You are on page 1of 9

Available online at www.sciencedirect.

com
Available online at www.sciencedirect.com
ScienceDirect
ScienceDirect
Structural
Available Integrity
online Procedia 00 (2019) 000–000
at www.sciencedirect.com
Structural Integrity Procedia 00 (2019) 000–000 www.elsevier.com/locate/procedia
www.elsevier.com/locate/procedia
ScienceDirect
Procedia Structural Integrity 28 (2020) 1892–1900

1st Virtual European Conference on Fracture


1st Virtual European Conference on Fracture
Modeling elastic properties of unidirectional composite materials
Modeling elastic properties of unidirectional composite materials
using Ansys Material Designer
using Ansys Material Designer
Saiaf Bin Rayhanaa*, Md Mazedur Rahmanbb
Saiaf Bin Rayhan *, Md Mazedur Rahman
a
School of Aeronautics, Northwestern Polytechnical University, Xi’an – 710072, China.
b
School
a of Mechanical Engineering, Northwestern Polytechnical University, Xi’an-710072, China.
School of Aeronautics, Northwestern Polytechnical University, Xi’an – 710072, China.
b
School of Mechanical Engineering, Northwestern Polytechnical University, Xi’an-710072, China.

Abstract
Abstract
Advanced reinforced composites are leading materials used in high tech industries (automotive, aerospace, naval, etc.) in recent
Advanced
years. Beforereinforced composites
manufacturing are leading
any structure madematerials used inmaterials,
by composite high techit industries (automotive,
is crucial to analyze theiraerospace, naval, etc.)
elastic properties in recent
by analytical
years.
meansBefore manufacturing
or computational any structure to
homogenization made by composite
reduce the overallmaterials, it is crucial
manufacturing to analyze
cost and time fortheir elastic
faster propertiesInby
production. analytical
this current
means
researchorpaper,
computational homogenization
Ansys Material Designer to is reduce
adoptedthe to overall
estimatemanufacturing cost and of
the elastic properties time for faster production.
unidirectional composite In this current
materials and
research paper,with
compare them Ansys Material
available Designer
analytical, is adopted
finite elementtoand estimate the elastic
experimental properties
results. of unidirectional
In general, it yields resultscomposite materials
with excellent and
accuracy
compare
comparedthem with available
to experimental analytical,
data finite
for all RVE element
shapes and experimental
(square, diamond andresults. In general,
hexagonal). it yields
However, resultsloss
accuracy with
is excellent
found foraccuracy
in-plane
compared to experimental
shear modulus (G12) when data for all
the fiber RVE shapes
volume fraction(square,
is morediamond
than 0.6.and hexagonal). However, accuracy loss is found for in-plane
shear modulus (G12) when the fiber volume fraction is more than 0.6.
© 2020 The Authors. Published by ELSEVIER B.V.
© 2020 The Authors. Published by Elsevier B.V.
© 2020
This The
is an Authors.
open accessPublished by ELSEVIER
article under B.V.
the CC BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0)
This is an open access article under the CC BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0)
This is an
Peer-review open access
under article under
responsibility of the CC BY-NC-ND
European Structurallicense
Peer-review under responsibility of the European Structural Integrity (https://creativecommons.org/licenses/by-nc-nd/4.0)
Integrity
Society Society (ESIS) ExCo
(ESIS) ExCo
Peer-review under responsibility of the European Structural Integrity Society (ESIS) ExCo
Keywords: Micromechanics; Effective Elastic Properties; Computational Homogenization; Ansys Material Designer
Keywords: Micromechanics; Effective Elastic Properties; Computational Homogenization; Ansys Material Designer

1. Introduction
1. Introduction
Over the past 40 years, advanced reinforced composite materials have been widely adopted in various industries,
Over the
ranging frompast 40 years,
sporting advanced
goods reinforcedtocomposite
manufacturers aerospacematerials
designinghave been widely
companies. Theadopted
main in various industries,
motivation of using
ranging from sporting goods manufacturers to aerospace designing companies. The main motivation of using

* Corresponding author. Tel.: +86-13080933760


* Corresponding rayhan.saiaf@mail.nwpu.edu.cn
E-mail address:author. Tel.: +86-13080933760
E-mail address: rayhan.saiaf@mail.nwpu.edu.cn
2452-3216 © 2020 The Authors. Published by ELSEVIER B.V.
This is an open
2452-3216 access
© 2020 Thearticle under
Authors. the CC BY-NC-ND
Published by ELSEVIER license
B.V.(https://creativecommons.org/licenses/by-nc-nd/4.0)
Peer-review
This under
is an open responsibility
access of the European
article under Structural
CC BY-NC-ND Integrity
license Society (ESIS) ExCo
(https://creativecommons.org/licenses/by-nc-nd/4.0)
Peer-review under responsibility of the European Structural Integrity Society (ESIS) ExCo

2452-3216 © 2020 The Authors. Published by Elsevier B.V.


This is an open access article under the CC BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0)
Peer-review under responsibility of the European Structural Integrity Society (ESIS) ExCo
10.1016/j.prostr.2020.11.012
Saiaf Bin Rayhan et al. / Procedia Structural Integrity 28 (2020) 1892–1900 1893
2 Author name / Structural Integrity Procedia 00 (2019) 000–000

composites is weight reduction, corrosion resistance, part-count reduction, enhanced fatigue life, and wear resistance,
to name a few.
In composite micromechanics, homogenization is the first step towards the design and analysis of composite
structures (Yan (2003)). The main aim of homogenization is to establish the macroscopic behavior of composites
which are microscopically heterogeneous (Cioranescu and Donato (2010)) in order to describe the stiffness presented
by five elastic properties namely E11 (modulus of elasticity in the fiber direction), E22 (modulus of elasticity in the
direction transverse to the fibers), G12 (in-plane shear modulus), G23 (out of plane shear modulus) and υ12 (in-plane
Poisson’s ratio) (Barbero (2011)). Dating back from the 19th century, many researchers have proposed analytical,
semi-analytical and computational homogenization models based on representative volume element (RVE) to evaluate
the elastic properties of unidirectional composite materials (Saeb et al. (2016)). The simplest law to quickly estimate
the four elastic constants (E11, E22, υ12, and G12) is the rule of mixture (ROM), which is mainly established by Voigt
(1889) and Reuss (1929). However, since the predicted values of E22 and G12 adopting ROM disaccord with the
experiments, the modified rule of mixture (MROM) can be an alternative for better estimation (Younes et al. (2012)).
A similar attempt to correct the values of E22 and G12 is found by Halpin-Tsai equations (Affdl and Kardos (1976)).
However, the main limitation of using ROM, MROM and Halpin-Tsai models is that they cannot predict the value of
G23.
To formulate all five independent elastic properties, a well-accepted micromechanical model is the Chamis model
which uses the same formulation of ROM to estimate E11 and υ12; however, different approaches for other elastic
moduli E22, G12 and G23 (Sendeckyj et al. (1989)). Other widely used constitutive laws are elasticity approach model
(EAM) (Hashin and Rosen (1964); Christensen (1990)), self-consistent model (S-C) (Hershey (1954); Budiansky and
Wu (1962); Hill (1965)) and Mori-Tanaka model (M-T) (Mori and Tanaka (1973); Benveniste (1987); Luo and Weng
(1987)). More recently, a new analytical approach, termed as bridging micromechanics model was established based
on bridging matrix to predict both the stiffness (Huang (2001a)) and strength (Huang (2001b)) of unidirectional
composite materials.
Since the development of computer simulation programs, the finite element method (FEM) has been widely used
by researchers to estimate the elastic properties of composite materials (Alexander and Tzeng (1997); Michel et al.
(1999); Sun et al. (2001); Andreassen and Andreassen (2014); Otero et al. (2015)). FEM can be an effective tool to
predict macroscopic properties of pultruded composite lamina (Xin et al. (2019)). For more complex problems, like
the determination of effective properties of nonlinear composite materials reinforced with particles of different shapes,
a hybrid M-T/FEM method can be used with simplified geometry (Ogierman (2019)). In recent times, the virtual
element method (VEM) has gained popularity for computational materials homogenization, which is a powerful
generalization of FEM, capable of solving polygonal mesh elements including non-convex or highly distorted elements
(Lo et al. (2020)). Other than FEM, meshless methods such as Natural Neighbour Radial Point Interpolation Method
(NNRPIM) can be used efficiently to predict better results than FEM with an expense of high computation time
(Rodrigues et al. (2018)).
Even though the numerical methods are widely used to predict the mechanical properties of composite materials,
defining the geometry of RVE and boundary conditions can be a challenging and time-consuming task [Saeb et al.
(2016); Younes et al. (2012)). To overcome these problems, a preloaded geometric definition of RVEs and boundary
conditions can be employed via Ansys Material Designer to estimate the stiffness of composite lamina (Ansys Inc.
(2018)). The main objective of this current research paper is to compute the elastic moduli of the unidirectional
composite lamina (carbon/epoxy and polyethylene/epoxy) and compare them with available predictions of
analytical/semi-analytical models, namely, I. Chamis Model, II. EAM Model, III. Bridging Model, IV. M-T Model
and V. S-C Model; Computational model, namely, VI. Comsol Multiphysics FEA & VII. Experimental data found in
the literature (Younes et al. (2012)) to evaluate the efficiency and reliability of the FEM technique adopted by Ansys
Material Designer. To the best of the authors’ knowledge, there is no previously published work presenting the
modeling of elastic properties based on Ansys Material Designer.
1894 Saiaf Bin Rayhan et al. / Procedia Structural Integrity 28 (2020) 1892–1900
Author name / Structural Integrity Procedia 00 (2019) 000–000 3

2. Finite Element Modeling Technique

For a periodic unit cell, 6 load cases are considered: 3 tensile tests (X, Y, and Z) and 3 shear tests (XY, YZ, and
ZX). A corresponding macroscopic strain is applied in each case, and reaction forces in the boundary faces of the
RVE are used to assemble the stiffness matrix. Then the engineering constants are extracted as follows.

  x   D11 D12 D13 0 0 0   x 


    
  y   D21 D22 D23 0 0 0   y 
  z   D31 D32 D33 0 0 0   z 
    (1)
  xy   0 0 0 D44 0 0   xy 
   0 0 0 0 D55 0   yz 
 yz    
  zx   0 0 0 0 0 D66 
  zx 

If the strain in the X-direction is fixed to  x  0.01 and all other strains are set to zero, the first column of the stiffness
matrix is obtained as follows.

 D11   x 
   
D
 21  y 
 D31  1  z 
    (2)
 0  0.01   xy 
 0   
   yz 
 0    zx 

Assuming the RVE occupies the volume 0, Lx   0, Ly   0, Lz  , on the faces normal to the X-axis, enforce as follows.

 Lx , y, z  ux 0, y, z    Lx
ux 
u y  Lx , y , z   u y  0, y, z  (3)
uz  Lx , y , z   uz  0, y, z 

Similarly, on the faces normal to the Y-axis, enforce as follows.

ux  x, Ly , z   ux  x,0, z 
u y  x, Ly , z   u y  x,0, z                                                                    (4) 
uz  x, Ly , z   uz  x,0, z 

 
Besides, for Z-Axis, enforce as follows.

ux  x, y, Lz   ux  x, y,0 
u y  x, y , Lz   u y  x, y,0  (5)
uz  x, y, Lz   uz  x, y,0 

In addition to these periodicity conditions, rigid body motions must also be prevented. This is done by enforcing

ux ( a po int with x  0)  0
u y (a po int with y  0)  0 (6)
uz (a po int with z  0)  0

To compute macroscopic stresses, the forces on the top faces are integrated. Let us consider  x . The force in the X-
Saiaf Bin Rayhan et al. / Procedia Structural Integrity 28 (2020) 1892–1900 1895
4 Author name / Structural Integrity Procedia 00 (2019) 000–000

direction at the face x  Lx is integrated.  x is obtained by normalizing with the face area.  y and  y are obtained
similarly. Then the entries for D11 , D21 and D31 in the stiffness matrix are obtained. By repeating the steps for all other
load cases, all the entries for the stiffness matrix are obtained. The stiffness matrix is inverted to obtain the compliance
matrix as follows.

C    D 1  (7)

Finally, the engineering constants are computed from the following relationship.

 1  yx  zx 
 0 0 0 
 xE E y E z 
  1  zy 
 xy
0 0 0 
 Ex Ey Ez 
 

 xz  yz 1 
 E 0 0 0 
Ey Ez
 
C 
 x
 (8)
 1 
 0 0 0 0 0 
 G xy 
 1 
 0 0 0 0 0 
 G yz 
 1 
 0 0 0 0 0 
 Gxz 

3. Material Definition

To calculate the stiffness of composite materials, elastic constants of fiber and matrix must be defined. It is
generally assumed that fibers are orthotropic where the matrix is isotropic. To compare the stiffness results, the
following material properties are adopted from the literature (Younes et al. (2012)), Table 1 and 2.

Table 1. Elastic moduli of fibers [7]

Fiber material E11 , GPa E22 , GPa υ12 υ23 G12, GPa G23, GPa

Carbon 232 15 0.279 0.49 24 5.034

Polyethylene 60.4 4.68 0.38 0.55 1.65 1.51

Table 2. Elastic moduli of matrix [7]

Matrix material E, GPa υ

Epoxy (With Carbon) 5.35 0.354


Epoxy (With Polyethylene) 5.5 0.37

4. RVE geometry and boundary condition

The first step of homogenization is the modeling of RVE. For unidirectional composite materials, three different
RVE can be selected to calculate the stiffness of the material (Fig. 1). Then a blocked mesh is generated and a periodic
boundary condition is employed by default (Ansys Inc. (2018)). In case the mesh is not periodic, non-periodic
1896 Saiaf Bin Rayhan et al. / Procedia Structural Integrity 28 (2020) 1892–1900
Author name / Structural Integrity Procedia 00 (2019) 000–000 5

boundary conditions can also be employed to calculate the mechanical properties. Finally, the RVE is exposed to
several load cases and the responses are computed accordingly to obtain homogenized material data.

                A.  B.                            C. 

     

                          

Fig. 1. RVE geometries with block meshing A: Square shape; B: Diamond shape; C: Hexagonal shape.

5. Result analysis and discussion

5.1. Longitudinal Young’s Modulus, E11

At first, we will examine the longitudinal Young’s Modulus (E11) data for different fiber volume fraction, obtained
from Ansys and compare them with available analytical, experimental and finite element outcomes (Younes et al.
(2012)). It is evident that, for both Carbon-Epoxy, Fig. 2 (A) and Polyethylene-Epoxy, Fig. 2 (B), Ansys Material
Designer’s prediction agrees quite well with other outcomes and the estimated values are identical for all RVE shapes.

  250
                                          
65
A. B.
  200 52
 
  150 39
E11,GPa

 
E11,GPa

  100
Exp. 26
Exp.
Chamis/Bridging
  Chamis/Bridging
EAM EAM
  50
M-T
13
M-T
S-C S-C
  Comsol FE. (Sq./Dm./Hex.) Comsol FE. (Sq./Dm./Hex.)
  0
Ansys FE. (Sq./Dm./Hex.) Ansys FE. (Sq./Dm./Hex.)
0
  0 0,2 0,4 0,6 0,8 1 0 0,2 0,4 0,6 0,8 1
Fiber Volume Fraction,Vf Fiber Volume Fraction, Vf

Figure 2. A: Longitudinal Young’s Modulus (E11), Carbon-Epoxy; B: Longitudinal Young’s Modulus (E11), Polyethylene-Epoxy

5.2. Transversal Young’s Modulus, E22

Unlike the previous case study, for transversal Young’s Modulus, E22, disagreements are found among analytical,
experimental and numerical predictions, Fig. 3 (A) and 3 (B). More interestingly, RVE shapes of Material Designer
predict slightly different outcomes (maximum difference found 5.14%) for the Carbon-Epoxy case, Fig. 3 (A).
However, this is not true for Polyethylene-Epoxy where the results are identical, Fig. 3 (B). For the Carbon-Epoxy
case, among the three different RVE shapes, square-shaped RVE of Material Designer provides excellent agreement
with experimental data with a maximum difference of 1.85% found for 0.65 fiber volume fraction.
For Polyethylene-Epoxy, in general, Material Designer estimates better results than Comsol FE and other analytical
solutions in comparison to experiments, except the Chamis model, which shows slightly better predictions for this
specific case. However, the highest discrepancy of estimated data by Material Designer with the experiment is 5.85%
for 0.68 fiber volume fraction.
6 Author name
Saiaf Bin / Structural
Rayhan et al. /Integrity
ProcediaProcedia
Structural00Integrity
(2019) 000–000
28 (2020) 1892–1900 1897

  21
A. 6,5
                                            B.
  19 6,3
  17 6,1
  15
5,9
  5,7
E22,GPa

E22,GPa
13
  5,5
  11 5,3
  9 5,1
  4,9
7
  4,7
  5 4,5
0 0,2 0,4 0,6 0,8 1 0 0,2 0,4 0,6 0,8 1
 
  Fiber Volume Fraction, Vf Fiber Volume Fraction, Vf
Exp. Chamis EAM Exp. Chamis
  Bridging M-T S-C EAM Bridging
  Comsol FE. Sq. Comsol FE. Dm. Comsol FE. Hex. M-T S-C
Ansys FE. Sq. Ansys FE. Dm. Ansys FE. Hex.
  Comsol FE Sq./Dm./Hex. Ansys FE. Sq./Dm./Hex.

Figure 3. A: Transversal Young’s Modulus (E22), Carbon-Epoxy; B: Transversal Young’s Modulus (E22), Polyethylene-Epoxy 

5.3. In-plane shear modulus, G12

For the calculation of in-plane shear modulus (G12), Ansys Material Designer yields good agreement with the
experiments along with Comsol FE predictions and bridging analytical model, Fig. 4 (A) and Fig. 4 (B). For the
Carbon-Epoxy case, three RVEs used by the Material Designer to predict the results only differ when the fiber volume
fraction is 0.7. For the same case, square-shaped RVE loses around 13% accuracy with experiments. For the
Polyethylene-Epoxy case, outcomes due to different RVE shapes do not vary at all and the maximum difference with
the experiment was found to be 4% when the fiber volume fraction reaches 0.6 and this trend continues for further
fiber volume fraction cases.
  25                              
2,1
A. B.
  20
2
 
1,9
  15
 
G12,GPa
G12,GPa

1,8
 
10
  1,7
 
5
  1,6
 
  0 1,5
0 0,2 0,4 0,6 0,8 1 0 0,2 0,4 0,6 0,8 1
 
Fiber Volume Fraction, Vf Fiber Volume Fraction, Vf
Exp. Chamis EAM Exp. Chamis
Bridging M-T S-C EAM Bridging
Comsol FE. Sq. Comsol FE. Dm. Comsol FE. Hex. M-T S-C
Ansys FE. Sq. Ansys FE. Dm. Ansys FE. Hex. Comsol FE. (Sq./Dm./Hex.) Ansys FE. (Sq./Dm./Hex.)

Figure 4. A: In-plane shear modulus (G12), Carbon-Epoxy; B: In-plane shear modulus (G12), Polyethylene-Epoxy 
1898 Saiaf Bin Rayhan et al. / Procedia Structural Integrity 28 (2020) 1892–1900
Author name / Structural Integrity Procedia 00 (2019) 000–000 7

5.4. Out of plane shear modulus, G23

In comparison with the experimental data, for both Carbon-Epoxy, Fig. 5 (A) and Polyethylene-Epoxy, Fig. 5 (B)
case studies, Ansys slightly overestimates out of plane shear modulus (G23) results. In the case of Carbon-Epoxy
prediction, the difference between the experimental results never exceeds more than 5%. Nonetheless, an acceptable
agreement with the experiment is also achieved for the case of Polyethylene-Epoxy where the maximum discrepancy
is found to be 7% against all the calculated fiber volume fraction.

  5,5 2,1
A. B.
2
 
4,5
  1,9

G23,GPa
G23,GPa

  3,5 1,8
 
1,7
  2,5
  1,6

  1,5 1,5
  0 0,2 0,4 0,6 0,8 1 0 0,2 0,4 0,6 0,8 1
  Fiber Volume Fraction, Vf Fiber Volume Fraction, Vf
Exp. Chamis
  Exp.
Bridging
Chamis
M-T
EAM
S-C EAM Bridging
  Comsol FE. Sq. Comsol FE. Dm. Comsol FE. Hex. M-T
Comsol FE. (Sq./Dm./Hex.)
S-C
Ansys FE. (Sq./Dm./Hex.)
Ansys FE. Sq. Ansys FE. Dm. Ansys FE. Hex.

Figure 5. A: Out of plane shear modulus (G23), Carbon-Epoxy; B: Out of plane shear modulus (G23), Polyethylene-Epoxy 

5.5. In-plane Poisson’s ratio, υ12

Even though the utilized literature data only comprise in-plane Poisson’s ratio (υ12) for Carbon-Epoxy case without
experimental outcomes, the predicted data of Ansys FE Material Designer is nearly identical with all the analytical
models and Comsol FE results, Fig. 6 (A). Besides, the Polyethylene-Epoxy data was calculated by analytical means
and compared with Material Designer data, Fig 6 (B). An excellent agreement was achieved further among the models.

  0,37 0,395
In-plane Poisson's Ratio,υ12

In-plane Poisson's ratio, υ12

A. 0,39 B.
  0,345
0,385
 
0,32 0,38
 
  0,375
0,295
  0,37
  0,27 0,365
  0 0,2 0,4 0,6 0,8 1 0 0,2 0,4 0,6 0,8 1
  Fiber volume fraction, vf
Fiber Volume Fraction, vf
Chamis Model/Bridging Model
Chamis/Bridging M-T Ansys FE (Sq./Dm./Hex.)
S-C Comsol FE. (Sq./Dm./Hex.) M-T
Ansys FE. (Sq./Dm./Hex.) S-C

Figure 6. A: In-plane Poisson’s ratio (υ12), Carbon-Epoxy; B: In-plane Poisson’s ratio (υ12), Polyethylene-Epoxy 
Saiaf Bin Rayhan et al. / Procedia Structural Integrity 28 (2020) 1892–1900 1899
8 Author name / Structural Integrity Procedia 00 (2019) 000–000

5.6. Computation time

Finally, we have measured the estimated time required by Ansys Material Designer to solve one problem. Even
though CPU and RAM play a significant role in the time required to solve each problem, a general conclusion can be
found in Table 3.

Table 3. Computation time of all the elastic properties by Material Designer

Computer Type CPU RAM, GB Computation Time, s

Laptop Core i-7,5th Gen. 4 40

Desktop AMD Ryzen 3700x 16 22

6. Conclusion

The finite element tools are efficient enough to predict the homogenized properties of unidirectional composite
materials. Ansys Material Designer provides a very robust solution under one minute to calculate the stiffness of any
unidirectional composite materials. From the results section, it is revealed that except for the Bridging model, other
analytical methods cannot estimate all the mechanical properties with good accuracy and in most cases, Ansys FE
values are identical with the Bridging model. On the other hand, for most of the instances, specialized Material
Designer FE code predicts better outcomes than Comsol FE. In terms of computation time, Ansys performs well in
providing a solution in less than one minute in case preloaded RVEs are used. However, it is also found that the
Material Designer starts to lose accuracy while predicting in-plane shear modulus (G12) values with experiments, when
the fiber volume fraction of unidirectional composite is more than 0.6. In summary, it can be concluded that Material
Designer is a powerful and efficient tool to calculate the stiffness of any unidirectional composites. In the future, we
wish to continue our research on woven composite, which is more complex in terms of computation and geometry.

Acknowledgement

The authors are grateful to China Scholarship Council (CSC) to finance their research at Northwestern
Polytechnical University, China (CSC grant No. GXZ023506).

References

Affdl, J. and Kardos, J., 1976. The Halpin-Tsai equations: A review. Polymer Engineering and Science, 16(5), pp.344-352.
Alexander, A. and Tzeng, J., 1997. Three Dimensional Effective Properties of Composite Materials for Finite Element Applications. Journal of
Composite Materials, 31(5), pp.466-485.
Andreassen, E. and Andreasen, C., 2014. How to determine composite material properties using numerical homogenization. Computational
Materials Science, 83, pp.488-495.
Ansys Material Designer User’s Guide, Release 19.2, 2018. Ansys, Inc.
Barbero, E., 2011. Introduction to Composite Materials Design. 2nd ed. Boca Raton: Taylor & Francis, pp.91-105.
Benveniste, Y., 1987. A new approach to the application of Mori-Tanaka's theory in composite materials. Mechanics of Materials, 6(2), pp.147-
157.
Budiansky, B. and Wu, T., 1962. Theoretical prediction of plastic strains in polycrystals. In Proceedings of the 4th U.S. National Congress
Theoretical Applied Mechanics, pp. 1175.
Christensen, R., 1990. A critical evaluation for a class of micro-mechanics models. Journal of the Mechanics and Physics of Solids, 38(3), pp.379-
404.
Cioranescu, D. and Donato, P., 2010. An Introduction to Homogenization. Oxford: Oxford University Press. London, UK.
Hashin, Z. and Rosen, B., 1964. The Elastic Moduli of Fiber-Reinforced Materials. Journal of Applied Mechanics, 31(2), pp.223-232.
Hershey, A., 1954. The elasticity of an isotropic aggregate of anisotropic cubic crystals. ASME Journal of Applied Mechanics, 21, 236–240.
Hill, R., 1965. Theory of mechanical properties of fibre-strengthened materials—III. Self-consistent model. Journal of the Mechanics and Physics
of Solids, 13(4), pp.189-198.
Huang, Z., 2001a. Simulation of the mechanical properties of fibrous composites by the bridging micromechanics model. Composites Part A:
Applied Science and Manufacturing, 32(2), pp.143-172.
1900 Saiaf Bin Rayhan et al. / Procedia Structural Integrity 28 (2020) 1892–1900
Author name / Structural Integrity Procedia 00 (2019) 000–000 9

Huang, Z., 2001b. Micromechanical prediction of ultimate strength of transversely isotropic fibrous composites. International Journal of Solids and
Structures, 38(22-23), pp.4147-4172.
Lo Cascio, M., Milazzo, A. and Benedetti, I., 2020. Virtual element method for computational homogenization of composite and heterogeneous
materials. Composite Structures, 232, p.111523.
Luo, H. and Weng, G., 1987. On Eshelby's inclusion problem in a three-phase spherically concentric solid, and a modification of Mori-Tanaka's
method. Mechanics of Materials, 6(4), pp.347-361.
Michel, J., Moulinec, H. and Suquet, P., 1999. Effective properties of composite materials with periodic microstructure: a computational approach.
Computer Methods in Applied Mechanics and Engineering, 172(1-4), pp.109-143.
Mori, T. and Tanaka, K., 1973. Average stress in matrix and average elastic energy of materials with misfitting inclusions. Acta Metallurgica,
21(5), pp.571-574.
Ogierman, W., 2019. HYBRID MORI-TANAKA/FINITE ELEMENT METHOD IN HOMOGENIZATION OF COMPOSITE MATERIALS
WITH VARIOUS REINFORCEMENT SHAPE AND ORIENTATION. International Journal for Multiscale Computational Engineering,
17(3), pp.281-295.
Otero, F., Oller, S., Martinez, X. and Salomón, O., 2015. Numerical homogenization for composite materials analysis. Comparison with other
micro mechanical formulations. Composite Structures, 122, pp.405-416.
Reuss, A., 1929. Berechnung der Fließgrenze von Mischkristallen auf Grund der Plastizitätsbedingung für Einkristalle. ZAMM - Zeitschrift für
Angewandte Mathematik und Mechanik, 9(1), pp.49-58.
Rodrigues, D., Belinha, J., Pires, F., Dinis, L. and Jorge, R., 2018. Homogenization technique for heterogeneous composite materials using meshless
methods. Engineering Analysis with Boundary Elements, 92, pp.73-89.
Saeb, S., Steinmann, P. and Javili, A., 2016. Aspects of Computational Homogenization at Finite Deformations: A Unifying Review from Reuss'
to Voigt's Bound. Applied Mechanics Reviews, 68(5).
Sendeckyj, G., Wang, S., Steven Johnson, W., Stinchcomb, W. and Chamis, C., 1989. Mechanics of Composite Materials: Past, Present, and Future.
Journal of Composites Technology and Research, 11(1), p.3.
Sun, H., Di, S., Zhang, N. and Wu, C., 2001. Micromechanics of composite materials using multivariable finite element method and homogenization
theory. International Journal of Solids and Structures, 38(17), pp.3007-3020.
Voigt, W., 1889. Ueber die Beziehung zwischen den beiden Elasticitätsconstanten isotroper Körper. Annalen der Physik, 274(12), pp.573-587.
Xin, H., Mosallam, A., Liu, Y., Veljkovic, M. and He, J., 2019. Mechanical characterization of a unidirectional pultruded composite lamina using
micromechanics and numerical homogenization. Construction and Building Materials, 216, pp.101-118.
Yan, C., 2003. ON HOMOGENIZATION AND DE-HOMOGENIZATION OF COMPOSITE MATERIALS. Ph.D. Thesis. Drexel University,
PA, USA.
Younes, R., Hallal, A., Chehade, F. and Fardoun, F., 2012. Comparative Review Study on Elastic Properties Modeling for Unidirectional Composite
Materials. INTECH Open Access Publisher.

You might also like