σ1=100 MPa, σ12=1 MPa, σ2=10 MPa.

No.Lay-upCritical crack density (1/mm)Gc (J/m2)Gm/c=0.01 (1/mm)Gd/c=0.01 (1/mm)
1[0/90]s0.813.793.8133.785
2[0/902]s0.6111.0613.63611.032
3[0/903]s0.4723.4832.02823.477
4[45/−45]s0.813.793.8133.785
5[40/−452]s0.6111.0613.63611.032
6[45/−453]s0.4723.4832.02823.477
7[40/−40]s1.7313.08820.7713.59
8[40/−402]s0.7341.1949.38241.133
9[40/−403]s0.41184.1990.6684.076
10[30/−30]s1.97567.222140.50874.702
11[30/−302]s0.923222.014305.058224.559
12[30/−303]s0.529451.260512.55450.651

6 Conclusion

A variational technique for investigating initiation of matrix cracking and also delamination under general symmetric in-plane loading was provided. In the first step, a unit cell under general remote stress loading was considered, and then its stiffness reduction according to crack density was derived. In the following step, high standards were set for using general equations of energy release rates and then their values have been compared with the previous results.

For the first time, it was attempted to analyze the critical crack density of an angle-ply laminate. Heretofore, the energy release rates for cross-ply laminates had been calculated by some complex formulation for some special cases. However, this model covers all the laminates with different stacking sequences. The current variational approach is two-dimensional and does not explain the interlaminar normal stresses induced by a free edge. Therefore, it cannot be a good idea for growth of delamination.

Generally, when the sublaminate (S) (for a laminate with stacking sequence like [S/90n]s) is adequately stiff and n is not too small, the first phase of damage will be microcracking in the middle sublaminates. This trend will continue until the value of crack density reaches the critical crack density. It is clear that the more the cracked layers, the lower will be the critical crack density, and when the orientation of angles in the local coordinate gets far from the [0/90n]s, the energy release rates are bigger than the ordinary situations.

Specifying critical distances between cracks are essential because these values must be considered to anticipate when a crack can constitute a real danger for a structure.

Appendix

M1a=1ET+1λEA+(1ET1EA)cos2γλ(1ET1GA1+2υAEA)sin22γ4λ.

M2a=1GA(1+λ)λ+(1ET1GA1+2υAEA)sin22γλ.

M3a=1EA+1λET(1ET1EA)cos2γλ(1ET1GA1+2υAEA)sin22γ4λ.

M12a=(1ET1EA)sin2γλ+(1ET1GA1+2υAEA)sin4γ2λ.

M13a=2υAEA(1+λ)λ+(1ET1GA1+2υAEA)sin22γ2λ.

M23a=(1ET1EA)sin2γλ(1ET1GA1+2υAEA)sin4γ2λ.

M1b=13(1GT+λGA)+λ3(1GT1GA)cos2γ.

M2b=13(λGT+1GA)λ3(1GT1GA)cos2γ.

M12b=λ3(1GT1GA)sin2γ.

M1c=νTET(λ+23)υAλ3EAλ3(νTETυAEA)cos2γ.

M2c=νAEA(λ+23)υTλ3ET+λ3(νTETυAEA)cos2γ.

M3c=λ3(νTETυAEA)sin2γ.

M1d=(1+λ)(3λ2+12λ+8)60ET.

References

[1] Joffe R. Damage Accumulation and Stiffness Degradation in Composite Laminates, Thesis. Luleå University of Technology, 1999.Search in Google Scholar

[2] Blázquez A, Mantič V, París F, McCartney NL. Solids Struct. 2008, 45, 1632–1662.10.1016/j.ijsolstr.2007.10.013Search in Google Scholar

[3] Carlos G, Davila PPC, Rose, CA. J. Compos. Mater. 2005, 39, 323–345.10.1177/0021998305046452Search in Google Scholar

[4] Camanho PP, Dávila CG, Pinho ST, Iannucci L, Robinson P. Compos. Part A Appl. Sci. Manuf. 2006, 37, 165–176.10.1016/j.compositesa.2005.04.023Search in Google Scholar

[5] Hu S, Bark JS, Nairn JA. Compos. Sci. Technol. 1993, 47, 321–329.10.1016/0266-3538(93)90001-WSearch in Google Scholar

[6] Nairn JA. J. Compos. Mater. 1989, 23, 1106–1129.10.1177/002199838902301102Search in Google Scholar

[7] Rebière J-L. Cogent Eng. 2016, 3, 1175060.10.1080/23311916.2016.1175060Search in Google Scholar

[8] Hosseini-Toudeshky H, Farrokhabadi A, Mohammadi B. Proc. Eng. 2011, 10, 236–241.10.1016/j.proeng.2011.04.042Search in Google Scholar

[9] Akula VMK, Garnich MR. Mech. Adv. Mater. Struct. 2014, 21, 737–748.10.1080/15376494.2012.707302Search in Google Scholar

[10] Hashin Z. Mech. Mater. 1985, 4, 121–136.10.1016/0167-6636(85)90011-0Search in Google Scholar

[11] Hajikazemi M, Sadr MH. Int. J. Solids Struct. 2014, 51, 1483–1493.10.1016/j.ijsolstr.2013.12.040Search in Google Scholar

[12] Hajikazemi M, Sadr MH. Int. J. Solids Struct. 2014, 51, 516–529.10.1016/j.ijsolstr.2013.10.024Search in Google Scholar

[13] Wu X-F, Jenson RA, Zhao Y. Mech. Mater. 2014, 69, 195–203.10.1016/j.mechmat.2013.10.004Search in Google Scholar

[14] Kashtalyan M, Soutis C. Int. J. Solids Struct. 2002, 39, 1515–1537.10.1016/S0020-7683(02)00007-0Search in Google Scholar

[15] Chen H-S, Zhao Q-L, Gu Y, Yin Y-J, Fang D-N. Mech. Adv. Mater. Struct. 2013, 20, 564–570.10.1080/15376494.2011.643277Search in Google Scholar

[16] Chen Z, Yan W. Mech. Mater. 2015, 91, 119–135.10.1016/j.mechmat.2015.07.007Search in Google Scholar

[17] McCartney LN. J. Mech. Phys. Solids 1992, 40, 27–68.10.1016/0022-5096(92)90226-RSearch in Google Scholar

[18] McCartney LN. In Local Mechanics Concepts for Composite Material Systems: IUTAM Symposium, Blacksburg, VA, 1991, Reddy JN, Reifsnider KL, Eds., Springer: Berlin, Heidelberg, 1992, pp. 251–282.Search in Google Scholar

[19] McCartney LN. Compos. Sci. Technol. 2000, 60, 2255–2279.10.1016/S0266-3538(00)00086-5Search in Google Scholar

[20] Adolfsson E, Gudmundson P. Int. J. Solids Struct. 1999, 36, 3131–3169.10.1016/S0020-7683(98)00142-5Search in Google Scholar

[21] Okabe T, Imamura H, Sato Y, Higuchi R, Koyanagi J, Talreja R. Compos. Part A Appl. Sci. Manuf. 2015, 68, 81–89.10.1016/j.compositesa.2014.09.020Search in Google Scholar

[22] Hashin Z. Eng. Fracture Mech. 1986, 25, 771–778.10.1016/0013-7944(86)90040-8Search in Google Scholar

[23] Hashin Z. Compos. Mater. 1987, 54, 872–879.10.1115/1.3173131Search in Google Scholar

[24] Vinogradov V, Hashin Z. Compos. Sci. Technol. 2010, 70, 638–646.10.1016/j.compscitech.2009.12.018Search in Google Scholar

[25] Zhang H, Minnetyan L. J. Reinforced Plastics Compos. 2006, 25, 1919–1938.10.1177/0731684406069920Search in Google Scholar

[26] Nairn JA, Hu S. Eng. Fract. Mech. 1992, 41, 203–221.10.1016/0013-7944(92)90181-DSearch in Google Scholar

[27] McCartney LN. Compos. Sci. Technol. 1998, 58, 1069–1081.10.1016/S0266-3538(96)00142-XSearch in Google Scholar

[28] O’Brien TK. ASTM STP 876, 1985, 282–297.Search in Google Scholar

[29] O’Brien TK. ASTM STP 1059, 1990, 9, 7–33.Search in Google Scholar

[30] Dharani LR, Tang H. Int. J. Fract. 1990, 46, 123–140.10.1007/BF00041999Search in Google Scholar

[31] Nairn JA, Hu S. Int. J. Fract. 1992, 57, 1–24.10.1007/BF00013005Search in Google Scholar

[32] Li S, Lim S-H. Compos. Part A Appl. Sci. Manuf. 2005, 36, 353–365.10.1016/j.compositesa.2004.06.036Search in Google Scholar

[33] Lim S-H, Li S. Compos. Part A Appl. Sci. Manuf. 2005, 36, 1467–1476.10.1016/j.compositesa.2005.03.015Search in Google Scholar

[34] Huang ZQ, Zhou JC, He XQ, Liew KM. Int. J. Solids Struct. 2014, 51, 3669–3678.10.1016/j.ijsolstr.2014.06.028Search in Google Scholar

[35] Hashin Z. Mater. Struct. 1983, 50, 481–505.10.1115/1.3167081Search in Google Scholar

[36] Simmons G. Differential Equations with Applications and Historical Notes. McGraw-Hill, 1972.Search in Google Scholar

Received: 2016-12-16
Accepted: 2017-06-06
Published Online: 2017-10-17
Published in Print: 2018-09-25

©2018 Walter de Gruyter GmbH, Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Articles in the same Issue

  1. Frontmatter
  2. Review
  3. Progress in the research and applications of natural fiber-reinforced polymer matrix composites
  4. Original articles
  5. Damage assessment of random multiwalled carbon nanotube-reinforced polymer nanocomposites
  6. A variational approach for predicting initiation of matrix cracking and induced delamination in symmetric composite laminates under in-plane loading
  7. Processing-structure-property correlations of in situ Al/TiB2 composites processed by aluminothermic reduction process
  8. Interface transition layer interaction mechanism for ZTAP/HCCI composites
  9. Mechanical and electrical properties of polylactic acid/carbon nanotube composites by rolling process
  10. Development of hydroxyapatite-polylactic acid composite bone fixation plate
  11. Dielectric and thermal properties of magnesium oxide/poly(aryl ether ketone) nanocomposites
  12. A sustainable approach to optimum utilization of used foundry sand in concrete
  13. Improving flame-retardant, thermal, and mechanical properties of an epoxy using halogen-free fillers
  14. Optimization of the PDMS/biochar nanocomposite membranes using the response surface methodology
  15. Preparation and characterization of low-cost high-performance mullite-quartz ceramic proppants for coal bed methane wells
  16. Experimental investigation and analysis on the wear properties of glass fiber and CNT reinforced hybrid polymer composites
  17. Preparation of polyaniline-polyvinyl alcohol-silver nanocomposite and characterization of its mechanical and antibacterial properties
  18. An effective approach to synthesize carbon nanotube-reinforced Al matrix composite precursor
  19. Effect of oxygen plasma treatment on tensile strength of date palm fibers and their interfacial adhesion with epoxy matrix
  20. A novel characterization method of fiber reinforced polymers with clustered microstructures for time dependent mass transfer
  21. Stress relaxation behavior of annealed aluminum-carbon nanotube composite
  22. Restrained shrinkage cracking of self-consolidating concrete roads
  23. The effective ellipsoid: a method for calculating the permittivity of composites with multilayer ellipsoids
Downloaded on 12.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/secm-2016-0379/html
Scroll to top button