Você está na página 1de 93

Development of VB Code to Generate Randomly

Distributed Short Fiber Composites and


Estimation of Mechanical Properties using FEM
Thesis submitted in partial fulfillment of the
Requirements for the degree of

Master of Science
in

Mechanical Systems Design


By

(Signature)

BIJU BL
(Reg. No.102520085)
Under the guidance of

(Signature)
Dr. BADARI NARAYANA KANTHETI
AEROSTRUCTURES
UTC AEROSPACE SYSTEMS
BANGALORE

MANIPAL UNIVERSITY, MANIPAL

This document contains no EAR or ITAR technical data

Development of VB Code to Generate


Randomly Distributed Short Fiber
Composites and Estimation of Mechanical
Properties using FEM
Thesis submitted in partial fulfillment of the
Requirements for the degree of

Master of Science
in

Mechanical Systems Design


By

BIJU BL
(Reg. No.102520085)

Examiner 1

Examiner 2

Signature:

Signature:

Name:

Name:

This document contains no EAR or ITAR technical data

UTC Aerospace Systems


Netra Tech Park, EPIP Industrial Area
Sy.No.28 , Bengaluru 560 066, INDIA
City, State/Province, Postal Code
www.utcaerospacesystems.com

CERTIFICATE
This is to certify that this thesis work titled

Development of VB Code to Generate Randomly


Distributed Short Fiber Composites and
Estimation of Mechanical Properties using FEM
Is a bonafide record of the work done by
BIJU BL
102520085
In partial fulfillment of the requirements for the award of the degree of Master of
Science in Mechanical Systems Design under Manipal University, Manipal and the
same has not been submitted elsewhere for the award for any other degree

(Signature)
Dr. BADARI NARAYANA KANTHETI
AEROSTRUCTURES
UTC AEROSPACE SYSTEMS
BANGALORE

This document contains no EAR or ITAR technical data

ACKNOWLEDGMENTS
Foremost, I would like to express my sincere gratitude to my advisor Dr. Badari
Narayana Kantheti for the support and the immense knowledge. His support was
instrumental from the choosing of the thesis topic through getting the report
completed. I would also like to thank my manager Mr. Ananda Kumar for allowing
me to be off work at times for completing this thesis. I would also like to thank Mr
Pradip Kumar Pandey, SBU Head-Aerostructures and Ravishankar Mysore, Vice
President-Engineering for their approval of this thesis.
A special thanks to my family, especially to my parents and my wife Kamya
for supporting me always. Without her support I could not have gathered so much
time at home to spend on my studies and on this project. Also credits go to my
daughter Samiha for cheering me up whenever I was fatigued.

i
This document contains no EAR or ITAR technical data

ABSTRACT
Short fiber reinforced polymers were developed largely to fill the property gap
between continuous fiber laminates used as primary structures by the aircraft and
aerospace industry and non-reinforced polymers used largely in non-load bearing
applications. In some respects the short fiber systems couple advantages from each of
these property bounding engineering materials. If the fibers are sufficiently long,
stiffness levels approaching those for continuous fiber systems at the same fiber
loading can be achieved, while the ability of the non-reinforced polymer to be molded
into complex shapes is at least partially retained in the short fiber systems. Thus, short
fiber reinforced polymers have found their way into lightly loaded secondary
structures, in which stiffness dominates the design, but in which there must also be a
notable increase in strength over the non-reinforced polymer.
The physical properties may be determined by conducting suitable experiments as per
industry standards. However, a specific set of experiments can only inform us about a
specific Fiber/matrix system. Hence, to design a composite system by tuning its
volume fraction, or fiber/matrix combination, or orientation, then a very large number
of experiments may have to be conducted. Such a process for material property
determination is extremely tedious, prohibitively expensive, and time consuming. Still
further, exact fiber/matrix combinations may not be always available for testing.
Hence, there is a need for developing mathematical models, which can reliably predict
different mechanical properties of composite materials. Such approaches are very
useful for engineers since they provide significant savings in time and cost.
Existing solutions for determining physical properties of aligned short fiber
composites are studied and methods are identified to extend it to randomly oriented
short fiber composite. These were compared with experimental results available in
Literature.
A close correlation was observed between the properties obtained by various
empirical methods and by FE modelling of random fiber. The data obtained should be
a good starting point first degree accuracy which can be used for preliminary analysis
of components manufactured using short fiber composite with random orientation.

ii
This document contains no EAR or ITAR technical data

LIST OF TABLES
Table
No
2.7.6.1
2.7.7.1
2.7.7.2

Table Title
Values for KR used in Eq 2.7.6-8 for shear lag models
Correspondence between Halpin-Tsai Eq 2.7.7-1 and
generalized self-consistent predictions
Traditional Halpin-Tsai parameters for short-fiber
composites

iii
This document contains no EAR or ITAR technical data

Page
No
34
37
39

LIST OF FIGURES
Figure
No

Figure Title

Page
No

1.1 1

Family of Composites (Courtesy http://nptel.ac.in/)

1.1 2

Typical

aerospace

application

brackets

1
(Courtesy 2

http://www.gtweed.com/)
2.2 1

Loads on Composite, Fibres, & Matrix in a Unidirectional 6


Lamina

2.3 1

A Slab Like Model for Predicting Transverse Properties of 10


Unidirectional Composites

2.6 1

Force

Equilibrium

of

an

Infinitesimal

Portion

of 15

Discontinuous Fiber which is aligned to External Load


2.6 2

is a plot of variation of fiber strength for three different fiber 18


lengths.

2.7 1

Eshelby's inclusion problem.

20

2.7 2

Eshelby's equivalent inclusion problem.

21

2.7 3

Idealized fiber and matrix geometry used in shear lag models.

32

2.7 4

Fiber packing arrangements used to find R in shear lag 35


models. (a) Hexagonal (Cox, 1952). (b) Hexagonal (Rosen,
1964) (c) Square (Robinson & Robinson, 1994).

3.1 1

Comparison of Empirical models with experimental results

54

3.5 1

Flow chart for Random fiber generation VBA code

60

3.5 2

VBA code for Random number Generation I

61

3.5 3

VBA code for Random number Generation II

62

3.5 4

A Sample random short fiber composite specimen generated 63


by the VBA code

3.5 5

A Sample random short fiber composite specimen generated 64


by the VBA code

3.5 6

A Sample random short fiber composite specimen generated 64


by the VBA code

3.6 1

Patran Session file for creating random fibers of volume 65


fraction 0.2
iv
This document contains no EAR or ITAR technical data

3.6 2

(a) Geometric model (b) FE model of random fibers of 65


volume fraction 0.2

3.7 1

Patran Session file for creating random fibers of volume 66


fraction 0.15

3.7 2

(a) Geometric model (b) FE model of random fibers of 66


volume fraction 0.15

3.8 1

Patran Session file for creating random fibers of volume 67


fraction 0.1

3.8 2

(a) Geometric model (b) FE model of random fibers of 67


volume fraction 0.1

3.9 1

Load application in the FE model

68

3.10 1

Stress on model with same material applied for fiber and 69


matrix

3.10 2

Stress on cross sections with same material applied for fiber 69


and matrix

4.1 1

Comparison of Empirical model results for E with FE of 72


random fibers

4.2 1

Comparison of Empirical model results for G with FE of 73


random fibers

4.3 1

Comparison of Empirical model results for Strength for 74


random fibers

v
This document contains no EAR or ITAR technical data

List of Notations
A

Area

cross sectional dimension of fiber


A

Strain concentration tensor

Compliance

Fiber diameter

Elastic modulus
Eshelbys Tensor

Shear modulus

length of fiber

lc

Critical Fiber length

Load

Radius of fiber

Stiffness

Thickness

Volume fraction

Strain

efficiency factor

Poissons ratio

Normal Stress

Shear stress

vi
This document contains no EAR or ITAR technical data

Contents
Page No
i
ii
iii
iv

Acknowledgement
Abstract
List Of Tables
List Of Figures
Chapter 1

INTRODUCTION

Introduction

Motivation

Organization of Report

Chapter 2

LITERATURE REVIEW

The Need for Predictive Models for Determining Composite


Properties

Predicting Longitudinal Modulus of Unidirectional Lamina

Predicting Transverse Modulus of Unidirectional Lamina

Shear Modulus and Poissons Ratio

12

Transverse Strength

12

About Short-Fibre Composites

13

vii
This document contains no EAR or ITAR technical data

Modulus of Short-Fiber Composites

viii
This document contains no EAR or ITAR technical data

19

In Eq 3.7.7-2 the underlined term is typically negligible, and


dropping it gives the familiar rule of mixtures for E11of a
continuous-fiber composite. However, dropping the underlined
term in Eq 3.7.7-3 and using a rule of mixtures for 12 is not
necessarily accurate if the fiber and matrix Poisson ratios differ.
Halpin and Tsai argue for this latter approximation on the
grounds that laminate stiffnesses are insensitive to 12. In
adapting their approach to short-fiber composites, Halpin and
Tsai noted that must lie between 0 and . If =0 then Eq
3.7.7-1 reduces to the inverse rule of mixtures ,
1  
= +
  

while for = the Halpin-Tsai form becomes the rule of


mixtures,
 =   +  

Halpin and Tsai suggested that was correlated with the


geometry of the reinforcement and, when calculating E11, it
should vary from some small value to infinity as a function of
the fiber aspect ratio l/d. By comparing model predictions with
available 2-D finite element results, they found that =2(l/d)
gave good predictions for E11of short-fiber systems. Also, they
suggested that other engineering constants of short-fiber
composites were only weakly dependent on fiber aspect ratio,
and could be approximated using the continuous-fiber formulae.
The resulting equations are summarized in Table 3.7.7-2. The
early references and do not mention G23. When this property is
needed the usual approach is to use the value given in Table
3.7.7-1. While the Halpin-Tsai equations have been widely used
for isotropic fiber materials, the underlying results of Hermans
and Hill apply to transversely isotropic fibers, so the Halpin-Tsai
equations can also be used in this case. The Halpin-Tsai
equations are known to fit some data very well at low volume
ix some stiffnesses at high volume
fractions, but to under-predict
This document contains no EAR or ITAR technical data

fractions. This has prompted some modifications to their model.


proposed making a function of vf, and by curve fitting found

Strength of Short fiber composites

Chapter 3

METHODOLOGY

47

48

Stiffness Estimation

49

Strength Estimation

55

Calculation of SFC stiffness for Low aspect ratio PEEK Carbon


Fibre composite

57

Generating Random oriented short fibre composite stiffness


from FE

59

VBA Code for random fiber generation

60

FE Creation and analysis for Fibre volume fraction of 0.2

65

FE Creation and analysis for Fibre volume fraction of 0.15

66

FE Creation and analysis for Fibre volume fraction of 0.1

67

Loads and Boundary Conditions

68

Validation of Stress continuity in FE model

69

Calculation of Elastic Constants from FE Results

70

Chapter 4

RESULT ANALYSIS
x
This document contains no EAR or ITAR technical data

72

Modulus of Elasticity

72

Modulus of Rigidity

73

Strength of Composite

74

Chapter 5

CONCLUSION AND FUTURE SCOPE

75

Work Conclusion

75

Future Scope of Work

75

REFERENCES
ANNEXURES (OPTIONAL)
PROJECT DETAILS

76

xi
This document contains no EAR or ITAR technical data

CHAPTER 1
2. INTRODUCTION
2.1.Introduction

High strength to weight ratio is always a premium for an aircraft. The lower the
structural weight the higher payload it can carry with lesser fuel consumption. This is
the scenario in which Light weight alloys and composites becomes crucial for
aerospace industry.
Composites are the most important materials to be adapted for aviation since the use
of aluminium in the 1920s. Composites are multi-phase materials that are
combinations of two or more organic or inorganic components. One material with
continuous phase serves as a "matrix," which is the material that holds everything
together, while the other material with dispersed phase serves as reinforcement, in the
form of fibres embedded in the matrix. Until recently, the most common matrix
materials were "thermosetting" materials such as epoxy, bismaleimide, or polyimide.
The reinforcing materials can be glass fibre, boron fibre, carbon fibre, or other more
exotic mixtures. Classification of composites is shown in Figure 2.1-1.

Figure 2.1-1 Family of Composites (Courtesy http://nptel.ac.in/)

Even though modern aircrafts like Boeing 787, Airbus A380 and A350 feature large
composite structures, a gap still exists for metal-replacement of smaller lightly loaded
secondary structures with complex-shaped parts such as structural brackets, fittings or
frames/intercostals , where injection moulding has insufficient performance but use of
traditional continuous fiber composite materials is typically impractical due to
complex component geometry. Figure 2.1-2 shows some typical aerospace brackets
1
This document contains no EAR or ITAR technical data

which are traditionally made of metallic materials and which can be replaced by
chopped fiber composites.

Figure 2.1-2 Typical aerospace application brackets (Courtesy http://www.gtweed.com/)

Technologies to produce complex shaped near-net moulded components for a number


of commercial aerospace applications using chopped fiber composites is under
development. It is thus important to structurally validate the components as
experimental data for physical properties and strength for components made with such
composites are not widely available. For this accurate prediction of physical
properties like strength and stiffness is very important. It is here that the objective of
this project is trying to bridge the gap.

2
This document contains no EAR or ITAR technical data

2.2.Motivation

Proven experimental data is not available for physical and structural properties of
chopped fiber composites. Empirical, mathematical & numerical models exist but the
application of the same in aerospace industry is yet to be explored. This project deals
with the use of such models to predict the physical properties of chopped fiber
composites which can be used in FE simulations to structurally evaluate components
manufactured from such composites.

Identify close form solution for strength and stiffness for chopped fiber

composite with aligned fibers

Identify close form solution for strength and stiffness for non-aligned fibers

and to understand what more is needed for random orientation.

Generate a VBA code for forming a unit volume of composite with randomly

oriented short fibers

Validate the mechanical properties using FEM for single fiber or certain fiber

combinations with random orientations

3
This document contains no EAR or ITAR technical data

2.3. Organization of the report

This report is organised into 5 Chapters. Chapter 1-Introduction throws light


on the current situation in the industry on short fibre composites and its importance.
Followed by Chapter 2-Literature Survey in which a research on the existing
empirical methods for predicting mechanical properties is pursued. Chapter 3Methodolgy details the calculations and process followed in the current thesis work
followed by Chapter 4-Results which presents and analyses the results from the
study. This report is closed by providing the conclusion and the future scope of work
in Chapter 5- Conclusion and Future Scope of Work.

4
This document contains no EAR or ITAR technical data

CHAPTER 2
3. LITERATURE SURVEY

This Chapter discusses the existing theories pertaining to the physical property
prediction of aligned short fibre composite and compare it with experimental results
available in literature for different combination of fiber and matrix.

3.1.The Need for Predictive Models for Determining Composite Properties

Mechanical properties of a composite material depend on:


Properties of constituent materials
Orientations of each layer
Volume fractions of each constituent
Thickness of each layer
Nature of bonding between adjacent layers

These properties may be determined by conducting suitable experiments as per


industry standards. However, a specific set of experiments can only inform us about a
specific Fibre/matrix system. Hence, to design a composite system by tuning its
volume fraction, or fiber/matrix combination, or orientation, then a very large number
of experiments may have to be conducted. Such a process for material property
determination is extremely tedious, prohibitively expensive, and time consuming. Still
further, exact fibre/matrix combinations may not be always available for testing.
Hence, there is a need for developing mathematical models, which can reliably predict
different thermo-mechanical properties of composite materials. Such approaches are
very useful for engineers since they provide significant savings in time and cost.

5
This document contains no EAR or ITAR technical data

3.2.Predicting Longitudinal Modulus of Unidirectional Lamina

Consider a unidirectional composite lamina with fibres which are continuous and
uniform in geometric and mechanical properties, and mutually parallel throughout the
length of the lamina. It is also assumed that the bonding between fibre and matrix is
perfect, and thus strains experienced by fibre (f), matrix (f) and composite (c) are
same in longitudinal direction (1-direction). For such a composite, when loaded in 1direction, the total external load Pc will be shared partly by fibres, Pf, and partly by
matrix, Pm. This is shown in Figure 3.2-1

Figure 3.2-1 Loads on Composite, Fibres, & Matrix in a Unidirectional Lamina

It is further assumed that fibres and matrix behave elastically. Thus, the expression for
stress in fibres, and matrix can be written in terms of their moduli (Ef, and Em) and
strains as:
f = Ef f

Eq 3.2-1

m = Em m

Eq 3.2-2

and

Further, if total cross-sectional areas of fibres and matrix are Af and Am, respectively, then:
Pf = Aff = AfEf f,

Eq 3.2-3

Pm = Amm= AmEm

Eq 3.2-4

and

Further, we know that load on composite, Pc, is sum of Pf and Pm. Thus,
6
This document contains no EAR or ITAR technical data

Pc = Acc = Aff + Amm,

Eq 3.2-5

c = (Af/Ac)f + (Am/Ac)m

Eq 3.2-6

or

However, for a unidirectional composite, Af/Ac and Am/Ac are volume fractions for
fibre and matrix, respectively. Hence,
c = Vff + Vmm = (VfEf + VmEm)

Eq 3.2-7

And if Eq 3.2-7 is differentiated with respect to strain (which is same in fibre and
matrix) then,

dc /d = Vf(df /d) + Vm(dm/d), or


Ec = VfEf + VmEm

Eq 3.2-8

Equations Eq 3.2-7 and Eq 3.2-8 show that contributions of fibres and matrix to
average composite tensile modulus and stress are proportionately dependent on their
respective volume fractions. In general, matrix material has a nonlinear stress-strain
response curve. For unidirectional composites having such nonlinear matrix materials
Eq 3.2-7 works well in terms of predicting their stress-strain. However, the stressstrain response curve in such materials may not show up as strongly nonlinear, since
fibres, especially when their volume fractions are high, dominate their stress-strain
response. The higher the fibre volume fraction, the closer is the stress-strain curve for
a unidirectional lamina to that for the fibre.

Experimental data pertaining to tensile test specimens of lamina agree very well with
Eq 3.2-7 and Eq 3.2-8. However, the results for compressive tests are not all that
agreeable. This is because fibres under compression tend to buckle, and this tendency
is resisted by matrix material. This is analogous to a structure with several columns on
an elastic foundation. For a unidirectional composite, the compressive response is
strongly dependent on shear stiffness of matrix material.

Further, Eq 3.2-7 shows us that load shared by fibres may be increased either by
increasing fibre stiffness or by increasing its volume fraction. However, experimental
7
This document contains no EAR or ITAR technical data

data show that it becomes impractical to aim for fibre volume fractions in excess of
80% due to issues of poor fibre wetting and insufficient matrix impregnation between
fibres. To predict longitudinal strength of a unidirectional ply requires one to
understand the nature of deformation of such a ply as load increases. In general,
stress-the stress strain response of unidirectional plies under tension undergoes four
stages of change.
In first stage, when stresses are small, fibre as well as matrix materials exhibit
elastic behaviour.
Subsequently, matrix starts becoming plastic, while most of the fibres continue to
extend elastically.
In the third stage, both fibres and matrix deform plastically. This may not happen in
case of glass or graphite fibres, as they are brittle in nature.
Finally, the fibres fracture leading to sudden rise in matrix stress, which in turn
leads to overall composite failure.
A unidirectional lamina starts to fail in tension, when its fibres are stretched to their
ultimate fracture strain. Here it is assumed that all of its fibre fails at the same strain
level. If at this stage, the volume fraction of matrix is below a certain threshold, then
it will not be able to absorb extra stresses transferred to it due to breaking of fibres. In
such a scenario, the entire composite lamina will fail.

Thus, ultimate tensile strength of a unidirectional ply can be calculated as:


uc = Vfuf + (1-Vf)m

Eq 3.2-9

Where uc and uf are ultimate tensile strengths of ply and fibre, respectively, and m
is stress in matrix at a strain level equalling fracture strain in fibre.

If the fibre volume fraction does not exceed a certain threshold (Vmin), then even if all
the fibres break, the matrix will take the total load on the composite. In such a
condition, the ultimate tensile strength of composite may be written as:
uc = Vmum= (1-Vf)um

Eq 3.2-10

8
This document contains no EAR or ITAR technical data

The relation for Vmin can be developed by equating Eq 3.2-9 And Eq 3.2-10, replacing
Vf by Vmin, and solving for the latter. This is shown in Eq 3.2-11.
Vmin = (um - m )/(uf + um m)

Eq 3.2-11

Further, a well-designed unidirectional lamina requires that its ultimate tensile


strength should exceed that of matrix. This can happen only when;
uc = Vfuf + (1-Vf)m um ,
Where, um is ultimate tensile strength of matrix.
This equation is satisfied only if fibre volume fraction exceeds a certain critical value,
which is defined as:
Vcrit = (um - m )/(uf - m )

Eq 3.2-12

Thus, if:
Vf < Vmin, then failure of matrix will coincide with failure of composite, while fibres
will fail prior to failure of matrix.
Vf = Vmin, then failure of matrix, fibre and composite will happen at the same time.
Vf > Vmin, then failure of fibre, will immediately lead to failure of matrix as well as
of the composite.
Vf > Vcrit, then failure of fibre will immediately lead to failure of matrix and also the
composite. In such a case the strength of unidirectional composite will exceed that of
matrix.

3.3.Predicting Transverse Modulus of Unidirectional Lamina

Figure 17.1 shows a simple model for predicting transverse modulus of unidirectional
lamina. Here, the model constitutes of two slabs of materials, fibre and matrix, of
thicknesses tf and tm, respectively. The overall thickness of composite slab is tc,
which is sum of tf and tm. It may be noted here that these thicknesses of fibre and
matrix are directly proportional to their respective volume fractions.

9
This document contains no EAR or ITAR technical data

Figure 3.3-1 A Slab Like Model for Predicting Transverse Properties of Unidirectional
Composites

In such a system, externally imposed stress on the composite c) is assumed to be


same as that seen by fibre (f) and also by matrix (m). This is in contrast to the model
developed for predicting longitudinal modulus, where we had assumed that strains,
and not stresses, in composite, fibre and matrix are equal. Further, in such a model,
which is akin to springs in series, the overall displacement in composite c) in
transverse direction due to external load is a sum of displacement in fibre (f) and
displacement in matrix (m).
c = f + m

Eq 3.3-1

Further, recognizing the relation between strains in each constituent, and their
thicknesses, above equation can be rewritten as:
c tc = m tm + f tf

Eq 3.3-2

Dividing above equation by thickness of composite (tc), and realizing that tf/tc, and
tm/tc equal Vf and Vm, respectively, we get:
c = m Vm + f Vf

Eq 3.3-3

In linear-elastic range, strain is a ratio of stress and the modulus. Hence, above
equation can be further re-written as:
(c/Ec)= (m/Em)Vm + (f/Ef)Vf

Eq 3.3-4

However, we had earlier assumed that externally applied stress on the composite (c)
is same as that seen by fibre (f) and also by matrix (m). Thus, previous equation can
be rewritten as:
10
This document contains no EAR or ITAR technical data

1/Ec= Vm/Em + Vf/Ef

Eq 3.3-5

Or alternatively,
Ec = ( EfEm)/([(1-Vf)Ef + VfEm]

Eq 3.3-6

Ec = ( EfEm)/([(1-Vf)Ef + VfEm]

Eq 3.3-7

Eq 3.3-5 and Eq 3.3-6 give us an estimate for transverse modulus of unidirectional


lamina. The relation shows that a significant increase in fibre volume fraction is
required to raise overall transverse modulus in moderate amounts. This is in stark
contrast with longitudinal modulus, which is linearly dependent on fibre volume
fraction.
Eq 3.3-5and Eq 3.3-6, even though based on a simple model, is not borne out well be
experimental data. To address this inconsistency, several alternative models have been
developed.
However, in we will use simple and generalized expressions for transverse modulus
developed Halpin and Tsai. These are relatively simple relations, and hence easy to
use in design practice. The results from Halpin and Tsai are also quite accurate
especially if fibre volume fraction is not too close to unity. As per Halpin and Tsai,
transverse modulus (ET) can be written as:
ET/Em = (1 + Vf)/(1 - Vf)

Eq 3.3-8

Where,
= [(Ef/Em) - 1] / [(Ef/Em) + ]
Here, is a parameter that accounts for packing and fibre geometry, and loading
condition. Its values are given below for different fibre geometries. = 2 for fibres
with square and round cross-sections. = 2a/b for fibres with rectangular crosssection. Here a is the cross-sectional dimension of fibre in direction of loading,
while b is the other dimension of fibres cross-section.

11
This document contains no EAR or ITAR technical data

3.4.Shear Modulus and Poissons Ratio

A perfectly isotropic material has two fundamental elastic constants, E and . Its shear
modulus and bulk modulus can be expressed in terms of these two elastic constants.
Likewise, a transversely isotropic composite ply has four elastic constants. These are:
EL, i.e. elastic modulus in longitudinal direction.
ET i.e. elastic modulus in transverse direction.
GLT i.e. longitudinal shear modulus.
LT i.e. Poissons ratio
A detailed discussion on the mathematical logic underlying existence of these four
constants will be conducted in a subsequent lecture. Till so far, we have developed
relations for EL, and ET. Now we will learn about similar relationships for GLT and
LT. Halpin and Tsai have developed relations similar to Eq 3.3-7 which can be used
predict longitudinal shear modulus, GLT. This is shown below.
GLT/Gm = (1 + Vf)/(1 - Vf)

Eq 3.4-1

Where,
= [(Gf/Gm) - 1] / [(Gf/Gm) + 1]
For predicting Poissons ratio LT, we exploit the fact that a longitudinal tensile strain
in fibre direction, will generate Poisson contraction in transverse direction in both,
matrix and fibre materials. In this context, we also use the fact that relative strain
values for such a contraction will be proportional to each constituent materials
volume fraction. Thus, overall Poissons ratio LT for the composite can be written as:

LT = fVf + fVm

Eq 3.4-2

3.5. Transverse Strength

We have seen that a unidirectional ply, when put to tension in fibre direction tends to
break at stress values which exceed matrix tensile strength. This is particularly true
when fibre volume fraction exceeds Vcrit. Similarly, fibres play a central role in
significantly enhancing the stiffness of the ply in fibre direction, and the overall
stiffness of the system tends to far surpass that of pure matrix. This occurs because
12
This document contains no EAR or ITAR technical data

fibres, which are stronger and stiffer vis--vis matrix, carry a major portion of
external load, thereby enhancing composites stiffness and strength. However, the
same may not be said for a unidirectional ply loaded in tension in the transverse
direction. This is because load-sharing between fibre and matrix in a transversely
loaded ply is very less. In contrast, the extent of load sharing between fibre and matrix
in a longitudinally loaded ply is very significant. When a unidirectional load is
subjected to transverse tension, fibres which are far stiffer vis--vis matrix, act to
constrain matrix deformation. Such a constraint on matrix deformation tends to
increase plys transverse modulus, though only marginally (unless fibre volume
fraction is high). The deformation constraints imposed on matrix by fibres tend to
generate strain and stress concentrations in matrix material. These stress and strain
concentrations cause the matrix to fail at much lesser values of stress and strain, than
a sample of matrix material which has no fibres at all. Thus, unlike longitudinal
strength, transverse strength tends to get reduced for composites due to presence of
fibres. This reduction in transverse strength of a unidirectional ply is characterized by
a factor, S, the strength-reduction-factor. The exact value of this factor can be
calculated by using a combination of advanced elasticity formulations and numerical
solution techniques. The strength of unidirectional ply in transverse direction, uT, can
be written as:
uT = uf /S

Eq 3.5-1

3.6.About Short-Fibre Composites

It was seen earlier that unidirectional composites tend to be very stiff and strong in
fibre direction, but very weak in the transverse direction. Their weakness in transverse
direction is attributable to presence of significant stress concentration at the interface
of matrix and fibre. Given these attributes, unidirectional composites are very useful
in applications where state of stress is well known. In such applications, lamination
sequence of composite can be tailor-made to bear external loads optimally. However,
if externally applied loads are Omni-directional, or if their direction can vary in time,
then such laminates fabricated by stacking up unidirectional laminas may not
necessarily meet our design needs.
13
This document contains no EAR or ITAR technical data

We may still be able design a laminate for such cases (that is when loading is unidirectional) which is equally strong in all directions, but even in such a design, the top
and bottom layers will be weak in transverse directions, and failure could get initiated
from here. Hence, in such applications, it is useful to have laminas which have in
plane isotropy. One way to produce such lamina is by using short fibres which are
randomly oriented. Such composites, in general are significantly less expensive than
unidirectional composites. The fibre lengths in these are from 1 to 8 cm. Such
composites are used extensively in general purpose applications, such as car body
panels, boats, household goods, etc. In most of such applications, glass fibre is used as
the reinforcing material for matrix.
In composite materials, fibres are invariably surrounded by matrix material. Hence,
external load is directly applied to matrix, and from here, it gets transferred to fibres.
A part of this load gets transferred to fibres at their ends, while remaining portion of
this load gets transferred to fibres through their external cylindrical surfaces. For
unidirectional composites with continuous fibres, transfer of load at fibre ends may be
very small vis--vis load transfer through fibres external surface. This is because
fibres are very long, and hence their cylindrical surfaces, across which load gets
transferred through shear-mechanism, are sufficiently long. In such fibres, the effect
of load transfer through fibre ends may not significantly affect overall mechanics of
load transfer.
However, in short-fibre composites the same may not be necessarily true. In such
composites, the length of the fibre is not sufficiently long such that much of load
transfer happens across cylindrical surfaces of fibres. Thus, in such fibres, both the
ends, as well as external cylindrical surfaces of fibres play a significant role in matrixto-fibre load transfer. Hence, it is important to understand role of end-effects in
context of load transfer to fibres. Without this understanding, our understanding of
reinforcing effects in short-fibre composites will be inaccurate and flawed.
Consider a short-fibre of length l embedded in matrix which is shown in Figure 3.6-1.
The figure also shows the details of an infinitesimal portion of fibre of length dz,
which experiences normal stress in length direction, and shear stress, , along its
cylindrical surface. Please note that while normal stress at one end of infinitesimal
fibre is f, it is f +df at its other end. This variation in normal stress along the length
14
This document contains no EAR or ITAR technical data

of infinitesimally long fibre is because some of the load gets transferred from matrix
to fiber due to application of shear stress on its cylindrical surface.

Figure 3.6-1 Force Equilibrium of an Infinitesimal Portion of Discontinuous Fiber which is aligned to
External Load

From principles of static equilibrium, equation of force equilibrium for this


infinitesimally sized portion of fiber.
r2f + (2 r dz)f = r2(f + df )
Cancelling out term r2f from both sides, and rearranging remaining terms :
df /dz = 2/r
Integrating above equation yields,
f = fo+ (2/r) dz, where the integral limits are 0 to z.
Quite often, fiber separates from the matrix due to presence of large stress
concentration. In other cases, matrix yields at the fiber end. The implication of either
case is that the integration constant for above equation, fo, is zero. Thus, above
equation can be rewritten as: f = (2/r) dz
The integral equation shown earlier can be evaluated if variation of shear stress, ,
with respect to coordinate z, is known. At this point, an assumption is made that the
shear stress at the interface of fiber and matrix is constant along fiber length, and
equals matrix yield shear, i.e. y. Such an assumption may be made for a system

15
This document contains no EAR or ITAR technical data

where matrix material transmits maximum possible stress to fiber, which would be y.
For such a case, the integral equation can be simplified as:
f = 2y z/r

Eq 3.6-1

For short fibers, maximum fiber stress is expected to occur at mid-length, i.e. z = l/2,
while it will be zero at its extremities for reasons explained earlier. Hence, the
equation written above will hold good only for values of z = 0 to l/2, and for the
region z = l/2 to l, the equation will have to have a negative slope. Further, the
maximum value of fiber stress will be, as per above equation:
f_max = y l/r, corresponding to z = l/2

Eq 3.6-2

Eq 3.6-1 and Eq 3.6-2 place no limit on the upper bound for fiber stress, and can
approach very large values if l is made very large. However, in reality there will
indeed be a limit, which will correspond to the stress borne by continuous and
infinitely long fibers in unidirectional plies. This stress, as calculated earlier is
Ef/Emc. Equating this value to maximum fiber stress in short-fiber (as per Eq 3.6-2)
gives us a load-transfer length, lt, which is required to achieve maximum possible
stress in fiber. This is shown below.
f_max = y lt/r = Ef/Emc

Eq 3.6-3

f_max = y lt/r = Ef/Emc

Eq 3.6-4

Or,
lt/r = f_max /y = (Ef/Emc) / y

Eq 3.6-5

Thus, a fiber which is at least lt long develops maximum fiber stress (Ef/Emc) as
defined earlier, when the externally applied stress is c.
Hence, if we increase external stress c, we will have to increase lt to ensure
maximum load in fiber, as f_max, which equals Ef/Emc, will also increase. But, there

16
This document contains no EAR or ITAR technical data

is a limit beyond which external stress c cannot be increased. This limit corresponds
to a point when the stress in fiber equals its ultimate strength (uf),
At this limit, any further stress in external stress will lead to failure of fiber
Thus, the condition for maximum possible stress in fiber is:
f_max = uf = Ef/Emc

Eq 3.6-6

For such a limiting stress, there is a corresponding minimum fiber length which is
required to support such a level of stress. Mathematically, the value of minimum fiber
length can be calculated from Eq 3.6-5 and is given below.
lmin/r = uf/y

Eq 3.6-7

Thus, any design of a short fiber composite should ensure that its fiber is at least lmin
long, because in such a system the overall composite strength will be maximized. If
fibers are shorter than this critical length, then composite strength would not be at its
maximum value, thereby adding weight and cost to the structure. Finally, if l is very
large compared to lmin, then composite increasingly behaves as one with continuous
fibers.

Till so far, It was assumed that the matrix material in fiber-matrix interface region is
perfectly plastic. This is not entirely true. In reality, most matrix materials exhibit
elasto-plastic behaviour. Developing analytical solutions for such systems is not easy.
Hence, numerical methods may be used to solve such problems to get better
understanding of load transfer mechanisms in short-fiber composites. Several such
studies have shown that load transfer at fiber ends is not significant, and hence our
earlier assumption of fo being zero, stands validated, though in an approximate sense.

17
This document contains no EAR or ITAR technical data

Figure 3.6-2 is a plot of variation of fiber strength for three different fiber lengths.

Following observations can be made from Figure 3.6-2. If fiber length is less than lt,
then the normal stress in fiber is zero at either ends of fiber, and it reaches a peak
value at mid-fiber length. In such a case, the longer the fiber, the higher is the value of
peak normal stress which occurs at its mid-length. If fiber length equals lt, then
normal stress in fiber gets maximized. However, the shape of stress plot still remains
triangular. Finally, if fiber length exceeds lt, then normal stress in fiber:
Rises from zero to a maximum value over part of the fiber length.
Remains constant once it has maximized.
Falls back to zero, over remaining part of fiber length.
Utilization of fiber strength is maximized in the third configuration.

18
This document contains no EAR or ITAR technical data

3.7.Modulus of Short-Fiber Composites

There are many models to predict stiffness of uniaxial short fiber composites. In
selecting models for consideration, we impose the general requirements that each
model must include the effects of fiber and matrix properties and the fiber volume
fraction, include the effect of fiber aspect ratio, and predict a complete set of elastic
constants for the composite. Any model not meeting these criteria was excluded from
consideration. All of the models use the same basic assumptions:

The fibers and the matrix are linearly elastic, the matrix is isotropic, and the
fibers are either isotropic or transversely isotropic. .

The fibers are axisymmetric, identical in shape and size, and can be
characterized by an aspect ratio l/d

The fibers and matrix are well bonded at their interface, and remain that way
during deformation. Thus, we do not consider interfacial slip, fiber/matrix
debonding or matrix micro-cracking.

3.7.1. Eshelby's equivalent inclusion


A fundamental result used in several different models is Eshelby's equivalent
inclusion (Eshelby, 1959) & (Eshelby, 1961). Eshelby solved for the elastic stress
field in and around an ellipsoidal particle in an infinite matrix. By letting the particle
be a prolate ellipsoid of revolution, one can use Eshelby's result to model the stress
and strain fields around a cylindrical fiber. Eshelby first posed and solved a different
problem, that of a homogeneous inclusion Figure 3.7-1. Consider an in finite solid
body with stiffness Cm that is initially stress-free. All subsequent strains will be
measured from this state. A particular small region of the body will be called the
inclusion, and the rest of the body will be called the matrix. Suppose that the inclusion
undergoes

19
This document contains no EAR or ITAR technical data

Figure 3.7-1 Eshelby's inclusion problem.

Starting from the stress-free state (a), the inclusion undergoes a stress-free
transformation strain T(b). Fitting the inclusion and matrix back together (c) produces
the strain state C(x) in both the inclusion and the matrix.
Some type of transformation such that, if it were a separate body, it would acquire a
uniform strain T with no surface traction or stress. T is called the transformation
strain, or the eigen strain. This strain might be acquired through a phase
transformation, or by a combination of a temperature change and a different thermal
expansion coefficient in the inclusion. In fact the inclusion is bonded to the matrix, so
when the transformation occurs the whole body develops some complicated strain
field C(x) relative to its shape before the transformation. Within the matrix the stress
is simply the stiffness times this strain,
m(x)= Cm C(x)

Eq 3.7.1-1

But within the inclusion the transformation strain does not contribute to the stress, so
the inclusion stress is
I =Cm (C-T)

Eq 3.7.1-2

The key result of Eshelby was to show that within an ellipsoidal inclusion the strain C
is uniform, and is related to the transformation strain by
C= ET

Eq 3.7.1-3

20
This document contains no EAR or ITAR technical data

E is called Eshelby's tensor, and it depends only on the inclusion aspect ratio and the
matrix elastic constants. Mura gave a detailed derivation and applications (Mura,
1982) , and analytical expressions for Eshelby's tensor for an ellipsoid of revolution in
an isotropic matrix appear in many papers. The strain field C(x) in the matrix is
highly non-uniform (Eshelby, 1959), but this more complicated part of the solution
can often be ignored. The second step in Eshelby's approach is to demonstrate
equivalence between the homogeneous inclusion problem and an inhomogeneous
inclusion of the same shape. Consider two infinite bodies of matrix, as shown in Fig.
2. One has a homogeneous inclusion with some transformation strain T. The other
has an inclusion with a different stiffness Cf , but no transformation strain. Subject
both bodies to a uniform applied strain A at infinity. We wish to find the
transformation strain T that gives the two problems the same stress and strain
distributions.

Figure 3.7-2 Eshelby's equivalent inclusion problem.

The inclusion (a) with transformation strain T has the same stress T and strain as the
in- homogeneity (b) when both bodies are subject to a far-field strain A:

For the first problem the inclusion stress is just Eq 3.7.1-2 with the applied strain
added,

I = Cm (A + C - T )

Eq 3.7.1-4

21
This document contains no EAR or ITAR technical data

While the second problem has no T but a different stiffness, giving a stress of
I = Cf (A + C )

Eq 3.7.1-5

Equating these two expressions gives the transformation strain that makes the two
problems equivalent. Using Eq 3.7.1-3 and some rearrangement, the result is
-[ Cm + (Cf - Cm ) E] T = (Cf - Cm ) A

Eq 3.7.1-6

Note that T is proportional to A, which makes the stress in the equivalent


inhomogeneity proportional to the applied strain.

3.7.2. Dilute Eshelby model


One can use Eshelby's result to find the stiffness of a composite with ellipsoidal fibers
at dilute concentrations. To find the stiffness one only has to find the strainconcentration tensor A. To do this, first note that for a dilute composite the average
strain is identical to the applied strain,
^ = A

Eq 3.7.2-1

Since this is the strain at infinity. Also, from Eshelby, the fiber strain is uniform, and
is given by
^f = A + C

Eq 3.7.2-2

Where, the right-hand side is evaluated within the fiber. Now write the equivalence
between the stresses in the homogeneous and the inhomogeneous inclusions, Eq
3.7.1-4 and Eq 3.7.1-5,
Cf (A + C) = Cm (A + C - T )

Eq 3.7.2-3

Then use Eq 3.7.1-3, Eq 3.7.2-1 and Eq 3.7.2-2to eliminate T, A and C from this
equation, giving
22
This document contains no EAR or ITAR technical data

[I + ESm (Cf - Cm)] ^f = ^

Eq 3.7.2-4

Comparing this to
f = A

Eq 3.7.2-5

Shows that the strain-concentration tensor for Eshelby's equivalent inclusion is


AEshelby = [I + ESm (Cf - Cm)] -1

Eq 3.7.2-6

This can be used in


C= Cm + vf (Cf- Cm )A

Eq 3.7.2-7

To predict the moduli of aligned- fiber composites, a result (Russel, 1973) was
developed. Calculations using this model to explore the effects of particle aspect ratio
on stiffness (Chow, 1977) were also presented. While Eshelby's solution treats only
ellipsoidal fibers, the fibers in most short- fiber composites are much better
approximated as right circular cylinders. In their paper the relationship between
ellipsoidal and cylindrical particles (Steif & Hoysan, 1987) was considered , who
developed a very accurate finite element technique for determining the stiffening
effect of a single fiber of given shape. For very short particles, l/d = 4, they found
reasonable agreement for E11 by letting the cylinder and the ellipsoid have the same
l/d. The ellipsoidal particle gave a slightly stiffer composite, with the difference
between the two results increasing as the modulus ratio Ef = Em increased. Henceforth
we will use the cylinder aspect ratio in place of the ellipsoid aspect ratio in Eshelbytype models. Because Eshelby's solution only applies to a single particle surrounded
by an infinite matrix, AEshelby is independent of fiber volume fraction and the stiffness
predicted by this model increases linearly with fiber volume fraction. Modulus
predictions based on Eq 3.7.2-5 and Eq 3.7.2-7 should be accurate only at low
volume fractions, say up to vf of 1%. The more difficult problem is to find some way
to include interactions between fibers in the model, and so produce accurate results at
higher volume fractions. We next consider approaches for doing that.

23
This document contains no EAR or ITAR technical data

3.7.3. Mori-Tanaka Model


A family of models for non-dilute composite materials (Mori & Tanaka, 1973) had
evolved. Later a particularly simple and clear explanation (Benveniste, 1987) of the
Mori-Tanaka approach was given to introduce the approach. Suppose that a composite
is to be made of a certain type of reinforcing particle, and that, for a single particle in
an infinite matrix, we know the dilute strain-concentration tensor AEshelby
f=AEshelby^

Eq 3.7.3-1

The Mori-Tanaka assumption is that, when many identical particles are introduced in
the composite, the average fiber strain is given by
f =AEshelby^m

Eq 3.7.3-2

That is, within a concentrated composite each particle sees a far-field strain equal to
the average strain in the matrix. Using the alternate strain concentrator defined in eqn
f = Am

Eq 3.7.3-3

The Mori-Tanaka assumption can be re-stated as


A^MT=AEshelby

Eq 3.7.3-4

Then
A=A^ [(1-vf) I + vf A^]-1

Eq 3.7.3-5

Gives the Mori-Tanaka strain concentrator as


AMT=AEshelby[(1-vf) I + vf A]-1

Eq 3.7.3-6

This is the basic equation for implementing a Mori-Tanaka model. The Mori-Tanaka
approach for modelling composites was first introduced by Wakashima et. al
(Wakashima, Umekawa, & Otsuka, 1974) for modelling thermal expansions of
composites with aligned ellipsoidal inclusions. Mori and Tanaka treat only the
homogeneous inclusion problem (Mori & Tanaka, 1973), and say nothing about
24
This document contains no EAR or ITAR technical data

composites. Mori-Tanaka predictions for the longitudinal modulus of a short-fiber


composite were again developed to also include the effects of cracks and of a second
type of reinforcement by (Taya & Mura, 1981) and (Taya & Chou, 1981). Their
method was generalised (Weng, 1984), lead to the usage of the Mori-Tanaka approach
to develop equations for the complete set of elastic constants of a short-fiber
composite (Tandon & Weng, 1984). Tandon andWengs equations for the plane-strain
bulk modulus k23and the major Poisson ratio 12 must be solved iteratively.
The usual development of the Mori-Tanaka model differs somewhat from
Benvenistes explanation. For an average applied stress
, the reference strain 0 is

defined as the strain in a homogeneous body of matrix at this stress,

= Cm 0

Eq 3.7.3-7

Within the composite the average matrix strain differs from the reference strain by
some perturbation
= 0+ 

Eq 3.7.3-8

A fiber in the composite will have an additional strain perturbation ~ f , such that
 =  + + 

Eq 3.7.3-9

While the equivalent inclusion will have this strain plus the transformation strain T.
The stress equivalence between the inclusion and the fiber then becomes

 (  + +  ) =  (  + +   )

Eq 3.7.3-10

Compare this to the dilute version, Eq 3.7.2-3, noting that A in the dilute problem is
equivalent to ( 0 +   ) here. The development is completed by assuming that the

extra fiber perturbation is related to the transformation strain by Eshelbys tensor,


 =  

Eq 3.7.3-11

Combining this with Eq 3.7.3-8 and Eq 3.7.3-9 reveals that Eq 3.7.3-11 contains the
essential Mori-Tanaka assumption: the fiber in a concentrated composite sees the
25
This document contains no EAR or ITAR technical data

average strain of the matrix. Some other micromechanics models are equivalent to the
Mori-Tanaka approach, though this equivalence has not always been recognized.
Chow considered Eshelbys inclusion problem and conjectured that in a concentrated
composite the inclusion strain would be the sum of two terms: the dilute result given
by Eshelby and the average strain in the matrix (Chow, 1978).
(  ) =   + (  )

Eq 3.7.3-12

This can be combined with the definition of the average strain from eqn (7) to relate
the inclusion strain ( C)f to the transformation strain T
(  ) = (1  )  

Eq 3.7.3-13

Chow then extended this result to an inhomogeneity following the usual arguments,
Eq 3.7.1-4 to Eq 3.7.2-6. This produces a strain-concentration tensor


= [# + $1  %& '  (]


*+

Eq 3.7.3-14

Which is equivalent to the Mori-Tanaka result Eq 3.7.3-6. Chow was apparently


unaware of the connection between his approach and the Mori-Tanaka scheme, but he
seems to have been the first to apply the Mori-Tanaka approach to predict the stiffness
of short-fiber composites. A more recent development is the equivalent poly-inclusion
model (Ferrari, 1994). Rather than use the strain-concentration tensor A, Ferrari used
an effective Eshelby tensor , , defined as the tensor that relates inclusion strain to

transformation strain at finite volume fraction:


(  ) = , 

Eq 3.7.3-15

Once , has been defined, it is straightforward to derive a strain-concentration tensor

A and a composite modulus. Ferrari considered admissible forms for , , given the

requirements that , must (a) produce a symmetric stiffness tensor C, (b) approach
Eshelbys tensor E as volume fraction approaches zero, and (c) give a composite
stiffness that is independent of the matrix stiffness as volume fraction approaches
unity. He proposed a simple form that satisfies these criteria,
26
This document contains no EAR or ITAR technical data

 = $1  %

Eq 3.7.3-16

The combination of Eq 3.7.3-15 and Eq 3.7.3-16 is identical to Chows assumption


Eq 3.7.3-15 and, for aligned fibers of uniform length, Ferraris equivalent polyinclusion model, Chows model, and the Mori- Tanaka model are identical. Important
differences between the equivalent poly-inclusion model and the Mori-Tanaka model
arise when the fibers are not oriented or have different lengths.

3.7.4. Self-Consistent Models


A second approach to account for finite fiber volume fraction is the self-consistent
method (Hill, 1965) and (Budiansky, 1965). The original work focused on spherical
particles and continuous, aligned fibers. The application to short-fiber composites was
developed (Laws & McLaughlin, 1979) and (Chou, Nomura, & Taya, 1980). In the
self-consistent scheme one finds the properties of a composite in which a single
particle is embedded in an infinite matrix that has the average properties of the
composite. For this reason, self-consistent models are also called embedding models.
Again building on Eshelbys result for a ellipsoidal particle, we can create a selfconsistent version of Eq 3.7.2-6 by replacing the matrix stiffness and compliance
tensors by the corresponding properties of the composite. This gives the selfconsistent strain-concentration tensor as
. = [# + &$  %]*+

Eq 3.7.4-1

Of course the properties C and S of the embedding matrix are initially unknown.
When the reinforcing particle is a sphere or an infinite cylinder, the equations can be
manipulated algebraically to find explicit expressions for the overall properties. For
short fibers this has not proved possible, but numerical solutions are easily obtained
by an iterative scheme. One starts with an initial guess at the composite properties,
evaluates E and then ASC from Eq 3.7.4-1, and substitutes the result into Eq 3.7.2-7 to
get an improved value for the composite stiffness. The procedure is repeated using
this new value, and the iterations continue until the results for C converge. An
additional, but less obvious, change is that Eshelbys tensor E depends on the matrix
properties, which are now transversely isotropic. Expressions for Eshelbys tensor for
27
This document contains no EAR or ITAR technical data

an ellipsoid of revolution in a transversely isotropic matrix were given in (Chou,


Nomura, & Taya, 1980) and (Lin & Mura, 1973). With these expressions in hand one
can use Eq 3.7.4-1 together with Eq 3.7.2-7 to find the stiffness of the composite. This
is the self-consistent approach used for short-fiber composites. A closely-related
approach, called the generalized self-consistent model, also uses an embedding
approach. However, in these models the embedded object comprises both fiber and
matrix material. When the composite has spherical reinforcing particles, the
embedded object is a sphere of the reinforcement encased in a concentric spherical
shell of matrix; this is in turn surrounded by an infinite body with the average
composite properties. The generalized self-consistent model is sometimes referred to
as a double embedding approach. For continuous fibers the embedded object is a
cylindrical fiber surrounded by a cylindrical shell of matrix. The first generalized selfconsistent models were developed for spherical particles (Kerner, 1956), and for
cylindrical fibers (Hermans, 1967). Both of these papers contain an error, which was
discussed and corrected later (Christensen & Lo, 1979). While the generalized selfconsistent model is widely regarded as superior to the original self-consistent
approach, no such model has been developed for short fibers.

3.7.5. Bounding Models


A rather different approach to modelling stiffness is based on finding upper and lower
bounds for the composite moduli. All bounding methods are based on assuming an
approximate field for either the stress or the strain in the composite. The unknown
field is then found through a variational principle, by minimizing or maximizing some
functional of the stress and strain. The resulting composite stiffness is not exact, but it
can be guaranteed to be either greater than or less than the actual stiffness, depending
on the variational principle. This rigorous bounding property is the attraction of
bounding methods. Historically, the Voigt and Reuss averages were the first models
to be recognized as providing rigorous upper and lower bounds. To derive the Voigt
model,
/

012

=  +  $  % =    +  

28
This document contains no EAR or ITAR technical data

Eq 3.7.5-1

One assumes that the fiber and matrix have the same uniform strain, and then
minimizes the potential energy. Since the potential energy will have an absolute
minimum when the entire composite is in equilibrium, the potential energy under the
uniform strain assumption must be greater than or equal to the exact result, and the
calculated stiffness will be an upper bound on the actual stiffness. The Reuss model,
& 34566 = & +  $&  & % = &  +  &

Eq 3.7.5-2

is derived by assuming that the fiber and matrix have the same uniform stress, and
then maximizing the complementary energy. Since the complementary energy must
be a maximum at equilibrium, the model provides a lower bound on the composite
stiffness. Detailed derivations of these bounds are provided in (Wu C & McCullough,
1977). The Voigt and Reuss bounds provide isotropic results (provided the fiber and
matrix are themselves isotropic), when in fact we expect aligned-fiber composites to
be highly anisotropic. More importantly, when the fiber and matrix have substantially
different stiffness then the Voigt and Reuss bounds are quite far apart, and provide
little useful information about the actual composite stiffness. This latter point
motivated Hashin and Shtrikman to develop a way to construct tighter bounds. Hashin
and Shtrikman developed an alternate vibrational principle for heterogeneous
materials (Hashin & Shtrikman, 1963). Their method introduces a reference material,
and bases the subsequent development on the differences between this reference
material and the actual composite. Rather than requiring two variation principles, like
the Voigt and Reuss bounds, their single variation principle gives both the upper and
lower bounds by making appropriate choices of the reference material. For an upper
bound the reference material must be as stiff or stiffer than any phase in the composite
(fiber or matrix), and for a lower bound the reference material must have a stiffness
less than or equal to any phase. In most composites the fiber is stiffer than the matrix,
so choosing the fiber as the reference material gives an upper bound and choosing the
matrix as the reference material gives a lower bound. If the matrix is stiffer than the
fiber, the bounds are reversed. The resulting bounds are tighter than the Voigt and
Reuss bounds, which can be obtained from the Hashin-Shtrikman theory by giving the
reference material infinite or zero stiffness, respectively. Hashin and Shtrikmans
original bounds apply to isotropic composites with isotropic constituents. Frequently
the bounds are regarded as applying to composites with spherical particles, orientation
29
This document contains no EAR or ITAR technical data

must also obey the bounds. Walpole re-derived the Hashin-Shtrikman bounds using
classical energy principles (Walpole, 1966), and extended them to anisotropic
materials (Walpole, 1966). Walpole also derived results for infinitely long fibers and
infinitely thin disks in both aligned and 3-D random orientations (Walpole, 1969).
The Hashin-Shtrikman-Walpole bounds were extended to short-fiber composites in
(Willis, 1977) and (Wu C & McCullough, 1977). These workers introduced a twopoint correlation function into the bounding scheme, allowing aligned ellipsoidal
particles to be treated. Based on these extensions, explicit formulae for aligned
ellipsoids were developed in (Weng, 1992) and (Eduljee, McCullough, & Gillespie,
1994). The general bounding formula, shown here in the format developed by Weng,
gives the composite stiffness C as
 = 7   8  +   8 9[ 8  +  8 ]*+

Eq 3.7.5-3

Where the tensors Qf and Qm are defined as


1

8 = [# + 0 &0 ' 0 (] :;< 8 = [# + 0 &0 ' 0 (]


Here Eo is Eshelbys tensor associated with the properties of the reference material,
which has stiffness So and compliance Co When the matrix is chosen as the reference
material, Eq 3.7.5-3 gives a strain concentrator of
=>

!4?

= [# +  & $   %]*+

Eq 3.7.5-4

This result is labelled here as the lower bound, on the presumption that the fiber is
stiffer than the matrix. The composite stiffness is found by substituting 

@ABCD

into Eq

3.7.3-5 and Eq 3.7.2-7. Eduljee et. al argue that the lower bound provides the most
accurate estimate of composite properties (Eduljee, McCullough, & Gillespie, 1994),
and recommend it as a model. Note that this lower bound prediction is identical to the
Mori-Tanaka model, Eq 3.7.3-6 . This correspondence lends theoretical support to the
Mori- Tanaka approach, and guarantees that it will always obey the bounds. The other
bound, found by using Eq 3.7.5-3 with the fiber as the reference material, has a strain
concentrator of

=5EE4? = [# +   & $   %]

Eq 3.7.5-5

30
This document contains no EAR or ITAR technical data

Note that the Eshelby tensor Ef is now computed for inclusions of matrix material
surrounded by the fiber material. Eq 3.7.5-5 is labelled as the upper bound, presuming
that the fiber is stiffer than the matrix. An identical result can be obtained from the
Mori-Tanaka theory by assuming that ellipsoidal particles of the matrix material are
embedded in a continuous phase of the fiber material. If the matrix is stiffer than the
fibers, then the right-hand sides of Eq 3.7.5-4 and Eq 3.7.5-5 are unchanged but Eq
3.7.5-4 becomes the upper bound and Eq 3.7.5-5 becomes the lower bound. All of the
preceding bounding formulae have been given for two-component composites, but the
theory readily accommodates multiple reinforcements. At fiber volume fractions close
to unity, the matrix stiffness strongly influences the composite stiffness for the lower
bound/Mori-Tanaka models, despite the tiny amount of it that is present. Packing
considerations suggest that the only way to approach such high volume fractions is for
the fiber phase to become continuous, and Lielens suggest that at very high fiber
volume fractions the composite stiffness should be much closer to the upper bound, or
equivalently to the Mori-Tanaka prediction using the fiber as the continuous phase
(Lielens, 1997). This insight prompted Lielens and co-workers to propose a model
that interpolates between the upper and lower bounds, such that the lower bound
dominates at low volume fractions and the upper bound dominates at high volume
fractions (again presuming the fiber is the stiffer phase). They perform this
interpolation on the inverse of the strain-concentration tensor ^A, producing the
predictive equation
=F04>4G6 = {(1 )7=>

!4? *+

+ [7=5EE4? 9 }*+
*+

Eq 3.7.5-6

The interpolating factor f depends on fiber volume fraction, and they propose
=

JK L JK M
N

This theory reproduces the lower bound and Mori-Tanaka results at low volume
fractions, but is said to give improved results at reinforcement volume fractions in the
40 to 60% range.

31
This document contains no EAR or ITAR technical data

3.7.6. Shear Lag Models


Historically, shear lag models were the first micromechanics models for short-fiber
composites, as well as the first to examine behavior near the ends of broken fibers in a
continuous-fiber composite. Despite some serious theoretical flaws, shear lag models
have enjoyed enduring popularity, perhaps due to their algebraic simplicity and their
physical appeal. Classical shear lag models only predict the longitudinal modulus E11,
so they do not meet the criterion of predicting a complete set of elastic constants.
However, it is included here because of their historical importance and their
widespread use. One could obtain a complete stiffness model by using the shear lag
prediction for E11 and some continuous-fiber model (such as Hermans) for the
remaining elastic constants. If the fiber is anisotropic then its axial modulus should be
used in the shear lag equations. The shear lag analysis focuses on a single fiber of
length l and radius rf (Cox, 1952), which is encased in a concentric cylindrical shell of
matrix having radius R. The fiber is aligned parallel to the z-axis, as shown in Figure
3.7-3. Only the axial stress 11 and axial strain 11 are of interest, and Poisson effects
are neglected so that f11=Eff11. The outer cylindrical surface of the matrix is
subjected to displacement boundary conditions consistent with an average axial strain
11 and one solves for the fiber stress f11(z). (More rigorously, f11(z) is the average
stress over the fiber cross-section at z)

Figure 3.7-3 Idealized fiber and matrix geometry used in shear lag models.

32
This document contains no EAR or ITAR technical data

<
++ =


2P?Q
D

Eq 3.7.6-1

Where rz is the axial shear stress at the fiber surface. The key assumption of shear lag
theory is that rz is proportional to the difference in displacement w between the fiber
surface and the outer matrix surface:
P?Q (R) =

S
[B(U, R) B$D , R%]
2TD

Eq 3.7.6-2

Where H is a constant that depends on matrix properties and fiber volume fraction.
Solving Figure 3.7-3 for
++ (z) and applying boundary conditions of zero stress at the


fiber ends gives an average fiber stress of




++

Y>

W:; ' N (
=  ++ [1
]
Y>
'N(

With
ZN =

Eq 3.7.6-3

S
N
TD++


Eq 3.7.6-4

It is convenient to rewrite this as an expression for the average fiber strain,


++ = [>
++


Eq 3.7.6-5

where l is a length-dependent efficiency factor,


Y>

W:; ' N (
[> = [1
]
Y>
'N(

Eq 3.7.6-6

Note that l is a scalar analog of the strain-concentration tensor A defined in Eq


3.7.2-5, and (1/) is a characteristic length for stress transfer between the fiber and the
matrix.

Table 3.7.6-1Values for KR used in Eq 3.7.6-8 for shear lag models

33
This document contains no EAR or ITAR technical data

Fiber Packing

KR

Cox

2/3=3.628

Composite Cylinders

Hexagonal

/23=0.907

Square

/4=0.785

It was found that the coefficient H by solving a second idealized problem (Cox,
1952). The concentric cylinder geometry is maintained, but the outer cylindrical
surface of the matrix is held stationary and the inner cylinder, which is now rigid, is
subjected to a uniform axial displacement. An elasticity solution for the matrix layer
then gives
S=

2T^
3

@; ( )

Eq 3.7.6-7

?K

This part of the problem was simplified by assuming that the matrix shell was thin
compared to the fiber radius, (R- rf)<< rf, obtaining H (Rosen, 1964)
S=

2T^
3

_? ` 1

Eq 3.7.6-8

Rosens approximation gives an error in Hof 10% at vf= 0:6, with much larger errors
at lower volume fractions, and we will not consider it further. It remains to choose the
radius R of the matrix cylinder, and the exact choice is important. Several choices
have been used, all of which can be written in the form
U
b3
=a
D
c

Eq 3.7.6-9

Where KR is a constant that depends on the assumption used to find R. Table 3.7.6-1
summarizes the choices for KR. A hexagonal packing, and chose as the distance
between centres of nearest-neighbour fibers. (Cox, 1952) Figure 3.7-4 a. It seems
more realistic to let equal half of the distance between nearest neighbours Figure 3.7-4
b, a choice labelled hexagonal in Table 3.7.6-1. rf2/R2 = vf was chosen in (Rosen,
34
This document contains no EAR or ITAR technical data

1964), and later (Carman & Reifsnider, 1992) so that the concentric cylinder model in
Figure 3.7-3 would have the same fiber volume fraction as the composite. This is the
same R as the composite cylinders model explained in (Hashin & Rosen, 1964). More
recently, a square array of fibers was assumed (Robinson & Robinson, 1994), and
chose R as half the distance between centres of nearest neighbours (Figure 3.7-4 c).
Each of these choices gives a somewhat different dependence of l on fiber volume
fraction, with larger values of KR producing lower values of E11. Shear lag models are
usually completed by combining the average fiber stress in Eq 3.7.6-3 with an average
matrix stress to produce a modified rule of mixtures for the axial modulus:
++ = [> c  + $1 c %

Eq 3.7.6-10

Figure 3.7-4 Fiber packing arrangements used to find R in shear lag models. (a) Hexagonal
(Cox, 1952). (b) Hexagonal (Rosen, 1964) (c) Square (Robinson & Robinson, 1994).

However, the matrix stress in this formula is not consistent with the basic concepts of
average stress and average strain. Note that
= c  + c

Eq 3.7.6-11

must hold for 11, as for any other component of strain. Combining this with Eq
3.7.6-5 to find the average matrix strain, and following through to find the composite
stiffness (with Poisson effects neglected), gives a result that is consistent with both the
assumptions of shear lag theory and the basic concepts of average stress and strain:
35
This document contains no EAR or ITAR technical data

++ = [> c  + $1 [> c %


=  + c (  )[>

Eq 3.7.6-12

This equation is an exact scalar analog of the general tensorial stiffness formula, Eq
3.7.2-7. For the cases in this paper, the difference between Eq 3.7.6-10 and Eq
3.7.6-12 is small. A model by (Fukuda & Kawata, 1974) for the axial stiffness of
aligned short-fiber composites is closely related to shear lag theory. They begin with a
2-D elasticity solution for the shear stress around a single slender fiber in an infinite
matrix. The usual shear lag relation, Eq 3.7.6-1, is used to transform this into an
equation for the fiber stress distribution, which is then approximated by a Fourier
series. The coefficients of a truncated series are evaluated analytically using
Galerkins method. This is a dilute theory, in which modulus varies linearly with fiber
volume fraction.

Like any shear lag theory, Fukuda and Kawatas theory predicts that E11 approaches
the rule of mixtures result as the fiber aspect ratio approaches infinity. But for short
fibers Fukuda and Kawatas theory gives much lower E11 values than shear lag theory.
In Fukuda and Kawatas theory, the ratio of fiber strain to matrix strain is governed by
the parameter (l/d)(Em/Ef). In contrast, for shear lag theory, Eq 3.7.6-6, the governing
parameter is l/2, which is proportional to (l/d)(Em/Ef). Thus, for high modulus ratio
and low aspect ratio, Fukuda and Kawatas theory tends to under predict E11.

3.7.7. Halpin-Tsai Equations

The Halpin-Tsai equations (Ashton, Halpin, & Petit, 1969) have long been popular for
predicting the properties of short-fiber composites. A detailed review and derivation is
provided by (Halpin & Kardos, The Halpin-Tsai Equations: A Review, 1976), from
which the main points are summarized. The Halpin-Tsai equations were originally
developed with continuous-fiber composites in mind, and were derived from the work
of (Hermans, 1967) and (Hill, 1964). Hermans developed the first generalized selfconsistent model for a composite with continuous aligned fibers (see Section 3.7.4).
36
This document contains no EAR or ITAR technical data

Halpin and Tsai found that three of Hermans equations for stiffness could be
expressed in a common form:
f

K
'f ( 1
1 + d[

g
=
BeW [ = fK

1 [
' (+1

Eq 3.7.7-1

fg

Here P represents any one of the composite moduli listed in Table 1, and Pf and Pm
are the corresponding moduli of the fibers and matrix, while is a parameter that
depends on the matrix Poisson ratio and on the particular elastic property being
considered. Hermans derived expressions for the plane-strain bulk modulus k23, and
for the longitudinal and transverse shear moduli G12and G23. The parameters for
these properties are given in Table 1. Note that for an isotropic matrix

Table 3.7.7-1 Correspondence between Halpin-Tsai Eq 3.7.7-1 and generalized selfconsistent predictions of (Hermans, 1967) and (Kerner, 1956). After (Halpin & Kardos, The
Halpin-Tsai Equations: A Review, 1976)
P

Pf

Pm

k23

kf

km

G23

Gf

Gm

G12

Gf

Gm

Kf

Km

Gf

Gm

Comments

N
1  2
1 + 

1 + 
N
3  4
1

2(1 2 )
1 + 

Plane strain bulk modulus, aligned fibers


Transverse shear modulus, aligned fibers
Longitudinal shear modulus, aligned fibers
Bulk modulus, particulates

7 5 )
8 10

Shear modulus, particulates

(Hill, 1964) showed that for a continuous, aligned-fiber composite the remaining

stiffness parameters are given by

37
This document contains no EAR or ITAR technical data

++

 
 
1
=   +   4 l +
r
+ n op
p p
Nq
m
m
K

+N =   +   + l

 
+
mK

Eq 3.7.7-2

+ n op
Nq
mg

 

r
p p

Eq 3.7.7-3

This completes Hermans model for aligned-fiber composites; note that one must
know k23 to find E11and 12. We now know that Hermans result for G23 is incorrect,
in that it does not satisfy all of the fiber/matrix continuity conditions (Hashin, 1983).
It is, however, identical to a lower bound on G23 derived by (Hashin, 1965). Hermans
remaining results are identical to Hashin and Rosens composite cylinders assemblage
model (Hashin & Rosen, 1964), so Hermans k23, and thus his E11 and 12, are
identical to the self-consistent results of (Hill, 1965).

The Halpin-Tsai form can also be used to express equations for particulate
composites derived by (Kerner, 1956), who also used a generalized self-consistent
model. Table 3.7.7-1 gives the details. Kerners result for shear modulus G is also
known to be incorrect, but reproduces the Hashin- Shtrikman-Walpole lower bound
for isotropic composites, while Kerners result for bulk modulus K is identical to
Hashins composite spheres assemblage model (Hashin, 1962). See (Christensen &
Lo, 1979) and (Hashin, 1983) for further discussion of Kerners and Hermans results.
To transform these results into convenient forms for continuous-fiber composites,
Halpin and Tsai made three additional ad hoc approximations:

Eq 3.7.7-1 can be used directly to calculate selected engineering constants, with E11or
E22 replacing P.

The parameters in Table 3.7.7-1 are insensitive to m, and can be approximated by


constant values.

The underlined terms in Eq 3.7.7-2 and Eq 3.7.7-3 can be neglected.

38
This document contains no EAR or ITAR technical data

Table 3.7.7-2 Traditional Halpin-Tsai parameters for short-fiber composites, used in Eq


3.7.7-1. For G23 see Table 3.7.7-1.
P

Pf

Pm

Comments

E11

Ef

Em

2(l/d)

Longitudinal modulus

E22

Ef

Em

Transverse modulus

G12

Gf

Gm

Longitudinal shear modulus

12

Poisons Ratio = f f + m m

In Eq 3.7.7-2 the underlined term is typically negligible, and dropping it gives the
familiar rule of mixtures for E11of a continuous-fiber composite. However, dropping
the underlined term in Eq 3.7.7-3 and using a rule of mixtures for 12 is not
necessarily accurate if the fiber and matrix Poisson ratios differ. Halpin and Tsai
argue for this latter approximation on the grounds that laminate stiffnesses are
insensitive to 12. In adapting their approach to short-fiber composites, Halpin and
Tsai noted that must lie between 0 and . If =0 then Eq 3.7.7-1 reduces to the
inverse rule of mixtures (Halpin & Kardos, The Halpin-Tsai Equations: A Review,
1976),

1  
= +
  

Eq 3.7.7-4

while for = the Halpin-Tsai form becomes the rule of mixtures,


 =   +  

Eq 3.7.7-5

Halpin and Tsai suggested that was correlated with the geometry of the
reinforcement and, when calculating E11, it should vary from some small value to
infinity as a function of the fiber aspect ratio l/d. By comparing model predictions
with available 2-D finite element results, they found that =2(l/d) gave good
predictions for E11of short-fiber systems. Also, they suggested that other engineering
constants of short-fiber composites were only weakly dependent on fiber aspect ratio,
and could be approximated using the continuous-fiber formulae (Halpin, 1969). The
resulting equations are summarized in Table 3.7.7-2. The early references (Ashton,
39
This document contains no EAR or ITAR technical data

Halpin, & Petit, 1969) and (Halpin, 1969) do not mention G23. When this property is
needed the usual approach is to use the value given in Table 3.7.7-1. While the
Halpin-Tsai equations have been widely used for isotropic fiber materials, the
underlying results of Hermans and Hill apply to transversely isotropic fibers, so the
Halpin-Tsai equations can also be used in this case. The Halpin-Tsai equations are
known to fit some data very well at low volume fractions, but to under-predict some
stiffnesses at high volume fractions. This has prompted some modifications to their
model. (Hewitt & Malherbe, 1970) proposed making a function of vf, and by curve
fitting found that
d = 1 + 40 +

Eq 3.7.7-6

This gave good agreement with 2-D finite element results for G12 of continuous fiber
composites. (Lewis & Nielsen, 1970) & (Nielsen, 1970) focused on the analogy
between the stiffness G of a composite and the viscosity of a suspension of rigid
particles in a Newtonian fluid, noting that one should find / m = G / Gm when the
reinforcement is rigid (Gf/Gm) and the matrix is incompressible. They developed
an equation in which the stiffness not only matches dilute theory at low volume
fractions, but also displays G/Gm) as vf approaches a packing limit vfmax. This
leads to a modified Halpin-Tsai form

1 + d[

=
 1 s( )d[

Eq 3.7.7-7

with retaining its definition from Eq 3.7.7-1. Here the function (vf) contains the
maximum volume fraction vfmax as a parameter. is chosen to give the proper
behaviour at the upper and lower volume fraction limits, which leads to forms such as
s$ % = 1 + o

s$ % =

1  tu
r 
 tu N

Eq 3.7.7-8


1
v1 Cwx o
rz

1 ( / tu)

40
This document contains no EAR or ITAR technical data

Eq 3.7.7-9

The Nielsen and Lewis model improves on the Halpin-Tsai predictions, compared to
experimental data for G of particle-reinforced polymers (Lewis & Nielsen, 1970) and
to finite element calculations for G12 of continuous-fiber composites (Nielsen, 1970),
using vfmax values from 0.40 to 0.85.
Recently (Ingber & Papathanasiou, 1997) tested the Halpin-Tsai equation and
its modifications against boundary element calculations of E11 for aligned short fibers.
They found the Nielsen modification to be better than the original Halpin-Tsai form.
Hewitt and deMalherbes form could be adjusted to fit data for any single l/d, but was
not useful for predictions over a range of aspect ratios.
3.7.8. Fiber Efficiency Factor Approach
(Blumentritt, VU, & Cooper, 1975) proposed a method to calculate the ultimate
strength and the Youngs modulus of the composite in the plane of the fibers. Their
results are summarized below

5{ = b|
5 } +
(1 } )

Eq 3.7.8-1

{ = b~  } +  (1 } )

Eq 3.7.8-2

where, uc is the ultimate strength of the composite, K is the fiber efficiency


factor for strength, uf is the ultimate strength of the fiber, Vf is the fiber volume
fraction, m is the matrix stress at the fracture strain of the composite, Ec is the
modulus of the composite, KE is the fiber efficiency factor for modulus, Ef is the
modulus of the fiber and Em is the matrix modulus.
(Blumentritt, VU, & Cooper, 1975) measured the mechanical properties of

discontinuous fiber reinforced thermoplastics fabricated using six types of


reinforcement fiber and five types of thermoplastics matrix resin. The fibers used
were Dupont type 702 nylon 6/6, Dupont type 73 poly (ethylene terephthalate),
Kuralon poly (vinyl alcohol), Owens-Cornings type 801 E-glass, Dupont Kevlar-49,
and Union Carbide Thornel 300 graphite. The poly (vinyl alcohol) fibers were 5 mm
in length and the glass fibers were 6.3 mm in length. The other fibers were all 9.5 mm
in length. The five thermoplastics used were Dupont Surlyn 1558 type 30 ionomer,
Dupont Alathon 7140 high-density polyethylene, Huels grade L-1901 nylon 12,
General Electric Lexan 105-111 polycarbonate and Dupont Lucite 47 poly (methyl

41
This document contains no EAR or ITAR technical data

methacrylate) (PMMA). All together thirty different combinations of fiber/resin were


tried.
The specimens in their study were made by a hand lay-up process and then
compression moulded. The moulded panels had a thickness of about 1 mm. Tensile
tests were conducted at an elongation rate of 5.1 mm/min to determine KE and K.
From the measurements, KE had a range of 0.44 to 0.06, while K had a range
of 0.25 to 0. The average fiber efficiency factors were 0.19 for modulus and 0.11 for
strength.
Average values of 0.43 and 0.25 for KE and K, respectively were reported for
similar composites with unidirectional fiber orientation (Blumentritt, VU, & Cooper,
1974) They concluded that for similar materials, the fiber efficiency factor of
unidirectional fiber composites was approximately twice that of random-in-plane
composites.
3.7.9. Christensen and Waals Model
(Christensen & Waals, 1972) examined the behaviour of a composite system with
a three-dimensional random fiber orientation. Both fiber orientation and fiber-matrix
interaction effects were considered. For low fiber volume fractions, the modulus of the 3D composite was estimated to be

q*  + [1 + (1 + )]
6

Eq 3.7.9-1

where, c < 0.2 and m is the Poissons ratio of the matrix. c is the volume
concentration of the fiber phase, which is equivalent to the fiber volume fraction.


For a state of plane stress, the modulus is given as

 + [1 + ]
3 

Eq 3.7.9-2

where, c < 0.2.

Comparisons were made between the predictions given by Eq 3.7.9-2 and data
reported by (Lee, 1969). It was found that at low fiber volume fractions, predictions
from Eq 3.7.9-2 were within a range of 0 ~ +15% higher than the test data. The
difference between the prediction and test data was attributed to partially ineffective
bonds and/or end effects for the chopped fibers.

42
This document contains no EAR or ITAR technical data

3.7.10. Approximation Model by Manera


(Manera, 1977) proposed approximate equations to predict the elastic
properties of randomly oriented short fiber-glass composites. The invariant properties
of composites defined by (Tsai & Pagano, 1968) were used along with Pucks
micromechanics formulation (Manera, 1977). Manera made a few assumptions and
simplified Puck invariants equations. The assumptions included high fiber aspect ratio
(>300), two-dimensional random distribution of fibers and treatment of randomly
oriented discontinuous fiber composites as laminates with an infinite number layers
oriented in all directions. The approximate equations can be expressed as
16
8
 + 2 ` + 
45
9
2
3
1
^ = } _  +  ` + 
15
4
3
1
=
3
 = } _

Eq 3.7.10-1
Eq 3.7.10-2
Eq 3.7.10-3

where, Vf is the fiber volume fraction, m is the Poissons ratio of the matrix,
Em is the modulus of the matrix, Ef is the modulus of the fiber, ^E and ^G are the
tensile (flexural) and shear moduli of the composite, respectively and ^ is Poissons
ratio of the composite.
It can be seen from Eq.(2.22)-Eq.(2.24) that ^E, ^, and ^G satisfy the
relationship

^ =


2(1 + )

Eq 3.7.10-4

In order to get adequate precision in the results, Manera chose Vf to be within

the range 0.1 Vf 0.4 and Em within the range 2Gpa Em 4Gpa.
Predictions of composite modulus by Eq 3.7.10-1 were compared with test
data (Manera, 1977). The constituent properties of the composites in the tests were 5
cm chopped glass fiber with Ef=73Gpa, f =0.25 and polyester resin with Em=2.25Gpa
and m=0.40. The differences between the predictions and test data were less than 5%.

43
This document contains no EAR or ITAR technical data

3.8.Theories for Random Fiber Composites Based on the Calculated


Properties of Unidirectional Fiber-Reinforced Composites

3.8.1. Tsai and Pagano


The in-plane modulus of a random fiber composite was proposed by (Tsai & Pagano,
1968) to be
=

3
5
++ + NN
8
8

Eq 3.8.1-1

Where E11 and E22 are longitudinal and transverse modulus of unidirectional
composite obtained from Halpin Tsai equation (Section 3.7.7)
Although Eq 3.8.1-1 have very simple form, the predictions are only good at very low
fiber volume fractions. At high fiber volume fractions, the predicted modulus is much
higher than measured. (Blumentritt, VU, & Cooper, 1975) explained that this was
caused by the increase in concentration of defects within the composite as the fiber
content increases.

3.8.2.

Lavengood and Goettler

(Lavengood & Goettler, 1987) established a general procedure for predicting the
average Young's modulus for randomly oriented short fiber composites. When the
fibers are two dimensionally oriented, they derived the Reuss-type expression as:

Where,

 = 24++ NN /(7NN + 17++ )


++ =  + } (  )

NN =  [

2} (U 1) + (U + 2)
} (1 U)
+ (U + 2)}]

Eq 3.8.2-1

Eq 3.8.2-2
Eq 3.8.2-3

In which Em and Ef are Young's moduli of the matrix and fiber, respectively. Vf is the
volume fraction of the fiber; R is the ratio of transverse fiber modulus to matrix
modulus.

44
This document contains no EAR or ITAR technical data

3.8.3.

Piggot

(Piggott, 1980) suggested the modulus for composites having fibers which are random
in three dimensions as
{ = '( }  + } 
+

3.8.4.

Eq 3.8.3-1

Voigt and Reuss

The simplest models are those which make use of the rule of mixtures (combining
rules). Voigt assumed that each component was subject to the same strain (isostrain),
giving,

{ = }  + } 

Eq 3.8.4-1

Alternately, Reuss assumed that each phase was subject to the same stress (isostress),
giving,

{ =   /(}  + }  )

Eq 3.8.4-2

where E denotes modulus and V volume fraction, and the subscripts c, m and f
represent composite, matrix resin and fiber, respectively.

3.8.5.

Cox

(Cox, 1952) who used a shear lag formulation to model the longitudinal elastic
modulus showed that the modulus of short-fiber composites can be expressed as :
1
4
{ = ( )++ + ( )NN
5
5

Eq 3.8.5-1

where Ell and E22 are defined as E11=Ec from Voigt model ( Eq 3.8.4-1) and
E22=Ec from Reuss model (Eq 3.8.4-2)

45
This document contains no EAR or ITAR technical data

3.8.6.

Hori & Onogi

(Hori & Onogi, 1951) proposed the following:


{ = (++ NN )+/N

Eq 3.8.6-1

where Ell and E22 are defined as E11=Ec from Voigt model ( Eq 3.8.4-1) and
E22=Ec from Reuss model (Eq 3.8.4-2)

46
This document contains no EAR or ITAR technical data

3.9.Strength of Short fiber composites

3.9.1. Kelly and Tysons Model


(Kelly & Tyson, 1965) developed a theory to predict the strength of short fiber
composites. Basically, it is an extension of the rule of mixtures by taking into
account the effects of both the fiber length and fiber orientation. It was based on the
assumption that plastic flow will occur during stress transfer between matrix and
fibers, giving

{ =
 } _1

@{
` + }

2@

Eq 3.9.1-1

where c is ultimate tensile strength of composite f , m , strengths of fiber


and matrix, respectively and l, 1c, fiber length and critical length of
One problem associated with Kelly and Tysons theory is that the estimates of
strength are higher than measured (Peijs, Garkhail, Heijenrath, Oever, & Bos, 1998)

3.9.2. Piggot Model


(Piggott, 1980) accounted for both plastic and elastic effects in the matrix in
his fiber theory. Piggot's composite strength model is expressed by lengthy equations
that will not be presented here. For composites having fibers which are random in
three dimensions, he also suggested an upper strength bound critical length of fibers.
1

{ = _ `
 } + }

5

Eq 3.9.2-1

3.9.3. Rileys model


(Riley, 1968) considered interaction between fibers by taking into
consideration of the stress transfer between fibers in a rationalized fiber array such as
a hexagonal arrangement, and derived a strength equation as

 }
6

{ = _ `
+
(1 })
7 1 + '> (
>

47
This document contains no EAR or ITAR technical data

Eq 3.9.3-1

CHAPTER 3
4. METHODOLOGY
This project is carried out in following steps.

i.

All existing empirical solutions need to be compared with experimental results


from Literature for PEEK Carbon Fiber short fibre composite

ii.

The closest Empirical relation is chosen for further studies.

iii.

Using the chosen empirical solution the Elastic moduli for PEEK Carbon fiber
composite is calculated for low aspect ratio fiber.

iv.

A VBA code is created to generate block of composite with randomly oriented


fibers for fiber volume fraction up to 0.2.

v.

Discretization of the unit volume of SFC is executed in Patran and analysed to


pull and torsion in Msc Nastran

vi.

The results from empirical solution and FE is compared for Elastic modulus

vii.

If a close match is obtained then other elastic constants are also determined
using the same FE model.

viii.

As empirical relationships does not exist for randomly oriented short fiber
composite, comparison is done with results for aligned fiber composites

Comparison of Empirical relationships with Test results


This section details how the mechanical properties of random oriented shot
fiber composites can be determined and also compares the results between Tsai and
Pagano approximation and results from Finite Element Method.
Out of all the available empirical solutions a comparison is made with
experimental results to choose the best solution. PEEK as matrix and Carbon fiber
combination is chosen for the study. Test results are available for high aspect ratio
shot fiber composite. But the study in this report is about short fiber composite with
low aspect ratio.
48
This document contains no EAR or ITAR technical data

4.1.Stiffness Estimation

49
This document contains no EAR or ITAR technical data

50
This document contains no EAR or ITAR technical data

51
This document contains no EAR or ITAR technical data

52
This document contains no EAR or ITAR technical data

53
This document contains no EAR or ITAR technical data

Experimental vs Empirical (E)


1.90E+10

Modulus of elasticity

1.70E+10
1.50E+10

EXP
HO

1.30E+10

COX
Piggot

1.10E+10

L&G
T&P

9.00E+09
7.00E+09
0.1

0.12

0.14

0.16

0.18

0.2

Fiber Volume Fraction

Figure 4.1-1Comparison of Empirical models with experimental results

Out of all the above methods most of them closely match with the experimental
results as the fiber volume fraction considered in this study is on the lower side less
than 0.2. Still Tsai and Pagano model is the best suited for comparison with FE
modelling and subsequent stiffness estimation. This is for the reason that this model
considers aspect ratio also as a parameter. The FE modelling and analysis method
pursued in this study considers fibers of very low aspect ratio as well.

54
This document contains no EAR or ITAR technical data

4.2.Strength Estimation

55
This document contains no EAR or ITAR technical data

56
This document contains no EAR or ITAR technical data

4.3.Calculation of SFC stiffness for Low aspect ratio PEEK Carbon Fibre composite

57
This document contains no EAR or ITAR technical data

58
This document contains no EAR or ITAR technical data

4.4.Generating Random oriented short fibre composite stiffness from FE

MS Excel VBA code is developed to generate random locations and


orientations of the fiber. A unit cube of the composite is constructed using these fibers
in the matrix inside Msc Patran as Geometric model. This model is meshed and
converted into FE model inside Patran. Interface between fibers and Matrix is
modelled using Nastran Glued Contact. Mechanical properties of Fiber and Matrix are
separately fed into the FEM. This model is analysed in Nastran to obtain the
deformation and from which Mechanical properties are calculated.
The VBA code for generating random locations and orientations for fibers is shown
below.

59
This document contains no EAR or ITAR technical data

4.5.VBA Code for random fiber generation

Figure 4.5-1Flow chart for Random fiber generation VBA code

60
This document contains no EAR or ITAR technical data

Figure 4.5-2VBA code for Random number Generation I

61
This document contains no EAR or ITAR technical data

Figure 4.5-3VBA code for Random number Generation II

62
This document contains no EAR or ITAR technical data

The vba code is able to generate random short fiber composite specimens as shown in Figure
4.5-4 through Figure 4.5-6

Figure 4.5-4 A Sample random short fiber composite specimen generated by the VBA code

63
This document contains no EAR or ITAR technical data

Figure 4.5-5 A Sample random short fiber composite specimen generated by the VBA code

Figure 4.5-6 A Sample random short fiber composite specimen generated by the VBA code

64
This document contains no EAR or ITAR technical data

4.6.FE Creation and analysis for Fibre volume fraction of 0.2

The code generates a Patran session file which can be executed inside Patran to
generate a Geometric model as shown in the figure below

Figure 4.6-1Patran Session file for creating random fibers of volume fraction 0.2

Figure 4.6-2(a) Geometric model (b) FE model of random fibers of volume fraction 0.2

65
This document contains no EAR or ITAR technical data

4.7.FE Creation and analysis for Fibre volume fraction of 0.15

Figure 4.7-1Patran Session file for creating random fibers of volume fraction 0.15

Figure 4.7-2(a) Geometric model (b) FE model of random fibers of volume fraction 0.15

66
This document contains no EAR or ITAR technical data

4.8.FE Creation and analysis for Fibre volume fraction of 0.1

Figure 4.8-1Patran Session file for creating random fibers of volume fraction 0.1

Figure 4.8-2(a) Geometric model (b) FE model of random fibers of volume fraction 0.1

67
This document contains no EAR or ITAR technical data

4.9.Loads and Boundary Conditions

Figure 4.9-1Load application in the FE model

The geometric model is discretized using 10 noded tetrahedral elements for both fibre and
matrix. As it is difficult to maintain node to node connectivity between fibre and matrix
elements permanent glued contact is used to model the interaction between them. The
displacement and stress continuity of the model using this model is tested by applying the
same material for fiber and matrix. As shown in 3.6 it is established that the permanent glued
contact can be used to model the fiber matrix interaction.

Once the model is discretized nodes one end is fixed in three translations and an RBE2 is
created to connect all nodes on the opposite side rigidly. A 100 N load is applied on the
master node of the RBE2 in axial direction for tensile and 100 N-m moments is applied for
torsion about the axis.

The model is analysed using MD Nastran and the deflections and rotations are obtained from
the Nastran results at the load application point

68
This document contains no EAR or ITAR technical data

4.10.

Validation of Stress continuity in FE model

To validate the model and the contacts used to represent the fiber-matrix
interaction a check is done for stress/ displacement continuity for the the same FE
model of 0.2 fiber volume fraction but with Fiber and Matrix applied with the same
isotropic material. The Patran insight is used to get the stresses on different planes
inside the FE model. As shown in Figure 4.10-1 the stresses are almost constant
throughout the model away from constraint and load application area. Figure 4.10-2
shows that the variation across cross section is minimal and the slight variation owes
to the proximity of constraints or load application regions. It is also evident that there
is no discontinuity of stress at all between fibre and matrix.

Figure 4.10-1 Stress on model with same material applied for fiber and matrix

Figure 4.10-2 Stress on cross sections with same material applied for fiber and matrix

69
This document contains no EAR or ITAR technical data

4.11.

Calculation of Elastic Constants from FE Results

4.11.1. Modulus of Elasticity

70
This document contains no EAR or ITAR technical data

4.11.2. Modulus of Rigidity


Torsion is used to derive the modulus of rigidity of the composite specimen.
For shafts of uniform cross section the torsion is given by the equation
=
^=


^

Or

Eq 4.11.2-1




Eq 4.11.2-2

Where T = Applied Torsion


JT = Torsion Constant for the section, for square 2.25a4
a = half the side length
L = Length of the specimen
= Angle of twist in radians

71
This document contains no EAR or ITAR technical data

CHAPTER 4
5. RESULT ANALYSIS
5.1.Modulus of Elasticity

Calculations from Section 4.3 (Using Empirical relationship) and Section 0


(Using FE Model) is compared for PEEK and Carbon fiber random oriented shot fiber
composite with low aspect ratio for volume fractions ranging from 0.1 to 0.2,

FE Vs Empirical (E)
1.00E+10

Modulus of Elasticity

9.00E+09
8.00E+09
FE
7.00E+09

E11 Halpin Tsai


T&P

6.00E+09

E22 Halpin Tsai


5.00E+09
4.00E+09
0.1

0.12

0.14

0.16

0.18

0.2

Fiber Volume Fraction

Figure 5.1-1Comparison of Empirical model results for E with FE of random fibers

It can be seen from the graph that the FE results match closely with Empirical results for the
fiber volume fraction under consideration for random short fiber composites. This can also be
compared with the uniaxial short fiber composite properties shown by the dotted lines. As
expected the short fiber composite property lies in between the longitudinal and lateral
stiffness of aligned short fiber composite.

72
This document contains no EAR or ITAR technical data

5.2.Modulus of Rigidity

FE Vs Empirical (G)
2.40E+09
2.20E+09

Shear Modulus

2.00E+09

Manera

1.80E+09
FE
1.60E+09
Aligned
(HalpinTsai)

1.40E+09
1.20E+09
1.00E+09
0.1

0.12

0.14

0.16

0.18

0.2

Fiber Volume Fraction

Figure 5.2-1 Comparison of Empirical model results for G with FE of random fibers

There is only one empirical model in the literature which predicts the Shear modulus and that
is put forward by (Manera, 1977). The trend followed by FE results is slightly off from the
ones predicted by empirical method. Graph also shows a comparison with uni-axially aligned
fibers by Hapin Tsai shown in dashed line.

73
This document contains no EAR or ITAR technical data

5.3.Strength of Composite

Strength
1.20E+09
1.00E+09

Strength

8.00E+08
Exp

6.00E+08

KT
4.00E+08

Piggot
Riley

2.00E+08
0.00E+00
0.1

0.2

0.3

0.4

Fiber Volume Fraction

Figure 5.3-1Comparison of Empirical model results for Strength for random fibers

Determining strength by FE modelling is not possible as all failure modes of random


short fiber composites are not known and it is very difficult to simulate its failure in
FE analysis. All of the empirical models predict higher strength than shown by
experimental results. Piggot predicts the closest result which again deviates from the
test results at higher fiber volume fractions.

74
This document contains no EAR or ITAR technical data

CHAPTER 5
6. CONCLUSION AND FUTURE SCOPE OF WORK
6.1.Work conclusion

An extensive literature survey was done to identify the various empirical models
available for the stiffness and strength of randomly oriented short fiber composites
(SFC). A different method was adopted to obtain the mechanical properties of SFC by
FE modelling of random oriented fibers in matrix. The results obtained from these
were providing a decent comparison.
Work methodology followed was

Identify close form solution for strength and stiffness for chopped fiber composite
with aligned fibers

Identify close form solution for strength and stiffness for non-aligned fibers and to
understand what more is needed for random orientation.

Validate the properties using FEM for single fiber or certain randomly orientated
fiber combinations.
6.2.Future Work

The results obtained can be used to simulate the structural responses of randomly
oriented short fiber composite structures.
The VBA code can be further extended to generate composites with high aspect ratio
and high fiber volume fraction
Better contact modelling to replicate the fiber matrix interaction can be done
Complexities like curved fibers, fibers with variable length, fibers sticking to each
other etc., can be studied by improving the VBA code.
The methodology can be extended to particulate composites and metal matrix
composites
Testing to study the response of a structure with Random Short Fiber Composites.
75
This document contains no EAR or ITAR technical data

7. REFERENCES
[1] Ashton, J., Halpin, J., & Petit, P. (1969). Primer on Composite Materials: Analysis. Stamford Conn:
Technomic.
[2] Baxter, W. J. (1992). The strength of metal matrix composites reinforced with randomly oriented
discontinuous fibers. Metallurgical Transactions A, Vol. 23A, 3045-3053.
[3] Benveniste, Y. (1987). A new approach to the application of Mori-Tanaka's theory in composite
materials. Mech Mater, 147-157.
[4] Blumentritt, B. F., VU, B. T., & Cooper, S. L. (1974). Mechanical properties of Oriented
Discontinuous Fiber Reinforced Thermoplastics. I. Unidirectional Fiber Orientation. Polymer
Engineering and Science, Vol. 14, No. 9, 633-640.
[5] Blumentritt, B. F., VU, B. T., & Cooper, S. L. (1975). Mechanical properties of Discontinuous Fiber
Reinforced Thermoplastics. II. Random-in-Plane Fiber Orientation. Polymer Engineering and Science,
Vol. 15, No. 6, 428-436.
[6] Carman, G., & Reifsnider, K. (1992). Micromechanics of short-fiber composites. Compos Sci Technol,
137-46.
[7] Chen, P. E. (1971). Strength Properties of Discontinuous Fiber Composites. Polymer Engineering and
Science, Vol. 11, No. 1, 51-56.
[8] Chou, T.-W., Nomura, S., & Taya, M. (1980). A self-consistent approach to the elastic stiffness of
short-fiber composites. J Compos Mater, 178-88.
[9] Chow, T. (1977). Elastic moduli of filled polymers: The effect of particle shape. J Appl Phys, 40724075.
[10] Chow, T. (1978). Effect of particle shape at finite concentration on the elastic modulus of filled
polymers. J Polym Sci: Polym Phys Ed, 959-965.
[11] Christensen, R. M., & Waals, F. M. (1972). Effective Stiffness of Randomly Oriented Fiber
Composites. Journal of Composite Materials, Vol. 6, 518-532.
[12] Christensen, R., & Lo, K. (1979). Solutions for effective shear properties in three phase sphere and
cylinder models. J Mech Phys Solids, 315-330.
[13] Cox, H. (1952). The elasticity and strength of paper and other fibrous materials. Brit J Appl Phys, 72-9.
[14] Curtis, P. T., Bader, M. G., & Bailey, J. E. (1978). The stiffness and strength of a polyamide
thermoplastic reinforced with glass and carbon fibers. Journal of Material Science, Vol. 13, 377-390.
[15] Eduljee. (1994). The Influence of Aggregated and Dispersed Textures on the Elastic Properties of
Discontinuous-Fiber Composites. Compos Sci Technol, 381-91.
[16] Eduljee, R., McCullough, R., & Gillespie, J. (1994). The Influence of Aggregated and Dispersed
Textures on the Elastic Properties of Discontinuous-Fiber Composites. Compos Sci Technol, 381-91.
[17] Eshelby, J. (1959). The elastic field outside an ellipsoidal inclusion. Proc Roy Soc A.
[18] Eshelby, J. (1961). Elastic inclusions and inhomogeneities. Hill R.
[19] Ferrari, M. (1994). Composite homogenization via the equivalent poly inclusion approach. Compos
Engr, 37-45.

76
This document contains no EAR or ITAR technical data

[20] Fukuda, H., & Kawata, K. (1974). On Young's Modulus of Short Fibre Composites. Fibre Sci Tech,
207-22.
[21] Hahn, H. T. (1975). On Approximations for Strength of Random Fiber Composites. Journal of
Composite Materials, Vol. 9, 316-326.
[22] Halpin, J. (1969). Stiffness and Expansion Estimates for Oriented Short Fiber Composites. J Compos
Mater, 3:732-4.
[23] Halpin, J., & Kardos, J. (1976). The Halpin-Tsai Equations: A Review. Polym Eng Sci, 16:344-52.
[24] Hashin, Z. (1962). The Elastic Moduli of Heterogeneous Materials. ASME J Applied Mech, 29:143-50.
[25] Hashin, Z. (1965). On the Elastic Behaviour of Fibre Reinforced Materials of Arbitrary Transverse
Phase Geometry. J Mech Phys Solids, 13:119-34.
[26] Hashin, Z. (1983). Analysis of composite materials a survey. ASME J Applied Mech, 50:481-505.
[27] Hashin, Z., & Rosen, B. (1964). The Elastic Moduli of Fiber-Reinforced Materials. ASME J Applied
Mech, 223-32.
[28] Hashin, Z., & Shtrikman, S. (1963). A Variational Approach to the Theory of The Elastic Behavior of
Multiphase Materials. J Mech Phys Solids, 127-140.
[29] Hermans, J. (1967). The elastic properties of fiber reinforced materials when the fibers are aligned.
Proc Kon Ned Akad v Wetensch B, 1-9.
[30] Hewitt, R., & Malherbe, M. (1970). An Approximation for the Longitudinal Shear Modulus of
Continuous Fibre Composites. J Compos Mater, 4:280-2.
[31] Hill, R. (1964). Theory of Mechanical Properties of Fibre-Strengthened Materials: I Elastic Behaviour.
J Mech Phys Solids, 12:199-212.
[32] Hill, R. (1965). A self-consistent mechanics of composite materials. Mech Phys Solids, 213-222.
[33] Hori, M., & Onogi, S. (1951). Dynamic Measurements of Physial Properties of Pulp and Paper by
Audiofrequency Sound. J. Appl. Phys, 22. , No. 7, 971-977.
[34] Ingber, M., & Papathanasiou, T. (1997). A Parallel-Supercomputing Investigation of the Stiffness of
Aligned Short-Fiber-Reinforced Composites using the Boundary Element Method. Int J Num Meth
Engr, 40:3477-91.
[35] Kelly, A., & Tyson, W. R. (1965). Journal of Mechanics and Physics of Solids, Vol. 13, 329-336.
[36] Kerner, E. (1956). The elastic and thermo-elastic properties of composite media. Proc Phys Soc B, 808813.
[37] Kuriger, R. J., & Alam, M. K. (2001). Strength Prediction of Partially Aligned Discontinuous Fiberreinforced Composites. Journal of Material Response, Vol. 16, No.1, 226-232.
[38] L. e. (1997). Prediction of thermo-mechanical properties for compression-moulded composites.
Composites A, 63-70.
[39] Lavengood, R. F., & Goettler, L. A. (1987). Contract Report,ONR/ARPA Contract No.NOOO14-67 C0218.
[40] Lee, L. H. (1969). Strength-Composition Relationships of Random Short Glass Fiber-Thermoplastics
Composites. Polymer Engineering and Science, Vol. 9, 213-219.
[41] Lewis, T., & Nielsen, L. (1970). Dynamic Mechanical Properties of Particulate-Filled Composites. J
Appl Poly Sci, 14:1449-71.

77
This document contains no EAR or ITAR technical data

[42] Lin, S., & Mura, T. (1973). Elastic fields of inclusions in anisotropic media (II). Phys Stat Sol (a), 2815.
[43] Manera, M. (1977). Elastic properties of randomly oriented short fiber-glass composites. Journal of
Composite Materials, Vol. 11, 235-247.
[44] Mori, T., & Tanaka, K. (1973). Average stress in matrix and average elastic energy of materials with
misfitting inclusions. Acta Metallurgica, 21:571-4.
[45] Mura, T. (1982). Micromechanics of Defects in Solids. The Hague: Martinus Nijhoff.
[46] Nielsen, L. (1970). Generalized Equation for the Elastic Moduli of Composite Materials. J Appl Phys,
41:4626-7.
[47] Pan, N. (1996). The elastic constants of randomly oriented fiber composite: A new approach to
prediction. Science and Engineering of composite materials, Vol. 5, No.2, 63-72.
[48] Peijs, T., Garkhail, S., Heijenrath, R., Oever, M., & Bos, H. (1998). Thermoplastic composites based
on flax fibers and polypropylene: Influence of fiber length and fiber volume fraction on mechanical
properties. Macromol. Symp., 193-203.
[49] Piggott, M. R. (1980). Load Bearing Fiber Composites. Pergamon Press.
[50] Riley, V. R. (1968). Fiber/Fiber Interaction. J. Compo. Mater., 2, 436-446.
[51] Robinson, I., & Robinson, J. (1994). The influence of fibre aspect ratio on the deformation of
discontinous fibre-reinforced composites. J Mater Sci, 4663-77.
[52] Rosen, B. (1964). Tensile Failure of Fibrous Composites. AIAA J, 1985-91.
[53] Russel, W. (1973). On the eective moduli of composite materials:effect of ber length and geometry
at dilute concentrations. J Appl Math Phys (ZAMP), 581-600.
[54] Steif, P., & Hoysan, S. (1987). An energy method for calculating the stiffness of aligned short-fiber
composites. Mech Mater, 197-210.
[55] Tandon, G., & Weng, G. (1984). The eect of aspect ratio of inclusions on the elastic properties of
unidirectionally aligned composites. Polym Compos, 327-33.
[56] Taya, M., & Chou, T.-W. (1981). On two kinds of ellipsoidal inhomogeneities in an infinite elastic
body an application to a hybrid composite. Int J Solids Structures, 553-563.
[57] Taya, M., & Mura, T. (1981). On stiffness and strength of an aligned short fiber reinforced composite
containing fiber-end cracks under uni-axial applied stress. ASME J Applied Mech, 361-367.
[58] Thompson, J. L., Vlug, M. A., Schipper, G., & Krikort, H. (1996). Influence of fiber length and
concentration on the properties of glass fiber-reinforced polypropylene: Part 3. Strength and strain at
failure. Composites Part A, Vol. 27A, 1075-1084.
[59] Tsai, S. W., & Pagano, N. J. (1968). Invariant properties of composite materials. In Composite
Materials Workshop (pp. 233-238). Stamford, Conn: Technomic Publishing Co.
[60] Wakashima, K., Umekawa, S., & Otsuka, S. (1974). Thermal expansion of heterogeneous solids
containing aligned ellipsoidal inclusions. J Compos Mater, 391-404.
[61] Walpole, L. (1966). On bounds for the overall elastic moduli for inhomogeneous systems-I. J Mech
Phys Solids, 151-62.
[62] Walpole, L. (1966). On bounds for the overall elastic moduli for inhomogeneous systems-II. J Mech
Phys Solids, 289-301.

78
This document contains no EAR or ITAR technical data

[63] Walpole, L. (1969). On the overall elastic moduli of composite materials. J Mech Phys Solids, 235-51.
[64] Weng, G. (1984). Some elastic properties of reinforced solids, with special reference to isotropic ones
containing spherical inclusions. Int J Engng Sci, 845-856.
[65] Weng, G. (1992). Explicit Evaluation of Willis' Bounds with Ellipsoidal Inclusions. Int J Engng Sci,
83-92.
[66] Willis, J. (1977). Bounds and Self-Consistent Estimates for the Overall Properties of Anisotropic
Composites. J Mech Phys Solids, 185-202.
[67] Wu C, T., & McCullough, R. (1977). Constitutive Relationships for Heterogeneous Materials. In
Developments in Composites Materials-1 Applied Science (pp. 118-186). London: GS Holister.
[68] http://nptel.ac.in/

79
This document contains no EAR or ITAR technical data

Você também pode gostar