You are on page 1of 81

MECHANICS OF THREE DIMENSIONAL CELLULAR

STRUCTURES

A Thesis Presented

by

Sahab Babaee

to

Department of Mechanical and Industrial Engineering

in partial fulfillment of the requirements


for the degree of

Master of Science

in

Mechanical Engineering

in the field of

Mechanics and Design

Northeastern University
Boston, Massachusetts

August 2011
ABSTRACT

Three-dimensional cellular materials are ubiquitous in nature and are also used

in a variety of engineering applications, ranging from sandwich structures with low

density cores for structural protection, sound and thermal insulation, and heat transfer

to scaffolds for tissue engineering and regenerative medicine. In many of these

applications, the mechanical properties and structural behavior of the cellular

materials play a key role in regulating the overall function of the system.

In this work, a series of analytical relationships is presented to predict the

mechanical properties and response of open three-dimensional Voronoi tessellation of

faced-centered cubic structures called rhombic dodecahedrons. The cell edge material

was assumed to be elastic-perfectly plastic and the effective mechanical properties of

the cellular structure were related to the cell edge material properties and the relative

density of the cellular structure. Detailed finite element models were carried out to

establish the validity of the analytical models. In the elastic regime, the monodisperse

cellular structure is orthotropic and near-incompressible in all loading directions and

its response is governed by bending deformation of the cell edges. However, the yield

strength of the cellular structure in all loading directions is qual. In the next part of the

work, we studied the role of irregularity in the organization of the cellular structure on

its mechanical properties. The irregularity in the cellular structure organization was

introduced by moving the vertices of a regular cellular structure in three orthogonal

directions by a random value within a predefined range called the ‘Irregularity index’.

At a constant overall relative density, increasing the level of irregularity increases the

effective elastic modulus, and significantly decreases the effective yield strength of

the cellular structure. We also studied the mechanical properties of the open and

I
closed cellular structures tied to rigid plates, in view of the application of cellular

structure as the core construction of sandwich panels. In this case, the cellular

structure is significantly stiffer and its mechanical response is dominated by cell wall

stretching.

II
ACKNOWLEDGEMENT

In the first place, I would like to express my deepest gratitude to professor

Ashkan Vaziri, my thesis advisor, for his guidance, inspiration and unlimited

encouragement during my graduate study at Northeastern university. His supervision

and support from the preliminary to the concluding level enabled me to develop an

understanding of the subject and made him a backbone of this research and so to this

thesis. During these two years, he gives me an extraordinary experiences throught out

this work. He provided me unflinching encouragement and support in various ways.

His truly scientist intuition has made him as a constant oasis of ideas and passions in

science, which exceptionally inspire and enrich my growth as a student, a researcher

and a scientist want to be. I also gratefully acknowledge professor Hamid Nayeb-

Hashemi for his advice, crucial contribution, and constructive comments on this

thesis.

Furthermore, I would like to thank my friends and colleagues at The High

Performance Materials and Structures Lab - Amin Ajdari, Babak Haghpanah Jahromi

and Hamid Ebrahimi - for sharing their enthusiasm and comments on my work. They

helped me out a lot in my research and shared useful knowledge with me. Without

their help and support, I wont’ be able to make this happen.

I dedicate my thesis to my family for their love and support.

III
TABLE OF CONTENTS

1. Introduction to Cellular Structures ...................................................1


1.1. Literature review and overview .........................................................................2

2. Open-cell Rhombic Dodecahedron Cellular Structures ..................7


2.1. Overview ...........................................................................................................8

2.2. Analytical predictions for fundamental mechanical properties .......................12

2.2.1. Elastic modulus and Poisson’s ratio in 2- or 3- direction ................................ 12


2.2.2. Elastic modulus and Poisson’s ratio in 1- direction......................................... 15
2.2.3. Yield strength .................................................................................................. 17
2.2.4. Elastic buckling strength ................................................................................. 18

2.3. Finite element modeling of three-dimensional cellular structure ....................20

2.4. Mechanical properties of open-cell rhombic dodecahedron cellular


structures..... .................................................................................................................24

2.5. Role of irregularity in the structural organization ...........................................29

2.6. Cellular structure tied to rigid plates (sandwich panel configuration) ............34

2.7. Concluding remarks .........................................................................................38

2.8. Acknowledgement ...........................................................................................40

2.9. Appendix A: Mechanical properties of the geometrical unit cell in the 1-


direction...... ................................................................................................................. 41

3. Closed-cell Rhombic Dodecahedron Cellular Structures ..............45


3.1. Overview .........................................................................................................46

3.2. Mechanical properties of cellular structure with regular organization ............49

3.2.1. Theoretical model ............................................................................................ 49


3.2.2. Numerical investigations ................................................................................. 53

3.3. Role of irregularity ..........................................................................................56

3.3.1. Cellular structures with random missing walls................................................ 56


3.3.2. Orderly missing cell walls ............................................................................... 58

IV
References................................................................................................62

V
LIST OF FIGURES

Figure 1: Schematic of two common three dimensional packable open cellular

structures. (a) A rhombic dodecahedron unit cell. (b) A tetrakaidecahedral

unit cell. (c) A tessellated rhombic dodecahedron cellular structure. (d) A

tessellated tetrakaidecahedral cellular structure. ........................................... 3

Figure 2: Some of the key biological cellular materials. Cork, toucan beak and elk

antler have random heterogeneity, while the bamboo has a functionally

graded cellular organization. ......................................................................... 5

Figure 3: Schematic of a regular rhombic dodecahedron structure. (a) A rhombic

dodecahedron unit cell has 12 identical rhombuses with edge length, , and

constant angles, 2  70.530


and 2  109.47
. In this figure, the top

and bottom surfaces (denoted by AJKE and CHMG) are perpendicular to

the 2- axis. The front and back surfaces (denoted by ABCD and KLMN) are

perpendicular to the 3- axis. The line connecting points I and F is in the

direction of 1- axis. (b) A rhombic dodecahedron unit cell in different

views. Due to symmetry in 2- and 3- directions, the top and front views of

the unit cell are identical. (c) A tessellated rhombic dodecahedron cellular

structure with 66 unit cells. The dimensions of the model relative to the cell

edge length of one unit cell are shown in the picture. ................................. 10

Figure 4: (a) Schematic of a rhombic dodecahedron unit cell under uniaxial

compression in 2- direction. The undeformed and the deformed shapes are

shown by dashed and solid lines, respectively. 22 is the displacement of

the nodes located on the top surface of the unit cell. 21 is the

VI
displacement of the nodes located in the mid-plane of the unit cell in 1-

direction due to the uniaxial compression in 2- direction. (b) Free body

diagram of linkage ABC. (c) Free body diagram of cell edge AB. .............. 14

Figure 5: (a) Schematic of a tessellated rhombic dodecahedron cellular structure under

uniaxial compression in 1- direction. (b) Free body diagram of the

tessellated structure. .................................................................................... 16

Figure 6: Schematic of the periodic boundary conditions. .......................................... 22

Figure 7: Mechanical properties of the rhombic dodecahedron tessellated cellular

structure. (a)Schematic of the finite element model of a tessellated structure

under uniaxial compression in 1- direction. (b) Normalized elastic modulus

of the tessellated cellular structure versus its relative density in 1- and 2-

directions. (c) Poisson’s ratios of the cellular structure against its relative

density. (d) Normalized yield strength of the tessellated structure versus the

relative density. The finite element results are presented for loading in both

directions. The analytical prediction is identical for loading in each

direction....................................................................................................... 27

Figure 8: Predicted collapse behavior of tessellated cellular structure for uniaxial

loading in 2- and 1- direction. ..................................................................... 28

Figure 9: Irregular cellular structures with different Irregularity indexes . ............... 31

Figure10: Mechanical properties of an irregular rhombic dodecahedron structure. (a)

and (b) Normalized elastic modulus of irregular cellular structure versus the

Irregularity index  for three different relative densities in loading in 1-

VII
and 2- direction, respectively. The dashed lines show the average results for

each relative density. (c) and (d) Normalized yield strength of the irregular

structure versus the irregularity index in loading in 1- and 2- direction,

respectively. The dashed lines show the average results for each relative

density. ........................................................................................................ 33

Figure 11: Free body diagram of two segments of a rhombic dodecahedron tessellated

structure tied to rigid plates under uniaxial loading in 2-direction. (a) and

(b) Free body diagrams of segments S1 and S2, respectively. The reaction

forces and moments applied to each node are shown. ................................ 35

Figure 12: Normalized elastic modulus of a rhombic dodecahedron tessellated cellular

structure tied to rigid plates versus its relative density. The analytical

estimates for structures with a single cell height and an infinite height are

shown by solid lines. The finite element results for one, two, four, and

twelve cells height are shown with square, rhombus, circle and triangle

markers, respectively................................................................................... 36

Figure 13:(a) Schematic of a rhombic dodecahedron unit cell under uniaxial

compression in 1- direction. The undeformed and the deformed unit cell are

shown by dashed and solid lines, respectively. 11 is the displacement of

node F due to the applied load. 12 is the displacement of the nodes

located on the top surface of the unit cell in 2- direction due to the uniaxial

compression in 1- direction. (b) Free body diagram of the unit cell. All the

reaction forces and moment due to loading 1 are shown. Also the local

coordinate system (1'2'3') used in the analysis is depicted. (c) Normalized

VIII
elastic modulus and yield strength of the rhombic dodecahedron unit cell

versus its relative density in 1- direction. ................................................... 44

Figure 14: Schematic of a closed rhombic dodecaheron cellular structure. (a) A

rhombic dodecahedron unit cell has 12 identical rhombuses plate with edge

length, , thickness, , and constant angles, 2  70.530


and 2 

109.47
. In this figure, the top and bottom surfaces (denoted by AJKE and

CHMG) are perpendicular to the 2- axis. The front and back surfaces

(denoted by ABCD and KLMN) are perpendicular to the 3- axis. The line

connecting points I and F is in the direction of 1- axis. (b) A rhombic

dodecahedron unit cell in different views. Due to symmetry in 2- and 3-

directions, the top and front views of the unit cell are identical. (c) A

tessellated rhombic dodecahedron cellular structure with 1367 unit cells.

The dimensions of the model as a function of the cell edge length of a unit

cell are shown in the picture. ...................................................................... 48

Figure 15: Schematic of a rhombic dodecahedron unit cell under uniaxial compression

in 2- direction. (a) Front view of deformed configuration of the unit cell

under uniaxial compression force F in 2- direction. The undeformed and

the deformed shapes are shown by solid and dashed lines, respectively. (b)

and (c) The deformed configuration of the traingular and rhombus plates. 50

Figure 16: Normalized elastic modulus and yield strength of the closed rhombic

dodecaheron cellular structure as a linear function of its relative density in

2- direction. The error between simulation and analytical results stems from

that the analytical soluton is calculated based on the model with infinite

IX
height while the finite element results are concerned with the model shown

in Fig. 14(c). ................................................................................................ 52

Figure 17: Effect of random missing cell walls on elastic modulus of the closed

rhombic dodecaheron cellular structure. The results are shown for the

model with three different relative densities in the terms of . 2 is the

effective elastic modulus in the 2- direction that is normalized in respect to

the effective elastic modulus of its regular cellular structure counterpart .

For each value of , we ran the simulation for three different models. Since

the results are very close, we just showed the results by one point. ........... 57

Figure 18: Effect of standard deviations, , on elastic modulus of the closed rhombic

dodecaheron cellular structure. (a) Three irregular cellular structures with

different standard variations. For   0 all the cell walls are removed from

the middle of the structure.   1 is called standard normal distribution

and the cell walls are removed according to that function. For   ∞ the

cell walls are removed uniformly (the same numbers of cell walls are

removed randomly from each row). (b) Normalaized elastic modulus of the

structure VS. standard deviasion for different percentage of missing cell

walls, . The relative density and mean value are kept fixed, 10% and 0,

respectively. Also normal distribution functions with different standard

deviasions are plotted. ................................................................................. 60

Figure 19: Normalaized elastic modulus of the closed rhombic dodecaheron cellular

structure VS. standard deviasion for different mean values,  (each mean

value is shown with the specific color). The relative density and percentage

X
of missing cell walls are kept fixed, 10%. The normal distribution functions

for   1 and different mean values are plotted inside the figure. ............ 61

XI
1. Introduction to Cellular

Structures

1
1.1. Literature review and overview

Three-dimensional cellular materials are ubiquitous in nature and are also used

in a variety of engineering applications, ranging from sandwich structures with low

density cores for structural protection [1-6], sound and thermal insulation [7-8], and

heat transfer [9-15] to scaffolds for tissue engineering and regenerative medicine [16-

18]. In many of these applications, the mechanical properties and structural behavior

of the cellular materials play a key role in regulating the overall function of the

system. In this context, the mechanics of two-dimensional cellular structures have

been studied extensively using theoretical approaches, computational models and

robust experiments [19-25]. These studies include investigating the role of structural

organization and hierarchy [19, 26-27], and heterogeneity and defects (e.g. missing

cell edges or cell clusters) on the behavior of cellular structures [28-35].

Studying the mechanical behavior of three-dimensional cellular structures is

inherently more challenging. A limited class of 3D base cells (the triangular, rhombic

and hexagonal prisms, the rhombic dodecahedron, and the tetrakaidecahedron) can be

packed together to generate a monodisperse cellular structure [19, 36-39]. The most

prevalent models of cellular solids are generated by Voronoi tessellation of

distributions of “seed-points” in space. Voronoi tessellation of a body-centered cubic

lattice (the tetrakaidecahedral cellular structures) and Voronoi tessellation of a face-

centered cubic lattice (the rhombic dodecahedron cellular structures) are two common

3D cellular structures that the latter was studied in this work. These two structures are

shown in Fig. 1. The available literature on the mechanics and material properties of

3D cellular structures is generally focused on the mechanics of tetrakaidecahedral

cellular structures, which have a unit cell with six square and eight hexagonal faces

[40-42]. The monodisperse open-cell tetrakaidecahedral cellular structures is isotropic

2
and incompressible, with the effective elastic modulus ~ 0.63*density2 and the

ultimate strength ~ 0.22*density 1.5 in all basic directions of loading [40, 42-47].

Figure 1: Schematic of two common three dimensional packable open cellular structures. (a)
A rhombic dodecahedron unit cell. (b) A tetrakaidecahedral unit cell. (c) A tessellated
rhombic dodecahedron cellular structure. (d) A tessellated tetrakaidecahedral cellular
structure.

Cellular structures with rhombic dodecahedron structural organization are

observed in nature. Honeybees use the geometry of rhombic dodecahedron to form

honeycomb from a tessellation of cells each of which is a hexagonal prism capped

with half a rhombic dodecahedron. Some minerals such as garnet (a semi-precious

3
stone) form a rhombic dodecahedral crystal habit. The rhombic dodecahedron also

appears in the unit cells of diamond and diamondoids [48]. Recently, the formation of

rhombic dodecahedral particles of C70 has been also observed [49].

The cellular structurs found in biological systems generally have strongly non-

uniform and heterogenous organization. Fig. 2 shows some examples of this

heterogenous cellular materials.

Heterogenous organization can affect the mechanical behaviour of the cellular

structure significantly.The deformation of the regular closed rhombic dodecaheron

cellular structure (generally closed-cell foams) is governed by cell wall stretching and

the elastic modulus and yield strength are scaled linearly as a function of relative

density. But several types of irregularities can produce bending deformations which

yields to considerable reduction in the mechanical properties. For instance, curvature

in the cell walls, fractured cell walls and irregularities in the cell shapes induce

bending but random cell geometry and non-uniform cell wall thickness do not

produce bending [50].

4
Figure 2: Some of the key biological cellular materials. Cork, toucan beak and elk antler have
random heterogeneity, while the bamboo has a functionally graded cellular organization.

For 3D cellular solids, the effect of geometrical defects on the fundamental

properties of open and closed cell foams has been investigated by several authors

using finite element and experimental approaches [30, 51-54]. The analytical

equations and formulas are mainly based on fitting curves on the emperical or the

numerical datas [53, 55]. The experimental studied were carried out on available

sythetic materials with irregular structural organization. We should emphasize that

thre are no liturature about analytical calculations for irregular cellular structure.

5
Most of these studies have considered tetrakaidecahedron cellular structure as

the initial configuration. These studies inclyding exploring the influence of non-

uniform cell shape , non-uniform cell wall thickness [56] and wavy distortion of cell

walls [57] on the elastic properties of 3D tetrakaidecahedron cellular structure. Most

cellular structures have more than one source of irregulaity (e.g. combination of non-

uniform cell shape and wall thickness). These irregularities could interactively control

the overall structural behavior of the cellular structure. For example for low density

open cell tetrakaidecahedron cellular structures with non-uniform cell shape and wall

thickness, the combined effect of cell shape and thickness irregularity on the elastic

modulus of open cell tetrakaidecahedron cellular structures have been studied by [28].

They have shown that for the low density cellular structure, increasing the

irregularities increases the elastic moduli but for the high density cellular structures,

influence of two co-existing irregularities is reversed and the elastic modulus

decreases.

6
2. Open-cell Rhombic

Dodecahedron Cellular Structures

7
2.1. Overview

A series of analytical relationships is presented to predict the mechanical

properties and response of open three-dimensional Voronoi tessellation of faced-

centered cubic structures called rhombic dodecahedrons. The cell edge material was

assumed to be elastic-perfectly plastic and the effective mechanical properties of the

cellular structure were related to the cell edge material properties and the relative

density of the cellular structure. Detailed finite element models were carried out to

establish the validity of the analytical models. In the elastic regime, the monodisperse

cellular structure is orthotropic and near-incompressible in all loading directions and

its response is governed by bending deformation of the cell edges. However, the yield

strength of the cellular structure in all loading directions is qual. In the next part of the

work, we studied the role of irregularity in the organization of the cellular structure on

its mechanical properties. The irregularity in the cellular structure organization was

introduced by moving the vertices of a regular cellular structure in three orthogonal

directions by a random value within a predefined range called the ‘Irregularity Index’.

At a constant overall relative density, increasing the level of irregularity increases the

effective elastic modulus, and significantly decreases the effective yield strength of

the cellular structure. We also studied the mechanical properties of the cellular

structure tied to rigid plates, in view of the application of cellular structure as the core

construction of sandwich panels. In this case, the cellular structure is significantly

stiffer and its mechanical response is dominated by cell wall stretching.

In this chapter, we studied the mechanical properties of an open-cell rhombic

dodecahedron cellular structure. We showed that this structure is orthotropic and

incompressible with its material properties being dominated by the bending of the cell

8
edges and its effective elastic modulus ~ density2 in three loading directions. The unit

cell of the structure is a space-filling convex polyhedron called a rhombic

dodecahedron (also called a rhomboidal dodecahedron). The unit cell of the structure

is shown in Fig. 3(a) and has 12 identical rhombic faces with 24 edges and 14

vertices. Each face of a rhombic dodecahedron is a rhombic, with angles, 2 

2   √2  70.53
and 2  2  √2  109.47 . The volume of a unit cell is

!  16# /3√3, where  is the edge length . For a square cell edge cross sectional

area, % & %, the relative density (volume fraction) of the rhombic dodecahedron unit

cells, '(  24% ) /!  9 √3 % ) /2) . For a three-dimensional rhombic dodecahedron

cellular structure of infinite size, each cell edge is shared by three adjacent unit cells,

so the effective relative density of a three-dimensional tessellated cellular structure,

denoted by ', is '  '( /3  3 √3 % ) /2) . Fig. 3(c) shows an example of a rhombic

dodecahedron cellular structure obtained by packing 66 rhombic dodecahedron unit

cells and has 864 identical cell edges. This monodisperse structure is the Voronoi

tessellation of the face-centered cubic lattice (i.e. the Voronoi tessellation of a face-

centered cubic lattice gives rhombic dodecahedron cellular structures). In the

coordinate system shown in Fig. 3, 2- and 3- axes are perpendicular and normal to the

top and front surfaces of rhombic dodecahedron in the outward direction. Cellular

structures with a similar structural organization are observed in nature. Honeybees use

the geometry of rhombic dodecahedron to form honeycomb from a tessellation of

cells each of which is a hexagonal prism capped with half a rhombic dodecahedron.

Some minerals such as garnet (a semi-precious stone) form a rhombic dodecahedral

crystal habit. The rhombic dodecahedron also appears in the unit cells of diamond and

diamondoids [48]. Recently, the formation of rhombic dodecahedral particles of C70

has been also observed [49].

9
dodecahedron unit cell has 12 identical rhombuses with edge length, , and constant angles,
Figure 3: Schematic of a regular rhombic dodecahedron structure. (a) A rhombic

2  70.53
and 2  109.47 . In this figure, the top and bottom surfaces (denoted by
AJKE and CHMG) are perpendicular to the 2- axis. The front and back surfaces (denoted by
ABCD and KLMN) are perpendicular to the 3- axis. The line connecting points I and F is in
the direction of 1- axis. (b) A rhombic dodecahedron unit cell in different views. Due to
symmetry in 2- and 3- directions, the top and front views of the unit cell are identical. (c) A
tessellated rhombic dodecahedron cellular structure with 66 unit cells. The dimensions of the
model relative to the cell edge length of one unit cell are shown in the picture.

In section 2.2., we present analytical models to calculate the effective elastic-

plastic properties of a rhombic dodecahedron unit cell and periodic tessellated cellular

structure. The cell edge material behavior was taken as elastic-perfectly plastic with

elastic modulus, * , yield strength, +, and Poisson’s ratio, -. Our analytical models

10
give estimates of the effective elastic moduli, Poisson’s ratios, yield strengths and

buckling strengths for loading along three orthogonal axes as shown in Fig. 3. In

section 2.3., we constructed finite element models of the cellular structures. The

comparison between the finite element results and the analytical predictions are

presented in section 2.4.. In Section 2.5., we developed finite element models of

irregular rhombic dodecahedron cellular structures, which were subsequently used to

investigate the role of irregularity on the basic mechanical properties of cellular

structures. The irregularity was induced in the cell edge organization of the structure

by randomly moving the vertices of the regular structure in three directions to create

cellular structures with different cell sizes (polydisperse foams). Finally, in section

2.6., we analyzed the mechanical properties of the cellular structures attached to two

rigid plates. This part of the work was carried out in view of recent interests in

developing novel lightweight and multifunctional sandwich structures with low

density core constructions [58-59]. We will show that a cellular structure attached to

rigid plates is significantly stiffer compared to the counterpart structure with periodic

boundary condition and its effective elastic modulus ~ density. Conclusions were

drawn in section 2.7..

11
2.2. Analytical predictions for fundamental mechanical properties

In this section, we derive analytical relationships for the effective mechanical

properties of an open rhomboidal dodecahedron unit cell and tessellated cellular

structure with periodic boundary conditions using fundamental concepts of mechanics

of materials (assuming small deformations). . and +/ denote the effective elastic

modulus and effective yield strength of the cellular structure in three orthogonal

directions (i = 1, 2, 3). In each loading direction, we also obtained the Poisson’s ratios

in the two directions normal to loading. The Poisson’s ratio is denoted by 0.1 , where i

is the loading direction and j is the direction normal to loading. (e.g. for loading in i =

1, we calculated 0) , 0# ). It should be noted that the cellular structure under study

has three orthogonal planes of symmetry and is orthotropic. Due to symmetry, the

mechanical properties in the 2 and 3 directions are identical (i.e. )  # , +2  +3 ,

0)  0# , 0)  0# , 0)#  0#) ).

2.2.1. Elastic modulus and Poisson’s ratio in 2- or 3- direction

Fig. 4(a) shows the deformed configuration of a rhombic dodecahedron unit

cell with twenty four cell edges (beams) subjected to uniaxial compression, 5) in 2-

direction. To obtain the effective elastic modulus of the unit cell, we calculated the

relative displacement of the top and bottom rhombic faces, as the unit cell is subjected

to uniaxial compression imposed by two rigid plates at the top and bottom of the unit

cell. Eight cell edges that construct the top and bottom rhombic faces are considered

to be in contact with the rigid plates, and thus, do not contribute to the mechanical

response of the unit cell. Moreover, the four cell edges located in the mid-plane cell of

the unit cell are neither bent nor stretched during compression and only undergo

12
translational rigid body motion in the 1- direction (and thus do not contribute in

stiffness). Six linkages, denoted by ABC, EFG, KLM, KNM, JIH and ADC in Fig.

4(a), have equal contribution in the mechanical response of the unit cell in 2- direction

and thus, it is adequate to just analyze the response of one linkage (e.g. ABC shown in

Fig. 4(b)). This pair of cell edges is subjected to force, 6  5) /6 in 2- direction and

moment M, which tends to bend the cell edges. The moment can be calculated using

Castiglione's theorem, 7  6 cos /2. The deflection of the unit cell in 2- direction,

can be obtained using Euler-Bernoulli beam theory. Neglecting the shear strain

energy, gives ))  6# ; ) /12* < = 6;>) /* ?, where < and ? denote the

second moment of inertia and area of the beam cross section, respectively, and

 54.735° . The strain in 2- direction is A))  )) /;> and the applied stress in

the same direction is ) =5) /4) sin 2 , where the effective area is assumed at the

mid-height of the unit cell. The effective elastic modulus of the unit cell parallel to 2-

direction is )( =) ⁄A)) , which gives,

)( 27√3 3
 E  ') F ') /√3 F 0.58')
*  24)
3√3 = '
< = ?
(1)

The cell edges mainly deform in bending, and the contribution of the

stretching stiffness of the cell edges in the overall stiffness is negligible.

For a three-dimensional rhombic dodecahedron cellular structure, each cell is

shared amongst three unit cells and the effective elastic modulus decreases by

increasing the number of unit cells. The lower limit for the elastic modulus of a

13
Figure 4: (a) Schematic of a rhombic dodecahedron unit cell under uniaxial compression in 2-

respectively. )) is the displacement of the nodes located on the top surface of the unit cell.
direction. The undeformed and the deformed shapes are shown by dashed and solid lines,

) is the displacement of the nodes located in the mid-plane of the unit cell in 1- direction
due to the uniaxial compression in 2- direction. (b) Free body diagram of linkage ABC. (c)
Free body diagram of cell edge AB.

tessellated cellular structure, ) , is 2/3 of the elastic modulus of the unit cell, which is

achieved for a cellular structure of infinite size. In this case,

) 2
F ') F 0.38 ')
* 3√3
(2)

14
To calculate 0) , we applied the Castiglione's theorem to find the

displacement in 1- direction as the three-dimensional tessellated cellular structure is

subjected to compression in 2- direction. This gives, )  6# sin 2 /48* < H

6 sin 2 /4* ?. The strain in the 1- direction due to uniaxial loading in 2- direction

is A)  ) /; and the Poisson’s ratio, 0)  HA) /A)) . By neglecting the axial

deformation terms for )) and ) (i.e. the second terms in the above equations),

0)  1. Similar calculation for the Poisson’s ratio in 3- direction under loading in 2-

direction yields 0)#  0.

2.2.2. Elastic modulus and Poisson’s ratio in 1- direction

The analytical solution for the elastic properties of a rhombic dodecahedron

unit cell shown in Fig. 4(a) subjected to uniaxial compression in 1- direction is

presented in Appendix A. This geometrical unit cell cannot be used to obtain the

mechanical properties of the tessellated cellular structure in direction- 1, as its

deformation does not represent the response of a three-dimensional rhombic

dodecahedron cellular structure of infinite size. For a tessellated structure, the

analytical solution will be different from the geometrical unit cell and is remarkably

straightforward. The tessellated structure shown in Fig. 5(a) can be divided into

segments shown in Fig. 5(b) with identical deformation. Thus, the analysis of the

segment shown in Fig. 5(b) is sufficient. In Fig. 5(b), I  5 /4   ? /4, where

?  4) ;>) is the effective cross-section area in 1- direction, and  is the

applied stress in 1- direction. The four cell edges shown in Fig. 5(b) have equal values

of strain energy and thus, the total strain energy of the segment is, J  4 K
7L ) /
M

2* < NO* = 4 K


P L ) /2* ? NO* , where 7L  H7 = √2IO* /2√3 and P L  HI/√3
M

and 0 Q O* Q  is the distance from one end of the cell edge as shown in Fig. 5(b).

15
Analogous to the calculation presented in Appendix A for the rhombic dodecahedron

unit cell, minimizing the total strain energy in respect to 7, gives 7  √2IO* /2√3

and application of the Castiglione's theorem gives   I# /9* <. Thus, the

effective elastic modulus of a three-dimensional rhombic dodecahedron cellular

structure in 1-direction is,

 1
 ') F 0.19 ')
* 3√3
(3)

Figure 5: (a) Schematic of a tessellated rhombic dodecahedron cellular structure under


uniaxial compression in 1- direction. (b) Free body diagram of the tessellated structure.

which is half of its effective elastic modulus in the 2- and 3- direction (i.e.   ) /

2).  is 29.14 % smaller than the effective elastic modulus of the unit cell in the

same direction, which is obtained in Appendix A. The Poisson's ratio, 0) and 0# for

a tessellated rhombic dodecahedron cellular structure can be obtained using the basic

relationship for an orthotropic material, (0) /  0) /) , 0# /#  0# / R, which

gives 0#  0)  0.5. It should be noted that 0) and 0# can be also calculated

using Castiglione's theorem, which gives the same result. The volume change per unit

16
volume of the cellular structure in this loading condition is equal to the summation of

the strains in three directions, which is equal to  S1/ = 0) /) = 0# /# R  0.

This shows that the cellular structure is incompressible under this loading condition.

A similar conclusion is valid for loading in 2- and 3- directions.

2.2.3. Yield strength

The proportional limit (elastic limit) stress of a unit cell, denoted here by TU. -

where i is the direction of loading -, is the maximum stress where the relationship

between the stress and strain (or equivalently force and displacement) is linear. The

nonlinear response initiates when the stress in one point of the cellular structure

reaches the yield stress of the cell edge material, +, . For a unit cell shown in Fig.

4(b), the maximum axial stress in the cell edges can be estimated as V)  7%/2< =

6;> /?, where 7  6 cos /2. The second term (contribution of P) is negligible

compared to the first term for a low density cellular structure. Neglecting the axial

term and equating V) = +, , gives TU) /+*  3√6 % # /8#  √3/6 '2 . The yield
3
W

strength of the unit cell in 2- direction, +2 ( , is approximately reached when the

bending moment reaches the fully plastic moment of the cell edge cross-section. In

our study, the cell edges have a square cross-section and the yield strength is 1.5 times

the proportional limit stress (by neglecting the axial term). Thus,

+2 ( 9√3 % # W√3 #


  ') F 0.33'.X
+* 8√2 # 4
(4)

For an infinite size tessellated structure, the yield strength in 2- direction is

denoted by +2 and is 2/3 of the yield strength of the unit cell,

17
+2 3√6 % # √3 #
W

  ') F 0.22'.X
+* 8 # 6
(5)

For a tessellated rhombic dodecahedron cellular structure in 1- direction, the

maximum axial stress happens in the two ends of each cell edge (beam), V 

7%/2< = I; /? . By neglecting the effect of the axial term, equating V  +,

gives TU /+*  √6 % # /4#  √3/9 '2 . Thus, the yield strength of the tessellated
3
W

structure in 1- direction, + , can be estimated from:

+ 3√6 % # √3 #
W

  ') F 0.22'.X
+* 8 # 6 (6)

Interestingly, while the cellular is orthotropic in the elastic regime, it has equal

yield strength in all loading directions and it is almost equal the yield strength of the

open-cell tetrakaidecahedral cellular structures.

2.2.4. Elastic buckling strength

The critical buckling load of the unit cell shown in Fig. 4(a) in 2- direction can

be obtained by using classical Euler’s buckling theory and applying the appropriate

boundary conditions for the pair of cell edges shown in Fig. 4(b). In the analysis, the

transverse force 6; and the moment 6 cos ⁄2 cause a pre-buckling transverse

deformation of the cell edge and the axial load of the cell edge is 6;> – See Fig.

4(c). The non-trivial solution for the general equation of the cell edge transverse

displacement gives the critical buckling load of the cell edge, 5YZ2  66YZ 

) [ ) * % E /2) ;> and YZ2 /*  ) [ ) % E /16E ; ;>) . For  54.735° and

18
  1 (i.e. first buckling mode), the critical elastic buckling stress of the unit cell,

denoted by YZ2 ( can be estimated from:

YZ2 ( 3√3[ ) % E [)
  ') F 0.24')
* 32 E 24√3
(7)

For a tessellated rhombic dodecahedron structure, the critical buckling stress,

denoted by YZ2 , is 2/3 of the critical buckling stress of the unit cell (similar to the

elastic modulus):

YZ2 √3[ ) % E [)
  ') F 0.16')
* 16 E 36√3
(8)

The critical buckling load of the tessellated cellular structure in 1- direction

shown in Fig. 5(b) is 5YZ\  π) * % E /3) ; and the critical buckling stress, denoted

by YZ\ , can be estimated from:

YZ\ √6π) % E [)
  ') F 0.22')
* 16 E 18√6
(9)

19
2.3. Finite element modeling of three-dimensional cellular structure

In this section, we developed finite element models of both unit cell and

tessellated cellular structures and used them to establish the validity of the analytical

models presented in section 2.2.. The simulations were carried out using the finite

element package Abaqus (SIMULIA, Providence, RI). The boundary condition for the

unit cell model is straight forward and identical to the boundary condition assumed in

the analytical investigations. For loading in each direction, two rigid plates were

attached to the two opposite ends of the cellular structure and were displaced towards

each other in the simulations. For a tessellated cellular structure, periodic boundary

conditions were applied in both directions normal to the loading direction to avoid the

influences of the model boundaries on the simulation results. To generate the periodic

boundary conditions in a three dimensional structural model, the opposite boundary

planes of the model must maintain the same shape during the deformation, as shown

schematically in Fig. 6. For our model (Fig. 3(c)), the periodic boundary condition

was applied in both 2- and 3- directions, when investigating the properties of the

structure in the 1- direction. Similarly, for analyzing the mechanical properties in the

2- direction, the period boundary condition was applied in 1- and 3- directions. To

define the above description mathematically, the following relationships have to be

specified in the finite element models for each pair of nodes located on the opposite

boundary planes:

^ _  ^ _
` a

(10)

b^ _ H b^ _  Sb^ RZcd H Sb^ RZcd e ,   1, 2, 3


` a _` _a

20
where ef and e are the opposite boundary planes of the cellular structure (e.g.

plane 1f and 1 , shown in Fig. 6). ^ _ and ^ _ are the rotation angles and b^ _
` a `

and b^ _ are the displacements of the nodes on ef and e planes in n direction.


a

Sb^ RZcd _ and Sb^ RZcd _ are the displacement of a pair of arbitrary reference points
` a

on the opposite boundary planes (e.g. g and gh in 1 and 1' planes, shown in Fig. 6).

The second equation implies that the difference of displacements in all three

directions must be equal for all pair nodes located on the opposite boundary planes

[41]. Similar to the simulations for the unit cell, two rigid plates were attached to the

two opposite ends of the cellular structure in the loading direction and were displaced

towards each other in the simulations. The cell edges were allowed to move relative to

the flat plates with no friction and can freely expand in the lateral directions.

The elastic properties and yield strength of the cellular structure were

calculated from the force-displacement response of the structure in each basic loading

direction. The effective elastic modulus is the initial slope of the response: the

Poisson’s ratios were calculated by dividing the negative value of lateral strain by the

21
Figure 6: Schematic of the periodic boundary conditions.

axial strain. The yield strength was obtained by plotting the stress-strain curve of the

structure and finding a point at which there is an increase in strain with no increase in

stress (the beginning of the plateau region of the curve). The subspace eigensolver

method was employed to estimate the critical elastic buckling loads of the unit cell in

each loading direction. In all calculations, we assumed the cell edge material to be

linear elastic–perfectly plastic with elastic modulus, *  70 i6, yield strength,

+,  130 76 and Poisson’s ratio, -  0.3. It should be noted that all the results are

expressed in the non-dimensional form and are independent of the values of the

material properties used in the calculations. The cell edges were meshed using a

standard Timoshenko beam element (element type B31 in Abaqus) that uses linear

interpolation (2-node linear beam) and allows for transverse shear deformation. A

mesh sensitivity analysis was performed to insure that the result is not sensitive to the

22
mesh size. The static general solver with general (standard) contact condition

available in Abaqus was used in the calculations.

23
2.4. Mechanical properties of open-cell rhombic dodecahedron cellular

structures

Fig. 7(a) shows the finite element model of a cellular structure subjected to

loading in 1- direction. Fig. 7(b) shows the effective elastic moduli of the tessellated

cellular structure in 1- and 2- directions as a function of the relative density of the

tessellated cellular structure, '. In the finite element calculations, the relative density

of the cellular structure was varied by changing the size of the square cross-section of

the cell edges, %. In the same figures, we have also plotted the theoretical estimates of

the effective elastic moduli in two directions from Eqs. (2) and (3). These analytical

estimates are based on considering only the bending deformation of cell edges. For

cellular structures with a low relative density, the finite element and theoretical results

are in good agreement, as bending is the dominant deformation mechanism. By

increasing the relative density of a cellular structure, the contribution of axial

deformation of cell edges in its overall stiffness becomes more significant which

results in the difference between the finite element and numerical solution.. Similar

comparison is made in Fig. 7(c) for the three different Poisson's ratios 0# , 0)# , and

0) of the tessellated cellular structure, which show good agreement between the

analytical predictions and finite element results. The results suggest that the cellular

structure behaves as a near-incompressible orthotropic material. Finally, in Fig. 7(d),

we compared the finite element results and analytical predictions for the yield strength

of the tessellated cellular structure in 1- and 2- directions. The results show good

agreement between the finite element and analytical results. The yield strength

obtained from the finite element analysis is slightly higher than the analytical

predictions.

24
To examine whether Euler buckling occurs prior to yield, we compared the

critical buckling stress and the yield strength of the structure under loading in each

direction. For loading in the 2- direction, using the buckling and yield stresses

equations (Eqs. (8) and (5), respectively) , gives the following relationship between

the cell edge material yield strength and elastic modulus for the condition where Euler

buckling and plastic collapsing occur simultaneously: +, /*  0.72 '
.X . At each

relative density, yield is the dominant mode for lower values of +, /* and buckling

is dominant for higher values of +, /* . The same approach for loading in 1-direction,

yields +, /*  1.02 '


.X. In Fig. 8, we have plotted these critical limits and showed

the predicted collapse mechanism in each region of the plot for each loading direction.

For almost all open-cell tessellated rhombic dodecahedron structures, yielding

proceeds Euler buckling since +, /* is generally < 0.01.

25
26
Figure 7: Mechanical properties of the rhombic dodecahedron tessellated cellular structure.
(a) Schematic of the finite element model of a tessellated structure under uniaxial
compression in 1- direction. (b) Normalized elastic modulus of the tessellated cellular
structure versus its relative density in 1- and 2- directions. (c) Poisson’s ratios of the cellular
structure against its relative density. (d) Normalized yield strength of the tessellated structure
versus the relative density. The finite element results are presented for loading in both
directions. The analytical prediction is identical for loading in each direction.

27
Figure 8: Predicted collapse behavior of tessellated cellular structure for uniaxial loading in 2-
and 1- direction.

28
2.5. Role of irregularity in the structural organization

Numerical simulations of the mechanical properties of three-dimensional

open-cell structures with regular, irregular and random structural arrangements have

been carried out by [28-30]. [55] used a finite element method to predict the Young’s

modulus and Poisson's ratio of four realistic random models of isotropic open cellular

solids. They proposed different relationships for low density and high density open-

cell cellular structures and compared their results with experimental data. In other

study, the tensile elastic properties of a regular dodecahedron with pentagonal faces

was studied to gain insight into the mechanical behavior of lungs [60-61]. In this part

of the study, we developed finite element models of irregular 3D rhombic

dodecahedron cellular structure to study the role of irregularity on the mechanical

properties of tessellated cellular structures. Specifically, we developed models with

irregular cellular organizations by perturbing the locations of the vertices of rhombic

dodecahedron structure which leads to models with various cell sizes (polydisperse

foam), Fig. 3(c). In developing the models, the vertices of the cellular structure

located on the boundaries were fixed and the periodic boundary conditions were

applied in two directions in each model as described in the previous section. In

generating each model of the irregular structure, each vertices of a regular structure

not located on the boundary were perturbed randomly according to:

jk1  j1 = l1 , mg n  1,2,3 (11)

where j1 and jk1 are coordinates of the vertices in the initial regular and final irregular

structures, respectively. 0 Q  Q 1 is a parameter used to quantify the degree of cell

shape irregularity, and is called the ‘Irregularity index’, and H1 o l1 o 1 is a

normally distributed random number. The geometrical characteristics of the cellular

29
structure such as the number of edges, vertices, and cells remain the same in regular

and irregular structures, but the rhombuses of the structure become irregular

quadrilaterals. We developed a MATLAB (Mathwork, Inc., Natick, MA) code to

generate models of irregular cellular structure with different Irregularity indexes. The

models were imported into Abaqus and similar to section 3 and mechanical properties

of the models were studied under loading in different directions. Fig. 9 shows three

examples of the developed models with different Irregularity indexes.

In this section, finite element models of irregular cellular structures with

different Irregularity indexes, 0 Q  Q 0.5 were created and analyzed. The

comparison between the mechanical behavior of regular and irregular cellular

structure was made at constant overall relative density. The total length of the cell

edges varies from one model to the other and the thickness was calculated for each

model to keep the overall (average) relative density the same as the regular cellular

structure. For irregular cellular structures with   0.1, 0.2, 0.3, three different models

of the cellular structures were constructed and analyzed. For   0.4 and 0.5, we

analyzed five models since the scatter in the calculated mechanical properties is

relatively large. The estimated effective elastic modulus and yield strength of the

irregular structures were normalized in respect to the effective elastic modulus and

yield strength of their regular cellular structure counterpart (i.e. with the same relative

density). The normalized effective elastic moduli in the 1- and 2- directions are

denoted by k and k) , respectively and the normalized yield strengths were denoted

by kr\ and kr2 , respectively. For a regular cellular structure, all these parameters are

equal to 1.

30
Figure 9: Irregular cellular structures with different Irregularity indexes .

The results are summarized in Fig. 10, where the dashed lines show the

average results of the finite element calculations for cellular structures at each relative

density. The results show a significant decrease in yield strength and an increase in

the effective elastic modulus of the structure by increasing the Irregularity index, . In

general, the role of irregularity on the mechanical properties of the cellular structure is

more pronounced for cellular structure with a low relative density. Moreover, the

mechanical properties in the 2- direction appear to be more sensitive to the

irregularity, compared to the properties in the 1- direction.

31
32
elastic modulus of irregular cellular structure versus the Irregularity index  for three
Figure 10: Mechanical properties of an irregular cellular structure. (a) and (b) Normalized

different relative densities in loading in 1- and 2- direction, respectively. The dashed lines
show the average results for each relative density. (c) and (d) Normalized yield strength of the
irregular structure versus the irregularity index in loading in 1- and 2- direction, respectively.
The dashed lines show the average results for each relative density.

33
2.6. Cellular structure tied to rigid plates (sandwich panel configuration)

We carried out a similar set of analytical and numerical analysis for a rhombic

dodecahedron cellular structure attached to two rigid plates at its two ends (e.g. plates

are normal to the loading direction and the cell edges are tied to the rigid plates). We

assumed that the cellular structure is infinite in both in plane directions. In this case,

the mechanical properties of the structure are height dependent. We derived analytical

estimates of the elastic modulus of the cellular structure in two extreme cases: i) a

cellular structure with single cell height and ii) a cellular structure with an infinite

height. The analytical model, as well as the finite element calculations are discussed

briefly, as many of the details are similar to the previous sections.

Fig. 11 shows two segments of the structure (denoted by S1 and S2 and shown

in Figs. 11(a) and 11(b), respectively) that were analyzed to obtain the analytical

solution for the elastic modulus of a single cell height tessellated cellular structure.

All the nodes of segment S1 can only move in the loading direction (2- direction) due

to symmetry. for segment S2, nodes A, B, C, and D can move only in 2- direction and

node E can move in both 1- and 2- directions. The reaction forces and moments for

each segment can be calculated by minimizing the total strain energy. For both

segments, minimizing the total strain energy of the segment shows that the internal

bending moment at each cross section of all cell edges is zero. Thus, the cell edges

undergo only stretching under loading in 2- direction. Using Euler-Bernoulli beam

theory and neglecting the shear strain energy, the deflection of the structure in 2-

direction can be obtained for each. For S1, ))  3sh/2* ?, and for S2, )) 

3s/* ?, where sh and s are the compression force applied to nodes C and D of S1

and S2, respectively, as shown in Fig. 11.

34
Figure 11: Free body diagram of two segments of a rhombic dodecahedron tessellated
structure tied to rigid plates under uniaxial loading in 2-direction. (a) and (b) Free body
diagrams of segments S1 and S2, respectively. The reaction forces and moments applied to
each node are shown.

The single cell height structure consists of two S1 and two S2 segments. Thus,

the total resisting force of the unit cell is equal to 2s = 2sh, and the effective elastic

modulus of the structure is,

) √3 % ) 1
 F '
* 4 ) 6
(12)

For the tessellated cellular structure with infinite height, the response only

relates to the deformation of S2 and the total resisting force of the unit cell is equal to

4s. The elastic modulus of the structure can be estimated from,

) √3 % ) 1
 F '
* 6 ) 9
(13)

35
Since the structure only deforms in stretching and the cell edges behave

effectively as truss elements and their bending moment and deformation is minimal.

The elastic modulus is scaled linearly as a function of relative density. Finite element

calculations were performed for one cell, two cells, four cells and twelve cells height

structures. Periodic boundary condition was applied in both directions normal to the

loading direction. Fig. 12 shows the theoretical and the finite element results for the

elastic modulus of the cellular structure in 2- direction. The solid lines represent the

analytical estimates for the elastic moduli of single cell and infinite height structure.

The results validate the descending trend of elastic modulus by increasing the height

of the structure. In general, the cellular structure with tied boundary conditions is

much stiffer and stronger compared to the cellular structure with the boundary

condition studied in the previous sections.

Figure 12: Normalized elastic modulus of a rhombic dodecahedron tessellated cellular


structure tied to rigid plates versus its relative density. The analytical estimates for structures
with a single cell height and an infinite height are shown by solid lines. The finite element

36
results for one, two, four, and twelve cells height are shown with square, rhombus, circle and
triangle markers, respectively.

37
2.7. Concluding remarks

We provided analytical estimates for the effective elastic moduli and yield

strength of three-dimensional Voronoi tessellation of faced-centered cubic (FCC)

structure called a rhombic dodecahedron in three different normal loading directions.

Detailed finite element calculations were carried out to establish the validity of the

analytical models. The cellular structure is two times stiffer in 2- and 3- directions,

compared to its stiffness in the 1- direction and is near incompressible in all loading

directions. The yield strength of the cellular structure was identical in all loading

directions, indicating that yielding in 1-direction occurs at a strain that is two times of

the yield strain in 2- and 3- directions. Comparison between the buckling load and

yielding of the cellular structure showed that for almost all open-cell tessellated

rhombic dodecahedron structures, yielding precedes Euler buckling. This would

suggest that the cellular structure has similar energy absorption capacity in different

loading directions if the loading is applied quasi-statically. However under dynamic

loading, the inertia effect and the contact between the cell edges could influence the

energy absorption capacity of the cellular structure in different directions ([6, 35, 58]).

We also extended our study to cellular structures with tied boundary condition

(attached to rigid plates) and showed that the deformation of the cellular structure is

dominated by cell wall stretching. In this case, the elastic modulus of the cellular

structure is a linear function of its relative density and is height dependent. In general,

this cellular structure tied to rigid plates is much stiffer compared to the counterpart

cellular structure with the periodic boundary condition.

The effect of irregularity in cell arrangements was also investigated. This

showed an increase in the effective elastic modulus and considerable decrease in the

38
yield strength of the cellular structure by increasing the level of irregularity in the

structural arrangement of the cell edges of the cellular structure. Our results are in

qualitative agreement with the results provided by [62] on the effects of cell

irregularity on the elastic properties of 3D open-cell random Voronoi foams.

Moreover, the microstructure irregularities in stochastic Voronoi structures is shown

to slightly elevate the effective elastic modulus and decrease the compressive strength

[63]. [31] also show that two-dimensional low density Voronoi structures are stiffer

and have lower yield strength compared to a regular hexagonal honeycomb. Similar

observation for the elastic properties of a two-dimensional irregular honeycomb was

also reported by [32, 64].

39
2.8. Acknowledgement

We thank John W. Hutchinson for fruitful discussions and the anonymous

reviewers for their constructive comments and suggestions. This work was supported

in part by the U.S. Department of Homeland Security (AA, AV) and in part by the

U.S. Air Force Office of Scientific Research under AFOSR YIP grant award, #FA

9550-10-1-0145 (SB, AV), and in part by the Department of Mechanical and

Industrial Engineering at Northeastern University (BHJ, HN). This material is based

upon work supported by the U.S. Department of Homeland Security under Award

Number 2008-ST-061-ED0001. The views and conclusions contained in this

document are those of the authors and should not be interpreted as necessarily

representing the official policies, either expressed or implied, of the U.S. Department

of Homeland Security.

40
2.9. Appendix A: Mechanical properties of the geometrical unit cell in the 1-

direction

Fig. 13(a) shows the deformed configuration of a rhombic dodecahedron unit

cell subjected to uniaxial compression, 5 , in 1- direction. The following linkages -

each consisting of three cell edges connected at one point to each other - have

identical deformation and response under this loading: AEKF, KJAI, CGMF, CHMI,

ABCF, ADCI, KLMF, and KNMI. Thus, we analyzed the deformation of one of the

linkages, as shown in Fig. 13(b). The force exerted to this linkage is   5 /4. The

total strain energy stored in this part of the unit cell is J  Jt = 2Ju , where Jt and

Ju are the strain energy stored in cell edge EF and cell edge AE (or EK), respectively.

J  J S , ) , # , 7 , 7) , 7# , 7E R, where ) and # are the reaction forces

and, 7 , 7) , 7# and 7E are the reaction moments applied at the boundary in Fig.

13(b). Using the Euler-Bernoulli beam theory and neglecting the shear strain energy,

the strain energy of cell edge EF can be estimated from

M7 ) M
Pt )
Jt  v NO = v NO
t3
2* < 2* ?
(14)

where Pt and 7t3 are the axial force and the internal moment around axis 3. From

equilibrium, Pt  H /√3 H √2) /√3 and 7t3  7E H √2 O/√3 = ) O/√3 ,

where x is the distance measured from the end of the cell edge where the external

force is applied (0 Q O Q ). To calculate Ju , we used a local coordinate system

(1'2'3'), where axis 3' is in the direction of the cell edge – see Fig. 13(b). In this local

coordinate, the total strain energy of the cell edge can be estimated from:

41
M7 M7 M7
Pu )
) ) ) M
u\w u2w u3w
Ju  v NO = v NO = v NO = v NO
2* < 2* < 2i* x 2* ?
(15)



where Pu , is the axial force and 7u w , 7u and 7u


\ 2w 3w
are the internal moments

applied to the cross section of the cell edge. From equilibrium, Pu  H /2√3 =

√2# /√3 and 7u w  7) H √2 y/2√3 H # y/√3,


\
7u w  H√27 /√3 =
2

7# /√3 = ) y/2, 7u w  7h /√3 = √27h# /√3, where y is measured from the left
3

end of the cell edge as shown in Fig. 13(b) ( 0 Q y Q ). The reaction forces and

moments are unknowns, which can be calculated by minimizing the total strain

energy in respect to each of them (e.g. zJ /z)  0). Solving the corresponding set

of equations gives:

)   , #  
√) √)
{  ) 

√3 7√6
7    , 7)   
12 72 
(16)

√6 5√6
7#   , 7E   
24 36 

Similar to the calculations presented in section 2.2.1.,  can be obtained

using the Castiglione's theorem. Neglecting the strain energy associated with shear

and axial deformations,   17 # /216* <. The strain in 1- direction is A 

2 / and the applied stress in the same direction is  = 5 /4) ;>) , where the

area is calculated at the middle of the cellular structure, which is a square with side

length 2 sin . The effective elastic modulus of the unit cell parallel to 1- direction,

(   ⁄A , is:

42
( 22 )
 ' F 0.27 ')
* 81
(17)

For a unit cell subjected to uniaxial compression in 1- direction, Fig. 13(a), the

maximum stress due to bending occurs in points A, E, F and K and can be estimated

from V  5 /√6 % # . Equating V  +, , gives TU /+*  3√6 % # /10# 

2√3/15 '2 . Similar to the previous section, the yield strength of the unit cell in 1-
3
W

direction, +\ ( , can be estimated from:

+\ ( 9√6 % # W√3 #


  ') F 0.26'.X
+* 20 # 5
(18)

The comparison between the analytical solution and finite element results for

normalized elastic modulus and yield strength of the rhombic dodecahedron unit cell

in 1- direction is shown in Fig. 13(c).

43
Figure 13: (a) Schematic of a rhombic dodecahedron unit cell under uniaxial compression in

respectively.  is the displacement of node F due to the applied load. ) is the
1- direction. The undeformed and the deformed unit cell are shown by dashed and solid lines,

displacement of the nodes located on the top surface of the unit cell in 2- direction due to the

forces and moment due to loading  are shown. Also the local coordinate system (1'2'3')
uniaxial compression in 1- direction. (b) Free body diagram of the unit cell. All the reaction

used in the analysis is depicted. (c) Normalized elastic modulus and yield strength of the
rhombic dodecahedron unit cell versus its relative density in 1- direction.

44
3. Closed-cell Rhombic
Dodecahedron Cellular Structures

45
3.1. Overview

In this chapter, first, we studied the mechanical properties of a monodispers

cellular structure with closed rhombic dodecahedron cell using analytial methods

based on the concepts of mechanical of materials and finite element analysis. This

monodisperse structure is the Voronoi tessellation of the face-centered cubic lattice

(i.e. the Voronoi tessellation of a face-centered cubic lattice gives rhombic

dodecahedron cellular structures). The unit cell of the structure is a space-filling

convex polyhedron called a rhombic dodecahedron (also called a rhomboidal

dodecahedron). The unit cell of the structure is shown in Fig. 14(a) and has 12

identical rhombic faces with 24 edges and 14 vertices. Each face of a rhombic

dodecahedron is a rhombic, with angles, 2  2   √2  70.53


and 2 

2  √2  109.47 . The volume of a unit cell is !  16# /3√3, where  is the

edge length . Considering each face as a plate with thickness , the relative density

(volume fraction) of the rhombic dodecahedron unit cells, '(  3% ) /!  3 √6/2.

For a three-dimensional rhombic dodecahedron cellular structure of infinite size, each

cell wall is shared by two adjacent unit cells, so the effective relative density of a

three-dimensional tessellated cellular structure with infinite size, denoted by ', is

'  '( /2  33 √6/4 F 1.84 / . Fig. 14(c) shows an example of a closed

rhombic dodecahedron cellular structure obtained by packing 1367 rhombic

dodecahedron unit cells. In the coordinate system shown in Fig. 14, 2- and 3- axes are

perpendicular and normal to the top and front surfaces of rhombic dodecahedron in

the outward direction. Our analytical models give estimates of the effective elastic

moduli and yield strength of the structure. We also constructed finite element models

of the closed rhombic dodecaheron cellular structure and made comparison between

the analytical predictions and the finite element results. The cell edge material

46
behavior was taken as elastic-perfectly plastic with elastic modulus, * , Poisson ratio,

-, and yield strength, +, .

In section 3.3., we developed finite element models of the irregular rhombic

dodecahedron cellular structures to examine the role of irregularities on the elastic

modulus of the cellular structures. The irregularity was induced by removing the cell

faces with random and Gaussian distributions. To find the best statistically

distrubution of missing cell walls, the influences of two cotrolling parameters of the

Gaussian dietribution function, standard deviations () and mean value (), were also

studied. Results and conclusions were presented in this section.

47
dodecahedron unit cell has 12 identical rhombuses plate with edge length, , thickness, , and
Figure 14: Schematic of a closed rhombic dodecaheron cellular structure. (a) A rhombic

constant angles, 2  70.53


and 2  109.47 . In this figure, the top and bottom surfaces
(denoted by AJKE and CHMG) are perpendicular to the 2- axis. The front and back surfaces
(denoted by ABCD and KLMN) are perpendicular to the 3- axis. The line connecting points I
and F is in the direction of 1- axis. (b) A rhombic dodecahedron unit cell in different views.
Due to symmetry in 2- and 3- directions, the top and front views of the unit cell are identical.
(c) A tessellated rhombic dodecahedron cellular structure with 1186 unit cells. The
dimensions of the model as a function of the cell edge length of a unit cell are shown in the
picture.

48
3.2. Mechanical properties of cellular structure with regular organization

3.2.1. Theoretical model

In this section, we provide an analytical estimation for the elastic modulus and

yield strength of a closed rhombic dodecahedron cellular structure in transverse

direction. We assume that the height of the cellular structure is very long and neglect

the effect of boundaries. To calculate the effective properties of the cellular structure,

we estiamted the total potential energy of the cellular structure under prescribed

displacement field due to transverse loading and used the principle of minimum

potential energy. Fig. 15(a) shows a unit cell of the structure and its deformed

configuration under transverse loading in 2-direction. In this schematic figure,  and

) denote the displacements of point A in 2- direction and point B in 1- direction,

respectively. Due to symmetry it is just adequate to analyze one subassembly of the

cellular structure comprising of a triangular plate and a rhombus plate shown in Fig.

15(b,c). To calculate the strain energy of this subassembly, we assumed a bilinear

displacement field in each of the two plates. The displacement fields in 1- and 2-

directions are denoted by b and | and were given by the following relations,

bSO, yR  
=  O = ) y and |SO, yR  
L = L O = )L y, where O and y are the

coordinate axes in 1- and 2- directions and 


,  , ) , 
L , L and )L are constants that

can be obtained by applying the appropriate boundary conditions. For the triangular

plate shown in Fig. 15(b), }bSO, yR|,t  0, }|SO, yR|,u  0, }bSO, yR|u  ) , and

}|SO, yR|t  H and The displacement field can be presented as bSO, yR  √3) O/

and |SO, yR  H √3 y/√2 . The associated strain field of the triangular plate is

calculated as

49
zb‹ )
… Š zO Ž
ˆ z|‹ ˆ √3 1
€‚ƒ  „ r †  zy  H  
…r ‰   √2
ˆzb‹ = z|‹ ˆ
(19)

‡ zy zOŒ 0

Figure 15: Schematic of a rhombic dodecahedron unit cell under uniaxial compression in 2-
direction. (a) Front view of deformed configuration of the unit cell under uniaxial
compression force F in 2- direction. The undeformed and the deformed shapes are shown by
solid and dashed lines, respectively. (b) and (c) The deformed configuration of the traingular
and rhombus plates.

The strain energy of an elastic body can be written,

J  ) K€‚‘ ’“€‚ N! , where ’“ is the stiffness matrix for for an elament with


plane stress condition. The elastic strain energy of the triangular plate can be estimate

from, Jƒ   S ) = ) ) H √2 -  ) R/S1 H - ) R.


√)
” *

50
For the rhombus plate, since the nodes B and F moves in 1- direction and the

nodes A and E moves in 2- direction, we can consider the plate as the line element AB.

The strain along this line due to uniaxial loading in 2- direction is tu  S H •R/

where •, the final length of the line AB, is equal to

–S√2/√3 H  R) = S/√3 = ) R). Neglecting the higher order terms, the strain of

the rhombus plate is

√2 )
tu  S H R
√3 √2
(20)

which gives the elastic strain energy of the rhombus plate as, JZ 

* S ) = ) ) ) H √2  ) R/S1 H - ) R.
)√) 
—

Then using the principle of minimum potential energy, [  J H ˜, where J

is the total elastic strain energy of the unit cell, J  4Jƒ = 4JZ and ˜  5 and

also minimizing the total potential energy, z[/z  0 and z[/ z)  0, the

displacement of the unit cell is calculated (-  0.3),

5
) F 0.778   F
2.54 * 
(21)

The applied stress in the same direction is )  5/4) sin 2 , where the

effective area is assumed at the mid-height of the unit cell (  54.735° ). The

effective elastic modulus of the unit cell parallel to 2- direction is ) = ) ⁄A) where

A)  √3 /√2, which gives,

) 
F 0.55  0.31 '
* 
(22)

51
Cells in cellular structures can collapse by elastic buckling, plastic yielding or

brittle crushing, depending on the nature of cell wall material [65]. For a closed

rhombic dodecahedron cellular structure made of elastic-perfectly plastic materials,

the collapse happens by elastic buckling (very low density cellular structures with

relative density less than 0.02) and plastic yielding (cellular structures with relative

density more than 0.02).

Figure 16: Normalized elastic modulus and yield strength of the closed rhombic dodecaheron
cellular structure as a linear function of its relative density in 2- direction. The error between
simulation and analytical results stems from that the analytical soluton is calculated based on
the model with infinite height while the finite element results are concerned with the model
shown in Fig. 14(c).

The yield strength of the structure, + , is calculated by equating the von Mises

stress to the yield strength of the cell wall material, +, . In the case of plane stress, the

von Mises stress is provide by the equation, ™š  › ) = ) ) H  ) , where  and

) are the stress components in 1- and 2- directions, respectively and they are

52
calculated for the both plates _ shown in Fig. 15(b,c)_ using Eqs. (19), (20), and

€‚  ’“€‚. Upon substitution of these values into von Mises equation and using

Eq. (19) and the von Mises criterion, ™š  +, , the displacements of the structure

that causes yielding of the traingular plate,  + and ) + , are  + /  0.0011 and

) + /  0.00085. These values for the rhombus plate are bigger than the traingular

plate and the yielding is initiated from the traingular plate. Considering the Hook's

law for plane stress conditions, the equivalent critical forces of these plates in 2-

direction is found as 5œZ ƒ  0.3  + * and 5œZ Z  35œZ ƒ /√2. So the effective yield

strength of the structure parallel to 2- direction is +  S25œZ ƒ = 45œZ Z R/4) sin 2 ,

which gives,

+ 
F 0.83 F 0.46 '
+ * 
(23)

3.2.2. Numerical investigations

In this section, we developed finite element models of closed rhombic

dodecahedron tessellated cellular structures. The finite elmenet models were used to

validate the analytical estimates for the effective elastic modulus and yield strength of

the cellular structure obtained in section 3.2.1.. The simulations were conducted using

the Finite Element package Abaqus (SIMULIA, Providence, RI). The model

dimensions are shown in Fig. 14(c) in the terms of edge length L. The model has

composed of 7&10&10 unit cells in 1-, 2- and 3- directions that were packed together

to build a cubic model. The edge length was assumed   √3 unit length. Two rigid

plates were tied to the two opposite ends of the cellular structure in the loading

direction (2- direction). The relative density (volume fraction) of our finite elment

53
models including 8220 faces with thickness t, is ' F 1.80 / which is almost equal

to the periodic model. All cell walls within a model have the same thickness. The cell

wall thickness was determined to obtain certain densities. The cell walls were meshed

using a standard shell element (element type S4R in Abaqus) which is a 4-node,

quadrilateral, first-order interpolation, stress/displacement with reduced integration

and a large-strain formulation. A mesh sensitivity analysis was performed to insure

that the result is minimally sensitive to the mesh size. The static general solver with

general (standard) contact condition available in Abaqus software package was used

in the calculations. In all calculations, we assumed the cell wall material to be linear

elastic–perfectly plastic with elastic modulus, *  70 i6, yield strength, +, 

130 76 and Poisson’s ratio, -  0.3. It should be noted that all the results are

expressed in the non-dimensional form and are independent of the values of the

material properties used in the calculations.

The elastic properties and yield strength of the cellular structure were

calculated from the force-displacement response of the structure in the transverse

direction. The effective elastic modulus is the initial slope of the response. The 0.2 %

offset method was used to determine the yield strength of the models. Fig. 16 shows

the elastic modulus and yield strength of the closed rhombic dodecahedron cellular

structure as a function of relative density which were compared to the analytical

estimation described in section 3.2.1.. The error between simulation and analytical

results of elastic modulus stems from that the analytical soluton is calculated based on

the sandwich panel with infinite height while the finite element results are based on

the finite element model shown in Fig. 14(c). The yield strength of the theory and

finite element simulation are in very good agreement and indicates that the yield

strength is slightly height dependent. We also carried out a parametric study to

54
examine the effect of height on the yield strength and elastic modulus of the structure.

It for the height of the structure more than five unit cells, the yield strength and elastic

modulus are both height independent. But when the height of the structure varies from

five to one, they will increase up to %15.

55
3.3. Role of irregularity

Most of the cellular materials in nature have irregular structural organization.

Trabecular bone, cork, cedar, wood, bamboo and sisal plant are some examples of

natural irregular cellular structures [65-68]. Also most commercially available

metallic foams have some defects in their structures [69]. In this section, we used

finite element method to investigate the effects of missing cell walls with random and

normal (Gaussian) distributions on the effective elastic modulus for closed rhombic

dodecahedron cellular structure. The simulations were carried out in abaqus and the

assupmtions for the cell walls material and elements type are similar to the section

3.2.2..

3.3.1. Cellular structures with random missing walls

Each of our regular models shown in Fig. 14(c) consist of 8220 plates that are

packed together to fill the space and build the regular sandwich panel. To generate the

random missing cell walls models, we need to remove some of the plates randomly. 

is a parameter defined as percentage of removed cell walls to the total cell walls. We

carried out the simulations for   0, 10, 20 and 30 percents in which   0 means

that there is no irregularity in the structure. By increasing the , the percentage of

deleted cell walls increases and the stricture becomes more irregular. For each value

of , three different models of the cellular structures were constructed and analyzed.

The models were made on ANSYS (ANSYS Inc., Canonsburg, PA). MATLAB

(Mathwork, Inc., Natick, MA) sofware was used to generate the random numbers to

rermove the plates from the models and finally they were imported into Abaqus to

conduct the finite element analysis.

56
Fig. 17 shows the effect of randomly missing cell walls on elastic modulus of

the sandwich panles for three relative densities, '  2, 4 and 10 percent. The effective

elastic moduli of the irregular structures were normalized in respect to the effective

elastic modulus of their regular cellular structure counterpart (i.e. with the same

relative density). For a regular structure, this parameter is equal to 1. In Fig. 17 the

dashed lines show the average results of the finite element calculations for cellular

structures at each relative density. As expected, increasing the removed cell walls (),

decreases the elastic modulus of the structure significantly. For example removing

20% of the cell walls,   0.2, results in a reduction of 50% in elastic modulus of the

models with relative density '  2%.

Figure 17: Effect of random missing cell walls on elastic modulus of the closed rhombic

relative densities in the terms of . k) is the effective elastic modulus in the 2- direction that
dodecaheron cellular structure. The results are shown for the model with three different

counterpart . For each value of , we ran the simulation for three different models. Since the
is normalized in respect to the effective elastic modulus of its regular cellular structure

results are very close, we just showed the results by one point.

57
3.3.2. Orderly missing cell walls

In this section the effect of missing cell walls with normal distribution is

investigated on elastic modulus of the closed rhombic dodecaheron cellular structure.

We developed finite element models of closed rhombic dodecahedron structure using

ANSYS (ANSYS Inc., Canonsburg, PA) software. The models were generated by

removing the cell walls according to the normal (Gaussian) distribution function [70],

1 S…¡R2
mSOR   

)¢2
√2[ )
(24)

where  is the standard deviation (height of the curve's peak) and  is the mean value

(position of the center of the peak). The graphs of Gaussian function with different 

and  are plotted inside the figures 18(b) and 19, respectively. As it is shown in Fig.

14(c), the model has 10 rows unit cell along 1- direction. The number of cell walls

that should be removed from each row is determined by the normal distribution

function which has been plotted between -5 to 5. Fig. 18(a) shows three examples of

the developed models with different standard deviation.

- Variation of standard deviation ()

Fig. 18(b) shows the effect of missing cell walls on the elastic modulus of the

structure at different standard deviations () and percentage of missing cell walls (γ).

In these models the mean value, , is assumed to be zero (middle of the structure) and

the standard deviation, , varies from zero to infinity.   0 refers to the case that all

the missed cell walls are removed from the middle and   ∞ means that the cell

walls are removed completely uniform (the same number of cell walls are removed

randomly from each row). The following results can be obtained:

58
1. As expected, increasing the percentage of missing cell walls decreases the

elastic modulus of the tessellated cellular structures significantly at a constant

standard deviation (relative density is kept fixed in all the models and equal to 10%).

2. The minimum elastic modulus of the heterogeneous structures happens

when the standard deviation tends to zero which means that cell walls are removed

from the middle of the structure (see Fig. 18(a)). Then by increasing , the

distribution of missing cell walls get more uniform and we have a sharp increase in

elastic modulus. Where  ¤ 1 there is slight increase in the elastic modulus.

3. For all the heterogeneous cellular structures, three different models were

made and analyzed. In the Fig. 18(b) when 0 Q  Q 5 all the three models have very

close elastic modulus, so just one point is shown in the figure but when standard

deviation increases, the scatter in the calculated elastic modulus becomes relatively

large.

4. For smaller values of γ, the elastic modulus of the structure approaches

more rapidly to the final value. For instance, for γ  5%, where  ¤ 1, the elastic

modulus becomes almost constant and is equal to the 0.85. For γ  10%, where

 ¤ 2, the elastic modulus is almost 0.75 and for γ  20%, where  ¤ 5, the elastic

modulus is 0.55. Also the results indicated that if the distributiion of the missing cell

walls in the heterogeneous structure is uniform, the stiffness of the structure will be

maximum.

59
Figure 18: Effect of standard deviations, , on elastic modulus of the closed rhombic

variations. For   0 all the cell walls are removed from the middle of the structure.   1
dodecaheron cellular structure. (a) Three irregular cellular structures with different standard

function. For   ∞ the cell walls are removed uniformly (the same numbers of cell walls
is called standard normal distribution and the cell walls are removed according to that

standard deviasion for different percentage of missing cell walls, . The relative density and
are removed randomly from each row). (b) Normalaized elastic modulus of the structure VS.

mean value are kept fixed, 10% and 0, respectively. Also normal distribution functions with
different standard deviasions are plotted.

60
- Variation of mean value ()

Effect of random missing cell walls on elastic modulus of the closed rhombic

dodecaheron cellular structure with different mean values () and standard deviasions

() is shown in Fig. 19. By increasing the standard deviasion the elastic modulus of

the structure increases. As standard deviasion tends to infinity, since the cell walls are

removed uniformly, the elastic modulus for the all mean values goes to the same

value of 0.7 (it does not depend on ).

VS. standard deviasion for different mean values,  (each mean value is shown with the
Figure 19: Normalaized elastic modulus of the closed rhombic dodecaheron cellular structure

The normal distribution functions for   1 and different mean values are plotted inside the
specific color). The relative density and percentage of missing cell walls are kept fixed, 10%.

figure.

61
References
1. Wei, Z., Deshpande, V.S., Evans, A.G., Dharmasena, K.P., Queheillalt, D.T.,
Wadley, H.N.G., Murty, Y.V., Elzey, R.K., Dudt, P., Chen, Y., Knight, D. and
Kiddy, K., The resistance of metallic plates to localized impulse. Journal of
the Mechanics and Physics of Solids, 2008. 56(5): p. 2074-2091.

2. Hohe, J. and W. Becker, Effective stress-strain relations for two-dimensional


cellular sandwich cores: Homogenization, material models, and properties.
Applied Mechanics Reviews, 2002. 55(1): p. 61-87.

3. Vaziri, A. and J.W. Hutchinson, Metal sandwich plates subject to intense air
shocks. International Journal of Solids and Structures, 2007. 44(6): p. 2021-
2035.

4. Radford, D.D., McShane, G.J., Deshpande, V.S., Fleck, N.A., Dynamic


Compressive Response of Stainless-Steel Square Honeycombs. Journal of
Applied Mechanics, 2007. 74(4): p. 658-667.

5. Deshpande, V.S. and N.A. Fleck, Energy absorption of an egg-box material.


Journal of the Mechanics and Physics of Solids, 2003. 51(1): p. 187-208.

6. Vaziri, A., Z. Xue, and J.W. Hutchinson, Performance and failure of metal
sandwich plates subjected to shock loading. J. Mechanics of Materials and
Structures, 2007. 2(10): p. 1947–1964.

7. Lu, T.J. and C. Chen, Thermal transport and fire retardance properties of
cellular aluminium alloys. Acta Materialia, 1999. 47(5): p. 1469-1485.

8. Vaziri, A., Z. Xue, and J.W. Hutchinson, Metal sandwich plates with
polymeric foam-filled cores. J. Mechanics of Materials and Structures, 2006.
1(1): p. 95-128.

9. Tian, J., et al., The effects of topology upon fluid-flow and heat-transfer within
cellular copper structures. International Journal of Heat and Mass Transfer,
2004. 47(14-16): p. 3171-3186.

62
10. Lu, T.J., H.A. Stone, and M.F. Ashby, Heat transfer in open-cell metal foams.
Vol. 46. 1998, Kidlington, Royaume-uni Elsevier.

11. Famodu, A.A., et al., Multifunctionality of cellular metal systems. Progress in


Materials Science, 1998. 43: p. 171-221.

12. Kooistra, G.W. and H.N.G. Wadley, Lattice truss structures from expanded
metal sheet. Materials & Design, 2007. 28(2): p. 507-514.

13. Queheillalt, D.T. and H.N.G. Wadley, Pyramidal lattice truss structures with
hollow trusses. Materials Science and Engineering A, 2005. 397(1-2): p. 132-
137.

14. Evans, A.G., et al., The topological design of multifunctional cellular metals.
Progress in Materials Science, 2001. 46(3-4): p. 309-327.

15. Gu, S., T.J. Lu, and A.G. Evans, On the design of two-dimensional cellular
metals for combined heat dissipation and structural load capacity.
International Journal of Heat and Mass Transfer, 2001. 44(11): p. 2163-2175.

16. Hollister, S.J., Porous scaffold design for tissue engineering. Nat Mater, 2005.
4(7): p. 518-524.

17. Hutmacher, D.W., Scaffolds in tissue engineering bone and cartilage.


Biomaterials, 2000. 21(24): p. 2529-2543.

18. Freyman, T.M., I.V. Yannas, and L.J. Gibson, Cellular materials as porous
scaffolds for tissue engineering. Progress in Materials Science, 2001. 46(3-4):
p. 273-282.

19. Gibson, L.J., Ashby, M.F., Cellular Solids: Structures and Properties. 2nd ed.
1997, Cambridge: Cambridge University Press.

20. Gibson, L.J., et al., The Mechanics of Two-Dimensional Cellular Materials.


Proceedings of the Royal Society of London. Series A, Mathematical and
Physical Sciences, 1982. 382(1782): p. 25-42.

63
21. Torquato, S., et al., Effective mechanical and transport properties of cellular
solids. International Journal of Mechanical Sciences, 1998. 40(1): p. 71-82.

22. Hohe, J. and W. Becker, A mechanical model for two-dimensional cellular


sandwich cores with general geometry. Computational Materials Science,
2000. 19(1-4): p. 108-115.

23. Zhu, H.X., J.R. Hobdell, and A.H. Windle, Effects of cell irregularity on the
elastic properties of 2D Voronoi honeycombs. Journal of the Mechanics and
Physics of Solids, 2001. 49(4): p. 857-870.

24. Klintworth, J.W. and W.J. Stronge, Elasto-plastic yield limits and deformation
laws for transversely crushed honeycombs. International J of Mech. Sciences,
1988. 30(3-4): p. 273-292.

25. Papka, S.D. and S. Kyriakides, In-plane compressive response and crushing of
honeycomb. Journal of the Mechanics and Physics of Solids, 1994. 42(10): p.
1499-1532.

26. Taylor, C.M., et al., The effects of hierarchy on the in-plane elastic properties
of honeycombs. International Journal of Solids and Structures, 2011. 48(9): p.
1330-1339.

27. Lira, C. and F. Scarpa, Transverse shear stiffness of thickness gradient


honeycombs. Composites Science and Technology, 2010. 70(6): p. 930-936.

28. Li, K., X.L. Gao, and G. Subhash, Effects of cell shape and strut cross-
sectional area variations on the elastic properties of three-dimensional open-
cell foams. Journal of the Mechanics and Physics of Solids, 2006. 54(4): p.
783-806.

29. Silva, M.J. and L.J. Gibson, The effects of non-periodic microstructure and
defects on the compressive strength of two-dimensional cellular solids.
International Journal of Mechanical Sciences, 1997. 39(5): p. 549-563.

64
30. Luxner, M.H., J. Stampfl, and H.E. Pettermann, Numerical simulations of 3D
open cell structures - influence of structural irregularities on elasto-plasticity
and deformation localization. International Journal of Solids and Structures,
2007. 44(9): p. 2990-3003.

31. Ajdari, A., et al., Effect of defects on elastic-plastic behavior of cellular


materials. Materials Science and Engineering: A, 2008. 487(1-2): p. 558-567.

32. Silva, M.J., W.C. Hayes, and L.J. Gibson, The effects of non-periodic
microstructure on the elastic properties of two-dimensional cellular solids.
International Journal of Mechanical Sciences, 1995. 37(11): p. 1161-1177.

33. Wang, A.-J. and D.L. McDowell, Effects of defects on in-plane properties of
periodic metal honeycombs. International Journal of Mechanical Sciences,
2003. 45(11): p. 1799-1813.

34. Zhu, H. and B.V. Sankar, Analysis of sandwich TPS panel with functionally
graded foam core by Galerkin method. Composite Structures, 2007. 77(3): p.
280-287.

35. Ajdari, A., H. Nayeb-Hashemi, and A. Vaziri, Dynamic crushing and energy
absorption of regular, irregular and functionally graded cellular structures.
International Journal of Solids and Structures, 2011. 48(3-4): p. 506-516.

36. Rosa, M.E. and M.A. Fortes, New types of space-filling polyhedra with
fourteen faces. Acta Crystallographica Section A, 1986. 42(4): p. 282-286.

37. Weaire, D. and N. Rivier, Soap, cells and statistics - random patterns in two
dimensions (Reprinted from Contemporary Physics, vol 25, pg 59, 1984).
Contemporary Physics, 2009. 50(1): p. 199-239.

38. Ko, W.L., Deformations of Foamed Elastomers. Journal of Cellular Plastics,


1965. 1(1): p. 45-50.

39. Thompson, D.W., On Growth and Form. Bonner, J.T ed. 1961: Cambridge
University Press.

65
40. Zhu, H.X., J.F. Knott, and N.J. Mills, Analysis of the elastic properties of
open-cell foams with tetrakaidecahedral cells. Journal of the Mechanics and
Physics of Solids, 1997. 45(3): p. 319-325.

41. Jang, W.-Y. and S. Kyriakides, On the crushing of aluminum open-cell foams:
Part II analysis. International Journal of Solids and Structures, 2009. 46(3-4):
p. 635-650.

42. Laroussi, M., K. Sab, and A. Alaoui, Foam mechanics: Nonlinear response of
an elastic 3D-periodic microstructure. International Journal of Solids and
Structures, 2002. 39(13-14): p. 3599-3623.

43. Gong, L., S. Kyriakides, and W.Y. Jang, Compressive response of open-cell
foams. Part I: Morphology and elastic properties. International Journal of
Solids and Structures, 2005. 42(5-6): p. 1355-1379.

44. Gong, L., S. Kyriakides, and N. Triantafyllidis, On the stability of Kelvin cell
foams under compressive loads. Journal of the Mechanics and Physics of
Solids, 2005. 53(4): p. 771-794.

45. Sullivan, R.M., L.J. Ghosn, and B.A. Lerch, A general tetrakaidecahedron
model for open-celled foams. International Journal of Solids and Structures,
2008. 45(6): p. 1754-1765.

46. Warren, W.E. and A.M. Kraynik, Linear Elastic Behavior of a Low-Density
Kelvin Foam With Open Cells. Journal of Applied Mechanics, 1997. 64(4): p.
787-794.

47. Gong, L. and S. Kyriakides, Compressive response of open cell foams Part II:
Initiation and evolution of crushing. International Journal of Solids and
Structures, 2005. 42(5-6): p. 1381-1399.

48. Webster, R. and P.G. Read, Gems: Their sources, descriptions and
identification. 5th ed. 1994, Great Britain: Butterworth-Heinemann.

66
49. Matsuoka, K.-i., T. Akiyama, and S. Yamada, Selective Formation and
Structural Properties of Rhombic Dodecahedral [70]Fullerene Microparticles
Formed by Reaction with Aliphatic Diamines. Langmuir, 2009. 26(6): p. 4274-
4280.

50. Gibson, L.J., Mechanical behavior of metallic foams. Annual review of


materials science, 2000. 30: p. 191-227.

51. Chen, C., T.J. Lu, and N.A. Fleck, Effect of imperfections on the yielding of
two-dimensional foams. Journal of the Mechanics and Physics of Solids, 1999.
47(11): p. 2235-2272.

52. Deshpande, V.S. and N.A. Fleck, Isotropic constitutive models for metallic
foams. Journal of the Mechanics and Physics of Solids, 2000. 48(6-7): p.
1253-1283.

53. Li, K., X.L. Gao, and G. Subhash, Effects of cell shape and cell wall thickness
variations on the elastic properties of two-dimensional cellular solids.
International Journal of Solids and Structures, 2005. 42(5-6): p. 1777-1795.

54. Chen, C., T.J. Lu, and N.A. Fleck, Effect of inclusions and holes on the
stiffness and strength of honeycombs. International Journal of Mechanical
Sciences, 2001. 43(2): p. 487-504.

55. Guo, X.E. and L.J. Gibson, Behavior of intact and damaged honeycombs: A
finite element study. International Journal of Mechanical Sciences, 1999.
41(1): p. 85-105.

56. Li, B., B. Wang, and S.R. Reid, Effective elastic properties of randomly
distributed void models for porous materials. International Journal of
Mechanical Sciences, 2010. 52(5): p. 726-732.

57. Kraynik, A.M., et al., Foam Micromechanics. NATO Advanced Studies,


Institute on, 1998: p. Medium: ED.

67
58. Markaki, A.E. and T.W. Clyne, The effect of cell wall microstructure on the
deformation and fracture of aluminium-based foams. Acta Materialia, 2001.
49(9): p. 1677-1686.

59. Roberts, A.P. and E.J. Garboczi, Elastic moduli of model random three-
dimensional closed-cell cellular solids. Acta Materialia, 2001. 49(2): p. 189-
197.

60. Jeon, I. and T. Asahina, The effect of structural defects on the compressive
behavior of closed-cell Aluminum foam. Acta Materialia, 2005. 53(12): p.
3415-3423.

61. Roberts, A.P. and E.J. Garboczi, Elastic properties of model random three-
dimensional open-cell solids. Journal of the Mechanics and Physics of Solids,
2002. 50(1): p. 33-55.

62. Grenestedt, J.L. and F. Bassinet, Influence of cell wall thickness variations on
elastic stiffness of closed-cell cellular solids. International Journal of
Mechanical Sciences, 2000. 42(7): p. 1327-1338.

63. Grenestedt, J.L., Influence of wavy imperfections in cell walls on elastic


stiffness of cellular solids. Journal of the Mechanics and Physics of Solids,
1998. 46(1): p. 29-50.

64. Xue, Z. and J.W. Hutchinson, Crush dynamics of square honeycomb sandwich
cores. International Journal for Numerical Methods in Engineering, 2006.
65(13): p. 2221-2245.

65. Xue, Z. and J.W. Hutchinson, Constitutive model for quasi-static deformation
of metallic sandwich cores. International Journal for Numerical Methods in
Engineering, 2004. 61(13): p. 2205-2238.

66. Budiansky, B. and E. Kimmel, Elastic Moduli of Lungs. Journal of Applied


Mechanics, 1987. 54(2): p. 351-358.

68
67. Kimmel, E. and B. Budiansky, Surface Tension and the Dodecahedron Model
for Lung Elasticity. Journal of Biomechanical Engineering, 1990. 112(2): p.
160-167.

68. Zhu, H.X., J.R. Hobdell, and A.H. Windle, Effects of cell irregularity on the
elastic properties of open-cell foams. Acta Materialia, 2000. 48(20): p. 4893-
4900.

69. Alkhader, M. and M. Vural, Mechanical response of cellular solids: Role of


cellular topology and microstructural irregularity. International Journal of
Engineering Science, 2008. 46(10): p. 1035-1051.

70. Gibson, L.J., M.F. Ashby, and B.A. Harley, Cellular Materials in Nature and
Medicine. 2010: Cambridge University Press.

71. Silva, E., M. Walters, and G. Paulino, Modeling bamboo as a functionally


graded material: Lessons for the analysis of affordable materials. Journal of
Materials Science, 2006. 41(21): p. 6991-7004.

72. Silva, F.d.A., N. Chawla, and R.D.d.T. Filho, Tensile behavior of high
performance natural (sisal) fibers. Composites Science and Technology,
2008. 68(15-16): p. 3438-3443.

73. Gibson, L.J., Biomechanics of cellular solids. Journal of Biomechanics, 2005.


38(3): p. 377-399.

74. Sugimura, Y., et al., On the mechanical performance of closed cell Al alloy
foams. Acta Materialia, 1997. 45(12): p. 5245-5259.

75. Walpole, R.E., Probability & statistics for engineers & scientists. 2007, Upper
Saddle River, NJ: Pearson Prentice Hall.

69

You might also like