You are on page 1of 72

Acknowledgement

We express deep sense of gratitude and profound regards to S Srilakshmi,

Asst. Professor, Dr. V Balakrishna Murthy, Professor, Department of Mechanical

Engineering, Prasad V Potluri Siddhratha Institute of Technology, Vijayawada, for

their keen interest, constant encouragement and invaluable guidance during the

course of the work.

We express our deep sense of gratitude to our beloved Principal, K

Srinivasu, for a kind patronage. We express our sincere thanks to Dr. K Sivaji Babu,

Head of the Department, Mechanical Engineering, for his encouragement

throughout the project.

We express our thanks to our programmer Mr G Rajendra Prasad in CAD lab,

Department of mechanical Engineering for providing the lab during the entire

course of our work.

Finally we thank others who assisted us directly or indirectly for successfully

completing this project.

Project Associates,

Prathyusha Achanta,

Kranthi Chand Musunuru,

Pranoy Yerraguntla,

Ashish Kalipatnapu.
Abstract
One of the most important damage mechanisms in composite

materials is the delamination between plies of the laminate. In industrial

applications, composite plates are sensitive to impact and delamination

occurs. Many composite components have curved shapes, tapered

thickness and plies with different orientations, which make the

delamination grow depending on the extent of the crack. It is therefore

important to analyse the delamination characteristics of composite

structures.

The main objective of the present investigation is the


characterization of the delamination growth in four layered
unidirectional fibre reinforced composite laminates under variations of
crack length, stacking sequence and different load applications. The
analysis has been carried out using Virtual Crack Closure Technique
(VCCT) in combination with Finite Element Methods (FEM) theoretically
and numerically with the help of commercially available Finite Element
Software, ANSYS.
<Summary/Conclusions Briefing>
This work can be useful in analysing the effect of various factors such
as location of the crack and nature of the load on fracture response of
FRP structures, saving a great amount of time for further research in this
field.
CONTENTS

1. INTRODUCTION
1.1 INTRODUCTION
1.2. FIBER REINFORCED POLYMERS
1.3. INTRODUCTION TO FRACTURE
1.3.1 MODES OF FAILURES
1.3.2EXAMPLES OF STRUCTURAL FAILURES CAUSED BY
FRACTURE• MECHANICAL, AERONAUTICAL, OR marine•
CIVIL ENGINEERING
1.4. FRACTURE MECHANICS VS. STRENGTH OF MATERIALS

2. LITERATURE SURVEY
2.1 INTRODUCTION
2.2 TYPOLOGY OF FRP DELAMINATION
2.2.1 INTERLAMINAR CRACKS AND LINEAR ELASTIC
FRACTURE MECHANICS
2.2.2 BASIC ANALYSIS OF INTERLAMINAR FRACTURE
TOUGHNESS
2.3. FRACTURE MODES
2.3.1. MICROSCOPIC ASPECTS
2.4. MAJOR HISTORICAL DEVELOPMENTS IN FRACTURE
MECHANICS
2.5. THE STUDY OF DELAMINATION USING NUMERICAL
METHODS

3 PROBLEM STATEMENT AND METHODOLOGY


3.1. INTRODUCTIONNTRODUCTION
3.2 PROBLEM STATEMENT
3.3 METHODOLOGY
3.4 ASSUMPTIONS

4.ANALYSIS OF EDGE CRACK LAMINATES


4.1 INTRODUCTION
4.2 PROBLEM MODELLING
4.3 CASE ONE: PRESSURE LOADING
ANALYSIS OF RESULTS:
4.3.1 ALL LONGITUDINAL FIBRES
4.3.2 ALL TRANSVERSE FIBRES
4.3.3 SYMMETRIC FIBRE ORIENTATION
4.3.4 ANTISYMMETRIC FIBRE ORIENTATION
4.4 CASE TWO: LINE LOADING
ANALYSIS OF RESULTS:
5.4.1 ALL LONGITUDINAL FIBRES
5.4.2 ALL TRANSVERSE FIBRES
5.4.3 SYMMETRIC FIBRE ORIENTATION
5.4.4 ANTISYMMETRIC FIBRE ORIENTATION

5 ANALYSIS OF CENTRE CRACK LAMINATES


5.1 INTRODUCTION
5.2 PROBLEM MODELING
5.3 CASE ONE: UNIFORM PRESSURE
5.3.1 ALL LONGITUDINAL FIBRES
5.3.2 ALL TRANSVERSE FIBRES
5.3.3 SYMMETRIC FIBRE ORIENTATION
5.3.4 ANTISYMMETRIC FIBRE ORIENTATION
5.4 CASE TWO: LINE LOADING
ANALYSIS OF RESULTS:
5.4.1 ALL LONGITUDINAL FIBRES
5.4.2 ALL TRANSVERSE FIBRES
5.4.3 SYMMETRIC FIBRE ORIENTATION
5.4.4 ANTISYMMETRIC FIBRE ORIENTATION

6 CONCLUSIONS AND SCOPE


ERROR! REFERENCE SOURCE NOT FOUND.
SCOPE

7 REFERENCES
List of Figures

Fig.1.1 Cracked Cantilevered Beam

Fig 1.2 Failure Envelope for a Cracked Cantilevered Beam

Fig.2.1.Replica of a cross-ply laminate with an inner delamination in the

0/90 ply interface

Fig.2.2 Internal delamination: (a) disposition across the laminate and (b)

effect on the overall stability

Fig.2.3.Near-surface delamination: (a) open in tension; (b) closed in tension;

(c) open buckled; (d) closed buckled; (e) edge buckled and (f) edge buckled

with secondary crack. [5]

Fig.2.4.Elastic variation of P versus and change in energy

Fig.2.5.Load-displacement curve for stable crack growth in (a) brittle matrix

and (b) tough matrix

Fig.2.6.Load-displacement curve for unstable crack growth

Fig.2.7.Crack propagation modes: (a) mode I; (b) mode II and (c) mode III

Fig.2.8.Stress field of a resin rich area point and matrix micro crack

formation ahead of the crack tip

Fig.2.9 Crack growth at (a) θ/0 and (b) 0/0 ply interfaces illustrating the

crack plane migration mechanism


Fig.2.10. Fibre bridging in a mode I inter-laminar crack. The vertical arrow

indicates a point 20 mm behind the crack tip

Fig.2.11.Formation and growth of a mode II delamination at the ply

interface: (a)micro crack formation ahead of the crack tip; (b) micro crack

growth and opening and (c) micro crack coalescence accompanied by shear

cusps

Fig.2.12.Development of shear micro cracks in mode II non-reversed and

reversed Delamination

Fig. 4.1 Geometry and Loaded model for edge crack at centre interface.

Fig. 4.2 Geometry of a 20 Node SOLID95 Element

Fig 4.3 Deformed model after cylindrical bending due to uniform pressure

Fig 4.4: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-0-0-0

laminate.

Fig 4.5: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-90-90-

90 laminate.

Fig 4.6: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-90-90-0

laminate.

Fig 4.7: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-0-0-90

laminate.

Fig 4.8: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-90-0-90

laminate.
Fig 4.9: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-0-90-0

laminate.

Fig: 4.10 Loaded model for edge crack opening at centre interface

Fig. 4.11 Deformed model for edge crack opening at centre interface

Fig 4.12 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-

0-0-0 laminate

Fig 4.13 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with

90-90-90-90 laminate

Fig 4.14 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-

90-90-0 laminate.

Fig 4.15 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with

90-0-0-90 laminate.

Fig 4.16 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-

90-0-90 laminate

Fig 4.17 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with

90-0-0-90 laminate.

Fig. 5.1 Geometric model for centre crack at centre interface.

Fig. 5.2 Geometry of a 20 Node SOLID95 Element

Fig. 5.3 Loaded model for centre crack at centre interface


Fig 5.4 Deformed model after laminate bending due to uniform pressure

load

Fig 5.5: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-0-0-0

laminate.

Fig 5.6: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-90-90-

90 laminate.

Fig 5.7: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-90-90-0

laminate.

Fig 5.8: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-0-0-90

laminate.

Fig 5.9: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-90-0-90

laminate

Fig 5.10: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-0-90-

0 laminate.

Fig. 5.11 Loaded model for edge crack at centre interface

Fig. 5.12 Deformed model for centre crack opening at centre interface

Fig 5.13 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-

0-0-0 laminate.

Fig 5.14 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with

90-90-90-90 laminate.
Fig 5.15 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-

90-90-0 laminate.

Fig 5.16 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with

90-0-0-90 laminate.

Fig 5.17 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-

90-0-90 laminate.

Fig 5.18 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with90-

0-90-0 laminate
NOMENCLATURE

E1 Young`s Modulus In longitudinal direction of the lamina

E2 Young`s Modulus in the in-plane transverse direction of the lamina

E3 Young`s Modulus in the out-of-plane transverse direction of the lamina

G12 Shear Modulus in 1-2 longitudinal plane of the lamina

G23 Shear Modulus in 1-3longitudinal plane of the lamina

G13 Shear Modulus in transverse plane of the lamina

Ѵ12 Major in- plane Poisson’s ratio

Ѵ23 Out- of- plane Poisson’s ratio

Ѵ13 Out-of-plane Poisson’s ratio

KI Stress intensity factor

KIc Critical stress intensity factor

VCCT Virtual crack closure technique

SERR Strain energy release rate

GI Strain energy release rate

GIC Critical strain energy release rate


1
INTRODUCTION

1.1 INTRODUCTION

The use of advanced composite materials has increased considerably in

the fabrication of structural elements. Advanced composite materials

progressively substitute traditional materials, such as steel, aluminium or

wood, due to their better specific properties. The excellent stiffness to weight

and strength to weight ratios of polymeric matrix composite materials, particularly

those reinforced with glass or carbon fibres make them very attractive for certain

manufacturing sectors. Initially, this type of materials was exclusively used in

technologically advanced applications, such as in aeronautical and aerospace

industries. Nowadays, due to technological development and reduction of

manufacturing costs, these materials are being used more and more in different

applications. These applications range from sportive items, biomedical

implants, car parts, fluid containers and pipes, small boats and road bridges

to advanced aircraft and space vehicles. It can be seen the important

increment in the use of this type of material by the general industrial sector.

Nevertheless, the especial characteristics of the design of a composite

part have limited a wider generalization of these materials. Designing a new

composite element not only requires the design of the element geometry, but the

design of the material itself. Traditionally, due to the reduced knowledge of the

behaviour of composite materials, this process was accomplished using methods


based on available empirical data. However, this methodology is limited to the

characterization of definite materials and stacking sequences, meanwhile the

number of material combinations is nearly unlimited. The experimental

characterization of the material is expensive and difficult to be extended to other

material configurations. Until a better knowledge about composite materials

behaviour and properties was achieved, the dependence on experimentation

limited, in part, a higher use of composites in more common applications.

1.2. FIBER REINFORCED POLYMERS

The oldest composite materials appeared long time ago in the nature.

Wood can be seen as a lignin matrix reinforced by cellulose fibres. Human and

animal bones can be described as fiber-like osteons embedded in an interstitial

bone matrix. The first man made composite was straw-reinforced clay for bricks

and pottery. Present composite materials use metal, ceramic or polymer binders

reinforced with different fibres or particles. Then, composite materials can be

defined as those materials resulting from the combination of two or more

materials (known as components or constituents), different in composition, form

or function at a macroscopic scale. In the resulting composite material, the

components conserve their initial identity without dissolving or mixing

completely. Usually, the components can be physically distinguished and it is

possible to identify the interface between components. Taking into account their

structural properties, composite materials can be defined as those materials having

a reinforcement component (fibre or particles) in an agglutinating component


(matrix). Reinforcement is responsible for the composite high structural properties;

meanwhile matrix gives physical support and ambient integrity.

With the combination of different matrices (usually polymeric matrices

or light metals) with different fibres (glass, carbon, organic and polymeric

fibres, among others), it is possible to obtain composite materials with

different mechanical properties specially designed for certain applications.

Thus, the great number of combinations results in a great number of composites.

They can be distinguished in function of their typology (long o short fibres, random

or oriented, single or multiple plies, etc.) or in function of their components

(thermoplastic or thermo set polymer matrix, aluminium or titanium metal

matrix, carbon matrix, inorganic or organic fibres, metal whiskers, etc.).

Usually, polymeric matrix fibre reinforced composite materials are found in

the shape of unidirectional laminates (all the reinforcement fibres in one

direction) or as bidirectional laminates (many unidirectional plies with different

fibre orientations). Different types of reinforcement can be used to improve the

properties of the resulting composite, which is then known as hybrid. This is the

case of reinforced concrete, a particle-reinforced composite (concrete) further

reinforced with steel rods. When a light core material is sandwiched between two

faces of stiff and strong materials, the result is an improved material called

sandwich.
1.3. INTRODUCTION TO FRACTURE

In this introductory stage of fracture, we shall start by reviewing the

various modes of structural failure and highlight the importance of fracture induced

failure and contrast it with the limited coverage given to fracture mechanics in

Engineering Education. In the next section we will discuss some examples of well

known failures/accidents attributed to cracking. Then, using a simple example we

shall compare the failure load predicted from linear elastic fracture mechanics with

the one predicted by “classical” strength of materials.

1.3.1 MODES OF FAILURES

The fundamental requirement of any structure is that it should be designed

to resist mechanical failure through any (or a combination of) the following modes:

1. Elastic instability (buckling)

2. Large elastic deformation (jamming)

3. Gross plastic deformation (yielding)

4. Tensile instability (necking)

5. Fracture
Most of these failure modes are relatively well understood, and proper

design procedures have been developed to resist them. However, fractures

occurring after earthquakes constitute the major source of structural damage, and

are the least well understood. In fact, fracture often has been overlooked as a

potential mode of failure at the expense of an overemphasis on strength. Such a

simplification is not new, and finds a very similar analogy in the critical load of a

column. If column strength is based entirely on a strength criterion, an unsafe

design may result as instability (or buckling) is overlooked for slender members.

Thus failure curves for columns show a smooth transition in the failure mode from

columns based on gross section yielding to columns based on instability. By

analogy, a cracked structure can be designed on the sole basis of strength as long as

the crack size does not exceed a critical value. Should the crack size exceed this

critical value, then a fracture-based failure results. Again, on the basis of those two

theories (strength of materials and fracture mechanics), one could draw a failure

curve that exhibits a smooth transition between those two modes.

1.3.2EXAMPLES OF STRUCTURAL FAILURES CAUSED BY FRACTURE

Some well-known, and classical, examples of fracture failures include:

• Mechanical, aeronautical, or marine

1. Fracture of train wheels, axles, and rails

2. Fracture of the Liberty ships during and after World War II


3. Fracture of airplanes, such as the Comet airliners, which exploded in mid-

air during the fifties, or more recently fatigue fracture of bulkhead in a Japan

Air Line Boeing 747

4. Fatigue fractures found in the Grumman buses in New York City, which

resulted in the recall of 637 of them

5. Fracture of the Glomar Java sea boat in 1984

6. Fatigue crack that triggered the sudden loss of the upper cockpit in the Air

Aloha plane in Hawaii in 1988

• Civil engineering

1. Fractures of bridge girders (Silver Bridge in Ohio)

2. Fracture of Stratford A platform concrete off-shore structure

3. Cracks in nuclear reactor piping systems

4. Fractures found in dams (usually unpublicized)

Despite the usually well-known detrimental effects of fractures, in many

cases fractures are man-made and induced for beneficial purposes examples

include:

1. Rock cutting in mining.

2. Hydro-fracturing for oil, gas, and geothermal energy recovery.

3. “Biting” of candies.
1.4. FRACTURE MECHANICS VS. STRENGTH OF MATERIALS

In order to highlight the fundamental differences between strength of

materials and fracture mechanics approaches, we consider a simple problem, a

cantilevered beam of length L, width B, height H, and subjected to a point load P at

its free end is given by

Fig.1.1 Cracked Cantilevered Beam

We will seek to determine its safe load-carrying capacity using the two

approaches.

1. Based on classical strength of materials the maximum flexural stress should

not exceed the yield stress σy, or

2. In applying a different approach, one based on fracture mechanics, the

structure cannot be assumed to be defect free. Rather, an initial crack must

be assumed. Governed failure; for the strength of materials approach in the


linear elastic fracture mechanics approach (as discussed in the next

chapter), failure is governed by:

KI≤ KIc

Where K Iis a measure of the stress singularity at the tip of the crack and KIc is the

critical value of K I. K I is related to σmax through:

The two equations governing the load capacity of the beam according to

Two different approaches, call for the following remarks:

1. Both equations are in terms of BH2 /6L

2. The strength of materials approach equation is a function of a material

property that is not size dependent.

3. The fracture mechanics approach is not only a function of an intrinsic

material property but also of crack size a.

On the basis of the above, we can schematically represent the failure envelope of

this beam in fig.where failure stress is clearly a function of the crack length.

Fig 1.2 Failure Envelope for a Cracked Cantilevered Beam


On the basis of this simple example, we can generalize our preliminary finding by

the curve shown in Fig. We thus identify four corners: on the lower left we have our

usual engineering design zone, where factors of safety are relatively high; on the

bottom right we have failure governed by yielding, or plasticity; on the upper left

failure is governed by linear elastic fracture mechanics; and on the upper right

failure is triggered by a combination of fracture mechanics and plasticity. This last

zone has been called elasto-plastic in metals, and nonlinear fracture in concrete.
2
LITERATURE SURVEY

2.1 INTRODUCTION

In the previous chapter, it has been stated that the main objective of the

present work is the study of inter-laminar fracture behaviour of four layered cross-

ply unidirectional continuous fibre reinforced composite laminates However, before

analysing the behaviour of inter-laminar cracks, it is necessary to address the

principles of the delamination mechanics, the onset and propagation of inter-

laminar cracks under static conditions, the interaction of delamination with

other micromechanics of composite laminates, etc. In this way, a better

understanding of the phenomenon can be achieved.

This chapter includes a sort of review on basic aspects of

composite eliminations. The review starts taking into account a classification of

inter-laminar cracks and continues with the application of fracture mechanics basic

concepts. Next, an overview on microscopic aspects of composite delamination is

considered. The historical approaches to the characterisation of this phenomenon

under static conditions are also presented and briefly discussed


2.2 TYPOLOGY OF FRP DELAMINATION

Crack formation between two adjacent plies, or delamination, is a damage

mechanism of composite laminates that can form during any moment of the life of

the structure: manufacturing, transport, mounting and service. According to [1]

and [2] the technological causes of the delamination can be grouped in two

categories. The first category includes delamination due to curved sections,

such as curved segments, tubular sections, cylinders and spheres, and

pressurised containers. In all these cases, the normal stresses in the interface of

two adjacent plies can originate the loss of adhesion and the initiation of the inter-

laminar crack. The second category includes abrupt changes of section, such as ply

drop-offs, unions between stiffeners and thin plates, free edges, and other

bonded and bolted joints. A third category related to temperature and moisture

effects can be added. The difference between the thermal coefficients of matrix

and reinforcement results in differential contractions between plies during the

curing process. The residual stresses originated by these differential contractions

may originate delamination [3]. Similarly, the differential inflation of the plies

during the absorption of moisture might be the cause of delamination [4].

Delamination can be also originated during the manufacturing stage

due to the shrinkage of the matrix, formation of resin-rich areas due to poor

quality in lying the plies, etc. [5-6]. Impact is an important source of delamination

in composite structures. Inter-laminar cracks can be originated by internal damage

in the interface between adjacent plies as a consequence of an impact in the


laminate, due to the drop of a tool during production, mounting or repairing, or

ballistics impacts in military planes or structures.

Location within the stacking sequence of the laminate has an important

effect on the growth of delamination [7]. According to [5-6], two types of

delamination can be considered: internal delamination and near-surface

delamination. Internal delamination originate in the inner ply interfaces of the

laminate and can be due to the interaction of matrix cracks and ply interfaces.

Delamination originated in the 0/90 interfaces by transversal matrix cracks in

the 90º plies of cross-ply laminates are common examples of this type of

delamination. Figure 2.1 shows a replica with an inner delamination growing from a

transverse crack to the left in a 0/90 interface of a carbon/epoxy cross-ply

laminate subjected to axial load [8]. In the replica, some fibre breaks in the 0º ply

can be seen due to the stress concentration near the transverse crack.

Fig.2.1.Replica of a cross-ply laminate with an inner delamination in the


0/90 ply interface [8]
Inner delaminations considerably reduce the load-capacity of composite

elements. In particular, when compression loads are applied, the overall flexural

behaviour of the laminate is significantly affected (as shown schematically in Figure

2.2). Although the delamination separates the laminate in two parts, there is an

interaction between the deformation of the one part of the laminate and the

other. Due to this interaction, both parts of the laminate deflect in a similar way.

Fig.2.2 Internal delamination: (a) disposition across the laminate and (b)

effect on the overall stability

Near-surface delaminations, as its name indicates, originate near the surface

of the laminate and represent a more complex scenario than internal

delaminations. The deformation of the delaminated part is less influenced by the

deformation of the rest of the laminate. Therefore, the deformation of the near-

surface delaminated part does not necessary follow the deformation of the rest
of the laminate. Consequently, not only the growth of the near-surface

delamination has to be taken into account but also its local stability. [5-6] classified

the different types of near-surface delaminations than can originate in plate

composite components in different load conditions. Figure 2.3 shows different

types of near-surface delaminations.

Fig.2.3.Near-surface delaminations: (a) open in tension; (b) closed in


tension; (c) open buckled; (d) closed buckled; (e) edge buckled and (f)
edge buckled with secondary crack. [5]

After initiation, both types of delaminations can propagate either under

static loads either under fatigue conditions. In both cases, the reduction in strength

and stability of the composite part to flexural loading is considerable.


2.2.1 INTERLAMINAR CRACKS AND LINEAR ELASTIC FRACTURE MECHANICS

Fracture mechanics is concerned with crack-dominated failures and

delamination is a fracture mechanism of composite laminates. Therefore, fracture

mechanics is a suitable methodology to approach the onset and propagation

of composite delaminations problem. In addition, usual composite laminates

are very stiff in the laminate plane and behave as linear elastic materials in

their gross deformation. Thus, it is reasonable to base the analysis of interlaminar

toughness on linear elastic-fracture mechanics (LEFM).

2.2.2 BASIC ANALYSIS OF INTERLAMINAR FRACTURE TOUGHNESS

Nowadays, composite materials are tailored in order to profit their

high in-plane tensile strength. However, the through-thickness properties of such

materials are in most cases very low compared to the in-plane tensile strength.

Therefore, the through-thickness stresses in laminated composite may initiate

delamination, especially if particular geometries (free edges, holes, ply drops

etc.) or previous damage (matrix cracks or micro-delaminations as a consequence

of impacts, fabrication problems, etc.) are present in the material [9]. After

delamination onset, the consequent propagation is not controlled by the through-

thickness strength any more but by the interlaminar fracture toughness.

If the interlaminar toughness is expressed in terms of energy

release rate, the delamination will propagate when the energy release rate

achieves a critical value, Gc. According to [10], for any form of elastic behaviour,
the energy release rate can be expressed as a function of the increment of

external work Ue, strain energy Us (kinetic energy is ignored in this case) and

crack increment ∆a. Therefore, for a crack of width b and length a, the

energy release rate can be expressed as

Figure 2.4 presents an elastic variation of the load P versus the

displacement for an interlaminar crack growing from an initial length a to a final

length a+∆a. In point A1 the applied load is P1, the displacement is 1 and

delamination length a. In point A2 the applied load and displacement are P2 and 2,

respectively, and the crack length is a+∆a.

Fig.2.4.Elastic variation of P versus and change in energy

Therefore, the external work and strain energy of the linear variation

shown in Figure 2.4 can be expressed, respectively, as


The change in energy is determined by the area OA1A2 (dashed area in the

figure). If linear deformation behaviour is assumed, the straight lines showed in the

figure are to be used and the change in energy becomes

For the considered crack increment and width, the increment in crack area

would be b∆a. Thus, as the critical energy release rate can be defined as the change

in energy per unit of new crack surface and denoting P1 as P, P2 as P+∆P, δ1 as δ

and δ2 as δ+∆δ, the expression for Gc can be written as

The compliance of the system depends on the crack length and is defined

as

Taking into account the increments of load and displacement and equation,

the increment in displacement can be expressed as

and combining equations a final expression for the critical energy

release rate can be found after mathematical manipulation as:


or in differential form as

For the experimental study of interlaminar crack propagation in composite

materials, the variation of the applied load with respect the obtained displacement,

as shown in Figure 2.4, is basic. This experimental data, together with the crack

length, is the basis for the calculation of G and the generation of the R-

curve. However, the experimental determination of the onset and propagation

values of G for an interlaminar crack is complicated and different methods

can be used. The first method is based in the determination of Gc by visual

observation of the crack onset.

Nevertheless, this method is imprecise and highly dependent on the

observer. The second method is based on the calculation of Gc at the point

of non-linearity of the load-displacement curve. According to [9], for brittle

matrix composites the non-linearity point coincides with the point at which

the initiation of the crack can be observed (see Figure 2.5(a)). However, for tough

matrices a region of non-linear behaviour may precede the observation of the crack

initiation (see Figure 2.5(b)). In the third method, Gc is determined as the

intersection between the load-displacement curve and the line that corresponds to

an increase by the 5 % to the original compliance of the system. If the maximum

load occurs before intersection, then the maximum load and corresponding

displacement are used to compute Gc.


Fig.2.5.Load-displacement curve for stable crack growth in (a) brittle

matrix and (b) tough matrix

The load-displacement curves represented in Figure 2.5 are for stable

crack growth cases. Unstable crack growth is characterised by one or more periods

without crack propagation (or very slow) followed by rapid propagations,

which results in sharp drops in the load-displacement curve. These rapid

propagations are normally followed by arrest and a reloading, which results in

a local peak load when delamination growth restarts. This behaviour is usually

known as stick-slip growth and results in typical saw-teeth load-displacement

curves [11]. Figure 2.6 shows a typical load-displacement curve for the case of

unstable interlaminar crack growth


Fig.2.6.Load-displacement curve for unstable crack growth

2.3. FRACTURE MODES

According to fracture mechanics, the growth or propagation of an

interlaminar crack, or delamination, may occur in mode I (opening), mode II

(shearing), mode III (tearing) and in any combination of these (see Figure 2.7). Every

mode has a fracture toughness value and an R-curve associated which are intrinsic

material characteristics. In the case of isotropic materials, only mode I toughness is

considered. For these materials, the fracture toughness is lowest in this mode

and even if the crack starts to grow under a different mode, the crack will

deviate and grow in mode I [9].

Fig.2.7.Crack propagation modes: (a) mode I; (b) mode II and (c) mode III
The propagation of delaminations in laminated composite materials is

mainly limited to lie between the strong fibre reinforced layers. In this way, it is

possible for a delamination to propagate in any combination of the three

propagation modes. A clear example is the case of transverse matrix cracks growing

in the 90º plies of cross-ply laminates loaded in tension (see Figure 2.1). Once the

crack reaches the strong fibres at the (0/90) interface, the crack is forced to deviate

and change direction in order to remain in the interface. Then, the propagation

mode is changed. In fact, composite delaminations are mostly studied under pure

mode I, pure mode II and mixed-mode I/II. It is generally accepted that the mode III

contribution in delamination growth is negligible. In fact, the mode III

contribution is typically quite small for composite structures as a consequence

of the constraints of adjacent plies, as shown by [12] for a layered structure and

by [13] in laminated lap-joints. In addition, the fracture toughness values for

delamination in composite laminates are higher in mode III than in the other

modes [9]. In the foregoing the term mixed-mode will stand for the mixed-

mode I/II condition.

In isotropic materials, toughness values are commonly expressed in

terms of the critical stress intensity factor. However, interlaminar fracture

toughness of laminated composites is normally expressed in terms of the

critical energy release rate. The stress intensity factor is governed by the local

crack-tip field and is extremely sensitive. It is difficult to obtain true values of

K at the crack tip due to the inhomogeneous composition of composite

laminates complicates. [2] state that the use of G for composite materials is
certainly more consistent with the analytical models in use than K, even

though the K-mix can be defined rigorously, in contrast to G. Therefore, the

majority of the studies about delaminations in composites use the critical

energy release rate, Gc, instead of the critical stress intensity factor, Kc, to

predict the initiation of the crack.

2.3.1. MICROSCOPIC ASPECTS

At the microscopic level, the growth of an interlaminar crack is

preceded by the formation of a damage zone ahead of the crack tip. This damage

zone is characterised by the formation of microcracks in the resin rich areas that

exist between the plies. According to [14], at this microscopic level, the matrix can

be seen as an isotropic and homogeneous material, which in general, like metals,

will only crack under tensile load conditions (local mode I). Therefore, matrix

microcracks will form and grow in the plane subjected to maximum tensile

stress. Figure 2.8 shows schematically a point of a resin rich area in the ply

interface subjected to mode I (opening) and mode II (shearing) loading and the

formation of a matrix microcrack in the plane subjected to the resulting maximum

tensile stress, σm.


Fig.2.8.Stress field of a resin rich area point and matrix microcrack

formation ahead of the crack tip

Under mixed-mode load condition, microcracks ahead of the crack tip form

at an angle from the plane of the plies and grow in this direction. According to [15],

when such a microcrack is located at 0/0 ply interface, where 0 stands for an off-

axis ply, fibres on the off-axis ply allow the propagation of the microcrack through

the ply. Consequently, the crack tip of the delamination migrates through the off-

axis ply. A change in the crack plane can be achieved if the crack tip

encounters the next ply interface. In this case, the study and characterisation

of the delamination become complicated. The crack plane migration mechanism is

represented in Figure 2.9(a). Oppositely, when the microcrack is located at 0/0

ply interface, the fibres at both 0º plies prevent the propagation of the crack

through the plies. The interlaminar crack is forced to remain adjacent to the fibres

of the ply. The mechanism is represented in Figure 2.9(b). In this case, no change in

the crack plane is present and the study and characterisation of the delamination

become easier. Actually, in order to avoid the crack plane migration, the study of
delaminations in composite laminates is usually carried out using unidirectional

laminates with the fibres parallel to the crack growth.

Fig.2.9 Crack growth at (a) θ/0 and (b) 0/0 ply interfaces illustrating the
crack plane migration mechanism [15]

As mentioned, since further growing would require fibre fracture, at 0/0 ply

interfaces the growth of the microcracks is arrested when they reach the

fibres of one of the boundaries of the interlaminar zone. In the general case of

mixed-mode loading, the propagation of the interlaminar cracks results from

the coalescence of these matrix microcracks [16]. For a greater contribution of

mode I, the matrix microcracks grow relatively parallel to the plane of the

plies but progressively displacing to one of the boundaries of the interply

zone. Consequently, the interlaminar crack progressively grows to one of the

interply boundaries, where the presence of the fibres modifies the damage zone

ahead of the crack tip and increases the stress concentration. This results in the

growth of the delamination by the peeling of the matrix from the fibres. According

to [14], this process justifies the presence of fibres in one of the fracture surfaces

while on the other only the fibre imprints are present.


However, the general scenario in delamination test specimens is different.

The presence of fibres bridging both fracture surfaces near the crack tip is

commonly observed. This phenomenon is known as fibre bridging and tends to

arrest or reduce the propagation of the delamination. In fact, the growth of the

crack involves pulling these bridging fibres from the resin under a tensile stress

state until they finally break. Accordingly, an artificial increment of the material

fracture toughness that depends on the crack extension is observed. For longer

crack lengths, more fibres from both fracture surfaces are bridging the crack. It has

been experimentally found that this effect is more important for higher mode I

contributions and less important for higher mode II dominated fractures [15],[17].

In this case, fibre breakage, broken pullout fibres, behind the crack tip can

be observed. According to [18], fibre bridging is a characteristic

micromechanism of unidirectional ply testing that will not occur in real

structures. Figure 2.10 shows an interlaminar crack with the presence of fibre

bridging and fibre breakage in a glass-fibre/vinyl ester composite [19].

Fig.2.10. Fibre bridging in a mode I interlaminar crack. The vertical arrow indicates a
point 20 mm behind the crack tip [19]
For a greater contribution of mode II the size of the damage zone

increases and matrix microcracks start to form at a relatively considerable

distance ahead of the crack tip. In addition, the angle between the direction

of the microcracks and the plane of the plies increases up to 45º. The coalescence

of the microcracks results in the growth of the interlaminar crack but with uneven

surfaces. These uneven surfaces are due to the formation of shear cusps or hackles.

For a greater contribution of mode II, more shear cusps form and deeper they

are. In addition, less influence of fibre bridging is observed [20]. The increased

area of the uneven fracture surfaces at microscopic level in mode II justifies

the increases of the measured fracture toughness for this mode, since more

atomic bonds have to be broken [14]. Figure 2.11 shows schematically the

formation and coalescence of mode II microcracks that result in the formation and

growth of a mode II delamination accompanied by some shears cusps. If these

surfaces are subsequently subjected to a fatigue process in mode II, the shear

cusps will degrade into matrix rollers due to the effect of the friction between

shear cusps of both surfaces.


Fig.2.11.Formation and growth of a mode II delamination at the ply
interface: (a)microcrack formation ahead of the crack tip; (b) microcrack
growth and opening and (c) microcrack coalescence accompanied by shear
cusps

The previous figure shows the formation of the damage zone ahead of

the crack tip for mode II delamination for a non-reversed loading condition. If a

reversed loading condition is considered, a second block of microcracks appear in

the normal direction to the previous, this is at -45º. Therefore, two sets of

microcracks form at approximately 90º [21]. Figure 2.12 shows the schema of the

microcrack formation ahead of the crack tip for mode II reversed and non-

reversed loading conditions.


Fig.2.12.Development of shear microcracks in mode II non-reversed and
reversed Delaminations

2.4. MAJOR HISTORICAL DEVELOPMENTS IN FRACTURE MECHANICS

As with any engineering discipline approached for the first time, it is helpful

to put fracture mechanics into perspective by first listing its major developments:

1. In 1898, a German Engineer by the name of Kirsch showed that a stress

concentration factor of 3 was found to exist around a circular hole in an infinite

plate subjected to uniform tensile stresses [22].

2. While investigating the unexpected failure of naval ships in 1913, [23] extended

the solution for stresses around a circular hole in an infinite plate to the more

general case of an ellipse. It should be noted that this problem was solved 3 years

earlier by Kolosoff (who was the mentor of Muschelisvili) in St Petersbourg,

however history remembers only Inglis who showed that a stress concentration

factor of

Prevails around the ellipse (where a is the half length of the major axis, and ρ is the

radius of curvature) 5.
3. Inglis’s early work was followed by the classical studies of Griffith, who was not

originally interested in the strength of cracked structures (fracture mechanics was

not yet a discipline), but rather in the tensile strength of crystalline solids and its

relation to the theory based on their lattice properties, which is approximately

equal to E/10 where E is the Young’s Modulus *24+. With his assistant Lockspeiser,

Griffith was then working at the Royal Aircraft Establishment (RAE) at Farnborough,

England (which had a tradition of tolerance for original and eccentric young

researchers), and was testing the strength of glass rods of different diameters at

different temperatures [25]. They found that the strength increased rapidly as the

size decreased. Asymptotic values of 1,600 and 25 Ksi were found for infinitesimally

small and bulk size specimens, respectively. On the basis of those two observations,

Griffith’s first major contribution to fracture mechanics was to suggest that internal

minute flaws acted as stress raisers in solids, thus strongly affecting their tensile

strengths. Thus, in reviewing Inglis’s early work, Griffith determined that the

presence of minute elliptical flaws were responsible in dramatically reducing the

glass strength from the theoretical value to the actually measured value.

4. The second major contribution made by Griffith was in deriving a thermo

dynamical criterion

for fracture by considering the total change in energy taking place during cracking.

During crack extension, potential energy (both external work and internal strain

energy) is released and “transferred” to form surface energy. Unfortunately, one

night Lockspeiser forgot to turn off the gas torch used for glass melting, resulting in

a fire. Following an investigation, (RAE) decided that Griffith should stop wasting his

time, and he was transferred to the engine department.


5. After Griffith’s work, the subject of fracture mechanics was relatively dormant for

about 20 years until 1939 when Westergaard [26] derived an expression for the

stress field near a sharp crack tip.

6. Up to this point, fracture mechanics was still a relatively obscure and esoteric

science. However, more than any other single factor, the large number of sudden

and catastrophic fractures that occurred in ships during and following World War II

gave the impetus for the development of fracture mechanics. Of approximately

5,000 welded ships constructed during the war, over 1,000 suffered structural

damage, with 150 of these being seriously damaged, and 10 fractured into two

parts. After the war, George Irwin, who was at the U.S. Naval Research Laboratory,

made use of Griffith’s idea, and thus set the foundations of fracture mechanics. He

made three major contributions: (a) He (and independently Orowan) extended the

Griffith’s original theory to metals by accounting for yielding at the crack tip. This

resulted in what is sometimes called the modified Griffith’s theory. (b) He altered

Westergaard’s general solution by introducing the concept of the stress intensity

factor (SIF). (c) He introduced the concept of energy release rate G

7. Subcritical crack growth was subsequently studied. This form of crack

propagation is driven by either applying repeated loading (fatigue) to a crack, or

surround it by a corrosive environment. In either case the original crack length, and

loading condition, taken separately, are below their critical value. Paris in 1961

proposed the first empirical equation relating the range of the stress intensity

factor to the rate of crack growth.

8. Non-linear considerations were further addressed by Wells, who around 1963

utilized the crack opening displacement (COD) as the parameter to characterize the
strength of a crack in an elasto-plastic solid, and by Rice, who introduce his J

integral in 1968 in probably the second most referenced paper in the field (after

Griffith); it introduced a path independent contour line integral that is the rate of

change of the potential energy for an elastic non-linear solid during a unit crack

extension.

9. Another major contribution was made by Erdogan and Sih in the mid ’60s when

they introduced the first model for mixed-mode crack propagation.

10. Other major advances have been made subsequently in a number of

subdisciplines of fracture mechanics: (i) dynamic crack growth; (ii) fracture of

laminates and composites; (iii) numerical techniques; (iv) design philosophies; and

others.

11. In 1976, [27] introduced the fictitious crack model in which residual tensile

stresses can be transmitted across a portion of the crack. Thus a new meaning was

given to cracks in cementitious materials.

12. In 1979 [28] showed that for the objective analysis of cracked concrete

structure, fracture mechanics concepts must be used, and that classical strength of

materials models would yield results that are mesh sensitive.


2.5. THE STUDY OF DELAMINATION USING NUMERICAL METHODS

Nowadays, many practical problems in delamination of composite

materials are solved using finite element methods (FEM) and other numerical

methods. Numerical methods were applied to the study of interlaminar cracks

shortly after the first studies about delamination in composites. [29] used a

numerical elasticity solution based on finite differences for the analysis of free

edge delaminations. [30] studied the same problem using a 3-D finite

element analysis combined with Maxwell stress functions and minimisation of

complementary energy. [31] developed a 2-D finite element model in which

the distribution of the axial displacements was taken into account. Until this work,

FEM solutions were only available under plane stress or strain conditions. [32]

were the first to obtain sufficiently accurate results of the free edge problem using

a 2-D finite element model. [33] were the first to report the calculation of the

stress singularities in the free edge of the laminate. This work, based on

Lekhnitskii’s stress potentials, was an important contribution and the basis for

other analytical models in the field of composite delaminations. This work was

followed by [34] with a solution to the free edge problem for a ±45 angle ply

laminate.

One of the numerical approaches more commonly used nowadays is the

virtual crack closure technique (VCCT). This technique is based on Irwin’s crack

closure integral [35-37] and assumes that the energy ∆E released when the crack

is extended by an increment ∆a, from a to a + ∆a, coincides with the energy

required to close the crack to its original condition, from a + ∆a to a. At present,


this technique is considered one of the most rigorous techniques for the analysis of

the propagation of interlaminar cracks [3], [38-40].

A current approach to model composite delaminations is based in

homogenisation methods [41-42]. These methods, also known as double scale

methods, are based on finite element methods and solve a submodel for

every integration point of the model. It is in the submodel where the mechanical

behaviour and damage characteristics of the material are defined.


3
PROBLEM STATEMENT AND METHODOLOGY

3.1. INTRODUCTION

In the previous chapter, the relevant literature available has been reviewed

and scope for the present work has been identified. In this chapter, statement of

the problem of present work and method used for solution of the problem has

been explained.

3.2 PROBLEM STATEMENT

The objective of the present work is to analyse the inter-laminar fracture

behaviour of four layered cross-ply unidirectional continuous fibre reinforced

composite laminates with delamination at two different locations (edge crack and

centre crack) using virtual crack closure technique (VCCT) in combination with finite

element method.

3.3 METHODOLOGY

Three dimensional finite element models are generated in ANSYS software

to represent four layered symmetric cross-ply laminate. The inter-laminar crack is

modelled as a longitudinal discontinuity with different nodes attached to the top

and bottom crack surfaces. The nodes at both discontinuity (crack) sides have the
same coordinates and are coupled through multi-point constraints. This multi-

point constraint type ties all the degrees of freedom at the tied node to the

corresponding degrees of freedom at the retained node.

The virtual crack closure technique, assumes that the energy ∆E

released when the crack is extended by an increment ∆a, from a to a + ∆a,

coincides with the energy required to close the crack to its original condition, from

a + ∆a to a. The energy release rate is determined as the work done by the

reaction forces developed due to the multi-point constraints at the crack tip per

unit area of virtual cracked surface for the considered loading condition.

The details of the finite element modelling, loading, boundary conditions,

material properties and crack length are explained in forth coming chapters.

3.4 ASSUMPTIONS

The assumptions in the present work are as follows

1. Possibility of branch cracks is ignored


2. State of plane strain is assumed.
3. Each layer of the laminate is a unidirectional continuous fibre reinforced
lamina
4. Each lamina behaves as a transversely isotropic material
5. The behaviour of the structure is geometrically linear
4

ANALYSIS OF EDGE CRACK LAMINATES

4.1 INTRODUCTION
This chapter presents the analysis of edge crack laminates supported along two

opposite edges and subjected to two different load cases. In first case, a uniform pressure

is applied on the top surface of the laminate. And in second case, a uniform line load is

applied along the top edge of the laminate parallel to the supported end.

4.2 PROBLEM MODELLING


In this section a four layered laminate structure is modelled with the longitudinal

dimension of the plate taken as 100 mm span (L) with a span/depth ratio of 10. Four layers

of equal thickness (10/4=2.5mm) are considered. The width of the plate is considered to be

infinite. The crack location is at one end of the laminate. The length of the crack is varied

from 15mm to 85mm. By trial and error method, comparison of theoretical and analytical

Energy release rate for isotropic material, the crack extension is found to be 0.11mm. . The

same value is extended to orthotropic material.

Fig. 4.1 Geometry and Loaded model for edge crack at centre interface.
Finite element mesh is generated using 20 node quadratic solid element SOLID95 in

ANSYS software. This element is defined by 20 nodes having three degrees of freedom per

node: translations in the nodal x, y and z directions. The element may have any spatial

orientation. SOLID95 has plasticity, creep, stress stiffening, large deflection, and large strain

capabilities. It has the capability to inherit orthotropic material properties and hence, best

suited for analysing FRP composites.

Fig. 4.2 Geometry of a 20 Node SOLID95 Element

The surfaces between the laminates, apart from those at the crack, are bonded

together using a contact pair. The coincident nodes at the surface of the crack extension

length are coupled. Simply supported boundary condition is applied on the bottom edges

of the x-z plane, i.e.; degrees of freedom along the z-axis are constrained. The plate is also

constrained in the y-direction to imply infinite length, and plane strain condition.

Unidirectional fibre reinforced layers of Carbon-Epoxy with the following properties

are used in the laminate. [45]

E1=147GPa; E2=E3=10.3GPa

v12= v13=0.27; v23=0.54

G12=G13=7GPa; G23=3.7GPa
4.3 CASE ONE: PRESSURE LOADING
The transverse uniform pressure of 1 MPa is applied on the surface along the top

edge of the laminate, parallel to the supported end.

Fig 4.3 Deformed model after cylindrical bending due to uniform pressure

The position of the crack is varied between different laminate interfaces, along

with the increase in the length of the crack. The nature of deflection in the Z-axis direction

is studied and the Strain Energy Release Rate in the plate is calculated from the element

tables.

ANALYSIS OF RESULTS:
With respect to Fig 4.4 to 4-9, the total amount of Energy is first used to deform

the crack region transversely and then for the extension of the crack longitudinally. Initially,

the energy required to deform is less. But, as the crack length increases, the energy

required to deform the crack region increases. Hence, more amount of energy has to be

supplied to shear the crack. After a certain crack length, the material resistance of the

laminate decreases. As a result of this the amount energy required to shear the crack

decreases.
4.3.1 ALL LONGITUDINAL FIBRES

120
Edge Crack 0-0-0-0
Top Interface
100 Bottom Interface
Centre Interface
80
G (J/m2)

60

40

20

0
0 20 40 60 80 100
a (mm)

Fig 4.4: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-0-0-0 laminate.

4.3.2 ALL TRANSVERSE FIBRES

1600
Edge Crack 90-90-90-90

1400 Top Interface


Bottom Interface
1200 Centre Interface

1000
G (J/m2)

800

600

400

200

0
0 20 40 60 80 100
a (mm)

Fig 4.5: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-90-90-90 laminate.

The amount of energy required shear the crack is very much less when all the fibres are
oriented longitudinally than transversely, as the fibre and crack extension are in the same
direction.
4.3.3 SYMMETRIC FIBRE ORIENTATION

700
Edge Crack 0-90-90-0
Top Interface
600
Bottom Interface
Centre Interface
500
G (J/m2)

400

300

200

100

0
0 20 40 60 80 100
a (mm)

Fig 4.6: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-90-90-0 laminate.

450
Edge Crack 90-0-0-90

400 Top Interface


Bottom Interface
350 Centre Interface
300
G (J/m2)

250

200

150

100

50

0
0 20 40 60 80 100
a (mm)

Fig 4.7: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-0-0-90 laminate.

In symmetric fibres there is a very slight difference in the Strain energy release rate at the
top and bottom interfaces. The symmetric nature of the Energy distribution is disrupted by
the proximity of the boundary conditions applied on the bottom layer of the laminate.
4.3.4 ANTISYMMETRIC FIBRE ORIENTATION

Edge Crack 0-90-0-90


600

Top Interface
500 Bottom Interface
Centre Interface
400
G (J/m2)

300

200

100

0
0 20 40 60 80 100
a (mm)

Fig 4.8: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-90-0-90 laminate.

600
Edge Crack 90-0-90-0
Top Interface
500 Bottom Interface
Centre Interface
400
G (J/m2)

300

200

100

0
0 20 40
a (mm) 60 80 100

Fig 4.9: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-0-90-0 laminate.

The structural resistance when the top layer is oriented 00 is greater than fibres oriented
with 900 at the top layer.
4.4 CASE TWO: LINE LOADING
A line load of 10 N is applied on the line along the top edge of the laminate, parallel
to the supported end.

Fig: 4.10 Loaded model for edge crack opening at centre interface

Fig. 4.11 Deformed model for edge crack opening at centre interface

Analysis of Results:

Fig 4-12 to 4-17 show that the opening up of a crack requires more amount of energy, as

the energy required for deformation of the crack dominates more than the energy required

for crack propagation.


4.4.1 ALL LONGITUDINAL FIBRES

Edge Crack Opening 0-0-0-0


0.8

0.7

0.6

0.5
G (J/m2)

0.4

0.3

0.2

0.1

0
0 10 20 30 40 50 60 70 80 90
a (mm)

Fig 4.12 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-0-0-0 laminate

4.4.2 ALL TRANSVERSE FIBRES

Edge Crack Opening 90-90-90-90


9
8
7
6
G (J/m2)

5
4
3
2
1
0
0 10 20 30 40 50 60 70 80 90
a (mm)

Fig 4.13 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 90-90-90-90
laminate
5.4.3 SYMMETRIC FIBRE ORIENTATION

Edge Crack Opening 0-90-90-0


3

2.5

2
G (J/m2)

1.5

0.5

0
0 10 20 30 40 50 60 70 80 90
a (mm)

Fig 4.14 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-90-90-0
laminate.

Edge Crack Opening 90-0-0-90


3

2.5

2
G (J/m2)

1.5

0.5

0
0 10 20 30 40 50 60 70 80 90
a (mm)

Fig 4.15 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 90-0-0-90
laminate.
4.4.4 ANTISYMMETRIC FIBRE ORIENTATION

Edge Crack Opening 0-90-0-90


3.5

2.5

2
G (J/m2)

1.5

0.5

0
0 10 20 30 40 50 60 70 80 90
a (mm)

Fig 4.16 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-90-0-90
laminate

Edge Crack Opening 90-0-90-0


3

2.5

2
G (J/m2)

1.5

0.5

0
0 10 20 30 40 50 60 70 80 90
a (mm)

Fig 4.17 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 90-0-0-90
laminate.
5

ANALYSIS OF CENTRE CRACK LAMINATES

5.1 INTRODUCTION
This chapter presents the analysis of centre crack laminates supported along two

opposite edges and subjected to two different load cases. In first case, a uniform pressure

is applied on the top surface of the laminate. And in second case, a uniform line load is

applied along the top edge of the laminate parallel to the supported end.

5.2 PROBLEM MODELING

In this model the cross-section of the rectangular dimension plate are taken as 100

mm span (L) with a span/depth ratio of 10. Four layers of equal thickness (10/4=2.5mm)

are considered. The width of the plate is considered to be infinite. The crack location is at

centre of the laminate. The length of the crack is varied from 15mm to 75mm. The crack

extension length is taken to be 0.11mm on both sides of the crack.

Fig. 5.1 Geometric model for centre crack at centre interface.


Finite element mesh is generated using 20 node quadratic solid element SOLID95 in

ANSYS software. This element is defined by 20 nodes having three degrees of freedom per

node; translations in the nodal x, y and z directions. The element may have any spatial

orientation. SOLID95 has plasticity, creep, stress stiffening, large deflection, and large strain

capabilities. It has the capability to inherit orthotropic material properties and hence, best

suited for analysing FRP composites.

Fig. 5.2 Geometry of a 20 Node SOLID95 Element

The surfaces between the laminates, apart from those at the crack, are bonded

together using a contact pair. The coincident nodes at the surface of the crack extension

length are coupled. Simply supported boundary condition is applied on the bottom edges

of the x-z plane, i.e.; degrees of freedom along the z-axis are constrained. The plate is also

constrained in the y-direction to imply infinite length, and plane strain condition.

Unidirectional fibre reinforced layers of Carbon-Epoxy with the following properties

are used in the laminate. [45]

E1=147GPa; E2=E3=10.3GPa

v12= v13=0.27; v23=0.54

G12=G13=7GPa; G23=3.7GPa
5.3 CASE ONE: UNIFORM PRESSURE
The transverse uniform pressure of 1 MPa is applied on the surface along the top

edge of the laminate, parallel to the supported end.

Fig. 5.3 Loaded model for centre crack at centre interface

The position of the crack is varied between different laminate interfaces, along

with the increase in the length of the crack. The nature of deflection is studied and the

Strain Energy Release Rate in the plate is calculated from the element tables.

Fig 5.4 Deformed model after laminate bending due to uniform pressure load
5.3.1 ALL LONGITUDINAL FIBRES

45
Centre Crack 0-0-0-0
40 Top Interface
Bottom Interface
35
Centre Interface
30
G (J/m2)

25
20
15
10
5
0
0 10 20 30 40 50 60 70 80
a (mm)

Fig 5.5: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-0-0-0 laminate.

5.3.2 ALL TRANSVERSE FIBRES

400
Centre Crack 90-90-90-90
Top Interface
350
Bottom Interface
300 Centre Interface

250
G (J/m2)

200

150

100

50

0
0 10 20 30 40 50 60 70 80
a (mm)

Fig 5.6: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-90-90-90 laminate.

Due to the symmetric nature of crack when at centre interface, the amount of resistance to
deformation is more. When the crack is at the top or bottom interface, the required energy
to propagate the crack is lesser. Due to the constraints on the bottom layer, forces develop
in opposite direction, hence propelling easier crack growth.
5.3.3 SYMMETRIC FIBRE ORIENTATION

180
Centre Crack 0-90-90-0
160 Top Interface
Bottom Interface
140
Centre Interface
120
G (J/m2)

100

80

60

40

20

0
0 10 20 30 40 50 60 70 80
a (mm)

Fig 5.7: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-90-90-0 laminate.

160
Centre Crack 90-0-0-90
Top Interface
140
Bottom Interface
120
Centre Interface
100
G (J/m2)

80

60

40

20

0
0 10 20 30 40 50 60 70 80
-20
a (mm)

Fig 5.8: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-0-0-90 laminate.

The symmetric nature of the orientation, the energy required at bottom and top surfaces is
almost the same amount. Minor variations in the value of G are only a result of constraints
applied. And also, it is easier to open up a crack at the interface of 0-0 fibres than 90-90.
5.3.4 ANTISYMMETRIC FIBRE ORIENTATION

180
Centre Crack 0-90-0-90
160 Top Interface
Bottom Interface
140
Centre Interface
120
G (J/m2)

100

80

60

40

20

0
0 10 20 30 40 50 60 70 80
a (mm)
Fig 5.9: Variation of ‘G’ with respect to ‘a’ from an edge crack with 0-90-0-90 laminate

Centre Crack 90-0-90-0


180
160 Top Interface
Bottom Interface
140
Centre Interface
120
100
G (J/m2)

80
60
40
20
0
0 10 20 30 40 50 60 70 80
-20
a (mm)

Fig 5.10: Variation of ‘G’ with respect to ‘a’ from an edge crack with 90-0-90-0 laminate.

The resistance when the top layer is oriented 00 is greater than when oriented at 900.
5.4 CASE TWO: LINE LOADING
A line load of 1000N is applied on the line along the centre of the laminate, parallel
to the supported end.

Fig. 5.11 Loaded model for edge crack at centre interface

Fig. 5.12 Deformed model for centre crack opening at centre interface

Analysis of Results:
The graphs 5.13-5.18 show the constant trend of the amount of energy required to open

up the crack to be a continuous increasing curve with respect to the crack length.
5.4.1 All Longitudinal Fibres

Centre Crack Opening 0-0-0-0


350

300

250
G (J/m2)

200

150

100

50

0
0 10 20 30 40 50 60 70 80
a (mm)

Fig 5.13 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-0-0-0 laminate.

5.4.2 ALL TRANSVERSE FIBRES

Centre Crack Opening 90-90-90-90


3000

2500

2000
G (J/m2)

1500

1000

500

0
0 10 20 30 40 50 60 70 80
a (mm)

Fig 5.14 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 90-90-90-90
laminate.
5.4.3 SYMMETRIC FIBRE ORIENTATION

Centre Crack Opening 0-90-90-0


1200

1000

800
G (J/m2)

600

400

200

0
0 10 20 30 40 50 60 70 80
a (mm)

Fig 5.15 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-90-90-0
laminate.

Centre Crack Opening 90-0-0-90


1200

1000

800
G (J/m2)

600

400

200

0
0 10 20 30 40 50 60 70 80
a (mm)

Fig 5.16 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 90-0-0-90
laminate.
5.4.4 ANTISYMMETRIC FIBRE ORIENTATION

Centre Crack Opening 0-90-0-90


900
800
700
600
G (J/m2)

500
400
300
200
100
0
0 10 20 30 40 50 60 70 80
a (mm)

Fig 5.17 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 0-90-0-90
laminate.

Centre Crack Opening 90-0-90-0


Fig 5.18 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with 90-0-0-90
1400
laminate.
1200

1000
G (J/m2)

800

600

400

200

0
0 10 20 30 40 50 60 70 80
a (mm)

Fig 5.18 Variation of ‘G’ with respect to ‘a’ of opening a centre crack with90-0-90-0
laminate
6

CONCLUSIONS AND SCOPE

INTRODUCTION

The energy required for the growth of a crack under shear and opening conditions is

analysed for a four layered cross ply unidirectional fibre reinforced composite laminate

with delamination at two different locations (edge crack and centre crack) using

commercially available Finite Element Software, ANSYS.

Trend Pattern

Edge Crack

The energy required to shear the crack in the laminate under shear follows an increasing

trend until a certain crack length and then decreases as the length of the crack increases.

When in opening mode, the energy required to open up the crack almost follows a linear

increasing trend until a definitive crack length and then shows a parabolic growth.

Centre Crack

The energy required to shear and also open up the crack are in a progressively in a

parabolic increasing pattern


CONCLUSIONS

 It is much safer to have centre crack structures than those with an edge crack.

 The presence of a crack at the centre interface is less harmful compared to

structures with a crack closer to the bottom or top layers.

 Transverse fibres are safer than longitudinal fibres in the case of presence of a

crack in a laminate structure.

SCOPE
The above analysis can be extend at various levels

 The energy release during delamination in angle ply laminates

 To study the effect of number of layers

 To study the effect of stacking sequence

 The effect of other boundary conditions can be studied

 The effect of thickness ratio of the laminate on the strain energy

 To separate individual energy modes


7

REFERENCES

[1]. Kedward, K.T., 1995. Mechanical Design Handbook. (ed.) Rothbart, H.A., Harold, A.

McGraw-Hill, New York (USA), pp. 15.01-15.29.

[2]. Pagano, N.J., Schoeppner, G.A., 2000. Delamination of polymer matrix composites:

problems and assessment. Comprensive Composite Materials, 2. (ed.) Kelly, A.,

Zweben, C. Elsevier Science Ltd., Oxford (UK).

[3]. Tay, T.E., Shen, F., 2002. Analysis of delamination growth in laminated composites with

consideration for residual thermal stress effects. Journal of Composite Materials 36

(11), pp. 1299-1320.

[4]. Crasto, A.S., Kim, R.Y., 1997. Hygrothermal influence on the free-edge delamination of

composites under compressive loading. Composite Materials: Fatigue and Fracture 6,

ASTM STP 1285. (ed.) Armanios, E.A. American Society for Testing and Materials,

Philadelphia (USA), pp. 381-393.

[5]. Bolotin, V.V., 1996. Delaminations in composite structures: Its origin, buckling,

growth and stability. Composites Part B-Engineering 27 (2), pp. 129-145.

[6]. Bolotin, V.V., 2001. Mechanics of delaminations in laminate composite structures.

Mechanics of Composite Materials 37 (5-6), pp. 367-380

[7]. Greenhalgh, E., Singh, S., 1999. Investigation of the failure mechanisms for

delamination growth from embedded defects. Proceedings of the 12th

International Conference on Composite Materials, Paris (France).

[8]. Gamstedt, E.K., Sjogren, B.A., 2002. An experimental investigation of the


sequence effect in block amplitude loading of cross-ply composite laminates.

International Journal of Fatigue 24 (2-4), pp. 437-446.

[9]. Robinson, P., Hodgkinson, J.M., 2000. Interlaminar fracture toughness. Mechanical

Testing of Advanced Fibre Composites. (ed.) Hodgkinson, J.M. Woodhead Publishing,

Cambridge (UK), pp. 170-210.

[10]. Hashemi, S., Kinloch, A.J., Williams, J.G., 1990a. The analysis of interlaminar

fracture in uniaxial fiber-polymer composites. Proceedings of the Royal Society of

London Series A-Mathematical Physical and Engineering Sciences 427 (1872), pp.

173-199.

[11]. Kusaka, T., Hojo, M., Mai, Y.W., et al., 1998. Rate dependence of mode I

fracture behaviour in carbon-fibre/epoxy composite laminates. Composites Science

and Technology 58 (3-4), pp. 591-602.

[12]. Jensen, H.M., Sheinman, I., 2001. Straight-sided, buckling-driven delamination of

thin films at high stress levels. International Journal of Fracture 110 (4), pp. 371-385.

[13]. Glaessgen, E.H., Raju, I.S., Poe, C.C., 2002. Analytical and experimental studies

of the debonding of stitched and unstitched composite joints. Journal of

Composite Materials 36 (23), pp. 2599-2622.

[14]. Singh, S., Greenhalgh, E., 1998. Micromechanics of interlaminar fracture in

carbon fibre reinforced plastics at multidirectional ply interfaces under static and

cyclic loading. Plastics Rubber and Composites Processing and Applications 27 (5), pp.

220-226.

[15]. Greenhalgh, E.S., 1998. Characterisation of mixed-mode delamination growth

in carbon-fibre composites. PhD Thesis. Imperial College of Science, Technology

and Medicine, London (UK).

[16]. Purslow, D., 1986. Matrix fractography of fiber-reinforced epoxy composites.

Composites 17 (4), pp. 289-303.


[17]. Tanaka, H., Tanaka, K., 1995. Mixed-mode growth of interlaminar cracks in

carbon/epoxy laminates under cyclic loading. Proceedings of the 10th

International Conference on Composite Materials, Whistler B.C. (Canada), pp. 181-189.

[18]. Olsson, R., Thesken, J.C., Brandt, F., Jonsson, N., Nilsson, S., 1996. Investigations of

delamination criticality and the transferability of growth criteria. Composite

Structures 36 (3-4), pp. 221-247.

[19]. Compston, P., Jar, P.Y.B., 1999. The influence of fibre volume fraction on the mode

I interlaminar fracture toughness of a glass-fibre/vinyl ester composite. Applied

Composite Materials 6 (6), pp. 353-368.

[20]. Tanaka, K., Tanaka, H., 1997. Stress-ratio effect on mode II propagation of

interlaminar fatigue cracks in graphite/epoxy composites. Composite Materials:

Fatigue and Fracture 6, ASTM STP 1285. (ed.) Armanios, E.A. American Society for

Testing and Materials, Philadelphia (USA), pp. 126-142.

[21]. Dahlen, C., Springer, G.S., 1994. Delamination growth in composites under

cyclic loads. Journal of Composite Materials 28 (8), pp. 732-781.

[22]. Timoshenko, S. and Goodier, J.: 1970, Theory of Elasticity, McGraw Hill.

[23]. Inglis, C.: 1913, Stresses in a plate due to the presence of cracks and sharp corners,

Trans. Inst. Naval Architects 55, 219–241.

[24]. Kelly, A.: 1974, Strong Solids, second edn, Oxford University Press.

[25]. Gordon, J.: 1988, The Science of Structures and Materials, Scientific American

Library.

[26]. Westergaard, H.: 1939a, Bearing pressures and cracks, J. Appl. Mech.

[27]. Hillerborg, A. and Mod´eer, M. and Petersson, P.E.: 1976, Analysis of crack

formation and crack growth in concrete by means of fracture mechanics and finite

elements, Cement and Concrete Research 6(6), 773–782.

[28]. Baˇzant, Z. and Cedolin, L.: 1991, Stability of Structures, Oxford University Press.
[29]. Pipes, R.B., Pagano, N.J., 1970. Interlaminar stresses in composite laminates under

uniform axial extension. Journal of Composite Materials 4 , pp. 538-548.

[30]. Rybicki, E.F., 1971. Approximate 3-dimensional solutions for symmetric

laminates under inplane loading. Journal of Composite Materials 5 (JUL), pp. 354-360.

[31]. Herakovich, C.T., Renieri, G.D., Brinson, H.F., 1976. Finite element analysis of

mechanical and thermal edge effects in composite laminates. Army Symposium

on Solid Mechanics, Composite Materials: The Influence of Mechanics of Failure

on Design, Cape Cod (USA), pp. 237-248.

[32]. Wang, A.S.D., Crossman, F.W., 1977. Some new results on edge effect in symmetric

composite laminates. Journal of Composite Materials 11 (JAN), pp. 92-106.

[33]. Wang, S.S., Choi, I., 1982. Boundary-layer effects in composite laminates: I Free-

edge stress singularities. Journal of Applied Mechanics-Transactions of the Asme

49 (3), pp. 541-548.

[34]. Wang, S.S., Choi, I., 1983. The interface crack between dissimilar anisotropic

composite-materials. Journal of Applied Mechanics-Transactions of the Asme 50 (1),

pp. 169-178.

[35]. Irwin, G.R., 1958. Fracture I. Handbuch der Physik, VI. (ed.) Flüge, S.

Springer-Verlag, New York (USA), pp. 558-590.

[36]. Rybicki, E.F., Kanninen, M.F., 1977. Finite-element calculation of stress

intensity factors by a modified crack closure integral. Engineering Fracture

Mechanics 9 (4), pp. 931-938.

[37]. Broek, D., 1986. Elementary engineering fracture mechanics. Noordhoff

International Publishing, Alphen aan den Rijn (The Nederlands).

[38]. Camanho, P.P., Dávila, C.G., 2002. Mixed-mode decohesion finite elements for
the simulation of delamination in composite materials. NASA-Technical Paper

211737, National Aeronautics and Space Agency, USA.

[39]. Kim, I.G., Kong, C.D., Uda, N., 2002. Generalized theoretical analysis method

for free-edge delaminations in composite laminates. Journal of Materials Science 37

(9), pp. 1875-1880.

[40]. Krueger, R., 2002. The virtual crack closure technique: history, approach and

applications. NASA/CR-2002-211628, National Aeronautics and Space Agency, USA.

[41]. Caiazzo, A.A., Costanzo, F., 2001. Modeling the constitutive behavior of

layered composites with evolving cracks. International Journal of Solids and

Structures 38 (20), pp. 3469-3485.

[42]. Rand, O., 2001. A multilevel analysis of solid laminated composite beams.

International Journal of Solids and Structures 38 (22-23), pp. 4017-4043.

[43] Villaverde, N.B., 2004. Variable mix mode delamination in composite laminate

under fatigue condition: Testing and Analysis

[44] Saouma, Victor E., 2000. Fracture Mechanics

[45] Daniel, Isaac M., Ishai, O. 2006. Engineering Mechanics of Composite Materials,
Oxford University Press.

[46] Prashant Kumar, 1999. Elements of Fracture Mechanics, Wheeler Publications

You might also like