Sie sind auf Seite 1von 23

CHAPTER - 4

COMPOSITE PROPERTIES - MICROMECHANICS

4.1 INTRODUCTION
4.2 UNIDIRECTIONAL COMPOSITES
4.2.1 Elastic Properties (Engineering Constants)
4.2.2 Strength Properties of Unidirectional Composites
4.2.3 Hygrothermal Properties
4.3 PARTICULATE AND SHORT FIBRE COMPOSITES
4.3.1 Particulate Composites
4.3.2 Short Fibre Composites
4.4 BIBLIOGRAPHY
4.5 EXERCISES

4.1 INTRODUCTION
The mechanical and hygrothermal properties of composites are of paramount
importance in the design and analysis of composite structures. The mechanical properties
constitute primarily the moduli and strength properties. The hygrothermal properties are
coefficient of expansion due to moisture (β), misture diffusion coefficient (d), coefficient of
thermal expansion (α), thermal conductivity (k) and heat capacity (c). Micromechanical
analyses concern with the theoretical prediction of these properties of constituent fibres and
matrices as well as several other parameters like the shape, size and distribution of fibres,
fibre misalignment, fibre-matrix interface properties, void content, fibre fracture, matrix
cracking and so on. The studies in micromechanics utilize micro-models, as the fibre
diameters usually vary in the microscopic scale between 5-140 �m. The micro-models
should simulate the microstructure of a realistic composite, but that usually makes the models
highly complex. The problems involving such complex models are normally tackled utilizing
advanced analytical methods as well as numerical analysis techniques(finite element and
finite difference methods). Even in the case of a complex model, a simplified idealization
with a reasonably good approximation of the real composite is desirable otherwise it may
lead to nowhere. It is not intended in this chapter to present the complete theoretical basis of
various micro-models used for the analytical prediction of all composite properties. The
presentation is limited to only a few simpler cases so as to acquaint the reader of the
background of the development in this area. Additional micromechanics relations for
unidirectional composites, that may find use in design applications, are listed in Table 4.1.
Typical properties of some of the common fibres and matrices are listed in Tables 4.2 and
4.3, respectively. The composite properties of a few composite systems derived using some
of the relations presented in this chapter are listed in Table 4.4. Tables 4.1 through 4.4 are
included at the end of this chapter.

4.2 UNIDIRECTIONAL COMPOSITES


4.2.1 Elastic Properties (Engineering Constants)
The stress-strain relation provides the basic interface between a material and a
structure. For a one dimensional isotropic, elastic body, the Hooke's law σ = E defines the
stress-strain behaviour. Here E is a material constant and is usually referred as elastic
constant (engineering constant) or Young's modulus. Besides E, the other conventional
engineering constant for a two-dimensional or three-dimensional isotropic body is Poisson's
ratio ν. The shear modulus G is not independent, but is related to E and ν as
G = E/2(1+ ν). A composite material is essentially heterogeneous in nature, therefore the
engineering constants, defined above, for an isotropic material are not valid. We consider
here a three-dimensional block of a unidirectional composite (Fig. 4.1), in which fibres are
aligned along the x'1 axis. The elastic behaviour for such a three-dimensional body

is orthotropic, and the engineering constants are , , (three Young's moduli along
three principal material axes x'1, x'2, x'3), ν'12, ν'13, ν'23, ν'21, ν'31, ν'32, (six Poisson's ratios) and
G'12, G'13, G'23, (three shear moduli). Of these, the first nine engineering constants i.e., three
Young`s moduli and six Poisson ratios are not independent. Due to symmetry of compliances
(see Eq. 6.18) these are related as given by

(4.1)
Note that,

(4.2)
Here ν'12 and ν'13 are usually referred as major Poisson ratios.
The 'mechanics of materials approach' provides convenient means to determine the
composite elastic properties. It is assumed that the composite is void free, the fibre-matrix
bond is perfect, the fibres are of uniform size and shape and are spaced regularly, and the
material behaviour is linear and elastic.
Consider a two-dimensional unidirectional lamina (Fig. 4.2), in which we define a
small volume element which represents not only the micro-level structural details but also the
overall behaviour of the composite. A simple representative volume element consists of an
isotropic fibre embedded in an isotropic matrix (Fig. 4.2b). This volume element is further
simplified as shown in Fig.4.2c, in which the fibre is assumed to have a rectangular cross-
section with the same thickness as the matrix. The width ratio is chosen to be the same as the
fibre volume fraction of the composite itself. The objective is to derive the composite
properties (E'11, E'22, ν'12, G12) in terms of the moduli, Poisson`s ratios and volume fractions
of the fibre and the matrix.

Longitudinal modulus , E'11


The micro-model (Fig. 4.2c) is subjected to a uniaxial tensile stress σ`11 as shown in
Fig. 4.3. It is assumed that plane sections remain plane after deformation. Hence,
'11= '11f = '11m = ΔL/L
and σ'11f = E'11f '11, σ'11m = E'11m '11, σ'11 = E'11 '11 (4.3)

Now, σ'11 W = σ'11f Wf + σ'11m Wm (4.4)

Substituting Eq. (4.3) into Eq. (4.4) and rearranging, we have


E'11 = E'11f Wf / W + E'11m Wm / W (4.5)

Noting that the volume fractions of the fibre and the matrix are
Vf = Wf /W and Vm = Wm / W respectively, Eq. 4.5 reduces to
E'11 = E'11f Vf + E'11m Vm (4.6)

Equation 4.6 defines the composite property as the 'weighted' sum of constituent properties
and is often termed as the 'rule of mixture'.

Transverse modulus, E'22


The tensile stress σ'22 is applied along the x'2 direction (Fig. 4.4) and the same is
assumed to act both on the fibre and the matrix. The strain on the fibre and the matrix are
'22 f = σ'22 / E'22f and '22m = σ'22 / E'22m (4.7)

also '22 = ΔW/W


and ΔW = '22f (Vf W) + '22 m (Vm W)

So, '22 = ΔW/W = '22 f Vf + '22 m Vm

or,

or,

or, (4.8)

Major Poisson 's Ratio, ν '12


The micro-model is stressed as in the case of determination of E`11 (Fig. 4.3). The
transverse contraction is noted as ΔW and is contributed by both the fibre and matrix. Thus,
ΔW = (ΔW)f + (ΔW)m
or, ΔW = W Vf ν '12f '11 + W Vm ν '12m '11 (4.9)

Now , (4.10)

and '22 = - ΔW/W (4.11)

Combining Eqs. 4.9 through 4.11, one obtains

or, ν '12 = Vf ν '12f + Vm ν '12 m (4.12)


Inplane Shear Modulus, G '12
The micro-model is now subjected to a shear stress σ '12 as shown in Fig. 4.5, and both
and the fibre and the matrix are assumed to experience the same shear stress.
'12f = σ '12 / G '12f and '12m = σ '12 / G '12m (4.13)

Now, Δ = '12W = W Vf '12f + W Vm '12m


or, '12 = Vf '12f + Vm '12m (4.14)

also, (4.15)
Substituting Eqs. 4.13 and 4.15 into Eq. 4.14 and eliminating σ '12 from both sides,
we get

or, (4.16)

Note that, for an isotropic fiber


E'11f = E'22f = Ef , ν'12f = νf

and (4.17)
and for an isotropic matrix
E'11m = E'22 m = Em , ν'12m = νm

and (4.18)

Equations 4.6 and 4.12 provide a reasonably accurate estimate of longitudinal modulus E�11
and ν�12, respectively. However, the transverse modulus E�22 and the shear modulus G�12,
estimated using Eqs. 4.8 and 4.16, are not so accurate mainly due to the reason that the
stresses in both the fibre and the matrix are assumed to be the same. The volume element
considered in the above mechanics of materials approach does not adequately represent the
micro structure of the composite. Advanced analytical methods employ better micro-models
along with the realistic material behaviour and boundry conditions. The analytical method
using a self-consistent field model provides a better estimation of composite properties in
comparison to the �mechanics of materials� approach. The model assumes the composite
to be a concentric cylinder (Fig. 4.6) in which a transversely isotropic matrix. Although the
assumed micro-model is simple, it permits formulation of the problem based on the theory of
elasticity so that it is possible to achieve the stress and strain variations in a realistic manner,
and the relations for the effective composite properties are then derived. These properties are
expressed as follows:

(4.19)

(4.20
)

(4.21)

(4.22)

(4.23)

where K' is the plane strain bulk modulus.

(4.24)
in which K', G'23, ν'122 and E'11are defined in Eqs. 4.19 through 4.23.

(4.25)

with E'11, E'22, K', and ν'12 defined in the above relations.

Note that for isotropic fibres and matrices,

and (4.26)

4.2.2 Strength Properties of Unidirectional Composites


The strength of a material is defined as the level of stress at which failure occurs. The
strength is a material constant. Most of the isotropic structural materials possess only one
constant i.e., the uniaxial tensile strength. The shear strength is normally related to the tensile
strength. A brittle isotropic material may have different strength values in tension and
compression and may be termed as a two-constant material. In contrast, a composite is a
multi-constant material. Referring to Fig. 4.1, it may be stated that a unidirectional composite
may possess three normal strengths X'11, X'22, X'33 and three shear strengths X'12 , X'13, X'23. A
normal strength may have different values in tension and compression, as the compressive
force usually induces premature failure due to buckling of fibres which have extremely high
slenderness ratio. So there are a total of nine independent strength constants X'11t, X'22t, X'33t,
X'11c, X'22c, X'33c, X'12, X'13, X'23.
Attempts made using micromechanical analyses to determine these strength constants,
met with little success. This is primarily due to the reason that the micro-models used in these
analyses are grossly unrealistic. In fact, it is extremely difficult to simulate the realistic
composite, as the initial microstructure changes continuously with the increase of applied
stress and propagation of failure in the form of fibre fracture, matrix cracking, fibre-matrix
debond and so on at several points located randomly within the composite. The brittleness of
the fibre and the matrix aggravates the situation. This is illustrated in Fig. 4.7. Note that lc is
the ineffective length. The presence of a single surface flaw in a brittle fibre causes the fibre
to fracture at A (Fig. 4.7a). This induces high shear stresses and causes the fibre-matrix
debond along the fibre direction (Fig. 4.7b). Also when a fibre fractures, a redistribution of
stresses in the vicinity results in the tensile fracture of the adjacent fibre due to stress
concentration. This process leads to the propagation of the crack in the direction transverse to
the propagation of the crack in the direction transverse to the fibres (Fig. 4.7c). In fact, the
final failure of a composite is resulted due to the cumulative damage caused by several micro
and macro-level failures.

Longitudinal Tensile Strength, X'11t


A simple relation can be derived for the longitudinal tensile composite strength X'11t
using the 'rule of mixtures' and is expressed as
X'11t = X'11f Vf + X'11m Vm (4.27)
Here it is assumed that, at a particular level of stress, all fibres fracture at the same time and
the failure occurs in the same plane. That this idealization is grossly unrealistic has already
been argued in the preceding paragraph.
Now, let us examine the validity of Eq. 4.27 for two composite systems: (i) a
carbon/epoxy composite, in which the fibre failure strain is less than the matrix failure strain,
i.e., '11fu < '11mu (Fig. 4.8a) and (ii) carbon/carbon composite when '11fu > '11mu
(Fig.4.8b). In these cases, both fibres and matrices are brittle. In the case of carbon/epoxy
composite, when Vf is much higher than Vm , the strength of the composite is primarily
controlled by the fibre fracture. Once the fibres fail, very little resistance is offered by the
matrix. So, the strength of the composite is given by
X'11t = X'11f V 'f + σ '11m Vm (4.28)
where σ '11m is the stress level in the matrix when the fibres fracture. On the other hand, when
Vf is low, there is a sufficient amount of matrix to resist the load after the failure of fibres. In
that case,
X'11t = X'11mt Vm (4.29)
It is therefore obvious that there exists a limiting value of Vf at which the final failure changes
from the fibre failure mode to the matrix failure mode. One may argue in a similar way to
identify the possible failure mechanisms in the case of a carbon/carbon composite also (Fig.
4.8b) as well as in the cases of other composites in which either fibres or matrices or both are
ductile. But the fact remains that there is no single relation which is able to define the
uniaxial tensile strength of a realistic composite.
However, Rosen's model of cumulative damage, which is based on the Weibull
distribution of the strength-length relationship, provides somewhat better estimation of X'11t,
when the fibres and the matrix exhibit brittle behaviour. This model assumes that the
composite consists of N fibres of original length L and the weaker fibres fracture due to the
applied tensile stress (Fig. 4.9). The original length is then divided into M segments, where
each segment (bundle or link) is of length 1c. Thus the composite forms a chain of M bundles
(links). When the number of fibres are very large (high Vf) the strength of each bundle or
chain link assumes the same value, i.e., the strength of the composite becomes equal to the
link strength. This is expressed as

(4.30)
where α and β are material constants and can be determined experimentally. The tensile
strength of the composite is then determined using

(4.31)
Note that lc is called the ineffective length or critical fibre length and is determined using the
shear lag stress. It is given by

(4.32)
where X'f is the tensile fracture strength of the fibre, d is the diameter of the fibre and Xi is the
fibre-matrix interfacial shear strength.
The longitudinal compressive strength X'11c of a unidirectional composite is primarily
affected by the buckling of fibres. In a simplified model, the fibres are treated as isotropic
thin plates lying in the x'1 x'2 plane (Fig. 4.10) and are supported on an isotrpic elastic
medium (matrix). Fibres may buckle in two modes-extension and shear. In
the extension mode, the matrix along the length of the fibre experiences alternate expansion
and contraction, whereas the matrix is subjected to shearing deformation in the shear mode.
The compressive strength is then determined employing the strain energy method. For the
extensional mode,

(4.33)

or, (4.34)

and for the shear mode

(4.35)
The transverse strength properties normally depend on the matrix properties. The
transverse tensile strength X'22t may also depend on the fibre-matrix interface strength, as
illustrated in Fig. 4.11. The experimental data for some composites confirm that the
transverse tensile strength enhances with the improvement in the fibre-matrix interface bond.
The actual fracture path, however, is a mixture of fibre-matrix debond, fibre splitting and
matrix cracking. A realistic model should be based on the variation of statistical data for all
these failure modes. Two simple relations, for the prediction of the transverse tensile strength
X'22t and transverse compressive strength X'22c of a unidirectional composite, are presented as
follows:

(4.36)

(4.37)
These relations assume that the transverse strength of a composite primarily depends on the
strength of the matrix.

4.2.3 Hygrothermal Properties


Transport Properties
The evaluation of transport properties like moisture diffusivity, heat conductivity,
electric conductivity, dielectric constant and magnetic permeability of a unidirectional
composite follows the similar procedure when one uses a self-consistent field model. The
resulting relations are, therefore, identical for all transport properties. The procedure is,
hence, illustrated considering only one case � the diffusion of moisture through a
unidirectional composite. Consider the concentric cylindrical model as shown in Fig. 4.6.
Both the fibre and the matrix are assumed to be moisture permeable. For example, aramid
fibres and polymer matrices are moisture permeable. For the diffusion of moisture along the
fibre direction (x'1 axis), the moisture diffusion equation assumes the form

(4.38)

where C is the moisture concentration per unit volume, t is time and d'11 is the longitudinal
moisture diffusion coefficient of the composite.

Steady state condition


Equation 4.38 takes the form

(4.39)

Assuming the boundary conditions (Fig. 4.6) to be as at


x'1 = 0, C = 0 and x'1 = L, C = C0 (4.40)

the solution is derived as

(4.41)

that satisfies Eqs. 4.39 and 4.40.


The direction of moisture diffusion per unit area parallel to the x'1 direction is defined
as

(4.42)

The total rate of moisture diffusing through the cross-section of the concentric
cylinder is given by

(4.43)

Note that

and (4.44)
where d'11f and d'11m are the longitudinal moisture diffusivities for the fibre and the matrix,
respectively. Substituting Eqs. 4.42 and 4.44 in Eq. 4.43 and noting that
Rf2/R2 = Vf and (R2 - Rf2) / R2 = Vm one obtains
d'11 = d '11f Vf + d'11m Vm (4.45)

when the fibres (e.g., glass, carbon, etc.) are impermeable to moisture
d'11 = d '11m Vm (4.46)

The transverse moisture diffusion coefficient d '22 can also be determined using a
similar self-consistent field model and is, given as

(4.47)

When fibres are impermeable to moisture, Eq. 4.47 reduces to

(4.48)

The longitudinal and transverse thermal conductivities k'11 and k'22 of the
unidirectional composite can be determined by replacing 'd ' with 'k' in Eqs.4.45 and 4.47,
respectively. Note that, in that case, heat conduction takes place both through the fibre and
the matrix. The other transport properties can also be derived in a similar way using Eqs. 4.45
and 4.47.

Expansional Strains
The longitudinal expansional strains (due to temperature or moisture) of a
unidirectional composite can be determined using the simple 'mechanics of materials
approach ' as discussed earlier. Consider the micro-model in Fig. 4.3. The total longitudinal
strains, after accounting for the mechanical strain and the expansional strain, are given as

and also and also (4.49)


Solving Eqs. (4.49) one gets

and (4.50)

Assuming free expansion '11 = '11e, the first relation of Eqs. 4.50 yields
σ '11 = 0 (4.51)

Therefore,
σ '11 W = σ '11f Wf + σ '11m Wm = 0

or, (4.
52)

Dividing Eq. (4.52) by W and noting that and and


Vm = Wm / W one obtains

(4.53)

Observing that the thermal expansional strain of a specimen of length L due to a rise
of temperature ΔT is given by e = LαΔT, the longitudinal thermal expansion coefficient α'11
of a unidirectional composite is derived from Eq. 4.53 as follows:

(4.54)

Similarly, the longitudinal moisture expansion coefficient β'11 of a unidirectional composite is


obtained from Eq. 4.54 replacing 'α' by 'β'.
For the transverse expansional strain '22e, the 'self-consistent field model ' approach
is, however, preferred. The expression for '22e can be derived as
(4.55)

The transverse thermal expansion coefficient α'22 is then derived from Eq, 4.55 in a similar
way

(4.56)

The transverse moisture expansion coefficient β'22 is obtained from Eq. (4.56) by replacing 'α'
with 'β'.

4.3 PARTICULATE AND SHORT FIBRE COMPOSITES


A unidirectional composite provides some sort of regularity in the microstructure, as
the fibres are continuous and aligned in one direction. This helps to assure a simple micro-
model with a constant strain or stress field and use the 'mechanics of materials ' approach to
determine the composite properties. Such a simple analytical treatment with constant stress or
constant strain field is not adequate in the case of particulate and short fibre composites. The
microstructure is not uniform through the composite medium. The point to point variation of
the microstructure is quite significant in many situations due to wide variations in the shape,
size and properties of fillers and reinforcements and their orientation and distribution in the
matrix phase. The discontinuous nature of some of these reinforcements adds to more
complexities. There exist innumerable high stress zones around irregular shaped particulate
reinforcements and at the tips of short fibres. The assumption of constant stress and strain
fields is no more valid. Further complications arise due to the anisotropy caused by the
alignment of short fibres and flake particulates.
All these preclude a general treatment of the problem. A single composite micro-
model, in no way can represent all composites of this category. Composites with different
reinforcements may require different micro-models and analytical treatments. This is
probably the main reason why the micromechanics analysis of this class of composites has
not received much attention from researchers. There is also another important reason for the
dearth of information in the area. In comparison to particulate and short fibre composites,
unidirectional composites find extensive uses in structural components in several engineering
disciplines. This has created more awareness and, in turn contributed to the growth of
knowledge in the micromechanics of unidirectional composites, while the understanding of
the micromechanical behaviour of particulate and short fibre composite still continues to
remain at its nascent stage.
4.3.1 Particulate Composites
The simplest mechanics of materials approach uses classical Voigt (constant strain)
and Reuss (constant stress) models to estimate the elastic properties for an isotropic
composite. With the Voigt model, the bulk modulus k and the shear modulus G are given as
P= Vf Pf + Vm Pm ,
where P=K,G
and E = 9 KG / (3K+G)
ν = (3K-2G) / (6K+2G) (4.58)

and with the Reuss model, the relations are

(4.59)

The properties predicted by Voigt model (highest) and Reuss model (lowest) are two
extremes to the real values. Several improved analytical models are known to exist, but are
not easily amenable to simple design uses. The Halpin-Tsai model, which is based on a semi-
empirical approach, is popular and provides both upper and lower bounds that fall within the
Voigt and Reuss limits. Simple relations that are developed based on an improved combining
rule are found to provide a reasonably good estimate of the properties of an isotropic
composite (Pf > Pm and 0 < νf < 0.5).
These are presented as follows:

(4.60)

with P = K,G.

For bulk modulus,

and for shear modulus, G: (4.61)


Young 's modulus E and Poisson 's ratio ν are then determined from Eqs.4.60 and 4.61
using Eqs.4.58.
The thermal expansion coefficient α is given by

(4.62)

where K is obtained using Eqs.4.60 and 4.61.

4.3.2 Short Fibre Composites


A simple model assumes a randomly oriented short fibre composite as a quasi-
isotropic micro-laminate in which each lamina consists of a group of short fibres oriented
along a particular direction. P� is determined using the modified Halpin-Tsai relation as
given by

(4.63)
where for
E '11, longitudinal modulus, ξ = 2l / D
E '22, transverse modulus, ξ = 2
G '12, inplane shear modulus, ξ =1

G '23, transverse shear modulus, ξ = (3 � 4 νm)-1


Note, that l and d are the length and the diameter of the short fibre, respectively. Both the
matrix and the fibre are isotropic in nature. The Poisson 's ratio ν'12 is estimated using the
simple mixture rule.
The longitudinal tensile strength is dependent on the critical fibre length lc (Eq. 4.32)
and is given by

where σ 'm is the stress on the matrix when the fibre breaks.

Table 4.1 Additional micromechanics relations for unidirectional composites


1. Volume fractions:
Vf + Vm + Vv =1 (1)
For a void free composite, Vv = 0; Vf + Vm = 1
2. Mass fractions
Mf + Mm = 1 (2)
3. Void volume fraction ρ
Vv = 1 - ρ[ (Mf / ρf ) + (Mm / ρm) ] (3)
4. Composite density
ρ= ρf Vf + ρmVm (4)
5. Fibre volume fraction

(5)
6. Matrix volume fraction

(6)
7. Transverse modulus

(7)
8. Shear moduli

(8a)

(8b)
9. Poisson's ratio

(9)
10. Longitudinal compression strength

(fibre crushing) (10a)

(microbuckling) (10b)
(10c)

11. Transverse thermal conductivity

(11a)

(11b)

where Kmf = 2(Km/Kf -1) for a cylindrical fibre.

(for MMCs) (11c)

12. Transverse moisture diffusivity

(12a)

(12b)

where dmf =2 (dm /df -1) for a cylindrical fibre.


13. Thermal expansion coefficients

(13a)
(13b)

(13c)

(for MMCs)
(13d)

14. Transverse moisture expansion coefficients

(14)

15. Heat capacity

(15)

Table 4.2: Typical properties of some common fibres

S. Property Boron Carbon(T300) Kelvar- S- E- Rayon


N0. 49 Glass Glass (T50)

1. Fibre diameter, d �m 140 8 12 9 9 8


2. Density, ρf gm /cm3 2.63 1.77 1.47 2.49 2.49 1.94
3. Longitudinal Modulus,
E'11f GPa 400 220 150 85 75 380
4. Transverse Modulus,
E'22f GPa 400 14 4.2 85 75 6.2
5. Longitudinal Shear 170 9 2.9 36 30 7.6
modulus, G'12f GPa
6. Transverse Shear
modulus, G'23f GPa 170 4.6 1.5 36 30 4.8
7. Longitudinal Poisson's
ratio, ν'12f 0.2 0.2 0.35 0.2 0.2 0.2
8. Transverse Poisson's
ratio, ν'23f 0.2 0.25 0.35 0.2 0.22 0.25
9. Heat capacity,
Cf kJ/(kg k) 1.30 0.92 1.05 0.71 0.71 0.84
10. Longitudinal Heat
conductivity k'11f
W/(mk) 38.0 1003.0 2.94 36.30 13.0 1003.0
11. Transverse Heat
conductivity k'22f
W/(mk) 38.0 100.3 2.94 36.30 13.0 100.3
12. Longitudinal thermal
Expansion coefficient,
α'11f 10-6 m/m/K 5.0 1.0 -4.0 5.0 5.0 7.7
13. Transverse thermal
expansion coefficient,
α'22f 10-6 m/m/K 5.0 10.1 54 5.0 5.0 10.1
S. Property Boron Carbon(T300) Kelvar- S- E- Rayon
N0. 49 Glass Glass (T50)
14. Longitudinal
compressive strength,
X'11ft MPa 4140 2415 2760 4140 2760 1730
15. Longitudinal
compressive strength,
X'11fc MPa 4830 1800 500 3450 2400 1380
16. Shear strength, X'12f
MPa 700 550 400 1050 690 350

Table 4.3 : Typical properties of some common matrices

S Property Poly- Epoxy Phe- Poly- Nylon 6061 Nickel Titan-


No. imide nolic ester Al ium

1. Density,ρm gm
/cm3 1.22 1.3 1.2 1.2 1.14 2.8 8.9 4.4

2. Young's
Modulus, Em
GPa 3.45 3.45 11 3 3.45 70 210 110

3. Shear
Modulus, Gm
GPa 1.28 1.28 4.07 1.11 1.28 26.12 81.40 44

4. Poisson's
ratio, νm 0.35 0.35 0.35 0.35 0.35 0.34 0.29 0.25

5. Heat capacity,
Cm kJ/(kgk) 1.05 0.96 1.30 1.15 1.67 0.96 0.46 0.39

6. Heat
conductivity,
km W/(mk) 2.16 0.18 0.21 0.25 0.19 171 62.0 7.0

7. Thermal
expansion
coefficient,
αm10-6 m/m/k 36.0 64.3 80.0 80.0 46.0 23.4 13.3 9.5

8. Moisture
diffusivity,
dm10-13 m2/s 0.39 1.637 1.20 1.80 1.10 0.0 0.0 0.0

9. Moisture
expansion
coefficient, βm
m/m/C 0.33 0.38 0.38 0.50 0.45 0.0 0.0 0.0

10. Tensile
strength,
Xtm MPa 120 90 60 60 81.4 310 760 1170

11. Compressive
210 130 200 140 60.7 310 760 1170
strength, Xcm
MPa

12. Shear
strength,
Xsm Mpa 90 60 80 50 66.2 180 440 675

Table 4.4 : Thermoelastic properties of three unidirectional composites (Vf = 0.6)

S. Property Kelvar/ T300/ Boron/ Fomulae


No Epoxy Epoxy polyimide used

1. Density, ρ gm /cm3 1.40 1.58 2.07 Eq.4*

2. Longitudinal modulus, E'11 GPa 91.38 133.38 241.38 Eq.4.6*

3. Transverse modulus, E'22 GPa 4.00 8.29 14.87 Eq.7*

4. Poisson 's ratio, ν'12 = ν'13 0.35 0.26 0.26 Eq.4.12*

5. Poisson 's ratio, ν'23 0.484 0.424 0.394 Eq.9*

6. Inplane shear modulus, G'12


2.26 3.81 5.53 Eq.8a*
= G'13 GPa

7. Transverse shear modulus, G '23


1.44 2.90 5.53 Eq.8b*
GPa

8. Longitudinal conductivity,
1.836 601.87 23.66 Eq.4.45*
k'11 W/ (mk)

9. Transverse conductivity,
0.57 0.72 6.95 Eq.11b*
k'22 W/(mk)

10. Heat capacity, c kJ/(kgk) 1.017 0.933 5.28 Eq.15*

11. Longitudinal thermal expansion


-2.48 1.99 5.28 Eq.4.54*
coefficient, α'11 (x10-6) m/m/k

12. Transverse thermal expansion


61.32 2.73 1.48 Eq.13b*
coefficient, α'22 (x10-6) m/m/k

* Eqs. Of Table 4.1

Das könnte Ihnen auch gefallen