You are on page 1of 55

Accepted Manuscript

Elastic properties of particle-reinforced composites containing


nonspherical particles of high packing density and interphase: DEM-FEM
simulation and micromechanical theory
Wenxiang Xu, Binbin Xu, Fenglin Guo

PII: S0045-7825(17)30590-X
DOI: http://dx.doi.org/10.1016/j.cma.2017.08.010
Reference: CMA 11549

To appear in: Comput. Methods Appl. Mech. Engrg.

Received date : 1 May 2017


Revised date : 10 July 2017
Accepted date : 9 August 2017

Please cite this article as: W. Xu, B. Xu, F. Guo, Elastic properties of particle-reinforced
composites containing nonspherical particles of high packing density and interphase: DEM-FEM
simulation and micromechanical theory, Comput. Methods Appl. Mech. Engrg. (2017),
http://dx.doi.org/10.1016/j.cma.2017.08.010

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form.
Please note that during the production process errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.
Highlights

H ighlights

1. PRCs consist of ellipses of high packing density, compliant interphase and matrix.

2. A robust coupling model of DEM and FEM is proposed to estimate the moduli of PRCs.

3. A micromechanical theory incorporating the interphase fraction is developed.

4. The numerical model and micromechanical framework have a reasonable accuracy.

5. The derived results provide a robust predictive toolkit for the design of PRCs.
Revised Manuscript
Click here to download Manuscript: Revised Manuscript.doc Click here to view linked References

E lastic properties of particle-reinforced composites containing nonspherical particles of high

packing density and interphase: D E M-F E M simulation and micromechanical theory


1
2
3
4
5
6 Wenxiang Xu*a, Binbin Xu a, b, Fenglin Guo b
7
8 a
Institute of Soft Matter Mechanics, College of Mechanics and Materials, Hohai University, Nanjing 211100, China
9
10
11 b
State Key Laboratory of Ocean Engineering, School of Naval Architecture, Ocean and Civil Engineering, Shanghai
12
13
14 Jiao Tong University, Shanghai 200240, China
15
16
17 A bstract: The physical and morphological properties of particles and interfaces can seriously impact
18
19
the whole mechanical behavior of particle-reinforced composites. In this work, we devise a robust
20
21
22 coupling model of the discrete element method (DEM) and the finite element method (FEM) to
23
24
25 numerically investigate the effective elastic moduli of particle-reinforced composites consisting of
26
27
28 elliptical particles of high packing density, compliant interfaces and homogeneous matrix. In the
29
30
numerical model, a novel parametric equation for the topological geometry of an interface that is
31
32
33 treated as an interphase model is formulated to realize a compliant (penetrable) layer with a constant
34
35
36 finite thickness coated surrounding each elliptical particle. Additionally, a convenient strategy is
37
38
39 implemented to cope with periodic boundary conditions containing numerous particles. On the other
40
41
42
hand, we also propose a micromechanical theoretical framework to derive the effective elastic
43
44 properties of such three-phase composites by incorporating the fraction of compliant interfaces that is
45
46
47 theoretically computed by using the statistical geometry of composites. It is shown that our numerical
48
49
50 and theoretical models lead to the prediction of elastic moduli of particle-reinforced composites with
51
52
53
nonspherical particles of high packing density to a reasonable accuracy by comparing with available
54
55 experimental data. Moreover, utilizing the proposed models, we systematically investigate the influence
56
57 
58 
Correspondence to: Institute of Soft Matter Mechanics, College of Mechanics and Materials, Hohai University, Nanjing 211100, China.
59
60 E-mail address: xuwenxiang@hhu.edu.cn (W. Xu).
61
62  1
63
64
65
of the characteristics of particles and interphase such as the high packing density and geometries of

elliptical particles, and interphase fraction, thickness and stiffness on the elastic moduli of
1
2
3 particle-reinforced composites. We find that these physical characteristics play significant roles in
4
5
6 determining the mechanical behavior of composites, suggesting that the properties of such materials
7
8 can be tailored via proper composites engineering and design.
9
10
11 K eywords: Particle-reinforced composites; Nonspherical particle; Interphase; Discrete element model;
12
13
14 Micromechanical theory; Elastic modulus
15
16
17 1. Introduction
18
19
It is well known that particles and interfaces are important components in various
20
21
22 particle-reinforced composites (PRCs) like polymers, colloids, ceramics, normal concrete and asphalt
23
24
25 concrete [1-5], which play important roles in the mechanical properties of PRCs. Owing to different
26
27
28 technologies of the manufacturing for materials, interface zones may be a compliant interphase
29
30
possessing a low stiffness or a relative high porosity with respect to matrix in MXD6-clay
31
32
33 polymer-composites [3] and cementitious composites (the so-called interfacial transition zones, ITZs
34
35
36 in concrete) [4, 5]. It is normally considered as a penetrable interfacial layer with a constant finite
37
38
39 thickness surrounding each particle in PRCs [1-5], leading to a famous core-shell structure [6].
40
41
42
Understanding the effects of particle and interphase characteristics on the effective elastic properties of
43
44 PRCs and rigorously incorporating such structural information into homogenization scheme for
45
46
47 material property predictions is a problem of great interest in materials research community. The
48
49
50 resulting rigorous structure-property relations are also valuable for material design and microstructure
51
52
53
optimization.
54
55 Over the past decades, significant efforts have been devoted to theoretical and numerical assessing
56
57
58 the effects of particle and interphase characteristics on the effective elastic properties of PRCs. In
59
60
61 theoretical aspects, Kerner [7] seminally developed a three-phase composite model including spherical
62  2
63
64
65
particles to study the effective elastic properties. Later, Christensen and Lo [8] applied the generalized

self-consistent scheme and the Eshelbys equivalent inclusion theory [9] to evaluate the effective shear
1
2
3 moduli of three-phase PRCs. For PRCs containing nonspherical particles (e.g., ellipsoids, cylinders or
4
5
6 planets) and interphase, some popular micromechanical schemes have been presented to formulate the
7
8 elastic properties of such PRCs, such as bound models [10, 11], generalized self-consistent method
9
10
11 (GSCM) [12, 13], Mori-Tanaka scheme (M-T) [14, 15], double-inclusion model [16, 17], differential
12
13
14 effective medium approximation (D-EMA) [18], and exact contrast expansions [19]. We note that the
15
16
17 bound models are based on the well-known variational principle originally introduced in the pioneering
18
19
work of Hashin and Shtrikman [20]. The effective medium models (e.g., the GSCM, M-T,
20
21
22 double-inclusion and D-EMA mentioned above) mainly stem from the famous Eshelbys equivalent
23
24
25 inclusion formulation [9], However, a potential limitation for these micromechanical schemes is that
26
27
28 the fraction (i.e., packing density) of particles is usually processed in a relative low level. In this work,
29
30
we try to address the issue via the combination of numerical and theoretical explorations.
31
32
33 Alternatively, numerical simulations provide an optional means to evaluate the elastic properties
34
35
36 of PRCs containing nonspherical particles and interphase, especially for the high packing density of
37
38
39 nonspherical particles. For instance, Zhou et al. [21] estimated the elastic modulus of concrete
40
41
42
composed of convex polygonal grains and interphase by random sequential additions (RSA) and finite
43
44 element method (FEM). Other similar studies can also be found in [22-25]. Additionally, Xu and
45
46
47 coworkers [26] and Abyaneh et al. [27] applied the three-dimensional RSA of spheroidal particles and
48
49
50 the lattice model to predict the effective properties of cementitious composites. Moreover, Patel and
51
52
53
Zohdi [28] developed a statistically numerical scheme to estimate the effective properties of PRCs
54
55 containing ellipsoidal particles. Recently, the authors [29] simulated the effective mechanical properties
56
57
58 of fully-grade concrete consisting of multisized convex polyhedral particles by the coupling of RSA
59
60
61 and FEM. Although the numerical studies above mentioned enriched our knowledge of the effective
62  3
63
64
65
properties of three-phase PRCs with anisotropic-shaped particles, nonspherical particles of high

packing density and interphase are rarely involved in the literature. It is our intention in the present
1
2
3 work to address this gap.
4
5
6 In this study, we mainly focus on elliptical particles, as a canonical model of nonspherical shapes
7
8 over a broad range of aspect ratios, which have been widely used in numerous material systems. We
9
10
11 attempt to present a numerical model that adopts the coupling of the discrete element method (DEM)
12
13
14 and FEM to estimate the effective elastic moduli of PRCs composed of elliptical particles of high
15
16
17 packing density, compliant interphase and homogeneous matrix. Perfect bonding conditions are here
18
19
used for both the particle/interphase and the interphase/matrix. Meanwhile, we also try to formulate a
20
21
22 theoretical model to predict the effective elastic properties of such PRCs through the combination of
23
24
25 the Mori-Tanaka average scheme and the Eshelbys equivalent inclusion theory on the basis of the
26
27
28 recent contribution [14]. Moreover, the interphase physical characteristics like the interfacial fraction
29
30
and thickness are derived and incorporated into the predictions of the effective elastic properties of
31
32
33 PRCs in the present numerical and theoretical work.
34
35
36 The rest of this article is organized as follows. Section 2 describes a numerical model for the
37
38
39 estimation of the effective elastic modulus of PRCs with elliptical particles of high packing density and
40
41
42
interphase, where the foundation of DEM, the topological geometry of an interface, the interphase
43
44 fraction, and the coupling of DEM and FEM are elaborated in detail, respectively. Section 3 illustrates a
45
46
47 micromechanical framework to theoretically compute the effective elastic properties of such PRCs,
48
49
50 among the fraction of compliant interfaces is formulated through the statistical geometry of composites
51
52
53
and geometrical probability. In Section 4, some results are given and discussed. Finally, this article is
54
55 completed with some concluding remarks in Section 5.
56
57
58 2. Numerical model
59
60
61 In this section, we firstly describe DEM to establish two-phase PRCs consisting of monodisperse
62  4
63
64
65
non-overlapping elliptical particles and homogeneous matrix. Subsequently, a parametric equation for

the topological geometry of an interfacial layer with a constant dimension is proposed to generate
1
2
3 three-phase PRCs. Previous works for the geometric construction of an interfacial layer, such as
4
5
6 [24-27], has demonstrated by uniformly enlarging the semi-axes (e.g., a1 and b1) of an elliptical particle
7
8 to a constant dimension t, namely, a1+t and b1+t. Unfortunately, such simplified fashion meets the
9
10
11 limitation of the accuracy of the core-shell structure that shells coating around cores are specified the
12
13
14 same thickness. To be specific, the previous operations cannot ensure the shell possessing the same
15
16
17 dimension around an elliptical particle, due to the difference of the surface curvature of an ellipse,
18
19
especially in the large aspect ratio. Then, the DEM-FEM coupling model is presented to estimate the
20
21
22 effective elastic modulus of three-phase PRCs.
23
24
25 2.1 The discrete element model (D EM)
26
27
28 DEM stemmed from Cundall and Strack [30] has been extensively used in granular media, PRCs,
29
30
chemical engineering and mineral engineering [31-38]. The main idea of DEM is to find the
31
32
33 equilibrium contact forces and displacements of an assembly of particles based on a series of
34
35
36 calculations tracing the movements of the individual particles. These movements attribute to the initial
37
38
39 unbalance forces like gravity or the propagation of disturbances originating from the boundaries. Thus,
40
41
42
there are two strategies to implement the accumulation of particles. One is that particles are randomly
43
44 casted in a sample space and allowed to move under the condition of gravity. Another is that the
45
46
47 boundaries are controlled to move inward and to push the particles together like bulldozers. Herein, the
48
49
50 later strategy is adopted in this work.
51
52
53
The classical Newtons second law and a force-displacement law are alternately applied in DEM,
54
55 where both laws are respectively used to derive the contact forces and the motion locus of a particle.
56
57
58 Suppose an overlap potential  between particle i and particle j can be decomposed into the normal and
59
60
61 tangential components n and t at a time step t . The normal and tangential directions depend on the
62  5
63
64
65
determination of the location of the specific point C, as shown in Fig. 1. Herein, we define the point C

as the midpoint of the segment joining two contact points C i and C j that follow the contribution from
1
2
3   
      C i and C j are also referred as the lowest geometric
4
5
6 potential points introduced in [36], which can be obtained through a series of linear transformations of
7
8 the coordinate system presented in the preliminary work [34]. As shown in Fig. 1, the normal direction
9
10
11 is prescribed from C i pointing to C j and its counterclockwise angle is defined as the positive value.
12
13
14 Then, the tangential direction is perpendicular to the normal direction and satisfies the right-hand rule.
15
16
17 Therefore, the segment distance between C i and C j is the normal component n(t), and the
18
19
20
corresponding tangential component can be obtained by the incremental method, namely, t(tt(t-t)
21
22 + vtt, where vt is the tangential velocity in the current time step. It is noted that vt equals to 0 at the
23
24
25 initial time whose value in next time step can be obtained by the coordinate transformation of the
26
27
28 previous velocities in the global coordinate system.
29
30
The corresponding normal and tangential forces between adjacent particles are thus expressed by
31
32
33 Fn (t )  kn n (t ), Ft (t )  Ft (t  t )  kt t (t ) (1)
34
35
36 where F n and kn are the normal force and stiffness, respectively. F t and k t are the tangential force and
37
38
39 stiffness, respectively. According to the Coulomb-type friction law, the magnitude of F t is checked
40
41
42
against the maximum possible value ( F t)max, that is
43
44 ( Ft )max  Fn tan u  c (2)
45
46
47 where u is the smaller value of the interparticle friction angles in contact and its value does not depend
48
49
50 only on the density and shape of particles, but also on the experimental conditions. For convenience of
51
52
53
simulation, we adopt the experimental data reported by Schanz and Vermeer [39] that the friction angle
54
55 was presented in the range of 40 to 50 degrees for dense sands. In this work, the friction angle is set to
56
57
58 be 40 degrees for the implementation of DEM. c is the smaller one of their cohesions. Once F t and F n
59
60
61 are determined for each particle, they can be resolved into components in the x and y directions of the
62  6
63
64
65
Cartesian coordinate system, as shown in Fig. 1. Also, the moment on particle i can be obtained by M =

F T i, where T i represents the position vector from point C to the centroid of particle i.
1
2
3 If an individual particle overlaps with more than one particles, the sums of these contact force
4
5
6 components  F x and  F y are applied to the individual particle. Plus, the resultant moment M is
7
8 just derived in terms of the scalar operation, since the direction of M is always perpendicular to the
9
10
11 plane of the paper. Consequently, the accelerations of the individual particle are calculated through
12
13
14 Newtons second law:
15
16
17 a   F / m,    M / I (3)
18
19
where a and m are the acceleration generating the displacement and the mass of the particle,
20
21
22 respectively.  and I are the angular acceleration and the moment of inertial of the particle, respectively.
23
24
25 In DEM, velocities and accelerations are normally assumed to be constant in one short time step t, so
26
27
28 that velocities and displacements of particles in t can be obtained as follows.
29
30 v  v 0  a   t ,   0     t (4)
31
32
33 u  u 0  v   t ,   0     t (5)
34
35
36 where v and  and the linear and angular velocities. u and  are the displacement and angular rotation.
37
38
39 The subscript 0 represents the corresponding quantity of the previous step. The velocities in the global
40
41
42
coordinate system can also be decomposed into two components in the local coordinate system again to
43
44 calculate the tangential overlap and the damping forces in the normal and tangential direction if needed.
45
46
47 The above process can be repeated until all particles reach equilibrium.
48
49
50 It is worth stressing that the key issue of the above implementation for DEM of elliptical particles
51
52
53
is to identify the contact detection and the overlap potential for adjacent ellipses. Fortunately, in our
54
55 previous works, a robust numerical algorithm for the quantitative contact of ellipses has been
56
57
58 developed [34], even more the algorithm can also be directly extended to the contact of
59
60
61 three-dimensional ellipsoids [35]. In the numerical algorithm, a series of linear transformations of the
62  7
63
64
65
coordinate system are executed to transform an ellipse (ellipsoid) into a unit circle (sphere) with its

center located at the origin of the coordinate system. Also, a novel golden section search algorithm
1
2
3 instead of solving a complex fourth-order (six-order) polynomial equation reported in the literature is
4
5
6 presented to numerically derive the overlap potential of adjacent ellipses (ellipsoids). More detailed
7
8 descriptions on the operation of the numerical algorithm can be found in [34, 35]. In addition, other
9
10
11 promising lines of theoretical research on DEM can also found in the literature [36-38].
12
13
14 In order to improve the efficiency of the present DEM, on the one hand, we divide a sample space
15
16
17 (e.g., a square area) into some subspaces, in which each subspace is numbered, as shown in Fig. 2(a).
18
19
The random generated particles in a subspace are just labeled as the same flag as the corresponding
20
21
22 subspace. In doing so, it is very convenient to compute the contact forces of a labeled particle, since the
23
24
25 search for adjacent particles is only implemented in the neighboring eight subspaces for the
26
27
28 two-dimensional (2D) case. As an introductory example, we can clearly see from Fig. 2(a) that an
29
30
ellipse within the subspace with the number 11 potentially contacts with those ellipses within the
31
32
33 subspaces with the number 6, 7, 8, 10, 12, 14, 15 or 16. On the other hand, we put forward to a
34
35
36 convenient periodic compensation scheme for satisfying periodic boundary conditions specified in this
37
38
39 work, as shown in Fig. 2(b). In one-time step t, we prescribe the upper and left boundaries move
40
41
42
inward, so that the particles within the subspaces of the first line and column are thereupon pushed to
43
44 move inward. When these particles update their locations, they are immediately copied to the bottom
45
46
47 and right regions, forming the borders of the compensatory additional particles, as shown in Fig. 2(b).
48
49
50 The formed two particle-borders move inward with the same velocity as the upper and left boundaries.
51
52
53
Finally, copying the particles in the subspaces of the last row and column to the left and top regions
54
55 generates a packing model with periodic boundary conditions, as shown in Fig. 2(b). Fig. 3 depicts two
56
57
58 introductory DEM structures representing two-phase PRCs with monodisperse non-overlapping
59
60
61 ellipses of different aspect ratios and homogeneous matrix.
62  8
63
64
65
2.2 The accurate construction of interphase

It is very crucial to precisely realize the topological geometry of an interfacial layer around an
1
2
3 elliptical particle, for the construction of three-phase PRCs and then the accurate estimation of
4
5
6 mechanical properties of such PRCs. Unlike polyhedrons or cylinders, whose interfacial layers can be
7
8 geometrically constructed by the Minkowsky sum manner [25, 40], the interfacial layer with a constant
9
10
11 dimension around an ellipse (ellipsoid) meets a huge challenge through the Minkowsky sum manner,
12
13
14 owing to the nature of the difference of the surface curvature of an ellipse. In this section, we will
15
16
17 develop a parametric equation for the accurate generation of the topological structure of an interfacial
18
19
layer with a constant dimension around an arbitrary ellipse.
20
21
22 First of all, we transform an arbitrary ellipse into a unit circle with its center located at the origin
23
24
25 through three coordinate transformations, namely, the rotation, translation and scaling transformations,
26
27
28 as shown in Fig. 4. The detailed operation can be found in [34]. Suppose that a straight-line L is tangent
29
30
to the unit circle in the coordinate system x3-y3, and the tangent point is denoted as A (cos(), sin()).
31
32
33 Thus, the linear equation for L can simply expressed as
34
35
36 sin( ) y  cos( ) x  1  0 (6)
37
38
39 In the coordinate system x2-y2, the corresponding tangent point A 1 and the slope of the
40
41
42
straight-line perpendicular to L 1 are written as (a1cos(), b1sin()) and a1sin() / (b1cos()), respectively.
43
44 We construct a unit vector n perpendicular to L 1 at the point A 1, meanwhile n is also the normal vector
45
46
47 of the ellipse at point A 1:
48
49 (b1 cos( ), a1 sin( ))
50 n (7)
51 a12 sin 2 ( )  b12 cos 2 ( )
52
53
54 Then, we can use the addition of the vector to get any one-point P1 on the interface in the coordinate
55
56
57 system x2-y2:
58
59 P1  A 1  n  t (8)
60
61
62  9
63
64
65
where t is the dimension of interface. Actually, Eq. (8) stands for the parametric equation for the

interface on the standard ellipse. Furthermore, the standard ellipse can be inversely transformed into the
1
2
3 original arbitrary ellipse. Consequently, the parametric equation for an interfacial layer with a constant t
4
5
6 around an arbitrary ellipse can be presented by
7
8 P  R TP1 (9)
9
10
11 where P represents the assembly of points on the interface around an arbitrary ellipse. R and T are the
12
13
14 rotation and translation matrices of coordinate transformations, of which the display-formats have been
15
16
17 introduced in the previous work [34]. Fig. 5 shows some typical patterns for the topological geometries
18
19
of the interfacial layer around an elliptical particle, where the green, blue and red curves represent the
20
21
22 profiles of the elliptical particle, the interface with a constant dimension realized by the proposed
23
24
25 parametric model and the approximate interface constructed by the simplified fashion mentioned above,
26
27
28 respectively. It can be explicitly seen that the difference between both interfacial topological geometries
29
30
realized by the above two methods, especially for an ellipse of the large aspect ratio.
31
32
33 2.3 The coupling of D EM and F EM
34
35
36 On the basis of the two-phase PRCs and the interfacial construction demonstrated above,
37
38
39 three-phase PRCs composed of elliptical particles of high packing density, compliant interphase and
40
41
42
homogeneous matrix can be established, among the intersections between an interfacial layer and
43
44 particles and neighboring interfacial layers may occur, as shown in Fig. 6. In doing so, the overlapping
45
46
47 interfacial fraction should be addressed before the prediction of elastic properties of PRCs. We adopt a
48
49
50 Monte Carlo random point sampling scheme to statistically evaluate the interfacial fraction, and its
51
52
53
operation strategy is elaborated in the Supplementary Information (see S1).
54
55 Subsequently, the geometric information about the particles and interfaces, such as their positions,
56
57
58 sizes and directions, is imported in the finite element software like  Due to the potential
59
60
61 overlap of interfaces, the definition of the material properties for different phases meets difficulty.
62  10
63
64
65
Herein, we discretize the whole three-phase structures by using the 6-node triangular element in the

       !, since the triangular element type has the great adaptation to
1
2
3 complex geometry. The mesh size of elements is set to be a half of the dimension of interfaces, which is
4
5
6 small enough to investigate the interphase effect. Fig. 7 displays a typical detailed meshing structure for
7
8 three-phase PRCs, where there are about 150,000 nodes and 70,000 elements, and the size ratio
9
10
11 between the cell and particles is set to be 8.5.
12
13
14 According to the average strain and stress theorems, the average stress ij and strainij of a
15
16
17 volume element V can be computed as average integral both on external boundary and volume:
18
1 1 1 1
19
20
 ij 
V V  ( x)dV  V  v Ti (S ) x j dS ,
ij 
 ij 
V V  ij ( x )dV 
V  v [ui (S )n j  u j (S )ni ]dS

(10)
21
22 where x, n, S, T (S) and u(S) denote the position vector, the outward unit normal vector, the volume
23
24
25 surface, the boundary traction and the boundary displacement fields, respectively. Through the
26
27
28      "           
  
 lowing:
29
30  ij  C ijkl  kl ,  ij  S ijkl kl (11)
31
32
33 where C ijkl and Sijkl are the apparent stiffness and compliance tensors, respectively. Then the boundary
34
35
36 displacement field of a PRC sample can be solved by ABAQUS as static mechanical problems. After
37
38
39 the average strainij is calculated by the above Eq. (10), the apparent stiffness tensor or compliance
40
41
42
tensor can be easily derived by Eq. (11). For the present 2D case, only the engineering moduli E 11,22
43
44 corresponding to the axial strain and stress are here solved by the coupling of FEM and DEM. It is
45
46
47 worth stressing that particles in a limited sample do not satisfy ideally the random orientation, so that
48
49
50 C ijkl or Sijkl are not perfectly isotropic unless the selected sample is large enough. In other words, the
51
52
53
sample meets the requirement of representative volume element (RVE), which will be concerned in
54
55 Section 4.2. Consequently, the apparent elastic properties are equal to the effective properties of PRCs
56
57
58 no matter what homogeneous displacement or traction boundary conditions are applied.
59
60
61 3. M icromechanical theory framewor k
62  11
63
64
65
At a microscopic scale, the average field theory is often used to evaluate the effective properties of

heterogeneous materials [12-17]. The average field theory can give the quantitative relation between
1
2
3 the average strain and stress of microscopic heterogeneous materials by defining the effective
4
5
6 mechanical properties, as depicted in Eqs. (10) and (11). Suppose a three-phase PRC is subjected to the
7
8 displacement boundary condition with a constant strain tensor 0. The elastic modulus tensors for
9
10
11 individual phase are marked as C p, C i and C m, in which the superscripts p, i , and m represent particles,
12
13
14 interphase and matrix, respectively. Also, the overall effective elasticity stiffnessC of PRC is defined
15
16
17 by
18
19
20
  C   C 0 (12)
21
22 where and are the volume average stress and strain tensors of PRC given by
23
24
25   f p  p  f i  i  f m m  f p C p  p  f i C i  i  f m C m  m (13)
26
27
28  = f p  p  fi  i  fm  m (14)
29
30
where fp, f i and fm are the fractions of particles, interphase and matrix, respectively. Normally, the
31
32
33 fraction of particles is given in a PRC. In other words, the fraction of interphase (interfaces) is strongly
34
35
36 related to the fraction of matrix:
37
38
39
fi  1 f p  fm (15)
40
41
42
Previous efforts have been spent analytically assessing the fraction of compliant interphase around
43
44 three-dimensional (3D) convex-shaped particles, like spheres [42], ellipsoids [43], convex polyhedra
45
46
47 [44] and spherocylinders [45]. Moreover, we recently developed a general formula for the fraction of
48
49
50 compliant interphase around 3D arbitrary convex-shaped particles [46]. All theoretical efforts are based
51
52
53
on the seminal work by Torquato and coworkers [47], who developed a theory of the nearest-surface
54
55 distribution functions to calculate the so-

  
 
   
 However, those
56
57
58 previous contributions are not suitable for the present 2D case, since the scaled-particle theory [48] for
59
60
61 the 3D case is entirely different from that for the 2D case in the derivation of the radial distribution
62  12
63
64
65
function at the contact value [47]. Therefore, it raises to concern on the fraction of compliant interphase

around 2D elliptical particles.


1
2
3 3.1 Theoretical model for the interfacial fraction in 2D case
4
5
6 By means of the statistical geometry of composites and the core-shell structure, the void exclusion
7
8 probability in two-phase composite media composed of 2D circular particles and voids is derived by
9
10
11 [47]
12
13
 r 2 +2 R r  R 2 r 2  
14
eV  r   1   exp    (16)
15
 1  1   
2
16
17
18 where e V(r ) is defined as the probability of find

V empty of composite media, which is a
19
20
21 circular cavity of radius r centered at an arbitrary point. In other words, the void exclusion probability
22
23
24 e V(r) represents the expected fraction of voids in such two-phase composite media.  is the number
25
26 density of circular particles with radius R.
 depicts a number-averaged treatment.  is a dimensionless
27
28
29 reduced density defined by
30
31
32  R2
 A  (17)
33   2
34
35
36 where
A is the average area of particles(x) is the gamma function. In fact, is the area fraction of
37
38
39 hard circular particles in two-phase media, i.e., = fp.
40
41
42
Considering a two-phase composite structure composed of circular particles and matrix, the void
43
44 exclusion probability e V(r) is essentially equivalent to the matrix exclusion probability, namely, the
45
46
47 fraction of matrix in the two-phase composite structure (i.e., e V(r ) = fm). Similarly, in a three-phase
48
49
50 composite structure with the core-shell model mentioned above, the quantity e V(r ) can still be mapped
51
52
53
into the fraction of matrix, since each particle and its surrounding layer with a constant dimension r can
54
55 be regarded as a composite particle, so that the three-phase structure is reduced to a two-phase system
56
57
58 containing composite particles and matrix. Also, fm  r 
   


59
60
61 overlap enough to fill up the remaining matrix. The quantity is designed to take into account the
62  13
63
64
65
overlap potential of compliant interphase. Consequently, the interfacial fraction around circular

particles is displayed the following formula on the basis of Eqs. (15) and (16).
1
2

t 2 +2 R t  R 2 t 2   
3    
4 f i  1  f p  1  exp 
 
 1 f p 1  f p   
2
5  
6 
(18)
 
 Pc t 2   
7
t2+ P t 2

= 1  f p  1  exp 

8 c
 
9
   1  f p  4 1  f 2   
10   p  
11
12
13
where t and P c are the interfacial dimension (i.e., r = t) and the perimeter of a circular particle,
14
15 respectively.
16
17
18 When the theory is applied to the three-phase elliptical particle composite structure, e V(t) still
19
20
21 stands for the fraction of matrix not occupied by all elliptical particles and interphase. However, e V(t ) is
22
23
24 subjected to the number density, size and the mean perimeter of elliptical particles. In terms of the
25
26 quantitative stereology [49], the number density of 2D solids equals to the ratio of the area fraction of
27
28
29 solid phase to the mean area of solids, namely,  = fp / A. On the other hand, an equivalent diameter
30
31
32 D eq, defined as the diameter of an equivalent sphere having the same volume as a non-spherical particle,
33
34
35 was introduced to characterize the size of the non-spherical particle in previous lines of research
36
37 [43-46]. We here adopt D eq as the size of an elliptical particle, so that the mean area of elliptical
38
39
40 particles can denoted as A =  D 2eq / 4. Plus, the perimeter P e of an ellipse can be approximately
41
42
43 expressed as [25]
44
 Deq 1   
1    ,   a / b
45 2
h h 2 h3 h4 
46 Pe  1      , h  (19)
1   
e e
2 
2
47 4 64 256 16384 
48
49
50 where a e and be are the lengths of the semi-major and semi-minor axes of an ellipse, respectively. It
51
52
53
should be pointed out that, in a structure cell, the shape of each ellipse is assigned to be identical so as
54
55 to clearly understand the impact of morphological detail of anisotropic particles. Consequently, upon
56
57
58 substituting Eq. (19) into Eq. (18), the fraction of compliant interphase around elliptical particles can be
59
60
61 derived by
62  14
63
64
65
 4f
t 2 + H   D t H 2   f D 2 t 2   
   
f i = 1  f p  1  exp   2p eq p eq
 (20)


 Deq



 p
1  f Deq 1  f p    
2 2


1
2
3 with
4
5
6 H   
1   
1  h  h2 
h3

h4



7 2  4 64 256 16384 
8
9
10 If a composite structure with the size of elliptical particles is designed as the mean value or
11
12
13
monodispersity, we may prescribe a geometric size factor  to reflect the coupling effect between the
14
15  
                    t / D eq. Thus, the interfacial fraction in such
16
17
18 three-phase PRCs is reduced to
19
20

 2 + H   f H 2    2   
21
22

f i = 1  f p  1  exp 4 f p
    p    (21)
23   1  f p  1  f p   
2

24  
25
26 3.2 Theoretical model for elastic moduli in 2D case
27
28
29 In the recent theoretical work [14], a two-phase PRC is Eshelby s inhomogeneous inclusion
30
31
32 problem and the total stress field is given by
33
34
35    m    C p (0   m   )  C m (0   m     ) (22)
36
37 where!m is an average strain perturbation in matrix from 0 due to the existence of particles, as well as
38
39
40 the corresponding average stress perturbation !m  
!m = C !! m.  and  are the stress
41
42
43 perturbation and strain perturbation, respectively. * is      stress-free .
44
45
46 Similarly, for a three-phase PRC, the presences of interphase and particles usually raise the
47
48 perturbed strains i and p, and the     stress-free s *i and *p for interphase
49
50
51 and particles with respect to matrix. Following up Eq. (22), the average strain of each phase in the PRC
52
53
54 can be expressed as
55
56
57  m  0   m ,  i  0   m   i ,  p  0   m   p (23)
58
59 i and p have been formulated by Hori [16]
60
61
62  15
63
64
65
fp i
 i  S i   i 
fi
 S  S p   p  i  ,  p  S p  p   Si  S p  i (24)

1
2
where Si and Sp are 

 
 s for interphase and particles, respectively. From the 


3
4 strain distribution is related to * as [9]
5
6
7     (0   m ),   p, i (25)
8
9
10 where M p and M i are the fourth-order tensors:
11
12 1
13

f
1

f 
14 M   S p   p   Si  S p  S p  p  Si  S p   i  S p  p  Si  S p    p  
p

15  f  f 
 i i 
16
17
18 1


f
1

f 
19
20
M    S i  S p    S p   p  S p  p  Si  S p    p  S p  p  Si  S p   i  
i

 f  f 
21  i i 
22
23
24 with
25
    C   C m  C m ,   p, i
1
26
27
28
29 when Eq. (25) is substituted into Eq. (24), i andp are rewritten as
30
31

f 
32  i  Si M i  p  Si  S p   0   m  ,  p 
S p M p   Si  S p  M i   0   m  (26)
33
fi 
34
35
36 by substituting Eq. (26) into Eq. (23), the average strain of each phase is further arranged by
37
38

f 
39  m  0   m ,  i  I  Si M i  p  Si  S p   m ,  p 
I  S p M p   Si  S p  M i   m (27)
40 fi 
41
42
43 where I is the fourth-order identity tensor. Upon substituting Eq. (27) into Eq. (13), the constitutive
44
45
46 relation of the three-phase PRC is formulated as
47
48 
fp  
49    f m C m  f p C p
I  S p M p   S i  S p  M i   f i C i I  S i M i   S i  S p     m (28)
50  fi  
51
52
53
From Eqs. (14) and (27), another form for the constitutive relation is given by
54
55 
fp  
56  =  f m I  f p
I  S p M p   S i  S p  M i   f i I  S i M i   S i  S p     m (29)
57  fi  
58
59 Upon substituting Eqs. (28) and (29) into Eq. (12), the effective elasticity stiffnessC of the three-phase
60
61
62  16
63
64
65
PRC is derived by

 f  
1 C   f m C m  f p C p  I  S p M p   S i  S p  M i   f i C i  I  S i M i  p  S i  S p   
2   fi  
3 1
(30)
 fp i p 

 f m I  f p  I  S M   S  S  M   f i  I  S M   S  S   
4 p p i p i i i
5
6   fi  
7
8 In the present three-phase PRCs, the   
 rs Si and Sp are same, namely, Si = Sp = S (see
9
10
11 Supplementary Information, S2), owing to the coaxial nature for an elliptical particle and its
12
13
14 surrounding interface. Thus, the fourth-order tensors M p and M i are reduced to
15
M     S     ,   p, i
16 1
17 (31)
18
19
with
20
21
22 Km Gm
    C   C m  C m  e1 I h  e2 I d , e1x 
1 
23 , e = ,   p, i
K  Km G  Gm
2
24
25
26 where I h and I d are the four-order tensors given by Ihijkl = (1/3)ijkl and Idijkl = 0.5(ikjl +ilj k -1.5ijkl).
27
28
29 K and G are the bulk and shear moduli of individual phase, respectively. Furthermore, Eq. (30) is
30
31
32 rearranged by
33
  
34 1

35 C  f m C m  f p C p  I  SM p   f i C i  I  SM i  f m I  f p  I  SM p   f i  I  SM i  (32)
36
37 On the other hand, the present PRC consist of elliptical particles and interphase randomly distributed in
38
39
40 a homogeneous matrix, as shown in Fig. 6. The effective mechanical properties of PRCs are
41
42
43 transversely isotropic. Consequently, the effective stiffness for PRCs is dependent of five independent
44
45
46 stiffness coefficients and its stiffness matrix can be written as [50]
47
48
C11 C12 C13 0 0 0 
49 
C12 C11 C13 0 0 0 
50
51 C13 C13 C 33 0 0 0 
52 C  0 0 0 C 44 0 0
 (33)
53 
54 0 0 0 0 C 44 0 
55 
0 1
56
0 0 0 0  C11  C12  
57 2 
58
59 Appendix A gives the detailed derivation for the effective stiffness components C ij of the present
60
61
62  17
63
64
65
three-phase PRCs. The stiffness components C ij are directly related to the effective mechanical

properties of PRCs, that is


1
2
C 33 C112  2C12 C132  2C11C132  C 33 C122 C C  C132 1
3
4
E eff  ,  eff  12 33 , G eff   C11  C12  (34)
C11C33  C13 2
C11C 33  C13
2
2
5
6
7 where E eff, eff and G eff are the effective elastic modulus, Poisson ratio and shear modulus of PRCs,
8
9
10 respectively.
11
12
13 4. Results and discussion
14
15 4.1 The esti mation of interfacial fraction
16
17
18 In accordance of the numerical and theoretical schemes for the interfacial fraction described above,
19
20
21 we observe the dependence of the interfacial fraction on the particle fraction and the aspect ratio, as
22
23
24 shown in Fig. 8. It can be seen from theoretical and numerical results that the dependence of the
25
26 interfacial fraction f i on the particle fraction fp possesses a well consistency for different aspect ratios.
27
28
29 To be specific, f i increases when fp increases up to about 0.6, which is also in good agreement with that
30
31
32 for 3D convex-shaped particles elucidated in previous contributions [43-46]. Nevertheless, as fp
33
34
35 continues to grow, f i yet gradually decreases. It seems to be the first time to our knowledge that such a
36
37 correlation under the high packing density of particles is explored. As mentioned above, f i is mapped
38
39
40 onto the probability of filling up matrix in composites. Actually, interfaces significantly occupy to the
41
42
43 remaining matrix as fp > 0.6. When the number of particles continues to increase, some interfacial
44
45
46 zones are occupied by the increased particles, so that the interfacial volume fraction gradually
47
48 decreases when fp exceeds 0.6. In practice, this is also one of theoretical reasons why interfacial
49
50
51 transition zones in high-performance concrete are normally lower than that in normal concrete [51].
52
53
54 It is worth noting that, for the numerical results of f i shown in Fig. 8, interfaces may generate a
55
56
57 percolating network as fp increases from 0.4 to 0.5, since the increase tendency of f i becomes slow
58
59 when fp exceeds 0.5. Such the percolation behavior can induce the dramatic change of elastic modulus
60
61
62  18
63
64
65
of PRCs, which will be discussed in the following section 4.3 (see Fig. 10). Unfortunately, the

percolation threshold of interfaces (or the percolation threshold of particles) cannot be unambiguously
1
2
3 determined in this work, though the issue is of importance that may be statistically analyzed through
4
5
6 the extensive finite-size scaling scheme [52] in the near future.
7
8 On the other hand, Fig. 8 manifests that fi increases with the increase of . When nonspherical
9
10
11 particles degenerate isotropic spheres, f i reaches the minimal value. In other words, elliptical particles
12
13
14 raise the increase of the interfacial fraction with respect to spherical particles. Such phenomena are
15
16
17 consistent with the 3D results [43, 44, 46], but opposite to the 2D results claimed by Zheng et al. [53].
18
19
This case is attributable to the circumference of an ellipse larger than that of the circle with the same
20
21
22 area. As such, in the process of the design of high performance concrete, the selections of gravels and
23
24
25 sands are usually required close to spherical particles as much as possible as excellent aggregates, in
26
27
28 order to reduce interfacial effects.
29
30
31
4.2 RVE size for elastic properties
32
33 In order to ensure the reliability of numerical results for elastic moduli, RVE for numerical
34
35
36 simulations should be determined before exploring the elastic properties of PRCs. A classical definition
37
38
39 for RVE reported by Hill [54] is introduced as a sample that is entirely representative of the whole
40
41
42
mixture on average and contains a sufficient number of particles. Thus, the overall properties are
43
44 macroscopically uniform and independent of boundary conditions. It is obvious that, in a certain
45
46
47 fraction of particles, the more the number of particles contains, the bigger the size of samples is. When
48
49
50 the size of samples satisfies the requirement of RVE, the measured value of elastic modulus is identical.
51
52
53
In this section, the effect of the number of particles on the elastic modulus of PRCs is numerically
54
55 analyzed, among the number of particles is applied to determine the size of RVE. In order to improve
56
57
58 the computational efficiency of DEM-FEM, a rectangular sample is divided into m cells along each
59
60
61 side. One particle is randomly placed in each cell before moving inside the rectangular boundaries.
62  19
63
64
65
Consequently, m2 random particles and the size of samples with periodic boundary conditions can be

obtained. The number of particles in samples is prescribed from 16 to 100 with an increment of 2(m+1).
1
2
3 Fig. 9 shows the change of the relative elastic modulus E eff / E m for PRCs with the number of particles
4
5
6 under different aspect ratios of particles. From Fig. 9, one can see that the dispersion of numerical
7
8 results for E eff decreases with the increase of the number of particles in samples. When the number of
9
10
11 particles is greater than 49, the average value of elastic moduli of 10 samples tends to stable. Also,
12
13
14 when the number of particles reaches 64, the size of numerical samples satisfies RVE and the
15
16
17 corresponding sample size is 8.5 times the size of particles.
18
19
4.3 Test for the numerical model and micromechanical theory framework
20
21
22 We apply the numerical model and micromechanical theoretical framework introduced above to
23
24
25 estimate the effective elastic moduli of PRCs and compare the derived numerical and theoretical results
26
27
28 with the experimental data for mortar reported by Yang [41], as shown in Fig. 10. In Yangs experiment,
29
30
31
cubic specimens (50  50  50 mm) and different particle fractions (fp = 0.1, 0.2, 0.3, 0.4 and 0.5) are
32
33 cast and tested by the compressive test, and both interfacial thicknesses of t = 20 m and 40 m are
34
35
36 examined in those experimental specimens. For purpose of comparison, we adopt the experimental
37
38
39 parameters into our numerical and theoretical models. Fig. 10 shows in the range of fp  0.4 that the
40
41
42
numerical and theoretical results for both interfacial thicknesses are in agreement with the experimental
43
44 data, specifically the numerical results. As fp > 0.4, the numerical results from the DEM-FEM model
45
46
47 are also basically consistent with the theoretical results, even though the numerical results are slightly
48
49
50 lower than the theoretical results, which their relative errors are limited within the scope of 12% for t =
51
52
53
20 m and 9% for t = 40 m, respectively. It is exceptionally important that suggesting the present
54
55 micromechanical framework is suitable for the prediction of elastic properties of PRCs containing
56
57
58 elliptical particles of high packing density.
59
60
61 In addition, Fig. 10 presents that elastic modulus versus the particle fraction almost satisfies the
62  20
63
64
65
linear-spring relationship within the scope of fp  0.4, but as fp > 0.4, the elastic modulus nonlinearly

increases with the increase of particle fraction. Such the phenomenon may attribute to the percolation
1
2
3 generation for the particle and interface phases when the particle packing density exceeds 0.4, which
4
5
6 corresponds to the percolation generation of interphase described in Fig.8. It is well known that the
7
8 dramatic changes in the physico-mechanical properties of PRCs occurs when inclusions form a
9
10
11 percolating network through the composite. Fig. 10 does not however present an obvious abrupt change
12
13
14 of elastic moduli of PRCs near the percolation thresholds of particles or interfaces, owing to the relative
15
16
17 small contrast of the moduli between the particle and matrix phases limited by the experimental data.
18
19
On the other hand, we also compute the elastic moduli of two-phase PRCs without interphase using the
20
21
22 micromechanical theory framework, as shown in Fig. 10. One can clearly see that the two-phase
23
24
25 theoretical results are apparently higher than the experimental data, numerical results and the
26
27
28 three-phase theoretical results. It means that the interphase effect dominates the overall elastic
29
30
properties of PRCs and needs to be considered in the design of the elastic properties of PRCs.
31
32
33 4.4 Effect of particle shape
34
35
36 As depicted in Fig. 8, the shape of elliptical particles characterized by the aspect ratio of  has
37
38
39 affected the interfacial fraction. In this section, we further explore the effect of particle shape on the
40
41
42
elastic moduli of three-phase PRCs through the DEM-FEM model and the micromechanical theory
43
44 framework. Since the elastic properties of PRCs under the relative low packing density of particles (e.g.,
45
46
47 fp  0.4) have been broadly studied in previous references [21-29], we just focus on the issue of the
48
49
50 high particle packing density. Fig. 11 displays respectively the numerical and theoretical results for the
51
52
53
relative elastic moduli E eff / E m of PRCs versus the high packing density of particles with  = 1, 2 and 3.
54
55 One can see from the theoretical and numerical results corresponding to a specific particle fraction, that
56
57
58 the change of the relative elastic moduli is very small with the increase from  = 1 to  = 3. It seems to
59
60
61 imply that the shape of elliptical particles has weak sensitive to the elastic moduli of PRCs with respect
62  21
63
64
65
to the high packing density of particles. Such the exploration is similar to the numerical estimation for

elastic moduli of concrete with the relative low packing density of aggregates reported by Zhou et al.
1
2
3 [21]. From Fig. 11, at first glance, the difference between theoretical and numerical results for a certain
4
5
6  seems to be distinct. Actually, the largest relative error between both results is only 10.7%
7
8 corresponding to  = 1 and fp = 0.55. Also, the reason attributes to the difference between both results
9
10
11 for the interfacial fraction shown in Fig. 8 and a certain machine error from implementing the
12
13
14 DEM-FEM model.
15
16
17 4.5 Effect of elastic modulus of interphase
18
19
Although Figs. 8 and 10 have manifested interphase as an important phase links particle to matrix
20
21
22 and its physical properties affected the overall elastic properties of PRCs, it is a huge challenge to
23
24
25 capture separately the elastic modulus of interphase through experiments. Recently, Shah and
26
27
28 coworkers [55] used Backscatter Electron and nanoindentation techniques to experimentally study the
29
30
elastic modulus of interphase (e.g., ITZs) in rock-filled concrete, and found that the elastic modulus of
31
32
33 ITZs is a variance with the porosity. In addition, in most of theoretical or numerical researches [21-29,
34
35
36 41, 42, 53], the elastic modulus of interphase is normally assumed to be a certain percentage of that of
37
38
39 bulk matrix. We here follow the previous predominant lines of study. Fig. 12 depicts the effect of the
40
41
42
ratio of the elastic modulus of interphase and that of bulk matrix on the effective elastic moduli of
43
44 PRCs under the high packing density of particles. One can clearly see that the selection of E i has a
45
46
47 pronounced influence on E eff. To be specific, E eff increases with the increase of E i as a matter of fact. On
48
49
50 the other hand, the relative error between the numerical and theoretical results gradually reduces from
51
52
53
14% to 0.9% with increasing the ratio of E i/ E m from 0.2 to 1.0. Such the deviation maybe originates
54
55 from the error of meshing interphase in the numerical simulations, since the three-phase PRCs
56
57
58 degenerate to two-phase structure consisting of particles and bulk matrix as E i/ E m = 1.0, reducing the
59
60
61 complexity of mesh generation scheme.
62  22
63
64
65
4.6 Effect of geometric size factor

In this section, we investigate the effect of geometric size factor  (i.e., the ratio of the interfacial
1
2
3 thickness and the average particle size, t / D eq) on the elastic moduli of PRCs with various high packing
4
5
6 densities of particles. Without loss of generality, we specified the constant interfacial thickness to
7
8 change the average particle size. Eventually, such the effect of  also reflects the impact of the average
9
10
11 particle size. Fig. 13 shows the elastic moduli of PRCs nonlinearly increase with the decrease of  or
12
13
14 the increase of D eq under different high fp. As an explanatory reason, the effect of  on the fraction of
15
16
17 compliant interfaces for 3D cases has been illuminated in our previous work [46] (see Fig. 2 in the
18
19
reference [46]), in which readers can clearly see that the interfacial fraction gradually increases to a
20
21
22 constant value with increasing  under a certain fp. As a weak phase, the enhancement of interfacial
23
24
25 fraction induces the reduction of the overall elastic properties of PRCs with increasing . From Fig. 13,
26
27
28 the discrepancy between theoretical and numerical results for a certain appears to be dramatic. But the
29
30
largest relative error between both results is only limited within the scope of 10% corresponding to 1/ 
31
32
33 = 5 and fp = 0.6. Such the error may attribute to two reasons. One is that, at the high packing density of
34
35
36 particles, the generation of percolating networks of the particle phase and interphase results in the
37
38
39 nonlinear change in the elastic moduli of PRCs. However, the present theoretical scheme does not
40
41
42
explicitly characterize the percolation-type behaviors near the percolation threshold, though the
43
44 simulated results shown in Figs. 8 and 10 seem to indicate the percolation threshold of particles is
45
46
47 within the interval of 0.4 and 0.5, of which the issue will be quantitatively addressed in the further work.
48
49
50 The other is the machine error for the statistical analysis of elastic moduli of PRCs in simulation. It
51
52
53
bears reiterating that these interesting explorations are very important to guide researchers to design
54
55 PRCs for their mechanical applications by controlling the geometric size factor.
56
57
58 5. Conclusions
59
60
61 In summary, we have presented a robust numerical model coupling of the discrete element method
62  23
63
64
65
(DEM) and finite element method (FEM) in details to predict the effective elastic moduli of

particle-reinforced composites (PRCs) containing elliptical particles of high packing density up to 0.8,
1
2
3 compliant interphase and homogeneous matrix. Whilst we devised a theoretical micromechanical
4
5
6 framework incorporating interphase physical characteristics like interfacial fraction, thickness, moduli
7
8 and Poisson ratio into the effective elastic properties of PRCs by using the combination of the
9
10
11 Mori-Tanaka scheme and the  equivalent inclusion theory. These numerical and theoretical
12
13
14 models addressed how the interphase characteristics and the geometries of particles dominate
15
16
17 quantitatively the elastic properties of PRCs with a high particle packing density. By comparing the
18
19
available experimental data, the proposed numerical and theoretical models are validated to be efficient
20
21
22 tools for the reasonable estimation of elastic properties of PRCs with a high particle packing density
23
24
25 and interphase. The derived numerical and theoretical results elucidated that, the high particle packing
26
27
28 density plays a predominant role in the interfacial fraction and elastic moduli of PRCs with respect to
29
30
the particle shape. Moreover, the elastic modulus of interphase and the reciprocal geometric size factor
31
32
33 have a positive impact on the elastic moduli of PRCs. These numerical and theoretical observations
34
35
36 provide researchers with significant insights on optimizing the particle geometries and interphase
37
38
39 characteristics for the design of particle-reinforced composites with the high packing density of
40
41
42
nonspherical particles.
43
44 A cknowledgments
45
46
47 This work was supported by the National Natural Science Foundation Project of China [grant
48
49
50 numbers 11402076, 51679078 and 51679265]; the Natural Science Foundation for Jiangsu Province
51
52
53
[grant number BK2017030243]; the Fundamental Research Funds for the Central Universities [grant
54
55 number 2016B06314]; and Research Fund of State Key Laboratory of Simulation and Regulation of
56
57
58 Water Cycle in River Basin [grant number IWHR-SKL-201709].
59
60
61 A ppendix A . Derivation of the elements C ij of the effective stiff matrix
62  24
63
64
65
In Eqs. (31) and (32), C m, C p and C i are isotropic material constant matrices of matrix, particle and

interface, respectively, and their elements can be expressed by individual engineering elastic constants
1
2
3 as
4
5  y 4 y 2 2

6 K 3G K y  Gy
3
K y  Gy
3
0 0 0
7 
8  K y  2Gy 4
K y  Gy
2
K y  Gy 0 0 0
9  3 3 3
10 
11 C  K y  2Gy
y
2
K y  Gy
4
K y  Gy 0 0 0 , y  m, p, i (A.1)
12  3 3 3
13 
14 0 0 0 Gy 0 0

15  0 0 0 0 Gy 0
16 
17 0 0 0 0 0 G y
18
19
Substituting of Eq. (A.1) into Eq. (31), the components of matrices of  in Eq. (31) are written by the
20
21
22 engineering elastic constants:
23
24
25  e1 +2e2 e1 -e2 e1 -e2 0 0 0

  
26
 e1 -e2 e1 +2e2 e1 -e2 0 0 0
27
28
 e1 -e2 e1 -e2 e1 +2e2 0 0 0
29 
1 0 3 
30   h  d
  e1 I  e2 I   0 0 e2 0 0 (A.2)
2
31 3
32
 0 3 
33 0 0 0 e2 0
34  2
35  3 
36  0 0 0 0 0 e2
37
2
38
39 Thus, the matrices M  of Eq. (31) are expressed by substituting S2 (Supplementary Information) and
40
41
42
Eq. (A.2) as
43
44  M 11 M 12 M 12 0 0 0

45  
46  M 21 M 22 M 23 0 0 0
47  M 21 M 23 M 22 0 0 0
48 M   S     
  1
(A.3)
49  0 0 0 M 44 0 0
50  0 0 0 0 M 55 0
51 
52  0 0 0 0 0 M 55
53
54
55 where the elements in the matrices are
56
57
58
59
60
61
62  25
63
64
65
1  1  1  1 
1
 e1  e2   s23  e1  2e2   s22
1
 e1  e2   s12  e1  e2   s12
M 11   3 3 , M 12   3 3
M0 1  1  1  1 
1
3
 e1  2e2   s22 3
 e1  e2   s23
M0
3
 e1  e2   s23 3
 e1  2e2   s22
2
3
4 1  1  1  1 
5

 e  e2   s21
1 3 1 3
 e1  e2   s23 1 3
 e1  e2   s12 3
 e1  2e2   s11
6 M 21 =  , M 22  
M0 1   1  1  1 
 e1  e2   s21  e1  2e2   s22  e1  2e2   s22  e1  e2   s21
7 M0
8
9
3 3 3 3
10
1  1 
 e  2e2   s11  e1  e2   s12
11
12
1 3 1 3 1 1
13 M 23   , M 44   
, M 55   
M0 1   1  0.5e2  s44 0.5e2  s55
14
15 3
 e1  e2   s21 3
 e1  e2   s23
16
17
1  1   1  
18
19 3
 e1  2e2   s11 3
 e1 -e2   s12 3
 e1 -e2   s12
20
1   1  1  
21
22
M 0 
3
 e1 -e2   s21 3
 e1  2e2   s22 3
 e1 -e2   s23
23
1   1   1 
24
25 3
 e1 -e2   s21 3
 e1 -e2   s23 3
 e1  2e2   s22
26
27
28 where sij are the components of the Eshelby tensor S (see S2). Subsequently, the terms I+SM  in Eq.
29
30
(32) are given by
31
32
33  a11 a12 a12 0 0 0

34  
35  a21 a22 a23 0 0 0
36  a21

a23 a22 0 0 0

37 I  SM   (A.4)
38  0 0 0 a44 0 0
39  0 0 0 0 a55 0
40 
41  0 0 0 0 0 a55
42
43
44 where the elements in the matrices are
45
46
47 a11  1  s11 M 11  2s12 M 21 , a12 =s11 M 12  s12 M 22  s12 M 23
48
49
50 a21  s21 M 11  s22 M 21  s23 M 21 , a22  1  s21 M 12  s22 M 22  s23 M 23
51
52 a23  a32  s21 M 12  s22 M 23  s23 M 22 , a44  1+s44 M 44 , a55  1+s55 M 55
53
54
55 As the matrices I+SM  have been known, the inverse matrix in Eq. (32) can be similarly formulated
56
57
58 by
59
60
61
62  26
63
64
65
b11 b12 b12 0 0 0 

b21 b22 b23 0 0 0 
b b22 0 0 0 

1 1 b
f m I  f p  I  SM p   f i  I  SM i  = 21 23  (A.5)
2
3 0 0 0 b44 0 0
4 0 0 0 0 b55 0 

5
0 0 0 0 0 b55 
6
7
8 where the elements in the matrix are
9
10
1 f m  f i a22  f p a22 f i a23  f p a23p f i a12i  f p a12p f i a12i  f p a12p
i p i
11 1
12 b11  , b12 
13 b0 f i a23
i
 f p a23p f m  f i a22
i
 f p a 22p b0 f m  f i a22i
 f p a22p f i a23
i
 f p a 23p
14
15 1 f i a21
i
 f p a21p f i a21
i
 f p a21p 1 f m  f i a11  f p a11
i p
f i a12i  f p a12p
16 b21  , b22 
17 b0 f m  f i a22
i
 f p a22p f i a23
i
 f p a 23p b0 f i a21
i
 f p a21p f m  f i a 22
i
 f p a 22p
18
19 1 f i a12i  f p a12p f i a23
i
 f p a 23p 1 1
20 b23  , b44  , b55 
21 b0 f m  f i a11i  f p a11p f i a21  f p a 21
i p
f m  f i a44  f p a44
i p
f m  f i a55i  f p a55p
22
23 f m  f i a11i  f p a11p f i a12i  f p a12p f i a12i  f p a12p
24
25 b0  f i a21 i
 f p a21p f m  f i a22i
 f p a 22p f i a 23
i
 f p a 23p
26 f i a21
i
 f p a21p f i a23
i
 f p a 23p f m  f i a 22
i
 f p a 22p
27
28
29 Therefore, on the substitution of Eqs. (A.1), (A.4) and (A.5) into Eq. (32), the elements C ij of the
30
31
32 effective stiff matrixC (see Eq. (33)) can be obtained by
33
1 1 1
34
35
C11   3c11  2c12  3c22  4c55  , C12   c11  6c12  c22  4c55  , C13   c12  c23  , C33  c22 ,
8 8 2
36
1
37 C 44   c44  c55 
38 2
39
40 with
41
42
4  4 4  2 2  
43 c11  K m  G m 
f   K   G   K m  G m  F11  2 K   G   K m  G m  F21 
44 3
45
  p ,i  3 3  3 3  
46
2  4 4  2 2  
c12  K m  G m 
f   K   G   K m  G m  F12  K   G   K m  G m   F22  F23  
47
48 3
49   p ,i  3 3  3 3  
50
4  4 4  2 2  
c22  K m  G m 
f   K   G   K m  G m  F22  K   G   K m  G m   F12  F23 
51
52
53
3   p ,i  3 3  3 3  
54
55 2  4 4  2 2  
56 c23  K m  G m 
f   K   G   K m  G m  F23  K   G   K m  G m   F12  F22 
57 3   p ,i  3 3  3 3  
58
59 c44  G m 
f G   G m  F  44
60  pi
 ,
61
62  27
63
64
65
c55  G m   f G   G m  F  55
 pi
 ,

where
1
2
3 F11 =a11 b11  2 a12 b21 , F12 =a11 b12  a12 b22  a12 b23 , F21 =a21

b11  a22 b21  a23 b21
4
5
6 F22 =a21 b12  a22 b22  a23 b23 , F23  a21
 
b12  a22 
b23  a23b22 , F44  a44 b44 , F55  a55 b55
7
8
References
9
10
11 [1] S.F. Li, X.W. Zeng, B. Ren, J. Qian, J.S. Zhang, A.K. Jha, An atomistic-based interphase zone
12
13
14 model for crystalline solids, Comput. Methods Appl. Mech. Engrg. 229 232 (2012) 87 109.
15
16
17 [2] H.L. Cheng, L. Meng, S.H. Chen, Y. Jiao, Y.M. Liu, Numerical investigation of microstructure
18
19
effect on mechanical properties of bi-continuous and particulate reinforced composite materials,
20
21
22 Comput. Mater. Sci. 122 (2016) 288 294.
23
24
25 [3] M. Kotal, A.K. Bhowmick, Polymer nanocomposites from modified clays: Recent advances and
26
27
28 challenges, Prog. Polym. Sci. 51 (2015) 127 187.
29
30
[4] V.P. Nguyen, M. Stroeven, L.J. Sluys, Multiscale failure modeling of concrete: Micromechanical
31
32
33 modeling, discontinuous homogenization and parallel computations, Comput. Methods Appl.
34
35
36 Mech. Engrg. 201 204 (2012) 139 156.
37
38
39 [5] T.T. Nguyen, J. Yvonnet, Q.-Z. Zhu, M. Bornert, C. Chateau, A phase-field method for
40
41
42
computational modeling of interfacial damage interacting with crack propagation in realistic
43
44 microstructures obtained microtomography, Comput. Methods Appl. Mech. Engrg. 312 (2016)
45
46
47 567 595.
48
49
50 [6] S. Torquato, Bulk properties of two-phase media. I. cluster expansion for the dielectric constant of
51
52
53
dispersions of fully penetrable spheres, J. Chem. Phys. 81 (11) (1984) 
 
54
55 [7] E.H. Kerner, The elastic and thermoelastic properties of composite media, Proc. Phys. Soc. B 69
56
57
58     
59
60
61 [8] R.M. Christensen, K.H. Lo, Solutions for effective shear properties in three phase sphere and
62  28
63
64
65
cylinder models, J. Mech. Phys. Solids 27 (4)   
 

[9] J.D. Eshelby, The determination of the elastic field of an ellipsoidal inclusion, and related problems,
1
2
3 Proc. R. Soc. Lond. A 241 (1226) 
   
4
5
6 [10] P.P. Castaeda, J.R. Willis, The effect of spatial distribution on the effective behavior of composite
7
8 materials and cracked media, J. Mech. Phys. Solids 43 (12) 


9
10
11 [11] H.L. Duan, J. Wang, B. Karihaloo, Theory of elasticity at the nanoscale, Adv. Appl. Mech. 42 (8)
12
13
14  
15
16
17 [12] X.K. Li, X. Zhang, J.B. Zhang, A generalized Hills lemma and micromechanically based
18
19
macroscopic constitutive model for heterogeneous granular materials, Comput. Methods Appl.
20
21
22 Mech. Engrg. 199 (4952) (2010) 31373152.
23
24
25 [13] W.X. Xu, H.F. Ma, S.Y. Ji, H.S. Chen, Analytical effective elastic properties of particulate
26
27
28 composites with soft interfaces around anisotropic particles, Compos. Sci. Technol. 129 (2016)
29
30
 
31
32
33 [14] W.X. Xu, F. Wu, Y. Jiao, M. Liu, A general micromechanical framework of effective moduli for the
34
35
36 design of nonspherical nano- and micro-particle reinforced composites with interface properties,
37
38
39 Mater. Design 127 (2017) 162172.
40
41
42
[15] J.F. Wang, L.W. Zhang, K.M. Liew, Multiscale simulation of mechanical properties and
43
44 microstructure of CNT-reinforced cement-based composites, Comput. Methods Appl. Mech. Engrg.
45
46
47 319 (2017) 393413.
48
49
50 [16] M. Hori, Double-inclusion model and overall moduli of multi-phase composites, Mech. Mater. 14
51
52
53
(3)    
54
55 [17] P. Lu, Y.W. Leong, P.K. Pallathadka, C.B. He, Effective moduli of nanoparticle reinforced
56
57
58 composites considering interphase effect by extended double-inclusion model-Theory and explicit
59
60
61    


62  29
63
64
65
[18] R. Xiao, X. Gou, W. Chen, Suppression of electromechanical instability in fiber-reinforced

dielectric elastomers, AIP Adv. 6 (3) (2016) 035321.


1
2
3 [19] S. Torquato, Effective stiffness tensor of composite media: I. exact series expansions, J. Mech.
4
5
6 Phys. Solids 45 (9)   
7
8 [20] Z. Hashin, S. Shtrikman, A variational approach to the theory of the effective magnetic
9
10
11 permeability of multiphase materials, J. Appl. Phys. 33 (10) 
 
12
13
14 [21] C. Zhou, K. Li, F. Ma, Numerical and statistical analysis of elastic modulus of concrete as a
15
16
17 three-phase heterogeneous composite, Comput. Struct. 139 (2014) 
18
19
[22] F. Grondin, M. Matallah, How to consider the Interfacial Transition Zones in the finite element
20
21
22 modelling of concrete, Cem. Concr. Res. 58 (2014)
 
23
24
25 [23] X.F. Wang, Z.J. Yang, J.R. Yates, A.P. Jikov, Ch. Zhang, Monte Carlo simulations of mesoscale
26
27
28 fracture modelling of concrete with random aggregates and pores, Constr. Build. Mater. 75 (2015)
29
30
3545.
31
32
33 [24] X. Li, Y. Xu, S. Chen, Computational homogenization of effective permeability in three-phase
34
35
36 mesoscale concrete, Constr. Build. Mater. 121 (2016) 100111.
37
38
39 [25] Z.G. Zhu, H.S. Chen, Overestimation of ITZ thickness around regular polygon and ellipse
40
41
42
aggregate, Comput. Struct. 182 (2017) 205218.
43
44 [26] L. Liu, D. Shen, H.S. Chen, W.X. Xu, Aggregate shape effect on the diffusivity of mortar: A 3D
45
46
47 numerical investigation by random packing models of ellipsoidal particles and of convex
48
49
50 polyhedral particles, Comput. Struct. 144 (2014) 4051.
51
52
53
[27] S.D. Abyaneh, H.S. Wong, N.R. Buenfeld, Modelling the diffusivity of mortar and concrete using a
54
55 three-dimensional mesostructure with several aggregate shapes, Comput. Mater. Sci. 78 (2013)
56
57
58 6373.
59
60
61 [28] B. Patel, T.I. Zohdi, Numerical estimation of effective electromagnetic properties for design of
62  30
63
64
65
particulate composites, Mater. Design 94 (2016) 546
553.

[29] H.F. Ma, W.X. Xu, Y. Li, Random aggregate model for mesoscopic structures and mechanical
1
2
3 analysis of fully-graded concrete, Comput. Struct. 177 (2016) 103
113.
4
5
6 [30] P.A. Cundall, O.D.L. Strack, A discrete numerical model for granular assemblies, Gotech. 29 (1)
7
8 (1979) 

9
10
11 [31] M. Michael, F. Vogel, B. Peters, DEM-FEM coupling simulations of the interactions between a tire
12
13
14 tread and granular terrain, Comput. Methods Appl. Mech. Engrg. 289 (2015) 227
248.
15
16
17 [32] K.J. Dong, C. Wang, A.B. Yu, A novel method based on orientation discretization for discrete
18
19
element modeling of non-spherical particles, Chem. Eng. Sci. 126 (2015) 500
516.
20
21
22 [33] K. Wang, W.C. Sun, A semi-implicit discrete-continuum coupling method for porous media based
23
24
25 on the effective stress principle at finite strain, Comput. Methods Appl. Mech. Engrg. 304 (2016)
26
27
28 546
583.
29
30
[34] W.X. Xu, H.S. Chen, Z. Lv, An overlapping detection algorithm for random sequential packing of
31
32
33 elliptical particles, Physica A 390 (13) (2011) 2452
2467.
34
35
36 [35] W.X. Xu, H.S. Chen, Mesostructural characterization of particulate composites via a contact
37
38
39 detection algorithm of ellipsoidal particles, Powder Technol. 221 (2012) 296
305.
40
41
42
[36] A. D iugys, B. Peters, A new approach to detect the contact of two-dimensional elliptical particles,
43
44 Int. J. Numer. Anal. Meth. Geomech. 25 (15) (2001) 14871500.
45
46
47 [37] T.-T. Ng, Shear strength of assemblies of ellipsoidal particles, Gotech. 54 (10) (2004) 659669.
48
49
50 [38] H.P. Zhu, Z.Y. Zhou, R.Y. Yang, A.B. Yu, Discrete particle simulation of particulate systems:
51
52
53
Theoretical developments, Chem. Eng. Sci. 62 (13) (2007) 33783396.
54
55 [39] T. Schanz, P.A. Vermeer, Angles of friction and dilatancy of sand, Gotech. 46 (1) (1996) 145151.
56
57
58 [40] W.X. Xu, H.S. Chen, Q. Duan, W. Chen, Strategy for interfacial overlapping degree in multiphase
59
60
61 materials with complex convex particles, Powder Technol. 283 (2015) 455461.
62  31
63
64
65
[41] C.C. Yang, Effect of the Transition Zone on the Elastic Moduli of Mortar, Cem. Concr. Res. 28 (5)

(1998) 727736.
1
2
3 [42] E.J. Garboczi, D.P. Bentz, Analytical formulas for interfacial transition zone properties, Adv. Cem.
4
5
6 Based Mater. 6 (34) (1997) 99108.
7
8 [43] W.X. Xu, H.S. Chen, W. Chen, L.H. Jiang, Prediction of transport behaviors of particulate
9
10
11 composites considering microstructures of soft interfacial layers around ellipsoidal aggregate
12
13
14 particles, Soft Matter 10 (4)      
15
16
17 [44] W. X. Xu, W. Chen, H.S. Chen, Modeling of soft interfacial volume fraction in composite materials
18
19
with complex convex particles, J. Chem. Phys. 140 (3) (2014) 34704.
20
21
22 [45] W.X. Xu, H. Wang, Y. Niu, J.T. Bai, Insight into interfacial effect on effective physical properties
23
24
25 of fibrous materials. I. The volume fraction of soft interfaces around anisotropic fibers, J. Chem.
26
27
28 Phys. 144 (1) (2016) 014703.
29
30
[46] W.X. Xu, Q. Duan, H. Ma, W. Chen, H. Chen, Interfacial effect on physical properties of
31
32
33 composite media: Interfacial volume fraction with non-spherical hard-core-soft-shell structured
34
35
36 particles, Sci. Rep. 5 (2015) 16003.
37
38
39 [47] B.L. Lu, S. Torquato, Nearest-surface distribution functions for polydispersed particle systems,
40
41
42
Phys. Rev. A 45 (8) (1992) 55305544.
43
44 [48] J.L. Lebowitz, E. Helfand, E. Praesegaad, Scaled particle theory of fluid mixtures, J. Chem. Phys.
45
46
47 43 (3) 
  
48
49
50 [49] E. E. Underwood, Quantitative Stereology, Addison-Wesley, University of California, 1968.
51
52
53
[50] S. Torquato, Random Heterogeneous Materials: Microstructure and Macroscopic Properties,
54
55 Springer-Verlag, New York, 2002.
56
57
58 [51] Z. J. Li, Advanced Concrete Technology, John Wiley & Sons, Inc., New Jersey, 2011.
59
60
61 [52] W. X. Xu, X. Su, Y. Jiao, Continuum percolation of congruent overlapping spherocylinders, Phys.
62  32
63
64
65
Rev. E 94 (3) (2016) 032122.

[53] J. J. Zheng, F. F. Xiong, Z. M. Wu, W. L. Jin, A numerical algorithm for the ITZ area fraction in
1
2
3 concrete with elliptical aggregate particles, Mag. Concr. Res. 61 (2) (2009) 109117.
4
5
6 [54] R. Hill, Elastic properties of reinforced solids: Some theoretical principles, J. Mech. Phys. Solids
7
8 11 (5) (1963) 
9
10
11 [55] Y. T. Xie, D. J. Corr, F. Jin, H. Zhou, S. P. Shah, Experimental study of the interfacial transition
12
13
14 zone (ITZ) of model rock-filled concrete (RFC), Cem. Concr. Compos. 55 (55) (2015) 
15
16
17
18
19
20
21
22
23

24
25
26
27
28 F igure captions:
29
30
31 F ig.1 Schematic view for the components of contact forces of adjacent particles
32
33
34 F ig.2 (a) the regional division for a sample space and (b) periodic boundary compensation scheme for
35
36
37 DEM
38
39 F ig.3 Two-phase PRCs with 400 overlapping elliptical particles realized by DEM: (a) the aspect ratio
40
41
42 of  = 1.5 and the packing density of p = 0.85 and (b) the aspect ratio of  = 2.0 and the packing
43
44
45 density of p = 0.80
46
47
48 F ig.4 Schematic diagram for the coordinate transformations of an arbitrary ellipse
49
50 F ig.5 Topological geometries of the interfacial layer around an ellipse (green curve) constructed by the
51
52
53 parametric equation (blue curve) and the the simplified manner (red curve)
54
55
56 F ig.6 Three-phase PRCs composed of the ellipse packing density of p = 0.8. The basic parameters
57
58
59    

 
 [41], in which the average size of particles is 450 , the average aspect
60
61
62  33
63
64
65
ratio of particles is  = 2.021, two interfacial dimensions of t = 20  (left) and t = 40  (right)

F ig.7 A detailed meshing structure for three-phase PRCs


1
2
3 F ig.8 The interfacial fraction f i versus the particle fraction fp with  = 1, 2 and 3, where t = 40  and
4
5
6 D eq = 450  followed   
 [41]
7
8 F ig.9 The relative elastic modulus E eff / E m for PRCs versus the number of particles in (a)  = 1.0, (b) 
9
10
11 = 1.5, (c)  = 2.0 and (d)  = 2.5, where p = 0.7, D eq = 1 mm, t = 40 m, E p / E m = 2.5 and E i / E m = 0.5
12
13
14 F ig.10 Comparisons of the elastic moduli from the present DEM-FEM model and micromechanical
15
16
17 framework with the experimental data (three experimental data corresponding to the individual fp)
18
19
reported by Yang [41] and with for the predicted moduli of two-phase PRCs without interphase from
20
21
22 the present micromechanical framework. Basic parameters follow Yangs experiment that E m = 20.76
23
24
25 GPa, E p = 80 GPa, m = 0.2, p = 0.21, i = 0.3,  = 2.021, D eq = 450 . (a) t = 20  and E i = 0.43 E m,
26
27
28 (b) t = 40  and E i = 0.65 E m [41]
29
30
F ig.11 The relative effective elastic moduli E eff/ E m versus high packing fraction fp with  = 1, 2 and 3.
31
32
33 Basic parameters are specified to be E m = 20 GPa, E p =2.5 E m, E i = 0.5 E m, m = p = i = 0.2, D eq = 450
34
35
36  and t = 40 
37
38
39 F ig.12 The effective elastic moduli E eff versus E i/ E m with the high packing density fp = 0.8. Basic
40
41
42
parameters are specified to be E m = 20 GPa, E p = 2.5 E m, m = p = i = 0.2, D eq = 450 , t = 40  and
43
44  = 1.5
45
46
47 F ig.13 The relative effective elastic moduli E eff/ E m versus the reciprocal geometric size factor 1/ 
48
49
50 ( D eq/t) with three high packing densities fp = 0.6, 0.7 and 0.8 (solid lines and points represent the
51
52
53
theoretical and numerical results with respect to the individual fp, respectively). Basic parameters are
54
55 specified to be E m = 20 GPa, E p = 2.5 E m, E i = 0.5 E m, m = p = i = 0.2, D eq = 450 , t = 100  and 
56
57
58 = 1.5
59
60
61
62  34
63
64
65
Fig.1
Click here to download high resolution image
Fig.2
Click here to download high resolution image
Fig.3(a)
Click here to download high resolution image
Fig.3(b)
Click here to download high resolution image
Fig.4
Click here to download high resolution image
Fig.5
Click here to download high resolution image
Fig.6(a)
Click here to download high resolution image
Fig.6(b)
Click here to download high resolution image
Fig.7
Click here to download high resolution image
Fig.8
Click here to download high resolution image
Fig.9(a)
Click here to download high resolution image
Fig.9(b)
Click here to download high resolution image
Fig.9(c)
Click here to download high resolution image
Fig.9(d)
Click here to download high resolution image
Fig.10(a)
Click here to download high resolution image
Fig.10(b)
Click here to download high resolution image
Fig.11
Click here to download high resolution image
Fig.12
Click here to download high resolution image
Fig.13
Click here to download high resolution image

You might also like