You are on page 1of 9

Materials Science and Engineering A 393 (2005) 170–178

On the mechanical properties of alumina–epoxy composites with an


interpenetrating network structure
M.T. Tilbrook, R.J. Moon, M. Hoffman∗
School of Materials Science and Engineering, University of New South Wales, Sydney, NSW 2052, Australia

Received 17 June 2004; received in revised form 4 October 2004; accepted 4 October 2004

Abstract

Methods of predicting effective mechanical properties of composites with an interpenetrating network structure are currently not well
understood, particularly for cases in which the constituent materials have widely differing properties. Alumina–epoxy composites with an
interpenetrating composite structure have been produced via an infiltration process and the elastic properties were measured via the impulse
excitation technique. A strong dependence of properties on composition and processing was observed. Properties were compared with several
mixing law predictions made using compositional data obtained from microstructural analysis. The effective medium approximation (EMA)
was shown to predict properties adequately, whilst others models proved inappropriate.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Interpenetrating network composite; Effective properties; Impulse excitation technique; Effective medium approximation

1. Introduction erated [1–5], whilst work on interpenetrating structures has


been more limited [8].
Composites potentially represent an advance in optimi- This issue is of particular interest for functionally graded
sation of material performance, by offering a combination materials (FGMs), which incorporate a spatial variation
of properties of several different constituent materials. of composition and properties [9–11], where experimental
Underpinning composite design is an understanding of the verification of local mechanical properties is difficult.
relation between mechanical properties of the constituent Furthermore, FGMs often possess an interpenetrating
materials and the effective mechanical properties of the network structure [12]. Recent interest in graded materials
resulting composite material. For a particular composite, has motivated the development of semi-empirical models,
this depends on the internal geometry and may be derived based on micromechanics and image analysis [13–15].
theoretically [1–5], though experimental verification of Several generic models may potentially be applica-
effective properties is crucial to accurately relate material ble to interpenetrating network-structured composites, in-
composition and properties. cluding the isostress- and isostrain-approximation models,
Composites with an interpenetrating network structure are Hashin–Shtrikman bounds, and the effective medium approx-
of particular interest as they have been found to demonstrate imation (EMA). In addition, several models have been pro-
higher strength, toughness and wear resistance when com- posed for composites with two continuous phases with simple
pared to other composite structures [6,7]. Traditionally, com- regular internal geometries [8,16]. For a more rigorous treat-
posite materials tended to display a fibre/matrix, layered, ment of mixing laws, see the work of Torquato et al. [17],
laminar or particle/matrix type structure. Correspondingly, Reiter et al. [11] or Suresh and Mortensen [10].
theoretical models based upon such geometries have prolif- The simple isostress and isostrain relations provide a first
approximation of the bounds for effective Young’s modulus
[17]. These are:
∗ Corresponding author. Tel.: +61 2 9385 4223; fax: +61 2 9385 5956.
E-mail address: mark.hoffman@unsw.edu.au (M. Hoffman). E∗ = v1 E1 + v2 E2 (1)

0921-5093/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2004.10.004
M.T. Tilbrook et al. / Materials Science and Engineering A 393 (2005) 170–178 171

 −1 The processing was conducted similarly to the method de-


∗ v1 v2
E = + (2) scribed by Cichocki et al. [23], which has been used to pro-
E1 E2
duce aluminium–alumina-interpenetrating composites [12].
Open-celled polyurethane (PU) foam (Bulpren S-31048, Eu-
where E is the Young’s modulus, (*) represents the effective rofoam, Troisdorf, Germany, 90 pores/in.) was used as an im-
composite property, v the volume fraction and subscripts 1 print for the interpenetrating network structure. It was com-
and 2 denote each phase. Replacing Young’s modulus, E, in pressed from an initial density of 2.5% to obtain the required
these equations with bulk and shear moduli, K and G, leads volume fractions of foam and air.
to the Voigt and Reuss model expressions. A colloidal suspension of alumina particles (99.99%
The Hashin–Shtrikman bounds [18] were used by Al2 O3 , Taimicron TM-DAR, Taimei Chemicals Co. Ltd.,
Torquato et al. to successfully predict the mechanical proper- Japan) was infiltrated into the PU foam under varying pres-
ties of interpenetrating boron carbide–aluminium composites sure, cast and allowed to dry. The foam was then pyrolysed at
[17]. The effective medium approximation (EMA) [4] was 800 ◦ C, leaving a ceramic green-body (approximate dimen-
used by Hoffman et al. to predict mechanical properties of sions: 50 mm × 30 mm × 6 mm) with a network of intercon-
aluminium–alumina-interpenetrating composites [19], al- necting pores. The ceramic was sintered at 1500 ◦ C for 1 h
though the validity of this was not confirmed experimentally. leading to ∼10% shrinkage in each dimension. Epoxy resin
Tuchinskii developed a simplified unit-cell model for com- (Epofix, Struers, Germany) was infiltrated into the ceramic
posites with two continuous phases [16]. This was observed under varying pressure then cured at room temperature. Spec-
by Jedamzik et al. [20] to fit experimental results for interpen- imens were then ground to size (50 mm × 30 mm × 5 mm)
etrating structured tungsten–copper composites. A similar with a 600 grit diamond-grinding wheel, and polished with
approach was applied by Feng et al. [21] to interpenetrating an automatic polisher and diamond paste down to a diamond
composites with more than two continuous phases. particle size of 1 ␮m in the final step. The plate specimens
This paper details the investigation of effective mechan- were later cut into small beams (50 mm × 5 mm × 4 mm ap-
ical properties of alumina–epoxy composites with an inter- prox.) with a low speed diamond saw for further testing. This
penetrating network structure. Experimental results obtained process results in two continuous interpenetrating phases; the
via the impulse excitation technique are presented and sev- interconnecting porosity of the foam perform is mimicked by
eral theoretical models are applied to this composite system. the alumina, which is a robust structure after firing, while the
The alumina–epoxy composite system has been adopted as a complete penetration of the epoxy with no discernable poros-
model system due to the significantly differing elastic proper- ity in the composite confirms the complete interpenetration
ties of the two phases. This enables a large stiffness variation of this phase.
to be attained across the composition range. Several samples of an alumina particulate/epoxy ma-
The large difference in properties of the two constituents trix composite in the composition range 75–90% were
leads to widely varying predictions from different mod- also produced, for comparison with the interpenetrating-
els. Considering for instance, a composite of 50% alumina network composites. Alumina powder (Taimicron TM-DAR,
(E = 400 GPa) and 50% epoxy (E = 4 GPa) [22], the predic- as above) was mixed with epoxy resin (Epofix) by stirring be-
tions from Eqs. (1) and (2) are 202 and 7.9 GPa, respectively. fore curing. Due to the viscosity of the mixture, air bubbles
These are clearly extremely disparate and a more precise pre- introduced during stirring were difficult to remove, which
diction is required. For a 50/50 composite of alumina and alu- lead to some porosity in these samples.
minium (E = 79 GPa), the predicted values, 240 and 132 GPa, Monolithic alumina and epoxy specimens were also pro-
respectively, differ less drastically though still significantly. duced for comparison. Alumina specimens were produced
An advantage of investigating an alumina–epoxy compos- from a suspension of alumina particles (Taimicron TM-
ite system is that the increased variation, between different DAR), which was slip-cast, dried and sintered at 1500 ◦ C
model predictions, enables improved identification of the ap- for 1 h. Epoxy specimens were produced from resin (Epofix)
propriate choice of model for composites with interpenetrat- cured at room temperature. Grinding and cutting was con-
ing network structure. ducted similarly to the composite specimens.

2.2. Microstructural analysis


2. Experimental techniques
Optical microscopy on the polished surfaces was used to
2.1. Sample processing and preparation characterise the alumina and epoxy phases and to quantify
the porosity. Images were obtained using a Nikon 200 micro-
Alumina–epoxy FGMs were produced by infiltration of scope and digital camera, and then processed using Adobe
epoxy into an open porosity alumina preform. Samples were Photoshop and Image Processing ToolKit (Version 3.0) [24].
produced in the composition range 5–50% epoxy, with an Examples of microstructures for a range of compositions,
interpenetrating network structure in which both phases are for both the interpenetrating-network and powder compos-
continuous. ites, are shown in Fig. 1. The continuity of each phase in the
172 M.T. Tilbrook et al. / Materials Science and Engineering A 393 (2005) 170–178

Fig. 1. Microstructural images of alumina epoxy composite samples. Interpenetrating network structures: (A) 5% epoxy; (B) 15% epoxy; (C) 30% epoxy; (D)
45% epoxy. Alumina appears as the light phase, epoxy is the grey phase and the darker spots are pores. Powder composites: (E) 75% epoxy and (F) 90% epoxy.
Alumina appears as the grey phase, epoxy is the light phase and the darker spots are pores.

interpenetrating microstructures cannot be observed in two- structured composites. For samples containing pores, mea-
dimensional micrographs. Fig. 1(a–d) show the interpene- surement of porosity volume fraction was incorporated into
trating network microstructures with the differences between the compositional analysis, by a second thresholding pro-
(a–b) and (c–d) attributable to the buckling of foam ligaments cess. Although a small amount of porosity was observed in
during compression which resulted in increasing microstruc- most samples, this was generally <2% and at most 5%. Error
tural irregularity for higher foam compression ratios in the margins for composition measurements were estimated via
resultant alumina phases (light regions). Fig. 1(e–f) shows perturbation during the thresholding step and from variation
a different morphology with some alumina particle contact in compositions observed across single samples.
(dark regions). After cutting and polishing, the densities were determined
Fig. 2 displays an original greyscale microscope image from mass and dimensional measurements. Composite densi-
and the corresponding black and white image obtained via ties were compared with those predicted using compositions
a digital thresholding process, which was used for volume obtained from image analysis and measured densities of con-
fraction estimation. The samples tested varied in composi- stituent phases, as shown in Fig. 3. The observed disparity
tion across the range 5–50% epoxy for the interpenetrating between the two density measurements, particularly for more
M.T. Tilbrook et al. / Materials Science and Engineering A 393 (2005) 170–178 173

Fig. 2. Microstructural image for 20% epoxy sample: (A) grey-scale image and (B) black and white image after thresholding.

Fig. 4. Comparison of theoretical alumina volume fractions calculated from


Fig. 3. Comparison of density values determined from dimensional
PU foam compression ratios with alumina volume fractions determined from
measurements and those calculated from microstructural data, for
microstructural analysis.
interpenetrating-network composites.

2.3. Elastic property characterisation


epoxy-rich samples, is most probably due to a higher volume
fraction of epoxy and pores near the surface where image For measuring elastic moduli, the impulse excitation tech-
analysis was undertaken. Several samples were progressively nique (IET) provides an alternative to tensile or flexural load-
ground down, polished and microanalysed at several depths, ing, either of which may damage brittle material samples.
and slightly higher alumina volume fractions were observed This technique involves lightly striking a rectangular-shaped
internally. For example, the density for the sample containing specimen to stimulate vibration in its natural modes [25–27].
∼40% epoxy was calculated, from surface microstructure, as Vibrational frequencies were measured with a micro-
2.55, and from internal microstructure, as 2.7. Similarly, the phone, amplifier and signal analyser (Analyzer 2000 V5+,
calculated densities for the sample containing ∼12% epoxy www.brownbear.de). Resonant frequencies were obtained for
were 3.5 at the surface and 3.65 internally. flexural and torsional modes, from which Young’s and shear
It should be noted that the compositions predicted from moduli, respectively were obtained. Samples were tested
the initial PU foam compression ratios differed significantly both as plates (50 mm × 30 mm × 5 mm approx.) and later as
from those realised experimentally, as shown in Fig. 4. The beams (50 mm × 5 mm × 4 mm approx.), to investigate the
assumption that final epoxy volume fraction is equal to ini- possibility of variation in properties across samples.
tial PU foam volume fraction appears to overly simplify the Values for Young’s modulus, E, were calculated from flex-
process of slip-casting and drying. It is possible that infiltra- ural frequencies, fflex , using the relation:
tion of the PU foam with alumina slip is incomplete due to
trapped air at vertices of foam ligaments, or that shrinkage mL3
E = 0.9465 T1 fflex
2
(3)
during drying leads to increased porosity in the ceramic body. bt 3
174 M.T. Tilbrook et al. / Materials Science and Engineering A 393 (2005) 170–178

Fig. 5. Young’s modulus values. Experimental results for interpenetrating- Fig. 6. Young’s modulus values. Experimental results for interpenetrating-
network composite () and powder composite (䊉) samples, and theoretical network composite () and powder composite (䊉) samples, computational
predictions: isostress (A); isostrain (B); Hashin–Shtrikman upper (C) and predictions () from Wegner and Gibson [8], and theoretical predictions:
lower (D) bounds. Tuchinskii unit-cell model upper and lower bounds, and effective medium
approximation with ψ = 1 and 6.
where L is the specimen length, b the width, t the thickness,
m the mass and T1 the geometry constant determined from b
and t [27]. Similarly, the relationship used to calculate shear
modulus, G, from the torsional frequency, ftors , was:
Lm B
G=4 f2 (4)
bt 1 + A tors
where A and B are geometry constants determined from L, b
and t [27]. Poisson’s ratio was subsequently calculated as:
E
ν= −1 (5)
2G
Error margins in modulus values were estimated from er-
rors in dimensional measurements and resonant frequency
readings. Errors in Poisson’s ratio results were relatively high
due to build-up of errors from Young’s modulus and shear
modulus calculations. Fig. 7. Shear modulus values. Experimental results for interpenetrating-
network composite samples, and EMA predictions (ψ = 6).

3. Results and discussion

3.1. Experimental results

Experimental results for Young’s modulus are shown plot-


ted against composition in Figs. 5 and 6, in comparison with
a number of mixing law predictions, which are discussed
shortly. The results for shear modulus and Poisson’s ratio are
shown in Figs. 7 and 8, respectively. Table 1 shows values ob-

Table 1
Experimentally obtained properties for alumina and epoxy
Property Alumina Epoxy
E (GPa) 390 3.4
Poisson’s ratio, ␯ 0.2 0.35
K (GPa) 280 3.8
Fig. 8. Poisson’s ratio values. Experimental results for interpenetrating-
G (GPa) 155 1.3
network composite samples, and EMA predictions (ψ = 6).
M.T. Tilbrook et al. / Materials Science and Engineering A 393 (2005) 170–178 175

density measurements and microstructural analysis, however,


this was difficult due to the disordered structure and residual
porosity, therefore, values are only approximate. The prop-
erties are apparently dominated by the epoxy matrix phase,
although the presence of alumina particles does increase stiff-
ness significantly.

3.2. Theoretical modelling predictions

The effective composite property models men-


tioned earlier have been applied to this alumina–epoxy
interpenetrating-structured composite system. The proper-
ties of alumina and epoxy given in Table 1 were used in all
calculations.
The isostress and isostrain bounds are shown in Fig. 5 as
curves A and B. Although these bounds do contain the experi-
Fig. 9. Variation of effective Young’s modulus with measured density. Ex- mental results, they differ widely, prohibiting their use for ac-
perimental results for interpenetrating-network composite and powder com-
curate predictions. The Hashin–Shtrikmann bounds [18] are
posite samples, and EMA predictions (ψ = 6).
shown in Fig. 5 as curves C and D. The results also lie within
these bounds; however, they provide only a slight improve-
tained for the monolithic alumina and epoxy samples. These ment on the isostress and isostrain bounds. The inadequacy of
are in agreement with results published elsewhere [22]. the Hashin–Shtrikman model is probably attributable to the
Fig. 9 shows a correlation between Young’s modulus and disparity in constituent properties, resulting in irregularities
density. It can also be seen that there is some variation in prop- in stress transfer at interfaces and stress distribution between
erties between beam samples cut from the same plate. This phases.
was attributed to compositional variation across the plates, The bounds predicted by Tuchinskii’s unit-cell model [16]
which was identifiable in the variation of measured density. are displayed in Fig. 6. Values predicted by this model are too
These results are of particular importance from a process- high. This may be attributed to the assumption of a simple
ing point-of-view and highlight the importance of ensuring regular cubic structure, which as shown in the microstructural
sample homogeneity. images, was not the case for these composites.
The possibility was considered that samples might display
anisotropic properties, as a result of the production process. If 3.3. Comparison with computational model
samples were anisotropic, their flexural rigidities in different
planes would differ, which would lead to differing Young’s Wegner and Gibson modelled interpenetrating network
modulus results for in-plane and out-of-plane measurements. structured two-phase composites using finite element anal-
As shown in Fig. 9, these results did not differ significantly. ysis [8]. They assumed one phase was comprised of slightly
This was corroborated by microanalysis of samples in planes overlapping spheres in a regular stacked array, whilst the other
perpendicular (Fig. 2), and parallel (Fig. 10), to the direction phase filled the interstitial volume. This compares well to
of foam-compression. No anisotropy was observed. the microstructures shown in Fig. 1 for lower epoxy volume
Results for the powder–matrix composite samples are in- fractions. They predicted a sharp drop-off in stiffness with
cluded in Figs. 5, 6 and 9. Compositions were estimated from increasing interstitial phase volume fraction, which agrees

Fig. 10. Representative microstructural images, on planes parallel to the direction of foam compaction for (A) 25% epoxy sample and (B) 10% epoxy specimen.
176 M.T. Tilbrook et al. / Materials Science and Engineering A 393 (2005) 170–178

with the experimental results as shown in Fig. 6. This is good By solving these self-consistently for K* and G* using
although such a computational model is time-consuming to MathCad, E* and ν* may be determined for spherical inclu-
produce and limited in applicability. sions (ψ = 1), as shown in Fig. 6. An iterative scheme was
used to calculate composite properties for non-spherical in-
3.4. Effective medium approximation clusions (ψ
= 1).
While it is possible to define different values of ψ for each
For a material with a random grain structure or a matrix- phase, resulting in different shape tensors, this was not pur-
inclusion topology with contacting inclusions, the effective sued in the current work, due to the uncertainty in selecting
medium approximation method may be appropriate [4]. This shape factor values. Rather, a single shape factor was used
technique, based on Hill’s self-consistent model [28], treats for both phases. As seen in Fig. 6, the closest fit was ob-
the composite material as being comprised of ellipsoidal in- tained with the EMA by altering the inclusion shape param-
clusions of each phase, surrounded by an effective medium eter, ψ, to a value of 6. This is not physically unreasonable
with the same properties as the total composite [29,30]. As as microstructural images indicated a fairly irregular inter-
well as composite elastic properties, this concept has been ap- nal geometry, particularly for the epoxy phase. Also, from
plied to effective electrical [31] and transport [32] properties considering the results in Fig. 5, the Poisson’s ratio of the
and to microcracked solids [33]. alumina was assumed to be 0.2, which is lower than values
The derivation, which is outlined here and given in full typically quoted in the literature [22]. This prediction is in-
by Kreher and Pompe [4], utilises the second-order tensor cluded in Figs. 6–9, and in each of these cases shows good
formulation for stress and strain. These are related by the agreement with measured results.
fourth-order compliance tensor: The success of the EMA in simulating elastic proper-
ties for this type of composite structure may be attributed
σ- = C- ε- (6) to its lack of inherent assumption regarding internal geom-
etry. Rather, inclusions are assumed to be surrounded by an
Tensors are denoted by an underline. Considering a single el-
effective medium. In the composites investigated, any partic-
lipsoidal inclusion with compliance, C- , in surrounding mate-
ular microstructural element of alumina or epoxy would be
rial with effective compliance, C- ∗ , and a prescribed far-field
in contact with both alumina and epoxy though not in any
strain, the local strain field is determined via the results of
regular configuration, and thus, is most accurately modelled
Eshelby [34].
via the effective medium concept. This is in agreement with
Applying this self-consistently to each phase of the com-
the conclusions of Christensen [35] and Munro [36].
posite results in a weighted mean for the effective compliance
It is possible that observed deviations from the predictions
tensor, C- ∗ , involving the compliance, C- (n) , and the inclusion
of the EMA could be explained by the observed variation in
shape-dependent strain relation tensor, T- (n) , for each phase:
microstructural morphology across the composition range.
(C- (n) − C- ∗ )T- (n) = 0 (7) Incorporating such effects into a model would be difficult,
however, and of questionable benefit, given the reasonably
where  denotes the mean over all phases. This expands to good fit obtained using the EMA in the formulation presented.
yield expressions that implicitly define the effective bulk and
shear moduli, K* and G* :

(Kn − K∗ )TK = 0, (Gn − G∗ )TG = 0


(n) (n)
(8) 4. Conclusions
(i) (i)
The strain relation tensor terms, TK and TG , are deter- The following conclusions were made:
mined from a tensor A - which is defined in terms of Eshelby’s
inclusion shape tensor P- and the shape functions f(ψ), where
ψ is the shape parameter describing the ellipsoid aspect ratio 1. The impulse excitation technique is a suitable method for
assumed for each phase (prolate for ψ < 1, spherical for ψ = 1, measuring mechanical properties of ceramic–epoxy com-
and oblate for ψ > 1). These tensors and functions are given posites.
in the appendix for completeness. For spherical inclusions, 2. The effective medium approximation provided the only
the strain relation tensors simplify considerably leading to accurate prediction of elastic properties of two-phase
the simplified equations: composites with an interpenetrating network structure.
    3. The Hashin–Shtrikman bounds and Tuchinskii model did
Kn − K∗ Gn − G∗
=0 =0 (9) not accurately predict elastic properties. This was at-
Kn + K̃∗ Gn + G̃∗ tributed to widely differing constituent properties and ir-
with the terms in the denominators given by: regularity in internal geometry.
4. Slight compositional variation resulted in significant vari-
4 ∗ ∗ 1 9K∗ + 8G∗ ation in mechanical properties, which highlighted the im-
K̃∗ = G G̃ = G∗ ∗ (10)
3 6 K + 2G∗ portance of careful processing.
M.T. Tilbrook et al. / Materials Science and Engineering A 393 (2005) 170–178 177

Appendix A where δ(α,β) is the Kronecker delta, equal to one for α = β,


and zero otherwise.The total magnitude are given by:
The full set of equations used in modelling composite
2 (n) (n) (n) (n) (n)
properties via the effective medium approximation technique (A(n) ) = A33 (A11 + A12 ) − 2A13 A31
is included here, for completeness. A more detailed deriva-
tion is given in [4] which follows the work of Eshelby [34] Inverting A (n)
- gives the corresponding strain relation tensor,
and others [28,32]. The shape functions, f1 and f2 , are defined T- :
(n)

in a piecewise manner, in terms of shape parameter, ψ: (n) −1


 T- (n) = (A )
   -
 −ψ 2 1 1
 cosh−1 − , 0 < ψ < 1 (prolate ellipsoids)
 2(1 − ψ2 )3/2 ψ 2(ψ2 − 1)

1
f1 (ψ) =  , ψ = 1 (spheres)
3
  
 ψ2
 −1 1 1
 2|1 − ψ2 |3/2 cos − , 1 < ψ (oblate ellipsoids)
ψ 2(ψ2 − 1)

 3f1 (ψ) − 1 The strain relation coefficients are then calculated as:
 , ψ
= 1
f2 (ψ) =  1 − ψ2 (n) 1 (n) (n) (n) (n) (n)
 0.2, TK = [A11 + 2A33 + A12 − 2(A13 + A31 )]
ψ=1 3(A(n) )
2

The shape polarisation tensor P- is given in condensed Voigt


notation as [34]: (n) 2 (n) (n) (n) (n) (n)
TG = 2
[A11 + 0.5A33 + A12 + A13 + A31 )]
3K∗ + G∗ 15(A(n) )
P11 =P22 = ((13 − 16ν∗ )f1 (ψ) − 3f2 (ψ))  
8G∗ (3K∗ + 4G∗ ) 1 2 1
+ (n) (n)
+ (n)
3K∗ + G∗
5 A11 − A12 A44
∗ ∗
P33 = (1 − 2ν − (1 − 4ν )f1 (ψ) − f2 (ψ))
G∗ (3K∗ + 4G∗ )
These calculations were done iteratively via MathCad for a
number of compositions. Initial guess values, K(0) and G(0) ,
−3K∗ − G∗
P12 = P21 = (f1 (ψ) + f2 (ψ)) were determined from Eq. (8), then the following iteration
8G∗ (3K∗ + 4G∗ ) scheme was utilised to obtain stable solutions:
−3K∗ − G∗ (n)
(Kn − K(i) )TK (K(i−1) , G(i−1) ) = 0
P13 =P31 =P23 = P32 = (f1 (ψ) − f2 (ψ))
2G∗ (3K∗ + 4G∗ )
(n)
(Gn − G(i) )TG (K(i−1) , G(i−1) ) = 0
3K∗ + G∗
P44 = P55 = where the strain relation coefficients were calculated using
2G∗ (3K∗ + 4G∗ )
the effective property values from the previous iteration.
×(1 − ν∗ − (2 − ν∗ )f1 (ψ) + f2 (ψ))

P66 = 0.5(P11 − P12 ) References


The summative terms are given as: [1] Z. Hashin, J. Appl. Mech. 50 (1983) 481.
3 [2] G.P. Sendeckyj, Elastic behaviour of composites, in: G.P. Sendeckyj
P1 = P2 = f1 (ψ), (Ed.), Mechanics of Composite Materials, vol. 2, Composite Mate-
3K + 4G∗

rials, Academic Press, 1974, p. 45.
3 [3] K.K.U. Stellbrink, Micromechanics of Composites: Composite Prop-
P3 = (1 − 2f1 (ψ)) erties of Fibre and Matrix Constituents, Hanser, 1996, p. 12.
3K∗ + 4G∗ [4] W. Kreher, W. Pompe, Internal Stresses in Heterogeneous Solids,
Akademie Verlag, 1989.
Tensor, P- (n) , for each phase is determined from the terms of
[5] S. Torquato, Appl. Mech. Rev. 44 (1991) 37;
the components of P- and the summative terms: S. Torquato, Int. J. Sol. Str. 35 (1998) 2385.
[6] H. Prielipp, M. Knechtel, N. Claussen, S.K. Streiffer, H. Müllejans,
Aαβ = δ(α, β) + 2(Gn − G∗ )Pαβ
(n)
M. Rühle, J. Rödel, Mater. Sci. Eng. A 197 (1995) 19.
[7] M. Sternitzke, M. Knetchel, M. Hoffman, E. Broszeit, J. Rödel, J.
+[Kn − K∗ − 23 (Gn − G∗ )]Pα , α, β = 1, 2, 3 Am. Ceram. Soc. 79 (1) (1996) 121.
[8] L.D. Wegner, L.J. Gibson, Int. J. Mech. Sci. 42 (2000) 925.
[9] A. Neubrand, J. Rödel, Zeit. f. Met. 88 (1997) 358.
A44 = 0.5 + 2(Gn − G∗ )P44
(n)
[10] S. Suresh, A. Mortensen, Int. Mater. Rev. 42 (3) (1997) 85.
178 M.T. Tilbrook et al. / Materials Science and Engineering A 393 (2005) 170–178

[11] T. Reiter, G.J. Dvorak, V. Tvergaard, J. Mech. Phys. Solids 45 (8) [23] F.R. Cichocki, K.P. Trumble, J. Rödel, J. Am. Ceram. Soc. 81 (6)
(1997) 1281. (1998) 1661.
[12] T.J. Chung, A. Neubrand, J. Rödel, T. Fett, Ceram. Trans. 114 (2001) [24] J.C. Russ, Image Analysis Handbook, third ed., CRC Press, Boca
789. Raton, FL, 1998.
[13] J.R. Cho, D.Y. Ha, Key, Eng. Mater. 183–187 (2000) 373. [25] ASTM Standard C1259-98. Standard Test Method for Dynamic
[14] P.-C. Zhai, Q.-J. Zhang, R.-Z. Yuan, Mater. Sci. Forum 308–311 Young’s Modulus, Shear Modulus and Poisson’s Ratio for Advanced
(1999) 995. Ceramics by Impulse Excitation of Vibration.
[15] M. Dong, P. Leßle, U. Weber, S. Schmauder, Mater. Sci. Forum [26] K. Heritage, C. Frisby, A. Wolfenden, Rev. Sci. Instr. 59 (6) (1998)
308–311 (1999) 1000. 973.
[16] L.I. Tuchinskii, Porosh Metall 7 (247) (1983) 85 (translated in: Pow- [27] R.R. Atri, K.S. Ravichandran, S.K. Jha, Mater. Sci. Eng. A 271
der Metallurgy and Metal Ceramics, Plenum, 1983). (1999) 150.
[17] S. Torquato, C.L.Y. Yeong, M.D. Rintoul, D.L. Milius, I.A. Aksay, [28] R. Hill, J. Mech. Phys. Solids 13 (1965) 213.
J. Am. Ceram. Soc. 82 (5) (1999) 1263. [29] B. Budiansky, J. Mech. Phys. Solids 13 (1965) 223.
[18] Z. Hashin, S. Shtrikman, J. Mech. Phys. Solids 11 (1963) 127. [30] L.J. Walpole, J. Mech. Phys. Solids 17 (1969) 235.
[19] M. Hoffman, S. Skirl, W. Pompe, J. Rödel, Acta Mater. 47 (2) (1999) [31] D. Stroud, Superlattices Microstruct. 23 (3–4) (1998) 567.
565. [32] J.R. Willis, J. Mech. Phys. Solids 25 (1977) 185.
[20] R. Jedamzik, A. Neubrand, J. Rödel, Mater. Sci. Forum 308–311 [33] Y. Huang, K.X. Hu, A. Chandra, J. Mech. Phys. Solids 42 (8) (1994)
(2000) 782. 1273.
[21] X.-Q. Feng, Z. Tian, Y.-H. Liu, S.-W. Yu, Appl. Comp. Mats. 11 [34] J.D. Eshelby, Proc. Roy. Soc. A 349 (1957) 376.
(2004) 33. [35] R.M. Christensen, J. Mech. Phys. Solids 38 (3) (1990) 379.
[22] M.F. Ashby, D.R.H. Jones, second ed., Engineering Materials 1: An [36] R.J. Munro, J. Am. Ceram. Soc. 84 (5) (2001) 1190.
Introduction to their Properties and Applications, vol. 2, Butterworth
Heinemann, 1996.

You might also like