Vous êtes sur la page 1sur 11

Materials Science and Engineering C 69 (2016) 1307–1317

Contents lists available at ScienceDirect

Materials Science and Engineering C

journal homepage: www.elsevier.com/locate/msec

Mechanical properties of additively manufactured


octagonal honeycombs
R. Hedayati a,b,⁎, M. Sadighi a, M. Mohammadi-Aghdam a, A.A. Zadpoor b
a
Department of Mechanical Engineering, Amirkabir University of Technology (Tehran Polytechnic), Hafez Ave, Tehran, Iran
b
Department of Biomechanical Engineering, Faculty of Mechanical, Maritime, and Materials Engineering, Delft University of Technology (TU Delft), Mekelweg 2, 2628 CD Delft, The Netherlands

a r t i c l e i n f o a b s t r a c t

Article history: Honeycomb structures have found numerous applications as structural and biomedical materials due to their
Received 21 April 2016 favourable properties such as low weight, high stiffness, and porosity. Application of additive manufacturing
Received in revised form 31 July 2016 and 3D printing techniques allows for manufacturing of honeycombs with arbitrary shape and wall thickness,
Accepted 7 August 2016
opening the way for optimizing the mechanical and physical properties for specific applications. In this study,
Available online 8 August 2016
the mechanical properties of honeycomb structures with a new geometry, called octagonal honeycomb, were in-
Keywords:
vestigated using analytical, numerical, and experimental approaches. An additive manufacturing technique,
Octagonal honeycomb namely fused deposition modelling, was used to fabricate the honeycomb from polylactic acid (PLA). The honey-
Timoshenko beam theory combs structures were then mechanically tested under compression and the mechanical properties of the struc-
Analytical relationships tures were determined. In addition, the Euler-Bernoulli and Timoshenko beam theories were used for deriving
Porous structures analytical relationships for elastic modulus, yield stress, Poisson's ratio, and buckling stress of this new design
of honeycomb structures. Finite element models were also created to analyse the mechanical behaviour of the
honeycombs computationally. The analytical solutions obtained using Timoshenko beam theory were close to
computational results in terms of elastic modulus, Poisson's ratio and yield stress, especially for relative densities
smaller than 25%. The analytical solutions based on the Timoshenko analytical solution and the computational
results were in good agreement with experimental observations. Finally, the elastic properties of the proposed
honeycomb structure were compared to those of other honeycomb structures such as square, triangular, hexag-
onal, mixed, diamond, and Kagome. The octagonal honeycomb showed yield stress and elastic modulus values
very close to those of regular hexagonal honeycombs and lower than the other considered honeycombs.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction a micro-architecture similar to that of the native tissue, e.g. bone, could
help tissue-substituting biomaterials to better mimic the mechanical re-
Cellular solids have been widely studied as temporary or permanent sponse of the native tissue they replace. The microstructure of trabecu-
scaffolds in tissue engineering [1,2]. In general, cellular solids can be cat- lar bone is shown to be close to that of honeycombs with thick walls in
egorized into at least two main groups: foams and honeycombs. Foams certain high density areas [7]. Recently, additive manufacturing tech-
have complex 3D micro-structural geometries with usually irregular niques have been used for fabrications of biomedical implants based
unit cell shapes [3]. Honeycombs, on the other hand, possess in-plane on honeycomb structures [8,9]. Moreover, collagen scaffold with honey-
tessellated and regular microstructures that form lightweight struc- comb-like micro-structure have been used for in vitro cell culture and in
tures. The advent of additive manufacturing techniques made it possible vivo studies [10–12].
to easily manufacture cellular solids with novel designs and precisely In this study, we investigate the in-plane mechanical properties of a
controlled micro-architectures. Several recent studies have investigated new design of honeycomb structures. Natural honeycombs constituting
the mechanical and biological properties of additively manufactured 3D the cancellous bone, cork, or wood are usually loaded in in-plane [13].
porous biomaterials with different unit cell types and densities [4–6]. To Understanding the in-plane mechanical properties of honeycomb struc-
integrate well within the host tissue, tissue-substituting biomaterials tures not only helps us better understand the form-function relation-
should provide enough space for de novo tissue regeneration while ship in biological materials but could be also used for better design
exhibiting stiffness values close to the native tissue they replace. Having and (additive) manufacturing of tissue-substituting biomaterials.
Several experimental [14–21] and numerical [22–27] works have
been published on the in-plane deformation of honeycombs. Hexagonal
⁎ Corresponding author at: Department of Mechanical Engineering, Amirkabir
University of Technology (Tehran Polytechnic), Hafez Ave, Tehran, Iran.
shape is the most widely studied geometry among the honeycombs.
E-mail addresses: rezahedayati@gmail.com, r.hedayati@tudelft.nl, However, other types of honeycombs have also been investigated by
rezahedayati@aut.ac.ir (R. Hedayati). other researchers [28]. Torquato et al. [29] investigated the effective

http://dx.doi.org/10.1016/j.msec.2016.08.020
0928-4931/© 2016 Elsevier B.V. All rights reserved.
1308 R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317

linear elastic properties of honeycombs with square, hexagonal, and tri- honeycomb structure to their corresponding value in the bulk material)
angular cell shapes using two different approaches, namely homogeni- are reported in the following. The mechanical properties of the bulk ma-
zation theory and discrete network analyses, in order to establish terial are required when calculating the normalized values of elastic
rigorous bounds on the effective elastic modulus of honeycombs in modulus and yield stress. Cylindrical specimens were manufactured
terms of the thermal conductivity and vice versa [28]. The effective from the same PLA materials and were tested under compressive load-
Young's modulus of triangular cell honeycombs was studied by Gulati ing to determine their mechanical properties. The measured elastic
[30] with reference to thermal shock resistance. modulus and yield stress of blue filaments were Es = 1.962 ±
In addition to numerical and experimental studies, several analytical 0.069 GPa and σys = 56.204 ± 1.2127 MPa, respectively.
relationships have also been obtained for honeycombs with different
micro-architectures. Analytical relationships provide fast (and accurate) 2.2. Relative density
estimation of the mechanical properties of honeycombs. Moreover, an-
alytical models allow for better understanding of the various geometric Relative density is defined as the ratio of the density of a porous
aspects and physical mechanisms behind the mechanical behaviour of structure to the density of the material it is made of. Expressing the me-
rationally designed materials such as honeycombs. For example, using chanical properties of a porous structure in terms of its relative density
analytical relationships, the user could quickly find the optimal micro- would be helpful in comparing its mechanical properties with those of
architecture of a honeycomb structure using simple optimization algo- porous structures with other morphologies. In order to obtain the rela-
rithms. The first analytical work published on the in-plane mechanical tive density, the volume of the bulk material present in one unit cell
properties of honeycombs was performed by El-Sayed et al. [31] in must be divided by the total volume occupied by the unit cell. Since
which the elastic properties of unfilled and filled hexagonal honeycomb the depth of the walls are constant throughout honeycomb structures,
sheets under in-plane loading and out-of-plane bending were studied. finding the filled and total areas of a unit cell in the cross-section of
However, their work had “numerous small errors” [13]. In 1982, Gibson the lattice structure is sufficient (see Fig. 1). If the cell walls are assumed
et al. [13] obtained analytical relationships for different mechanical to be thin enough, the area occupied by the materials constructing the
properties of hexagonal honeycombs (E1, E2 , υ12, υ21, σy1, σy2, G, and 2
walls of a unit cell is (4lt+ 4lt/2) = 6lt. The total area of a unit cell is l
σel). Their analytical results showed excellent agreement with their ex- pffiffiffi 2
perimental results for both the rubber- and metal-made honeycombs ð 2 þ 1Þ . Consequently, the approximate relative density of an octag-
but only for small values of relative density. A few other analytical stud- onal honeycomb is obtained as
ies [21,32–34] were carried out to improve the Gibson et al.'s relation-
ships for hexagonal honeycombs. Seven other types of honeycomb cell 
6 t
shapes (square, triangular, hexagonal, Kagome, mixed, diamond, and μ ¼ pffiffiffi 2 ð1Þ
l
rectangular) were investigated in a study by Wang and McDowell [28] 2þ1
using simple beam or column/truss elements, including linear and non-
linear theories to explore in-plane effective elastic stiffness and initial
yield strength. The exact area occupied by the material constructing the walls of a
In this paper, a design for honeycomb-shaped pores referred to as unit cell is (Fig. 1)
octagonal, is introduced and its mechanical response is studied in the
elastic range. An additive manufacturing technique, namely fused depo- ( pffiffiffi pffiffiffi !  )
1 l t 2 l t 2 t t 1 l t l t
sition modelling, was used to fabricate the honeycombs from polylactic Am ¼ 8 − þ − þ þ þ þ − t ð2Þ
2 2 2 2 2 2 2 2 2 2 2 2
acid (PLA). The honeycomb structures were then mechanically tested
under compression and the mechanical properties of the structures
were determined. The elastic modulus, Poisson's ratio, yield stress, and
buckling stress of this structure is obtained analytically. The Euler-
Bernoulli and Timoshenko beam theories are used for deriving the ana-
lytical relationships. Two finite element (FE) models are also created,
one consisting of 1/4 of a unit cell with periodic boundary conditions
and the other consisting of a large set of unit cells. The experimental, an-
alytical, and computational results are compared with each other.

2. Materials and methods

2.1. Experimental tests

We used an additive manufacturing technique, namely fused depo-


sition modelling (FDM), to create octagonal honeycomb structures.
Three samples from each relative density were manufactured and
were mechanically tested under compression (Fig. 5). The wall thick-
ness to length ratio of the samples was varied between t/l = 0.2727, t/
l =0.4091, t/l = 0.5454, and t/l = 0.6817 to create four different relative
densities. The specimens were made of PLA using 5th generation
Replicator Desktop Makerbot 3D printers. All the samples had heights
and widths of 77.26 mm and depths of 25 mm. The mechanical tests
were done using Instron ElectroPuls E10000 machine with a 10 kN
load cell. The samples were compressed with a displacement rate of
6 mm/min. The test procedures followed the specifications laid out by
the ISO standard 13314:2011.
To have a better impression of the elastic modulus and yield stress
parameters, their normalized values (i.e., ratio of their value in the Fig. 1. Dimensions of an octagonal unit cell.
R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317 1309

or To calculate the displacements and rotations of the inclined edge


 (edge BC in Fig. 3a), its free body diagram is plotted (Fig. 3b). Edge AB
pffiffiffi t   
t pffiffiffi must remain horizontal and, thus, the angle of vertex B must remain un-
Am ¼ 2 l−t 2 þ t þ 2lt ¼ 2t 3l þ −t 2 ð3Þ
2 2 changed. Edge BC can therefore be considered as a cantilever beam at
the end of which an axial load F/2sin θ, a lateral load F/2 cos θ, and a mo-
pffiffiffi
2 2
and the area of a unit cell is again l ð 2 þ 1Þ . Therefore, the exact rel- ment M are applied, where θ = π/4. As already stated, the loads F/2sin θ
ative density of this structure is calculated as and F/2 cosθ are imposed on edge BC by edge CD of the cell, while mo-
ment M is applied by the adjacent cell located in the right side of the
 pffiffiffi 2t  
t t t pffiffiffi cell. Since the vertical edge remains vertical, and given that angle C
2t 3l þ −t 2 3þ − 2
2 l 2l l must remain constant, the rotation of beam BC at its end C must be
μ¼ pffiffiffi 2 ¼ pffiffiffi 2 ð4Þ
2 zero after deformation. The axial load F/2 sinθ does not rotate the end
l 2þ1 2þ1
C. Therefore, the rotations produced by the lateral load F/2 cos θ and
the moment M must be equal and in the reverse directions from
which the value of M can be calculated as:
2.3. Euler-Bernoulli beam theory
pffiffiffi
F 2 2 pffiffiffi
In this section, we use the Euler-Bernoulli beam theory to derive an- l Ml Fl 2
2 2 ¼ ⇒ M¼ ð5Þ
alytical relationships for the elastic modulus, Poisson's ratio, and yield 2Es I Es I 8
stress of an octagonal honeycomb structure as functions of the elastic
modulus and yield stress of the matrix material (Es , σys) and the relative 3
where I ¼ 121
bt is the area moment of inertia of the edges in which b is
density (μ) of the honeycomb.
Each of the unit cells constructing a 2D lattice structure (Fig. 2) con- the depth of the honeycomb structure in the out-of-plane direction. In
sists of two vertical, two horizontal, and four inclined edges. Due to the order to obtain the vertical strain of the unit cell as a function of the ap-
symmetry existent in the octagonal structure (Fig. 2), considering 1/4 of plied load F (which is necessary for obtaining the elastic modulus and
the structure (here the lower right portion) is sufficient for the analyti- Poisson's ratio of the unit cell), first the vertical component of the struc-
cal study (Fig. 3a). If the total load applied to each unit cell is F, then the ture displacement must be obtained. Decrease in the length of half of the
axial load applied to each vertical edge is F/2. The load applied to edge vertical edge CD (i.e., CC′) is
CD is F/2, and since the lattice structure is free to move laterally, the hor-
izontal loads are zero. Due to the symmetry existent between each of Fl
Fl
two adjacent cells, the vertical and horizontal edges have to respectively δCC 0 ¼ 2 ¼
2 ð6Þ
A 2AEs
remain vertical and horizontal during elastic deformation. That is, if one Es
of the noted edges rotates by θ°, satisfying the symmetry conditions re- 2
quires that it rotates −θ° in the adjacent cell. The rotation −θ° equals θ°,
only if θ = 0. That holds only for the vertical and horizontal edges, be- where A = bt is the cross-sectional area of the edges. The lateral load F/
cause they are the only edges that are shared between adjacent cells. 2 cos θ tends to decrease the distance GC (Fig. 3b) by
A concentrated moment must have been applied to vertex C of edge pffiffiffi
CD by the adjacent cell located in the right side of the cell (Fig. 3a) to F 2 3 pffiffiffi
l 2 Fl
3
keep the angle of vertex C unchanged. A moment with equal magnitude δBC F;bending ¼2 2 ¼ ð7Þ
and reverse directions is applied by the cell to the noted edge of the 3Es I 2 12Es I
right cell due to symmetry.
On the other hand, the moment M tends to increase the distance GC
(Fig. 3b) by

2
pffiffiffi
Ml 2
δBC M ¼ ð8Þ
2Es I 2

Substituting M from Eq. (5) into Eq. (8) gives

3
Fl
δBC M ¼ ð9Þ
16Es I

Finally, the axial load F/2 sin θ tends to decrease the distance GC
(Fig. 3b) by

pffiffiffi
F 2 pffiffiffi
l 2 Fl
δBC F;axial ¼2 2 ¼ ð10Þ
AEs 2 4AEs

Therefore, the total vertical displacement of the unit cell is

δuc ¼ 2 δCC 0 þ δBC F;axial þ δBC F;bending −δ


!BCM
3 3
Fl Fl Fl Fl
Fig. 2. A 2D cut showing an octagonal unit cell before and after deformation under ¼2 þ þ − ð11Þ
2AEs 4AEs 12Es I 16Es I
compressive loading.
1310 R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317

Fig. 3. The free body diagram and deformation of (a) 1/4 portion of, and (b) the inclined edge of an octagonal unit cell under compressive loading.

which after being simplified becomes In order to find the yield stress in the structure, first the maximum
stress in each cell wall must be calculated. If the load F is applied to
3
3Fl Fl each unit cell, the normal compressive stress in edge CD is
δuc ¼ þ ð12Þ
2AEs 24Es I
F
2 F
In order to find the elastic modulus of the octagonal honeycomb σ max;CD ¼− ¼− ð18Þ
A A
structure, the simple equation E ¼ AFucLδucuc can be used in which F is the 2
pffiffiffi pffiffiffi
load applied to the unit cell, and Luc ¼ ð1 þ 2Þl and Auc ¼ blð1 þ 2Þ
are the length and the cross-section area of the unit cell, respectively. and the normal stress in edge BC is:
Substituting Eq. (12), Luc, and Auc in E ¼ AFucLδucuc gives the elastic modulus 2pffiffiffi! 3
F 2
of the octagonal honeycomb structure pffiffiffi 6 lc7
F 2 6 Mc 2 2 7
σ BC ¼ σ BC;axial  σ BC;bending ¼− 6
6 I − 7
7 ð19Þ
F 2A 2 4 I 5
E¼ " 3
# ð13Þ
3Fl Fl
b þ
2AEs 24Es I
where c is the distance between the neutral plane and the farthest por-
and the relative elastic modulus is tion of the beam cross-section. Substituting M from Eq. (5) gives:
2 pffiffiffi! 3
pffiffiffi F 2
E 1 2t
Eμ ¼ ¼ " #¼ h
2 i ð14Þ pffiffiffi 6Fl 2 c lc7 pffiffiffi pffiffiffi
Es 3l l
3
l 3 þ tl F 2 6 2 2 7
7 ¼ − F 2 − Fl 2c
b þ σ BC; max ¼− 6
6 I
8 − 7 ð20Þ
2A 24I 2A 2 4 I 5 4A 8I

For obtaining the Poisson's ratio, it is first necessary to calculate the


lateral displacement of the unit cell. The lateral displacement of the unit Considering Eq. (18) and Eq. (20), the maximum compressive stress
cell is twice the horizontal displacement of point C in Fig. 3, which is in the structure is found by:
given by ( pffiffiffi pffiffiffi )

F F 2 Fl 2c
σ max ¼ max ; þ ð21Þ
δlateral ¼ 2 −δBC F;axial þ δBC F;bending −δBC M ð15Þ A 4A 8I

Substituting Eqs. (7), (9) and (10) into Eq. (15) gives Our calculations showed that the second term is larger than the first
! term in all the relative densities. If an applied external stress ð1þpF ffiffi2Þlb on
3 3 3
Fl Fl Fl Fl Fl pffiffi pffiffi
δlateral ¼ 2 − þ − ¼− þ ð16Þ the unit cell creates the maximum local stress of F4A2 þ Fl 8I2c in the struc-
4AEs 12Es I 16Es I 2AEs 24Es I
ture, the external stress σy which causes the maximum local stress in
the structure reach σys can be found by a simple cross multiplication.
The Poisson's ratio is therefore given by:
By dividing the resulted relationship by σys, the relative yield stress of
3
the structure can be found as
Fl Fl
− þ 2 2 "pffiffiffi pffiffiffi #−1 pffiffiffi
δ 2AEs 24Es I −12I þ Al −t 2 þ l
ν ¼ lateral ¼ ¼ ¼ ð17Þ 1 2 l 2c 2 2t 2
δuc 3Fl Fl
3
36I þ Al
2
3t 2 þ l
2
σ y rel ¼  pffiffiffi þ ¼  pffiffiffi ð22Þ
þ 1 þ 2 lb 4A 8I 1 þ 2 ðt þ 3lÞl
2AEs 24Es I
R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317 1311

2.4. Timoshenko beam theory On the other hand, the lateral load F/2 cos θ tends to decrease the
length GC (Fig. 3b) by
The Timoshenko beam theory takes into account shear deformation
and rotational inertia effects, making it suitable for describing the behav- pffiffiffi ! pffiffiffi !
3 3
iour of thick beams. For a homogenous beam of constant cross-section, 2 l l 2 F l l
δBC F;bending ¼ F þ ¼ þ ð28Þ
the Timoshenko beam governing equations are: 4 3Es I κAGs 2 4 3Es I κAGs

2  
d dφ
E s I ¼ qðx; t Þ The displacements resulted from the axial loads, i.e. δBCF,axial and δCC′
dx2 dx
are identical to those for the Euler-Bernoulli beam theory. The total ver-
tical displacement of the unit cell is therefore

δuc ¼ 2(δCC 0 þ δBC F;axial þ δBC F;bending −δBCM


! !)
3 3
  Fl Fl F l l F l l
dw 1 d dφ ¼2 þ þ þ − þ
¼ φ− Es I ð23Þ 2AEs 4AEs 4 3Es I κAGs 8 2Es I κAGs
dx κAG dx dx 3
3Fl Fl Fl
¼ þ þ ð29Þ
2AEs 4κAGs 24Es I
where φ is the angle of rotation of the normal to the mid-surface of the
beam, w is the lateral displacement of the mid-surface, and κ is the Like before, the simple equation E ¼ AFucLδucuc can be used to find the elas-
shear coefficient factor. The shear coefficient factor is 10(1 + ν)/ tic modulus of the octagonal honeycomb structure. Substituting Eq.
(12+11ν) for a beam with rectangular cross-section (which is the case (29), the elastic modulus of the octagonal honeycomb structure is
in this study). In a linear elastic Timoshenko beam, the bending moment
Mxx and the shear force Qx are related to the angle of rotation, φ, and the F
E¼ " 3
# ð30Þ
displacement, w, by 3Fl Fl Fl
b þ þ
2AEs 4κAGs 24Es I

∂φ
Mxx ¼ −EI and the relative elastic modulus is
∂x

t
2
E 1 l
Eμ ¼ ¼ " #¼  2 ð31Þ
  Es 3l lð1 þ ν s Þ l
3
1 þ νs l
∂w b þ þ 3þ þ
Q x ¼ κAG −φ þ ð24Þ 2A 2κA 24I κ t
∂x

To obtain the Poisson's ratio of the structure, it is first necessary to


For a cantilever Timoshenko beam with a lateral point load P at its
calculate its lateral displacement. The lateral displacement of the struc-
free end, Eqs. (24) yield
ture is twice the horizontal displacement of point C in Fig. 3b, which is
given by
Pl
3
Pl Pl
2
P

δ¼ þ and θ¼ þ ð25Þ δlateral ¼ 2 −δBC F;axial þ δBC F;bending −δBC M ð32Þ


3Es I κAGs 2Es I κAGs

Substituting Eqs. (10), (27), and (28) into Eq. (32) gives
which have additional terms compared to those in the Euler-Bernoulli " ! !#
3 3
beam. However, for a beam with a concentrated moment M at its end, −Fl F l l F l l
δlateral ¼ 2 þ þ − þ
the displacement and rotation are identical to those of Euler-Bernoulli 4AEs 4 3Es I κAG 8 2Es I κAG
beam. As in the case of the Euler-Bernoulli beam theory, the rotations −Fl Flð1 þ ν s Þ Fl
3
produced by the lateral load P = F/2 cos θ and moment M must be ¼ þ þ ð33Þ
2AEs 2κAEs 24Es I
equal and opposite from which the value of M for a structure based on
the Timoshenko beam theory can be determined:
The Poisson's ratio is then easily found as:

2
pffiffiffi" 2 # −Fl Flð1 þ ν s Þ Fl
3
Aκl
2
Pl P 2 l 1 Ml þ þ −Iκ þ Ið1 þ ν s Þ þ
þ ¼F þ ¼ ⇒ M δlateral 2AEs 2κAEs 24Es I 12
2Es I κAGs 4 2Es I κAGs Es I ν¼ ¼ ¼
pffiffiffi  δ 3Fl Fl Fl
3
Aκl
2
2 l Es I þ þ 3Iκ þ Ið1 þ νs Þ þ
¼F þ ð26Þ 2AEs 4κAG  s 24Es I 12
4 2 κAGs l 2 1 þ νs
−t 2 þ l þ t 2
κ
¼   ð34Þ
2 2 1 þ νs
3t þ l þ t
2
κ
The moment M tends to increase the distance GC (Fig. 3b) by
Now, we find the yield stress for the structure based on the Timo-
pffiffiffi  shenko beam theory. The normal stress in edge CD is
2 l Es I 2 pffiffiffi !
F þ l 3
4 2 κAGl 2 F l l
δBC M ¼ ¼ þ ð27Þ F
2Es I 2 8 2Es I κAGs σ max;CD ¼ − ð35Þ
A
1312 R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317

and the normal stress in edge BC is: from which the critical stress is readily obtained

2 pffiffiffi! 3  3
F 2 4π2 Es I 1 4π2 Es I π2 Es t
pffiffiffi 6 lc7 σ cr ¼ 2
 pffiffiffi ¼  pffiffiffi 3
¼  pffiffiffi ð42Þ
F 2 6 6Mc 2 2 7
7 l 1 þ 2 lb 1 þ 2 bl 3 1þ 2 l
σ BC ¼ σ BC;axial  σ BC;bending ¼− 6 − 7
2A 2 4 I I 5

2 pffiffiffi  pffiffiffi! 3
2 l 2 ð1 þ ν ÞI F 2
pffiffiffi 6 F þ
s
c lc7 2.6. FE modelling
F 2 6 4 2 κAl 2 2 7
¼− 6
6 − 7
7 ð36Þ
2A 2 4 I I 5 The main mechanism of deformation in the microstructure of a hon-
eycomb structure is bending of the cell walls making the beam elements
the natural choice for modelling them. They are computationally inex-
pensive and can be used to construct models with many cells [35]. The
The maximum compressive stress in the cell wall is therefore cell edges were discretized using the standard Timoshenko beam ele-
ments (element type 189 in ANSYS) that uses linear interpolation
pffiffiffi pffiffiffi
F 2 2ð1 þ ν s Þc F 2lc (two-node linear beam). Beams that were rigidly connected at the ver-
σ BC; max ¼ 1− þ ð37Þ
4A κl 8I tices mechanically represented all the edges in the honeycomb struc-
ture. In all the calculations, the cell edge material was assumed to be
linear elastic with mechanical properties similar to the mechanical
Consequently, the maximum stress in the structure is found by: properties of the bulk PLA material determined in the experiments of
( cylindrical specimens.
pffiffiffi pffiffiffi )
F F 2 2ð1 þ νs Þc F 2lc The static nonlinear implicit solver available in ANSYS FE code was
σ max ¼ max ; 1− þ ð38Þ used for the calculations. Two types of models were created for the nu-
A 4A κl 8I
merical modelling. The first model was a small model consisting of three
edges corresponding to 1/4 of a unit cell (Fig. 4a). In this structure, all
Again, the calculations showed that the second term is larger than the degrees of freedom of the bottom edge were fixed and the right
the first term in all the relative densities. A simple cross-multiplication edge was only allowed to move transitionally. A concentred force F
could be used to determine the relative yield stress in an octagonal was applied to the top right side of the structure (Fig. 4a).
unit cell based on the Timoshenko beam theory: The other model was a large lattice structure with several unit cells.
The number of elements per cell wall was changed from 1 to larger
"pffiffiffi pffiffiffi #−1 values and no considerable difference was observed. However, the pre-
1 2 2ð1 þ ν s Þc 2lc
σ y rel ¼ pffiffiffi 1− þ dicted results were affected by the number of unit cells in each direction
1 þ 2 lb 4A κl 8I of the FE model. The results of the lattice structure converged for lattice
t sizes larger than 9 × 9 × 9. A lattice structure size of 14 × 14 × 14 unit
1 l cells was chosen for the numerical analysis. The lowermost part of the
¼ pffiffiffi "pffiffiffi pffiffiffi # ð39Þ
1þ 2 2 2ð1 þ ν s Þt 3 2l lattice structure was not allowed to move in the Y direction, while the
1− þ
4 2κl 4t uppermost part of the lattice structure was displaced downward in
such a way that the structure underwent 0.2% strain. The lowermost
and the uppermost nodes were allowed to move in the X direction, to
better capture the lateral behaviour of the structure (for example for
2.5. Buckling load obtaining the Poisson's ratio parameter). One of the lowermost nodes
of the lattice structure was fixed in all the directions to prevent the
Since the honeycomb structures are usually used in compression, rigid body motions of the structure. The effective elastic modulus was
buckling instability can also cause sudden structural deformations calculated using the simple equation E ¼ APLattice
lattice Llattice P lattice
δlattice ¼ b δlattice . To obtain
which can lead to catastrophic failure of the honeycombs. The elastic
the Poisson's ratio, the lateral displacement was divided by the axial dis-
buckling limit of the octagonal honeycomb can be obtained using the
placement of the structure. For the yield stress, similar to the analytical
classical Euler's buckling theory. Since the Timoshenko and Euler-
solution, first the maximum stress in the structure was found and then
Bernoulli theories predict similar deformations under axial loading,
inserted into a simple relationship for cross-multiplication.
one buckling relationship will be obtained for both theories. The Euler's
formula for buckling load of columns under compressive loading is:
3. Results

π2 Es I There were few variations in the load-displacement curves obtained


P cr;column ¼ ð40Þ
ðKLÞ2 for various samples with the same relative density (Fig. 6). The stress
levels and failure loads were several times higher for specimens with
higher values of relative density as compared to those with the lowest
where KL is the effective length of the column and is 0.5 L when both
values of relative density (Fig. 6). All of the octagonal test samples
ends are fixed. In the octagonal honeycomb unit cell, the vertical
showed 45° failure patterns (Fig. 7). Comparison of the analytical solu-
edges are the most susceptible edges to buckling. These vertical edges
tions and the computational results showed that the FE models and
can be considered as columns with both ends fixed, due to their purely
both the analytical models give close results in terms of elastic modulus,
transitional lateral movement in pre-buckling deformation. By
Poisson's ratio, and yield stress, especially for relative densities smaller
substituting KL=0.5 L in Eq. (40), the critical load of a unit cell under
than 20% (Fig. 8a–c). The presence of porosity in the honeycomb struc-
the compressive load F is given by
ture decreases the relative elastic modulus and relative yield stress se-
verely, in such a way that even at relative densities as large as 50%, the
4π2 Es I relative elastic modulus and relative yield stress did not exceed 15% of
P cr;unit cell ¼ 2
ð41Þ
l those in the bulk material (Fig. 8a,c).
R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317 1313

Fig. 4. The geometries, applied loads, and applied boundary conditions used in the FE (a) single unit cell, and (b) lattice octagonal structure.

At relative densities larger than 25%, the numerical and analytical 4. Discussion
values of elastic modulus started to deviate from each other (Fig. 8a).
However, even at large relative densities, each set of analytical (Euler- In this study, two types of FE models were used. A simple model sim-
Bernoulli and Timoshenko) and numerical (lattice structure and single ilar to the one considered for obtaining the analytical relationships and a
unit cell) elastic moduli showed very close predictions. The experimental large 2D tessellated structure with a high number of cell walls (1204 cell
values of the elastic modulus were very close to the numerical predictions walls). The elastic moduli obtained from both the computational
for three out of four relative densities (Fig. 8a). There was a better agree- models were very close with a maximum difference below 4% even for
ment between computational results and the analytical solutions based a relative density as large as 60%.
on the Timoshenko beam theory (as compared to those based on the It is customary to present the analytical relationships used for pre-
Euler-Bernoulli beam theory) (Fig. 8a). diction of mechanical properties in terms of the relative density, μ, rath-
Both analytical models and the computational models predicted er than t/l ratio. Since the exact relative density formula (Eq. (4))
Poisson's ratios that started from 1 for very low values of relative density obtained here is a quadratic polynomial, it cannot be readily inserted
and gradually decreased (Fig. 8b). Similar to case of elastic modulus, the into the relationships obtained for the elastic modulus, yield stress,
analytical values of Poisson's ratio based on the Timoshenko beam theory and Poisson's ratio. However, the approximate relative density formula
were closer to the numerical results for all values of relative density (as (i.e. Eq. (1)) may be inserted into the elastic property equations to de-
compared to those based on the Euler-Bernoulli beam theory). At a rela- rive those equations as a function of (approximate) relative density, μ.
tive density of 60%, the Poisson's ratio of the octagonal structure de- The problem is that in the approximate formula for relative density,
creased to ≈0.4 (Fig. 8b). the material located in the vertices of the cells are counted multiple
Similar to the case of elastic modulus and Poisson's ratio, the pre- times, which can lead to inaccurate results [36]. Therefore, before
dicted values of yield stress were very close for both the FE models doing that, it is necessary to find the influence of multiple counting of
(Fig. 8c). The experimental values of the yield stress were always material located in the cell vertices on the obtained relative density. In
lower than the yield stress values predicted by both the analytical solu- one of our previous studies [36], the effect of multiple counting on the
tions and computational models (Fig. 8c). Reviewing Fig. 8a–c shows elastic properties of 3D open-cell porous structures with different unit
that the elastic modulus, Poisson's ratio, and yield stresses predicted cell types was investigated. As it can be seen in Fig. 10, as the ratio of
by the Timoshenko analytical solution is lower, higher, and higher, re- t/l increases, the difference between the calculated relative densities in-
spectively, than those calculated by the Euler-Bernoulli analytical creases. However, the relative densities do not deviate from each other
solution. more than 20% even for a large t/l ratio such as 0.5.

Fig. 5. Octagonal honeycomb samples with (a) t/l = 0.2727, (b) t/l = 0.4091, (c) t/l = 0.5454, and (d) t/l = 0.6817.
1314 R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317

10000 0.35
t/l=0.2727 (sample 1) Finite element (Lattice structure)
t/l=0.2727 (sample 2)
9000 Finite element (Single unit cell)
t/l=0.2727 (sample 3) 0.3
t/l=0.4091 (sample 1) Analytical (Euler−Bernolli)
t/l=0.4091 (sample 2) Analytical (Timoshenko)

Relative elastic modulus


8000
t/l=0.4091 (sample 3) 0.25 Experiment
t/l=0.5454 (sample 1)
7000 t/l=0.5454 (sample 2)
t/l=0.5454 (sample 3) 0.2
t/l=0.6817 (sample 1)
6000
t/l=0.6817 (sample 2)
Load (N)

t/l=0.6817 (sample 3) 0.15


5000

0.1
4000

0.05
3000

2000 0
0 0.1 0.2 0.3 0.4 0.5 0.6
Relative density
1000
(a)
0
0 1 2 3 4 5 6 7 8
Displacement (mm)
1

Fig. 6. Load-displacement curves of octagonal honeycombs with different t/l ratios.


0.8

Poisson‘s ratio
0.6

0.4

Finite element (Lattice structure)


0.2 Finite element (Single unit cell)
Analytical (Euler−Bernolli)
Analytical (Timoshenko)
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Relative density

(b)
0.25
Finite element (Lattice structure)
Finite element (Single unit cell)
Analytical (Euler−Bernolli)
0.2
Analytical (Timoshenko)
Experiments
Relative yield stress

0.15

0.1

0.05

0
0 0.1 0.2 0.3 0.4 0.5 0.6
Relative density

(c)

Fig. 8. Variation of (a) relative elastic modulus, (b) Poisson's ratio, and (c) relative yield
stress vs. relative density for octagonal honeycomb.

The main reason for the 45° failure band can be the fact that inclined
walls are the critical edges in the octagonal structure. The computation-
al results as well as the analytical results (Eqs. (38) and (39)) both show
the fact that the inclined edges are the parts with the most vulnerability.
All the experimental data points for both the elastic modulus and
yield stress correlate well with the numerical and analytical values
(Fig. 8). The only exception is the experimental results corresponding
to the weakest sample (i.e. the sample with t/l = 0.2727). Compared
to the numerical and analytical results, the experimentally determined
Fig. 7. 45° failure bands formed in honeycomb samples with (a) t/l = 0.2727 and (b) t/l = values of elastic properties are very small for the samples with this rel-
0.5454. ative density. The additively manufactured porous structures usually
R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317 1315

−3 0.18
x 10
5 Square
Buckling stress 0.16 Triangular
4.5 Euler−Bernolli yield stress Hexagonal
0.14 Kagome

Relative elastic modulus


4
Diamond
Relative stress (σ/σ )

0.12 Mixed
ys

3.5
Octagonal
3 0.1

2.5 0.08

2 0.06

1.5 0.04
1
0.02
0.5
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
0
0 0.02 0.04 0.06 0.08 0.1 Relative density
Relative density (a)
0.18
Fig. 9. Variation of yield and buckling stresses vs. relative density for octagonal Square
honeycomb. 0.16 Triangular
Hexagonal
0.14 Kagome
have relatively rough external surfaces. The roughness of the samples Diamond

Relative yield stress


with different relative densities is not essentially different because the 0.12 Mixed
roughness of the product is mostly determined by the manufacturing Octagonal
0.1
parameters of the manufacturing device rather than the geometry of
the product. Therefore, the larger the thickness of the cell walls, the 0.08
lower the effect of surface roughness. This is because the proportion of
0.06
completely filled central parts to the weaker external parts in the
cross-section of the cell walls increases by increasing the thickness of 0.04
the cell walls. This fact is verified by our experimental results: the ex-
0.02
perimental/analytical correlation is much improved for higher relative
densities. The reason behind the very small mechanical properties of 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
the most porous structure (i.e. the sample with t/l = 0.2727) might be Relative density
the fact that in this structure, the cell walls are much thinner than the (b)
weaker external parts of the cell as compared to the thickness of the
completely filled central parts. Therefore, a large fraction of the cross- Fig. 11. Comparison of (a) elastic modulus, and (b) yield stress of the octagonal
section of the cell walls is composed of external weak parts that cannot honeycomb with other types of honeycombs.
bear load.
Deformations of open-cell porous materials are usually dominated
by mechanisms such as linear elastic deformation, buckling of the which of the two failure mechanisms occur first. Plotting the yield stress
edges, plasticity of the edges, contact between interfaces of neighbour (Eq. (39)) against the buckling stress (Eq. (42)) demonstrated that the
cell edges, etc. In the elastic regime, the two mechanisms of local yield- buckling stress of this structure (if it is made of steel) is far smaller
ing of the edges and buckling of the edges parallel to the loading direc- than its yield stress (Fig. 9).
tion are the mechanisms dominating the start of plateau region in the After the advent of additive manufacturing techniques, they have
stress-strain curve of the porous structure. It is necessary to determine been mostly used in production of 3D complex geometries that could
not be otherwise manufactured. Among the different 3D foams, struc-
tures such as diamond [37], rhombic dodecahedron [36], truncated
Exact formula cuboctahedron [38,39], truncated cube [40], octahedral [41], and
Approximate formula rhombicuboctahedron [42] have been previously studied by our group
0.5 as candidates for biomedical implants used in orthopedics. Convention-
al manufacturing techniques for the production of honeycombs have
0.4 several limitations. First, honeycombs with very small cell sizes cannot
Relative density

be produced. Second, since honeycomb cell walls are usually shaped


using metal forming methods, their thickness to cell size ratio cannot
0.3
be larger than a specific value. Third, for each unit cell type, a different
production method is necessary and production of honeycombs with
0.2 arbitrary geometry is not always possible. Additive manufacturing
(AM) techniques overcome all the noted limitations to a great extent.
In recent years, selective laser melting (SLM) has been used to create
0.1
hexagonal honeycombs [43,44]. The octagonal honeycomb studied in
this paper is one of the many types of honeycombs that can be produced
0 using additive manufacturing.
0 0.1 0.2 0.3 0.4 0.5
t/l In this study, the elastic modulus, yield strength, and buckling stress
of the in-plane deformation of the octagonal honeycomb were
Fig. 10. Comparison of exact and approximate formulas for relative density versus the t/l expressed in terms of relative density to enable direct comparison of
ratio. the obtained results with other types of honeycombs. In a study carried
1316 R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317

hexagonal honeycomb. Fig. 12 shows the different post-yielding defor-


mation modes of a hexagonal honeycomb in its two main in-plane di-
rections. The square honeycombs have the highest elastic properties
among all the geometries. It must be stated that the low elastic proper-
ties of the octagonal honeycomb is not necessarily a weakness for this
type of structures. In fact, depending on the application of this structure,
such as energy absorption in in-plane or out-of-plane loading, the low
elastic properties of the honeycomb can be insignificant or even desir-
able. In most cellular structures, the energy absorption capacity ability
of the structure is the main criteria for selection of a specific design.
Therefore, investigation of the response of octagonal honeycombs in
the post-elastic region is necessary for implementing this geometry in
practice.
In the earliest phases of tissue regeneration, when blood vessels are
not yet regenerated, nutrition and oxygenation of the cells must occur
through diffusion. If honeycomb structures are used for tissue regener-
ation applications, the entire plane of their two-dimensional structure
should be in contact with the culture media from both sides to ensure
diffusion can effectively perform the required mass transport. Even
then, the thickness of the honeycombs cannot exceed what can be effec-
tively nurtured and oxygenated through diffusion. Additive
manufacturing techniques similar to the ones used in the current
study could be employed to introduce holes in the walls of the honey-
comb structure with very large thicknesses. However, the analytical for-
mulas derived here should be then modified to take the presence of
those holes into account.

5. Conclusions

In this study, analytical relationships were obtained for the mechan-


ical properties (i.e. elastic modulus, Poisson's ratio, yield stress, and
buckling stress) of octagonal honeycomb structures. Both the Euler-
Bernoulli and Timoshenko beam theories were considered for deriving
the analytical relationships. Two FE models were also created, one
consisting of only 1/4 of a unit cell and the other consisting of a large
set of unit cells. The simulation results obtained from the two FE models
revealed that their results were very close and the difference between
the results was less than 4% for elastic modulus. Comparison of the
Euler-Bernoulli and Timoshenko analytical results in one hand and the
numerical results in the other hand showed that the FE models and
the analytical Timoshenko results were close to each other in terms of
all the properties of elastic modulus, Poisson's ratio, and yield stress, es-
pecially at relative densities smaller than 25%. Plotting the yield stress
against the buckling stress demonstrated that the buckling stress of
this structure made of steel is far smaller than its yield stress. The results
Fig. 12. Different post-yielding deformation modes of the two main directions of a of the Timoshenko analytical solution as well as the two finite element
hexagonal honeycomb: (a) an undeformed aluminium honeycomb, (b) the post- models also showed good correlation with the experimental results. Fi-
yielding deformation of the honeycomb in the X1 direction, and (c) the post-yielding nally, the elastic properties of the octagonal honeycomb structure was
deformation of the honeycomb in the X2 direction [3].
compared to those of honeycomb structures having square, triangular,
hexagonal, mixed, diamond, and Kagome unit cell shapes. The octagonal
out by Wang and McDowell [28], several geometries such as square, tri- honeycomb showed yield stress and elastic modulus values close to
angular, hexagonal, Kagome, diamond, and mixed were considered for those of hexagonal honeycomb and lower than the other above-men-
honeycomb structures and were investigated analytically using Euler- tioned honeycombs.
Bernoulli beam theory. Comparison of the elastic properties of the oc-
tagonal honeycomb with other types of honeycombs provides good in- References
sight on the potential use of this honeycomb to the designer.
Considering Fig. 11, it is obvious that the octagonal honeycomb has a [1] M. Tanaka, K. Yoshizawa, A. Tsuruma, H. Sunami, S. Yamamoto, M. Shimomura, For-
mation of hydroxyapatite on a self-organized 3D honeycomb-patterned biodegrad-
relatively smaller yield stress and elastic modulus compared to most
able polymer film, Colloids Surf. A Physicochem. Eng. Asp. 313 (2008) 515–519.
of the above-mentioned geometries. In fact, the response of an octago- [2] N. Yang, L. Gao, K. Zhou, Simple method to generate and fabricate stochastic porous
nal honeycomb is very close to that of a hexagonal honeycomb which scaffolds, Mater. Sci. Eng. C 56 (2015) 444–450.
[3] L.J. Gibson, M.F. Ashby, Cellular Solids: Structure and Properties, Cambridge Univer-
is the most prevalent geometry for honeycombs. Although the elastic
sity Press, 1997.
modulus and yield stress of an octagonal honeycomb is very similar to [4] N. Chantarapanich, P. Puttawibul, S. Sucharitpwatskul, P. Jeamwatthanachai, S.
hexagonal honeycomb, the symmetry of the microstructure of octago- Inglam, K. Sitthiseripratip, Scaffold library for tissue engineering: a geometric eval-
nal honeycomb in its two main directions can lead to more similar uation, Comput. Math. Methods Med. 2012 (2012).
[5] X.-b. Su, Y.-q. Yang, Y. Peng, J.-f. Sun, Development of porous medical implant scaf-
post-yielding deformation in its two main directions than is the case folds via laser additive manufacturing, Trans. Nonferrous Metals Soc. China 22
for the post yielding deformation of the two main directions of (2012) s181–s187.
R. Hedayati et al. / Materials Science and Engineering C 69 (2016) 1307–1317 1317

[6] S. Van Bael, Y.C. Chai, S. Truscello, M. Moesen, G. Kerckhofs, H. Van Oosterwyck, J.-P. [27] Z. Zheng, J. Yu, J. Li, Dynamic crushing of 2D cellular structures: a finite element
Kruth, J. Schrooten, The effect of pore geometry on the in vitro biological behavior of study, Int. J. Impact Eng. 32 (1) (2005) 650–664.
human periosteum-derived cells seeded on selective laser-melted Ti6Al4V bone [28] A.-J. Wang, D. McDowell, In-plane stiffness and yield strength of periodic metal hon-
scaffolds, Acta Biomater. 8 (7) (2012) 2824–2834. eycombs, J. Eng. Mater. Technol. 126 (2) (2004) 137–156.
[7] I. Singh, The architecture of cancellous bone, J. Anat. 127 (Pt 2) (1978) 305. [29] S. Torquato, L. Gibiansky, M. Silva, L. Gibson, Effective mechanical and transport
[8] X. Li, C. Wang, W. Zhang, Y. Li, Fabrication and compressive properties of Ti6Al4V properties of cellular solids, Int. J. Mech. Sci. 40 (1) (1998) 71–82.
implant with honeycomb-like structure for biomedical applications, Rapid Prototyp. [30] S. Gulati, Effects of Cell Geometry on Thermal Shock Resistance of Catalytic Mono-
J. 16 (1) (2010) 44–49. liths, 1975 (SAE Technical Paper).
[9] J.-H. Bae, H.-R. Song, H.-J. Kim, H.-C. Lim, J.-H. Park, Y. Liu, S.-H. Teoh, Discontinuous [31] F.A. El-Sayed, R. Jones, I. Burgess, A theoretical approach to the deformation of hon-
release of bone morphogenetic protein-2 loaded within interconnected pores of eycomb based composite materials, Composites 10 (4) (1979) 209–214.
honeycomb-like polycaprolactone scaffold promotes bone healing in a large bone [32] I. Masters, K. Evans, Models for the elastic deformation of honeycombs, Compos.
defect of rabbit ulna, Tissue Eng. A 17 (19–20) (2011) 2389–2397. Struct. 35 (4) (1996) 403–422.
[10] N. Kakudo, A. Shimotsuma, S. Miyake, S. Kushida, K. Kusumoto, Bone tissue engi- [33] S. Goswami, On the prediction of effective material properties of cellular hexagonal
neering using human adipose-derived stem cells and honeycomb collagen scaffold, honeycomb core, J. Reinf. Plast. Compos. 25 (4) (2006) 393–405.
J. Biomed. Mater. Res. A 84 (1) (2008) 191–197. [34] S. Balawi, J. Abot, A refined model for the effective in-plane elastic moduli of hexag-
[11] A.P. Rodriguez, L. Missana, H. Nagatsuka, M. Gunduz, H. Tsujigiwa, R. Rivera, N. onal honeycombs, Compos. Struct. 84 (2) (2008) 147–158.
Nagai, Efficacy of atelocollagen honeycomb scaffold in bone formation using [35] T. Daxner, Finite element modeling of cellular materials, In Cellular and Porous Ma-
KUSA/A1 cells, J. Biomed. Mater. Res. A 77 (4) (2006) 707–717. terials in Structures and Processes, Springer 2010, pp. 47–106.
[12] J. George, Y. Kuboki, T. Miyata, Differentiation of mesenchymal stem cells into oste- [36] R. Hedayati, M. Sadighi, M. Mohammadi-Aghdam, A.A. Zadpoor, Effect of mass mul-
oblasts on honeycomb collagen scaffolds, Biotechnol. Bioeng. 95 (3) (2006) tiple counting on the elastic properties of open-cell regular porous biomaterials,
404–411. Mater. Des. 89 (2016) 9–20.
[13] L. Gibson, M. Ashby, G. Schajer, C. Robertson, The mechanics of two-dimensional cel- [37] S. Ahmadi, G. Campoli, S. Amin Yavari, B. Sajadi, R. Wauthlé, J. Schrooten, H.
lular materials, Proc. R. Soc. Lond. A Math. Phys. Sci. 382 (1782) (1982) 25–42. Weinans, A.A. Zadpoor, Mechanical behavior of regular open-cell porous biomate-
[14] S.D. Papka, S. Kyriakides, In-plane compressive response and crushing of honey- rials made of diamond lattice unit cells, J. Mech. Behav. Biomed. Mater. 34 (2014)
comb, J. Mech. Phys. Solids 42 (10) (1994) 1499–1532. 106–115.
[15] S.D. Papka, S. Kyriakides, In-plane crushing of a polycarbonate honeycomb, Int. J. [38] R. Hedayati, M. Sadighi, M. Mohammadi-Aghdam, A. Zadpoor, Mechanical behavior
Solids Struct. 35 (3) (1998) 239–267. of additively manufactured porous biomaterials made from truncated
[16] A. Karakoç, K. Santaoja, J. Freund, Simulation experiments on the effective in-plane cuboctahedron unit cells, Int. J. Mech. Sci. 106 (2016) 19–38.
compliance of the honeycomb materials, Compos. Struct. 96 (2013) 312–320. [39] S. Ahmadi, S. Yavari, R. Wauthle, B. Pouran, J. Schrooten, H. Weinans, A. Zadpoor, Ad-
[17] C.C. Foo, G.B. Chai, L.K. Seah, Mechanical properties of Nomex material and Nomex ditively manufactured open-cell porous biomaterials made from six different space-
honeycomb structure, Compos. Struct. 80 (4) (2007) 588–594. filling unit cells: the mechanical and morphological properties, Materials 8 (4)
[18] F. Scarpa, P. Panayiotou, G. Tomlinson, Numerical and experimental uniaxial loading (2015) 1871–1896.
on in-plane auxetic honeycombs, J. Strain Anal. Eng. Des. 35 (5) (2000) 383–388. [40] R. Hedayati, M. Sadighi, M. Mohammadi-Aghdam, A.A. Zadpoor, Mechanical proper-
[19] S.D. Papka, S. Kyriakides, Experiments and full-scale numerical simulations of in- ties of regular porous biomaterials made from truncated cube repeating unit cells:
plane crushing of a honeycomb, Acta Mater. 46 (8) (1998) 2765–2776. analytical solutions and computational models, Mater. Sci. Eng. C 60 (2016)
[20] H.X. Zhu, N. Mills, The in-plane non-linear compression of regular honeycombs, Int. 163–183.
J. Solids Struct. 37 (13) (2000) 1931–1949. [41] R. Hedayati, M. Sadighi, A. Zadpoor, Analytical Relationships for the Mechanical
[21] R. Hedayati, M. Sadighi, M. Mohammadi Aghdam, A.A. Zadpoor, Mechanical proper- Properties of Additively Manufactured Porous Biomaterials Based on Octahedral
ties of additively manufactured thick honeycombs, Materials 9 (8) (2016) 613. Unit Cells(Submitted) 2015.
[22] D. Ruan, G. Lu, B. Wang, T. Yu, In-plane dynamic crushing of honeycombs—a finite [42] R. Hedayati, M. Sadighi, M. Mohammadi-Aghdam, A.A. Zadpoor, Mechanics of addi-
element study, Int. J. Impact Eng. 28 (2) (2003) 161–182. tively manufactured porous biomaterials based on the rhombicuboctahedron unit
[23] J. Chung, A.M. Waas, Compressive response of honeycombs under in-plane uniaxial cell, J. Mech. Behav. Biomed. Mater. 53 (2016) 272–294.
static and dynamic loading, part 2: simulations, AIAA J. 40 (5) (2002) 974–980. [43] N. Chantarapanich, A. Laohaprapanon, S. Wisutmethangoon, P. Jiamwatthanachai, P.
[24] Y. Sun, N.M. Pugno, In plane stiffness of multifunctional hierarchical honeycombs Chalermkarnnon, S. Sucharitpwatskul, P. Puttawibul, K. Sitthiseripratip, Fabrication
with negative Poisson's ratio sub-structures, Compos. Struct. 106 (2013) 681–689. of three-dimensional honeycomb structure for aeronautical applications using se-
[25] M. Grediac, A finite element study of the transverse shear in honeycomb cores, Int. J. lective laser melting: a preliminary investigation, Rapid Prototyp. J. 20 (6) (2014)
Solids Struct. 30 (13) (1993) 1777–1788. 551–558.
[26] X.E. Guo, L.J. Gibson, Behavior of intact and damaged honeycombs: a finite element [44] O. Rehme, C. Emmelmann, Selective laser melting of honeycombs with negative
study, Int. J. Mech. Sci. 41 (1) (1999) 85–105. Poisson's ratio, J. Laser Micro/Nanoengineering 4 (2009) 128–134.

Vous aimerez peut-être aussi