failure analysis of a quasi-isotropic laminated composite ......with the increased use of laminated...

107
Failure Analysis of a Quasi-isotropic Laminated Composite Plate with a Hole in Compression Dr. O.H. Griffin, Jr. by Nirmal Iyengar ' Thesis submitted to the Faculty of the Virginia Polytechnic Institute and State University in partial fulfillment of the requirements for the degree of Master of Science in Engineering Mechanics Dr. Z. Gurdal, Chairman June 1992 Blacksburg, Virginia Dr. S.L. Hendricks

Upload: others

Post on 19-Feb-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

  • Failure Analysis of a Quasi-isotropic Laminated Composite Plate with a

    Hole in Compression

    Dr. O.H. Griffin, Jr.

    by

    Nirmal Iyengar

    '

    Thesis submitted to the Faculty of the

    Virginia Polytechnic Institute and State University

    in partial fulfillment of the requirements for the degree of

    Master of Science

    in

    Engineering Mechanics

    Dr. Z. Gurdal, Chairman

    June 1992

    Blacksburg, Virginia

    Dr. S.L. Hendricks

  • FAILURE ANALYSIS OF A QUASI-ISOTROPIC

    LAMINATED COMPOSITE PLATE WITH A

    HOLE IN COMPRESSION

    ABSTRACT

    The ability to predict failure of laminated composites in compression has been doggedly pursued

    by researchers for many years. Most have, to a limited extent, been able to predict failure for a

    narrow range of laminates. No means, as yet, exist for predicting the strength of generic laminates

    under various load conditions. Of primary concern has been the need to establish the mode at

    failure in compression. Even this has been known to vary for fiber and matrix dominated laminates.

    This study has been carried out to analyze the failure of specimens with a hole made of laminates

    with various quasi-isotropic stacking sequences. Different stacking sequences are achieved by

    rotating a [±45/90/0]9 stacking sequence laminate as a whole with respect to the loading axis of the

    specimens. Two- and three-dimensional finite element models, using commercial packages, were

    ge~erated to evaluate the stresses in the region of the hole. Two different compressive failure

    prediction techniques based on distinctly different failure modes have been used. The validity of

    these techniques was measured against experimental data of quasi-isotropic specimens tested.

    To investigate the applicability of the failure criteria for different laminated composite plates,

    analyses were repeated for specimens with different stacking sequences resulting from the rotation

    of the laminate.

    The study shows the need for the use of three-dimensional analysis of the stress state in the

    vicinity of the hole in order to be able to accurately predict failure. It also shows that no one mode

    of failure is responsible for limiting the strength for all laminate orientations but rather the mode

    changes with change in stacking sequence. The failure of the laminate with a hole was seen to be

    very sensitive to the stacking sequence. Experimental data presented also shows that the peak

    strength obtainable from the laminate analyzed, [±45/90/0]9 , is going to be in the off-axis

    configuration rather than on-axis placement of the stacking sequence with respect to the loading

    direction.

  • ! ~

    I I I I

    I I I I

    ACKNOWLEDGEMENTS

    The author would like to thank the following people for their, direct or indirect, contribution to this

    work:

    Dr.Z. Gurdal, committee chairman, for taking on the author as his student and for

    his advice and patience during the course of this work.

    Ors. O.H. Griffin, Jr. and S.L. Hendricks for serving on the authors graduate

    committee.

    Many colleagues of the author who helped him through the learning curve on

    PATRAN and ABACUS, notably and

    and who were innocent victims of the authors

    frustrations of trying to understand compressive failure, and who in spite of it all

    provided valuable insights to the problem.

    and for their hospitality during the defense.

    The •permanent• occupants of ESM Computer Lab namely,

    and who helped maintain a sense of humour when all was not

    going right.

    , the author's in-laws who provided invaluable family support

    when it was most needed.

    and who provided the author and his family with valuable

    professional and domestic assistance whenever needed.

    , of NKF Engineering, who showed the author that curiosity is a

    trademark of a good engineer, and that the need to learn should never stop.

    for their yoghurt, after all one has to eat, and healthy too.

    Finally the last but, most definitely, not the least, to the authors wife , who

    provided the moral and financial support through it all, his daughter , who

    is still trying to figure out why she sees her father only every three weeks, and

    their dogs, . and , who provided the family security and companionship

    during the past few years.

    iii

  • TABLE OF CONTENTS

    1 INTRODUCTION................................................................................................................... 1

    1.1 Compression Failure of Laminated Composites............................................................. 2 1.2 Failure of Panel with Stress Raisers............................................................................... 3 1.3 Review of Literature......................................................................................................... 3

    1.3.1 Failure Response of Laminated Composites in Compression.................................. 4 1.3.2 Failure Criteria for Laminated Composites

    with Stress Concentrations in Compression ................................................................ 10 1.3.3 Summary of Literature Review ................................................................................... 13

    1.4 Compression Strength Test Data.................................................................................... 13 1.5 Scope of Investigation ...................................................................................................... 19

    2 :'FINITE ELEMENT MODELS ................................................................................................ 21

    2.1 Introduction ....................................................................................................................... 21 2.2 Model Description ............................................................................................................ 23

    2.2.1 Two-Dimensional Model. ............................................................................................ 23 2.2.2 Three-dimensional Model. .......................................................................................... 25

    2.3 Boundary Conditions ........................................................................................................ 27 2.3.1 Symmetry .................................................................................................................... 27 2.3.2 Traction Free Boundary Conditions........................................................................... 29

    2.4 Comparison Between 2-D and 3-D Models .................................................................... 32

    3 FAILURE CRITERIA ............................................................................................................. 40 3.1 Introduction ....................................................................................................................... 40 3.2 Mechanical Properties ...................................................................................................... 41

    3.2.1 Characteristic Distance ............................................................................................... 42 3.2.2 Buckling Wavelength .................................................................................................. 43

    3.3 Failure Prediction ............................................................................................................. 45

    iv

  • TABLE OF CONTENTS (contd)

    3.3.1 Kinking Model. ............................................................................................................ 45

    3.3.2 Delamination Criterion .................................. ;················-'············································ 50

    4 STRESS ANALYSIS ............................................................................................................. 52 4.1 Introduction ....................................................................................................................... 52

    4.2 Two-Dimensional Model Stresses ................................................................................... 57

    4.3 Three-Dimensional Model Stresses................................................................................ 63

    4.3.1 Three-Dimensional Model Stress Correction ............................................................. 75

    5 FAILURE ANALYSIS ........................................................................................................... 80

    5.1 Introduction ....................................................................................................................... 80

    5.2 Initiation of Kinking ........................................................................................................... 82

    5.2.1 Two-dimensional Model. ............................................................................................. 82

    5.2.2 Three-dimensional Model.. ......................................................................................... 83

    5.3' Inception of Delamination ................................................................................................ 90

    6 CONCLUSIONS AND RECOMMENDATIONS .................................................................... 93

    7 REFERENCES ...................................................................................................................... 95

    VITA..................................................................................................................................... 99

    v

  • LIST OF FIGURES

    Figure 1-1: Kink Band Formation .................................................................................................. 7

    Figure 1-2: Laminate Configuration ............................................................................................... 14

    Figure 1-3: Compressive Strength of [45/-45/90/0bs Laminate .................................................... 16

    Figure 2-1: Plate Specimen Dimensions and Loading ................................................................. 22

    Figure 2-2: Finite Element Model. ................................................................................................. 24

    Figure 2-3: Comparison of cr33 and -r23 in Half and Quarter Symmetric models .......................... 28

    Figure 2-4: 3-D finite element model - undefromed state ............................................................ 30

    Figure 2-5: 3-D finite element model - deformed state ................................................................ 31

    Figure 2-6: Through the thickness strain distribution for 2-D and 3-D models ........................... 33

    Figure 2-7a: Comparison of strains between 2-D and 3-D models, 0=0 and =O ....................... 34 Figure 2-7b: Comparison of strains between 2-D and 3-D models, 0=90 and =O ..................... 34 Figure 2-7c: Comparison of strains between 2-D and 3-D models, 0=-45 and =O .................... 35 Figu~.e 2-8: Comparison of .stresses obtained from 2-D and 3-D Models at q, = 0.... ...... ......... 36 Figure 2-9: Comparison of stresses obtained from 2-D and 3-0 Models at q, = 45............ ...... 36 Figure 2-10: Comparison of stresses obtained from 2-D and 3-D Models at q, = 90 .................. 37 Figure 2-11: Comparison of stresses obtained from 2-D and 3-D Models at q, = -45 ................. 37 Figure 3-t: Model of a fiber under combined axial and shear forces .......................................... 44

    Figure 3-2: Forces on fiber adjacent to hole ................................................................................ 49

    Figure 4-1a: Distribution of cr11 around hole at mid laminae at = 0 .......................................... 53

    Figure 4-1b: Distribution of -r12 around hole at mid laminae at = 0 .......................................... 53

    Figure 4-2a: Distribution of cr11 around hole at mid laminae at = 90 ......................................... 54

    Figure 4-2b: Distribution of -r12 around hole at mid laminae at = 90 ......................................... 54

    Figure 4-3a: Distribution of cr11 around hole at mid laminae at q, = -45 ........................................ 55 Figure 4-3b: Distribution of -r12 around hole at mid laminae at = -45 ........................................ 55

    Figure 4-4a: Distribution of cr11 around hole at mid laminae at= 45 ......................................... 56

    Figure 4-4b: Distribution of -r12 around hole at mid laminae at = 45 ......................................... 56

    Figure 4-5a: Distribution of cr11 around hole at mid laminae at = 15 ......................................... 58

    Figure 4-5b: Distribution of -r12 around hole at mid laminae at = 15 ......................................... 58

    vi

  • LIST OF FIGURES (contd)

    Figure 4-6a: Distribution of a11 around hole at mid laminae at cl>= 0 .......................................... 59 Figure 4-6b: Distribution of t 12 around hole at mid laminae at cl> = 0................................... ....... 59 Figure 4-6c: Distribution of t 13 around hole at mid laminae at cl> = 0.......................................... 60 Figure 4-7a: Distribution of 0 11 around hole at mid laminae at cl>= 45 ......................................... 61 Figure 4-7b: Distribution of t 12 around hole at mid laminae at cl>= 45 ......................................... 61 Figure 4-7c: Distribution of t 13 around hole at mid laminae at cl>= 45 ......................................... 62 Figure 4-8: Variation in axial stress, a 11 with laminate orientation (cl>= Oto 45) ......................... 64 Figure 4-9: Variation in axial stress, a11 with laminate orientation (cl>= Oto -45) ........................ 64 Figure 4-10: Variation in shear stress, t 12 with laminate orientation (cl>= Oto 45) ...................... 65 Figure 4-11: Variation in shear stress, t 12 with laminate orientation (cl>= Oto -45) ..................... 65 Figure 4-12: Variation in shear stress, t 13 with laminate orientation (cl>= Oto 45) ...................... 66 Figure 4-13: Variation in shear stress, t 13 with laminate orientation (cl>= Oto -45) .................... 66 Figure 4-14a: Distribution of a11 around hole at mid interfaces at cl>= 0 ...................................... 67 Figure 4-14b: Distribution of t 12 around hole at mid interfaces at cl>= 0 ...................................... 67

    z Figure 4-14c: Distribution of t 13 around hole at mid interfaces at cl>= 0 ...................................... 68 Figure 4-15a: Distribution of a11 around hole at mid interfaces at q, = 0 ...................................... 69 Figure 4-15b: Distribution of t 12 around hole at mid interfaces at cl> = 0 ...................................... 69 Figure 4-15c: Distribution of t 13 around hole at mid interfaces at cl> = 0 ...................................... 70 Figure 4-16: Variation in axial stress, a 11 , at midplane, with laminate orientation

    (cl> = O to 45) ................................................................................................................. 72 Figure 4-17: Variation in axial stress, a11 , at midplane, with laminate orientation

    (cl>= 0 to -45) ................................................................................................................ 72 Figure 4-18: Variation in shear stress, t 12, at midplane, with laminate orientation

    (cl> = Oto 45) ................................................................................................................. 73 Figure 4-19: Variation in shear stress, t 12, at midplane, with laminate orientation

    (cl>= Oto -45) ................................................................................................................ 73 Figure 4-20: Variation in shear stress, t 13, at midplane, with laminate orientation

    (cl>= Oto 45) ................................................................................................................. 74 Figure 4-21: Variation in shear stress, t 13, at midplane, with laminate orientation

    (cl>= 0 to -45) ................................................................................................................ 74

    vii

  • LIST OF FIGURES (contd)

    Figure 4-22a: Comparison of 't°12 and t 12 for 9=0 and ip:o .......................................................... 77 Figure 4-22b: Comparison of 't°12 and t 12 for 9=0 and' fP=15 ......................................................... 77 Figure 4-22c: Comparison of t 0 12 and t 12 for 9=0 and fP=15 ......................................................... 78

    Figure 5-1: Compressive failure distribution based on fiber kinking (2-D FEM Model) .............. 81

    Figure 5-2: Compressive failure distribution based on fiber kinking (3-D FEM Model).............. 84

    Figure 5-3: Compressive failure distribution based on fiber kinking using corrected t 12 • •• • ••••••• 85

    Figure 5-4: Initiation of matrix failure at mid laminae (3-0 FEM Model) ..................................... 87

    Figure 5-5: Compressive failure distribution based on fiber kinking at the interfaces ................ 88

    Figure 5-6: Compressive failure distribution based on fiber kinking at the interfaces

    using corrected t 12 ••• •••••••••••••••••••• •• •••••••••• ••••••••••• •••••••• ••••••• ••••••• ••••• •• ••••• ••••• •• • •••• •• • •••• ••• 89

    Figure 5-7: Inception of delamination ........................................................................................... 91

    LIST OF TABLES

    Table 1-1 : Experimental Results of [±45/90/0]38 Laminate Specimens ....................................... 15

    Table 2-1: Material Properties of AS1/T3201 .............................................................................. 23

    Table 3-1: AS1/T3201 Lamina Strengths .................................................................................... 42

    viii

  • 1 INTRODUCTION

    With the increased use of laminated composite materials for primary structural components, there

    exists a need to fully understand their capabilities. What makes laminated composites attractive

    as structural materials is their ability to be tailored to achieve specific material properties and

    strength for a given laminate configuration.

    For a material to be successfully utilized for structural purposes it must satisfy certain fundamental

    requirements; one being the ability to maintain structural integrity in the presence of discontinuities.

    Discontinuities, in the form of holes, are necessary in structural panels to allow for joining, access

    openings for maintenance, and piping and electrical penetrations. The existence of these

    discontinuities produce areas of local stress concentration, and it is necessary to be able to predict

    the response of these panels under various loading conditions if they are to be used for structural

    purposes.

    Experimental research [1-4] has shown the detrimental effect of discontinuities such as notches,

    holes etc., on the load carrying capacity of laminated composites. The effect of holes on the load

    carrying capacity is greater in laminated composites than in metals. Being made of brittle fibers,

    the ability of high performance laminated composites to undergo local plastic deformation in an

    area of high stress concentration is limited [1]. It has also been shown [3-10] that the failure

    mechanism is more complex in compression loading than in tension. The modes of failure in

    compression are particularly sensitive to the presence of holes, notches, and other similar stress

    raisers. Once the mechanism of failure propagation can be identified and the load carrying

    capability predicted, the material can then be tailored to accommodate designed discontinuities.

    Also the effect of discontinuities, introduced subsequent to initial design, can readily be assessed.

    Introduction 1

  • 1.1 Compression Failure of Laminated Composites

    Prediction of strength of a laminated composite depends on two factors, which are material strength

    and stacking sequence. The material strength of a laminate is dependent not only on the properties

    of its constituents, but also on a number of other faqors governed by the manufacturing process.

    Flaws incorporated during manufacturing that can affect laminate strength are non-uniform fiber

    distribution, fiber waviness, and matrix voids. Another factor whose effect on strength is not known

    but at present is under investigation is fiber-matrix interaction. Due to the number of ingredients

    influencing compressive strength there is no completely satisfactory account, as yet, of

    compressive failure [6].

    In order to assess the compressive strength of a laminate, it is necessary that the mode of failure

    be understood. Studies of failure on plates in compression [1,2, 11-26], for a variety of laminates,

    accompanied by experimental verification, have been presented over the years. None of the

    predictions, however, seem general enough to be able to predict the failure over a range of

    laminate configurations. There does seem to be some consensus on the nature of failure being

    localized and was initially attributed to a loss of fiber stability [11-21 ].

    Unlike their isotropic counterparts where stability, whether global or local, is based on geometric

    considerations, stability of laminated composites is to be evaluated at global, macro and micro

    levels [3]. Due to the makeup of fiber reinforced composites, inhibiting the loss of stability at the

    global (orthotropic failure) and macro (lamina failure) levels, will precipitate it at the micro

    (fiber/matrix failure) level. At the micro level the loss of stability is, as mentioned earlier, only to a

    limited extent dependent on the geometry of the fiber. Research [2,22-24] has shown evidence that

    fiber failure may not be due to this •1oss of stability• at the micro level, but rather a phenomenon

    called fiber •kinking• based on observations of kink bands in the failed specimens.

    The theory of compressive strength of a laminate being dependent on kinking has been studied

    by many researchers [2,4,9,10,26]. Experimental evidence has shown [15,17,20,21] that the

    kinkbands originate at a free edge and propagate towards the center of the plate eventually

    resulting in catastrophic failure. Kinking is also said to initiate from a defect, notch, or stress raiser,

    and grows in a manner analogous to a dislocation. Experimental evidence also seems to indicate

    that the initiation of kinking and subsequent catastrophic failure, are often not very far apart, thus

    making it difficult to chart the progress of this phenomenon [21 ]. Recent studies, however, indicate

    Introduction 2

  • that failure in compression cannot be attributed to one mode for all laminates, as suspected earlier,

    but is more likely to be progressive in nature [2,27]. These researchers (2,27] have recorded

    delamination along with kinking in specimens in which the failure process was partly captured by

    interrupting loading prior to catastrophic failure. The exact order and effect of one failure mode on

    another has never been evaluated.

    1.2 Failure of Panels with Stress Raisers

    As mentioned earlier the presence of stress raisers is detrimental to the load carrying capacity of

    a laminated composite panel in compression. The failure of these panels has, in general, been

    assessed at the macro and micro levels. The studies on the macro level use a fracture mechanics

    approach to obtain Stress Correlation Factors (SCF) or Stress Intensity Factors (SIF) for a variety

    of laminates. Experimental studies [28] Were also carried out to evaluate the effect of finite width

    and hole size. These studies were confined to the lo5s in load carrying capacity and did not

    address the mode of failure. To provide a generalized assessment of the failure, researchers have

    focused on the micromechanics in order to evaluate the initial mode of failure [4,9, 1 O].

    ~-

    Most discussions on failure have been restricted to unidirectional loading. Many practical loading

    situations are unlikely to be unidirectional and the existence of combined compressive and shear

    loads is a reality. Even under unidirectional loading shear stresses generated due to complex

    stresses existing at edges of stress raisers [9, 1 O] brings up the need to understand the effect of

    combined loading on such panels with geometric discontinuities. The response of a laminated plate

    to combined loading can be simulated by analyzing a plate with a hole. The hole or other

    geometric discontinuities provide the complex stress state to be analyzed, and the failure process

    can be scrutinized better by focussing on the region around the hole.

    1.3 Review of Literature

    It is apparent that the failure process in compression is very complex. Lack of knowledge of the

    micromechanical behavior of a laminated composite plate under compression inhibits proper

    understanding of the failure process. This, however, has not deterred researchers in attempting to

    analyze the problem. Excellent surveys of existing literature have been presented by Gurdal (4],

    Greszczuk (20], and Shuart (29] to name a few. However, in order to bring the task on hand into

    Introduction 3

  • perspective the process is repeated. The literature survey is discussed in two parts, namely, the

    failure response and the failure criteria of laminated composites in compression. In the former, the

    modes of failure in compression with experimental evidence are discussed. In the latter, analytical

    models and criteria used to predict failure of laminated composites with geometric discontinuities

    have been discussed

    1.3.1 Failure Response of Laminated Composites in Compression

    For the past two decades, since it was first recognized, the failure of a laminated composite in

    compression was attributed to short wavelength buckling (microbuckling) and had been the focus

    of many researchers. Most analytical work done so far is based on the models proposed by Rosen

    [11 ]. He proposed two modes under which microbuckling could take place, the extension mode and

    the shear mode. Rosen's model were based on the stability of a beam (representing the fiber) on

    an elastic foundation (representing the matrix). Analytically he presented the two failure modes as

    follows:

    ~ension Mode: (1.01)

    Shear Mode: (1.02)

    where:

    a,cr - Compressive strength of composite Ei. Em - Elastic modulus of fiber and matrix respectively V1 - Volume fraction of fiber

    Gm - Shear modulus of matrix

    h - Width of fiber

    L - Buckling wavelength of fiber

    m - Fiber buckle mode number

    Rosen [11] neglected the second term of the latter equation on the basis of its magnitude; the

    argument being that the buckling wave length to fiber diameter ratio was very large. Thus reducing

    Introduction 4

  • the shear mode equation to :

    (1.03)

    Rosen [11] however recognized that the predictions obtained by his models overestimated the

    strength of the laminates. He attributed this difference to the possible loss of shear modulus

    resulting from plasticity of the matrix [6]. Schuerch [12] also obtained similar equations as Rosen,

    but he extended his model to include inelastic microbuckling as he noted that the strain levels

    exceeded the yield strain of his matrix.

    Hayashi [13] introduced the concept of "shear instability" associated with compressive loading. He

    proposed the idea that in the case of structural members with high flexural rigidity compared to

    shear rigidity, buckling along with shear deformation will take place. The buckling strength is

    therefore attributed largely to the shear modulus of the laminate. Foye [14] was also soon to come

    to the conclusion that, for unidirectional composites, the ultimate strength in longitudinal

    compression is limited by. their shear modulus. He obtained a modified stress strain law for an

    element under shear, indicating that in compression the effective shear modulus would decrease

    With the increase in load.

    (1.04)

    At failure the element would experience a complete loss of shear stiffness resulting in shear

    instability or crippling failure. Noting that his predicted strengths were higher than experimental

    values, he included in his analysis model the effect of voids and matrix fillers.

    In order to fully understand the behavior of a fiber in an elastic matrix Hermann and Mason [15]

    and Sadowsky et al. [16] experimented with single fibers isolated in a matrix to study short

    wavelength buckling of the fiber. Hermann and Mason [17] found that predicted failure of their

    model related well to Rosen's [11] extension mode. They also showed that the initial waviness of

    the beam (fiber) had a significant impact on failure. Sadowsky et al. [16] took into consideration

    residual thermal stresses, sensing that these stresses could cause waviness and fiber buckling and

    came to the conclusion that shear deformation had negligible effect on the compressive strength.

    Introduction 5

  • They also measured the strain during buckling and found that the predicted strains were more than

    an order of magnitude greater than that measured experimentally. Crawford [17) was the first to

    incorporate initial waviness of the lamina into the expression of stiffness. This resulted in

    interlaminar stresses being generated. Based on this, he concluded that the laminate failure was

    a consequence of these interlaminar stresses, rather than short wavelength buckling, though the

    exact mode of such a failure was not suggested.

    To further study the effect of the matrix in fiber buckling, Lager and June [18) tested Boron fibers

    in two epoxy systems. They based their model on that of Rosen's [11 ], but replaced the Elongation

    and Shear Modulii in Equations 1.01 and 1.03 by their respective Tangent Modulii. Based on

    experimental data they then replaced the tangent modulii by an "effective" modulus which is

    defined as a product of an "influence coefficient" and the material modulus. The value of the

    influence coefficient was determined experimentally to be 0.63. They also established that the

    strength of a composite in compression is largely dependent on matrix strength. The study also

    showed that for fiber volume fractions less than 10%, failure correlated well with the extension

    model while for large volume fractions the shear mode was observed. Since most of the

    experimental work was still over predicting the failure of a laminated composite in compression De

    F~rran and Harris [19] conducted experiments on polyester reinforced steel wire. They questioned

    the validity of Rosen's premise, that planar buckling was occurring. They concurred with what Lager

    and June [19) had alluded to, in that the buckling had an out of plane component and so

    considered a three-dimensional helical buckling approach. The hypothesis, of microbuckling being

    affected by interlaminar stresses, having been presented it was not until much later that Guynn and

    Bradley [27] and Waas et al. [2] would provide experimental evidence that in- and out-of-plane

    microbuckling does indeed occur. This type of microbuckling could be a result of the interlaminar

    stresses.

    To investigate the effect of matrix properties on the strength of a laminated composite in

    compression Greszczuk [20,21,30) performed many experiments by varying the matrices and

    volume fractions and considering different failure modes to come to the following conclusions. For

    composites with a low-modulus resin, failure in compression is generally governed by

    microbuckling, however, as the resin modulus is increased the transformation to a non

    microbuckling failure takes place. His data showed the effect of fiber matrix interface strength on compressive strength indicating an interaction type failure. Nonlinear material and geometric effects

    of the fiber and matrix were not considered.

    Introduction 6

  • T' T' -····~

    _______ ...

    . 111 " 1111

    T'

    Figure 1-1: Kink Band Formation (After Ref. 22)

    7 Introduction

  • At this point, with still no means for accurately accounting for the compressive strength Berg and

    Salama [22) through their studies of fatigue of graphite epoxy in compression, introduced the

    concept of kinking, to laminated composites. They showed the existence of kink bands due to

    microbuckling, in the specimens failed as a result of compressive fatigue. They concluded that axial

    matrix cracking was essential to the initiation of microbuckling instability. This longitudinal cracking

    permits kink bands on conjugate planes to link up leading to catastrophic failure. They postulated

    a kink process by which microbuckled fibers undergo shear deformation to the point that they break

    off into bits (Figure 1-1) due to tensile failure of the fibers.

    Based on work by Berg and Salama [22), and Weaver and Williams [23), Evans and Adler [24)

    produced an exhaustive study on kinking. They minimized the plastic work done to obtain the kink

    inclination and minimized the elastic strain energy to determine the kink boundary. They also

    utilized a model for statistical fiber fracture and matrix enhancement to determine parameters and

    an expression for critical kink formation stress. This expression confirmed the need for high matrix

    strength to suppress kink formation. Chaplin [1] attributed the low compressive strength to the

    presence of a defect from which local failure can propagate. He suggested use of fracture

    mechanics approach observing that the failure occurs at one region and then rapidly propagates

    across the specimen. Defects which act as stress concentrators are responsible for the initial ·:;;·

    failure. Chaplin [1] compared the failure to shear instability that occurs in un-reinforced resin and

    so felt the term microbuckling to be inappropriate.

    Utilizing the above information Parry and Wronski [25) investigated the compressive strength of a

    plate with a notch. They supported the Berg and Salama [22) claim that the principal mode of

    failure at the notch was due to the formation of a kink band. Later, experiments carried out by

    Rhodes et al. [26), and Waas et al. [2] also support this assertion. Rhodes [26) described the failure

    sequence beginning with matrix failure at 85% load which was interfiber in nature. Following matrix

    failure, upon further loading shear crippling in plies of the same orientation were initiated resulting

    in specimen failure. Waas et al. [2] also observed failure at the hole edge in the load carrying ply

    (0 degree) and at the location of the maximum compressive stress. They also found the

    microbuckling to originate at the hole surface and persist well into the interior of the specimen

    leading to catastrophic failure.

    Departing from the traditional view of fiber failure, Shuart and Williams [31] conducted experiments

    on angle ply laminates, with ±45 degree and ±45 degree dominated plies, with holes to conclude that the failure of the former was due to matrix shear and the latter was a combination of matrix

    Introduction 8

  • failure and delamination. On close inspection of the damaged specimens they observed kink band

    formation and transverse matrix cracking. Continuing work in compression failure of laminated

    composites, Shuart [29] developed a model with the fibers in the lamina as a plate and the matrix

    acting as an elastic foundation. He used linear analysis to determine stress, strain and mode shape

    using short wave buckling criteria. He extended this study to include the effect of volume fraction '

    on the buckling stress and then incorporated the effect of fiber imperfections in a non-linear

    analysis. He considered,the failure of the lamina based on the buckling of the outer lamina, in-

    plane matrix shearing and interlaminar shear strains from the fiber imperfections. He also provided

    a method to evaluate the load carrying capacity of a O degree dominated lamina in compression.

    Shuart [32] extended his previous non-linear model to include the effects of out-of-plane ply

    waviness, in-plane fiber waviness and fiber scissoring. He used this model to predict the failure

    of [±9/•9ls. composite laminates in compression. Using experimental data, he showed his model

    to provided excellent agreement ,for laminates with fiber angles greater than 45 degrees, 9 > 45, while that for those with fiber angles less than 45 degrees, 9 < 45, was not as good. Again he found the dominant failure modes to be interlaminar shearing, in-plane matrix shearing and matrix

    compression failure.

    So far the failure of a laminated composite in compression had been attributed to one failure mode.

    The possible interaction of two modes of failure were first recorded by Guynn and Bradley [27].

    Guynn and Bradley [27] provided evidence, based on experiments on compressive failure of

    AS4/PEEK laminate specimens with open holes, of the existence of two failure modes but

    discounted their interaction. Based on their observations of specimens that were inspected on

    interruption of the tests, prior to catastrophic failure, that the principal mode of failure was shear

    crippling. They also found evidence of in- and out-of-plane microbuckling prior to the shear

    crippling, which was also accompanied by local delamination. They concluded that this local

    de lamination was a result of the shear deformation of the fibers, caused by microbuckling, however,

    they did not believe that the delamination propagated to become macroscopic delamination, until

    final compressive failure when large scale brooming was observed. Waas et al. [2] on the other

    hand observed, depending on the specimens tested, the formation of delamination buckling prior

    to actual final failure. The damage initiated by microbuckling led to delamination cracks and

    suQ&equently to catastrophic failure. The growth of the delamination was in large part dependent

    on the percentage of 0 degree plies. In laminates with a small percentage of 0 degree plies

    delamination size grew till the area buckled leading to failure. The larger percentage of O degree

    plies led to brittle failure while a small percentage of 0 degree plies led to a more ductile failure.

    Beyond failure initiation by microbuckling, the delamination and buckling growth stage was

    Introduction 9

  • influenced by the overall stiffness. Waas et al. [2] also presented initial 0 degree strains which

    seemed to initiate failure.

    1.3.2 Failure Criteria for Laminated Composites with Stress Concentrations in Compression

    The failure criteria of plates with cutouts, holes, and notches have been studied for years, however,

    most of them relate to tensile loading. Many of the criteria proposed have been based on failure

    at the laminate level without any emphasis on the failure modes. Whitney and Nuismer [3]

    continued work by Waddoups et al. [5] to obtain average and point stress criteria for the failure of

    laminated plates with holes. They presented characteristic distances for both criteria, based on

    work by Bowie [33] on isotropic material. They, however, postulated that the characteristic

    distances were constant for various laminates and discontinuities. Nuismer and Labor [7] extended

    the concepts established above to laminated plates under compressive loads. However, the

    characteristic distance was found to be much greater in compression than was observed in tension.

    They attributed this difference to the greater redistribution of stress prior to failure in compression.

    Garbo and Rogonowski [8] conducted experiments and provided strength analysis for composite ·:~

    laminates under compressive and tensile loading. They postulated that the failure of a composite

    laminate can be correlated with analytical predictions of point stress at a characteristic distance

    away from the edge of the stress concentration. Garbo and Rogonowski [8] also extended the

    above theory to predict strength of a laminate, by utilizing the point stress criteria to evaluate the

    strength of each of the laminae. They recommended that, since, in a laminated composite, the

    peak stress does not occur at the same point at the hole edge, it is necessary to evaluate the

    stresses adjacent to the hole to obtain the combination of stresses that will lead to failure.

    Gurdal [4] expanded the work on failure criteria of laminated plates with cutouts, holes and notches

    by combining micromechanics and the laminate point stress criterion. Using information provided

    by Rhodes et al. [26], on the occurrence of kinking at the boundary of a stress raiser and the work

    by Greszczuk [21,30] on point stress criteria for plates with holes, he recommended that the two

    could be coupled to evaluate the failure strength of plates with notches. He also noticed that most

    of the prior work had restricted itself to the effect of the local compressive stresses only. Supported

    by Greszczuk's [21] observations that microbuckling occurs at a higher load than experimentally

    obtained, Gurdal [4] suggested that the discrepancy may be caused by shear stresses induced by

    the presence of the discontinuity. He then used an energy formulation to analyze a beam on an

    Introduction 10

  • elastic foundation, subject to in-plane and shear loading and obtained equations for stress in the

    fiber due to bending. Details of the formulation are presented in Chapter 3. The model assumed

    failure due to bending rather than microbuckling instability, as had been the norm till then. The

    shear loading in the model helped account for the shear stresses generated due to the presence

    of the notch. To prove the validity of the model he used the same variational formulation to do a

    stability analysis to obtain a buckling expression similar to that of Rosen's [ 11 J microbuckling model in the shear mode. Whereas Rosen [11 J neglected the second term Gurdal [4J retained it. He

    argued that experimental evidence provided by Evans and Adler [24J had shown that the buckling

    wave length to fiber diameter ratio did not justify ignoring the second term in Equation 1.02. He

    used two-dimensional finite element models to obtain the stress distribution in notched plates and

    using experimental data he evaluated the characteristic distance for the quasi-isotropic laminate

    to be 1.5 mm for plates with notches. Similarly for plates with holes the distance was shown to be

    2.5 mm. Gurdal [4J extended his model of plates with cracks to represent those with holes as more

    data was available for that purpose. He supported his decision by correlating the analysis with

    experiments conducted by Haftka and Starnes [9J for laminates with varying degrees of orthotropy.

    The results, however, showed the predicted values to be 40% greater than the experimental

    values. Gurdal and Haftka [1 OJ extended the earlier work to predict failure of a plate with a hole

    }lnder combined loading. They modeled the plate so as to obtain predictions for various laminate

    orientations. This was done by rotating the principal axis of orthotropy of the laminate with respect

    to the load direction. The stresses around the hole were evaluated using techniques proposed by

    Savin [34J and corrected for finite width from Hong and Crews [28J. The study clearly showed the

    anisotropy of strength of a quasi-isotropic plate under compression. Gurdal and Haftka [1 OJ also

    concluded that for a plate under combined in-plane and shear loading, rotating the laminate axis

    of orthotropy in a direction opposite to the applied shear load, can increase its load carrying

    capacity, thus suggesting a possible use of orthotropy for strength improvements of plates with

    holes.

    Some other attempts to predict the compression strength of laminated composites plates with

    holes, using more deterministic approaches are described below. Bums et al. [35J analyzed the

    compressive failure of notched angle ply laminated composites utilizing a three-dimensional finite

    element model and compared the results with experimental data. A point stress method and

    classical quadratic tensor and maximum stress failure criteria were used to evaluate the failure.

    Burns et al. [35J showed that failure analysis at the hole edge gave extremely conservative

    strengths. They concluded that the lack of consistent results was due to the point stress criteria

    being used and that failure in a laminate was not dependent on point failure. They found that the

    Introduction 11

  • in-plane shear stress, 't12, was dominant in the midplane with the out-of-plane shear stress, 't13,

    dominant at the interface. Observing the points at which failure was initiated, they concluded, that

    a combination of the two shear stresses were required to initiate failure. The characteristic distance

    theory had limited success in predicting failure.

    Chang and Lessard [36] considered the progressive failure of a laminated composite containing

    a hole subject to compressive loading. The stress analysis was performed using a finite element

    model considering the material and geometric non-linearities based on finite deformation theory.

    Failure criteria were also formulated for first ply failure, based on a semi statistical method and a

    damage progression model. Property degradation models for various failure have been provided.

    Good correlation was found when compared with experimental data for the laminates in question.

    The model considered ply thicknesses, clustering and finite widths, however the possibility of

    delamination was not considered.

    In a very different approach, Soutis et al. [37] use a modified Linear Elastic Fracture Mechanics

    technique to predict failure of a Carbon Fiber-Epoxy laminate, [(±45/0J3]., with a hole. Based on

    experimental observations, they inferred that failure is initiated by microbuckling at the hole edge

    )Yhich they equated with a "crush zone". This crush zone was considered to grow stably under

    increasing load till a critical length was reached, when unstable growth results in catastrophic

    failure. Using a non-linear approach to crack growth they predicted good results for the

    compressive strength of the laminate, though their prediction of length of the microbuckled region

    (crush zone) was greater than observed. This technique did not in any way address progressive

    failure and, being based on a fiber failure mode only, may have its limitations when utilized for

    laminates without fiber dominant layers.

    Considering the hypothesis that interlaminar stresses may play a part in compression failure of

    laminated composites, and that the magnitude of these stresses [38] in the vicinity of a hole cannot

    be ignored, Lagace and Saeger [34] proposed a failure criteria based on the inception of

    delamination at a hole edge. They used a methodology based on an eigen function solution

    technique to evaluate the interlaminar stresses near a hole for a plate in tension. They then used

    Hashim-Rotem type quadratic failure criteria proposed earlier (Lagace and Brewer [39]) to

    determine the load at the onset of delamination. They plotted the load at onset of delamination

    versus the laminate orientation, ~. to show the anisotropic and asymmetric distribution of strength of the quasi-isotropic laminates. The failure criterion was based on a single point stress and no

    consideration was given to progressive failure initiated by either microbuckling or matrix cracking.

    Introduction 12

  • 1.3.3 Summary of Literature Review

    In summary, the literature review shows the extensive research done on the failure prediction of

    laminated composites in compression. Most of the research, however, has been restricted to the

    compressive strength of unidirectional laminates a~ has been based on the assumption of short

    wavelength fiber buckling (microbuckling) in the plane of the lamina. Attempts at correlating

    analytical with experimental data by incorporating manufacturing defects such as voids, fiber

    waviness and scissoring and debonding have been studied. Research has also covered the failure

    of the fiber as a result of kinking.

    Based on evidence that the mode of failure in compression is related to the type of laminate (fiber

    or matrix dominated), research on matrix initiated failures have also been done. Having accepted

    the fact that compression failure of a laminate is initiated at a notch or similar stress raiser

    researchers have developed models to predict compression strength of laminates with holes with

    reasonable success. The influence of interlaminar stresses on compression strength has also to

    a limited extent been covered, though it has long been suspected to play a part in the failure

    mechanism.

    Most of the above studies were based on planar failure models and those that studied interlaminar

    stresses did not include in-plane failure modes. As a result, in spite of the extensive work being

    done there is no general means of predicting failure of a laminated composite with a hole in

    compression.

    1.4 Compressive Strength Test Data

    As discussed in the previous section, existing techniques for the prediction of strength of laminated

    composite plates with holes were applicable to only specific laminate systems, with none being

    general enough to be applicable for a range of laminate configurations. The above prediction

    techniques are very sensitive to specific laminate configurations, whether unidirectional or angle

    plies. In order to assess the capability of predictive models for laminates in compression, it was necessary to obtain experimental data with which the analytical predictions of failure could be

    compared. Also in order that the generality of the analysis technique could be demonstrated, it was

    decided that the use of a quasi-isotropic laminate was necessary. Earlier models had been used

    for predicting the strength of unidirectional and angle ply laminates separately; the response of a

    Introduction 13

  • Figure 1-2: Laminate Configurations

    Introduction 14

    .I

  • Table 1-1: Experimental Results of [±45/9010],. Laminate Specimens

    Orientation Number of Off-axis Failure Loads Average Failure Angle Specimens of Specimens Loads

    ,

    -80 2 6728.8 7105.3 7481.7

    -67.5 4 6405.9 6640.2 6700.2 6875.0 6875.0

    -45 5 6446.8 6941.2 6703.2 6647.0 6813.96 6667.2

    -22.5 3 . 8296.5 7923.3 8078.1 8014.5

    -15 5 8126.5 7800.9 . 7919.9 7941.9 7803.0 7926.9

    -10 5 8155.2 8416.7 8256.8 8227.5 8218.9 8265.4

    -5. 5 7561.6 n38.o 7597.0 7492.0 7857.0 7336.1

    0 5 7200.0 7080.0 7199.8 7266.2 7122.8 7330.0

    10 5 7042.0 6n9.8 6940.7 6872.6 7094.1 6914.9

    22.5 5 6756.9 7078.8 6970.6 6693.4 7378.8 6945.2

    45 5 6929.6 7189.6 6740.5 6861.3 6325.4 6396.8

    Introduction 15

  • --T·--r·---]-1--r-

    l l

    i l

    i :

    : :

    : :

    l 1

    l l

    ~ ....................... i ........................ ! ....................... j··· ··················1·······--...... o-······f···············-···

    l l

    0 co

    0 C")

    0 0 C") I

    0 co I

    0 O> I

    G)

    - ca c ·-E ca ..J en C") ...... 0 - 0 0) - It) oi::t' I - It) -&

    -oi::t' .....

    c:: 0 -..::::; - 0 as .... 0

    .c -

    .... c::n

    - 0 Q) c G) ..

    15> c::

    -"' <

    ( G

    ) >

    ·c;; en G) .. c. E 0 CJ Cf) I .

    I T""

    G)

    I .. ~

    I c::n i!

    16

  • quasi-isotropic laminate is a combination of the above two configurations. Further the generality

    of the technique could be corroborated by applying it to various laminate orientations that will be

    described later.

    Results of a test program carried out earlier [40], for quasi-isotropic laminates of [±45/90/0b. and

    [±45/90/0k. stacking sequences are presented here (See Table 1-1 ). Specimens in Reference [40]

    were cut from a large panel with a [±45/90/0J. layup so that laminates with varying angles of

    laminate orientation, cl> (as defined in Figure 1-2), were obtained. That is, a specimen cut at cl>= 15 degrees would have a stacking sequence of [60,-30,-75,15]. with respect to the principal axes

    of the specimen. This layup is still quasi-isotropic, however it is suspected that the strength of the

    specimen would be different than a specimen cut at cl> = O degrees because of the change in the stacking sequence which would affect the local stresses around the hole. The loading direction was

    kept constant and the laminate orientation cl> changed; this was analogous to keeping the laminate orientation constant and varying loading angle thus simulating the effect of combined shear and

    axial loading on the laminate. The counter clockwise direction was considered to be a positive

    rotation and is referred to as the laminate orientation cj>. The specimens were 1.5 inches long, 1.0 inch wide and 0.133 inch thick (24 ply) with a circular hole of 0.20 inch diameter.

    'The compressive strength obtained at off-axis laminate orientations (cl>* 0) were normalized by the strength at cl> = O and shown in Figure 1-3. On plotting the normalized strengths, P/P 0 versus laminate orientation, cj>, the curves obtained for the specimens with two laminate configurations tested were practically coincidental with one another. The only difference between the two laminate

    configurations being the total number of plies, and therefore the total laminate thickness. For this

    reason the following discussion has been based on the experimental data points obtained from the

    testing of [±45/90/0b. laminate specimens only. The results of the experimental data shown in

    Figure 1-3 are described below:

    • The single roost distinctive feature of the failure envelope, at laminate orientations -90 <

    cl> < 90, is that its is anisotropic and asymmetrical about cl> = 0. By anisotropic, it is meant that the strength varies as the laminate is rotated about the z-axis.

    • The failure loads at cl> = 45 and cl> = -45 are approximately the same, within experimental scatter.

    • The trend at the extremes indicates that the load at cl> = 90 would be close to or equal to that at cl>= 0.

    • The failure load (strength) peaks at about cl> = -15 degrees and is approximately 12% greater than that at cl> = O and about 19% greater than the value at cl> = 15 degrees. This is rather unusual because based on CL T, the axial stiffness of the laminate does not

    change if it is rotated in the +15 or -15 degree direction. At the lamina level the axial

    Introduction 17

  • " I ~ If

    1:

    I I I

    stiffness coefficients, A1• are the same for a lamina with a fiber angle of +0 as that with -a. The difference, however, in the two laminate orientations,~= 15 and -15 is the stacking sequences. In these orientations the former becomes a [60,30,-75,15}. laminate and the

    latter [30,-60,75,-151. laminate respectively. As can be seen in these two laminate

    configurations, for each 0 in the former there exists a -0 in the latter and vice versa. So the lamina stiffness coefficients of the two are the same. The relative through-the-thickness

    locations of the two laminates are different. The stacking sequence seems be the only

    feature that probably accounts for the difference in compressive strengths of the two

    laminate orientations. It can be conjectured that the strength of the laminate in

    compression is not only a function of the total laminate stiffness but also one of the relative

    location, i.e. stacking sequence, of the various axial stiffness components. It is for this

    reason that it is essential to determine the three-dimensional stress state for this laminate

    in various orientations. A two-dimensional model is not able to capture these subtle

    differences.

    Another manner in which the stacking..sequence can be considered to have significant

    impact on compressive strength is in the through-the-thickness location of the primary load

    bearing laminae. This can be seen by comparing the orientations at angles in steps of 45

    degrees, that is at ~ = 0, 45, 90 and -45 degrees. It seems to indicate that when the primary load carrying laminae, i.e. one in which a = ~. is at the top and bottom of the laminate during rotation, there is a great reduction in the strength. For example, when ~ = 0 the laminate configuration is [±45/90/0J. while in the ~ = ±45 orientations they are

    [0/90/±45]. and [90/0/•451. respectively. In the former condition (~ = 0) the primary load

    bearing plies are at the midplane of the laminate while in the latter (~ = ±45) they are on or near the surface. This could result in out-of-plane buckling due to the lack of transverse

    support for the surface layers. This hypothesis is in part supported by experimental

    observations of Guynn et al. [27), who suggest that the out-of-plane buckling does occur

    and may be due to the influence of transverse stresses on the laminate.

    The asymmetry of compression strength has tremendous significance to real life applications. In

    order to optimize the design of a quasi-isotropic material for a structure that may experience

    tension and compression loading it may be necessary to utilize the material in an off-axis

    configuration. Sun and Zhou [41] have also presented research showing the anisotropic behavior

    of tensile strength in off-axis configurations of quasi-isotropic laminates. The angle of laminate

    orientation, ~. exhibiting maximum strength in tension, was different from that shown in Figure 1-3 for compression.

    Introduction 18

  • 1. I

    1.5 Scope of Investigation

    There seems to be adequate evidence supporting the hypothesis that compression failure of a

    plate with a hole is initiated by fiber kinking at the hole surface [2,4,9, 10,27,37]. This kinking leads

    to fiber breakage causing kink bands and eventually resulting in catastrophic failure. Research

    [4,9, 1 O] shows the significant contribution of shear to this mode of failure. Evidence of this is more

    apparent when considering laminates that have no O degree plies in the loading direction, i.e. all

    plies are angle plies. Even under a uniaxial compressive load combined compressive and shear

    stresses are developed due to the stiffness coupling in the laminae and the presence of geometric

    discontinuities. Analyses to date have had reasonable success in predicting the failure strength of

    specific laminated composites under a compressive load, though a more general approach is

    desired. The objective of the present work is to evaluate two existing models of compression failure

    and attempt to predict the strength of a laminated composite plate with a hole under varying

    laminate orientations, ~·

    As discussed earlier, to date most work on microbuckling and kinking has utilized a two-

    dimensional approach based on the assumption that the response is planar. Based on inference

    ·from limited experimental data, as discussed in Section 1 .4 above, that the variation in normalized

    strength, P/P 0 , was the same for the two laminate configurations tested regardless of their

    thickness, and for the sake of simplicity, this study is based on the analysis of a [±45/90/0].

    laminate. Also for the reasons discussed in Section 1.4 above, in this study a three-dimensional

    linear finite element model of the above laminate is used to analyze the stress distribution around

    a hole. Comparison between the stress distribution in this model and a similar two-dimensional

    model is presented. The stresses from the finite element models, along with the kinking model

    proposed by Gurdal [4], are used to predict failure of a quasi-isotropic laminate under compression.

    The kinking model failure criteria is applied to the stresses at the edge of the hole to evaluate

    failure of the laminate, at varying angles of rotation of the quasi-isotropic laminate. The analysis

    is then compared with experimental data available as described in Section 1.3. As it was suspected

    that transverse stresses may contribute to compressive failure, the kinking model was modified to

    incorporate them. Also in order to fully evaluate the effect of the inter1aminar stresses, the onset

    of delamination was evaluated based on a criteria proposed by Lagace and Saeger [34]. Even

    though it is suspected [2,39] that the failure is progressive in nature, the present study ignores the

    progressive nature of failure and does not use ply discount technique. Also interaction between the

    two modes of failure will not be considered.

    Introduction 19

  • In Chapter 2 the two finite element models, two-dimensional and three-dimensional, are described,

    while Chapter 3 discusses the kinking and delamination models along with the failure criteria that

    have been used. In Chapter 4 the stresses obtained from the two finite element models have been

    discussed while Chapter 5 discusses the result of the failure analysis using criteria and models

    presented in Chapters 2 and 3. Conclusions and recommendations have been covered in Chapter

    6.

    Introduction 20

  • 2 FINITE ELEMENT MODELS

    2.1 Introduction

    In order to compare the effect of different analysis models on the failure prediction of a plate with

    a hole loaded in compression, two finite element models were generated. The first is a two-

    dimensional model which could account for in-plane stresses only. The second model is a three-

    dimensional model capable of providing all six stresses at a point. The primary objective of creating

    the two-dimensional model was to provide a basis by corroborating data obtained by Gurdal [4],

    as to the effect of in-plane compressive and shear forces on the failure strength of the laminated

    plate in compression. Due to discrepancies between the analytical and experimental values

    observed by Gurdal and Haftka [1 O], the natural progression seemed to be the study of the effects

    of out-of-plane stresses on the failure of the plate. It has long been suspected [18] that these

    stresses may play a significant role in the mechanism contributing to failure of the abovementioned

    plate due to the out-of-plane nature of the kink formation. Recently experimental evidence by

    Guynn et al. [27] have substantiated this claim. In addition to the out-of-plane microbuckling/kink

    band formation, researchers [2,27,37] have also detected the existence of delamination during the

    failure process. The prediction of the inception of delamination also requires evaluation of the

    transverse stresses which makes a three-dimensional analysis necessary.

    Many techniques have been put forth for the evaluation of the complex stress state along the edge

    of a hole in a plate. These vary from being based on an elasticity approach [42], to using

    Lekhnitskii complex function based solution [43], and finally the three-dimensional finite element

    approach. Of the above, the three-dimensional finite element approach was chosen in order to

    compare and contrast with similar techniques (two-dimensional) utilized earlier. It will be shown

    later that the use of this model not only provides the out-of-plane stresses but also provides in-

    plane stresses that are quite different those obtained from a similar two-dimensional model.

    Finite Element Models 21

  • y

    c .£ x 0

    ~ .-

    ------- 1 .50 in --------

    z

    0.048 in

    Figure 2-1: Plate Specimen Dimensions and Loading

    Finite Element Models 22

  • Both finite element models were generated using a commercially available software, PATRAN [44)

    (version 2.5), for pre- and post-processing while the finite element analysis was carried out using

    ABAQUS [45) (version 4.8). The dimensions of the plate are as shown in Figure 2-1. The

    distribution of stresses around the hole are described by angle a., with a. = 0 being at x = 0.1 O and y = 0.00. The laminate orientation, as described in Figure 1-2, is denoted by cj>. The properties of the laminae are assumed to be hornogenous and are based on those of a commercially available

    graphite/epoxy (AS1/T3201) provided in Table 2-1. The layup of the plate is [±45/90/0). which

    exhibits quasi-isotropy at the laminate level.

    Table 2-1: Material Properties of AS1/T320f

    E11 E22 E33 \>12 \>13 \>23 G12 G13 G23

    20.0 1.30 1.25 0.30 0.30 0.49 1.03 0.90 0.90

    * All rnodulii in msi

    The above properties are not the exact properties of the specimens described in Chapter 1, but

    rather reflects those of similar commercially available. material.

    2.2 Model Description

    2.2.1 Two-dimensional Model

    The two-dimensional model consists of a 0.048 in thick plate in which the properties of the various

    laminae were smeared to provide the global properties of the laminate. By virtue of the layup it is

    quasi-isotropic in the x-y plane. The analysis of the two-dimensional model is based on Classical

    Lamination Theory (CL T) and a plane stress state is considered. Since no through-the-thickness

    variation of strains are allowed, a single element was used to model the laminate in the z direction. Eight node quadratic elements were used in this model, except at places where due to geometric

    necessity (i.e. at the fine-coarse mesh interface), triangular elements were utilized. These elements

    consist of quadratic lagrangian interpolation functions, so are capable of providing better

    Finite Element Models 23

    I 'i

    '' '

  • r

    Finite Element Models

    -G> "C 0 E ... c G> E G> -G> G> ~ c ii: .. N

    I N

    ~ ::J C> ii:

    24

  • compliance to the plate in the vicinity of the hole, than linear lagrangian function elements. Since

    interlaminar stresses are ignored in Classical Lamination Theory (CL T), considerable freedom can

    be exercised in generating this model due to the numerous planes of symmetry which will be

    elaborated on later. However to facilitate comparison, the nodal distribution of the two models in

    the x-y plane has been kept the same. The nodal points were located so as to limit computation

    time and also prevent distortion of the elements. The maximum aspect ratio (Vt) of the elements

    was 41.667, which is well within the limits for these quadratic elements, as established by

    Vidussoni [46].

    The distribution of the nodes at the point of interest, surrounding the hole, has been kept fine and

    then gradually expanded away from it. The nodes were distributed radially along the hole to a point

    0.15 inches, beyond which the distribution was on a cartesian basis as shown in Figure 2-2. The

    nodal distribution was chosen so as to be sensitive to the radial variation of stresses along the hole

    edge.

    2.2.2 Three-dimensional Model

    The primary objective of this model was to study the magnitude and effect of the three-dimensional

    stress state on the compression failure of a plate with a hole. The modeling procedure is thus

    critical for proper evaluation of stresses. Three-dimensional finite element models are often not

    given consideration as analytical tools as, at the present state of the art of computers, they are not

    cost effective. Many techniques have been presented over the years, with an emphasis on

    improved cost-effectiveness, to provide the three-dimensional information within the area of interest,

    yet utilize a two-dimensional model away from it. A study by Dong [47] provides Qood overview on alternatives to complete three-dimensional modeling that can be utilized. Of these the Local-Global

    technique has been the widely accepted [46,48] procedure. Bums et al. [35] recommends that for

    Local-Global analysis of a notched composite laminate, the local-global interface should be at least

    five times the laminate thickness away from the hole boundary. This interface for our model would

    have to be 0.240 inches from the hole boundary or 0.340 inches from the center of the hole. Given

    the dimensions of our specimen, this technique, would not have provided any substantial reduction

    in size of the finite element model. Since the objective of this study, however, is principally to

    evaluate the three-dimensional effect on compressive strength, the cost effectiveness of this

    technique will not be belabored.

    Finite Element Models 25

  • Limitations on the size of the model are imposed by the capability of the computing system to

    process the data. Most of the limitations experienced were hardware based, primarily inadequate

    swap diskspace, resulting in the use of a barely acceptable model. The number of elements

    chosen, particularly in the region of the hole, was such so as to provide accurate representation

    of the stress distribution, and satisfaction of boundary conditions. The greater the number of nodes

    in the x-y plane the better the representation of the transverse stresses. Similarly the greater the

    number of nodes through the thickness the more accurate satisfaction of the boundary conditions

    at the hole edge. It can be see that as the degree of accuracy of the stresses obtained from the

    finite element model increases the size of the model grows exponentially. Since both of the above

    requirements cannot be satisfied due to hardware limitations the model has to be compromised.

    Pagano [38) has shown that the effect of interlaminar stresses are confined to a distance of

    approximately one lamina thickness away from the free edge, which in our case is the hole

    boundary where the in-plane stresses are the greatest. As the stresses in this region are known

    to be singular, a large number of nodes is needed in this region to properly represent the

    distribution of transverse stresses. Too few nodes would cause the model to be too stiff and so

    over/under estimate the magnitude of these stresses, while too many nodes would rapidly increase

    the size of the model. -·'

    In order to ensure the proper modeling of the stress gradients, and to satisfy the traction-free

    boundary conditions as closely as possible, it is desirable to use two elements per layer through

    the thickness. Hardware limitations permitted only one element per layer to be used to represent

    the thickness of each lamina. Vidussoni [46] has shown that reasonably adequate representation

    can be provided with one 20 node element per layer for cross ply laminates, though, two such

    elements per layer would definitely provide more accurate results. This has been corroborated by

    Thompson [48) though she used more elements. Depending on the local distribution, the difference

    between using one or two elements per layer may not be great due to the fact that, in the event

    of using two elements, the aspect ratio of the element for the same planar dimensions increases

    as compared to the one element case. Thus the gain in sensitivity may not have the desired result,

    as the change in aspect ratio effects the stiffness characteristic of the element [46]. Also, in order

    to maintain adequate sensitivity in the x-y plane at least two elements have been accommodated

    within one laminate thickness from the hole edge.

    In the x-y plane, the distribution of the nodes is the same as that in the two-dimensional model and

    is shown in Figure 2-2. At least two rings of elements are located within a region of one laminate

    Finite Element Models 26

  • thickness from the hole edge. The density of radial distribution of the nodes was made to vary with

    the distance from the hole edge. Away from the hole, where the stresses are assumed to

    gradually reduce to far field values, larger aspect ratios (Vt) have been used. Twenty node brick

    elements have been used all over except at the location where the distribution changes from radial

    to cartesian where, appropriately, 15 node triangular wedge elements have been used.

    2.3 Boundary Conditions

    2.3.1 Symmetry

    The use of boundary conditions to represent planes of symmetry on the three-dimensional model

    is far more critical than that of the two-dimensional model. In a two-dimensional model of a quasi-

    isotropic plate, two planes of symmetry can be exploited. As such a quarter plate model can be

    used for the two-dimensional model. By way of definition a model with a midplane of symmetry

    (symmetric laminate) is called a half symmetric model. The increase in the number of planes of

    symmetry reduces the order of symmetry of the plate by a factor of two, e.g. a two-dimensional

    symmetric laminate plate will be an eighth symmetric model due to assuming symmetry about the

    planes through the three orthogonal axes. In a three-dimensional model the number of elements

    and the total degrees of freedom being large, symmetry provides a necessary means of reducing

    the size of the model. For all symmetric laminates the midplane automatically presents a plane of

    symmetry. Burns [35] has shown that no other plane of true symmetry exists for a three-

    dimensional models of quasi-isotropic laminates.

    Despite the lack of true symmetry, in order to reduce the computational effort other symmetry

    conditions were investigated. One such condition which proved to yield accurate stress distribution

    was introduction of the symmetry about the x-z plane. For the 0 and 90 degree laminae the

    coefficients of mutual influence 1111.i• which are defined as [35]

    Finite Element Models

    (2511-2513-551 ) cos38sin8- (2522-2512 -SH) cos8sin38 S11cos'8+ (2S12+S11 ) sin28cos26+S22sin'8

    (2.01)

    27

  • 'tx z

    0.01.-----.----...-....... ---....-...-....... -..----..-----.-----.

    0

    -0.011

    .. .f~ .. l : ! l i ~ !

    ~

    I ......... -............ 1 ................... .

    ; l

    ··········f···-·····--··+······-·····-···············+·····················-··-···+····-·······················

    ~'. I 11~--i I --& - Ctr. Sym. ~ -0.1 ----------------------

    0 10

    Angle a 1311

    Figure 2-3a: Comparison of a In half and zz

    quarter symmetric models

    110 ~

    0.1 ..-........ -.----...-........ -.--....-..-........ -..---.-..-................ ---.---..

    0.011 . . ···················-----···-·--··-···········-··········· .. ······· ··········-·· ............................... ..

    0

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    411 10

    Angle a 1311

    Figure 2-3b: Comparison of 't In half and xz quarter symmetric models

    110 ~

    Finite Element Models 28

  • where:

    'Y1i - Shear strains

    £1 - Normal strains

    S11 - Compliance coefficients

    9 - Fiber angle are zero about the plane perpendicular to the loading axis. However, this is not the case with angle

    ply laminae. Thus, an in-plane strain will result in shear strains in the off-axis plies. The effect of

    imposing symmetry on the x-z plane causes erratic behavior of the through-the-thickness stresses,

    az and 'txz• in the immediate vicinity of the hole along the x-axis.

    Figures 2-3 show the distribution of these stresses along the periphery of the hole in three-

    dimensional models, with and without assumed symmetry about the x-z plane. The stresses are

    normalized by the average far-field stress, att• which is obtained by dividing the force reactions of the nodes at the supports, i.e. X=-0.75, by the laminate area at that location. The location of the

    support is at a distance such that the stresses there are unaffected by the existence of the hole.

    The stress distribution of both, the half and quarter symmetric models coincide over most of the

    hole boundary except when approaching the plane of assumed symmetry (x-axis) due to the

    imposed boundary constraint, Ui = 0.00. Similar behavior has also been shown by Burns [35], which -~

    he describes to be of localized nature. This imposition of symmetry is therefore not recommended

    where absolute accurate representation of the stress state is necessary throughout the edge of the

    hole. Thus a quarter symmetry model of a quasi-isotropic laminate would be considered adequate

    for a failure analysis of an axially loaded condition as the point of interest is along the y axis, i.e.

    away from the axis of symmetry. The use of a half symmetry model would result in significantly

    more computation time with no particular gain in accuracy.

    2.3.2 Traction-free Boundary Conditions

    In finite element modeling, isoparametric displacement elements satisfy equilibrium by means of

    static geometric equations between the internal forces and the external loads. The finite elements

    are based on a displacement based variational principle and as such continuity of displacements

    is Satisfied. Displacements are thus generated at each point of integration. The strains at these

    points are then generated, on an elemental basis, as derivatives of the displacements. The

    stresses at each strain are then obtained by multiplying the strains by the stiffness coefficient

    matrix. It is thus seen that the stresses are discontinuous at the interfaces due to the difference

    Finite Element Models 29

  • ,.. I

    Finite Element M

    odels

    CD ... cu ... 0 "C CD E ... 0 """" CD "C c :s I -CD "C 0 E -cu c 0 ·-0 c CD E ·-"C I CD CD ... .c I-.. ~ I C\I CD ... :s en ·-LL

    30

  • s1ap

    ow 1

    uawa

    13 a

    11u1

    :1

    "T1 -· cc c ~ CD N I c.n •• -f ::r ~ CD CD I c. -·

    . ~": 3 CD :J tn -· 0 :J m - 3 0 c. CD -I c. a 0 ~ 3 CD c. en .. m .. CD

  • in material properties in the two adjacent layers and due to the lack of continuity of the derivatives

    of the displacements across the element boundaries. This is acceptable for in-plane stresses and

    strains, however, equilibrium conditions require interlaminar stresses to be continuous at the

    interfaces. These have been taken care of by using an averaging interpolation technique. By using

    these displacement elements it is not possible to satisfy the boundary conditions at the traction free

    edges. They will not be identically zero [48]. It is for'this reason that, in order to attain convergence,

    the number of elements used to model the thickness be increased in order to attain traction free

    edges. It has been shown that in spite of system limitations and the small number of elements per

    layer used in the models, satisfactory results can be obtained to provide valuable information

    [35,48].

    2.4 Comparison Between 2-D and 3-D Models

    So far most of the work done [3,4, 1 O], on the ~nalysis of failure of a plate with a hole has been

    restricted to the use of two-dimensional finite element models, assuming planar failure. Due to the

    disparity between experimental and analytical values of strength obtained, it has become necessary

    to evaluate the three-dimensional effect on the response of a laminate to in-plane loading. A two-

    Climensional model was used to correlate the results obtained earlier [1 O] and compare with a

    three-dimensional model that is used to build upon it. Prior to performing a failure analysis it is

    necessary to understand the response of the respective models under the uniaxial load. Figures

    2-4 and 2-5 show the "undeformed" and "deformed" shapes of the three-dimensional finite element

    model. In Figure 2-5 an "exagerration factor" (maximum model dimension/maximum displacement)

    of 1 O was used.

    In order to generalize a failure theory for quasi-isotropic laminates, it is necessary to validate it for

    all laminate orientations,~- To ensure this Gurdal & Haftka [10] recommended that the model of a quasi-isotropic laminate [±45/90/0]. be subject to a load that was rotated through 180 degrees.

    This is analogous to keeping the load constant and varying the fiber angles of the laminae by the

    same amount resulting in a variety of quasi-isotropic laminates, as described in Section 1.3. In this

    case, they showed, that it was necessary to perform only one two-dimensional finite element

    analysis, subsequent orientations of the laminate could be accomplished by transformation of the

    loading axis. This process is efficient and economical and was considered for the three-dimensional

    model as well. However, as will be shown later this procedure is not conducive for use with three-

    dimensional analysis.

    Finite Element Models 32

  • r

    z

    3h/8 ··································f····································1····:············~················l.-.......................... .

    i + ! I ;

    ·································-'····································•····· ·····A·················::'=,,,::····································

    i I ~ ~

    h/4

    h/8 ··································f····································l·················4r··············f··········· i i I i

    ..--~~~-'~ : I

    1~~=1 * 0 .__.__.__.__.__,&,__..__..__..__..__"'-___ ~,..._...__,_..._~ ............ ~

    -0.00045 -0.0004 -0.00035

    Strain £ xx

    -0.0003 -0.00025 lg24.pm

    Figure 2-6: Through the thickness strain distribution for '" 2-D and 3-D models at x=0.50 and y=-0.33

    (symmetric about z-axis)

    Finite Element Models 33

  • O.OOO&r-....-....... --...--.,.--.--...-.--....-....... --...--..--.--...-.--....--.

    Eu -0.0005

    -0.001

    ., I i

    " i I i I i

    " i ••..•. ,. ........... r······························ I i

    _:_j ____ _ --.- 3-0 Model (0)

    - -ll - 2·0Model(O) -0.002 .__ ____ ..__ __ _..._.__....__._ __ ..__...a..._..._.__....__,

    0 46 10 135

    Angle a

    Figure 2-7a: Comparison of strains between 2-D and 3-D models, 9=0 and '=0

    -0.0005 ___ ....... __....__ ............ __..._...._ __ ....... __. ___ ............ __._...._ __ ..... 0 45 10 135 180

    Angle a

    Figure 2· 7b: Comparison of strains between 2-D and 3-D models, 9::90 and t = O

    Finite Element Models 34

  • r

    0.001 .--...--.---.--...-------.-----------.-----.---....-..---e-- 3-D Modi! (..CS) - -II - 2-0 Modi! (..CS)

    0.0005

    . ' : ········-r-···················:a:·········r······· . . . . i ~ i . . . . . . i \ i i ..

    -0.0005

    -0.001 L-.......__;,..__,.....,.&-...................... ..._...L-_.___.__,..._....L.. .............. ,__...._....,1 0 45 10 135 180

    Angle a

    Figure 2-7c: Comparison of strains between 2-D and 3-D models, 9=-45 and '=0

    Finite Element Models 35

  • r

    au

    4r;:::c:::c:::z:==::::::i====~,....----...--."""."""r-----_,..-, _..__ 3-0 Model (0)

    - -b - 3-0 Modi! (45)- •• -45

    : Jrq::~:::=::··~+~··:::~::~::~:-= : ! !

    1 ······························!·················· ············!············ .................. L. .......................... .

    I I : !

    0 ~ .... ~ ...... .e-tS::·-··-1··-··--··-·1 ~...._ ........ _..__.,__.._...._ ........ ......i.. ....... -....__._....a... ........ __ ..._....__~

    0 45 80 135 180

    . Angle ex if29.pll Figure 2·8: Comparison of stresses obtained by

    rotation and transformation of 3·0 model· at • = O degrees

    __._ 3-0 Model (45)- •· 0

    - -b - 3-0 Model to>- t•45 -+- ~D Model (0)- + • Onmd. 45

    . ~~:=~--~~:::~··-~:~~::= : : : I I . i oa... ............... 4=:tP-'. .. :'.'.':--r-.

    :

    i 0 80 135 180

    Angle Cl

    Figure 2·9: Comparison of stresses obtained by rotation and transformation of 3-0 model at • = 45 degrees

    Finite Element Models 36

  • a.,

    0.1------.-. ....... --------------------------------. . : ., . __ ;;,_;;_~;~~--J_L.fL.: ...