석사학위논문 Master’s Thesis

손상경계면을 갖는 복합재료의 탄성 및 탄소성 다단계 손상 모델

Elastic and Elastoplastic Multi-Level Damage Models for Composites with Imperfect Interface



석 훈 (表 錫 勳 Pyo, Suk-Hoon) 건설 및 환경공학과

Department of Civil and Environmental Engineering 한 국 과 학 기 술 원 Korea Advanced Institute of Science and Technology 2008

손상경계면을 갖는 복합재료의 탄성 및 탄소성 다단계 손상 모델

Elastic and Elastoplastic Multi-Level Damage Models for Composites with Imperfect Interface

Elastic and Elastoplastic Multi-Level Damage Models for Composites with Imperfect Interface

Advisor: Professor Haeng-Ki Lee by Suk-Hoon Pyo Department of Civil and Environmental Engineering Korea Advanced Institute of Science and Technology

A thesis submitted to the faculty of the Korea Advanced Institute of Science and Technology in partial fulfillment of the requirements for the degree of Master of Science in Engineering in the Department of Civil and Environmental Engineering. Daejeon, Korea 2007. 12. 10. Approved by

_____________________ Professor Haeng-Ki Lee

손상경계면을 갖는 복합재료의 탄성 및 탄소성 다단계 손상 모델

표 석 훈

위 논문은 한국과학기술원 석사학위논문으로 학위논문 심사 위원회에서 심사 통과하였음.

2007 년 12 월 10 일

심사 위원장

이 행 기

(인)

심 사 위 원

김 진 근

(인)

심 사 위 원

곽 효 경

(인)

i

MCE

표 석 훈. Pyo, Suk-Hoon. Elastic and Elastoplastic Multi-Level Damage Models

20063601

for Composites with Imperfect Interface. 손상경계면을 갖는 복합재료의 탄성 및 탄소성 다단계 손상 모델. Department of Civil and Environmental Engineering. 2008. 134p. Advisor: Prof. Lee, Haeng-Ki. Text in English.

ABSTRACT Micromechanics-based multi-level damage models are presented to predict elastic and elastoplastic overall behavior and damage evolution in composite materials. First of all, Eshelby’s tensor for an ellipsoidal inclusion with slightly weakened interface is adopted to model particles with imperfect interface in particle reinforced composites and fibers with imperfect interface in fiber reinforced composites, respectively. Imperfect interfaces between inclusion and matrix in composites occur as deformations or loadings continue to increase. As an effort to realistically reflect the effect of loading history on the progression of imperfect interface, a multi-level damage model in accordance with the Weibull’s probabilistic function is developed. It is assumed that the progression of imperfect interface is governed by the average stress of inclusion. To the estimate overall elastoplastic behavior of ductile matrix composites, an effective yield criterion is derived based on the ensemble-volume averaging process and the first-order effects of eigenstrains due to the existence of spherical inclusions. As applications, the overall elastic and elastoplastic stress-strain curves of particle reinforced brittle matrix composites, particle reinforced ductile matrix composites, and unidirectional fiber reinforced brittle matrix composites under uniaxial, biaxial, and triaxial loading are investigated in detail. A series of parametric analysis are carried out to investigate the influence of model parameters. Furthermore, the proposed multi-level damage models are compared with available experimental data in the literature to verify the accuracy of the models. Finally, proposed micromechanics-based evolutionary multi-level damage model for unidirectional fiber reinforced composites is implemented into a finite element program ABAQUS to predict behavior of laminated composites under various loading conditions.

ii

TABLE OF CONTENTS ABSTRACT..................................................................................................................................................i TABLE OF CONTENTS ............................................................................................................................ ii LIST OF FIGURES .....................................................................................................................................v

CHAPTER 1 INTRODUCTION ..............................................................................................................1 1.1 Review and Scope .......................................................................................................................1 1.2 Overview of the present study .....................................................................................................3

CHAPTER 2 MICROMECHANICS-BASED ELASTIC DAMAGE MODELING OF PARTICULATE COMPOSITES WITH IMPERFECT INTERFACES ....................................................................................................................5 2.1 Introduction .................................................................................................................................5 2.2 A micromechanics-based constitutive model for particulate composites with imperfect interfaces ....................................................................................................................8 2.2.1 Recapitulation of the Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface ..............................................................................................8 2.2.2 Effective elastic moduli of three-phase particulate composites ........................................9 2.3 Damage modeling ..................................................................................................................... 12 2.4 Numerical simulation ................................................................................................................ 14 2.5 Comparisons with other numerical and analytical predictions and experiment ........................ 22 2.6 Appendix ................................................................................................................................... 27

CHAPTER 3 MULTI-LEVEL MODELING OF EFFECTIVE ELASTIC AND PROGRESSIVE IMPERFECT INTERFAVE IN PARTICULATE COMPOSITES ................................................................................................................. 28 3.1 Introduction ............................................................................................................................... 28 3.2 Multi-level damage modeling ................................................................................................... 31 3.2.1 Level 0 of two-phase composite state ............................................................................. 31 3.2.2 Level 1 of three-phase composite state ........................................................................... 31

iii

3.2.3 Level 2 of four-phase composite state ............................................................................. 34 3.2.4 Level 3 of five-phase composite state ............................................................................. 35 3.3 A micromechanics-based, multi-level elastic model for particulate composites with imperfect interfaces .......................................................................................................... 36 3.3.1 Level 0 of two-phase composite state ............................................................................. 36 3.3.2 Level 1 of three-phase composite state ........................................................................... 37 3.3.3 Level 2 of four-phase composite state ............................................................................. 37 3.3.4 Level 3 of five-phase composite state ............................................................................. 38 3.4 Numerical example ................................................................................................................... 39 3.5 Parametric analysis ................................................................................................................... 41 3.6 Experimental comparisons ........................................................................................................ 44 3.7 Appendix ................................................................................................................................... 48

CHAPTER 4 AN ELASTOPLASTIC MULTI-LEVEL DAMAGE MODEL FOR DUCTILE MATRIX COMPOSITES EVOLUTIONARY IMPERFECT INTERFACE ............................................................................................ 50 4.1 Introduction ............................................................................................................................... 50 4.2 Ensemble-volume-average homogenization process ................................................................ 52 4.2.1 Overview ......................................................................................................................... 52 4.2.2 A first-order formulation of the current stress norm of multi-phase ductile matrix composites ................................................................................................ 52 4.2.3 A first-order formulation of effective elastoplastic behavior of multiphase ductile matrix composites ...................................................................................... 54 4.3 Numerical simulation ................................................................................................................ 60 4.4 Parametric analysis ................................................................................................................... 66 4.5 Experimental comparison.......................................................................................................... 69

CHAPTER 5 A MICROMECHANICS-BASED EVOLUTIONARY DAMAGE MODEL FOR CONTINUOUS, FIBER-REINFORCED BRITTLE MATRIX COMPOSITES WITH MICROCRACKS AND IMPERFECT INTERFACES .......................................................................................... 71

iv

5.1 Introduction ............................................................................................................................... 71 5.2 Effective elastic moduli of continuous, fiber-reinforced brittle matrix composites with microcracks and imperfect interfaces ............................................................ 74 5.3 Damage modeling ..................................................................................................................... 80 5.4 Numerical simulations .............................................................................................................. 83 5.5 Experimental comparison.......................................................................................................... 90 5.6 Appendix ................................................................................................................................... 93

CHAPTER 6 NUMERICAL ANALYSIS OF LAMINATED COMPOSITE STRUCTURES BASED ON MICROMECHANICAL MULTI-LEVEL DAMAGE MODEL .......................................................................................................... 97 6.1 Introduction ............................................................................................................................... 97 6.2 Micromechanics-based constitutive damage model for off-axis unidirectional fiber reinforced composites ............................................................................... 99 6.3 Damage modeling ................................................................................................................... 101 6.4 Numerical simulations and experimental comparison ............................................................ 104 6.4.1 Verification of the computer code ................................................................................. 104 6.4.2 Experimental comparison of the response and damage behavior of a laminated composite under bending ............................................................................. 106 6.4.3 Experimental comparison of the response and damage behavior of a laminated composite under shear .................................................................................. 113 6.5 Appendix ................................................................................................................................. 118

CHAPTER 7 CONCLUSIONS ............................................................................................................. 119

SUMMARY (IN KOREAN) .................................................................................................................. 124

REFERENCES ....................................................................................................................................... 125

ACKNOWLEDGEMENTS CURRICULUM VITAE

v

LIST OF FIGURES 2.1

Schematics of a particulate composite subjected to uniaxial tension: (a) the initial state; (b) the damaged state .......................................................................................................................6

2.2

The present predicted stress-strain responses of particulate composites with imperfect interfaces under uniaxial tension with various S 0 values ............................................... 14 B

2.3

B

The predicted evolution of volume fractions of perfectly bonded particles and particles with imperfect interface corresponding to Figure 2.2 (S 0 = 1.09*σ y ) ............................................ 15 B

2.4

B

B

B

The predicted evolution of volume fractions of perfectly bonded particles and particles with imperfect interface corresponding to Figure 2.2 (S 0 = 2.18*σ y ) ............................................ 16 B

2.5

B

B

B

The predicted evolution of volume fractions of perfectly bonded particles and particles with imperfect interface corresponding to Figure 2.2 (S 0 = 3.27*σ y ) ............................................ 17 B

2.6

B

B

B

The present predicted stress-strain responses of particulate composites with imperfect interfaces under biaxial tension with various S 0 values ................................................................. 18 B

2.7

B

The predicted evolution of volume fraction of particles with imperfect interface corresponding to Figure 2.6 ........................................................................................................... 19

2.8

The present predicted stress-strain responses of particulate composites with imperfect interfaces under hydrostatic tension with various S 0 values .......................................................... 20 B

2.9

B

The predicted evolution of volume fraction of particles with imperfect interface corresponding to Figure 2.8 ........................................................................................................... 20

2.10

The present predicted elastic responses of particulate composites with imperfect interfaces under uniaxial tension with various α and β values ....................................................... 21

2.11

The comparison of predicted overall elastic uniaxial responses of 6061-T6 aluminum alloy matrix/silicon-carbide particle composites between the present perfect bonding and imperfect interface models and Ju and Lee’s (2001) partial debonding model .............................. 23

2.12

The comparison between the present predictions with various combinations of volume fractions of perfectly bonded particles and particles of imperfect interface and the Halpin-Tsai’s theoretical prediction for effective (normalized) Young’s modulus versus volume fraction of particles ........................................................................................................... 24

2.13

The comparison between the present prediction and experimental data (Zhou et al., 2004) for overall uniaxial tensile responses of Mg-Al matrix with Al 2 O 3 particulate B

B

B

B

vi

composites...................................................................................................................................... 25 2.14

The predicted evolution of volume fractions of perfectly bonded particles and particles with imperfect interface corresponding to Figure 2.13 .................................................................. 26

3.1

Schematics of a particulate composite subjected to uniaxial tension: (a) Level 0 of twophase composite state (initial state); (b) Level 1 of three-phase composite state; (c) Level 2 of four-phase composite state; (d) Level 3 of five-phase composite state ................................... 30

3.2

The present predicted stress-strain responses of particulate composites under uniaxial tension ............................................................................................................................................ 39

3.3

The predicted evolution of volume fractions of perfectly bonded particles various types of damaged particles corresponding to Figure 3.2 ......................................................................... 40

3.4

The present predicted stress-strain responses of particulate composites under uniaxial tension at the four-phase composite state with various S 0 values .................................................. 41 B

3.5

B

The predicted evolution of volume fractions of perfectly bonded particles, particles with mild imperfect interface and particles with severe imperfect interface corresponding to Figure 3.4 ....................................................................................................................................... 42

3.6

The present predicted stress-strain responses of particulate composites under uniaxial tension at four-phase composite state with various α 2 and β 2 values ............................................ 43 B

3.7

B

B

B

The comparison between the present prediction and experimental data (Sørensen, 1993) for overall uniaxial tensile responses of SiC fiber-reinforced CAS glass-ceramic matrix composites...................................................................................................................................... 44

3.8

The predicted evolution of volume fractions of perfectly bonded particles, particles with mild imperfect interface, severe imperfect interface and completely debonded particles corresponding to the present prediction in Figure 3.7 .................................................................... 45

3.9

The comparison between the present prediction and experimental data (Geiger and Welch, 1997) for overall uniaxial tensile responses of silicon carbide particle-reinforced 2009 aluminum matrix composites ................................................................................................ 46

3.10

The predicted evolution of volume fractions of perfectly bonded particles, particles with mild imperfect interface, severe imperfect interface and completely debonded particles corresponding to the present prediction in Figure 3.9 .................................................................... 47

4.1

The present predicted elastoplastic stress-strain responses of particulate composites under uniaxial tension .................................................................................................................... 61

vii

4.2

(a) The predicted evolution of volume fractions of perfectly bonded particles and particles with mild impefect interface of the composites at the three-phase composite state; (b) The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles at the four-phase composite state; (c) The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles at the five-phase composite state corresponding to Figure 4.1 ......................... 62

4.3

The present predicted elastoplastic stress-strain responses of particulate composites under biaxial tension ...................................................................................................................... 63

4.4

The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles corresponding to Figure 4.3 ................................................................ 64

4.5

The present predicted elastoplastic stress-strain responses of particulate composites under triaxial tension (Γ = 0.6)....................................................................................................... 65

4.6

The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles corresponding to Figure 4.5 ................................................................ 66

4.7

The present predicted elastoplastic stress-strain responses of particulate composites under uniaxial tension at the four-phase composite state with various S 0 values .......................... 67 B

4.8

B

The present predicted elastoplastic stress-strain responses of particulate composites under uniaxial tension at the four-phase composite state with various α 2 and β 2 values ............... 68 B

4.9

B

B

B

The comparison between the present prediction and experimental data (Papazian and Adler, 1990) for overall uniaxial tensile responses of SiC particulate-reinforced 5456 aluminum alloy matrix composites ................................................................................................ 70

4.10

The predicted evolution of volume fractions of perfectly bonded particles, particles with mild imperfect interface, severe imperfect interface and completely debonded particles corresponding to the present prediction in Figure 4.9 .................................................................... 71

5.1

The schematic of a continuous, fiber-reinforced composite subjected to uniaxial loading in the longitudinal direction ........................................................................................................... 73

5.2

The schematic of a penny-shape crack modeled as an oblate spheroidal void (a 1 = a 2 >> B

B

B

B

a 3 ) and aspect ratio a 3 /a 1 →0 ......................................................................................................... 75 B

5.3

B

B

B

B

B

The schematic of multi-level damage transition in a continuous, fiber-reinforced composite: (a) Level 0 of three-phase composite state (initial state); (b) Level 1 of threephase composite state; (c) Level 2 of five-phase composite state; (d) Level 3 of six-phase

viii

composite state ............................................................................................................................... 76 5.4

The present predicted stress-strain responses of unidirectional composites under uniaxial tension in the longitudinal direction with various S 0 values .......................................................... 84 B

5.5

B

The predicted evolution of volume fractions of perfectly bonded fibers, fibers with mild imperfect interface, fibers with severe imperfect interface and completely debonded fibers corresponding to Figure 5.4 ................................................................................................. 85

5.6

The present predicted stress-strain responses of unidirectional composites under uniaxial tension in the transverse direction with various S 0 values ............................................................. 86 B

5.7

B

The present predicted stress-strain responses of particulate composites under uniaxial tension in the transverse direction with various α 2 and β 2 values.................................................. 87 B

5.8

B

B

B

The present predicted stress-strain responses of unidirectional composites under uniaxial tension in the longitudinal direction with various c 1 values .......................................................... 88 B

B

5.9

The predicted evolution of microcrack density corresponding to Figure 5.8 ................................. 89

5.10

The present predicted stress-strain responses of unidirectional composites under uniaxial tension in the longitudinal direction with various c 2 values .......................................................... 90 B

B

5.11

The predicted evolution of microcrack density corresponding to Figure 5.10 ............................... 90

5.12

The comparison between the present predictions (with and without microcracks) and experimental data (Luo and Daniel, 2000) for overall uniaxial tensile responses in the longitudinal direction of silicon carbide/calcium aluminosilicate (SiC/CAS) composites ............ 91

5.13

The predicted evolution of volume fractions of perfectly bonded fibers, fibers with mild imperfect interface, severe imperfect interface and completely debonded fibers (fivephase composite state) corresponding to Figure 5.12 .................................................................... 92

5.14

The predicted evolution of microcrack density (six-phase composite state) corresponding to Figure 5.12 ................................................................................................................................. 93

6.1

Coordinates for a transversely isotropic material ........................................................................... 99

6.2

The comparison between the present prediction and experimental data (Weeks and Sun, 1998) for overall uniaxial tensile responses with three fiber orientations of AS4/PEEK composite ..................................................................................................................................... 105

6.3

The predicted evolution of volume fraction of perfectly bonded fibers, fibers with mild imperfect interface, fibers with severe imperfect interface, completely debonded fibers and microcrack corresponding to 15º fiber orientation case in Figure 6.2 ................................... 106

ix

6.4

The finite element discretization employed to simulate the bending response and damage evolution in laminated composite beam ....................................................................................... 108

6.5

The comparison of load-displacement curves between the present prediction and experimental data (Huang, 2007) under three-point bending ....................................................... 109

6.6

The predicted damage evolution versus displacement of laminated composite beam under three-point bending corresponding to Figure 6.5: (a) Segment A; (b) Segment B; (c) Segment C; (d) Segment D; (e) Segment E; (f) Segment F .................................................... 110

6.7

The sequence of the deformed shape and von-Mises effective stress of laminated composite beam during three-point bending ................................................................................ 113

6.8

FE model of Iosipescu shear test .................................................................................................. 114

6.9

The comparison of load-displacement curves between the present prediction and experimental data (Khashaba, 2004) under shear loading: (a) [75/-15] 2s laminated B

B

composite; (b) [45/-45] 2s laminated composite; (c) [30/-60] 2s laminated composite .................. 115 B

6.10

B

B

B

The sequence of the deformed shape and von-Mises effective stress of laminated composite under shear loading in case of [75/-15] 2s .................................................................... 116 B

6.11

B

The sequence of the deformed shape and von-Mises effective stress of laminated composite under shear loading in case of [45/-45] 2s .................................................................... 117 B

6.12

B

The sequence of the deformed shape and von-Mises effective stress of laminated composite under shear loading in case of [30/-60] 2s .................................................................... 118 B

B

1

CHAPTER 1 INTRODUCTION

1.1 Review and Scope This dissertation is primarily related to the mechanical behavior of composite materials. Composite structures offer an attractive alternative to more conventional forms of construction due to its high strengthto-weight ratio and resistance to corrosion (Park et al., 2006). Development of new composite materials and new applications of composite materials in various engineering fields are now accelerating. Composite materials mean that two or more materials, called constituents, are combined in a certain configuration on a macro- or micro-scopic scale to form a new material with specific properties that combine the best features of each constituent so as to maximize a given set of properties (stiffness, strength-to-weight ratio, tensile strength, etc.) and minimize others (weight and cost) (Schoutens, 1982). In terms of reinforcement profile, composite materials are classified as particulate composites (e.g., whisker) and fibrous composites (e.g., short-fiber reinforced composites, unidirectional composites, and woven fiber reinforced composites). Furthermore, in terms of matrix profile, composite materials are also classified as ductile matrix composites (e.g., metal matrix composites) and brittle matrix composites (e.g., ceramic matrix composites and polymer matrix composites). As composite materials undergo damage, the corresponding stress and strain responses do not remain constant even for the case of elastic deformation (Voyiadjis and Park 1996). Interfacial bonding condition is one of the most important factors that control the local elastic fields and the overall properties of composite materials (Duan et al., 2007). The imperfect interfacial bonding may be due to damage in a very thin interfacial layer known as interphase or interface (Nie and Basaran, 2005). Therefore, to evaluate accurately the effective property of a composite, the properties and structures of interfaces must be taken into consideration (Qu, 1993b). Most analytical and numerical models assume that the bond between the reinforcement and matrix is perfect and can be modeled using the continuity of tractions and displacements across a discrete interface. When the perfect interface conditions are present, the primary result is the well-known Eshelby (1957) solution of the ellipsoidal inclusion problem. Eshelby’s micromechanics model provides a relatively easy tool to solve three-dimensional problems in elasticity, and thus is suitable for such a composite materials. In the literature, two kinds of model are often used to simulate the mechanical properties of interface

Chapter 1. Introduction

2

regions in particle- or fiber-reinforced composites. The first kind of model can be referred to as interface models, for example, linear spring model (Hashin, 1991; Qu, 1993b; Zhong and Meguid, 1997; Duan et al., 2007), free sliding model (Ghahremani, 1980; Jasiuk et al., 1987), and interface stress model (Sharma et al., 2003; Duan et al., 2005), in which displacement and/or stress discontinuities are assumed to exist at an interface. The second is the interphase model which describes the interface region as a layer, called an interphase, between the particle and matrix. The interphase model (three-phase model) with two perfect bonding interfaces was first proposed by Walpole (1978). Later, many studies concerned with homogeneous and inhomogeneous interphases have been done in the damage analysis and the prediction of effective moduli of composite materials (Qiu and Weng, 1991; Wang and Jasiuk, 1998; Dai et al., 1999; Wu et al., 2004; Zhong et al., 2004). The objective of this study is to determine the overall elastic and elastoplastic behavior of particleand fiber-reinforced composite materials containing imperfect interface based on micromechanical formulation proposed by Ju and Chen (1994a, 1994b, 1994c). A multi-level damage model in accordance with the Weibull's probabilistic function is developed in this study to describe the aforementioned sequential, progressive imperfect interface in the composites. Furthermore, in order to model spherical particles and cylindrical fibers having imperfect interfaces in the composites, the Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted. To verify proposed damage models, comparisons between numerical predictions and experimental data in the literature are conducted. Finally proposed damage models are implemented into ABAQUS non-linear FE code using a user-subroutine UMAT to solve various engineering boundary value problems and to numerically characterize the elastic behavior of laminated composites.

Chapter 1. Introduction

3

1.2 Overview of the present study An outline of the present work is as follows. In Chapter 2, a micromechanical framework is proposed to predict the effective elastic behavior and imperfect interface evolution of particulate composites. The Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted to model spherical particles having imperfect interfaces in the composites and is incorporated into the micromechanical framework. Based on the Eshelby's micromechanics, the effective elastic moduli of three-phase particulate composites are derived. A damage model is subsequently considered in accordance with the Weibull's probabilistic function to characterize the varying probability of evolution of imperfect interface between the inclusion and the matrix. The proposed micromechanical elastic damage model is applied to the uniaxial, biaxial and triaxial tensile loadings to predict the various stress-strain responses. Comparisons between the present predictions with other numerical and analytical predictions and available experimental data in the literature are conducted to assess the potential of the present framework.

In Chapter 3, a multi-level elastic damage model is proposed based on a combination of a micromechanical formulation and a multi-level damage model to predict the effective elastic behavior and progressive imperfect interface in particulate composites. The progression of imperfect interface is assumed to be gradual and sequential. A multi-level damage model in accordance with the Weibull's probabilistic function is developed to describe a sequential, progressive imperfect interface in the composites. The Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted to model particles having mild or severe imperfect interface and is incorporated into the micromechanical formulation by Ju and Chen (1994). A numerical example corresponding to uniaxial tension loading is solved to illustrate the potential of the proposed multi-level elastic damage model. A parametric analysis is also carried out to address the influence of model parameters on the progressive imperfect interface in the composites. Furthermore, the present predictions are compared with available experimental data in the literature to further illustrate the elastic damage behavior of the present framework and to verify the validity of the proposed multi-level elastic damage model.

In Chapter 4, an elastoplastic multi-level damage model considering evolutionary imperfect interface is developed in this work to predict the effective elastoplastic behavior and multi-level damage evolution in particle reinforced ductile matrix composites (PRDMCs). The elastoplastic multi-level damage model is micromechanically derived on the basis of the ensemble-volume averaging procedure and the first-order effects of eigenstrains. The Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface

Chapter 1. Introduction

4

(Qu, 1993a, 1993b) is adopted to model particles having mild or severe imperfect interface, and a multilevel damage model (Lee and Pyo, in press) in accordance with the Weibull's probabilistic function is employed to describe the sequential, progressive imperfect interface in the composites. Numerical examples corresponding to uniaxial, biaxial and triaxial tension loadings are solved to illustrate the potential of the proposed micromechanical framework. A series of parametric analysis are carried out to investigate the influence of model parameters on the progression of imperfect interface in the composites. Furthermore, the present prediction is compared with available experimental data in the literature to verify the proposed elastoplastic multi-level damage model.

In Chapter 5, a 3D micromechanics-based evolutionary damage model is proposed to predict the effective elastic behavior of continuous, fiber-reinforced brittle matrix composites with microcracks and imperfect interfaces. The Eshelby's tensor for a circular cylindrical inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted to model continuous fibers with imperfect interfaces. The nucleation of microcracks is simulated by employing the continuum damage model proposed by Karihaloo and Fu (1989). A multi-level damage modeling process in accordance with the Weibull's probabilistic function is incorporated into the micromechanical framework to describe the sequential evolution of imperfect interfaces in the composites. Numerical examples corresponding to uniaxial loadings in the longitudinal and transverse directions are solved to illustrate the potential of the proposed damage model. Furthermore, the present prediction is compared with available experimental data in the literature to verify the accuracy of the proposed damage model.

In Chapter 6, a 3D micromechanics-based evolutionary damage model (Lee and Pyo, submitted for publication) is implemented into a finite element program to numerically predict the overall behavior and damage evolution of laminated composite structures. A multi-level damage modeling process in accordance with the Weibull's probabilistic function is incorporated into the micromechanical framework to describe the sequential evolution of imperfect interfaces in the composites. To verify the implemented computational model, computational studies are undertaken by comparing the predicted stress-strain curve and experimental data available in literature under uniaxial tension with various fiber orientations. The implemented computational model provides results that show good agreement with available benchmark problems having different material compositions, thickness and boundary conditions.

In Chapter 7, concluding remarks are summarized.

5

CHAPTER 2 MICROMECHANICS-BASED ELASITC DAMAGE MODELING OF PARTICULATE COMPOSITES WITH IMPERFECT INTERFACES

2.1 Introduction Interfacial bonding condition is one of the most important factors that control the local elastic fields and the overall properties of composites (Duan et al., 2007). The imperfect interfacial bonding may be due to damage in a very thin interfacial layer known as interphase or interface (Nie and Basaran, 2005). The term imperfect interface is used to characterize a situation in which the displacements are discontinuous at the interface between the inclusion and the matrix (Crouch and Mogilevskaya, 2006). The linear springlayer of vanishing thickness, which is the simplest model of an imperfect interface, assumes that interfacial traction becomes continuous, but the displacements at the two sides of the interphase-layer become discontinuous (Crouch and Mogilevskaya, 2006). Refer to Benveniste (1985), Aboudi (1987), Achenbach and Zhu (1989, 1990), Hashin (1990, 1991), Zhu and Achenbach (1991), Qu (1993a, 1993b), Sudak et al. (1999), and Mogilevskaya and Croush (2002) for a further literature review on imperfect interface. Micromechanical damage models for composites considering interfacial debonding were proposed by many researchers (e.g., Zhao and Weng, 1997; Ju and Lee, 2000, 2001; Lee, 2001; Liang et al., 2006; Liu et al., 2006; Lee and Kim, 2007). They replaced the isotropic debonded inclusions by the perfectly debonded inclusions with yet unknown transversely isotropic properties to describe the loss of the loadtransfer capacity of debonded interface between inclusion and matrix. Schjudt-Thomsen and Pyrz (2000) studied creep modeling of short fiber reinforced composites with weakened interfaces and complex fiber orientation. Their approach was based upon the Mori-Tanaka mean field theory and the modified Eshelby's tensor for an inclusion with weakened interface proposed by Qu (1993a, 1993b). A similar work was previously conducted by Jun and Haian (1997). Zhong et al. (2004) studied three -dimensional micromechanical modeling of particulate composites with imperfect interface. In their derivations, the imperfect interface was characterized by a spring-type model assuming that the traction continuity remains intact, while the displacement experienced a jump proportional to the interfacial traction. Recently, Ju and Ko (in press) proposed a new micromechanical elastoplastic progressive damage model where the partial

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

6

debonding at the fiber interface was represented by the growing debonding angles of arc microcracks. A micromechanical framework is proposed in the present study to predict the effective elastic behavior and imperfect interface evolution of particulate composites. The schematic of a particulate composite with imperfect interfaces is illustrated in Figure 2.1. In order to model spherical particles having imperfect interfaces in the composite, the Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted. Based on the Eshelby's micromechanics, the effective elastic moduli of three-phase particulate composites are derived. A damage model is subsequently considered in accordance with the Weibull's probabilistic function to describe the varying probability of evolution of imperfect interface between the inclusion and the matrix. The novelty in the present derivation is in the simplicity and flexibility of the model to represent the imperfect interface evolution with limited number of model parameters. It is also straightforward to extend the proposed model to be used as a multi-level damage model, where some initially perfectly bonded particles may be transformed to particles with mild imperfect interface, some particles with mild imperfect interface may then be transformed to particles with severe imperfect interface, and all particles may be transformed to completely debonded particles asymptotically. In our derivations, the particles are assumed to be elastic spheres that are randomly dispersed in an elastic matrix. All particles are assumed to be non-interacting and initially embedded firmly in the matrix with perfect interfaces. It is also assumed that the evolution of imperfect interface is controlled by the

Matrix (phase 0)

Particle with imperfect interface (phase 2)

Perfectly bonded particle (phase 1)

(a)

(b)

Figure 2.1 Schematics of a particulate composite subjected to uniaxial tension: (a) the initial state; (b) the damaged state

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

7

average internal stresses of the particles as well as the Weibull parameters (Weibull, 1951). The proposed micromechanical elastic damage model is then applied to the uniaxial, biaxial and triaxial tensile loadings to predict the various stress-strain responses. Comparisons between the present predictions with other numerical and analytical predictions and available experimental data are conducted to further assess the potential of the present framework.

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

8

2.2 A micromechanics-based constitutive model for particulate composites with imperfect interfaces

2.2.1 Recapitulation of the Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface Qu (1993a, 1993b) derived the Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface in an elastic matrix of infinite extent. The imperfect interface between the inclusion and the matrix was modeled by a spring layer of vanishing thickness. The Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted to model spherical particles having imperfect interfaces in particulate composites and is incorporated into the micromechanical framework. The summary of the Eshelby's tensor derived by Qu (1993a, 1993b) is repeated here for completeness of the proposed elastic damage model. With the help of the Green's function in the unbounded domain, the strain field inside an elliptical inclusion in a homogeneous linearly elastic solid can be represented by (Qu, 1993b)

 ij  x   Sijkl  kl*  Tijst  x   I stkl  Sstkl   kl*

(2.1)

M  Sijkl  x   kl*

where S is the (original) Eshelby inclusion tensor, ε * is the uniform eigenstrain, I is the fourth-rank identity P

P

tensor and the fourth-rank tensor T is given by

Tijpq  x   Lklmn Lpqst

S ksGijmn ξ  xnt nl dS ξ 

(2.2)

in which L is the fourth-rank elasticity tensor, G(x) is related to the Green's function  given in Eq. (2.6a) and the Appendix of Qu (1993b), and n denotes the unit outward normal vector. In addition, the second-order tensor ij denoting the compliance of the interface spring layer is given by (Qu, 1993a)

ij  ij       ni n j

(2.3)

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

9

where  and  represent the compliance in the tangential and normal directions of the interface and

 ij signifies the Kronecker delta. Since Sijkl   Lklmn  Gijkl  ξ  x  dV  ξ  , one arrive at (Qu, 1993b) 

M Sijkl 





1 Sijkl  x dV  x   Sijkl  Iijpq  Sijpq H pqrs Lrsmn  I mnkl  Smnkl   

(2.4)

where

H pqrs 





1 ik n j nl   jk ni nl  il n j nk   jl ni nk dS 4 S

(2.5)

In the case of spherical inclusions, H ijkl can be simplified as (Qu, 1993b)

Hijkl   Pijkl       Qijkl

(2.6)

in which

Pijkl 

1 Iijkl a

Qijkl 

1 2 Iijkl  ij  kl 5a



(2.7)



(2.8)

Here, a denotes the radius of inclusion. Further details of the derivation of the Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interfaces can be found in Qu (1993a, 1993b).

2.2.2 Effective elastic moduli of three-phase particulate composites Let us consider an initially perfectly bonded two-phase composite consisting of an elastic matrix (phase 0) with bulk modulus  0 and shear modulus 0 , and randomly dispersed elastic spherical particles (phase 1) with bulk modulus 1 and shear modulus 1 . As loadings or deformations are applied, some

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

10

particles may be imperfect in their interfaces (phase 2). The effective stiffness tensor C* for the three-phase composite can be derived based on the governing equations for linear elastic composites containing arbitrarily non-aligned and/or dissimilar inclusions (Ju and Chen, 1994a) as

2 1    1 1 C*  C0  I   r  A r  Sr    I  r S r  A r  S r        r 1  

(2.9)

where C0 is the elasticity tensor of the phase 0 (matrix), “  ” is the tensor multiplication,  r denotes the volume fraction of the r-phase inclusion, and S r signifies the Eshelby's tensor for the r-phase. The fourthrank tensor A r is defined as 1

Ar  Cr  C0   C0

(2.10)

where Cr is the elasticity tensor of the r-phase. The components of Eshelby's tensor S1 for a (perfectly bonded) spherical particle embedded in an isotropic linear elastic and infinite matrix can be derived as

 S1 ijkl 





1 (5 0  1)  ij kl  (4  5 0 ) ( ik  jl   il  jk ) 15 1  0 

(2.11)

where  0 denotes the Poisson's ratio of the matrix. By carrying out the lengthy algebra, the Eshelby's tensors S 2 for a spherical particle with imperfect interface embedded in an isotropic linear elastic and infinite matrix given in Eq. (2.4) can be rephrased as

 S2 ijkl



1 1200a 1  0 

2





 80  480  400 2 a  500 1  2 2    0 0 0 0 

 98  140  50      268  1240  1300       320  720  400  a   3  2   7  5         2

0

2

0

0

0

0

0

ij kl

2

2

0

0

0

0

ik

jl

il

jk

(2.12)

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

11

where  0 denotes the Lame constant of the matrix. Substituting Eqs. (2.11) and (2.12) into Eq. (2.9) yields the effective stiffness tensor C* for the three-phase particulate composite as



C*  *ij kl   * ik  jl  il  jk



(2.13)

Accordingly, the effective Lame constant  * and the shear modulus  * for the three-phase composite can be expressed as

1 2



 *   30  20  1  3   20    2   4  1 2





 *  2 0    2   4  

(2.14)

(2.15)

where the components 1 , ... ,  4 are listed in the Appendix. Alternatively, the effective Young's modulus E * and Poisson's ratio  * of the three-phase particulate composite are easily obtained through the following relations

E  *

* 



 * 3 *  2 *   *



*

2 *  *

*





(2.16)

(2.17)

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

12

2.3 Damage modeling Imperfect interfaces between inclusion and matrix in particulate composites may occur as deformations or loadings continue to increase, and may affect the load-carrying capacity and the overall stress-strain behavior of the composites. For convenience, following Tohgo and Weng (1994) and Zhao and Weng (1995, 1996, 1997), the probability of imperfect interface is modeled as a two-parameter Weibull process. Assuming that the Weibull (1951) statistics governs and the average internal stresses of particles are the controlling factor of the Weibull function, one can express the current volume fraction of particles with imperfect interface  2 at a given level of 11 1 (see also Ju and Lee, 2001)

      M      2   Pd  11  1    1  exp    11 1      S0       

(2.18)

where  is the original particle volume fraction, P d denotes the cumulative probability distribution B

B

function of imperfect interface for the uniaxial tensile loading (in the 1-direction), 11 1 is the internal stress of particles (phase 1) in the 1-direction, the subscript  1 denotes the particle phase, and S 0 and M are B

B

the Weibull parameters. For the biaxial and hydrostatic tensile loadings, 11 1 in Eq. (2.18) are replaced by

 22    33   / 2 and 11    22    33   / 3 , respectively. 1 1 1 1 1   Within the context of the first-order (noninteracting) approximation, the stresses inside particles should be uniform. The internal stresses of particles required for the initiation of the imperfect interface were explicitly derived by Ju and Lee (2000) as

2 1  1  1  C1   I  S1   A1  S1    I  r S r   A r  S r      r 1  

1

:ε  U : ε

(2.19)

By carrying out the lengthy algebra, the components of the positive definite fourth-rank tensor U for the present model are explicitly given by



Uijkl  U1ij kl  U 2 ik  jl  il  jk



(2.20)

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

13

with

U1 

1  22  24 1 3 1 2 1   1  1 22   2 11  3   1 22   11  22  24  1  31  22  33  24 

(2.21)

U2 

 1 1 2 2   11  2 2  2 4 

(2.22)

where  1 is the current volume fraction of perfectly bonded particles,  1 is the Lame constant of particles, and the components 1 , ... ,  4 are listed in the Appendix. In the case of tensile loading, the averaged internal stresses of particles can be obtained as follows (see also Ju and Lee, 2000):

11 1  U1  2U2  11  U1 22  U133

(2.23)

 22 1  U111  U1  2U2   22  U133

(2.24)

 33 1  U111  U1 22  U1  2U2  33

(2.25)

where 11 ,  22 and  33 are the total strains in the 1, 2 and 3 directions, respectively.

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

14

2.4 Numerical simulation To illustrate the proposed micromechanical elastic damage model for particulate composites, let us first consider the uniaxial tension loading. In this case, the components of the macroscopic stress  can be expressed as 11  0 and  ij  0 for all other stress components. According to the linear elasticity, the macroscopic incremental elastic strain for this case reads (see also Ju and Lee, 2001)

1 0     0  * 0 0 

0    0  *11 E  * 

(2.26)

For convenience, we adopt the material properties and volume fraction of particles for the 6061-T6 aluminum alloy matrix/silicon-carbide particle composites according to Ju and Lee (2001) and Zhao and

800

700

600

Stress [σ11] (MPa)

500

400

300

200

perfect bonding perfectly bonded imperfect σy*3.27 interface (S 0 = 3.27*σ y ) σy*2.18 interface (S 0 = 2.18*σ y ) imperfect σy*1.09 interface (S 0 = 1.09*σ y ) imperfect

100

B

B

B

B

B

B

B

B

B

B

B

B

B

0 0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

Strain [ε11]

Figure 2.2 The present predicted stress-strain responses of particulate composites with imperfect interfaces under uniaxial tension with various S 0 values B

B

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

15

Weng (1995): E 0 = 68.3 GPa,  0 = 0.33, E 1 = 490 GPa, 1 = 0.17,  y = 250 MPa;  = 0.2. The compliance B

B

B

B

parameters  and  given in Eq. (2.3), which are related to the components of Eshelby's tensor S 2 for spherical inclusions with imperfect interfaces, are assumed to be  = 2.0 and  = 3.0. Moreover, to address the influence of the Weibull parameter S 0 on the evolution of imperfect interface and to evaluate the B

B

proposed elastic damage model sensitivity to the Weibull parameter S 0 , three different sets of Weibull B

B

parameters are used: S 0 =1.09×  y and M = 5; S 0 = 2.18×  y and M = 5; S 0 = 3.27×  y and M = 5. B

B

B

B

B

B

We plot the present predicted stress-strain responses of particulate composites with imperfect interfaces (imperfect interface model) under uniaxial tension with various S 0 values in Figure 2.2. We also B

B

plot the present predicted response of the perfectly bonded particulate composites (perfect bonding model) in the figure for comparison. Figures 2.3 – 2.5 exhibit the evolutions of perfectly bonded particles and

0.2

Particle Volume Fraction

0.15

0.1

0.05

perfect bonding imperfect interface 0 0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

Strain [ε11]

Figure 2.3 The predicted evolution of volume fractions of perfectly bonded particles and particles with imperfect interface corresponding to Figure 2.2 (S 0 = 1.09*σ y ) B

B

B

B

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

16

particles with imperfect interface as functions of the uniaxial strains corresponding to Figure 2.2. From Figures 2.2 – 2.5, it is clear that lower S 0 leads to faster evolution of imperfect interface, indicating that B

B

most particles are imperfect in their interfaces in early stage with lower S 0 and the composites show the B

B

nonlinear stress-strain behavior. It is also noted that the responses with the imperfect interface are lower than the response without the imperfect interface. To further examine the effect of varying S 0 values on the stress-strain responses of the composites B

B

under different loading cases, we consider biaxial and triaxial loading cases on next. In biaxial tensile loading case, the components of the macroscopic stress  can be expressed as  22   33  0 and  ij  0 for all other components. From the linear elasticity, the macroscopic incremental elastic strain for a biaxial tensile loading is given by (see also Ju and Lee, 2001)

0.2

Particle Volume Fraction

0.15

perfect bonding

0.1

imperfect interface

0.05

0 0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

Strain [ε11]

Figure 2.4 The predicted evolution of volume fractions of perfectly bonded particles and particles with imperfect interface corresponding to Figure 2.2 (S 0 = 2.18*σ y ) B

B

B

B

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

17

 2 * 0 0     *    0 1  0  *22   E  0 0 1  *   

(2.27)

In triaxial tensile loading case, the components of the macroscopic stress  can be expressed as  22   33  11  0 and  ij  0 for all other components, where  is the proportional loading ratio. In particular,  = 1 indicates a pure hydrostatic loading. From the linear elasticity, the macroscopic incremental elastic strain for a triaxial tensile loading can be expressed as (see also Ju and Lee, 2001)

0.2

Particle Volume Fraction

0.15

0.1

perfect bonding imperfect interface 0.05

0 0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

Strain [ε11]

Figure 2.5 The predicted evolution of volume fractions of perfectly bonded particles and particles with imperfect interface corresponding to Figure 2.2 (S 0 = 3.27*σ y ) B

B

B

B

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

1  2 *     0   0 

0   1   

18

    *11 0  E   1    *   0

*

0

(2.28)

Figures 2.6 and 2.8 show the present predicted stress-strain responses of particulate composites with imperfect interfaces under biaxial and hydrostatic tension loadings with various S 0 values, respectively. B

B

Figures 2.7 and 2.9 exhibit the evolution of particles with imperfect interface as functions of the biaxial and hydrostatic strains corresponding to Figures 2.6 and 2.8, respectively. Similar to the uniaxial tensile loading case, lower S 0 leads to faster evolution of imperfect interface and lower stress-strain response. It is also B

B

observed that hydrostatic tension results in higher stress-strain behavior, which corresponds to findings from Ju and Lee's (2001) observations.

700

600

Stress [σ22(=σ33)] (MPa)

500

400

300

200

perfect bonding perfectly bonded imperfect interface (S 0 = 3.27*σ y ) 3.27 imperfect interface (S 0 = 2.18*σ y ) 2.18 imperfect interface (S 0 = 1.09*σ y ) 1.09

100

B

B

B

B

B

B

B

B

B

B

B

B

B

0 0.000

0.001

0.002

0.003

0.004

0.005

Strain [ε22(=ε33)]

Figure 2.6 The present predicted stress-strain responses of particulate composites with imperfect interfaces under biaxial tension with various S 0 values B

B

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

19

Lastly, to examine the effect of varying  and  values on the stress-strain response of the proposed micromechanical framework, we conduct a parametric analysis of  and  . Four different sets of the compliance parameters are used:  = 2.0 and  = 3.0;   2.0 104 and   3.0 104 ;   2.0 105 and   3.0 105 ;   2.0 106 and   3.0 106 . Figure 2. 10 exhibit the present predicted stress-strain responses of particulate composites with imperfect interfaces under uniaxial tension with various values of  and  . It is seen from the figure that as the values of  and  become higher, the effect of imperfect interface on the stress-strain behavior is more pronounced. Further, it is shown from the parametric analysis that the converged values of the parameters for the lower bound are  = 2.0 and  = 3.0. It is also worthy mentioning that if  =  = 0, the material shows a perfect bonding material characteristic.

0.2

Particle Volume Fraction

0.15

0.1

0.05

S3.27 0 = 3.27*σ y S2.18 0 = 2.18*σ y S1.09 0 = 1.09*σ y

0 0.000

0.001

0.002

0.003

B

B

B

B

B

B

B

B

B

B

B

0.004

0.005

Strain [ε22(=ε33)]

Figure 2.7 The predicted evolution of volume fraction of particles with imperfect interface corresponding to Figure 2.6

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

20

1400

1200

Stress [σ11(=σ22=σ33)] (MPa)

1000

800

600

400

perfect bonding bonded 3.27 imperfect interface (S 0 = 3.27*σ y ) 2.18 imperfect interface (S 0 = 2.18*σ y ) 1.09 imperfect interface (S 0 = 1.09*σ y )

200

0 0.000

0.001

0.002

B

B

B

B

B

B

B

B

B

B

B

B

0.003

0.004

B

0.005

Strain [ε11(=ε22=ε33)]

Figure 2.8 The present predicted stress-strain responses of particulate composites with imperfect interfaces under hydrostatic tension with various S 0 values B

B

0.2

Particle Volume Fraction

0.15

0.1

0.05

S3.27 0 = 3.27*σ y S2.18 0 = 2.18*σ y S1.09 0 = 1.09*σ y B

B

B

B

B

B

B

B

B

B

B

0 0.000

0.001

0.002

0.003

0.004

0.005

Strain [ε11(=ε22=ε33)]

Figure 2.9 The predicted evolution of volume fraction of particles with imperfect interface corresponding to Figure 2.8

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

21

500

450

400

Stress [σ11] (MPa)

350

300

250

200

150

perfect bonding perfectly bonded α=2.0, β=3.0 standard α=2.E-4, case 1 β=3.E-4 α=2.E-5, case 2 β=3.E-5 α=2.E-6, case 3 β=3.E-6

100

B

50

B

B

0 0.000

0.001

0.002

0.003

0.004

0.005

Strain [ε11]

Figure 2.10 The present predicted elastic responses of particulate composites with imperfect interfaces under uniaxial tension with various α and β values

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

22

2.5 Comparisons with other numerical and analytical predictions and experiment To verify and further assess the proposed micromechanical framework, we compare the present predictions with other numerical (Ju and Lee, 2001) and analytical (Kerner, 1956; Halpin and Kardos, 1976) predictions and available experimental data (Zhou et al., 2004). We first compare the present prediction with the numerical prediction given in Ju and Lee (2001) for 6061-T6 aluminum alloy matrix/silicon-carbide particle composites under uniaxial tension in Figure 2.11. Similar to Section 2.4, the elastic properties of the composites used in this simulation are (see Arsenault, 1984; Nieh and Chellman, 1984; Zhao and Weng, 1995; Ju and Lee, 2001) E 0 = 68.3 GPa,  0 = 0.33, E 1 = B

B

B

B

490 GPa,  1 = 0.17,  y = 250 MPa;  = 0.2. The compliance parameters  and  are assumed to be  =2.0 and  = 3.0 (lower bound) to clearly show the effect of imperfect interface on the stress-strain behavior of imperfect particulate composites. The Weibull parameters S 0 and M are assumed to be S 0 = 1.09×  y , M = B

B

B

B

5. Clearly, the present imperfect interface model shows slightly high stress values during the evolution of imperfect interface in comparison with the partial debonding model (Ju and Lee, 2001). It should be noted that we do not display our predictions beyond the strain value 11 = 0.004, which corresponds to the yielding stress of the Ju and Lee's (2001) predicted response, in the figure since the present model does not consider plastic formulations. Our analytical predictions are compared with Halpin-Tsai's analytical solutions (Kerner, 1956; Halpin and Kardos, 1976) for particulate composites to validate the proposed micromechanical framework. The Halpin-Tsai's analytical equations predict the elastic moduli of composites with different reinforcement geometries changing from a sphere (aspect ratio of one in the principal material direction) to a long fiber (aspect ratio approaching infinity) (Halpin and Kardos, 1976). The analytical equation for the Young's modulus of particulate composites is given by (Kerner, 1956; Halpin and Kardos, 1976; Lee and Simunovic, 2001)



E*  2 1  *



f f   m    (7  5  )  (8  10  )  15(1   ) m m f m      f m m     (7  5 m )  (8  10 m )  f 15(1  f ) 

(2.29)

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

23

where subscripts m and f denote the matrix and particle phases, respectively, and the Poisson ratio * is given by a rule of mixtures

 *   m m  f  f

(2.30)

The following constituent elastic properties for carbon fiber polymer matrix composites are considered for comparison: E 0 = 3 GPa, 0 = 0.35, E 1 = 380 GPa and  1 = 0.25 (Lee and Simunovic, 2001). B

B

B

B

Figure 2.12 shows the effective (normalized) Young's modulus versus volume fraction of particles predicted from the present perfect bonding and imperfect interface models. In particular, various combinations of volume fractions of perfectly bonded particles (phase 1) and particles with imperfect

400

partial debonding (Ju and Lee, 2001) imperfect interface (present prediction) perfect bonding (present prediction)

Stress [σ11] (MPa)

300

200

100

0 0.000

0.001

0.002

0.003

0.004

Strain [ε11]

Figure 2.11 The comparison of predicted overall elastic uniaxial responses of 6061-T6 aluminum alloy matrix/silicon-carbide particle composites between the present perfect bonding and imperfect interface models and Ju and Lee’s (2001) partial debonding model

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

24

interface (phase 2) are considered in the imperfect interface model as:  1 /  2 = 9/1, 7/3, 5/5, 0/10 (lower bound). The compliance parameters are assumed to be  = 2.0,  = 3.0. It is also assumed that the present imperfect interface model is stationary (no damage evolution). The theoretical prediction based on the Halpin-Tsai's analytical solution is also plotted in the figure. It is seen from Figure 2.12 that the present prediction from the perfect bonding model (upper bound) is almost identical with the Halpin-Tsai's analytical solution. The present predictions from the imperfect interface model are well within the bounds. To further assess the potential of the present framework, we compare the present prediction with experimental data reported by Zhou et al. (2004) for the uniaxial stress-strain behavior of Mg-Al matrix with Al 2 O 3 particulate composites. Here, we adopt the elastic properties according to Zhou et al. (2004) as B

B

B

B

140

120

Halpin-Tsai’s analytical solution Kerner perfect bonding bonding imperfect interface (  1:  2 =9:1) 9:1 imperfect interface (  1:  2 =7:3) 7:3 imperfect interface (  1:  2 =5:5) 5:5 imperfect interface (  1:  2 =0:10) 0:10

100

B

B

B

B

E*/E0

80

60

40

20

0 0

0.2

0.4

0.6

0.8

1

Volume fraction of particle

Figure 2.12 The comparison between the present predictions with various combinations of volume fractions of perfectly bonded particles and particles of imperfect interface and the Halpin-Tsai’s theoretical prediction for effective (normalized) Young’s modulus versus volume fraction of particles

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

25

follow: E 0 = 73 GPa,  0 = 0.33, E 1 = 400 GPa,  1 = 0.24,  y = 250 MPa,  = 0.48. It is emphasized that we B

B

B

B

need to estimate the compliance parameters  and  given in Eq. (2.12) as well as the Weibull parameters S 0 and M in Eq. (2.18) due Zhou et al. (2004) did not report those parameters. In accordance with the B

B

experimental data reported by Zhou et al. (2004), the above compliance and Weibull parameters are estimated to be:  = 2.0,  = 3.0; S 0 = 625 MPa, M = 5. B

B

Based on the above material properties and parameters, we depict our uniaxial stress-strain predictions on the composites with and without evolution of imperfect interface against experimental data provided by Zhou et al. (2004) in Figure 2.13. Overall, the present prediction from the imperfect interface model and the experimental data match well. The predicted evolution of imperfect interface corresponding to Figure 2.12 is shown in Figure 2.14. The good agreement between the present prediction and the experiment encourages for possible use of the proposed micromechanical framework to modeling the imperfect interface in (quasi-brittle) particulate composites. 400

Stress [σ11] (MPa)

300

200

100

experiment imperfect interface perfect bonding 0 0.0000

0.0005

0.0010

0.0015

0.0020

0.0025

Strain [ε11]

Figure 2.13 The comparison between the present prediction and experimental data (Zhou et al., 2004) for overall uniaxial tensile responses of Mg-Al matrix with Al 2 O 3 particulate B

composites

B

B

B

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

26

0.5

Fiber Volume Fraction

0.4

0.3

0.2

perfect bonding 0.1

imperfect interface

0 0.0000

0.0005

0.0010

0.0015

0.0020

0.0025

Strain [ε11]

Figure 2.14 The predicted evolution of volume fractions of perfectly bonded particles and particles with imperfect interface corresponding to Figure 2.13

Chapter 2. Elastic damage modeling of composites with imperfect interfaces

27

2.6 Appendix. Parameters Λ 1 , . . , Λ 4 in Eqs. (2.14), (2.15), (2.21), and (2.22) B

B

B

B

  1  22  1  1  31  2 1    1  15 1  0   1     2 1  31  2 1 1  2 2 1  31  2 2    2 

(2.31)

15 1  0   1

(2.32)

4 1 1  2 2 

  2  24  1  3  32  2 2  2   3  600a 1  0   2    2  32  2 2 1  24  1  33  24    

(2.33)

300a 1  0   2 2

4 

(2.34)

 2 1  24 

in which

1 

  1  51 1  0    5 0  1 ,  2 1  3   2  1 1   

2 

3 

2 2 2

    2  31  2  2    1  ,  3   2  2 2    

4 

 1 15 1  0 

(2.36)

2 2

 1  0 3   1   0   2  1  0  2

(2.35)

 2 2

1 0  0  1

 2  1200a 1  0  1 

 1  4  5 0  2 1

 5 0  1

(2.37)

 2  0  0  2  1  2  0 3   2   0   2  2  0 

(2.38)

 7  5 0  0  8  10 0  1 2  1  0 

 2  600a 1  0 

2

0

2  0

(2.39)

 2

(2.40)

with









1  80  480 0  400 02 a  500 1  2 0  0   98  140 0  50 02 0



2



 268  1240 0  1300 02 0 





2  320  720 0  400 02 a  0  3  2   7  5 0 

2

(2.41)

(2.42)

28

CHAPTER 3 MULTI-LEVEL MODELING OF EFFECTIVE ELASTIC AND PROGRESSIVE IMPERFECT INTERFAVE IN PARTICULATE COMPOSITES

3.1 Introduction Debonding phenomenon existed in between particles and matrix is one of well-known damage mechanisms in particulate composites and its effect on the mechanical behavior of the composites has to be well addressed for an accurate analysis of the composites. The progressive interfacial debonding between particles and matrix may occur under increasing deformation or stresses from the viewpoint of failure analysis. One-step Weibull distribution has been used by many researchers to describe the interfacial debonding between particles (fibers) and matrix (e.g., Tohgo and Weng, 1994; Zhao and Weng, 1995, 1996, 1997; Ju and Lee, 2000, 2001; Lee and Simunovic, 2000, 2001, 2002; Liu et al., 2005; Liang et al., 2006; Lee and Pyo, 2007). However, a sequential probabilistic debonding analysis is necessary to realistically reflect the effect of loading history on the interfacial debonding. Gurvich and Pipes (1995, 1996), Tabiei and Sun (1999, 2000) and Hwang et al. (2003) introduced sequential multi-step failure models to model the sequential probabilistic failure between fibers and matrix in laminated composites. In their sequential multi-step failure models, a structure was modeled as a parallel arrangement of elements based on the fiber bundle theory (Hwang et al., 2003). When an element fails, the load is redistributed among the remaining elements and final failure occurs when all of the elements fail (Hwang et al., 2003). Another approach to model the sequential probabilistic failure is proposed by Liu et al. (2004, 2006) and Ju and Ko (in press). In their formulations, a progressive interfacial debonding model was developed based on Weibull distribution function (Liu et al., 2004, Ju and Ko, in press) and one damage mode was transformed to another as applied external loads increased. That is, another damage mode is initiated when the previous local principal radial normal stress reaches the critical interfacial debonding strength (Liu et al., 2004, Ju and Ko, in press). Recently, a multi-level model for damage evolution in microstructurally debonded composites was developed by Ghosh et al. (2007) for multi-scale analysis of damaged composite structures due to microstructural damage induced by interfacial debonding.

Chapter 3. Multi-level modeling of particulate composites

29

In this study, a multi-level elastic damage model based on a combination of a micromechanical formulation and a multi-level damage model is proposed to predict the effective elastic behavior and progressive imperfect interface in particulate composites. Since the progression of imperfect interface is assumed to be gradual and sequential, as loads or deformations continue to increase, some initially perfectly bonded particles are transformed to particles with mild imperfect interface. Some particles with mild imperfect interface are then transformed to particles with severe imperfect interface, and all particles are transformed to completely debonded particles asymptotically within the proposed framework. A multilevel damage model in accordance with the Weibull's probabilistic function is developed to describe the aforementioned sequential, progressive imperfect interface in the composites. The Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adapted to model particles having mild or severe imperfect interface and is incorporated into the micromechanical formulation by Ju and Chen (1994a). The schematic of a particulate composite under uniaxial tension, illustrating a sequential progression of imperfect interface in the composite, is shown in Figure 3.1. In particular, the following four-level elastic damage model is considered for a complete description of the sequential progression of imperfect interface in the composite: (1) level-0 of two-phase composite state consisting of matrix and perfectly bonded particles; (2) level-1 of three-phase composite state consisting of matrix, perfectly bonded particles and particles having mild imperfect interface; (3) level-2 of four-phase composite state consisting of matrix, perfectly bonded particles, particles having mild imperfect interface and particles having severe imperfect interface; (4) level-3 of five-phase composite state consisting of matrix, perfectly bonded particles, particles having mild imperfect interface, particles having severe imperfect interface and completely debonded particles. We assume that the level of severity of imperfect interface can be classified simplistically as either mild (moderately) imperfect interface or severe imperfect interface. It is obvious that the severity of imperfect interface may vary as loads or deformations continue to increase. However, our premise is that its cumulative effect on the effective stress-strain behavior can be quantitatively described by using this simple approximation. The two imperfect states, mild imperfect interface and severe imperfect interface, are also assumed to occur sequentially. Further, particles are assumed to be randomly dispersed, non-interacting elastic spheres that are initially embedded firmly in an elastic matrix with perfect interfaces. It is also assumed that the progression of imperfect interface is governed by the average internal stresses of the particles as well as the Weibull parameters (Weibull, 1951). A numerical example corresponding to uniaxial tension loadings is solved to illustrate the potential of the proposed multi- level elastic damage model. A parametric analysis is also carried out to address the influence of model parameters on the progressive imperfect interface in the

Chapter 3. Multi-level modeling of particulate composites

30

composites. Comparisons between the present prediction with available experimental data are conducted to further illustrate the elastic damage behavior of the present framework and to verify the validity of the proposed multi-level elastic damage model.

a

Perfectly bonded

b

particle (phase 1)

Matrix (phase 0)

Particle with mild imperfect interface (phase 2)

c

d

Particle with severe

Void (phase 4)

imperfect interface (phase 3)

Figure 3.1 Schematics of a particulate composite subjected to uniaxial tension: (a) Level 0 of twophase composite state (initial state); (b) Level 1 of three-phase composite state; (c) Level 2 of four-phase composite state; (d) Level 3 of five-phase composite state

Chapter 3. Multi-level modeling of particulate composites

31

3.2 Multi-level damage modeling Imperfect interfaces between particles and matrix in particulate composites may occur as deformations or loadings continue to increase, and may affect the load-carrying capacity and overall stressstrain behavior of the composites. As an effort to realistically reflect the effect of loading history on the progression of imperfect interface, a multi-level damage model in accordance with the Weibull's probabilistic function is developed. Specifically, a four-level damage model is considered in the present study, in the order of sequence of progressive imperfect interface, for a complete description of the sequential, progressive imperfect interface in the composite as illustrated in Figure 3.1 and is discussed in detail in this section.

3.2.1 Level 0 of two-phase composite state As illustrated in Figure 3.1 (a), in the initial state, an initially perfectly bonded two-phase composite state consists of an elastic matrix (phase 0), and randomly dispersed elastic spherical particles (phase 1) with the volume fraction  .

3.2.2 Level 1 of three-phase composite state Following Tohgo and Weng (1994) and Zhao and Weng (1995, 1996, 1997), the probability of mild imperfect interface is modeled as a two-parameter Weibull process. Assuming that the Weibull (1951) statistics governs and the average internal stresses of perfect bonded particles (phase 1) are the controlling factor of the Weibull function, the current volume fractions of particles having mild imperfect interface  2 and perfectly bonded particles  1 in the three-phase composite at a given level of 11 1 are (see also Ju and Lee, 2001)

 

     M    11 1     S0     

 2   Pd  11  1    1  exp   

(3.1)

1     2

(3.2)

 

Chapter 3. Multi-level modeling of particulate composites

32

where  is the original particle volume fraction, P d signifies the cumulative probability distribution B

B

function of mild imperfect interface for the uniaxial tensile loading (in the 1-direction), internal stress of particles (phase 1) in the 1-direction, the subscript

11 1 is

the

1 denotes the particle phase, and S 0 B

B

and M are the Weibull parameters. For the biaxial and hydrostatic tensile loadings, 11 1 in Eq. (3.1) are replaced by  22 1   33 1  / 2 and 11 1   22 1   33 1  / 3 , respectively. The internal stresses of particles required for the initiation of the imperfect interface were explicitly derived by Ju and Lee (2000) as

n 1  1  1  C1   I  S1   A1  S1    I  r S r   A r  S r      r 1  

1

:ε  U : ε

(3.3)

where C 1 is the elasticity tensor of the perfectly bonded particles, “  ” is the tensor multiplication, I is the B

B

fourth-rank identity tensor, and  r denotes the volume fraction of the r-phase inclusion. The fourth-rank tensor A r is defined as B

B

1

Ar  Cr  C0   C0

(3.4)

The components of Eshelby's tensor S 1 for a (perfectly bonded) spherical particle and S 4 for a B

B

B

B

(completely debonded) spherical particle embedded in an isotropic linear elastic and infinite matrix can be derived as

 S1,4 ijkl  15 11  (5 0  1) ij kl  (4  5 0 ) (ik jl  il jk )

(3.5)

0

where  0 denotes the Poisson's ratio of the matrix and  ij signifies the Kronecker delta. Following Chapter 3, the Eshelby's tensors S m+1 for a spherical particle with imperfect interface B

B

embedded in an isotropic linear elastic and infinite matrix can be expressed as

Chapter 3. Multi-level modeling of particulate composites

 Sm1 ijkl 

1 1200a 1  0 

2



33



 80  480  400 2 a  500 1  2 2    0 0 0 0 m 

 98  140  50      268  1240  1300       320  720  400  a   3  2  7  5         2

0

2

0

0 m

0

0

ij kl

(3.6)

2

2

0

0 m

0

0

m

m

0

ik

jl

il

jk

where m = 1, 2 in which 1 indicates the mild imperfect interface stage, while 2 denotes the severe imperfect interface stage. In addition, a denotes the radius of particles,  0 and 0 denote the Lame constants of the matrix, and  and  represent the compliance parameters in the tangential and normal directions of the interface. Solutions to the case of small  with  = 0 yield approximations to the free-sliding interfaces, whereas the free-sliding can be achieved by setting    with  = 0 (Qu, 1993b). Following Qu's (Qu, 1993b) definitions on  and  , the severity of imperfect interface is assumed to be controlled by the values of  and  in the present study. It should also be noted that those parameters are not position dependent, indicating that no interfacial partial debonding is considered in this framework. By carrying out the lengthy algebra, the components of the positive definite fourth-rank tensor U for the multi-phase composite state are explicitly given by



Uijkl  U1ij kl  U 2 ik  jl  il  jk



(3.7)

with

U1 

1  22  24 1 3 1 2 1   1  1 22   2 11  3   1 22   11  22  24  1  31  22  33  24 

(3.8)

U2 

 1 1 2 2   11  2 2  2 4 

(3.9)

where  1 and 1 are the Lame constants of particles and the components 1 , ... ,  4 are listed in the Appendix.

Chapter 3. Multi-level modeling of particulate composites

34

3.2.3 Level 2 of four-phase composite state Similar to the Level 1 of three-phase composite state, assuming that the Weibull (1951) statistics governs and some particles with mild imperfect interface are transformed to particles with severe imperfect interface as deformations or loads increase, the current volume fractions of particles having severe imperfect interface  3 , particles having mild imperfect interface  2 and perfectly bonded particles  1 in the four-phase composite state at a given level of 11 1 can be derived through the following two-step Weibull approach

      M      2   1  exp    11 1      S0       

(3.10)

      M      3   2 1  exp    11 1      S 0       

(3.11)

 2   2  3

(3.12)

1     2

(3.13)

The components of the positive definite fourth-rank tensor U for the four-phase composite state in Eq. (3.7), denoted by U 1 and U 2 , can be expressed as B

B

B

B

U1 

1  22  24  26 1 3 1 2 1   1  1 22   2 11  3  5   1 22   11  22  24  26 1  31  22  33  24  35  26 

U2 

 1 1 22   11  2 2  2 4  26 

where the components 1 , ... ,  6 are listed in the Appendix.

(3.14)

(3.15)

Chapter 3. Multi-level modeling of particulate composites

35

3.2.4 Level 3 of five-phase composite state Similar to the Level 2 of four-phase composite state, assuming that the Weibull (1951) statistics governs and some particles with severe imperfect interface are transformed to completely debonded particles as deformations or loads increase, the current volume fractions of completely debonded particles  4 particles having severe imperfect interface  3 , particles having mild imperfect interface  2 and perfectly bonded particles  1 in the five-phase composite at a given level of 11 1 can be derived through the following three-step Weibull approach

      M      2   1  exp    11 1      S0       

(3.16)

      M      3   2 1  exp    11 1      S 0       

(3.17)

 

     M    11 1     S0     

 4   3 1  exp   

(3.18)

3  3   4

(3.19)

 2   2  3

(3.20)

1     2

(3.21)

 

The components of the positive definite fourth-rank tensor U for the five-phase composite state in Eq. (3.7), denoted by U 1 and U 2 , can be expressed as B

U1 

B

B

B

1  22  24  26  28 1 3 1 2 1   1  1 22   2 11  3  5  7   1 22   11  22  24  26  28 1  31  2 2  33  2 4  35  26  37  28 

U2 

 1 1 22   11  22  24  26  28 

where the components 1 , ... , 8 are listed in the Appendix.

(3.22)

(3.23)

Chapter 3. Multi-level modeling of particulate composites

36

3.3 A micromechanics-based, multi-level elastic model for particulate composites with imperfect interfaces We start with considering an initially perfectly bonded two-phase composite consisting of an elastic matrix (phase 0) with bulk modulus  0 and shear modulus 0 , and randomly dispersed elastic spherical particles (phase 1) with bulk modulus 1 and shear modulus 1 . As loads or deformations continue to increase, some initially perfectly bonded particles are transformed to particles with mild imperfect interface, some particles with mild imperfect interface are then transformed to particles with severe imperfect interface, and all particles are transformed to completely debonded particles asymptotically within the proposed framework. The effective stiffness tensor C* for the multi-phase particulate composite can be derived based on the governing equations for linear elastic composites containing arbitrarily non-aligned and/or dissimilar inclusions (Ju and Chen, 1994a) as

n 1    1 1 C*  C0  I   r  A r  Sr    I  r S r  A r  S r        r 1  

(3.24)

where Cr is the elasticity tensor of the r-phase. Substituting Eqs. (3.5) and (3.6) into Eq. (3.24) yields the effective stiffness tensor C* for the multiphase particulate composite as



C*  *ij kl   * ik  jl  il  jk



(3.25)

3.3.1 Level 0 of two-phase composite state The effective Lame constants  * and  * in Eq. (3.25) for the two-phase composite state can be expressed as

1 2



 *   30  20  1  20    2  

(3.26)

Chapter 3. Multi-level modeling of particulate composites

1 2

37



 *  2 0    2 

(3.27)



where the components 1 and  2 are listed in the Appendix.

3.3.2 Level 1 of three-phase composite state The effective Lame constants  * and  * in Eq. (3.25) for the three-phase composite state can be expressed as

1 2



 *   30  20  1  3   20    2   4  1 2

(3.28)





 *  2 0    2   4 

(3.29)



where the components 1 ,..,  4 are listed in the Appendix.

3.3.3 Level 2 of four-phase composite state The effective Lame constants  * and  * in Eq. (3.25) for the four-phase composite state can be expressed as

1 2



 *   30  20  1  3  5   20    2   4   6  1 2



 *  2 0    2   4   6  

where the components 1 ,..,  6 are listed in the Appendix.



(3.30)

(3.31)

Chapter 3. Multi-level modeling of particulate composites

38

3.3.4 Level 3 of five-phase composite state The effective Lame constants  * and  * in Eq. (3.25) for the five-phase composite state can be expressed as

1 2



 *   30  20  1  3  5  7   20    2   4   6  8  1 2



 *  2 0    2   4   6   8  

where the components 1 ,.., 8 are listed in the Appendix.



(3.32)

(3.33)

Chapter 3. Multi-level modeling of particulate composites

39

3.4 Numerical example To illustrate the elastic damage behavior of the present framework, we consider the uniaxial tension loading case ( 11  0 and  ij  0 for all other stress components). For brevity, we adopt the material properties for the 6061-T6 aluminum alloy matrix/silicon-carbide particle composites according to Zhao and Weng (1995) and Ju and Lee (2001): E 0 = 68.3 GPa,  0 = 0.33, E 1 = 490 GPa, 1 = 0.17,  y = 250 B

B

B

B

MPa,  = 0.2. As a result of the parametric analysis on compliance parameters  and  reported in Chapter 2, the compliance parameters of the mild imperfect interface, denoted by 1 and 1 , and the severe imperfect interface, denoted by  2 and  2 , in Eq. (3.6) are chosen to be 1  2.0 105 , 1  3.0 105 ;  2 = 2.0,  2 = 3.0. Moreover, the Weibull parameters in Eq. (3.1) are assumed to be S 0 = 1.09×  y and M = 5. B

B

500

450

2-phase composite state 3-phase composite state

400

4-phase composite state 5-phase composite state

Stress [σ11] (MPa)

350

300

250

200

150

100

50

0 0.000

0.001

0.002

0.003

0.004

0.005

Strain [ε11]

Figure 3.2 The present predicted stress-strain responses of particulate composites under uniaxial tension

Chapter 3. Multi-level modeling of particulate composites

40

To investigate the influence of the level of damage on the stress-strain response of the composites, four different phase composite states illustrated in Figure 3.1 are considered in this simulation. Figure 3.2 shows the present predicted stress-strain responses of particulate composites under uniaxial tension. The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles (e.g., particles with mild imperfect interface, particles with severe imperfect interface, completely debonded particles) at the three-phase, four-phase and five-phase composite states are exhibited in Figure 3.3. At the early stage of loading, the stress-strain curves are shown to be linear since the volume fractions of the damaged particles are low (Figure 3.2). As strains or stresses continue to increase, the volume fractions of the damaged particles increase gradually and the composites show nonlinear stress-strain responses as shown in Figure 3.2. Clearly, the high-level composite states give rise to substantial weaker stress-strain response in comparison with the low-level composite states. It is demonstrated through this simulation that the current multi-level elastic damage model naturally captures the gradual transition from previous damage level to next damage level and various types of damaged particles exist simultaneously within the proposed framework. 0.2

Particle Volume Fraction

0.15

perfectly bonded (3-phase)

0.1

mild imperfect (3-phase) perfectly bonded (4-phase) mild imperfect (4-phase) severe imperfect (4-phase) perfectly bonded (5-phase) mild imperfect (5-phase) severe imperfect (5-phase)

0.05

completely debonded (5-phase)

0 0.000

0.001

0.002

0.003

0.004

0.005

Strain [ε11]

Figure 3.3 The predicted evolution of volume fractions of perfectly bonded particles various types of damaged particles corresponding to Figure 3.2

Chapter 3. Multi-level modeling of particulate composites

41

3.5 Parametric analysis The Weibull parameter S 0 in Eq. (3.1) is closely related to the strength at the particle-matrix interface B

B

in particulate composites. A parametric analysis is conducted to examine the influence of the Weibull parameter S 0 on the progression of imperfect interface in the composites. As primary parameters governing B

B

the progressive imperfect interface, three representative sets of Weibull parameters are used: S 0 = 1.09×  y , B

B

M = 5; S 0 = 2.18×  y , M = 5; S 0 = 3.27×  y , M = 5. For comparison, a uniaxial tensile simulation is B

B

B

B

conducted using the same material properties and model parameters as used in Chapter 3.4. The present predicted stress-strain responses of particulate composites under uniaxial tension at the four-phase composite state with various S 0 values are shown in Figure 3.4. Figure 3.5 exhibits the B

B

predicted evolution of volume fractions of perfectly bonded particles, particles with mild imperfect

500

Stress [σ11] (MPa)

400

300

200

S3.27 0 = 3.27*σ y B

100

B

B

S2.18 0 = 2.18*σ y B

B

B

S1.09 0 = 1.09*σ y B

B

B

B

B

0 0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

Strain [ε11]

Figure 3.4 The present predicted stress-strain responses of particulate composites under uniaxial tension at the four-phase composite state with various S 0 values B

B

Chapter 3. Multi-level modeling of particulate composites

42

interface, and particles with severe imperfect interface, respectively, as a function of the uniaxial strain corresponding to Figure 3.4. It is clearly shown from Figures 3.4 and 3.5 that if the strength at the particlematrix interface is low (lower S 0 ), most particles are mild imperfect or severe imperfect in their interfaces B

B

in early stage and the composites this parametric analysis that the influence of the Weibull parameter S 0 on B

B

the progression of show a nonlinear stress-strain behavior even in early stage of loading. It can be concluded from imperfect interface and stress-strain behavior of the composites is quite significant; thus, an accurate characterization of S 0 would be essential to realistically predict the progressive imperfect B

B

interface in particulate composites. Another parametric analysis is conducted to evaluate the proposed multi-level elastic damage model sensitivity to the compliance parameters  and  and to address the influence of varying  and  values on the stress-strain response of the model. As parameters controlling the severity of imperfect interface, four

0.3 perfectly bonded (S 0 = 1.09*σ y ) B

m ild im perfect severe im perfect perfectly bonded m ild im perfect

0.25

severe im perfect perfectly bonded m ild im perfect severe im perfect

(S 0 (S 0 (S 0 (S 0 (S 0 (S 0 (S 0 (S 0

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

= 1.09*σ y ) = 1.09*σ y ) = 2.18*σ y ) = 2.18*σ y ) = 2.18*σ y ) = 3.27*σ y ) = 3.27*σ y ) = 3.27*σ y ) B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

Particle Volume Fraction

0.2

0.15

0.1

0.05

0 0.000

0.002

0.004

0.006

0.008

0.010

Strain [ε 11]

Figure 3.5 The predicted evolution of volume fractions of perfectly bonded particles, particles with mild imperfect interface and particles with severe imperfect interface corresponding to Figure 3.4

Chapter 3. Multi-level modeling of particulate composites

43

sets of the compliance parameters for the severe imperfect interface are used:  2 = 2.0,  2 = 3.0;  2  2.0 104 , 2  3.0 104 ;  2  6.0 105 , 2  7.0 105 ;  2  2.0 105 , 2  3.0 105 . For simplicity, fixed values of the compliance parameters for the mild imperfect interface are used as follows: 1  2.0 105 , 1  3.0 105 . The present predicted stress-strain responses of particulate composites with severe imperfect interfaces under uniaxial tension with various values of  2 and  2 are shown in Figure 3.6. As the values of the compliance parameters  2 and  2 become higher, the effect of the severe imperfect interface on the stress-strain behavior of the composites is more pronounced.

350

300

Stress [σ11] (MPa)

250

200

150

100

α2.0,3.0 2 =2.0, β 2 =3.0 α2E-4, 3E-4β 2 =3.E-4 2 =2.E-4, α6E-5,7E-5 =6.E-5, β 2 =7.E-5 2 α2E-5,3E-5 =2.E-5, β 2 =3.E-5 2

50

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

0 0

0.0005

0.001

0.0015

0.002

0.0025

0.003

0.0035

0.004

Strain [ε11]

Figure 3.6 The present predicted stress-strain responses of particulate composites under uniaxial tension at four-phase composite state with various α 2 and β 2 values B

B

B

B

Chapter 3. Multi-level modeling of particulate composites

44

3.6 Experimental comparisons The proposed multi-level elastic damage model is further exercised to predict the behavior of brittle ceramic composites. The uniaxial stress-strain behavior of unidirectional SiC fiber-reinforced CAS glassceramic matrix composites tested by Sørensen (1993) is numerically predicted using the multi-level elastic damage model. The unidirectional fibers in the composites are modeled as spherical particles in this simulation. The elastic properties of the composites used in this simulation are as follows (Sørensen, 1993; Sørensen and Talreja, 1993; Sørensen et al., 1993; Vanswijgenhoven and Van Der Biest, 1997): E 0 = 98 B

B

GPa, 0 = 0.3, E 1 = 200 GPa,  1 = 0.15,  = 0.35. According to the experimentally obtained stress-strain B

B

curve in the longitudinal direction (Sørensen, 1993), the compliance parameters 1 , 1 ,  2 and  2 given in

500

450

400

Stress [σ11] (MPa)

350

300

250

200

150

experiment 100

present prediction (5-phase composite)

50

0 0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

Strain [ε11]

Figure 3.7 The comparison between the present prediction and experimental data (Sørensen, 1993) for overall uniaxial tensile responses of SiC fiber-reinforced CAS glass-ceramic matrix composites

Chapter 3. Multi-level modeling of particulate composites

45

Eq. (3.6) as well as the Weibull parameters S 0 and M in Eq. (3.1) are estimated to be: 1  2.0 105 , B

B

1  3.0 105 ,  2 = 2.0,  2 = 3.0; S 0 = 450 MPa, M = 2. B

B

The present predicted uniaxial stress-strain curve of the composites at the five-phase composite state is compared with the experimentally obtained stress-strain curve in the longitudinal (fiber) direction (Sørensen, 1993) in Figure 3.7. It is noted that the present prediction and the experimental data match well within the small strain region (the corresponding threshold strain  th = 0.005), but the predicted stress-strain curve starts to deviate from the experimentally obtained one after the threshold strain. The experiment in Sørensen (1993) was designed to show the brittle behavior of ceramic composites reinforced with unidirectional fibers. It was shown from Figure 3.7 that the experimentally obtained stress of the composites was increased substantially after the threshold strain due the unidirectional fibers may play the

0.4

perfectly bonded mild imperfect severe imperfect completely debonded

0.35

Particle Volume Fraction

0.3

0.25

0.2

0.15

0.1

0.05

0 0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

Strain [ε11]

Figure 3.8 The predicted evolution of volume fractions of perfectly bonded particles, particles with mild imperfect interface, severe imperfect interface and completely debonded particles corresponding to the present prediction in Figure 3.7

Chapter 3. Multi-level modeling of particulate composites

46

primary role in the increase of stress. On the other hand, inclusions are limited to be spherical particles in our framework and therefore our predictions cannot be compared with his experimental data after the threshold strain. The predicted evolution of volume fractions of particles corresponding to Figures 3.7 is shown in Figures 3.8. To assess and demonstrate the accuracy of the proposed multi-level elastic damage model, the present predictions are compared with experimental data reported by Geiger and Welch (1997) for the uniaxial stress-strain behavior of silicon carbide particulate-reinforced 2009 aluminum matrix composites. The matrix is assumed to behave elastically. We adopt the same material properties for the composites as those in Geiger and Welch (1997) as follows: E 0 = 72.4 GPa, 0 = 0.33, E 1 =450 GPa, 1 = 0.19,  = 0.294. Since B

B

B

B

model parameters of the proposed model were not reported by Geiger and Welch (1997), one needs to

600

500

Stress [σ11] (MPa)

400

300

200

experiment 100

present prediction (5-phase composite state)

0 0

0.002

0.004

0.006

0.008

0.01

0.012

Strain [ε11]

Figure 3.9 The comparison between the present prediction and experimental data (Geiger and Welch, 1997) for overall uniaxial tensile responses of silicon carbide particle-reinforced 2009 aluminum matrix composites

Chapter 3. Multi-level modeling of particulate composites

47

estimate the compliance parameters 1 , 1 ,  2 and  2 given in Eq. (3.6) as well as the Weibull parameters S 0 and M in Eq. (3.1). B

B

Those compliance and Weibull parameters are fitted according to the experimentally obtained stressstrain curve (Geiger and Welch, 1997). The fitted model parameters are: 1  2.0 105 , 1  3.0 105 ,

 2 = 2.0,  2 = 3.0; S 0 = 825 MPa, M = 3. The predicted uniaxial stress-strain curve of the composites at the B

B

five-phase composite state based on the above material properties and parameters is shown in Figure 3.9. The experimentally obtained stress-strain curve is also plotted in the figure for comparison. As a whole, the present prediction and the experimental data match well. The predicted evolution of volume fractions of particles corresponding to Figure 3.9 is shown in Figure 3.10.

0.3

perfectly bonded mild imperfect severe imperfect completely debonded

0.25

Particle Volume Fraction

0.2

0.15

0.1

0.05

0 0

0.002

0.004

0.006

0.008

0.01

0.012

Strain [ε11]

Figure 3.10 The predicted evolution of volume fractions of perfectly bonded particles, particles with mild imperfect interface, severe imperfect interface and completely debonded particles corresponding to the present prediction in Figure 3.9

Chapter 3. Multi-level modeling of particulate composites

48

3.7 Appendix. Parameters Λ 1 , . . , Λ 8 in Eqs. (3.8) - (3.23), and (3.26) - (3.33) B

B

B

B

  1  22  1  1  31  2 1    1  15 1  0   1     2 1  31  2 1 1  2 2 1  31  2 2    2 

15 1  0   1

(3.34)

(3.35)

4 1 1  2 2 

  2  24  1  3  32  2 2  2   3  600a 1  0   2    2  32  2 2 1  24  1  33  24     300a 1  0   2

(3.36)

2

4 

(3.37)

 2 1  24 

  3  26  1  5  33  2 3  2   5  600a 1  0   3     3  33  2 3 1  26 1  35  26    300a 1  0   3

(3.38)

2

6 

(3.39)

 3 1  26 

  4  28  1  7  34  2 4    7  15 1  0   4   2  3   2  1  2  1  3   2        4  4 4 8 7 8   8 

15 1  0   4

(3.40)

(3.41)

4 4 1  28 

in which

1 

  1  51 1  0    5 0  1 ,  2 1    31  2 1 

2 

3 

2 2 2

    2  31  2  2    1  ,  3   2  2 2    

4 

5 

3 2 3

    3  33  2  4    3  ,  3   2  3 3    

6 

7 

  4  54 1  0    5 0  1 ,  2 4   3 4  2 4  

8 

 1  4  5 0  2 1  2 2 2 2

3 4 2 3

 4  4  5 0  2 4

(3.42)

(3.43)

(3.44)

(3.45)

Chapter 3. Multi-level modeling of particulate composites

 1 15 1  0 

1 0  0  1

 1  0 3   1   0   2  1  0 

 2  1200a 1  0 

 3 1200a 1  0 

2

2

49

 5 0  1

(3.46)

 2  0  0  2  1  2  0 3   2   0   2  2  0 

(3.47)

3 0  0  3  3  3  0 3   3   0   2  3  0 

(3.48)

 4  5 0  1 1 

(3.49)

 7  5 0  0  8  10 0  1 2  1  0 

(3.50)

0  2 2  0

(3.51)

 2  600a 1  0 

 3  600a 1  0  4 

2

0

2

3  0

 4

(3.52)

5 0  7 2

(3.53)

with









1  80  480 0  400 02 a  500 1  2 0  0 1  98  140 0  50 02 01



2



 268  1240 0  1300 02 0 1

    80  480  400  a  500 1  2   268  1240  1300   

2  320  720 0  400 02 a  0  31  21   7  5 0 

3

2

0

0

0

2

2 0 2   98  140 0  50 02  0 2

2

0



0



(3.54)

(3.55)

(3.56)

0 2

4  320  720 0  400 02 a  0  3 2  22   7  5 0 

2

(3.57)

50

CHAPTER 4 AN ELASTOPLASIC MULTI-LEVEL DAMAGE MODEL FOR DUCTILE MATRIX COMPOSITES EVOLUTIONARY IMPERFECT INTERFACE

4.1 Introduction Ductile matrix composites (e.g., metal matrix composites, MMCs) reinforced with various shapes of particles, fibers or whiskers have been used in aerospace, electronics and a variety of other engineering applications. In particular, MMCs combine the properties of high strength and high modulus ceramics with those of high ductile metals or alloys to produce enhanced mechanical properties of the composite over metal or alloys (Ravi et al., 2007). Particle or fiber reinforced ductile matrix composites are in general subjected to a number of damage modes on the microscale (Drabek and Bohm, 2004). Debonding phenomenon existed in between particles and the matrix is one of the major damage modes in particle reinforced ductile matrix composites (PRDMCs) and its effect on the mechanical behavior of the composites has to be well addressed for an accurate analysis of the composites. Ju and Chen (1994c) developed a micromechanical formulation to predict the effective elastoplastic behavior of two-phase PRDMCs under arbitrary loading histories by considering the first-order stress perturbations of elastic particles on the ductile matrix. Ju and Tseng (1996, 1997) further improved Ju and Chen's (1994c) work by incorporating second-order stress perturbations due to pairwise particle interactions. Qu (1993a, 1993b) derived the Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface in an elastic matrix of infinite extent where the imperfect interface between the inclusion and the matrix was modeled by a spring layer of vanishing thickness. Lee and Pyo (2007) proposed a micromechanics-based elastic damage model to predict the effective elastic behavior and imperfect interface evolution in particle composites. The Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface (Qu, 1993a, 1993b) was adopted in their formulations to model the imperfect interface. Most recently, Lee and Pyo (in press) proposed a multi-level elastic damage model based on a combination of a micromechanical formulation and a multi-level damage model to predict the effective elastic behavior

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

51

and progressive imperfect interface in particulate (brittle) composites. Their multi-level damage model in accordance with the Weibull's probabilistic function described the sequential, progressive imperfect interface in the composites. The primary objective of the present work is the extension of the framework of Lee and Pyo (in press) to accommodate the elastoplastic behavior and multi-level damage evolution in PRDMCs. Following Lee and Pyo (in press), the four-level elastic damage model, which was illustrated in Figure 3.1, is adopted for a complete description of the sequential progression of imperfect interface in the composite: (1) level-0 of two-phase composite state consisting of a ductile matrix and perfectly bonded particles; (2) level-1 of three-phase composite state consisting of a ductile matrix, perfectly bonded particles and particles with mild imperfect interface; (3) level-2 of four-phase composite state consisting of a ductile matrix, perfectly bonded particles, particles with mild imperfect interface and particles with severe imperfect interface; (4) level-3 of five-phase composite state consisting of a ductile matrix, perfectly bonded particles, particles with mild imperfect interface, particles with severe imperfect interface and completely debonded particles. Since two different damage modes of imperfect interface (mild imperfect interface and severe imperfect interface) are assumed to occur sequentially, imperfect interface would be developed and transformed from one mode to another as deformations or loadings continue to increase. Accordingly, there may be two different modes of imperfect interface (mild imperfect interface and severe imperfect interface) occurring simultaneously at the Level 2 of four-phase composite state. It will be followed by three different modes of imperfect interface (mild imperfect interface, severe imperfect interface and complete debonding) occurring simultaneously at the Level 3 of five-phase composite state at the next step. The two-parameter Weibull statistics is adopted to estimate the volume fraction of particles with different modes of imperfect interface. Particles are assumed to be non-interacting, randomly dispersed, elastic spheres that are initially embedded firmly in the ductile matrix with perfect interfaces. It is also assumed that the progression of imperfect interface is governed by the average internal stresses of perfect bonded particles (phase 1) as well as the Weibull parameters (Weibull, 1951). Numerical examples corresponding to uniaxial, biaxial and triaxial tension loadings are solved to illustrate the potential of the proposed micromechanical framework. A series of parametric analysis are carried out to investigate the influence of model parameters on the progression of imperfect interface in the composites. Furthermore, the present prediction is compared with available experimental data in the literature to further illustrate the elastoplastic damage behavior of the present framework and to verify the proposed elastoplastic multi-level damage model.

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

52

4.2 Ensemble-volume-average homogenization process

4.2.1 Overview The homogenized elastoplastic damage responses of ductile matrix composites with imperfect interfaces are formulated in this section. Specifically, five-phase composites consisting of a ductile matrix (phase 0), perfectly bonded particles (phase 1), particles with mild imperfect interface (phase 2), particles with severe imperfect interface (phase 3) and completely debonded particles (voids) (phase 4) are considered. The von-Mises yield criterion with isotropic hardening law is adopted here for the ductile matrix; thus, the stress σ and the equivalent plastic strain e p at any matrix point satisfy the following yield function (see, e.g., Ju and Lee, 2000; Ju and Ko, in press):





 

F σ, e p  H  σ   K 2 e p  0

(4.1)

 

where K e p denotes the isotropic hardening function of the ductile matrix and H  σ   σ : I d : σ signifies the square of the deviatoric stress norm in which I d denotes the deviatoric part of the fourth-rank B

B

identity tensor. Following the work of Ju and Chen (1994c), Ju and Lee (2000, 2001), Sun et al. (2003), Ju et al. (2006) and Ju and Ko (in press), a micromechanical framework, where an ensemble-averaged yield criterion is constructed for the entire composite and only the first order effects are considered in the formulation of effective plastic response, is employed.

4.2.2 A first-order formulation of the current stress norm of multi-phase ductile matrix composites Following Ju and Chen (1994c), Ju and Tseng (1996) and Ju and Lee (2000), the square of the “current stress norm” at the local point x, which determines the plastic strain in ductile matrix composites for a given phase configuration  , can be written as

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

53

  x|  : I d :   x|  , if x in the matrix; H  |x    0, otherwise. 

(4.2)

The ensemble average of H  |x  over all possible realizations, denoted by H

m

 x ,

can be

obtained by integrating H over all possible particle and void configurations (cf. Ju and Lee, 2000)

H

m

H  x|   H  P   d   H  x|   H  P   d   H  x|   H  P   d    H  x|   H  P   d   

 x   H 0  

0

0

1

1

1

2

2

0

0

3

3

2

3

4

4

(4.3)

4

 

where P  q is the probability density function for finding the q-phase (q = 1, 2, 3, 4) configuration  q in the composites and H 0  σ0 : I d : σ0 is the square of the far-field stress norm in the matrix. Moreover, the total stress at any point x in the matrix can be decomposed into two parts: the far-field stress σ 0 and the perturbed stress σ ' due to the presence of the particles and voids as:

σ  x   σ0  σ '  x 

(4.4)

with σ0  C0 : ε0 where C0 denotes the fourth-rank elasticity tensor of the matrix and ε 0 is the elastic strain field induced by the far-field loading. Following Ju and Chen (1994c) and Ju and Lee (2000), the perturbed stress for any matrix point x due to a typical isolated q-phase inhomogeneity centered at xq(1) reads









 : ε*0 σ ' x|x(1)  C0  G x  x(1) q q   q

(4.5)

* where ε*0 q is the solution of the eigenstrain ε q for the single inclusion problem of the q-phase, and

 

G rˆq 



1 q3H1  q5 H 2 30 1  0 



The components of H 1 and H 2 are given by P

P

P

P

(4.6)

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

54

 

(4.7)

 

(4.8)

H1ijkl rˆq  5Fˆijkl  15,3 0 ,3,3  6 0 , 1  2 0 ,1  2 0  2 Hijkl rˆq  3Fˆijkl  35, 5, 5, 5,1,1

ˆq , q  a / rˆq , a is the radius of a spherical particle or void, and  0 is the where rˆq  x  x(1) q , rˆq  r Poisson's ratio of the matrix material. The components of the fourth-rank tensor Fˆ - which depends on six scalar quantities B 1 , B 2 , B 3 , B 4 , B 5 , B 6 - were previously defined by (Ju and Lee, 2000): B

B

B

B

B

B

B

B

B

B

B

B





Fˆijkl  Bm   B1nˆi nˆ j nˆk nˆl  B2  ik nˆ j nˆl   il nˆ j nˆk   jk nˆi nˆl   jl nˆi nˆk  B3 ij nˆk nˆl



 B4 kl nˆi nˆ j  B5 ij kl  B6  ik  jl   il  jk



(4.9)

with the unit outer normal vector nˆ  rˆq / rˆq and index m = 1 to 6. In addition, the noninteracting eigenstrain ε*0 q in Eq. (4.5) is given by (see, e.g., Ju and Chen, 1994a)



ε*0 q   Aq  Sq



1

: 0,

q  1, 2,3, 4

(4.10)

where S q is the Eshelby's tensor for the q-phase inclusion and the fourth-rank tensor A q is defined as B

B



Aq  Cq  C0

B



1

B

 C0 in which C q signifies the elasticity tensor of the q-phase. B

B

4.2.3 A first-order formulation of effective elastoplastic behavior of multi-phase ductile matrix composites Since a matrix point receives the perturbations from particles and voids, the ensemble-average stress norm for any matrix point x can be evaluated by collecting and summing up all the current stress norm perturbations produced by any typical particle or void centered at x(1) q in the particle or void domains and averaging over all possible locations of x(1) q (cf. Ju and Lee, 2000). Accordingly, H be rephrased as

m

 x  in Eq. (4.3) can

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

H

m

55

(1) (1) 0  x   H 0   xx(1) a H  x|x1(1)   H 0  P  x1(1)  dx1(1)   x x(1) a H  x|x(1) 2   H  P  x 2  dx 2 1



(1) x  x3  a

H  x|x   H  P  x  dx (1) 3

0

(1) 3

2

(1) 3



(1) x  x4  a

H  x|x   H  P  x  dx (1) 4

0

(1) 4

(1) 4

 ...

(4.11)

 

where P x(1) signifies the probability density functions for finding a particle or a void centered at x(1) q . It q

 

is assumed that the probability density functions take the form P x(1)  q

Nq

, where N q is the total B

V

B

numbers of particles or voids dispersed in a representative volume V. Further, Eq. (4.11) can be recast into a more simplified form:

H

m

 x  H 0  

N3 V

N1 V

0 0 2 rˆ1 drˆ1 A(rˆ1) H rˆ1   H  dA  V rˆ 2 drˆ2 A(rˆ 2 ) H rˆ2   H  dA

N

rˆ 3 A(rˆ 3 )  drˆ3



H  rˆ3   H 0 dA 

N4 V

rˆ 4 A(rˆ 4 )  drˆ4



H  rˆ4   H 0 dA  ...

(4.12)

 

where A rˆq is a spherical surface of radius rˆq (q = 1,2,3,4). Using the two identities Eqs. (28) and (29) in Ju and Chen (1994c) and the perturbed stress given in Eq. (4.5), the ensemble-averaged current stress norm at any matrix point can be obtained as:

H

m

 x   σ0 : T(q) : σ0

(4.13)

The components of the positive definite fourth-rank tensor T (q) for the “q”-phase composite state read P



(q) Tijkl  T1( q)ij kl  T2( q) ik  jl  il  jk

P



(4.14)

where T1(5) and T2(5) for the five-phase composite state are given by

2 2   1 802 a 2 1  0   2 802 a 2 1  0   3 4 2  (4.15) 3T1(5)  2T2(5)  50 1  2 0      2 2 2   3  2 2 3   2  3   2  3   2        1 1 2 2 4 4 3 3  

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

T2(5) 

2 2   402 a 2 1  0   2 402 a 2 1  0   3  4  1 1   23  50 0  35 02  2     4 1 2  22  32 4 42   





where the volume fraction for the q-phase is defined as q 

56

(4.16)

4 a3 N q , and the parameters 1 , ... , 3 V

 4 and  1 , ...,  4 are listed in Appendix of Chapter 3. Similarly, T1(2) and T2(2) , T1(3) and T2(3) , T1(4) and T2(4) for the two-phase, three-phase, and four-phase composite states, respectively, read:

3T1(2)  2T2(2)  50 1  2 0 

T2(2) 

1

2

 31  21 2

1 1  23  50 0  35 02 2 4 12





2  1 802 a 2 1  0   2  2  3T1(3)  2T2(3)  50 1  2 0    2   3  2 2  3   2    1 1 2 2  

T2(3)

2   402 a 2 1  0   2  1 2  1    23  50 0  35 0   4 12  2  22  

3T1(4)



 2T2(4)

T2(4) 

 50 1  2 0 

2

(4.18)

(4.19)



(4.20)

2 2  1 802 a 2 1  0   2 802 a 2 1  0   3      2 2   3  2 2  3   2  3   2     3 1 1 2 2 3  

(4.21)

2 2   402 a 2 1  0   2 402 a 2 1  0   3  1 1   23  50 0  35 02  2    4 1  2  22  32  



(4.17)



(4.22)

Following Ju and Chen (1994c), the relation between the far-field stress σ 0 and the macroscopic stress σ takes the form

σ 0  P( q ) : σ

where the components of P( q ) for the “q”-phase composite state are

(4.23)

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites



(q) Pijkl  P1( q)ij kl  P2( q) ik  jl  il  jk

57



(4.24)

with

3P1( q )  2 P2( q )  P2( q ) 



1

(4.25)

1  1( q )

1

2 1  2( q )

(4.26)



and the coefficients 1(5) and 2(5) for the five-phase composite state read:

101 1  2 0 

1(5) 

31  2 1



2(5)







 2 31  2  2  1200a 1  0 

2

32  2 2

10 4 1  2 0 

   3  2 3

3

4

 1200a 1  0 

33  2 3

2

 (4.27)

34  2 4



1  7  5 0  2 600a 1  0    2 2 1



2

2

   600a 1   3

0

3

2

 4

   7  5  4

0

2 4

(4.28)

where the parameters 1 , ..,  4 are listed in Appendix of Chapter 3. Similarly, the coefficients 1(2) and 2(2) ,

1(3) and 2(3) , 1(4) and 2(4) for the two-phase, three-phase, four-phase composite states, respectively, read:

1(2) 

2(2) 

1(3)



2(3) 

101 1  2 0 

(4.29)

31  2 1

1  7  5 0 

(4.30)

2 1 101 1  2 0  31  2 1





 2 31  2  2  1200a 1  0  32  2 2



1  7  5 0  2 600a 1  0    2 2 1



2

2



2



(4.31)

(4.32)

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

1(4)

2(4)

101 1  2 0 



31  2 1





 2 31  2  2  1200a 1  0  32  2 2



1  7  5 0  2 600a 1  0    2





2 1

2

2

2

   3  2 3

3

58

4

 1200a 1  0 

2



33  2 3

   600a 1   3

0

2

 4



3

(4.33)

(4.34)

By combining Eqs. (4.13) and (4.23), the ensemble-averaged current stress norm in a matrix point can be rephrased as:

H

m

 x   σ : T(q) : σ

(4.35)

where the positive definite fourth-rank tensor T( q ) for the “q”-phase composite state is defined as

T

T( q )   P( q )   T( q )  P( q)  

(4.36)

and can be shown to be



(q) Tijkl  T1( q)ij kl  T2( q) ik  jl  il  jk



(4.37)

where

3T1( q )  2T2( q ) 

T2( q ) 



T2( q )

1  2( q )

3T1( q )  2T2( q )

1    (q) 1



2

2

(4.38)

(4.39)

Following Ju and Lee (2000), the effective yield function for the multi-phase ductile matrix composites can be proposed as

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

 

F  1  1   : T :   K 2 e p 2

59

(4.40)

where 1 is the current (perfectly bonded) particle volume fraction. The effective equivalent plastic strain rate for the composites is defined as (Ju and Lee, 2000)

ep 

2 p 1  :T : 3

p



 2 1  1



2



2 σ:T:σ 3

(4.41)

in which  denotes the plastic consistency parameter. The  together with the yield function F must obey the Kuhn-Tucker loading/unloading conditions. In our derivations, the simple power-law type isotropic hardening function is employed (see also Ju and Lee, 2000):

 

K2 e p 

  

2 p  y  h e 3

q

(4.42)

where  y denotes the initial yield stress, and h and q are the linear and exponential isotropic hardening parameters, respectively, for the multi-phase composites. Furthermore, a four-level damage model in the order of sequence of progressive imperfect interface as illustrated in Figure 3.1 is employed here to model the sequential, progressive imperfect interface in ductile matrix composites (see detail descriptions in Chapter 3.).

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

60

4.3 Numerical simulations To show the potential of the proposed micromechanical framework, we first predict the behavior of silicon-carbide particle reinforced 6061-T6 aluminum alloy matrix composites under uniaxial tension loading ( 11  0 and  ij  0 for all other stress components). The elastic properties of the composites are adopted according to Zhao and Weng (1995), Ju and Lee (2001) and Lee and Pyo (2007) as: E 0 = 68.3 B

B

GPa, 0 = 0.33, E 1 = 490 GPa,  1 = 0.17,  = 0.2. Following the parametric analysis on the compliance B

B

parameters  and  reported in preceding Chapters, the compliance parameters of the mild imperfect interface, denoted by 1 and 1 , and the severe imperfect interface, denoted by  2 and  2 , in Eq. (3.6) are chosen to be 1  2.0 105 , 1  3.0 105 ;  2 = 2.0,  2 = 3.0. Moreover, the plastic parameters and

600

500

Stress [σ11] (MPa)

400

300

200

2-phase composite state 3-phase composite state

100

4-phase composite state 5-phase composite state 0 0.000

0.002

0.004

0.006

0.008

0.010

Strain [ε11]

Figure 4.1 The present predicted elastoplastic stress-strain responses of particulate composites under uniaxial tension

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

0.2

0.2

a

b

0.15

0.15

Particle Volume Fraction

Particle Volume Fraction

61

perfectly bonded

0.1

mild imperfect

0.1

perfectly bonded mild imperfect severe imperfect

0.05

0.05

0 0.000

0 0.000

0.002

0.004

0.006

0.008

0.010

0.002

0.004

0.006

0.008

0.010

Strain [ε 11 ]

Strain [ε11]

0.2

c

Particle Volume Fraction

0.15

0.1

perfectly bond mild imperfect severe imperfect 0.05

perfectly debonded

0 0.000

0.002

0.004

0.006

0.008

0.010

Strain [ε11]

Figure 4.2 (a) The predicted evolution of volume fractions of perfectly bonded particles and particles with mild impefect interface of the composites at the three-phase composite state; (b) The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles at the four-phase composite state; (c) The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles at the five-phase composite state corresponding to Figure 4.1

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

62

Weibull parameters are assumed to be  y = 400 MPa, h = 220 MPa, q = 0.45; S 0 = 380 MPa, M = 5. B

B

Four different phase composite states as illustrated in Figure 3.1 are considered in this simulation to examine the effect of the level of damage on the elastoplastic stress-strain response of the composites. The present predicted elastoplastic stress-strain responses of the particulate composites under uniaxial tension considering the four different phase composite states are plotted in Figure 4.1. Figure 4.2(a) exhibits the predicted evolution of volume fractions of perfectly bonded particles and particles with mild imperfect interface of the composites at the three-phase composite state versus uniaxial strains, which is corresponding to Figure 4.1. The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles at the four-phase and five-phase composite states versus uniaxial strains corresponding to Figure 4.1 are exhibited in Figures 4.2(b) and 4.2(c), respectively. In Figures 4.1 and 4.2, the elastoplastic stress-strain responses corresponding to various phase

600

500

Stress [σ22(=σ33)] (MPa)

400

300

200

2-phase composite state 3-phase composite state

100

4-phase composite state 5-phase composite state 0 0

0.002

0.004

0.006

0.008

0.01

Strain [ε22(=ε33)]

Figure 4.3 The present predicted elastoplastic stress-strain responses of particulate composites under biaxial tension

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

63

composite states are shown to be bounded by the responses of the two-phase and five-phase composite states. It is observed that the responses with the higher phase composite states are lower than those with the lower phase composite states. It is also noted from the figures that a smooth and gradual transition from the perfectly bonded interface to the perfectly debonded interface occurs within the present framework. In the present numerical simulations corresponding to Figure 4.2, the lower value of the Weibull parameters, which relates to the interfacial strength, (S 0 = 380 MPa and M = 5.0) is used to clearly show B

B

the effect of imperfect interface on the stress-strain behavior of imperfect particulate composites. Experimental characterization of the parameters is, however, required for more accurate prediction of imperfect interface evolution. Details of the relationship between the parameter and interfacial strength can be found in Lee (2001). The proposed multi-level elastic damage model is further exercised to predict the behavior of the composites under biaxial tension (  22   33  0 and  ij  0 for all other components) and triaxial tension

0.2

Particle Volume Fraction

0.15

0.1

perfectly bonded (3-phase) mild imperfect (3-phase) perfectly bonded (4-phase) mild imperfect (4-phase) severe imperfect (4-phase) perfectly bonded (5-phase) mild imperfect (5-phase) severe imperfect (5-phase) completely debonded (5-phase)

0.05

0 0.000

0.002

0.004

0.006

0.008

0.010

Strain [ε22(=ε33)]

Figure 4.4 The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles corresponding to Figure 4.3

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

64

(  22   33  11  0 and  ij  0 for all other components) loading cases. The proportional loading ratio Γ is chosen to be 0.6 for the axisymmetric triaxial tension in this simulation. The predicted stress-strain responses of the composites considering the four different phase composite states under biaxial tension and triaxial tension loading cases are exhibited in Figures 4.3 and 4.5, respectively. The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles under biaxial tension loading corresponding to Figure 4.3 are shown in Figure 4.4. Figure 4.6 exhibits the predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles under triaxial tension loading corresponding to Figure 4.5. It is observed from Figures 4.1, 4.3 and 4.5 that triaxial tension results in a higher elastoplastic stress-strain behavior, which corresponds to findings from Ju and Lee's (2000) and Lee and Pyo's (2007) observations.

1400

2-phase composite state 1200

3-phase composite state 4-phase composite state

Stress [σ11(=σ22/Γ=σ33/Γ)] (MPa)

1000

5-phase composite state

800

600

400

200

0 0.000

0.002

0.004

0.006

0.008

0.010

Strain [ε11]

Figure 4.5 The present predicted elastoplastic stress-strain responses of particulate composites under triaxial tension (Γ = 0.6)

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

65

0.2

Paricle Volume Fraction

0.15

perfectly bonded (3-phase) mild imperfect (3-phase)

0.1

perfectly bonded (4-phase) mild imperfect (4-phase) severe imperfect (4-phase) perfectly bonded (5-phase) mild imperfect (5-phase) 0.05

severe imperfect (5-phase) completely debonded (5-phase)

0 0.000

0.002

0.004

0.006

0.008

0.010

Strain [ε11]

Figure 4.6 The predicted evolution of volume fractions of perfectly bonded particles and various types of damaged particles corresponding to Figure 4.5

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

66

4.4 Parametric analysis A series of parametric analysis are conducted to examine the influence of the Weibull parameters S 0 B

B

and M and the compliance parameters  and  on the elastoplastic behavior and progression of imperfect interface in PRDMCs. First, following Lee's (2001) investigation on the Weibull parameters, three sets of Weibull parameters that are closely related to the strength at the particle-matrix interface in the PRDMCs are used: S 0 = 1.09×  y , M = 5; S 0 = 2.18×  y , M = 5; S 0 = 3.27×  y , M = 5. For simplicity, a uniaxial B

B

B

B

B

B

tensile simulation is conducted using the same elastic properties as used in Section 4.3. The compliance parameters and the plastic parameters are assume to be 1  2.0 105 , 1  3.0 105 ,  2  2.0 ,

2  3.0 ;  y = 250 MPa, h = 220 MPa, q = 0.45.

400

350

300

Stress [σ11] (MPa)

250

200

150

S3.27 0 = 3.27*σ y

100

B

B

B

S2.18 0 = 2.18*σ y B

B

B

B

S1.09 0 = 1.09*σ y B

50

B

B

2-phase composite compositesstate state

0 0.000

0.002

0.004

0.006

0.008

0.010

Strain [ε11]

Figure 4.7 The present predicted elastoplastic stress-strain responses of particulate composites under uniaxial tension at the four-phase composite state with various S 0 values B

B

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

67

Figure 4.7 shows the predicted elastoplastic stress-strain responses of the PRDMCs under uniaxial tension at the four-phase composite state with various S 0 values. It is clear from the figure that higher B

B

interfacial strength parameter S 0 leads to higher elastoplastic stress-strain response. It is also found from B

B

this parametric analysis that the influence of the Weibull parameter S 0 on the elastoplastic response is quite B

B

influential. Thus, the Weibull parameter S 0 needs to be experimentally characterized via the measurement B

B

of the interfacial strength of PRDMCs. The methodology for determining model parameters including the Weibull parameters can be found in Chapter 3. A parametric analysis conducted to determine lower and upper limits of the Weibull parameters can also be found in Lee (2001). To evaluate the proposed elastoplastic multi-level damage model sensitivity to the compliance parameters  and  , a parametric analysis of  and  is also carried out. Following Chapter 3, four sets of the compliance parameters for the severe imperfect interface are used:  2  2.0 , 2  3.0 ;  2  2.0 104 ,

300

250

Stress [σ11] (MPa)

200

150

100

α2.0,3.0 2 =2.0, β 2 =3.0 α2E-4, 2 =2.E-4, 3E-4β 2 =3.E-4 α6E-5,7E-5 2 =6.E-5, β 2 =7.E-5 α2E-5,3E-5 2 =2.E-5, β 2 =3.E-5

50

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

0 0

0.002

0.004

0.006

0.008

0.01

Strain [ε11]

Figure 4.8 The present predicted elastoplastic stress-strain responses of particulate composites under uniaxial tension at the four-phase composite state with various α 2 and β 2 values B

B

B

B

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

68

2  3.0 104 ;  2  6.0 105 , 2  7.0 105 ;  2  2.0 105 , 2  3.0 105 . The compliance parameters for the mild imperfect interface are fixed as follows: 1  2.0 105 , 1  3.0 105 . Moreover, the Weibull parameters and plastic parameters are assumed to be S 0 = 1.09×  y , M = 5;  y = 250 MPa, h B

B

= 220 MPa and q = 0.45. The predicted stress-strain responses of PRDMCs with severe imperfect interfaces under uniaxial tension with various values of  2 and  2 are shown in Figure 4.8. It is seen from the figure that the effect of the severe imperfect interface on the elastoplastic stress-strain behavior of the composites become more noticeable as  2 and  2 become higher.

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

69

4.5 Experimental comparison To assess the predictive capability of the proposed elastoplastic multi-level damage model, the present prediction is compared with the experimental data on SiC particulate-reinforced 5456 aluminum alloy matrix composites reported by Papazian and Adler (1990). We adopt the material properties of the composites as follows (Papazian and Adler, 1990; Bonfoh and Lipinski, 2007): E 0 =73 GPa,  0 = 0.33, E 1 B

B

B

B

=485 GPa,  1 = 0.2,  = 0.2,  y = 230 MPa. Since the model parameters of the proposed model were not reported by Papazian and Adler (1990), the model parameters are estimated by fitting the experimentally obtained stress-strain curve (Papazian and Adler, 1990) with the prediction. The fitted model parameters are: 1  2.0 105 , 1  3.0 105 ,  2  2.0 , 2  3.0 ; h = 1,350 MPa and q = 0.24; S 0 = 530 MPa, M B

B

500

450

400

Stress [σ11] (MPa)

350

300

250

200

150

experiment

100

present prediction (5-phase composite) 50

0 0

0.002

0.004

0.006

0.008

0.01

0.012

0.014

Strain [ε11]

Figure 4.9 The comparison between the present prediction and experimental data (Papazian and Adler, 1990) for overall uniaxial tensile responses of SiC particulate-reinforced 5456 aluminum alloy matrix composites

Chapter 4. Elastoplastic multi-level modeling for ductile matrix composites

70

= 5. Figure 4.9 shows the comparison between the present prediction and the experimental data on the overall elastoplastic stress-strain response of the composites at the five-phase composite state on the based of the above material properties and parameters. Overall, the present prediction and the experimental data match well. The predicted evolution of volume fractions of particles corresponding to Figure 4.9 is shown in Figure 4.10.

0.2

Particle Volume Fraction

0.15

0.1

perfectly bonded mild imperfect severe imperfect completely debonded 0.05

0 0

0.002

0.004

0.006

0.008

0.01

0.012

0.014

Strain [ε11]

Figure 4.10 The predicted evolution of volume fractions of perfectly bonded particles, particles with mild imperfect interface, severe imperfect interface and completely debonded particles corresponding to the present prediction in Figure 4.9

71

CHAPTER 5 A MICROMECHANICS-BASED EVOLUTIONARY DAMAGE MODEL FOR CONTINUOUS, FIBER-REINFORCED BRITTLE MATRIX COMPOSITES WITH MICROCRACKS AND IMPERFECT INTERFACES

5.1 Introduction The mechanical behavior of continuous, fiber-reinforced brittle matrix composites (e.g., ceramic matrix composites, CMCs) is shown to be affected by several factors such as fiber bridging, crack branching, microcracking, delamination, fiber-matrix debonding, fiber slip, and fiber fracture, etc. (Miller et al., 2001). The fiber/matrix interfacial debonding usually dominates the damage behavior at a lower level of loading, and will be followed by the subsequent nucleation of microcracks as the load or displacement continues to increase (Lee et al. 2004). Many studies have been conducted to understand the influence of the fiber/matrix interface and matrix cracking on the mechanical behavior of the composites (e.g., Ahn et al., 1998; Curtin et al., 1998; Paar et al., 1998; Lamon, 2001; González and LLorca, 2007). One-step or multi-step damage models based on the Weibull probabilistic function have been proposed to predict the progressive interfacial debonding and its effect on the mechanical behavior of a wide range of composites (Zhao and Weng, 1995, 1996, 1997; Liang et al., 2006; Lee and Kim, 2007; Ju and Ko, in press, etc.). Another approach has also been proposed to model the interfacial behavior of particle or fiber-reinforced composites with imperfect interface (Qu, 1993a, 1993b; Jun and Haian, 1997; Lee and Pyo, 2007, in press). In their derivations, the imperfect interface was characterized by a spring-type model assuming that the traction continuity remains intact, while the displacement experienced a jump proportional to the interfacial traction (Qu, 1993a, 1993b; Lee and Pyo, 2007, in press). Crack nucleation is another important damage mechanism affecting the mechanical behavior of brittle matrix composites. The (stochastic) crack nucleation process is related to the distribution of the weak sites in the composites and the microcracks initiate uniformly across the material volume (Herrmann and Roux, 1990; Isichenko, 1992; Chakrabarti and Benguigui, 1997; Krajcinovic and Vujosevic, 1998; Lee and

Chapter 5. Evolutionary damage model for fiber-reinforced composites

72

Simunovic, 2006), allowing us to model the nucleation process using a simple scalar field that implies uncorrelated crack position and orientation (Lee and Simunovic, 2006). Lee and Simunovic (2006) developed a constitutive model based on a combination of a fracture-mechanics based model and micromechanical formulations for the prediction of the crack evolution and effective mechanical behavior of damage-tolerant brittle composites. Liang et al. (2006) employed a continuum damage model proposed by Karihaloo and Fu (1989) to simulate the nucleation of microcracks in laminated composites. In the present study, a three-dimensional (3D) micromechanics-based evolutionary damage model is proposed to predict the effective elastic behavior of continuous, fiber-reinforced brittle matrix composites with microcracks and imperfect interfaces. The Eshelby's tensor for a circular cylindrical inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted to model continuous fibers with imperfect interfaces and is incorporated into the micromechanical framework. The progression of imperfect interfaces is assumed to be governed by the average internal stresses of the fibers as well as the Weibull parameters (Weibull, 1951), and nucleation of microcracks is simulated by employing the continuum damage model proposed by Karihaloo and Fu (1989). The multi-level elastic damage model proposed by Lee and Pyo (in press) is considered for a complete description of the sequential progression of imperfect interface in the composites. In the present derivation, continuous fibers are modeled to be elastic cylindrical inclusions that are randomly dispersed yet unidirectionally aligned in an elastic matrix. Microcracks are assumed to be located perpendicular to the fibers and the size of microcracks is uniform (see Figure 5.1). Numerical examples

1

3 2 Figure 5.1 The schematic of a continuous, fiber-reinforced composite subjected to uniaxial loading in the longitudinal direction

Chapter 5. Evolutionary damage model for fiber-reinforced composites

73

corresponding to uniaxial loadings in the longitudinal and transverse directions are solved to illustrate the potential of the proposed damage model. A series of parametric analysis are also carried out to address the influence of model parameters on the elastic damage behavior of the present framework. Comparisons between the present prediction with available experimental data are conducted to verify the accuracy of the proposed damage model.

Chapter 5. Evolutionary damage model for fiber-reinforced composites

74

5.2 Effective elastic moduli of continuous, fiber-reinforced brittle matrix composites with microcracks and imperfect interfaces Let us consider an initially perfectly bonded three-phase composite consisting of an elastic matrix (phase 0) with bulk modulus  0 and shear modulus 0 , and randomly located unidirectionally aligned elastic circular fibers (phase 1) with bulk modulus 1 and shear modulus 1 , and penny-shaped microcracks (phase 5) of radius c (see Figure 5.1). The penny-shaped microcrack is modeled as an oblate spheroidal void (a 1 = a 2 >> a 3 ) and aspect ratio a 3 /a 1 →0 as shown in Figure 5.2. The fibers are assumed B

B

B

B

B

B

B

B

B

B

to be non-interacting and initially embedded firmly in the matrix with perfect interface. As loads or deformations continue to increase, some initially perfectly bonded fibers are transformed to fibers with mild imperfect interface (phase 2), some fibers with mild imperfect interface are then transformed to fibers with severe imperfect interface (phase 3), and all fibers are transformed to completely debonded fibers that are regarded as cylindrical voids (phase 4), and microcracks are nucleated asymptotically within the proposed framework. The schematic of this transition is shown in Figure 5.3.The effective stiffness tensor

C* for multi-phase, linear elastic composites containing arbitrarily non-aligned and/or dissimilar inclusions can be derived as (Ju and Chen, 1994a)

a2 B

a3 B

B

a1 B

B

B

Figure 5.2 The schematic of a penny-shape crack modeled as an oblate spheroidal void (a 1 = a 2 >> B

a 3 ) and aspect ratio a 3 /a 1 →0 B

B

B

B

B

B

B

B

B

Chapter 5. Evolutionary damage model for fiber-reinforced composites

5 1    1 1 C*  C0  I   r  A r  Sr    I  r S r  A r  S r        r 1  

75

(5.1)

where Cr is the elasticity tensor of the r-phase, “  ” is the tensor multiplication, I is the fourth-rank identity tensor, and  r denotes the volume fraction of the r-phase inclusion, and S r signifies the Eshelby's tensor for the r-phase. The fourth-rank tensor A r is defined as

a

b

Matrix (phase 0)

Fiber with mild imperfect interface (phase 2)

Perfectly bonded fiber (phase 1) Penny-shape microcrack (phase 5)

d

Void (phase 4)

c

1

2

Fiber with severe imperfect interface (phase 3)

Figure 5.3 The schematic of multi-level damage transition in a continuous, fiber-reinforced composite: (a) Level 0 of three-phase composite state (initial state); (b) Level 1 of three-phase composite state; (c) Level 2 of five-phase composite state; (d) Level 3 of six-phase composite state

Chapter 5. Evolutionary damage model for fiber-reinforced composites

76

1

Ar  Cr  C0   C0

(5.2)

The fourth-rank Eshelby's tensor S1 for perfectly bonded cylindrical fibers and S 4 for completely debonded cylindrical voids embedded in an isotropic linear elastic and infinite matrix can be derived as (Ju et al., 2006)

 S1,4 ijkl  4 1 1  SIK(1) ij  kl  SIJ(2) ik  jl  il  jk 

(5.3)

0

where  0 denotes the Poisson's ratio of the matrix and  ij signifies the Kronecker delta. The second-rank (1) tensor S IJ(2) is symmetric, while the second-rank tensor S IK is not symmetric. Details of the definition of (1) the second-rank tensors S IK and S IJ(2) can be found in Ju and Sun (1999). The components of the second(1) rank tensors S IK and S IJ(2) are given by (Ju et al., 2006)

(1) (1) (1) (1) S11  S22  S12  S21 

1  4 0  1 , 2

(1) (1) S13  S23  2 0 ,

(2) (2) (2) (2) (1) (1) (1)  S22  S12  S21  S31  S32  S33  0 , S11 (2) (2) (2) (2) S13  S23  S31  S32  1  0  ,

1  3  4 0  , 2

(5.4)

(2) S33 0

The Eshelby's tensor for a circular cylindrical inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted to model continuous fibers with imperfect interface. The Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface can be written as (Qu, 1993a, 1993b)

M Sijkl 







1 O O O O Sijkl (x) dV (x)  Sijkl  Iijpq  Sijpq H pqrs Lrsmn I mnkl  Smnkl  



(5.5)

where Ω denotes an ellipsoidal subdomain in a homogeneous and linearly elastic solid, L signifies the O fourth-rank elasticity tensor, Sijkl is the original Eshelby's tensor and H is given by

Chapter 5. Evolutionary damage model for fiber-reinforced composites

H pqrs 





1 nik n j nl  n jk ni nl  nil n j nk  n jl ni nk dS 4 S

77

(5.6)

In case of circular cylindrical inclusions (a 1 = a 2 = a, a 3 →∞), H ijkl can be simplified as (Qu, 1993b) B

B

B

B

B

B

B

B

Hijkl   Pijkl       Qijkl

(5.7)

in which

P1111  P2222  4P2323  4P1313  2P1212 

3 , 8a

Q1111  Q2222  3Q1122  3Q2211  3Q1212 

9 , 32a

(5.8)

others = 0

By carrying out the lengthy algebra, the Eshelby's tensors S l+1 for a circular cylindrical inclusion with B

B

imperfect interface embedded in an isotropic linear elastic and infinite matrix in Eq. (5.5) can be recast as

 Sl 1 ijkl 

1 256a 1  0 

2

S

M (2l 1)  ij IK



 kl  S IJM (2l )  ik  jl   il  jk 

(5.9)

where l = 1, 2 in which 1 indicates the mild imperfect interface stage, while 2 denotes the severe imperfect M (2l 1) interface stage. In addition, the components of the second-rank tensors S IK and S IJM (2l ) are given by



M (2l 1) M (2l 1) M (2l 1) M (2l 1) S11  S22  S12  S21  32a 1  5 0  4 02



 30 l  24 1  2 0  0 l  3 7  32 0  32 02 2

M (2l 1) S13



M (2l 1) S23



M (2l ) S22



0 l

 128a 0 1  0   48 1  2 0  0 1  2 0  0  0 0  ,

M (2l 1) M (2l 1) M (2l 1) S31  S32  S33 0, M (2l ) S11

   ,

M (2l ) S12



M (2l ) S21

 32a



3  7 0  4 02

  3

0

(5.10)

l  l  ,

M (2l ) M (2l ) M (2l ) M (2l ) S13  S23  S31  S32  4 1  0  16a  30l  , 2

M (2l ) S33 0

Chapter 5. Evolutionary damage model for fiber-reinforced composites

78

Following Sun and Ju (2004), the fourth-rank Eshelby's tensor S 5 for an oblate spheroidal void (a 1 = B

B

B

B

a 2 >> a 3 ) is employed to model penny-shaped microcracks (the aspect ratio a 3 /a 1 → 0) as B

B

B

B

B

B

B

B

 S5 ijkl  SIK(3) ij  kl  SIJ(4) ik  jl  il  jk 

(5.11)

(3) The components of the second-rank tensors S IK and S IJ(4) are given by

 4 2  1 (3) (3) (3) (3) S11  S22  S12  S21   2 0   4  2 1 





 2  g    ,  2  2 1 





 2 2  1  2 2 (3) (3) , S13  S23   2 0  2  g    2   1   1   2 2  1  2 2 (3) (3) S31  S32   4 0  2 ,  g    4 0  2   1   1   2  4 (3) , S33   4 0  2  g    4 0  2  1 3  1 



(5.12)



 4 2  7 (4) (4) (4) (4) S11  S22  S12  S21   2 0   4  2 1 





 2  g    ,  2  2 1 





  2  2 2 (4) (4) (4) (4) S13  S23  S31  S32   0  2 ,  g    2 0  2   1  1    4 2  2  12 2  8 (4) S33   4 0  2  g    4 0    1  3  2 1 





where γ denotes the aspect ratio of the spheroidal void and g(γ) is given by (Sun and Ju, 2004) 1/ 2      1  2  cos 1   ,   1  3/ 2     1   2 g ( )   1/ 2     cosh 1     2  1  ,   1  3/ 2  2      1

















(5.13)

Chapter 5. Evolutionary damage model for fiber-reinforced composites

79

Substituting Eqs. (5.3), (5.9) and (5.11) into Eq. (5.1) yields the effective stiffness tensor C* for the six-phase unidirectional composite as





(1) C*  CIK ij kl  CIJ(2)  ik  jl  il  jk



(5.14)

where 3

(1) (2) (1) (1) , CIK  20  KK  20  IK   0  nK n 1

(2) CIJ(2)  2 0  IJ

(5.15)

(1) in which 0 denotes the Lame constant of the matrix, and the components  IK and  IJ(2) are listed in

Appendix. Accordingly, the stress-strain relation for the (unidirectional) continuous, fiber-reinforced brittle matrix composites can be written as

(2) C (1)  2C11  11   11 (1)    C21   22   (1)   33   C31   0   23      0  13    12   0 

(1) C12

(1) C13

0

0

(1) (2) C22  2C22

(1) C23

0

0

0

0

(1) C32

(1) C33

(2)  2C33

0

0

(2) C23

0

0

0

0

(2) C13

0

0

0

0

0     11  0      22  0    33    0   2 23    0   213    (2)   212  C12 

(5.16)

Chapter 5. Evolutionary damage model for fiber-reinforced composites

80

5.3 Damage modeling The multi-level elastic damage model proposed by Lee and Pyo (in press) is considered here for a complete description of the sequential progression of imperfect interface in the composites. The probability of imperfect interface is modeled as a two-parameter Weibull process and the average internal stresses of perfect bonded fibers (phase 1) are chosen to be the controlling factor of the Weibull function as

      M      2   1  exp    11 1      S0       

(5.17)

      M      3   2 1  exp    11 1      S0       

(5.18)

      M      4   3 1  exp    11 1      S 0       

(5.19)

3  3   4 ,

(5.20)

 2   2  3 ,

1     2

where  is the original particle volume fraction,

11 1 is the internal stress of fibers (phase 1) in the 1-

direction, the subscript  1 denotes the particle phase, and S 0 and M are the Weibull parameters. In case of B

the uniaxial loading in the longitudinal direction,

B

11 1 is replaced by  33 1 .

The internal stresses of fibers required for the initiation of the imperfect interface were explicitly derived by Ju and Lee (2000) as

n 1  1  1  C1   I  S1   A1  S1    I  r S r   A r  S r      r 1  

1

:ε  U : ε

(5.21)

By carrying out the lengthy algebra, the components of fourth-rank tensor U for the present model are explicitly given by



(1) Uijkl  U IK ij  kl  U IJ(2) ik  jl  il  jk



(5.22)

Chapter 5. Evolutionary damage model for fiber-reinforced composites

81

(1) where the second-rank tensors U IK and U IJ(2) can be expressed as

(1) U IK 

(1) EIK

1  2 (2) KK



2 EII(2)WIK 1  2 (2) II

3



(1) EIn WnK

(2) n 1 1  2 nn

U IJ(2) 

,

EIJ(2)

(5.23)

1  2 (2) IJ

in which

(1) EIK

 2 (2)  1 1  KK  1 

(1)  21 IK 1 3 (1)   nK ,   1 1 n 1 

 2 (2) EIJ(2)  1 1  IJ  1 

(1) (2) (1)   (1)  31 WI 1  1  11  211 21    (1) (1) (2) (1) 12 1   22  2 22  32  WI 2    W   (1) (1) (2)  13  (1) 1   33  2 33  I 3   23 





(1) (3) (5) (7) (1) IK    IK   IK   IK   IK ,

1

 ,  

 (1)  I1  (1)   I 2  ,  (1)    I 3 



(2) (4) (6) (8) (2) IJ    IJ   IJ   IJ   IJ

(5.24)



(1) where the subscript I = 1, 2, 3 and the components  IK , ...,  IJ(8) in Eq. (5.24) are listed in Appendix.

In case of tensile loading, the averaged internal stresses of fibers can be obtained as

(1) (2) (1) (1)   11   U11  2U11 U12 U13 1       (1) (1) (2) (1) U 21 U 22  2U12 U 23   22 1       (1) (1) (1) (2)  U 31 U 32 U 33  2U 33   33 1    

1

 11     22     33 

(5.25)

In accordance with the continuum damage model for the nucleation of microvoids or microcracks in brittle materials (Karihaloo and Fu, 1989), the density of nucleated microcracks in the composite is modeled using the following isotropic scalar function (Liang et al., 2006)

 v 0 ,  5     th v 0  c1 1  a   

 a   th c

2 a th  ,    

(5.26)

Chapter 5. Evolutionary damage model for fiber-reinforced composites

82

where v 0 is the initial density of microcracks.  th   ijth ijth is the threshold effective strain below which no nucleation occurs, c 1 and c 2 are material (nucleation) parameters depending on the specific shape and B

B

B

B





1/ 2

2 2 2 2 2 2  distribution of microcracks, and  a   ij  ij  11   22   33  2  23  13  12  

accumulated effective strain (Karihaloo and Fu, 1989).

is the current

Chapter 5. Evolutionary damage model for fiber-reinforced composites

83

5.4 Numerical simulations To illustrate the elastic damage behavior of the present framework and to address the effect of the Weibull parameter S 0 on the evolution of imperfect interface in uncracked, continuous, fiber-reinforced brittle B

B

matrix composites (five-phase composite state), we first consider a uniaxial tension in the longitudinal direction case (  33  0 and  ij  0 for all other stress components). We adopt the material properties for the Nicalon fiber-reinforced Barium Magnesium Aluminosilicate (BMAS) glass ceramic composite in accordance with Dannemann and Mandell (1994) in this simulation: E m = 106 GPa,  m = 0.222, E f = 200 B

B

B

B

GPa,  f = 0.299 and  = 0.4, where the subscripts m and f represent the matrix and fibers, respectively. The compliance parameters 1 , 1 ,  2 and  2 given in Eq. (5.10) are assumed to be 1  2.0  107 ,

300

250

Stress [σ33] (MPa)

200

150

100

So S 0 == 100 100MPa MPa B

50

B

B

So S 0 == 200 200MPa MPa B

B

S 0 == 300 300MPa MPa So B

B

0 -0.003

-0.002

[ε 11 ] B

B

-0.001

0 Strain

0.001

0.002

0.003

0.004

0.005

[ε 33 ] B

B

Figure 5.4 The present predicted stress-strain responses of unidirectional composites under uniaxial tension in the longitudinal direction with various S 0 values B

B

Chapter 5. Evolutionary damage model for fiber-reinforced composites

84

1  3.0 107 ,  2  2.0 , 2  3.0 . Moreover, three different sets of Weibull parameters are used to evaluate the proposed elastic damage model sensitivity to the Weibull parameter S 0 : S 0 = 100 MPa and M B

B

B

B

= 2; S 0 = 200 MPa and M = 2; S 0 = 300 MPa and M = 2. B

B

B

B

The present predicted stress-strain responses of the composites with imperfect interfaces under uniaxial tension in the longitudinal direction with various S 0 values are shown in Figure 5.4. Figure 5.5 B

B

exhibits the evolutions of perfectly bonded fibers, fibers with mild imperfect interface, severe imperfect interface and completely debonded fibers as functions of the uniaxial strains corresponding to Figure 5.4. From Figure 5.5, it is clear that lower S 0 leads to faster evolution of imperfect interface, indicating that B

B

most fibers are damaged in their interfaces in early stage with lower S 0 and the composites exhibit a B

B

nonlinear stress-strain behavior. We also plot the present predicted stress-strain responses of unidirectional

0.6

0.5

perfect bonding (So=100)

mild imperfect (So=100)

severe imperfect (So=100)

complete debonded (So=100)

perfect bonding (So=200)

mild imperfect (So=200)

severe imperfect (So=200)

complete debonded (So=200)

perfect bonding (So=300)

mild imperfect (So=300)

severe imperfect (So=300)

complete debonded (So=300)

Fiber Volume Fraction

0.4

0.3

0.2

0.1

0 0

0.001

0.002

0.003

0.004

Strain [ε 33]

Figure 5.5 The predicted evolution of volume fractions of perfectly bonded fibers, fibers with mild imperfect interface, fibers with severe imperfect interface and completely debonded fibers corresponding to Figure 5.4

Chapter 5. Evolutionary damage model for fiber-reinforced composites

85

composites with imperfect interfaces under uniaxial tension in the transverse direction( 11  0 and

 ij  0 for all other stress components) with various S 0 values in Figure 5.6. As shown in Figures 5.4 and B

B

5.6, the stress-strain responses due to uniaxial tension in the transverse direction are lower than those due to uniaxial tension in the longitudinal direction. It is concluded that the progressive evolution of imperfect interface affects significantly the overall elastic damage behavior of the composites. To further examine the proposed evolutionary damage model sensitivity to the compliance parameters

 and  , we conduct a parametric analysis of  and  . As parameters controlling the severity of imperfect interface, three sets of the compliance parameters for the severe imperfect interface condition are used:  2 = 2.0,  2 = 3.0;  2  2.0 104 , 2  3.0 104 ;  2  2.0 107 , 2  3.0 107 . For simplicity, fixed values of the compliance parameters for the mild imperfect interface condition and

400

Stress [σ11] (MPa)

300

200

= 100 100 MPa SSo MPa 0 =

100

B

B

B

= 200 200 MPa SSo MPa 0 = B

B

= 300 300 MPa SSo MPa 0 = B

B

0 -0.003

-0.001

[ε 33 ] B

B

0.001 Strain

0.003

0.005

0.007

0.009

0.011

[ε 11 ] B

B

Figure 5.6 The present predicted stress-strain responses of unidirectional composites under uniaxial tension in the transverse direction with various S 0 values B

B

Chapter 5. Evolutionary damage model for fiber-reinforced composites

86

Weibull parameters are used: 1  2.0 107 , 1  3.0 107 ; S 0 = 200 MPa, M = 2. The present predicted B

B

stress-strain responses of the composites under uniaxial tension in the transverse direction with various values of  2 and  2 are shown in Figure 5.7. As the values of the compliance parameters  2 and

 2 become higher, the effect of the severity of imperfect interface on the stress-strain behavior of the composites is more pronounced. Another parametric study is also conducted using the same material properties as used in the previous simulations to investigate the influence of nucleation parameters c 1 and c 2 in Eq. (5.26) on the elastic B

B

B

B

damage behavior of cracked, continuous, fiber-reinforced brittle matrix composites (six-phase composite state). Six different sets of c 1 and c 2 are used in the simulation: c 1 = 0.2 and c 2 = 1.5; c 1 = 0.25 and c 2 = B

B

B

B

B

B

B

B

B

B

B

B

1.5; c 1 = 0.3 and c 2 = 1.5; c 1 = 0.3 and c 2 = 1.1; c 1 = 0.3 and c 2 = 1.6; c 1 = 0.3 and c 2 = 2.3. For brevity, B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

B

fixed values of the compliance parameters and Weibull parameters are used: 1  2.0 107 ,

300

Stress [σ11] (MPa)

200

100

α 2 =2.E-7, β 2 =3.E-7 So = 100 MPa B

B

B

B

α 2 =2.E-4, β 2 =3.E-4 So = 200 MPa B

B

B

B

α 2 =2.0, So = 300βMPa 2 =3.0 B

B

B

B

B

B

B

0 0

0.002

0.004

0.006

0.008

Strain [ε11]

Figure 5.7 The present predicted stress-strain responses of particulate composites under uniaxial tension in the transverse direction with various α 2 and β 2 values B

B

B

B

Chapter 5. Evolutionary damage model for fiber-reinforced composites

87

1  3.0 107 ,  2 = 2.0,  2 = 3.0; S 0 = 180 MPa, M = 1.6.  th in Eq. (5.26) is assumed to be  th = B

B

0.001. We plot the present predicted stress-strain responses of the composites with microcracks and imperfect interfaces under uniaxial tension in the longitudinal direction with various c 1 and c 2 values in B

B

B

B

Figures 5.8 and 5.10, respectively. Figures 5.9 and 5.11 exhibit evolutions of microcrack density corresponding to Figures 5.10 and 5.12, respectively. As the values of the material constants c 1 becomes B

B

higher and c 2 becomes lower, the effect of nucleation of microcracks on the stress-strain behavior of the B

B

composites is shown to be more noticeable.

300

250

Stress [σ33] (MPa)

200

150

c 1 =0.2 c1 = 0.2 B

100

B

B

c 1 =0.25 c1 = 0.25 B

B

c 1 =0.3 c1 = 0.3 B

B

B

50

0 -0.004

-0.002

[ε 11 ] B

B

0 Strain

0.002

0.004

0.006

0.008

[ε 33 ] B

B

Figure 5.8 The present predicted stress-strain responses of unidirectional composites under uniaxial tension in the longitudinal direction with various c 1 values B

B

Chapter 5. Evolutionary damage model for fiber-reinforced composites

88

0.300

0.250

Microcrack density

0.200

0.150

0.100

ccrack1 1 =0.2 B

B

B

ccrack2 1 =0.25 B

0.050

B

ccrack3 1 =0.3 B

B

B

0.000 0

0.002

0.004 Strain [ε

3

0.006 3

0.008

]

Figure 5.9 The predicted evolution of microcrack density corresponding to Figure 5.8

Chapter 5. Evolutionary damage model for fiber-reinforced composites

89

300

250

Stress [σ33] (MPa)

200

150

100

cc22 =1.1 = 1.1 B

B

B

= 1.6 cc22 =1.6 B

B

= 2.3 cc22 =2.3

50

B

B

B

0 -0.004

-0.002

[ε 11 ] B

B

0

0.002

0.004

0.006

0.008

[ε 33 ]

Strain

B

B

Figure 5.10 The present predicted stress-strain responses of unidirectional composites under uniaxial tension in the longitudinal direction with various c 2 values B

B

0.300

0.250

Microcrack density

0.200

0.150

ccrack1 2 =1.1 B

0.100

B

B

ccrack2 2 =1.6 B

B

ccrack3 2 =2.3 B

B

B

0.050

0.000 0

0.002

0.004 Strain [ε

3

0.006 3

0.008

]

Figure 5.11 The predicted evolution of microcrack density corresponding to Figure 5.10

Chapter 5. Evolutionary damage model for fiber-reinforced composites

90

5.5 Experimental comparison The proposed evolutionary damage model is further exercised to predict the behavior of continuous, fiber-reinforce brittle matrix composites, and the prediction is compared with experimental data reported by Luo and Daniel (2000) to verify the accuracy of the proposed model. The stress-strain behavior of unidirectional silicon carbide/calcium aluminosilicate (SiC/CAS) composites tested by Luo and Daniel (2000) is numerically predicted using the proposed evolutionary damage model. The elastic properties of the composites used in this simulation are (Luo and Daniel, 2000): E m = 97.9 GPa,  m = 0.25, E f = 188 B

B

B

B

GPa,  f = 0.3,  = 0.36. According to the experimentally obtained stress-strain curve in the longitudinal

500

Stress [σ33] (MPa)

400

300

200

experiment 100

present prediction (no-microcracks) present prediction (with microcracks)

0 0

0.002

0.004 Strain [ε33

0.006

0.008

]

Figure 5.12 The comparison between the present predictions (with and without microcracks) and experimental data (Luo and Daniel, 2000) for overall uniaxial tensile responses in the longitudinal direction of silicon carbide/calcium aluminosilicate (SiC/CAS) composites

Chapter 5. Evolutionary damage model for fiber-reinforced composites

91

direction (Luo and Daniel, 2000), the compliance parameters 1 , 1 ,  2 and  2 given in Eq. (5.10) as well as the Weibull parameters S 0 and M in Eqs. (5.17)-(5.19) for the uncracked, five-phase composite state are B

B

estimated to be: 1  2.0 107 and 1  3.0 107 ,  2 = 2.0,  2 =3.0; S 0 = 280 MPa, M = 1.3. For the B

B

cracked, six-phase composite state, additional parameters related to nucleation of microcracks c 1 , c 2 , B

B

B

B

v 0 and  th in Eq. (5.26) are estimated to be: c 1 = 0.25, c 2 = 1.6, v 0 = 0.0 and  th =0.002. B

B

B

B

The present predicted (longitudinal) uniaxial stress-strain responses of the two different composite states are compared with the experimentally obtained stress-strain curve by Luo and Daniel (2000) in Figure 5.12. The predicted evolutions of volume fractions of fibers (five-state composite state) and microcrack density (six-phase composite state) corresponding to Figures 5.12 are shown in Figures 5.13 and 5.14, respectively. It is observed that the response with the model considering nucleation of

0.4

0.35

Fiber Volume Fraction

0.3

0.25

0.2

perfect bonding mild imperfect interface

0.15

severe imperfect interface completely debonded

0.1

0.05

0 0

0.002

0.004 Strain [ε33

0.006

0.008

]

Figure 5.13 The predicted evolution of volume fractions of perfectly bonded fibers, fibers with mild imperfect interface, severe imperfect interface and completely debonded fibers (fivephase composite state) corresponding to Figure 5.12

Chapter 5. Evolutionary damage model for fiber-reinforced composites

92

microcracks is lower than that without nucleation of microcracks. Clearly, in comparison with the model without nucleation of microcracks, the present prediction with nucleation of microcracks matches well with the experimentally obtained stress-strain curve. 0.250

Microcrack density

0.200

0.150

0.100

0.050

0.000 0

0.002

0.004

0.006

0.008

Strain [ε33]

Figure 5.14 The predicted evolution of microcrack density (six-phase composite state) corresponding to Figure 5.12

Chapter 5. Evolutionary damage model for fiber-reinforced composites

93

(1) (10) (2) (1) 5.6 Appendix. Parameters  IK and  IJ in Eq. (5.15), and  IK , … ,  IJ in

Eq. (5.24) These parameters in Eqs. (5.15) and (5.24) take the form:

 IJ(2) 

(1) (3) (5) (7) (9)  IK  (1) IK   IK   IK   IK   IK ,

1 (4) (6) (8) (10)  (2) IJ   IJ   IJ   IJ   IJ 2

(5.27)

with 3  (2r 1) (2 r )   r 1) (2 r ) (2 r 1) (2 r 1) (2 r 1)  (2   2    2     In nK  (r = 1, 2, 3, 4, 5)  r  IK IK KK II IK   n 1  

(5.28)

r) (2 r ) (2 r ) (2 IJ  2 rIJ IJ

(5.29)

(r = 1, 2, 3, 4, 5)

and (2 r 1) IK 

 (IKr ) 1  2 II(2r )

,

IJ(2r ) 

1

(5.30)

2  4 IJ(2r )

(2 r 1) (2 r ) (2 r 1) (2 r 1)  (Ir1)  1  11   211  21 31      (r )  (2 r 1) (2 r 1) (2 r ) (2 r 1) 12 1   22  2 22 32   I 2     (r )   (2 r 1) (2 r 1) (2 r 1) (2 r )  I3   13  23 1  33  233     

1

 I(21 r 1)   (2r 1)    I 2   (2r 1)     I 3 

(5.31)

(1) where I = 1, 2, 3. In addition, the parameters  IK , .... ,  IJ(10) read (1)  IK 

(3)  IK 

(5)  IK 

(7)  IK 

3   1  (1) (2) (2) (1) (1) (1)  2S IK  KK  2S II  IK   S In nK  , 4 1  0    n 1  

2

256a 1  0 

2

3 256a 1  0 

2

 IJ(2) 

 1S IJ(2) IJ(2) 2 1  0 

3  M (1) (4)   M (2) (3) M (1) (3)  2S IK  KK  2S II  IK  S In nK  ,   n 1  

 IJ(4) 

3   M (3) (6)  M (4) (5) M (3) (5)  2S IK  KK  2S II  IK  S In nK  ,   n 1  

 IJ(6) 





3  4  (1) (8)  (2) (7) (1) (7)  2S IK  KK  2S II  IK   S In nK  , 4 1  0    n 1  

 

3

 

 

n 1

 

(9) (3) (10) (9) (3) (9)  IK   5 2S IK KK  2S II(4)IK   S In nK  ,

(8)  IJ 

(5.32)

 2 S IJM (2) IJ(4) 128a 1  0 

 3 S IJM (4) IJ(6) 128a 1  0 

(2) (8)  4 S IJ  IJ 2 1  0 

(10) (4) (10)  IJ  2 5 SIJ IJ

2

2

(5.33)

(5.34)

(5.35)

(5.36)

Chapter 5. Evolutionary damage model for fiber-reinforced composites

94

with

(1)  IK 

(3)  IK 

(5)  IK 

(7)  IK 

(9)  IK 

(1) YIK

  1 2  1'  S II(2)    4 1  0   

,

 IJ(2) 

(2) YIK

  1 M (2)  2  2'  S II 2   256a 1  0    (3) YIK

  1 M (4)  2  3'  S 2 II   256a 1  0    (4) YIK , 1 (2) 1  S 2 1  0  II

(5) YIK

1  2S II(4)

,

 IJ(10) 

1  '  1 4  1  S IJ(2)    4 1  0   

,

 IJ(4) 

,

 IJ(6) 

 IJ(8) 

1   1 M (2)  4  2'  S IJ 2   256a 1  0    1   1 M (4)  4  3'  S 2 IJ   256a 1  0   

1 1 2  S (2) 1  0 IJ

1 2  4S IJ(4)

(5.37)

(5.38)

(5.39)

(5.40)

(5.41)

in which

YI(1r )   ( r )  YI 2   Ar B r  (r )  YI 3 

(5.42)

with

 '  1 1 1 (1) (2)  (1) (1) '  S11  2S11 1'  S21 1'  S31 1  21     4 1   4 1   4 1    0  0  0     1 1 1 (1) (1) (2) (1) '  S22  2S22   A1   1'  S12 1'  21'    S32 1  4 1  0  4 1  0   4 1  0      1 1 1 (1) (1) (1) (2) ' ' ' '   S33  2S33   1  S13 1  S23 1  21    4 1  0  4 1  0  4 1  0    

1

(5.43)

Chapter 5. Evolutionary damage model for fiber-reinforced composites

95

  1 1 1 M (2)   S M (1)  2S11 2'  22'   2'  S M (1) 2'  S M (1) 2  11 2 21 2 31    256a 1  0  256a 1  0  256a 1  0    1 1 1   M (2)  M (1) '  S M (1)  2S22 A2   2'  S M (1) 2'  22'    S32 2  2 12 2  22 2  256a 1  0  256a 1  0  256a 1  0      1 1 1 M (1) M (1) M (1) M (2)  ' ' ' '    2  S 2  S 2  22  S  2S33 2 13 2 23 2  33   256a 1  0  256a 1  0  256a 1  0   

1

(5.44)

  1 1 1 M (4)   S M (3)  2S11 3'  23'   3'  S M (3) 3'  S M (3) 2  11 2 21 2 31    256a 1  0  256a 1  0  256a 1  0    1 1 1   M (4)   S M (3)  2S22 A3   3'  S M (3) 3'  23'  3'  S M (3)  2 12 2  22 2 32  256 a 1   256 a 1   256 a 1    0  0  0     1 1 1 M (3) M (3) M (3) M (4)  ' ' ' '    3  S 3  S 3  23  S  2S33 2 13 2 23 2  33   256a 1  0  256a 1  0  256a 1  0   

1

(5.45)

  1 1 1 (1) (2)  (1) (1)  S11  2S11 S21 S31  1     4 1  0  4 1  0  4 1  0      1 1 1 (1) (1) (2)  (1)  S22  A4   S12 1   2S22 S32  4 1  0  4 1  0   4 1  0      1 1 1 (1) (1) (1) (2)     S33 S13 S23 1   2S33   4 1  0  4 1  0  4 1  0    

(3) (4)  1  S11  2S11  (3) A5   S12  (3) S13 

(3) S21 (3) (4) 1  S22  2S22

 '  1 S I(1) 1  1  4 1  0      1 B1  1'  S I(1)2  , 4 1  0      1 1'   S I(1) 4 1  0  3  

(3) S23

  (3) S32  (3) (4)  1  S33  2S33  (3) S31

1

(5.46)

1

  1 M (1)  2'  S 2 I1   256a 1  0    1  M (1)  B2  2'  S 2 I2  256a 1  0      1 M (1)   '  S  2 256a 1  2 I 3  0  

(5.47)

(5.48)

Chapter 5. Evolutionary damage model for fiber-reinforced composites

  1   1 3'  S M (3)  S I(1)  2 I1 1    256a 1  0   4 1  0    S I(3)    1    (3)  1 1  '  M (3)  (1)  , , B3  3  S B  S B  S      4 5 I 2 I 2 I 2 2 256a 1  0     (3)   4 1  0   S      I 3  1 (1) 1 M (3)   '    S S I3  3 256a 1  2 I 3   4 1  0   0  

96

(5.49)

where

r'  

r 0  0  r

 r  0 3   r  0   2  r  0 

,

r' 

0

2   r  0 

(5.50)

97

CHAPTER 6 NUMERICAL ANALYSIS OF LAMINATED COMPOSITE STRUCTURES BASED ON MICROMECHANICAL MULTI-LEVEL DAMAGE MODEL

6.1. Introduction Due to their great potential for weight savings, fiber reinforced laminated composites are becoming increasingly important for structural applications concerned with light weight design, e.g. in the fields of aerospace, transportation and other industrial applications. Despite these desirable physical properties, composites are fragile and susceptible to damage from a number of sources, both during initial processing and in service (Lapczyk and Hurtado, 2007). Even though the damage does not cause catastrophic failure, it significantly reduced structural stiffness (Hassan and Batra, 2008). In order to exploit the advantages of laminated composites, reliable prediction of the mechanical behavior is essential. Many micromechanical models have been developed in order to obtain the homogenized elastic properties of composite materials (Qu, 1993a, 1993b, Tohgo and Weng, 1994; Zhao and Weng, 1995, 1996, 1997; Ju and Lee, 2000, 2001; Liu et al., 2004, 2006; Lee and Pyo, 2007, in press) or to study the effect of microcracks on the elastic response of brittle solids (Karihaloo and Fu, 1989; Lee and Liang, 2004; Lee and Simunovic, 2006; Lee and Pyo, submitted for publication). Damage propagation cannot easily be modelled by more conventional finite element models due to the size and complexity of these microstructures (Grufman and Ellyin, in press). Many finite element analyses about damage and failure in laminated composites under various loading cases have been carried out (Davalos et al., 1996; Luciano and Zinno, 2000; Maa and Cheng, 2002; Kilic and Haj-Ali, 2003; Rodriguez and Ochoa, 2004; Liang et al., 2006; Guo, 2007; Lee and Kim, 2007; Maimi et al., 2007; Teng, 2007). Davalos et al. (1996) conducted the analysis of pultruded FRP beams in bending using finite element (FE) analysis. Liang et al. (2006) and Lee and Kim (2007) implemented micromechanics based damage model into FE code to simulate laminated plates containing a cutout under uniaxial compression. Guo (2007) conducted numerical and experimental studies to investigate the effect of reinforcements around cutouts on the stress concentration and buckling behavior of a carbon/epoxy composite panel under in-plane shear load.

Chapter 6. Damage modeling of laminate composite structures

98

However, there still remain various uncertainties about the failure mechanism (Suemasu et al., 2006). Development of computational methodologies for the accurate prediction of damage accumulation and growth in continuous fiber composite laminates is presently an active area of research (Basu et al., 2007). In the present study, continuous fibers are modeled to be elastic cylindrical inclusions that are randomly dispersed yet unidirectionally aligned in an elastic matrix. Microcracks are assumed to be located perpendicular to the fibers and the size of microcracks is uniform (see also Lee and Pyo, submitted for publication). It is also assumed that the progression of imperfect interface is governed by the average internal stresses of the fibers as well as the Weibull parameters (Lee and Pyo, in press). A micromechanics based constitutive model incorporating damage models was implemented into ABAQUS non-linear FE code (ref. 116) using a user-subroutine UMAT to solve boundary value problems. Numerical examples corresponding to various loading conditions and geometries are solved to illustrate the potential of the proposed implemented damage model. Comparisons between the present prediction with available experimental data in the literature are conducted to verify the accuracy of the proposed computational model.

X2 B

X1 B

X3 B

Figure 6.1 Coordinates for a transversely isotropic material

Chapter 6. Damage modeling of laminate composite structures

6.2

Micromechanics-based

constitutive

99

damage

model

for

off-axis

unidirectional fiber reinforced composites The 3-direction is chosen to be fiber direction and the plane 1-2 corresponds to the transversely isotropic plane in the unidirectional fiber reinforced composites in this study (see Figure 5.1). Now we consider rotations through an angle  about the X 2 -axis to derive the stiffness of off-axis unidirectional B

B

fiber-reinforced composites (see Figure 6.1). Following the stiffness transformation law by Herakovich (1998), we have the following stress and strain transformations (see also Liang et al., 2006):

   C :    T1 :  

(6.1)

   T2 : 

(6.2)

where the subscript ρ signifies stress and strain matrices in the principal coordinate system, the subscript ς denotes stress and strain matrices in the global coordinate system, and the transformation matrices T 1 and B

B

T 2 are given as B

B

 cos 2   0   sin 2  T1    0  cos  sin   0   cos 2   0   sin 2  T2    0   2 cos  sin   0 

0

sin 2 

0

2 cos  sin 

1

0

0

0

0

cos 2 

0

2 cos  sin 

0

0

cos 

0

0

cos 2   sin 2 

 sin 

0

0  cos  sin  0

0

0   0   0  sin    0   cos  

0

sin 2 

0

 cos  sin 

1

0

0

0

0

cos 

0

cos  sin 

0

0

cos 

0

0

cos 2   sin 2 

 sin 

0

2

0 2 cos  sin  0

0

(6.3)

0   0   0  sin    0   cos  

(6.4)

By combining Eqs. (6.1) and (6.2), the stiffness matrix of off-axis unidirectional fiber-reinforced composites C , which is the transformed stiffness matrix through an arbitrary angle  about X 2 -axis, is B

B

Chapter 6. Damage modeling of laminate composite structures

100

derived as

   T11  C  T2 :    C : 

(6.5)

where the stiffness matrix of the unidirectional fiber-reinforced composites with the fibers oriented off-axis takes the form:

 C11 C12  C21 C22 C C32 C   31 0 0  C C52  51 0  0

C13

0

C15

C23

0

C25

C33

0

C35

0

C44

0

C53

0

C55

0

C64

0

0   0  0   C46  0   C66 

The components of C are given in Appendix.

(6.6)

Chapter 6. Damage modeling of laminate composite structures

101

6.3 Damage modeling The multi-level elastic damage model proposed by Lee and Pyo (in press) is considered here for a complete description of the sequential progression of imperfect interface in the composites. The probability of imperfect interface is modeled as a two-parameter Weibull process and the average internal stresses of perfect bonded fibers (phase 1) are chosen to be the controlling factor of the Weibull function as

   f   2   1  exp     S0    

       

  m 

M

(6.7)

   f   3   2 1  exp     S0    

       

(6.8)

   f   4   3 1  exp     S0    

       

(6.9)

  m    m 

3  3   4 ,

 2   2  3 ,

M

M

1     2

where  is the original particle volume fraction,

(6.10)

 f m is the internal stress of fibers, and S

0

B

B

and M are

the Weibull parameters. The internal stresses of fibers required for the initiation of the imperfect interface were explicitly derived by Ju and Lee (2000) as 1

1 1 1  C1   I  S1   A1  S1    I   S1   A1  S1   : ε  U : ε    

(6.11)

By carrying out the lengthy algebra, the components of fourth-rank tensor U for the present model are explicitly given by



(1) Uijkl  U IK ij  kl  U IJ(2) ik  jl  il  jk



(6.12)

Chapter 6. Damage modeling of laminate composite structures

102

(1) where the second-rank tensors U IK and U IJ(2) can be expressed as

3

(1) (1) f (2) (1) f (1) U IK  2 EIK KK  2 EII(2)IKf (1)   EIn nK , n 1

U IJ(2)  2EIJ(2)IJf (2)

(6.13)

in which

(1) EIK

 2 f (2)  1 1  KK   

IKf (1) 

f (1)  IK

1  2 IIf (2)

 21 IKf (1) 1      

IJf (2) 

,

3

nKf (1) , n 1

 2 f (2) EIJ(2)  1 1  IJ   

1

(6.14)

(6.15)

2  4 IJf (2)

 If1(1)  1  11f (1)  211f (2)   21f (1) 31f (1)      f (1)  12f (1) 1   22f (1)  2 22f (2) 32f (1)   I 2      f (1)   13f (1)  23f (1) 1  33f (1)  233f (2)   I 3     

 IKf (1) 

   

3  (1) (2) (2) (1) (1) (1)   2S IK KK  2S II IK   S In nK  , 4 1  0    n 1



 IJf ( 2 )

1

 If1(1)   f (1)   I 2   f (1)    I 3 

(6.16)

 S IJ( 2) IJ ( 2 ) 2 1  0 

(6.17)

(1) (1) where the subscript I = 1, 2, 3 and the components S IK , S IJ(2) ,  IK and  IJ(2) in Eq. (6.17) are listed in

Chapter 5. In case of tensile loading, the averaged internal stresses of fibers can be obtained as

 σ f m   f 11   f 22   f 33  2 f 23   f 13   f 12  2

with

2

2

2

2

2

1/ 2

(6.18)

Chapter 6. Damage modeling of laminate composite structures

 f 11   (1) (2) U11  2U11  f 22   U (1) 21  f 33   U31(1)   f 23   0    f 13   0  0  f 12  

            

(1) U12

103

(1) U13

0

0

(1) U 23 (1) (2) U 33  2U 33

0

0

0

0

0

0

(2) U 23

0

0

0

0

(2) U13

0

0

0

0

(2)  2U12 (1) U 32

(1) U 22

0     11  0      22  0    33    0  2 23    0   213   (2)    212  U12 

(6.19)

In accordance with the continuum damage model for the nucleation of microvoids or microcracks in brittle materials (Karihaloo and Fu, 1989), the density of nucleated microcracks in the composite is modeled using the following isotropic scalar function (Liang et al., 2006)

 v 0 ,  5     th v 0  c1 1  a   

 a   th c

2 a th  ,    

(6.20)

where v 0 is the initial density of microcracks.  th   ijth ijth is the threshold effective strain below which no nucleation occurs, c 1 and c 2 are material (nucleation) parameters depending on the specific shape and B

B

B

B





1/ 2

2 2 2 2 2 2    22   33  2  23  13  12 distribution of microcracks, and  a   ij  ij  11  

accumulated effective strain (Karihaloo and Fu, 1989).

is the current

Chapter 6. Damage modeling of laminate composite structures

104

6.4 Numerical simulations and experimental comparison In this section some comparisons with both numerical and experimental data available in literature are carried out to show the effectiveness of the proposed constitutive and damage model presented previous sections. The proposed constitutive and damage model for laminate composites are implemented into ABAQUS non-linear FE code using a user-subroutine UMAT.

6.4.1 Verification of the computer code The implemented computational model is validated by comparing the predicted stress-strain curve and experimental data under uniaxial tension with various fiber orientations. The stress-strain curve of

400

350

Stress [σ33] (MPa)

300

prediction-15º

experiment-15º

prediction-30º

experiment-30º

prediction-45º

experiment-45º

250

200

150

100

50

0 0

0.005

0.01

0.015

Strain [ε33]

Figure 6.2 The comparison between the present prediction and experimental data (Weeks and Sun, 1998) for overall uniaxial tensile responses with three fiber orientations of AS4/PEEK composite

Chapter 6. Damage modeling of laminate composite structures

105

AS4/PEEK composite test by Weeks and Sun (1998) is numerically predicted using implemented computational model. The elastic properties of the composites used in this simulation are (Weeks and Sun, 1998; Hassan and Batra, 2008): E m = 6.14 GPa,  m = 0.356, E f = 214 GPa,  f = 0.263 and  = 0.6, B

B

B

B

where the subscript m and f represent the matrix and fibers, respectively. According to the experimentally obtained stress-strain curve (Weeks and Sun, 1998), the model parameters (see detail descriptions of Chapter 5) are estimated as: 1  2.0 107 and 1  3.0 107 ,  2 = 2.0,  2 = 3.0; S 0 = 290 MPa, M = 0.8, B

B

c 1 = 0.25, c 2 = 1.1,  v 0 = 0.0 and  th = 0.005. B

B

B

B

The three-dimensional, eight-node, linear brick solid element (C3D8) in ABAQUS is used for modeling the laminated composites. Following Hassan and Batra (2008), four-ply composites of overall dimensions 216 mm × 15.9 mm × 2.54 mm, divided into a uniform 84 × 6 × 4 FE mesh, were loaded by pulling the opposite end faces. Figure 6.2 exhibits the comparison between present predicted uniaxial stress-strain responses and the experimentally obtained stress-strain curve (Weeks and Sun, 1998) with 0.7

0.6

Volume Fraction

0.5

perfect bonding 0.4

mild imperfect interface severe imperfect interface completely debonded

0.3

microcrack

0.2

0.1

0 0

0.005

0.01

0.015

Strain [ε33]

Figure 6.3 The predicted evolution of volume fraction of perfectly bonded fibers, fibers with mild imperfect interface, fibers with severe imperfect interface, completely debonded fibers and microcrack corresponding to 15º fiber orientation case in Figure 6.2

Chapter 6. Damage modeling of laminate composite structures

106

various fiber orientations (15°, 30° and 45°). Overall, good predictions are shown by the FE models compared with the experimental results. This confirms the ability of the proposed implemented computational model to predict the overall behavior of laminated composites. The predicted evolutions of volume fractions of fibers and microcrack corresponding to 15° fiber orientation case of Figure 6.2 are shown in Figure 6.3. Smooth and gradual transitions from the perfectly bonded fibers to various types of damaged fibers and evolution of microcracks occur within the present framework; as a result, nonlinear stress-strain response can be obtained during the damage evolution.

6.4.2 Experimental comparison of the response and damage behavior of a laminated composite under bending To further examine the proposed implemented computational model, we conduct comparison between proposed model and 12-ply carbon fiber reinforced epoxy matrix laminated beam, [0/±45/0/90/0] s , B

B

subjected to three-point bending test by Huang (2007). The beam had a span of 84 mm, a width of 15.1 mm and a thickness of 2.76 mm (Huang, 2007). Figure 6.4 illustrates the finite element discretization employed to simulate the bending response and damage evolution in laminate composite. The material properties of laminated composite are (Huang, 2007): E m = 3.685 GPa,  m = 0.35, E f = 210 GPa,  f = 0.2 and  = B

B

B

B

0.44. According to the experimentally obtained load-displacement curve (Huang, 2007), the model parameters are estimated as: 1  2.0 107 and 1  3.0 107 ,  2 = 2.0,  2 = 3.0; S 0 = 450 MPa, M = B

B

3.0, c 1 = 0.35, c 2 = 1.15,  v 0 = 0.0 and  th = 0.005. B

B

B

B

The predicted load-displacement curve is compared with experimental data (Huang, 2007) under three-point bending in Figure 6.5. It is worth noting that the prediction of the proposed implemented computational model for laminated composites agrees well with the experimental data. The predicted evolution of volume fractions of perfectly bonded fibers, various types of damaged fibers and microcracks versus displacement corresponding to various segments depicted in Figure 6.5 are exhibited in Figure 6.6. The damage evolutions of segment A are shown to be more noticeable compared with those of the other segment. Figure 6.7 shows the sequence of the deformed shape and von-Mises effective stress of laminated composite beam during three-point bending.

Chapter 6. Damage modeling of laminate composite structures

107

Segment D Segment C Segment B Segment A

Segment A

Segment E Segment F

Figure 6.4 The finite element discretization employed to simulate the bending response and damage evolution in laminated composite beam

Chapter 6. Damage modeling of laminate composite structures

108

0.4

0.35

0.3

Load (KN)

0.25

0.2

0.15 prediction experiment #1

0.1

experiment #2 experiment #3 experiment #4

0.05

0 0

1

2

3

4

5

displacement (mm)

Figure 6.5 The comparison of load-displacement curves between the present prediction and experimental data (Huang, 2007) under three-point bending

Chapter 6. Damage modeling of laminate composite structures

a

109

0.5

Segment A

0.45 0.4

perfect bonding mild imperfect interface severe imperfect interface completely debonded microcrack

volume fraction

0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0

1

2

3

4

5

displacement (mm)

b

0.5

Segment B

0.45 0.4

perfect bonding mild imperfect interface severe imperfect interface completely debonded microcrack

volume fraction

0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0

1

2

3

4

5

displacement (mm)

Figure 6.6 The predicted damage evolution versus displacement of laminated composite beam under three-point bending corresponding to Figure 6.5: (a) Segment A; (b) Segment B; (c) Segment C; (d) Segment D; (e) Segment E; (f) Segment F

Chapter 6. Damage modeling of laminate composite structures

110

0.5

c

Segment C

0.45 0.4

volume fraction

0.35 0.3 perfect bonding

0.25

mild imperfect interface 0.2 severe imperfect interface 0.15

completely debonded

0.1

microcrack

0.05 0 0

1

2

3

4

5

4

5

displacement (mm)

0.5

d

Segment D

0.45 0.4

volume fraction

0.35 0.3

perfect bonding mild imperfect interface

0.25

severe imperfect interface 0.2

completely debonded

0.15

microcrack

0.1 0.05 0 0

1

2

3

displacement (mm)

Figure 6.6 The predicted damage evolution versus displacement of laminated composite beam under three-point bending corresponding to Figure 5 (continued)

Chapter 6. Damage modeling of laminate composite structures

e

0.5

Segment E

0.45

perfect bonding mild imperfect interface

0.4

severe imperfect interface

0.35 volume fraction

111

completely debonded

0.3

microcrack

0.25 0.2 0.15 0.1 0.05 0 0

1

2

3

4

5

displacement (mm)

f

0.5

Segment F

0.45 0.4

volume fraction

0.35 0.3

perfect bonding mild imperfect interface

0.25

severe imperfect interface

0.2

completely debonded

0.15

microcrack

0.1 0.05 0 0

1

2

3

4

5

displacement (mm)

Figure 6.6 The predicted damage evolution versus displacement of laminated composite beam under three-point bending corresponding to Figure 5 (continued)

Chapter 6. Damage modeling of laminate composite structures

112

10/50

20/50

30/50

40/50

50/50 Figure 6.7 The sequence of the deformed shape and von-Mises effective stress of laminated composite beam during three-point bending

Chapter 6. Damage modeling of laminate composite structures

113

6.4.3 Experimental comparison of the response and damage behavior of a laminated composite under shear Another numerical study is also conducted to predict shear behavior of laminated composites. In this simulation, the Iosipescu (V-notch) shear test was selected, as one of the most efficient method, to determine the shear properties of laminated composites (Broughton et al., 1990; Bhatnagar et al., 1995; Khashaba, 2004). Figure 8 illustrates FE model of Iosipescu shear test in present simulations. The material properties of glass fiber reinforced epoxy matrix laminated composite are (Khashaba, 2004): E m = 3.2 GPa, B

B

 m = 0.36, E f = 72.4 GPa,  f = 0.22 and  = 0.4267. The comparisons are conducted with the cross-ply B

B

laminates with different stacking sequences ([75/-15] 2s , [45/-45] 2s and [30/-60] 2s ). According to the B

B

B

B

B

B

experimentally obtained load-displacement curve (Khashaba, 2004), the model parameters are estimated as: 1  2.0 107 and 1  3.0 107 ,  2 = 2.0,  2 = 3.0; S 0 = 1400 MPa, M = 3.0, c 1 = 0.35, c 2 = 1.1, B

B

B

B

B

B

 v 0 = 0.0 and  th = 0.0001. Figures 6.9 shows the comparisons of load-displacement curves between the present prediction and experimental data (Khashaba, 2004) under shear loading for cross-ply laminated composite. It is evident that computed results are close to those experimentally obtained data. Figures 6.10 – 6.12 illustrate the sequence of the deformed shape and von-Mises effective stress of cross-ply laminated composite under shear loading.

Figure 6.8 FE model of Iosipescu shear test

Chapter 6. Damage modeling of laminate composite structures

114

9

a

8

[75/-15] 2s

7

B

B

Load (KN)

6 5 4 3

prediction

2

experiment

1 0 0

1

2

3

4

5

displacement (mm)

10

b

9

[45/-45] 2s

8

B

B

Load (KN)

7 6 5 4 prediction 3 experiment

2 1 0 0

1

2

3

4

displacement (mm)

10 9

c

[30/-60] 2s

8

B

B

Load (KN)

7 6 5 4 prediction

3

experiment

2 1 0 0

1

2

3

4

displacement (mm)

Figure 6.9 The comparison of load-displacement curves between the present prediction and experimental data (Khashaba, 2004) under shear loading: (a) [75/-15] 2s laminated B

B

composite; (b) [45/-45] 2s laminated composite; (c) [30/-60] 2s laminated composite B

B

B

B

Chapter 6. Damage modeling of laminate composite structures

115

[75/-15] 2s B

B

10/40

20/40

30/40

40/40

Figure 6.10 The sequence of the deformed shape and von-Mises effective stress of laminated composite under shear loading in case of [75/-15] 2s B

B

Chapter 6. Damage modeling of laminate composite structures

116

[45/-45] 2s B

B

10/40

20/40

40/40

30/40

Figure 6.11 The sequence of the deformed shape and von-Mises effective stress of laminated composite under shear loading in case of [45/-45] 2s B

B

Chapter 6. Damage modeling of laminate composite structures

117

[30/-60] 2s B

B

10/40

20/40

40/40

30/40

Figure 6.12 The sequence of the deformed shape and von-Mises effective stress of laminated composite under shear loading in case of [30/-60] 2s B

B

Chapter 6. Damage modeling of laminate composite structures

118

6.5 Appendix. Components of C in Eq. (6.7) These components are











(1) (2) (1) (1) (2) (1) (2) C11  c4 C11  2C11  c 2 s 2 C13  C31  4 C13  s 4 C33  2C33



(6.21)

(1) (1) C12  c2C12  s 2C32

(6.22)





(1) (1) (2) (1) (2) (2) (1) C13  c4 C13  c 2 s 2 C11  2C11  C33  2C33  4 C13  s 4 C31

 



(6.23)



(1) (2) (1) (2) (1) (1) (2) (2) C15  cs c 2 C11  2C11  C13  2C13  s 2 C31  C33  2C33  2 C13



(6.24)

(1) (1) C21  c2C21  s 2C23

(6.25)

(1) C22  C22

(2)  2C22

(6.26)

s

(1) C21

(6.27)

(1)  C21 )

(6.28)

C23  c

2

(1) C23

(1) C25  cs(C23

2





(1) (1) (2) (1) (2) (2) (1) C31  c4 C31  c 2 s 2 C11  2C11  C33  2C33  4 C13  s 4 C13

(6.29)

(1) (1) C32  c2C32  s 2C12

(6.30)











(1) (2) (1) (1) (2) (1) (2) C33  c4 C33  2C33  c 2 s 2 C13  C31  4 C13  s 4 C11  2C11

 





(6.31)



(1) (1) (2) (2) (1) (2) (1) (2) C35  cs c2 C31  C33  2C33  2C13  s 2 C11  2C11  C13  2 C13



(2) (2) C44  c2C23  s 2C12



(2) (2) C46  cs C23  C12

(6.33)



(6.34)

 





(1) (2) (1) (2) (1) (1) (2) (2) C51  cs c 2 C11  2C11  C31  2C13  s 2 C13  C33  2C33  2 C13



(1) (1) C52  cs C32  C12







(6.35) (6.36)

 



(2) (1) (2) (1) (1) (1) (2) (2) C55  c 4  s 4 C13  c 2 s 2 C11  2C11  C13  C31  C33  2C33  2C13







(1) (1) (2) (2) (1) (2) (1) (2) C53  cs c2 C13  C33  2C33  2C13  s 2 C11  2C11  C31  2 C13



(6.32)





(6.37) (6.38)



(6.39)

(2) (2) C66  c2C12  s 2C23

(6.40)

(2) (2) C64  cs C23  C12

in which c = cos  and s = sin 

119

CHAPTER 7 CONCLUSIONS

Elastic and elastoplastic multi-level damage models for composites with imperfect interface have been presented.

In Chapter 2, a micromechanical framework to predict the effective elastic behavior and imperfect interface evolution of particulate composites is presented. The Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface (Qu, 1993a, 1993b) is employed to model spherical particles having imperfect interfaces in the composites and is incorporated into the micromechanical framework. The effective elastic moduli of three-phase particulate composites are derived based on the Eshelby's micromechanics. A damage model is subsequently considered in accordance with the Weibull's probabilistic function to characterize the varying probability of evolution of imperfect interface between the inclusion and the matrix. The proposed micromechanical elastic damage model is applied to the uniaxial, biaxial and triaxial tensile loadings to predict the various stress-strain responses. Further, the present predictions are compared with other numerical and analytical predictions and available experimental data to verify and assess the proposed micromechanical framework. The successful incorporation of the weakened interface model by a spring layer of vanishing thickness (Qu, 1993a, 1993b) into governing micromechanical ensemble-volume averaged field equations (Ju and Chen, 1994a) yields computational frameworks that enable us to predict the effective elastic behavior and imperfect interface evolution of particulate composites, which can be portrayed as the first noninteracting, micromechanics-based approximation for imperfect interface modeling. It is observed from Figures 2.2 – 2.10 that the effect of imperfect interfaces on the effective stress-strain responses is more pronounced as the damage evolution becomes rapid. It is found from the parametric study on the compliance and Weibull parameters that higher  and  and lower S 0 , which is related the B

B

interfacial strength of composites, lead to faster evolution of imperfect interface. By contrast, if the  and  are low and the S 0 is high, then the effects of imperfect interface are not pronounced when B

B

compared with the perfect bonding model. The proposed micromechanical framework is assessed by comparing the present predictions with

Chapter 7. Conclusions

120

other numerical (Ju and Lee, 2001) and analytical (Kerner, 1956; Halpin and Kardos, 1976) predictions as well as experimental data (Zhou et al., 2004). The predicted stress-strain behavior of particulate composites featuring the current evolution in imperfect interface is observed to be in good qualitative agreement with the experimental data as shown in Figure 13. However, further comparisons between numerical predictions based on the proposed elastic damage model and other numerical simulations and experiments will be needed for the calibration of the model parameters (  ,  , S 0 , M). B

B

In Chapter 3, a multi-level elastic damage model based on a combination of a micromechanical formulation and a multi-level damage model has been presented to predict the effective elastic behavior and progressive imperfect interface in particulate composites. A four-level damage model based on the Weibull's probabilistic function is developed to describe the progressive imperfect interface in the composites. The Eshelby's tensor for a slightly weakened ellipsoidal inclusion (Qu, 1993a, 1993b) to model particles having mild or severe imperfect interface is incorporated into a micromechanics-based multi-level elastic model. The proposed micromechanical elastic damage model is applied to the uniaxial tension loading to predict the corresponding stress-strain responses. A parametric analysis is also carried out to address the influence of the compliance and Weibull parameters on the progressive imperfect interface in the composites. Finally, the present predictions are compared with experimental data (Sørensen, 1993; Geiger and Welch, 1997) to further illustrate the elastic damage behavior of the present framework and to verify the validity of the proposed multi-level elastic damage model. The stress-strain curves of the composites are linear at the early stage of loading since the volume fractions of damaged particles are low. As strains or stresses continue to increase, the volume fractions of damaged particles increase gradually, resulting in a nonlinear stress-strain response. The proposed multilevel elastic damage model naturally captures the gradual transition from previous damage level to next damage level and various types of damaged particles exist simultaneously within the proposed framework. If the strength at the particle-matrix interface is low (lower S 0 ), most particles are mild imperfect, severe B

B

imperfect or completely debonded in their interfaces even in early stage of loading. The influence of the Weibull parameter S 0 on the progression of imperfect interface and stress-strain behavior of the composites B

B

is quite significant. As the values of the compliance parameters  2 and  2 become higher, the effect of the severe imperfect interface on the stress-strain behavior of the composites is more pronounced. The predicted stress-strain behavior of particulate composites featuring the multi-level damage progression of imperfect interface is observed to be in good qualitative agreement with experimental data (Geiger and Welch, 1997).

Chapter 7. Conclusions

121

In Chapter 4, an elastoplastic multi-level damage model considering evolutionary imperfect interface to predict the effective elastoplastic behavior and multi-level damage evolution in particle reinforced ductile matrix composites (PRDMCs) has been presented. The Eshelby's tensor for an ellipsoidal inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted to model particles having mild or severe imperfect interface, and a multi-level damage model (Lee and Pyo, in press) in accordance with the Weibull's probabilistic function is employed to describe a sequential, progressive imperfect interface in the composites. The proposed micromechanical elastoplastic damage model is applied to the uniaxial, biaxial and triaxial tension loadings to predict the corresponding elastoplastic stress-strain responses. A series of parametric analysis are carried out to investigate the influence of model parameters on the elastoplastic behavior and progression of imperfect interface in the composites. Furthermore, the present prediction is compared with available experimental data in the literature to verify the proposed elastoplastic multi-level damage model. The elastoplastic stress-strain responses corresponding to various phase composite states are bounded by the responses of the two-phase and five-phase composite states. The higher phase composite states are lower than those with the lower phase composite states. A smooth and gradual transition from the perfectly bonded interface to the perfectly debonded interface occurs within the present framework, resulting in a nonlinear stress-strain response within the elastic range. The influence of the Weibull parameters S 0 and M B

B

and the compliance parameters  and  on the elastoplastic response and progression of imperfect interface in the composites is shown to be significant. Most particles are mild imperfect, severe imperfect or completely debonded in their interfaces even in early stage of loading with the lower S 0 . As the values B

B

of the compliance parameters  2 and  2 become higher, the effect of the severe imperfect interface on the elastoplastic stress-strain behavior of the composites is more pronounced. The predicted elastoplastic stress-strain behavior of PRDMCs is shown to have a very good correlation with the experimental data (Papazian and Adler, 1990). It should be noted that particles having different levels of imperfect interface (different levels of homogeneously distributed interfacial damage), giving rise to isotropic behavior, under general loading conditions are a certainly modeling assumption made in this framework. However, as deformations proceed under a uniaxial or biaxial loading, the composites progressively become transversely isotropic after evolutionary partial interfacial debonding as stated in Ju and Lee (2001). This issue is beyond the scope of the present work; nevertheless, it should be subject of future research for more accurate prediction of particulate composites behavior. It is obvious that the probabilities of perfectly bonded particles transforming to particles with mild imperfect interface or particles with mild imperfect interface transforming to particles with severe

Chapter 7. Conclusions

122

imperfect interface are dependent upon the two Weibull parameters S 0 and M and the internal stress of B

B

those particles. For simplicity, the internal stress of perfectly bonded particles has been considered as the controlling factor of the Weibull statistical function for the initiation of various stages of imperfect interface of the multi-phase composite state in the present study. The above assumption would, however, tend to lead to a slight overestimation of the volume fraction of particles that have reached Levels 2 and 3.

In Chapter 5, a 3D micromechanics-based evolutionary damage model has been proposed to predict the effective elastic behavior of continuous, fiber-reinforced brittle matrix composites with microcracks and imperfect interfaces. The Eshelby's tensor for a circular cylindrical inclusion with slightly weakened interface (Qu, 1993a, 1993b) is adopted to model continuous fibers with imperfect interfaces. A multi-level damage modeling process in accordance with the Weibull's probabilistic function is incorporated into the micromechanical framework to describe the sequential evolution of imperfect interfaces in the composites. The nucleation of microcracks is simulated by employing the continuum damage model proposed by Karihaloo and Fu (1989). Numerical examples are solved to illustrate the potential of the proposed damage model and comparisons between the present prediction with available experimental data are conducted to verify the accuracy of the proposed damage model. The progressive evolution of imperfect interface has significant influence on the overall elastic damage behavior of continuous, fiber-reinforced brittle matrix composites. The stress-strain responses due to uniaxial tension in the transverse direction are lower than those due to uniaxial tension in the longitudinal direction. It is found from the parametric study on the Weibull and compliance parameters that lower S 0 and higher  2 and  2 , which is related the interfacial strength of composites, lead to faster B

B

evolution of imperfect interface. Furthermore, as the values of the nucleation parameters c 1 becomes B

B

higher and c 2 becomes lower, the effect of the penny-shape microcrack nucleation on the stress-strain B

B

behavior of the composites is more noticeable. In comparison with the model without nucleation of microcracks, the present prediction with nucleation of microcracks is observed to be in good qualitative agreement with the experimentally obtained stress-strain curve (Luo and Daniel, 2000). It is observed from the numerical simulations and parametric analysis that the influence of the nucleation parameters c 1 and c 2 , Weibull parameter S 0 , and compliance parameters  2 and  2 is quite B

B

B

B

B

B

remarkable. Therefore, experimental verifications on those parameters will be needed for more accurate prediction of the elastic damage behavior of continuous, fiber-reinforced brittle matrix composites.

In Chapter 6, The micromechanics based constitutive model (Lee and Pyo, submitted for publication) for laminated composites, the multi-level damage model (Lee and Pyo, in press) for sequential progression

Chapter 7. Conclusions

123

of imperfect interface, and the continuum damage model for the nucleation of microcracks (Karihallo and Fu, 1989) are implemented into ABAQUS non-linear FE code using a user-subroutine UMAT to solve boundary value problems and to numerically characterize the elastic behavior of laminated composites. To verify the implemented computational model, computational studies are undertaken by comparing the predicted stress-strain curve and experimental data available in literature under uniaxial tension with various fiber orientations. Furthermore, three-point bending test and Iosipescu (V-notch) shear test are also numerically conducted and compared with experiment data to predict overall behavior and damage evolution. Numerical validations for widely different loadings and geometries have shown that the model can obtain satisfactory solutions that agree well with results available in the literature. In particular, multi-level elastic damage model naturally captures the gradual transition from previous damage level to next damage level and various types of damaged fibers exist simultaneously within the proposed framework. However, the damage model needs to be extended to accommodate other damage mechanism (e.g., fiber breakage, delamination) to realistically characterize the damage mechanisms in laminated composites.

As an extension of the present study, a series of characterization tests on a particulate composite and fiber reinforced composites need to be conducted in the future for an accurate calibration of the model parameters (S 0 , M, 1 , 1 ,  2 and  2 ). Data points will be sampled from the tests (e.g., stress-strain curves, B

B

load-deflection curves, etc.) and will be used to obtain an optimal set of the parameters by minimizing the sum of the squares of differences between the predicted and measured loads at all data points. Since the lower and upper limits of the parameters can be determined from a series of parametric analysis, this would be a nonlinear least-square constrained minimization problem. Alternatively, these parameters can be determined simultaneously with a trial-and-error procedure.

124

SUMMARY (IN KOREAN) 요 약 문

손상경계면을 갖는 복합재료의 탄성 및 탄소성 다단계 손상 모델

복합재료의 탄성 및 탄소성 거동과 손상 진전을 예측하기 위해 미세역학을 기반으로 한 다단계 손상 모델을 제안하였다. 먼저, 손상경계면을 갖는 입자강화 복합재료와 섬유강화 복 합재료의 손상을 모사하기 위해 미소한 손상경계면을 갖는 타원체형 강화입자 Eshelby 텐서 를 적용하였다. 변형이나 외부하중이 증가함에 따라 강화입자와 기지재료 사이의 손상경계면 이 발생한다. 이러한 손상경계면의 손상의 변화 정도를 실제적으로 모사하기 위하여 Weibull 확률함수를 기본으로 한 다단계 손상모델을 개발하였다. 다단계 손상 확률함수에서 손상의 진전은 강화입자의 내부응력에 의해 지배된다고 가정하였다. 연성의 기지재료를 갖는 복합재 료의 탄소성 거동을 예측하기 위하여 Ensemble-volume average와 first-order effects of eigenstrain 의 개념을 적용하여 유효 항복 기준을 유도하였다. 제안된 다단계 손상모델들을 적용하였다. 먼저, 취성의 기지재료를 갖는 입자강화 복합재 료, 연성의 기지재료를 갖는 입자강화 복합재료, 그리고 취성의 기지재료를 갖는 섬유강화 복 합재료를 1축, 2축, 3축 응력 하에서의 응력-변형률 관계를 조사하였다. 모델상수들이 거동에 미치는 영향을 조사하기 위해 변수검증을 실시하였다. 또한, 문헌조사를 통하여 얻은 기존 실 험결과와의 비교를 통하여 제안된 모델들의 검증을 실시하였다. 최종적으로 비선형 유한요소 프로그램인 ABAQUS에 적용하여 다양한 경계조건을 갖는 적층된 복합재료의 거동을 예측하 였다.

125

REFERENCES [1]

Aboudi, J., “Damage in composites modeling of imperfect bonding,” Composites Science and Technology, Vol. 28, pp. 103-128, 1987.

[2]

Achenbach, J. D. and Zhu, H., “Effect of interfacial zone on mechanical behavior and failure of fiber-reinforced composites,” Journal of the Mechanics and Physics of Solids, Vol. 37, pp. 381-393, 1989.

[3]

Achenbach, J. D. and Zhu, H., “Effect of interphases on micro and macromechanical behavior of hexagonal-array fiber composites,” Journal of Applied Mechanics, Vol. 57, pp. 956-963, 1990.

[4]

Ahn, B. K., Curtin, W. A., Parthasarathy, T. A., Dutton, R. E., “Criteria for f deflection/penetration criteria for fiber-reinforced ceramic matrix composites,” Composites Science and Technology, Vol. 58, pp. 1775-1784, 1998.

[5]

Arsenault, R. J., “The strength of aluminum alloy 6061 by fiber and platelet silicon carbide,” Materials Science and Engineering, Vol. 64, pp. 171-181, 1984.

[6]

Basu, S., Waas, A. M., Ambur, D. R., “Prediction of progressive failure in multidirectional composite laminated panels,” International Journal of Solids and Structures, Vol. 44, pp. 2648-2676, 2007.

[7]

Benveniste, Y., “The effective mechanical behavior of composite materials with imperfect contact between the constituents,” Mechanics of Materials, Vol. 4, pp. 197-208, 1985.

[8]

Bhatnagar, N., Ramakrishnan, N., Naik, N. K., Komanduri, R., “On the machining of fiber reinforced plastic (FRP) composite laminates,” International Journal of Machine Tools and Manufacture, Vol. 35, pp. 701-716, 1995.

[9]

Bonfoh, N. and Lipinski, P., “Ductile damage micromodeling by particles’ debonding in metal matrix composite,” International Journal of Mechanical Sciences, Vol. 49, pp. 151-160, 2007.

[10]

Broughton, W. R., Kumosa, M., Hull, D., “Analysis of the Iosipescu shear test as applied to unidirectional carbon-fibre reinforced composites,” Composites Science and Technology, Vol. 38, pp. 299-325, 1990.

[11]

Chakrabarti, B. K. and Benguigui, L. G., “Statistical Physics of Fracture and Breakdown in Disordered System,” Clarendon Press, 1997.

Reference

[12]

126

Crouch, S. L. and Mogilevskaya, S. G., “Loosening of elastic inclusions,” International Journal of Solids and Structures, Vol. 43, pp. 1638-1668, 2006.

[13]

Curtin, W. A., Ahn, B. K., Takeda, N., “Modeling brittle and tough stress-strain behavior in unidirectional ceramic matrix composites,” Acta Materialia, Vol. 46, pp. 3409-3420, 1998.

[14]

Dai, L. H., Huang, Z. P., Wang, R., “Explicit expressions for bounds for the effective moduli of multi-phased composites by the generalized self-consistent method,” Composites Science and Technology, Vol. 59, pp. 1691–1699, 1999.

[15]

Davalos, J. F., Salim, H. A., Qiao, P., Lopez-Anido, R., Barbero, E. J., “Analysis and design of pultruded FRP shapes under bending,” Composites Part B: Engineering, Vol. 27, pp. 295-305, 1996.

[16]

Dannemann, K. A. and Mandell, J. F., “Tensile damage development in fiber-reinforced ceramic matrix composites,” Materials Science and Engineering A, Vol. 177, pp. 95-114, 1994.

[17]

Drabek, T. and Bohm, H. J., “Simulation of Ductile Damage in Metal Matrix Composites,” Proceedings of the 8th International Conference on Numerical Methods in Industrial Forming Processes. AIP Conference Proceedings, Vol. 712, pp. 1857-1862, 2004.

[18]

Duan, H. L., Wang, J., Huang, Z. P., Luo, Z. Y., “ Stress concentration tensors of inhomogeneities T

TT

TT

TT

with interface effects,” Mechanics of Materials, Vol. 37, pp. 723-736, 2005. TT

[19]

T

Duan, H. L., Yi, X., Huang, Z. P., Wang, J., “A unified scheme for prediction of effective moduli of multiphase composites with interface effects. Part I: Theoretical framework,” Mechanics of Materials, Vol. 39, pp. 81-93, 2007.

[20]

Geiger, A. L. and Welch, P., “Tensile properties and thermal expansion of discontinuously reinforced aluminium composites at subambient temperatures,” Journal of Materials Science, Vol. 32, pp. 2611-2616, 1997.

[21]

Ghahremani, F., “Effect of grain boundary sliding on anelasticity of polycrystals,” International Journal of Solids and Structures, Vol. 16, pp. 825–845, 1980.

[22]

Ghosh, S., Bai, J. Raghavan, P., “Concurrent multi-level model for damage evolution in microstructurally debonding composites,” Mechanics of Materials, Vol. 39, pp. 241-266, 2007.

[23]

González, C. and LLorca, J., “Mechanical behavior of unidirectional fiber-reinforced polymers under transverse compression: Microscopic mechanisms and modeling,” Composites Science and Technology, Vol. 67, pp. 2795-2806, 2007.

Reference

[24]

127

Grufman, C. and Ellyin, F., “Numerical modelling of damage susceptibility of an inhomogeneous representative material volume element of polymer composites,” Composites Science and Technology, in press.

[25]

Guo, S. J., “Stress concentration and buckling behaviour of shear loaded composite panels with reinforced cutouts,” Composite Structures, Vol. 80, pp. 1-9, 2007.

[26]

Gurvich, M. R. and Pipes, R. B., “Strength size effect of laminated composites,” Composites Science and Technology, Vol. 55, pp. 93-105, 1995.

[27]

Gurvich, M. R. and Pipes, R. B., “Probabilistic strength analysis of four-directional laminated composites,” Composites Science and Technology, Vol. 56, pp. 649-656, 1996.

[28]

Halpin, J. C. and Kardos, J. L., “The Halpin-Tsai equations: A review,” Polymer Engineering and Science, Vol. 16, pp. 344-352, 1976.

[29]

Hashin, Z., “Thermoelastic properties of fiber composites with imperfect interface,” Mechanics of Materials, Vol. 8, pp. 333-348, 1990.

[30]

Hashin, Z., “The spherical inclusion with imperfect interface,” Journal of Applied Mechanics, Vol. 58, pp. 444-449, 1991.

[31]

Hassan, N. M. and Batra, R. C., “Modeling damage in polymeric composites,” Composites Part B: Engineering, Vol. 39, pp. 66-82, 2008.

[32]

Herakovich, C.T., “Mechanics of Fibrous Composites,” John Wiley and Sons, New York, 1998.

[33]

Herrmann, H. J. and Roux, S., “Statistical Models for the Fracture of Disordered Media,” NorthHolland, 1990.

[34]

Huang, Z. M., “Failure analysis of laminated structures by FEM based on nonlinear constitutive relationship,” Composite Structures, Vol. 77, pp. 270-279, 2007.

[35]

Hwang, T. K., Hong, C. S., Kim, C. G., “Size effect on the fiber strength of composite pressure vessels,” Reviews of Modern Physics, Vol. 59, pp. 489-498, 2003.

[36]

Isichenko, M. B., “Percolation, statistical topography, and transport in random media,” Reviews of Modern Physics, Vol. 64, pp. 961-1043, 1992.

[37]

Jasiuk, I., Tsuchida, E., Mura, T., “The sliding inclusion under shear,” International Journal of Solids and Structures, Vol. 23, pp. 1373–1385, 1987. T

[38]

T

Ju, J. W. and Chen, T. M., “Micromechanics and effective moduli of elastic composites containing randomly dispersed ellipsoidal inhomogeneities,” Acta Mechanica, Vol. 103, pp. 103-121, 1994a.

Reference

[39]

128

Ju, J. W. and Chen, T. M., “Effective elastic moduli of two-phase composites containing randomly dispersed spherical inhomogeneities,” Acta Mechanica, Vol. 103, pp. 123-144, 1994b.

[40]

Ju, J. W. and Chen, T. M., “Micromechanics and effective elastoplastic behavior of two-phase metal matrix composites,” Journal of Engineering Materials and Technology, Vol. 116, pp. 310-318, 1994c.

[41]

Ju, J. W. and Ko, Y. F., “Micromechanical elastoplastic damage modeling for progressive interfacial arc debonding for fiber reinforced composites,” International Journal of Damage Mechanics, in press.

[42]

Ju, J. W., Ko, Y. F., Ruan, H. N., “Effective elastoplastic damage mechanics for fiber-reinforced composites with evolutionary complete fiber debonding,” International Journal of Damage Mechanics, Vol. 15, pp. 237-265, 2006.

[43]

Ju, J. W. and Lee, H. K., “A micromechanical damage model for effective elastoplastic behavior of ductile matrix composites considering evolutionary complete particle debonding,” Computer Methods in Applied Mechanics and Engineering, Vol. 183, pp. 201-222, 2000.

[44]

Ju, J. W. and Lee, H. K., “A micromechanical damage model for effective elastoplastic behavior of partially debonded ductile matrix composites,” International Journal of Solids and Structures, Vol. 38, pp. 6307-6332, 2001.

[45]

Ju, J. W. and Sun, L. Z., “A novel formulation for the exterior-point Eshelby’s tensor of an ellipsoidal inclusion,” Journal of Applied Mechanics, Transactions of ASME, Vol. 66, pp. 570-574, 1999.

[46]

Ju, J. W. and Tseng, K. H., “Effective elastoplastic behavior of two-phase ductile matrix composites: A micromechanical framework,” International Journal of Solids and Structures, Vol. 33, pp. 4267-4291, 1996.

[47]

Ju, J. W. and Tseng, K. H., “Effective elastoplastic algorithms for two-phase ductile matrix composites,” Journal of Engineering Mechanics, Vol. 123, pp. 260-266, 1997.

[48]

Jun, Z. and Haian, L., “The validity of the modified Mori-Tanaka method for composites with slightly weakened interface,” Acta Mechanica Solida Sinica, Vol. 10, pp. 108-117, 1997.

[49]

Karihaloo, B. L. and Fu, D., “A damage-based constitutive law for plain concrete in tension,” European Journal of Mechanics, A-Solids, Vol. 8, pp. 373-384, 1989.

[50]

Kerner, E. H., “The elastic and thermo-elastic properties of composite media,” Proceedings of the Physical Society. Section B, Vol. 69, pp. 808-813, 1956.

Reference

[51]

129

Khashaba, U. A., “In-plane shear properties of cross-ply composite laminates with different off-axis angles,” Composite Structures, Vol. 65, pp. 167-177, 2004.

[52]

Kilic, H. and Haj-Ali, R., “Elastic-degrading analysis of pultruded composite structures,” Composite Structures, Vol. 60, pp. 43-55, 2003.

[53]

Krajcinovic, D. and Vujosevic, M., “Intrinsic failure modes of brittle materials,” International Journal of Solids and Structures, Vol. 35, pp. 2487-2503, 1998.

[54]

Lamon, J., “A micromechanics-based approach to the mechanical behavior of brittle-matrix composites,” Composites Science and Technology, Vol. 61, pp. 2259-2272, 2001.

[55]

Lapczyk, I. and Hurtado, J. A., “Progressive damage modeling in fiber-reinforced materials,” Composites Part A: Applied Science and Manufacturing, Vol. 38, pp. 2333-2341, 2007.

[56]

Lee, H. K., “A computational approach to the investigation of impact damage evolution in discontinuously reinforced fiber composites,” Computational Mechanics, Vol. 27, pp. 504-512, 2001.

[57]

Lee, H. K. and Kim, B. R., “Numerical characterization of compressive response and damage evolution in laminated plates containing a cutout,” Composites Science and Technology, Vol. 67, pp. 2221-2230, 2007.

[58]

Lee, H. K. and Liang, Z., “Computational Modeling of the Response and Damage Behavior of Fiber-Reinforced Cellular Concrete,” Computers and Structures, Vol. 82, pp. 581-592, 2004.

[59]

Lee, H. K. and Pyo, S. H., “Micromechanics-based elastic damage modeling of particulate composites with weakened interfaces,” International Journal of Solids and Structures, Vol. 44, pp. 8390-8406, 2007.

[60]

Lee, H. K. and Pyo, S. H., “Multi-level modeling of effective elastic behavior and progressive weakened interface in particulate composites,” Composites Science and Technology, in press.

[61]

Lee, H. K. and Pyo, S. H., “A micromechanics-based evolutionary damage model for continuous, fiber-reinforced brittle matrix composites with microcracks and imperfect interfaces,” Composites Science and Technology, submitted for publication.

[62]

Lee, H. K. and Simunovic, S, “Modeling of progressive damage in aligned and randomly oriented discontinuous fiber polymer matrix composites,” Composites Part B: Engineering, Vol. 31, pp. 7786, 2000.

Reference

[63]

130

Lee, H. K. and Simunovic, S, “A damage constitutive model of progressive debonding in aligned discontinuous fiber composites,” International Journal of Solids and Structures, Vol. 38, pp. 875895, 2001.

[64]

Lee, H. K. and Simunovic, S, “A damage mechanics model of crack-weakened, chopped fiber composites under impact loading,” Composites Part B: Engineering, Vol. 33, pp. 25-34, 2002.

[65]

Lee, H. K. and Simunovic, S, “Prediction of crack evolution and effective elastic behavior of damage-tolerant brittle composites,” Computer Methods in Applied Mechanics and Engineering, Vol. 196, pp. 118-133, 2006.

[66]

Lee, H. K., Simunovic, S., Shin, D. K., “A computational approach for prediction of the damage evolution and crushing behavior of chopped random fiber composites,” Computational Materials Science, Vol. 29, pp. 459-474, 2004.

[67]

Liang, Z., Lee, H. K., Suaris, W., “Micromechanics-based constitutive modeling for unidirectional laminated composites,” International Journal of Solids and Structures, Vol. 43, pp. 5674-5689, 2006.

[68]

Liu, H. T., Sun, L. Z., Ju, J. W., “An interfacial debonding model for particle-reinforced composites,” International Journal of Damage Mechanics, Vol. 13, pp. 163-185, 2004.

[69]

Liu, H. T., Sun, L. Z., Ju, J. W., “Elastoplastic modeling of progressive interfacial debonding for particle-reinforced metal matrix composites,” Acta Mechanica, Vol. 181, pp. 1-17, 2006.

[70]

Liu, H. T., Sun, L. Z., Wu, H. C., “Monte Carlo simulation of particle-cracking damage evolution in metal matrix composites,” Journal of Engineering Materials and Technology, Transactions of ASME, Vol. 127, pp. 318-324, 2005.

[71]

Luciano, R. and Zinno, R., “A micromechanical approach for the analysis of damage in laminated composite structures,” Computers and Structures, Vol. 74, pp. 201-214, 2000.

[72]

Luo, J. J. and Daniel, I. M., “A cylinder model for characterization of deformation and damage development in a unidirectional composite,” Composites Science and Technology, Vol. 60, pp. 2791-2802, 2000.

[73]

Maa, R. H. and Cheng, J. H., “A CDM-based failure model for predicting strength of notched composite laminates,” Composites Part B: Engineering, Vol. 33, pp. 479-489, 2002.

[74]

Maimi, P., Camanho, P. P., Mayug, J. A., Davila, C. G., “A continuum damage model for composite laminates: Part II - Computational implementation and validation,” Mechanics of Materials, Vol. 39, pp. 909-919, 2007.

Reference

[75]

131

Miller, J. H., Liaw, P. K., Landes, J. D., “Influence of fiber coating thickness on fracture behavior of continuous woven Nicalon fabric-reinforced silicon-carbide matrix ceramic composites,” Materials Science and Engineering A, Vol. 317, pp. 49-58, 2001.

[76]

Mogilevskaya, S. G. and Crouch, S. L., “A Galerkin boundary integral method for multiple circular elastic inclusions with homogeneously imperfect interfaces,” International Journal of Solids and Structures, Vol. 39, pp. 4723-4746, 2002.

[77]

Nie, S. and Basaran, C., “A micromechanical model for effective elastic properties of particulate composites with imperfect interfacial bonds,” International Journal of Solids and Structures, Vol. 42, pp. 4179-4191, 2005.

[78]

Nieh, T. G. and Chellman, D. J., “Modulus measurements in discontinuous reinforced aluminum composites,” Scripta Metallurgica, Vol. 18, pp. 925-928, 1984.

[79]

Paar, R., Vallés, J.-L., Danzer, R., “Influence of fibre properties on the mechanical behaviour of unidirectionally-reinforced ceramic matrix composites,” Materials Science and Engineering A, Vol. 250, pp. 209-216, 1998.

[80]

Papazian, J. M. and Adler, P. N., “Tensile properties of short fiber-reinforced SiC/Al composites, Part I: effect of matrix precipitates,” Metallurgical Transactions A, Vol. 21, pp. 401-410, 1990.

[81]

Park, T., Kim, K., Han, S., “Linear static and dynamic analysis of laminated composite plates and shells using a 4-node quasi-conforming shell element,” Composites Part B: Engineering, Vol. 37, pp. 237-248, 2006.

[82]

Qu, J., “Eshelby tensor for an elastic inclusion with slightly weakened interfaces,” Journal of Applied Mechanics, Vol. 60, pp. 1048-1050, 1993a.

[83]

Qu, J., “The effect of slightly weakened interfaces on the overall elastic properties of composite materials,” Mechanics of Materials, Vol. 14, pp. 269-281, 1993b.

[84]

Qiu, Y. P. and Weng, G. J., “Elastic moduli of thickly coated particle and fiber-reinforced composites,” Journal of Applied Mechanics Transactions of the ASME, Vol. 58, pp. 388–398, 1991.

[85]

Ravi, K. R., Sreekumar, V. M., Pillai, R. M., Mahato, C., Amaranathan, K. R., Arul kumar, R., Pai, B.C., “Optimization of mixing parameters through a water model for metal matrix composites synthesis,” Materials and Design, Vol. 28, pp. 871-881, 2007.

[86]

Rodriguez, D. E. and Ochoa, O. O., “Flexural response of spoolable composite tubulars: an integrated experimental and computational assessment,” Composites Science and Technology, Vol. 64, pp. 2075-2088, 2004.

Reference

[87]

132

Schjudt-Thomsen, J. and Pyrz, R., “Overall creep modelling of short fibre reinforced composites with weakened interfaces and complex fibre orientation,” Mechanics of Materials, Vol. 32, pp. 349359, 2000.

[88]

Schoutens, J. E., Introduction to metal matrix composite materials, DOD Metal Matrix Composites Information and Analysis Center (MMCIAC) Tutorial Series, No. 272, 1982.

[89]

Sharma, P., Ganti, S., Bhate, N., “Effect of surfaces on the size-dependent elastic state of nanoinhomogeneities,” Applied Physics Letters, Vol. 82, pp. 535–537, 2003.

[90]

Sørensen, B. F., “Effect of fibre roughness on the overall stress-transverse strain response of ceramic composites,” Scripta Metallurgica et Materialia, Vol. 28, pp. 435-439, 1993.

[91]

Sørensen, B. F. and Talreja, R., “Analysis of Damage in a Ceramic Matrix Composite,” International Journal of Damage Mechanics, Vol. 2, pp. 246-271, 1993.

[92]

Sørensen, B. F., Talreja, R., Sørensen, O. T., “Micromechanical Analysis of Damage Mechanisms in Ceramic Matrix Composites During Mechanical and Thermal Loading,” Composites, Vol. 24, pp. 129-140, 1993.

[93]

Suemasu, H., Takahashi, H., Ishikawa, T., “On failure mechanisms of composite laminates with an open hole subjected to compressive load,” Composites Science and Technology, Vol. 66, pp. 634641, 2006.

[94]

Sudak, L. J., Ru, C. Q., Schiavone, P., Mioduchowski, A., “A circular inclusion with inhomogeneously imperfect interface in plane elasticity,” Journal of Elasticity, Vol. 55, pp. 19-41, 1999.

[95]

Sun, L.Z. and Ju, J.W., “Elastoplastic modeling of metal matrix composites containing randomly located and oriented spheroidal particles,” Journal of Applied Mechanics, Transactions of ASME, Vol. 71, pp. 774-785, 2004.

[96]

Sun, L.Z., Ju, J.W., Liu, H.T., “Elastoplastic modeling of metal matrix composites with evolutionary particle debonding,” Mechanics of Materials, Vol. 35, pp. 559-569, 2003.

[97]

Tabiei, A. and Sun, J., “Statistical aspects of strength size effect of laminated composite materials,” Composite Structures, Vol. 46, pp. 209-216, 1999.

[98]

Tabiei, A. and Sun, J., “Analytical simulation of strength size effect in composite materials,” Composites Part B: Engineering, Vol. 31, pp. 133-139, 2000.

[99]

Teng, H., “Transverse stiffness properties of unidirectional fiber composites containing debonded fibers,” Composites Part A: Applied Science and Manufacturing, Vol. 38, pp. 682-690, 2007.

Reference

133

[100] Tohgo, K. and Weng, G. J., “A progress damage mechanics in particle-reinforced metal-matrix composites under high triaxial tension,” Journal of Engineering Materials and Technology, Vol. 116, pp. 414-420, 1994. [101] Voyiadjis, G. Z. and Park, T., “Elasto-Plastic Stress and Strain Concentration Tensors for Damaged Fibrous Composites,” Studies in Applied Mechanics, Vol. 44, Damage and Interfacial Debonding in Composites, edited by Voyiadjis, G. Z. and Allen, D. H., Elsevier, Amsterdam, 1996, pp.81-106. [102] Vanswijgenhoven, E. and Van Der Biest, O., “The relationship between longitudinal stress and transverse strain during tensile testing of unidirectional fibre toughened ceramic matrix composites,” Acta materialia, Vol. 45, pp. 3349-3362, 1997. [103] Walpole, L. J., “A coated inclusion in an elastic medium,” Mathematical Proceedings of the Cambridge Philosophical Society, Vol. 83, pp. 495–506, 1978. [104] Wang, W. and Jasiuk, I., “Effective elastic constants of particulate composites with inhomogeneous interphases,” Journal of Composite Materials, Vol. 32, pp. 1391–1424, 1998. [105] Weeks, C. A. and Sun, C. T., “Modeling non-linear rate-dependent behavior in fiber-reinforced composites,” Composites Science and Technology, Vol. 58, pp. 603-611, 1998. [106] Weibull, W., “A statistical distribution function of wide applicability,” Journal of Applied Mechanics, Vol. 18, pp. 293-297, 1951. [107] Wu, Y. M., Huang, Z. P., Zhong, Y., Wang, J., “Effective moduli of particle-filled composite with inhomogeneous interphase Part I bounds,” Composites Science and Technology, Vol. 64, pp. 1345– 1351, 2004. [108] Zhao, Y. H. and Weng, G. J., “A theory of inclusion debonding and its influence on the stress-strain relations of a ductile matrix,” International Journal of Damage Mechanics, Vol. 4, pp. 196-211, 1995. [109] Zhao, Y. H. and Weng, G. J., “Plasticity of a two-phase composite with partially debonded inclusions,” International Journal of Plasticity, Vol. 12, pp. 781-804, 1996. [110] Zhao, Y. H. and Weng, G. J., “Transversely isotropic moduli of two partially debonded composites,” International Journal of Solids and Structures, Vol. 34, pp. 493-507, 1997. [111] Zhong, Z. and Meguid, S. A., “On the elastic field of a spherical inhomogeneity with an imperfectly bonded interface,” Journal of Elasticity, Vol. 46, pp. 91–113, 1997.

Reference

134

[112] Zhong, Y., Wang, J., Wu, Y. M., Huang, Z. P., “Effective moduli of particle-filled composite with inhomogeneous interphase Part II mapping method and evaluation,” Composites Science and Technology, Vol. 64, pp. 1353–1362, 2004. [113] Zhong, Z., Yu, X. B., Meguid, S. A., “3D micromechanical modeling of particulate composite materials with imperfect interface,” International Journal of Multiscale Computational Engineering, Vol. 2, pp. 493-507, 2004. [114] Zhou, C. W., Yang, W., Fang, D. N., “Mesofracture of metal matrix composites reinforced by particles of large volume fraction,” Theoretical and Applied Fracture Mechanics, Vol. 41, pp. 311326, 2004. [115] Zhu, H. and Achenbach, J.D., “Radial matrix cracking and interphase failure in transversely loaded fiber composites,” Mechanics of Materials, Vol. 11, pp. 347-356, 1991. [116] ABAQUS User's Manual, Version 6.5, Hibbitt, Karlsson & Sorensen, Inc., Pawtucket, RI, 2004.

ACKNOWLEDGEMENTS I rarely plan my research it plans me.

Max Perutz

I gratefully acknowledge the supervision and guidance of my advisor, Prof. Haeng-Ki Lee, and committee members, Prof. Jin-Keun Kim and Prof. Hyo-Gyoung Kwak. I am thankful to Prof. Taehyo Park of Hanyang University, who is the first person to instruct me on how to do academic study. I want to express my love and gratitude to my family and my love, Seong-Kum, for their support and care during the development of this work. I also wish to express my gratitude to my colleagues, Mr. Su-Yong Song, BongRae Kim, Sung-Kug Ha, Sang-Hyeon Cheong, Muhammad Afzal, Rudy Tawie, and Hyeong-Ki Kim at the KAIST for their support throughout this study.

감사의 글 어느덧 지난 2년을 돌아보는 시간이 되었습니다. 항상 훗날 뒤를 돌아보았을 때 후회하 는 일이 없어야겠다는 생각을 하지만 처음생각과 같은 결과를 얻기란 쉽지 않다는 것을 새삼 깨닫게 됩니다. 가장 먼저 2년 동안 부족한 제자에게 학문적 소양과 사회적인 가르침을 주신 이행기 교 수님께 깊은 감사의 마음을 전합니다. 많이 부족한 석사학위논문에 많은 가르침을 주신 김진 근 교수님과 곽효경 교수님께도 감사 드립니다. 또한 석사과정 동안 많은 지도를 아끼지 않 으신 카이스트 건설 및 환경공학과 모든 교수님들께도 감사의 말씀을 드립니다. 저에게 공부의 큰 뜻을 일깨워 주시고, 이끌어 주신 한양대 토목공학과 박대효 교수님께 머리 숙여 감사의 뜻을 전합니다. ‘훨훨 날아라’라고 말씀 해주시는 아버지, 항상 사랑으로 돌봐주시고 어머니, 저를 믿 고 후원해주는 형제들, 우리 가족 항상 감사하고 사랑합니다. 항상 바쁘다는 핑계로 자주 못 가도 투정하지 않고 나 하나만 믿고 따라준 성금양 사랑합니다. 지난 대학원 생활 동안 기쁜 일, 힘든 일, 슬픈 일, 즐거운 일을 항상 함께 해준 SAM Lab. 식구들에게 진심으로 고마움을 전합니다. 먼저 졸업하신 수용형(실험동에서 웃고 떠들 면서 실험했던 기억이 생생합니다. 얼렁 2세 만드셔요), 이 긴 학위논문이 시작할 수 있도록

도와준 랩장 봉래형, 연구실 궂은일 도맡아서 해준 연구실 살림꾼 성국이, 나이 어린 동생을 동기로 둔 죄로 고생 많았던 상현형(형 마음 다 압니다), 어디로 튈지 모르고 사고 잘치는 형기, 차세대 연구실 대표주자 은아, 많은걸 전수해 주지 못해 미안한 은화, 모든 식구들 여 러분 덕분에 이렇게 마무리를 웃으면서 할 수 있는 것 같습니다. I’m grateful for all the help given me by Brother Afzal & Rudy. I’ll pray for you & your family. 감사의 글에 제일 먼저 이름을 올리고 싶었지만 일에는 순서가 있는 것 같습니다. 광식 형! 형 덕분에 대전생활이 외롭지 않았습니다. 제게 큰 꿈을 그릴 수 있도록 도와주고, 형의 에너지를 직접 느낄 수 있어서 너무 좋았습니다. 고마운 나의 든든한 응원군 정식이, 열심히 군복무중인 종혁이, 열심히 공부중인 대환이, 병조, 세홍이, 항상 동네에서 받겨주던 종훈이, 종삼이, 승준이, 성민이, 원태, 영수, 영일이, 나의 소중한 친구들에게 감사의 말을 전하고 싶습니다. 석사 2년을 함께 해준 한밭회 식구들 준래형님, 범섭형, 강욱형, 영훈형, 준호형, 형주형, 영훈이, 승훈이, 효정이, 잊지 못할 추억이 많은 동기들 상민형, 주원형, 낙타문주, 성우, 인엽 이, 엠마, 윤정이, 지영이, 신애, 지현이, 저희 연구실 식구인 것 같은 호영형, 옥상에서 담배 같이 피던 창길이 여러분 덕분에 카이스트 생활이 외롭지 않았던 것 같습니다. 한양 토목쟁이들 근영이 아빠 인환형, 승환형, 수현형, 형섭형, 희찬이, 문규, 지원이, 현 진이, 승현형, 우헌형, 배수용이, 최석이, 현준, 재홍, 정훈, 류상, 무철, 병민, 수진, 준우 다들 토목계에서 한자리씩 하실 거라 믿습니다. 파이팅 입니다. 왕십리 술집에 받겨주는 마음씨 좋은 동아리 동기들 의태, 현태, 용민, 성철, 태한, 현정, 연희, 관희, 정혁, 현진, 희창, 호형 다들 바쁠 때라 자주 연락도 못하고 미안하구나 위에 이름이 빠졌다고 서운해 하거나 미워하지 말아주셔요. 저의 불찰이니 너그러운 마 음으로 이해해 주시고, 제가 이 자리에 있게 해주신 모든 분들 감사합니다. 사랑합니다.

Friendship is the inexpressible comfort of feeling safe with a person, having neither to weigh thoughts nor measure words. George Eliot

CURRICULUM VITAE Name

: Pyo, Suk-Hoon

Date of Birth

: February 18, 1982

Place of Birth

: Gochang, Korea

EDUCATION M.S.

Department of Civil and Environmental Engineering Korea Advanced Institute of Science and Technology (KAIST), Daejeon, Korea March 2006-February 2008

B.S.

Department of Civil Engineering Hanyang University, Seoul, Korea March 2000-February 2006

EXPERIENCE Graduate Research Assistant

March 2006 to December 2007

Department of Civil and Environmental Engineering, KAIST, Korea

LICENSE Engineer

Civil Engineering, Human Resources Development Service of Korea (License No. 05201070585T)

Engineer

Information Processing, Human Resources Development Service of Korea (License No. 052020219630)

LIST OF PUBLICATIONS Thesis 1.

Pyo, Suk-Hoon, Elastic and Elastoplastic Multi-Level Damage Models for Composites with Imperfect Interface, Master’s Thesis, Department of Civil and Environmental Engineering, Korea Advanced Institute of Science and Technology, Daejeon, Korea, February 2008.

Journal Papers 1.

Lee, H. K. and Pyo, S. H., “Micromechanics-based elastic damage modeling of particulate composites with weakened interfaces,” International Journal of Solids and Structures, Vol. 44, Issue 25-26, pp. 8390-8406, 2007.

2.

Lee, H. K. and Pyo, S. H., “Multi-level modeling of effective elastic behavior and progressive weakened interface in particulate composites,” Composites Science and Technology, in press.

3.

Lee, H. K. and Pyo, S. H., “An elastoplastic multi-level damage model for ductile matrix composites considering evolutionary weakened interface,” International Journal of Solids and Structures, in press.

4.

Lee, H. K. and Pyo, S. H., “A micromechanics-based evolutionary damage model for continuous, fiber-reinforced brittle matrix composites with microcracks and imperfect interfaces,” Composites Science and Technology, submitted for possible publication

5.

Lee, H. K., Pyo, S. H., Kim, B. R., “On joint strength, peel stress and failure modes in adhesively bonded double-strap and supported single-lap GFRP joints,” Composite Structures, submitted for possible publication

6.

Lee, H. K., Pyo, S. H., Kim, H. K., “A Micromechanics-based Elastic Constitutive Model for ParticleReinforced Composites Containing Weakened Interfaces and Microcracks,” Journal of the Computational Structural Engineering Institute of Korea, in press (in Korean).

Conference Papers 1.

Lee, H. K., Pyo, S. H., Kim, B. R., Cho, J. R., “Experimental study on mechanical behavior of adhesively bonded GFRP joints,” Conference on Korean Society of Composite Mechanics, Kumi, Korea, pp. 269-272, November 23, 2006 (in Korean).

2.

Lee, H. K. and Pyo, S. H., “A Micromechanics-based Elastic Model for Particle-Reinforced Composites Containing Slightly Weakened Interfaces,” Proceedings of COSEIK Annual Conference 2007, Seoul, Korea, pp. 441-444, April 12-13, 2007 (in Korean).

3.

Lee, H. K., Pyo, S. H., Kim, B. R., “Investigation on joint strength of adhesively bonded, supported single-lap GFRP joints,” Conference on Korean Society of Composite Mechanics, Seoul, Korea, pp. 45-48, April 28, 2007 (in Korean).

4.

Lee, H. K., Pyo, S. H., Kim, B. R., “Investigation on peel stresses of adhesively bonded double-lap and supported single-lap GFRP joints,” Third International Conference on Structural Engineering, Mechanics and Computation, Cape Town, South African, September 10-12, 2007.

5.

Lee, H. K., and Pyo, S. H., “A Numerical Approach to Multi-Level, Elastic Modeling of Weakened Particle-Reinforced Brittle Composites,” Proceedings of 20th KKCNN Symposium on Civil Engineering, Jeju, Korea, pp. 283-286, October 4-5, 2007.

6.

Lee, H. K., and Pyo, S. H., “Micromechanics-based Elastoplastic Multi-Level Damage Modeling of Ductile Polymeric Composites with Weakened Interface,” Proceedings of the 6th Korea-Japan Joint Symposium on Composite Materials, Pohang, Korea, pp. 115-116, November 1, 2007.

7.

Lee, H. K., and Pyo, S. H., “Study on Micromechanics-based Elastic and Damage Model for fiberreinforced composites,” 14th Guided Weapon Conference 2007, Daejeon, Korea, pp. 644-647 (in CDROM), November 1, 2007 (in Korean).

MS_dissertation-pyo-KAIST.pdf

Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. MS_dissertation-pyo-KAIST.pdf. MS_dissertation-pyo-KAIST.pdf.

4MB Sizes 5 Downloads 90 Views

Recommend Documents

No documents