Mechanics of Materials: F. Ongaro, P. de Falco, E. Barbieri, N.M. Pugno

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Mechanics of Materials 97 (2016) 26–47

Contents lists available at ScienceDirect

Mechanics of Materials
journal homepage: www.elsevier.com/locate/mechmat

Mechanics of filled cellular materials


F. Ongaro a, P. De Falco a, E. Barbieri a,∗, N.M. Pugno a,b,c,∗∗
a
School of Engineering and Materials Science, Queen Mary University of London, London, UK
b
Laboratory of Bio-Inspired and Graphene Nanomechanics, Department of Civil, Environmental and Mechanical Engineering,
University of Trento, Trento, Italy
c
Center for Materials and Microsystems, Fondazione Bruno Kessler, Povo (Trento), Italy

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

Article history: Many natural systems display a peculiar honeycomb structure at the microscale and nu-
Received 21 July 2015 merous existing studies assume empty cells. In reality, and certainly for biological tissues,
Revised 15 January 2016
the internal volumes are instead filled with fluids, fibers or other bulk materials.
Available online 15 February 2016
Inspired by these architectures, this paper presents a continuum model for composite
cellular materials. A series of closely spaced independent linear-elastic springs approxi-
Keywords:
Composite mates the filling material.
Cellular material Firstly, finite element simulations performed on the microstructure and numerical ho-
Winkler model mogenization demonstrate a convergence towards non-micro polarity, in contrast to clas-
Linear elasticity sical empty materials. Secondly, theoretical homogenization, in its most general setting,
Parenchyma tissue confirms that the gradients of micro-rotations disappear in the continuum limit.
Finite element method In addition, the resulting constitutive model remains isotropic, as for the non-filled
cellular structures, and reconciles with existing studies where the filling material is absent.
Finally, the model is applied for estimating the mechanical properties of parenchyma
tissues (carrots, apples and potatoes). The theory provides values for Young moduli rea-
sonably close to the ones measured experimentally for turgid samples.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction protection and support for the body. Natural cellular ma-
terials have been studied widely, the pioneering textbook
The use of cellular structures allows good mechanical by Gibson and Ashby (2001) and the exhaustive work by
properties at low weight. Nature adopts this advantageous Gibson et al. (2010).
strategy on numerous occasions in biological systems The factors influencing the mechanical properties of
such as wood, bone, tooth, mollusk shells, crustaceans a cellular material are the apparent density, defined as
and many siliceous skeleton species like radiolarians, sea the ratio between the density of the cellular solid and
sponges and diatoms (Gordon et al., 2008; Neethirajan the density of the material, the internal architecture and
et al., 2009). Cellular structures have a diversity of func- the material properties of the microstructure. In its most
tions but a fundamental one is mechanical, providing sophisticated form, natural cellular materials are even
able to adapt their architectures to changing mechan-

Corresponding author at: School of Engineering and Materials Science,
ical environments (Fratzl and Weinkamer, 2007; 2011).
Queen Mary University of London, London, UK. Tel.: +4402078826310. One example is the combination of dense and strong
∗∗
Corresponding author at: Laboratory of Bio-Inspired and Graphene fibre-composite outer faces and an inner foam-like core
Nanomechanics, Department of Civil, Environmental and Mechanical En- in the leaves of monocotyledon plants, like irises and
gineering, University of Trento, Trento, Italy.
maize (Gibson, 2012), which better resist bending and
E-mail addresses: f.ongaro@qmul.ac.uk (F. Ongaro), p.defalco@
buckling. The lightweight core, in particular, is a simple
qmul.ac.uk (P. De Falco), e.barbieri@qmul.ac.uk (E. Barbieri),
nicola.pugno@unitn.it (N.M. Pugno). plant tissue, known as parenchyma (Gibson, 2012; Gibson

http://dx.doi.org/10.1016/j.mechmat.2016.01.013
0167-6636/© 2016 Elsevier Ltd. All rights reserved.
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 27

Fig. 1. (a) Parenchyma tissue from the moss Plagiomnium Affine, (b) Natural tubular structures: celery cross section.

et al., 2010; Bruce, 2003; Warner et al., 20 0 0; Georget (2002) and Wang and Stronge (1999) obtained the equiv-
et al., 2003) made up of thin-walled polyhedral cells filled alent (micropolar) constitutive equations associated with
by liquid and, notwithstanding its low relative density the homogenized continuum model of the discrete lattice.
and low mechanical properties, can significantly improve Dos Reis and Ganghoffer (2012) elaborated a variant of
the resistance of the leaf. Similar composite solutions are the asymptotic homogenization technique introduced by
found in the natural tubular structures, like plant stems or Caillerie et al. (2006) and calculated the effective behav-
animal quills. In these systems, the inner honeycomb or ior of a 2D square and hexagonal lattices through the
foam-like core behaves like an elastic foundation support- analysis of discrete sums. In accordance with the papers
ing the dense outer cylindrical shell and makes it more quoted above, the adopted technique leads to a micropolar
performant (Gibson, 2005; Dawson and Gibson, 2007), equivalent continuum. Adopting the viewpoint of homog-
Fig. 1. enization theory, Davini and Ongaro (2011) constructed a
It is well known (Gibson and Ashby, 2001) that the continuum model for the in-plane deformations of a hon-
main problem with cellular materials is to extract an eycomb material from general theorems of  -convergence.
equivalent continuum model. In fact, the formulation of Differently from (Chen et al., 1998), (Kumar and McDowell,
a continuum model is hindered by two types of diffi- 2004), (Warren and Byskov, 2002), (Wang and Stronge,
culties: the spatial variability of size and morphology of 1999), it emerges that the limit model is a pseudo-polar
the microscopic architecture, on one side, and the crucial continuum, that is a material that can undergo applied
passage from the microscopic discrete description to the distributed couples without developing couple stresses.
coarse continuous one, on the other. The typical approach Finally, Gonella and Ruzzene (2008) investigated the
to the continuum modeling of cellular materials includes equivalent in-plane properties of a square, hexagonal and
the assumption of periodicity and the selection of a repre- re-entrant lattices through homogenization techniques. In
sentative volume element (RVE). Moreover, it necessitates particular, the adopted homogenization approach yields
the application of micro-macro relationships in terms of the continuum set of partial differential equations associ-
forces and displacements and energy equivalence concepts. ated with their equivalent continuum model. The assump-
Furthermore, the material properties are typically assumed tion of no concentrated couples acting to the structure
to be as simple as possible, for example linear-elastic and in conjunction with the appropriate elasticity equations
isotropic (Altenbach and Oechsner, 2010). lead to the equivalent Young’s moduli and Poisson’s
Various authors extensively studied cellular materials. ratios.
Among them, Gibson and Ashby (2001), Gibson et al. In the literature, few investigations concern the char-
(2010), Gibson (1989), Gibson et al. (1982) obtained simple acterization of cellular materials having filled cells. Niklas
power-law relations between the density of a wide range (1989), for example, dealt with the mechanical behavior of
of honeycombs and foams and their mechanical proper- plant tissues and provided an analytical description of the
ties, studying regular cellular structures and assuming a influence of the turgor pressure on the effective stiffness.
prevalent mode of deformation and failure. Other authors, Georget et al. (2003) considered the stiffness of the carrot
like Chen et al. (1998), Davini and Ongaro (2011), Kumar tissue as a function of the turgor pressure and the mechan-
and McDowell (2004), Warren and Byskov (2002), Wang ical properties of the cell walls. By modeling the tissue
and Stronge (1999), Dos Reis and Ganghoffer (2012), Pugno as a fluid-filled foam, the authors found good agreement
(2006), Chen and Pugno (2012, 2011), Gonella and Ruzzene between their predictions and the experimental values.
(2008) focused on the passage from discrete to continuum Warner et al. (20 0 0), which goes in this direction, inves-
and deduced the constitutive model for the in-plane de- tigated a range of deformation mechanisms of closed-cell
formation of various two-dimensional microstructures (tri- cellular solids, unfilled and filled with liquid, to charac-
angle, hexagonal, rectangular, square, re-entrant or mixed terize elasticity and failure in foods. In particular, above a
lattices). (Chen et al., 1998) and (Kumar and McDowell, critical strain, it emerged that the filling liquid forced the
2004), in particular, considered an infinite arrangement of walls to stretch rather than to bend as they do when a
equal cells and derived a continuum model (a micropolar dry cellular solid is deformed. In the context of sandwich
continuum) introducing asymptotic Taylor expansion of panels, Burlayenko and Sadowski (2010) presented a finite
the displacement and rotation fields in the strain energy of element-based technique to evaluate the structural perfor-
the system. Applying the principles of structural analysis mance of sandwich plates with a hexagonal honeycomb
to the representative volume element, Warren and Byskov core made of aluminium alloy and filled with PVC foam.
28 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

Fig. 2. (a) The hexagonal microstructure, (b) The unit cell, (c) The triplet of elastic beams, (d) The beam on Winkler elastic foundation: degrees of freedom
in the local reference system.

Specifically, the investigation suggests that the effective between numerical and theoretical homogenization are
elastic properties of a honeycomb sandwich panel are also provided. Finally, Section 5 illustrates the application
improved by the presence of the filling PVC foam. The in- of the theoretical model to the biological parenchyma
plane crush response and energy absorption of a circular tissue of carrot, apple and potato.
cell honeycomb filled with PDMS elastomer is studied in To the authors’ best knowledge, this is the first time to
(D’Mello and Waas, 2013). It emerged an increase in the report such results. In fact, the beam on Winkler founda-
load carrying capacity and an improvement in the energy tion model has never been applied to represent the mi-
absorption capability due to the presence of the filling crostructure of a 2D composite cellular material.
material. More recently, Guiducci et al. (2014) investigated
the mechanics of a diamond-shaped honeycomb internally
pressurized by a fluid phase. The authors proposed a theo- 2. Equivalence between the biphasic continuum and
retical model based on the Born model, as well as a finite the hybrid system continuum - springs
element-based analysis to study the consequences of the
pressure within the cells. Two different effects emerged: a A sequence of elastic beams of length , forming a peri-
marked change in the lattice’s geometry and an improve- odic array of hexagonal cells (Fig. 2a), simulates a cellular
ment in the load-bearing capacity of the material. composite material having a hexagonal microstructure. The
This paper, inspired by the aforementioned high effi- Euler-Bernoulli beam on Winkler foundation model sim-
ciency of the composite structures in nature, deals with ulates each beam. Specifically, a series of closely spaced
the mechanical behavior of a 2D composite cellular mate- independent linear elastic springs of stiffness kw , the
rial subjected to in-plane loads. In particular, the analysis Winkler foundation constant per unit width, approximate
focuses on a honeycomb-like microstructure having the the elastic material filling the cells (Fig. 2c and d).
cells filled by a generic elastic material, represented by It should be noted that modeling the material within
a sequence of beams on Winkler elastic foundation. The the cells by a Winkler foundation is a simplification aimed
paper proceeds as follows. Firstly, Section 2 describes the at obtaining a more mathematically tractable problem.
equivalence between the hybrid system continuum-springs However, this could affect the prediction ability of the pro-
and the biphasic continuum-continuum. Numerical simu- posed model.
lations on a single composite cell show the implications Numerical simulations on a computational model of a
of the modeling approach based on the Winkler model. single composite cell provide the right value of the kw con-
Then, Section 3 focuses on the passage from the discrete stant. Also, the Finite Element (FE) simulations illustrate
description to the continuum one and the effective con- the implications of the modeling approach based on the
stitutive equations and elastic moduli are derived. Some Winkler model, Figs. 5 and 6, as well as the deformation
considerations about the influence of the microstructure mechanisms of the system continuum-springs, Fig. 7. As
parameters on the overall elastic constants are reported in Fig. 3 shows, two different configurations are considered.
Section 4, as well as a comparison between our model and In the first one, Fig. 3a, the filling material is represented
those available in literature. The results of the comparison by the Winkler foundation while in the second, Fig. 3b,
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 29

stress tensor and elastic modulus tensor of the material


within the cell, satisfying the relation

σ f = C f f , (7)

or

σ11 Ef
σ22 =
( 1 + ν f )(1 − 2ν f )
σ12

(1 − ν f ) νf 0
× νf (1 − ν f ) 0 . (8)
Fig. 3. Equivalence between the elastic moduli of the filling material and 0 0 ( 1 − 2ν f )
corresponding spring. (a) Filling material as a Winkler foundation, (b) Fill-
ing material as a classical continuum.
From classical continuum mechanics, the deformation
of the filling material in the generic ni direction is (Fig. 3)

by a classical continuum having Young’s modulus Ef and  di


nTi  f ni = , i = 1, 2, 3 (9)
Poisson’s ratio of ν f . d
First of all, let us focus on a suitable relation between where  di is the elongation in the ni direction and d =

Ef , ν f and the Winkler foundation constant kw . As Fig. 3 3. The assumption
shows, the elastic energy of the cell, Wc , is obtained by
summing the contribution of the walls, Ww , and of the fill- Ui =  di , i = 1, 2, 3 (10)
ing material, Wf : leads to
Wc = Ww + W f . (1) Ui
nTi  f ni = → Ui = (nTi  f ni )d, i = 1, 2, 3. (11)
Assuming that the elastic energy of the cell walls is the d
same of the elastic beams, Fig. 3a, and in the case of filling Substituting (11) into (3) gives
material as a classical continuum, Fig. 3b, the equivalence
 1
T
3
1
d (nTi  f ni )T nTi Kw ni (nTi  f ni ) d =  C  f A. (12)
Ww, beams ≡ Ww, continuum (2) 2 2 f f
i=1

takes the form Considering the deformation states


  
W f, W inkler ≡ W f, continuum . (3) 1 0 0
f = 0 , f = 1 , f = 0 (13)
Note that in (2) and (3) Ww, beams , Wf, Winkler and
0 0 1
Ww, continuum , Wf, continuum stand for the elastic energies of
the cell walls and of the filling material in the cases of and substituting, in turn, (13) into (12) provides
Winkler foundation model and of continuum model, re- √
3 3 Kw E f (ν f − 1 )
spectively. In particular, Wf, Winkler derives from the elastic = ,
energies of the three sets of springs in the directions n1 , 8 2(2ν f − 1 )(ν f + 1 )

n2 , n3 (Fig. 3a): 3 Kw Ef
  = . (14)

3 2 (ν f + 1 )
1
W f, W inkler =  UTi · Kw  Ui b, (4)
Accordingly,
2
i=1 √
5 3 Kw
with  Ui the elongation of the springs in the ni direction, ν f = 0.25, Ef = (15)
b the width, 8
  and, in the case of isotropic filling material, the shear mod-
Kw 0 ulus takes the form
Kw = (5)
0 Kw √
Ef 3 Kw
the stiffness matrix of the elastic foundation, Kw = kw  the Gf = = . (16)
2 (ν f + 1 ) 4
Winkler constant and  the length of the walls. It should
be noted that in Fig. 3a, for ease of reading, the series of As clearly seen, this theory leads to a fixed Poisson’s ratio,
springs are represented by three single springs having stiff- known result from the Spring Network Theory (Alzebdeh
ness Kw in the directions n1 , n2 , n3 . and Ostoja-Starzewski, 1999), (Ostoja-Starzewski, 2002).
Moreover, Regarding the cell considered in the numerical simula-
  tions, Kw = 10−1 GPa. The walls, assumed to be isotropic
1 1
W f, continuum = σ Tf  f dV = Tf C f  f dV, (6) linear elastic, have Young’s modulus Es = 79 GPa, Poisson’s
2 V 2 V ratio νs = 0.35, thickness h = 1 mm, length  = 10 mm and

3 3 2 unitary width. As more fully described in Section 4, each
been V = A b and A = , respectively, the volume and node has three degrees of freedom: two in-plane displace-
2
the area of the cell, f , σ f , Cf , in turn, the strain tensor, ments and the rotational component. The three loading
30 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

states analyzed, Fig. 4, are simulated by applying uniax-


ial forces of the same intensity at the boundary nodes.
Specifically, forces of intensity, 10−3 N in the directions e1 ,
Fig. 4a, and e2 , Fig. 4b, and shear forces of 10−5 N, Fig. 4c
are adopted. As Fig. 4 shows, constrained nodes are intro-
duced to avoid rigid body motions that could lead to an er-
roneous comparison between the deformed configurations
of the two considered models.
The results of the analysis, summarized in Figs. 5 and
6, illustrate that the predictions of the Winkler foundation
model are in accordance with those obtained in the case
of filling material as a classical continuum.
Specifically, in terms of the horizontal and vertical dis-
placements of the nodes, UX and UY respectively, the dif-
ference between the two estimates, UX and UY , is gen-
erally 1 − 3 % (Fig. 6).
Furthermore, in the case of uniaxial compression in
the e1 direction it emerges an higher difference between
the horizontal displacements of node 1 predicted by the
two models, Fig. 6a. This is due to the limitations of the
Winkler foundation model, where the elastic springs only
connect two opposite beams, 1-2 and 4-5, 6-1 and 3-4
(Fig. 3a). The beams 1-2 and 3-4, 6–1 and 4-5 that, in re- Fig. 4. Finite element implementation of the composite cell, the load con-
ditions. (a) Uniaxial compression in the e1 direction, (b) Uniaxial com-
ality, are coupled by the presence of the filling material,
pression in the e2 direction, (c) Shear forces.
in the Winkler model are not connected. However, in view
of Figs. 5a and 6a, the consequences of this simplification
(3), linked by the elastic beams (0)-(1), (0)-(2) and (0)-(3),
slightly affect the prediction ability of the proposed model.
represented by the vectors b1 = l1 − s, b2 = l2√− s, b3 =
Similarly, in the case of vertical compression, Fig. 4b,
−s. The area of the unit cell is A0 = |l1 × l2 | = 3 2 3 2 .
the high values of UX and UY at nodes 1, 4 are related
The analysis of the representative cell of the mi-
to the limitations induced by the Winkler model.
crostructure provides the strain energy density of the dis-
Regarding Fig. 6c, the same considerations apply. That
crete structure. Its continuum approximation is then the
is to say, the missing influence of the filling material on
consequence of a particular assumption.
the beams 1-6 and 4-5 provides an high value of UX at
the nodes 5 and 6.
As a conclusion, it can be said that the analysis reveals 3.1.2. The hybrid system: Euler–Bernoulli beam on Winkler
the validity of the modeling approach based on the Win- foundation
kler model, notwithstanding the simplifications introduced. Let us consider the beam element in the 2D space. Ne-
glecting, for simplicity, the second order effects and the
3. Theoretical model and homogenization of the shear deformability, it is subjected to bending and axial
continuum - springs system deformation. The hypothesis of two-dimensional structure
also prevents the possibility of torsion and bending in the
3.1. Geometry and energy of the discrete problem normal plane. The fields of axial and transverse displace-
ment of its axis and the rotation of its section describe
3.1.1. Geometric description the kinematics of a generic element. In a global reference
The hexagonal microstructure of the composite mate- system, defined by the unit orthonormal vectors e1 and
rial considered here can be described as the union of two e2 with origin in O and by the Cartesian coordinate sys-
simple Bravais lattices, tem X = (X, Y )T , the configuration of the eth structural el-
ement is known by specifying the coordinates of its end
L1 ( ) = x ∈ R2 : x = n1 l1 + n2 l2 , with (n1 , n2 ) ∈ Z2 ,
nodes I and J. According to the kinematics of the incident
L2 (  ) = s + L1 (  ), (17) beams, each node has three degrees of freedom, two trans-
lations and one rotation (Fig. 2d). Finally, forces arbitrar-
simply shifted with respect to each other by the shift vector
ily oriented and couples act at the extreme nodes of each
s. In Cartesian coordinates,
beam.
√ √ √
s = ( 3 /2, /2 ), l1 = ( 3 , 0 ), l2 = ( 3 /2, 3 /2 ) To analyze the generic structural element is more con-
venient to use a local reference system, specific to the
(18)
considered beam and closely dependent to its geometry.
describe the shift vector and the lattice vectors, l1 and l2 Such reference consists of two orthonormal unit vectors
(Fig. 2a). (ηe1 , ηe2 ), Figs. 2c and d, and a coordinate system (x, y).
The lattice vectors and the shift vector define the unit Note that, in the sequel, the extreme nodes of the beam
cell of the periodic array (Fig. 2b), which is composed by in the local notation are denoted by the indices i and j,
the central node (0) and the three external nodes (1), (2), Fig. 2d.
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 31

a 20
Filling material as a
Winkler foundation
15
mm Filling material as a
10 classical continuum

-5
-5 0 5 10 15 20 25
mm
b 20
Filling material as a
Winkler foundation
15
mm Filling material as a
10 classical continuum

-5
-5 0 5 10 15 20 25
mm
c 20
Filling material as a
Winkler foundation
15
mm
10 Filling material as a
classical continuum
5

-5
-5 0 5 10 15 20 25
mm
Fig. 5. Filling material as a Winkler foundation vs. filling material as a classical continuum, superposition of the deformed configurations in the case of
(a) Uniaxial compression in the e1 direction with forces of 10−3 N, (b) Uniaxial compression in the e2 direction with forces of 10−3 N, (c) Shear forces of
10−5 N, and with Kw = 10−1 Pa, Es = 79 GPa, ν s = 0.35, h = 1 mm, l = 10 mm.
32 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

a 4 ∆ UX
∆ UY
3
%
2

0
1 2 3 4 5 6 1
node number
b 3
∆ UX
∆ UY
2
%
1

0
1 2 3 4 5 6 1
node number
c 2.5
∆ UX
2
∆ UY
%
1.5

0.5

0
1 2 3 4 5 6 1
node number

Fig. 6. Filling material as a Winkler foundation vs. filling material as a classical continuum, comparison between the nodal displacements in the case of
(a) Uniaxial compression in the e1 direction with forces of 10−3 N, (b) Uniaxial compression in the e2 direction with forces of 10−3 N, (c) Shear forces of
10−5 N, and with Kw = 10−1 Pa, Es = 79 GPa, ν s = 0.35, h = 1 mm, l = 10 mm.

The introduction of the local reference system allows us with Kw = kw , kw the Winkler foundation constant per
to define more easily the stiffness matrices keb and kew f that Es h Es h3
unit width, C = and D = , respectively, the
1−νs2 12(1−νs2 )
are, respectively, the stiffness matrix of the classical elastic tensile and bending stiffness (per unit width) of the beams,
beam and of the Winkler foundation (Janco, 2010), denoted Es , ν s , h and , in turn, the Young’s modulus, Poisson’s ra-
by lowercase letters since they are expressed in the local tio, thickness and length of the cell walls.
reference. Their components are
⎡ C / 0 0 −C / 0 0


⎢ 0 12D /3 6D /2 0 −12D /3 6D /2 ⎥
⎢ ⎥
⎢ 0 6D / 2
4D / −6D / 2
2D / ⎥
⎢ 0 ⎥
keb = ⎢ ⎥ (19)
⎢−C / 0 0 C / 0 0 ⎥
⎢ ⎥
⎣ 0 −12D /3 −6D /2 0 12D /3 −6D / 2⎦

0 6D /2 2D / 0 −6D /2 4D /


and
⎡0 0 0 0 0 0

⎢0 13 Kw /35 11 Kw /210 0 9 Kw /70 −13 Kw /420⎥
⎢ ⎥
⎢0 11 Kw /210 Kw 2 /105 13 Kw /420 −Kw 2 /140 ⎥
⎢ 0 ⎥
kew f =⎢ ⎥, (20)
⎢0 0 0 0 0 0 ⎥
⎢ ⎥
⎣0 9 Kw /70 13 Kw /420 0 13 Kw /35 −11 Kw /210 ⎦
0 −13 Kw /420 −Kw 2 /140 0 −11 Kw /210 Kw 2 /105
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 33

The elastic energy of each beam, obtained by superposi-


tion principle due to the assumption of linear elastic beam,
is
1 e T e e 1 1
 
we = ( u ) · kb u + ( ue,a )T · kew f  ue,a
2 2 2

1 1

+ ( ue,b )T · kew f  ue,b , (21)
2 2
where ue = [ui , u j ]T = [ui , vi , ϕi , u j , v j , ϕ j ]T is the general-
ized vector of nodal displacement expressed in the local ref-
erence and
 T
 ue,a =  uai ,  uaj
 T
=  uai ,  vai ,  ϕia ,  uaj ,  vaj ,  ϕ aj , (22)

 T
 ue,b =  ubi ,  ubj
 T
=  ubi ,  vbi ,  ϕib ,  ubj ,  vbj ,  ϕ bj (23)

is the elongation of the springs a, the first, and of


the springs b, the second (Figs. 8 and 9). In particular,
the springs a and the springs b connect each beam with
the opposite one in the −ηe2 and +ηe2 direction, respec-
tively. See Appendix A for further details. Note that the fac-
tor 1/2 in the second and third term of (21), is due to the
fact that each spring is shared by two opposite beams and
contribute only half of its strain energy to the unit cell.
In the local reference, the forces and couples acting at
the end of each beam are

fe = keb ue + kew f  u + kew f  u ,


e,a e,b
(24)

with fe = [fi , f j ]T = [ fxi , fyi , mi , fx j , fy j , m j ]T the vector of


nodal forces and couples and ue ,  ue, a ,  ue, b , keb and kew f
Fig. 7. The hybrid system continuum-springs: filling material as a Win-
the same as before. Also in this case, forces and couples
kler foundation. Longitudinal stress (GPa) along the beams with Kw =
are obtained by superposition principle and are the sum of 10−1 GPa in the case of (a) Uniaxial compression in the e1 direction and
three terms. The first one, keb ue , corresponding to the clas- applied forces of 10−3 N, (b) Uniaxial compression in the e2 direction and
sical elastic beam, the second and the third, kew f  u and
e,a applied forces of 10−3 N, (c) Shear forces of 10−5 N.

e,b
kew f  u , related to the Winkler foundation.
where f0 and fj are, respectively, the forces and moments
of the central and external nodes j = 1, 2, 3. A mere parti-
3.1.3. Energetics of the discrete system
tion of keb and kew f leads to the matrices
The elastic energy of the unit cell, W, derives from that
of the three beams it consists of and it depends on the 
C / 0 0
displacements and rotations of the external nodes. kb,11 = 0 12D /3 −6D /2 ,
First of all, it is not difficult to see that the first node 0 −6D /2 4D /
of each beam coincides with the central node (0). There- 
fore, denoted by u0 the displacements of the node (0) and 0 0 0
a b kw f,11 = 0 13Kw /35 −11Kw /210 ,
by  u0 ,  u0 the elongation of the springs in (0), fol-
a a b b 0 −11Kw /210 Kw 2 /105
lows ui = u0 ,  ui =  u0 and  ui =  u0 . Moreover, (24) 
takes the form −C / 0 0
    kb,10 = 0 −12D /3 −6D /2 ,
f0 kb,01 u j + kb,00 u0 0 6D /2 2D /
fe = =
fj kb,11 u j + kb,10 u0 
  0 0 0
kw f,01  + kw f,00 
a a
uj u0 kw f,10 = 0 9Kw /70 13Kw /420 ,
+
kw f,11  + kw f,10 
a a
uj u0 0 −13Kw /420 −Kw 2 /140
  
k  u j + kw f,00  u0
b b −C / 0 0
+ w f,01 b , (25) kb,01 = 0 −12D /3 6D /2 ,
kw f,11  u j + kw f,10  u0
b
0 −6D /2 2D /
34 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

Fig. 8. The triplet of elastic beams with focus on springs. (a) Beam (0)-(1), (b) Beam (0)-(2), (c) Beam (0)-(3).

Fig. 9. The two sets of springs connecting the triplet of elastic beams. (a) Springs a, (b) Springs b.


0 0 0 3.2. The continuum model
kw f,01 = 0 9Kw /70 −13Kw /420 ,
0 13Kw /420 −Kw 2 /140 3.2.1. Energy
 The assumption that in the limit  → 0 there exist the
C /ell 0 0
continuous displacement and microrotation fields, u ˆ (· ) and
kb,00 = 0 2D /3 6D /2 ,
ϕˆ (· ), and that the discrete variables (uj , ϕ j ) previously in-
0 6D /2 4D /
 troduced to represent the degrees of freedom (displace-
0 0 0 ments and rotations) of the external nodes of the unit cell
kw f,00 = 0 13Kw /35 11Kw /210 . (26) can be expressed by
0 11Kw /210 Kw 2 /105
uj = u
ˆ0 + ∇ uˆ b j , ϕ j = ϕˆ 0 + ∇ ϕˆ b j , j = 1, 2, 3, (28)

Then, expressing (24) in the global reference (see provides the continuum description of the discrete struc-
Appendix B), adding up forces at the central node (0) and ture. The terms bj in (28) are the vectors formerly defined
condensing the corresponding degrees of freedom to take while u ˆ 0 and ϕˆ 0 are the values of u ˆ (· ) and ϕˆ (· ) at the cen-
account of the forces balance in (0), as in (Davini and On- tral point of the cell in the continuum description. Substi-
garo, 2011), leads to tuting (28) into (7) gives the strain energy of the unit cell
as a function of the fields u ˆ and ϕˆ . Note that the contin-
uous displacement u ˆ (· ) is referred to the global reference
W = W ( u 1 , u 2 , u 3 ,  u 1 ,  u2 ,  u3 ,  u1 ,  u 2 ,  u 3 ) .
a a a b b b
and uˆ (· ) stands for U ˆ 0 and ∇ u
ˆ (· ). Also in (28), uj , u ˆ stands
(27) for Uj , Uˆ 0 and ∇ U ˆ . The use of lowercase letters simplifies
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 35

the notation and, in what follows, u ˆ and ϕˆ stand for u ˆ0 In particular, it emerges that σ is a non-symmetric ten-
and ϕˆ 0 . sor and its components are
Dividing the expression that turns out from the calcula- ∂w ∂ w ∂εαβ ∂ w ∂ω
σγ δ = = + , α , β , γ , δ = 1, 2, (35)
tion by the area of the unit cell, A0 , gives the strain energy ∂ uˆγ ,δ ∂εαβ ∂ uˆγ ,δ ∂ω ∂ uˆγ ,δ
density in the continuum approximation w. The following where w is the strain energy density defined in (33). By
relation summarize the adopted procedure: observing that  
W ( U, ϕ ) ∼ W ( u
ˆ , ϕˆ ) ∂w ∂εαβ 1 ∂w ∂w ∂w
= = w. (29) = + = ≡ σγsym
δ ,
A0 A0 ∂εαβ ∂ uˆγ ,δ 2 ∂εγ δ ∂εδγ ∂εγ δ
In particular, the resulting energy density ∂w ∂ω 1 ∂w


= δ1γ δ2δ − δ2γ δ1δ
w=w εαβ , (ω − ϕˆ ), ϕˆ ,α (30) ∂ω ∂ uˆγ ,δ 2 ∂ω
1 ∂w
is a function of the infinitesimal strains εαβ = 2 ( u α ,β
1
+ ˆ = e ≡ σγskw δ , (36)
2 ∂ω γ δ
uˆβ ,α ) and the infinitesimal rotation ω = 12 (uˆ1,2 − uˆ2,1 ) that with eγ δ the alternating symbol (e11 = e22 = 0, e12 =
represent, respectively, the symmetric and skew-symmetric −e21 = 1) and δ ij the Kronecker delta (δi j = 1 if i = j, δi j =
part of ∇ u ˆ as in the classical continuum mechanics. ϕˆ ,α 0 if i = j), follows
are the microrotation gradients. Its explicit expression is


2 4
ε11
2
+ ε22
2
(C  + 36DC 2 ) + 2ε11 ε22 (C2 4 − 12DC 2 ) + 96DC 2 ε12 2
+ 48DC 3 ε12 ϕˆ ,2
w= √
4 33 (12D + C 2 )


2
24DC  (ε22 − ε11 )ϕˆ ,1 + 8D  (3D + C 2 ) ϕˆ ,21 + ϕˆ ,22 + 12D (12D + C 2 ) ω − ϕˆ
3 2
+ √
4 33 (12D + C 2 )


2

Kw 813 ε11 + ε22 2
+ 1224ε12 2
+ 402ε11 ε22 Kw  −12ε12 ϕˆ ,2 + (ϕˆ ,21 + ϕˆ ,22 ) + 6ϕˆ ,1 (ε11 − ε22 )
+ √ + √ . (31)
2496 3 sym 96 3
σγ δ = σγ δ + σγ δ , γ , δ = 1 , 2 ,
skw
(37)
Es (h/ )
After rewriting (31) in terms of c ≡ C / = and d ≡
1−νs2 that is
Es (h/ )3    
D /3 = , it emerges that in the energy obtained ∂w 1 ∂w
12(1−νs2 ) σ1δ = σ1sym +
1
e , σ2 δ = σ
sym
+ e2δ , δ = 1 , 2 . (38)
from the calculation the coefficients scale with different δ 2 ∂ω 1δ 2δ 2 ∂ω

order in , as in (Davini and Ongaro, 2011):



2
ε11
2
+ ε22
2
(c + 36 cd ) + 2ε11 ε22 (c2 − 12 cd ) + 96 cd ε12 2
+ 48 cd ε12 ϕˆ ,2
w= √
4 3(12 d + c )


2
24 cd (ε22 − ε11 ) ϕˆ ,1 + 8 d (3d + c ) 2 ϕˆ ,21 + ϕˆ ,22 + 12 d (12d + c ) ω − ϕˆ
+ √
4 3(12d + c )


2

Kw 813 ε11 + ε22 2
+ 1224ε12 2
+ 402ε11 ε22 Kw −12ε12 ϕˆ ,2 + 2 (ϕˆ ,21 + ϕˆ ,22 ) + 6 ϕˆ ,1 (ε11 − ε22 )
+ √ + √ . (32)
2496 3 96 3
Also,
Specifically, the coefficients in (32) are independent of ∂w ∂w ∂ (ω − ϕˆ ) ∂w
, with the exception of the microrotation gradients that = = . (39)
∂ω ∂ (ω − ϕˆ ) ∂ω ∂ (ω − ϕˆ )
scale with first order in . Consequently, in the limit  →
0 the contribution of the microrotation gradients is miss- Accordingly,
ing and the equivalent continuum is non-polar, differently
from (Chen et al., 1998). Accordingly, the strain energy
density in the continuum approximation takes the form:

2 4
ε11
2
+ ε22
2
(C  + 36DC 2 ) + 2ε11 ε22 (C2 4 − 12DC 2 ) + 96DC 2 ε12
w= √
4 33 (12D + C 2 )

2

2
3D ω − ϕˆ Kw 271 ε11 + ε22
2
+ 408ε12 2
+ 134ε11 ε22
+ √ + √ . (33)
3 3 832 3 (C2 2 + 36DC )ε11 + (C2 2 − 12DC )ε22
σ11 = σ11sym = √
2 3 (12D + C 2 )
3.2.2. Constitutive equations Kw (271ε11 + 67ε22 )
The constitutive equations + √ ,
416 3
1 ∂W
σ = , (34) (C 2 2 + 36DC )ε22 + (C2 2 − 12DC )ε11
A0 ∂∇ uˆ σ22 = σ22
sym
=  √
with σ the Cauchy-type stress tensor, ensue from (33). 2 3 (12D + C 2 )
36 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

Kw (271ε22 + 67ε11 ) or, in terms of strain,


+ √ ,   ∗ 
416 3 ε11 C11 ∗
C12 ∗
C13 σ11sym
48DC ε12 51Kw ε12 ε22 = C21
∗ ∗
C22 ∗
C23 σ22
sym
(45)
σ12
sym
= σ21
sym
= √ + √ ,
2 3 (12D + C 2 ) 52 3 ε12 ∗
C31 ∗
C32 ∗
C33 σ12
sym


3 D ω − ϕˆ with
σ12
skw
= −σ21skw
= ,
3 C22 C33
∗ ∗
σ12 = σ12 + σ12 , σ21 = σ21 + σ21
sym skw sym skw
, (40) C11 = C22 =
33 − C12 C33
2 C
C22 2

been σγ δ and σγskw
sym 2 3(1 − νs2 )(271(1 + λ2 )Kw (1 − νs2 ) + 104(λ + 3λ3 )Es )
δ , in turn, the symmetric and skew- = ,
symmetric parts of σ . (13Kw (1 − νs2 ) + 8λEs )(51(1 + λ2 )Kw (1 − νs2 ) + 104λ3 Es )
∗ ∗ C12 C33
C12 = C21 =−
33 − C12 C33
2 C
C22 2
3.2.3. Elastic constants √
The constitutive equations derived so far enables us to 4 3(1 − νs2 )(271(1 + λ2 )Kw (1 − νs2 ) − 208(λ3 − λ )Es )
=− ,
get the effective elastic constants in the continuum ap- (13Kw (1 − νs2 ) + 16λEs )(51(1 + λ2 )Kw (1 − νs2 ) + 208λ3 Es )
proximation by adopting the following approach. ∗
2
C22 − C12
2
C33 =
First of all, let us consider a stress state where σ 11 = 0, 2 C
C22 33 − C12 C33
2

σ22 = σ12 = σ21 = 0. In view of (40) and from Hooke’s law 52 3(1 − νs2 )(1 + λ2 )
= ,
σ11
sym
= E1∗ ε11 , the Young’s modulus in the e1 direction is: 51Kw (1 − νs2 )(1 + λ2 ) + 104 Es λ3

∗ ∗ ∗ ∗
C13 = C23 = C31 = C32 = 0. (46)
(13 Kw (1 − νs2 ) + 16λEs )(51(1 + λ2 )Kw (1 − νs2 ) + 208λ3 Es )
E1∗ = √ ,
4 3(1 − νs2 )(271(1 + λ2 )Kw (1 − νs2 ) + 208(λ + 3λ3 )Es )
The constants Cij , derived in Section 3, are listed as
(41)
C 2 2 + 36DC 271 Kw
with Es , ν s and λ = (h/ ) respectively, the Young’s modu- C11 = C22 = √ + √ ,
lus and the Poisson’s ratio of the cell walls material, and
2 3 (12D + C 2 ) 416 3
the ratio between the thickness and the length of the cell C 2 2 − 12DC 67 Kw
C12 = C21 = − √  + √ ,
arms. 2 3 (12D + C  ) 416 3
2
The related Poisson’s ratio ν12 ∗ = −ε /ε
22 11 takes the 48DC 51 Kw
form: C33 = √ + √ ,
2 3 (12D + C 2 ) 52 3
67(1 + λ2 )Kw (1 − νs2 ) − 208λ(λ2 − 1 )Es
ν12

= . (42)
C13 = C23 = C31 = C32 = 0.
271(1 + λ2 )Kw (1 − νs2 ) + 208λ(1 + 3λ2 )Es (47)

A stress state defined as σ 22 = 0, σ11 = σ12 = σ21 = 0 Finite element simulations on a computational model of
leads, with analogous calculations, to the Young’s modu- the microstructure assess the analytical approach. In par-
lus in the e2 direction and to the related Poisson’s ratio ticular, the Euler-Bernoulli beam on Winkler foundation
ν21
∗ = −ε /ε . In particular, it emerges that E ∗ = E ∗ ≡ E ∗
11 22 1 2
elements represent the composite hexagonal microstruc-
and ν12
∗ = ν ∗ ≡ ν ∗ , where E∗ and ν ∗ stands for the Young’s
21
ture. The cell wall material, isotropic linear elastic for as-
modulus, the first, and the Poisson’s ratio, the second, of sumption, has Young’s modulus Es and Poisson’s ratio ν s ,
the approximated continuum. respectively, of 79 GPa and 0.35, thickness h = 0.1 and
Moreover, the stress state σ12 = 0, σ11 = σ22 = 0
sym
Kw = 0.0 0 01Es .
yields the tangential elastic modulus G∗ = σ12 /2 ε12 :
sym
The displacements and the derived quantities at every
point within the beam are obtained by interpolation, using
51(1 + λ2 )Kw (1 − νs2 ) + 208λ3 Es the same shape functions:
G∗ = √ . (43)
208 3(1 + λ2 )(1 − νs2 )
u ( ξ ) = N ( ξ ) ue , (48)
Finally, the above moduli E∗ , ν∗, G∗
satisfy the classical re-
E∗ been ξ a parametric coordinate along the length of
lation G∗ = , typical of the isotropic materials. the beam (−1 ≤ ξ ≤ 1 ), u(ξ ) = [u(ξ ) v(ξ )] and ue =
2 (1 + ν ∗ )
[uei vei ϕie uej vej ϕ ej ], respectively, the displacements at any
4. Discussion value of ξ and the local displacements of the extreme
nodes of the beam,
4.1. Comparison between the analytical and numerical
 
N1 0 0 N2 0 0
homogenization N (ξ ) = (49)
0 N3 N4 0 N5 N6

The compact expression of the constitutive equations the shape functions matrix. In particular,
derived so far is
 sym   1−ξ 1+ξ
σ11 C11 C12 C13 ε11 N1 = , N2 = ,
2 2
σ22
sym
= C21 C22 C23 ε22 (44)
1 − ξ ( 3 − ξ 2 )/2
σ12
sym
C31 C32 C33 ε12 N3 = ,
2
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 37

(1 − ξ 2 )(1 − ξ ) 1 + ξ ( 3 − ξ 2 )/2 In Cartesian coordinates, s = (cos θ , sin θ ), been θ the


N4 = , N5 = , angle of inclination of the beam with respect to the x-axis.
8 2
(1 − ξ )(1 + ξ )
2 Accordingly,
N6 = − . (50)
8 ∂s ∂s
= cos θ δ1 j + sin θ δ2 j , = cos θ δ1i + sin θ δ2i ,
This study involves a 75 × 50 mm rectangular domain ∂xj ∂ xi
discretized in an increasing number of hexagonal cells of i, j = 1, 2, (59)
gradually smaller length . The load conditions considered
are the uniaxial compression and uniaxial traction in the with δ ij the Kroneker delta. The classical continuum me-
directions e1 and e2 , and pure shear. Also, forces acting at chanics provides the Young’s modulus, E1∗ , and the related
the unconstrained boundary nodes of the domain simulate Poisson’s ratio ν12
∗ :

the basic loading states. The corresponding effective stiff- (1 )


σ11 ε22
ness components are calculated as the ratio between the E1∗ = (1 )
, ν12

=− (1 )
. (60)
average volume strain, ε 11 ε11

1 With analogous calculations, forces acting vertically
εi j = εi j dV, i, j = 1, 2, (51)
V V provide
and the applied stress. Specifically, when the forces act ⎡ (2 ) ⎤   
horizontally, (45) leads to
ε 11 ∗
C11 ∗
C12 ∗
C13 0 ∗
C12 σ22
ε (2 ) ⎣ ( 2) ⎦
= ε 22 = C21∗ ∗
C22 ∗
C23 σ22 = C22 σ22

⎡ ⎤   
(1 )
ε11 ∗
C11 ∗
C12 ∗
C13 σ11 ∗
C11 σ11
(2 )
ε 12

C31 ∗
C32 ∗
C33 0 ∗
C32 σ22
(1 ) ⎣ (1 ) ⎦
ε = ε22 = C21 ∗ ∗
C22 ∗
C23 0 = C21 σ11 ,

(61)
(1 )
ε12

C31 ∗
C32 ∗
C33 0 ∗
C31 σ11
and
(52) (2 ) (2 ) (2 )

ε 11 ∗
ε22 ∗
ε 12
been σ 11 the applied stress and ε (1 ) the corresponding C12 = , C22 = , C32 = . (62)
σ22 σ22 σ22
strain vector. Consequently,
As before, σ 22 is the applied stress, ε (2 ) the corresponding

ε (1 ) ∗
ε (1 ) ∗
ε (1 ) strain vector and ε i(j2 ) the average volume strain defined in
C11 = 11 , C21 = 22 , C31 = 12 , (53)
σ11 σ11 σ11 (58). Regarding the elastic moduli,
where (2 )
 σ22 ε 11
(1 ) 1 (1 ) E2∗ = , ν21

=− . (63)
εi j = εi j dV, i, j = 1, 2, (54) (2 )
ε 22 (2 )
ε 22
V V

and V is the volume of the domain. Because of the present Finally, in the case of in-plane shear
analysis involves a domain with unitary width, composed ⎡ (3 ) ⎤   
by a sequence of discrete beams having the same length  ε 11 ∗
C11 ∗
C12 ∗
C13 0 ∗
C13 σ12
and the same thickness h, (54) turns out to be ε (3) = ⎣ε 22
(3 ) ⎦ ∗
= C21 ∗
C22 ∗
C23 0 ∗
= C23 σ12 ,
n b
 n b
 (3 )
ε 12

C31 ∗
C32 ∗
C33 σ12 ∗
C33 σ12
(1 ) m=1
εi(j1) (s ) ds m=1
εi(j1) (s ) ds
εi j = n b
 m
= m
, (64)
ds nb 
m=1 m
that leads to
i, j = 1, 2, (55)
(3 ) (3 ) (3 )

ε 11 ∗
ε22 ∗
ε 12
with s a parametric coordinate along the length of the C13 = , C23 = , C33 = . (65)
beam (0 ≤ s ≤ ),  the area of the domain and nb the
σ12 σ12 σ12
number of the beams. Again, σ 12 is the applied stress, ε (3 ) the corresponding
Remembering that strain vector and ε i(j3 ) the average volume strain defined in
 
1 ∂ ui ( s ) ∂ u j ( s ) (58). The relation
εi j ( s ) = + (56) σ12
2 ∂xj ∂ xi G∗ = (66)
(3 )
2 ε 12
and that
gives the tangential elastic modulus.
∂ ui ( s ) ∂ ui ( s ) ∂ s ∂ u j ( s ) ∂ u j ( s ) ∂ s
= , = , i, j = 1, 2, The results of our analysis are presented in Tables 1 and
∂xj ∂s ∂xj ∂ xi ∂ s ∂ xi 2, that provide a comparison between the analytical and
(57) numerical values of the Ci∗j constants, Tables 1, and of the
gives elastic moduli, Tables 2. As it can be seen, the results from
the continuum formulation compare reasonably well with
nb

 the numerical solutions.
1
m=1 2 (ui ( ) − ui (0 ) ) ∂∂xsj + u j ( ) − u j (0 ) ∂∂xsi
εi(j1) = m
. Nevertheless, the difference between the analytical re-
nb  sults and the numerical solutions slightly increases when
(58) the number of cells increases. This is not surprising, since
38 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

Table 1
Comparison between the analytical and numerical approach, Ci∗j constants.
∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗
No. cells  (mm) C11 C22 C12 C21 C33 C13 = C23 = C31 = C32

10 × 7 5 2.4 2.8 −2.2 −2.1 10.1 0


50 × 35 1 2.3 2.3 −2.3 −2.3 9.7 0
100 × 70 0.5 2.4 2.4 −2.3 −2.3 5.9 0
200 × 140 0.25 2.3 2.4 −2.3 −2.3 4.3 0
250 × 175 0.2 2.3 2.3 −2.2 −2.2 3.9 0
400 × 280 0.125 2.2 2.2 −2.1 −2.1 3.9 0
500 × 350 0.1 2.2 2.2 −2.1 −2.1 3.9 0
Analytical results 2.5 2.5 −2.4 −2.4 4.8 0

Table 2
Comparison between the analytical and numerical approach, elastic moduli.

No. cells  (mm) E1∗ (GPa) E2∗ (GPa) ν12



ν21

G∗ (GPa)

10 × 7 5 0.42 0.44 0.91 0.92 0.03


50 × 35 1 0.43 0.43 1.00 0.99 0.05
100 × 70 0.5 0.42 0.42 0.96 0.95 0.08
200 × 140 0.25 0.43 0.42 0.96 0.95 0.11
250 × 175 0.2 0.43 0.44 0.95 0.96 0.13
400 × 280 0.125 0.44 0.45 0.95 0.96 0.13
500 × 350 0.1 0.44 0.44 0.95 0.95 0.13
Analytical results 0.40 0.40 0.96 0.96 0.10

the assumptions of the theoretical model inevitably intro- (41), (42), (43) are now listed as
duce approximations that are as significant as the number √
∗ 4 3λ3 Es
of cells increases. That is, the condensation of the degrees E = , (67)
of freedom of the central node, the introduction of the lin- 3 (1 − νs2 )(1 + 3λ2 )
ear interpolants of nodal displacements and rotations in 1 − λ2
the discrete energy of the system to obtain the contin- ν∗ = , (68)
1 + 3λ2
uum model, the filling material represented by a Winkler

foundation. In particular, condensing the degrees of free-
∗ 3λ3 Es
dom of the central node subordinates the freedom of the G = . (69)
3 (1 − νs2 )(1 + λ2 )
nodes of lattice L2 to that of lattice L1 and, in contrast
to the numerical model, gives the two lattices a different Fig. 10 and Tables 3, based on a honeycomb made of an
role (Davini and Ongaro, 2011). Regarding the approxima- aluminum alloy with Es = 79 GPa and ν s = 0.35, illustrate
tion introduced by the linear interpolants, one could im- the results of the analysis. It emerges that the above ex-
prove the analytical predictions by the use of asymptotic pressions coincide with those in (Davini and Ongaro, 2011),
expansions as (Chen et al., 1998), (Gonella and Ruzzene, where the authors deduced the constitutive model for the
2008), (Dos Reis and Ganghoffer, 2012). Also, this could in-plane deformations of a honeycomb through an energy
lead to a micropolar continuum as in (Chen et al., 1998) approach in conjunction with the homogenization theory.
and (Dos Reis and Ganghoffer, 2012). Finally, as stated, In particular, they represented the microstructure as a se-
representing the filling material by a Winkler foundation quence of elastic beams and expressed the energy in the
is a simplification aimed at obtaining a mathematically continuum form by introducing the linear interpolants of
tractable problem. This obviously implies approximations nodal displacements and rotations in the discrete energy of
that could be improved by modeling the filling material by the system. From general theorems of  -convergence, the
a more complex model, as the Winkler–Pasternak founda- authors obtained the homogenized model that, in contrast
tion, (Limkatanyu et al., 2014), (Civalek, 2007), (Kerr, 1964). to (Chen et al., 1998), does not develop couple stresses.
Similarly to the present paper, (Davini and Ongaro, 2011)
4.2. Comparison with other works considered the equilibrium condition at the central node
and condensed the corresponding degrees of freedom.
Recent works focus on the mechanical characteriza- Furthermore, as Fig. 10 shows, (67) and (69) are in ac-
tion of cellular materials with an unfilled honeycomb mi- cordance with (Gibson and Ashby, 2001), where the equiv-
crostructure. A comparison between the proposed results alent elastic moduli are obtained by applying the principles
and the available analytical solutions for unfilled honey- of structural analysis to the representative volume element
combs verify the adopted modeling strategy. Obviously, it and by assuming a prevalent mode of deformation and fail-
is necessary to neglect the presence of the filling ma- ure.
terial, represented by a Winkler foundation, and assume Regarding (Gonella and Ruzzene, 2008), the authors de-
kw = 0. So, the composite cellular material of the present rived the equivalent properties of the lattice by focusing on
approach reduces to a traditional cellular material with a the partial differential equations associated with the corre-
honeycomb-like microstructure and the elastic moduli in sponding homogenized model. Differently from the present
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 39

Fig. 10. Comparison between the elastic constants obtained in the present paper and those of the other authors: (a) Young’s modulus, (b) Shear modulus,
(c) Poisson’s ratio.

Table 3
Expressions for the Young’s modulus, shear modulus and Poisson’s ratio derived from
the models that we have compared.

Chen et al. (1998)



∗ 2 λ ( 1 + λ2 ) ∗ 3 λ ( 1 + λ2 ) 1 + λ2
ECH
= √
GCH
= νCH

=
Es
3 (1 − νs2 )(3 + λ2 )
Es
12 (1 − νs2 ) 3 + λ2
Davini and Ongaro (2011)
√ √
∗ 4 3λ3 G∗DO 3 λ3 1 − λ2
EDO
= = νDO

=
Es
3 (1 − νs2 )(1 + 3λ2 ) Es
3 (1 − νs2 )(1 + λ2 ) 1 + 3 λ2
Dos Reis and Ganghoffer (2012)
√ √
∗ λ + 3 λ3 G∗DG 4 3 λ3 12 − λ2
EDG
= = νDG

=
Es
6 (1 − νs2 )(1 + λ2 ) Es
(1 − νs2 )(12 + λ2 ) 3 ( 4 + λ2 )
Gibson and Ashby (2001)
∗ 4 λ3 G∗GA λ3
EGA
Es
= √ Es
= √ νGA

=1
3 3
Gonella and Ruzzene (2008)
√ √
∗ 4 3λ3 G∗GR 3 λ3 1 − λ2
EGR
= = νGR

=
Es
3 ( 1 + 3 λ2 ) Es
3 ( 1 + λ2 ) 1 + 3 λ2
40 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

approach, their homogenization technique involves Taylor present study and those of the other authors, in the case
series expansions truncated to the second order of the dis- of filled honeycomb.
placements and rotations of the boundary nodes. Then, the As it can be seen, our results are in good accordance
assumption of no concentrated couples acting to the struc- with the predictions in (Murray et al., 2009), derived from
ture leads to a non polar continuum, slightly less stiff than a 2D finite element analysis of an aluminum honeycomb
that of the present paper. filled with a polymeric material. Similarly to the present
An alternative technique for the analysis of periodic approach, (Murray et al., 2009) modeled the cell walls as a
lattices is proposed in (Dos Reis and Ganghoffer, 2012). Euler-Bernoulli beam elements while the infill as a plane-
The authors derived the homogenized stress-strain rela- stress shell element.
tions under compression and shear for a 2D hexagonal Regarding (Burlayenko and Sadowski, 2010), the authors
lattice composed of extensional and flexural elements. In considered an aluminum honeycomb filled with a PVC
particular, the homogenization problem is performed in foam and obtained the equivalent elastic moduli using a
the framework of micropolar elasticity, where the inter- 3D finite element analysis and the strain energy homoge-
actions between two neighboring points involve both the nization technique of periodic media. Also, shell elements
Cauchy stress, as in classical mechanics, and the couple and brick elements represent, respectively, the cell walls
stress tensor (Eremeyev et al., 2013). Also, contrary to clas- and the filling material, both of them assumed isotropic. As
sical elasticity, the Cauchy tensor is not symmetrical. As Table 4 shows, the approach in (Burlayenko and Sadowski,
Fig. 10 shows, the homogenized elastic moduli proposed in 2010) leads to a slightly more stiff equivalent continuum
(Dos Reis and Ganghoffer, 2012) EDG ∗ , G∗ , ν
DG DG agree with than that of the present paper.
ours, E , G , ν , in the limit of slender beams. For instance,
∗ ∗ ∗ It should be noted that (15) provides the relation
after writing (67) in terms of the Young’s modulus in between the values of Kw and the Young’s modulus
(Dos Reis and Ganghoffer, 2012), EDG ∗ , it emerges that the of the filling material in (Murray et al., 2009), Ef, M ,
two estimates differ by a quantity related to the ratio λ be- and in (Burlayenko and Sadowski, 2010), Ef, B , listed in
tween the thickness and the length of the cell walls, that Table 4. Also, for sake of clarity, in Table 4 E1∗,M , E2∗,M
reduces as λ → 0: and E1∗,B , E2∗,B , G∗B stand for the elastic constants derived in
(Murray et al., 2009) and in (Burlayenko and Sadowski,
8λ2 (1 + λ )
∗ 2
EDG 2010), respectively.
E∗ =
2 . (70)
1 + 3λ2
4.3. The influence of microstructure parameters in the overall
In the case of (Gibson and Ashby, 2001) Young’s modu- properties
lus an analogous consideration applies:
As it can be noted from the expressions (41)–(43), the

EGA elastic moduli in the continuum approximation are obvi-

E =
. (71)
1 + 3λ2 (1 − νs2 ) ously related to the microstructure parameters. Such as the
Young’s modulus Es and the Poisson’s ratio ν s of the cell
An analysis of the micropolar behavior of the honey- walls material, the product Kw = kw  of the constant kw
comb microstructure is also provided in (Chen et al., 1998). of the Winkler foundation model and the length of the
The authors represented the lattice as a sequence of elastic cell arms, the ratio λ = h/ between the thickness and the
beams and derived the continuum model by means of an length of the beams.
asymptotic Taylor’s expansion of the nodal displacements Figs. 11 and 12, based on an aluminum alloy with
and rotations in the strain energy of the discrete structure. Es = 79 GPa and ν s = 0.35 as before, show, in turn, the
Differently from (Davini and Ongaro, 2011), (Gonella and influence of Kw and of λ in the macroscopic Young’s mod-
Ruzzene, 2008), (Dos Reis and Ganghoffer, 2012), the ap- ulus E∗ , shear modulus G∗ and Poisson’s ratio ν ∗ .
proach in (Chen et al., 1998) ignores the connectivity of the When λ is fixed, Fig. 11a and b suggest that both the ra-
beams and, in particular, the equilibrium conditions and tio E∗ /Es and G∗ /Es increase with increasing Kw . That is to
the condensation of the degrees of freedom of the central say, both the Young’s modulus and shear modulus increase
node. They also calculated the elastic energy of the discrete with increasing Kw and this is consistent with the result
structure by a superposition of the strain energies of the that one expects by increasing the stiffness of the material
individual beams. This technique predicts a much stiffer filling the cells (the parameter Kw ). Specifically, the big-
behavior than the results of the other authors (Fig. 10). ger λ, that corresponds to a beam that becomes more and
As a conclusion, the present approach leads to elastic more thick, the larger will be the increase of the moduli
moduli that are generally in accordance with those avail- E∗ and G∗ . Also, an high value of λ (λ = 0.2 ) leads to an
able in the literature, with the exception of (Chen et al., higher initial value of both E∗ and G∗ than that which oc-
1998). As pointed out, in (Chen et al., 1998) the equilib- curs for a small value of λ (λ = 0.02 ). Regarding the Pois-
rium conditions and the degrees of freedom of the central son’s ratio ν ∗ , Fig. 11c shows that for fixed λ, an increase
node are not considered. This states the major difference in Kw yields a decrease in the Poisson’s ratio. In particular,
with respect to the present approach, in conjunction with the smaller λ, the larger will be the decrease and, as it can
the use of an asymptotic expansions of the displacements be seen, the smaller λ, the smaller will be the value of Kw
and rotation fields. in correspondence of which ν ∗ reaches its minimum value:
In addition, Table 4 summarizes the outcome of the Kw ≈ 0.025 Es , Kw ≈ 0.35 Es , Kw ≈ 0.47 Es , respectively for
comparison between the effective elastic constants of the λ = 0.02, 0.05, 0.1. The initial decrease in ν ∗ , followed by
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 41

Table 4
Comparison between the elastic constants of the present paper and those of the other authors
in the case of filled microstructure.

(Murray et al., 2009) Present


Es = 70 GPa, Es = 70 GPa,
ν s = 0.35, ν s = 0.35,
λ = 0.1 λ = 0.1

Ef, M E1∗,M = E2∗,M Kw E1∗ = E2∗


(GPa) (GPa) (GPa) (GPa)
0.001 0.3÷0.4 0.001 0.35
0.01 0.3÷0.4 0.009 0.37
0.1 0.5 0.094 0.45
1 1.2÷1.3 0.94 1.3
(Burlayenko and Sadowski, 2010) Present
Es = 72.2 GPa, Es = 72.2 GPa,
ν s = 0.34, ν s = 0.34,
λ = 0.0125 λ = 0.0125
Ef, B E1∗,B = E2∗,B G∗B Kw E1∗ = E2∗ G∗
(GPa) (MPa) (MPa) (GPa) (MPa) (MPa)

0.056 0.627 0.238 0.053 0.610 0.235


0.105 0.788 0.282 0.099 0.760 0.267
0.230 1.061 0.386 0.216 0.885 0.338

Fig. 11. The influence of Kw in the elastic constants: (a) Young’s modulus, (b) Shear modulus, (c) Poisson’s ratio.
42 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

Fig. 12. The influence of λ in the elastic constants: (a) Young’s modulus, (b) Shear modulus, (c) Poisson’s ratio.

an almost horizontal line once Kw reaches a specific value, to say that the bigger Kw , the smaller will be the influence
is larger for high values of λ. In fact, the initial slope of the of λ. Only when the filling material is missing, Kw = 0, an
curves corresponding to λ = 0.1 and λ = 0.2 is bigger than increase in λ leads to a decrease in ν ∗
that corresponding to λ = 0.02 and λ = 0.05.
In terms of the influence of λ in the overall properties 5. Some practical applications
E∗ , G∗ , ν ∗ , Fig. 12a and b suggest that when Kw is fixed, the
ratio E∗ /Es and G∗ /Es generally increase with increasing λ. The high efficiency of the composite structures in na-
Namely, when the beam become thicker there will be an ture, like plants stems and parenchyma tissues inspired
increase in E∗ and G∗ . Nevertheless, to high values of Kw this paper. In particular, the parenchyma is a plant tissue
(0.25 Es , 0.5 Es for the Young’s modulus, 0.08 Es , 0.04 Es for composed by thin-walled polyhedral cells filled by a quasi-
the shear modulus) corresponds to an higher initial value incompressible fluid that exerts the hydrostatic pressure Pi
of both E∗ and G∗ than that which occurs for small values (known as turgor pressure) on the cell walls and is respon-
of Kw (0 Es , 0.05 Es for the Young’s modulus, 0 Es , 0.008 Es sible for the strength and rigidity of the cell (Van Liedek-
for the shear modulus). Regarding the Poisson’s ratio, from erke et al., 2010). It is acknowledged that the mechani-
Fig. 12c globally emerges that when Kw is fixed, an in- cal properties of the whole tissue are related to those of
crease in λ leads to an increase in ν ∗ . In particular, for high the individual components, as the constitutive equations of
values of Kw (0.5 Es , 0.25 Es ) the increase is less significant the cell walls, the turgor pressure and the cell-to-cell in-
than that occurring for small values of it (0.05 Es ). That is teractions (Zhu and Melrose, 2003). However, the physical
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 43

Table 5
Young’s modulus for apple parenchyma tissue.

E∗ (MPa)

Present
h/l = 0.02 ÷ 0.2 (by assumption) 1÷1.6
 kw = 0.3 ÷ 4.1MPa, Pi = 0.8MPa (Georget et al., 2003)
Es = 52.8MPa (Wu and Pitts, 1999), ν s = 0.24 (Wu and Pitts, 1999)
 = 1.5 μm (Sanchis Gritsch and Murphy, 2005)
Gibson et al. (2010) 0.31÷3.46

Table 6
Young’s modulus for potato parenchyma tissue.

E∗ (MPa)

Present
h/ = 0.0087,  = 1.5μm (Sanchis Gritsch and Murphy, 2005) 4÷4.2
 kw = 9.6 ÷ 9.62MPa, Pi = 0.8 MPa (Georget et al., 2003)
Es = 500 ÷ 600MPa, ν s = 0.5 (Niklas, 1992)
Gibson (2012) 5÷6
h/ = 0.0087, Es = 500 ÷ 600MPa
Experimental value (Gibson, 2012) 3.5÷5.5

Table 7
Young’s modulus for carrot parenchyma tissue, with Es = 100 MPa, ν s = 0.33.

E∗ (MPa)

Present
h/ = 0.02 ÷ 0.2,  kw = 0.3 ÷ 4.1MPa 1.7÷2
 = 1.5μ m (Sanchis Gritsch and Murphy, 2005)
Pi = 0.8MPa (Georget et al., 2003)
Gibson, Ashby model (Georget et al., 2003) 2.4÷32.8
h/ = 0.02 ÷ 0.35
Warner, Edwards model (Georget et al., 2003) Lower limit: 10−5 ÷ 1.6
h/ = 0.02 ÷ 0.35 Upper limit: 0.03÷14.2
Experimental value (Georget et al., 2003) 7±1

properties of the living cells are, in general, very difficult However, Tables 5–7 indicate that the proposed theoretical
to measure (Wu and Pitts, 1999) and, as a consequence, model, based on a 2D configuration, could be a useful tool
the analysis of the parenchyma tissue is extremely com- to gain some quantitative information on the mechanical
plex without great simplifications and assumptions. properties of some vegetative tissues.
Applying the results of our model to the parenchyma
tissue of carrot, potato and apple is a tool to validate the 6. Conclusions
present modeling approach. Specifically, the expression in
(41) gives the Young’s modulus of the whole tissue, start- This paper, inspired by the high efficiency of composite
ing from the experimental measurements of the cell prop- structures in nature, focuses on the analysis of the me-
erties, Young’s modulus, Poisson’s ratio, dimensions of the chanical behavior of a 2D composite cellular material, hav-
walls, turgor pressure values, and some considerations (see ing a honeycomb texture and the cells filled by a generic
Appendix C) to obtain an approximation of kw , the Winkler elastic material.
foundation constant. An approximate value of the length Within the framework of linear elasticity and by
of the carrot, potato and apple parenchyma cell walls de- modeling the microstructure as a sequence of beams on
rives from (Sanchis Gritsch and Murphy, 2005), where the Winkler elastic foundation, the constitutive equations and
authors describe the structure of the parenchyma cell at elastic moduli in the macroscopic continuum description
different stages of development. In the present paper, the are derived. Specifically, an energetic approach leads to
mature cell is considered. Regarding the value of the turgor the continuum representation. The assumption that, in
pressure, (Georget et al., 2003) obtain its values by focus- the limit, the variables defined to represent displacements
ing on the carrot parenchyma tissue. In our investigation, and rotations of the nodes can be expressed in terms of
for simplicity, the assumption of the same turgor pressure two continuous fields solved the passage from discrete to
value for all the considered parenchyma tissues holds true. continuum. Finally, the introduction of the aforementioned
Tables 5–7 show that our results generally agree with continuous fields in the strain energy function of the unit
the experimental published data and with the predictions cell and simple mathematical manipulations yield the
of other authors, derived from 3D models of fluid-filled, stress-strain relations and elastic constants. Obviously, the
closed-cell foams. It should be noted that a rigorous ex- overall properties are related to the microstructure param-
amination of the parenchyma tissue is beyond our aim. eters. The analysis developed to investigate such influence,
44 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

a b

Fig. A1. The triplet of elastic beams with focus on springs. (a) Beam (0)-(1), (b) Beam (0)-(2), (c) Beam (0)-(3).

reveals that the macroscopic mechanical behavior of the and, as it can be seen, is the sum of three terms. The first
medium is improved by increasing the stiffness of the one,
material that fills the cells. Also, from the finite element 1 e T e e
simulations performed to assess the analytical model, it ( u ) · kb u , (2)
2
emerges that the results from the continuum formulation
corresponding to the classical elastic beam, while the sec-
compare reasonably well with the numerical solutions.
ond and the third,
In addition, this paper presents the application of the
analytical results to the vegetative parenchyma tissue of

1 1

1 1
 
( ue,a )T · kew f  ue,a , ( ue,b )T · kew f  ue,b , (3)
carrot, apple and potato to derive the Young’s modulus of 2 2 2 2
the whole tissue. The values obtained generally agree with
related to the Winkler foundation and, in particular, to the
the published data and this modeling strategy could pro-
elongation of the springs a, the first, and of the springs b,
vide some qualitative information on the mechanical prop-
the second (Fig. A2).
erties of biological tissues.
Similarly, the forces and couples acting at the end of
To the authors’ best knowledge, the beam on Win-
each beam are
kler foundation has never been applied to model the mi-
fe = keb ue + kew f  u + kew f  u .
e,a e,b
crostructure of a cellular material to obtain an analytical (4)
expression of the equivalent elastic moduli of a filled cel-
In particular,
lular solid.
 T  T
 u1,a =  ua0  ua1 ,  u2,a =  ua0  ua2 ,
 T
Acknowledgment  u3,a =  ua0  ua3 (5)

E. Barbieri is supported by the Queen Mary University and


 T  T
of London Start-Up grant for new academics.  u1,b =  ub0  ub1 ,  u2,b =  ub0  ub2 ,
N. M. Pugno is supported by the European Research  T
Council (ERC StG Ideas 2011 BIHSNAM n. 279985, ERC PoC  u3,b =  ub0  ub3 . (6)
2013 KNOTOUGH n. 632277, ERC PoC 2015 SILKENE nr.
The elastic energy of the unit cell, W, derives from that of
693670), by the European Commission under the Graphene
the three beams it consists of. As stated, adding up forces
Flagship (Nanocomposites, n. 604391).
at the central node (0) and condensing the corresponding
degrees of freedom to take account of the forces balance
Appendix A in (0), leads to

W = W (u1 , u2 , u3 ,  u1 ,  u2 ,  u3 ,  u1 ,  u2 ,  u3 ).
a a a b b b
(7)
The elastic energy of each beam in the discrete system
is The assumption that in the limit  → 0 the discrete vari-
1 1 1
  ables (uj , ϕ j ) can be written as
w = (ue )T · keb ue +
e
( ue,a )T · kew f  ue,a
2 2 2 uj = uˆ 0 + ∇u

1 1
 ˆ bj
+ ( ue,b )T · kew f  ue,b , (1) ϕ j = ϕˆ 0 + ∇ ϕˆ b j , j = 1, 2, 3, (8)
2 2
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 45

a (0 iii ) iv
(0 )
b
i
(3,2 )
ii (2) (0 iii ) iv
(0 )
i
(2 ,1 )
ii i
(3,2 )
ii (2)
i ii
(2 ,1 )
ii
(0 )v
(0) (0 )
(0 v) (0 ii)
(3) (1) (0)
(3) (1)

e 2 e 2
(0 )
i
(0vi ) (0 )
i
(0vi )
e i ii e i ii
1 (1 ,3 ) 1 (1 ,3 )

Fig. A2. The two sets of springs connecting the triplet of elastic beams. (a) Springs a, (b) Springs b.

- Beam (0)-(3)
Discrete system
   
u 3 − u 3i u3 − u3ii
 ua3 = ,  ub3 = . (15)
ϕ3 − ϕ3i ϕ3 − ϕ3ii
Continuum description
u3 = uˆ + ∇u
ˆ b3 , ϕ3 = ϕˆ + ∇ ϕˆ b3 ,
ˆ + ∇u
u 3i = u ˆ b f 3, ϕ3i = ϕˆ + ∇ ϕˆ b f 3 , (16)
Fig. A3. The bfi vectors. ˆ + ∇u
u3ii = u ˆ b f 1, ϕ3ii = ϕˆ + ∇ ϕˆ b f 1 ,
and
   
provides the continuum description of the discrete struc- ∇ uˆ b3 − ∇ uˆ b f 3 ∇ uˆ b3 − ∇ uˆ b f 1
ture. The terms u ˆ 0 and ϕˆ 0 are the values of u ˆ (· ) and ϕˆ (· )  ua3 = ,  u3 =
b
. (17)
∇ ϕˆ b3 − ∇ ϕˆ b f 3 ∇ ϕˆ b3 − ∇ ϕˆ b f 1
at the central point of the cell in the continuum descrip-
tion. Note that in what follows, to simplify the notation, u ˆ Finally, the vectors bfi (Fig. A3) are
and ϕˆ stand for u ˆ 0 and ϕˆ 0 . Substituting (8) into (7) gives
b f 1 = −l2 + (l1 − s ),
the strain energy of the unit cell as a function of the fields
b f 2 = l1 + (l2 − s ), (18)
ˆ and ϕˆ .
u
b f 3 = ( l2 − l1 ) − s.
Specifically, the aforementioned quantities are (Figs. A1
and A3):
Appendix B
- Beam (0)-(1)
Discrete system
T
T
    Let Ue = UI , UJ = UI , VI , ϕI , UJ , VJ , ϕJ be the vector
u 1 − u 1i u1 − u1ii of nodal displacement expressed in the global reference and
 ua1 = ,  ub1 = . (9)
ϕ1 − ϕ1i ϕ1 − ϕ1ii
T
 Ue,a =  UaI ,  UaJ
In the continuum description,
T
u1 = uˆ + ∇u
ˆ b1 , ϕ1 = ϕˆ + ∇ ϕˆ b1 , =  UIa ,  VIa ,  ϕIa ,  UJa ,  VJa ,  ϕJa , (19)
ˆ + ∇u
u 1i = u ˆ b f 1, ϕ1i = ϕˆ + ∇ ϕˆ b f 1 , (10)  T
ˆ + ∇u
u1ii = u ˆ b f 2, ϕ1ii = ϕˆ + ∇ ϕˆ b f 2 ,  Ue,b =  UbI ,  UbJ
that, substituted in (9), lead to
T
    =  UIb ,  VIb ,  ϕIb ,  UJb ,  VJb ,  ϕJb , (20)
∇ uˆ b1 − ∇ uˆ b f 1 ∇ uˆ b1 − ∇ uˆ b f 2
 a
u1 = ,  u1 =
b
. (11) the elongation of the springs a, the first, and of the springs
∇ ϕˆ b1 − ∇ ϕˆ b f 1 ∇ ϕˆ b1 − ∇ ϕˆ b f 2
b, the second. One can prove that the relations
- Beam (0)-(2)
u e = Qe Ue ,  ue,a = Qe  Ue,a ,  ue,b = Qe  Ue,b (21)
Discrete system
    hold true, with Qe the rotation matrix that rotates the axis
u 2 − u 2i u2 − u2ii e1 , e2 in ηe1 , ηe2 , given by
 a
u2 = ,  b
u2 = . (12)
ϕ2 − ϕ2i ϕ2 − ϕ2ii ⎡ ⎤
ηe1 · e1 ηe1 · e2 0 0 0 0
Continuum description
⎢ηe2 · e1 ηe2 · e2 0 0 0 0⎥
u2 = uˆ + ∇u
ˆ b2 , ϕ2 = ϕˆ + ∇ ϕˆ b2 , ⎢ 0 0 1 0 0 0⎥
Qe = ⎢
⎢ 0
⎥, (22)
ˆ + ∇u
u 2i = u ˆ b f 2, ϕ2i = ϕˆ + ∇ ϕˆ b f 2 , (13) 0 0 ηe1 · e1 ηe1 · e2 0⎥
⎣ ⎦
ˆ + ∇u
u2ii = u ˆ b f 3, ϕ2ii = ϕˆ + ∇ ϕˆ b f 3 , 0 0 0 ηe2 · e1 ηe2 · e2 0
0 0 0 0 0 1
and
     T  T
∇ uˆ b2 − ∇ uˆ b f 2 ∇ uˆ b2 − ∇ uˆ b f 3 ue = ui , u j = ui , vi , ϕi , u j , v j , ϕ j the generalized vec-
 ua2 = ,  u2 =
b
. (14)
∇ ϕˆ b2 − ∇ ϕˆ b f 2 ∇ ϕˆ b2 − ∇ ϕˆ b f 3 tor of nodal displacement expressed in the local reference,
46 F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47

Fig. A4. (a) Degrees of freedom in the global reference system, (b) Forces and couples in the global reference. .

(Niklas, 1989) like a cylinder of length L with inner radius


ri and outer radius ro . Every cell is filled by a fluid that ex-
erts the hydrostatic pressure Pi , turgor pressure, on the cell
walls (Fig. A5).
Restricting the analysis to the 2D space, a circular thin-
walled flexible ring stiffened by the internal pressure Pi
can represent the parenchyma cell. Furthermore, when the
external pressure Po exceeds the internal pressure Pi , the
flexible ring fails through elastic instability. At the onset
of buckling, Po = Pi , the flexible ring behaves as a fictitious
stiff ring whose critical buckling load is
 
Fig. A5. The internal and external pressures in a parenchyma cell. 3 Ew I 1
Pcr = = Pi , (28)
 
a T
ri3 
u e,a
= 
a
ui ,  uj
 T been Ew and I = π (ro4 − ri4 )/4, respectively, the Young’s
=  uai ,  vai ,  ϕia ,  uaj ,  vaj ,  ϕ aj , (23) modulus of the cell walls of the stiff ring and the moment
of inertia of the cross section.
 T
From (28), the fictitious Young’s modulus is
 ue,b =  ubi ,  ubj
 T Pi ri3 
=  ubi ,  vbi ,  ϕib ,  ubj ,  vbj ,  ϕ bj (24) Ew (Pi ) = . (29)
3I
the elongation of the springs a, the first, and of the springs A 2D parenchyma cell can also be represented as a
b, the second. Similarly, the stiffness matrices expressed in composite material. That is, as a stiffened thin-walled cell
the global reference Keb and Kew f , can be easily derived by filled by an elastic material having Young’s modulus Ef .
the stiffness matrices expressed in the local reference, keb and From the rule of mixtures,
kew f : E eq = E f f + Es (1 − f ), (30)

Keb = (Qe )T keb Qe , Kew f = (Qe )T kew f Qe . (25) been Eeq


and Es , respectively, the Young’s modulus of the
composite structure and of the cell walls, f the porosity
Furthermore, the elastic energy in (1) takes the form given by
1 e T 1 1
 
We = (U ) · Keb Ue + ( Ue,a )T · Kew f  Ue,a Vf
2 2 2 f = . (31)
  V f + Vs
1 1
+ ( Ue,b )T · Kew f  Ue,b , (26) Also, Vs and Vf stand, in turn, for the volume of the walls
2 2
and of the material within the cell. In the case of cell thick-
if referred to the global reference. ness tending to zero Vs Vf . Accordingly, f ≈1 and
As regards the forces and couples acting at the ends of
each beam (Fig. A4b), in the global reference are E eq = E f . (32)
Furthermore, the equivalence between the biphasic contin-
Fe = Keb Ue + Kew f  U + Kew f  U ,
e,a e,b
(27)
uum and the hybrid system continuum-spring of Section 2,
been Fe = [FI , FJ ]T = [FXI , FY I , MI , FXJ , FY J , MJ ]T the vector of provide
nodal forces and couples expressed in the global reference, √
5 3
Ue ,  Ue, a ,  Ue, b , Keb and Kew f the quantities previously Ef = Kw , (33)
8
defined.
with Kw the stiffness of the filling material if modeled by a
Appendix C Winkler foundation. Substituting (33) into (32) and impos-
ing the equivalence between the Young’s modulus derived
A sequence of beams on Winkler elastic foundation from the two modeling strategies, Eeq and Ew , provides
simulates the microstructure of the composite material an-
8 Pi ri3 
alyzed in the present paper. Kw = √ . (34)
Focusing on a generic parenchyma cell provides a suit- 15 3 I
able approximation of Kw , the Winkler’s foundation con- Note that the relation in (29) is obtained in the case of
stant. Specifically, a parenchyma cell can be modeled cylindrical cell. Consequently, to correctly apply (29) to our
F. Ongaro et al. / Mechanics of Materials 97 (2016) 26–47 47

Gibson, L.J., 2005. Biomechanics of cellular solids. J. Biomech. 38, 377–399.


Gibson, L.J., 2012. The hierarchical structure and mechanics of plant ma-
terials,. J. R. Soc. Interface 9, 2749–2766.
Gibson, L.J., Ashby, M.F., 2001. Cellular Solids. Structure and Properties.
Cambridge University Press.
Gibson, L.J., Ashby, M.F., Harley, B.A., 2010. Cellular Materials in Nature
and Medicine. Cambridge University Press.
Gibson, L.J., Ashby, M.F., Schajer, G.S., Robertson, C.I., 1982. The mechanics
of two-dimensional cellular materials. Proc. R. Soc. Lond. A 382, 25–
42.
Gonella, S., Ruzzene, M., 2008. Homogenization and equivalent in-plane
properties of two-dimensional periodic lattices. Int. J. Solids Struct.
Fig. A6. The annulus approximating the hexagonal cell.
45, 2897–2915.
Gordon, R., Losic, D., Tiffany, M.A., Nagy, S.S., Sterrenburg, F.A.S., 2008. The
model, it is necessary to approximate the hexagonal cell of glass menagerie: diatoms for novel applications in nanotechnology.
the present paper with an annulus. Specifically, the annu- Trends Biotechnol. 27 (2), 116–127.
Guiducci, L., Fratzl, P., Brechet, Y.J.M., Dunlop, J.W.C., 2014. Pressurized
lus (Fig.√A6) has inner and
√ outer radii given, respectively, honeycombs as soft-actuators: a theoretical study. J. R. Soc. Interface
by ri = 3/2 and ro = 3/2 + h, where  is the length 11 (98).
and h the thickness of the beams in the original model. Janco, R., 2010. Solution methods for beam and frames on elastic founda-
tion using the finite element method. In: Proceedings of Mechanical
Structures and Foundation Engineering, International Scientific Con-
References
ference MSFE.
Kerr, A.B., 1964. Elastic and viscoelastic foundation models. J. Appl. Mech.
Altenbach, H., Oechsner, A., 2010. Cellular and Porous Materials in Struc- 31, 491–498.
tures and Processes. CISM. Kumar, R.S., McDowell, D.L., 2004. Generalized continuum modeling of 2-
Alzebdeh, K., Ostoja-Starzewski, M., 1999. On a spring-network model and d periodic cellular solids. Int. J. Solids Struct. 41, 7399–7422.
effective elastic moduli of granular materals. J. Appl. Mech. 66, 172– Limkatanyu, S., Ponbunyanon, P., Prachasaree, W., Kuntiyawichai, K.,
180. Kwon, M., 2014. Correlation between beam on winkler-pasternak
Bruce, D.M., 2003. Mathematical modelling of the cellular machanics of foundation and beam on elastic substrate medium with inclusion of
plants. Phil. Trans. R. Soc. Lond. B 358, 1437–1444. microstructure and surface effects. J. Mech. Sci. Technol. 28, 3653–
Burlayenko, V.N., Sadowski, T., 2010. Effective elastic properties of foam- 3665.
filled honeycomb cores of sandwich panels. Compos. Struct. 92, 2890– Murray, G., Gandhi, F., Hayden, E., 2009. Polymer filled honeycombs to
2900. achieve a structural material with appreciable damping. Am. Instit.
Caillerie, D., Mourad, A., Raoult, A., 2006. Discrete homogenization in Aeronaut. Astronaut..
graphene sheet modeling. J. Elast. 84, 33–68. Neethirajan, S., Gordon, R., Wang, L., 2009. Potential of silica bodies (phy-
Chen, J.Y., Huang, Y., Ortiz, M., 1998. Fracture analysis of cellular materi- toliths) for nanotechnology. Trends Biotechnol. 27 (8), 461–467.
als: a strain gradient model. J. Mech. Phys. Solids 46 (5), 789–828. Niklas, K.J., 1989. Mechanical behavior of plant tissues as inferred from
Chen, Q., Pugno, N.M., 2011. A parametrical analysis on the elastic the theory of pressurized cellular solids. Am. J. Bot. 76 (6), 929–937.
anisotropy of woven hierarchical tissues. Adv. Biomater. 13 (10), 337– Niklas, K.J., 1992. Plant Biomechanics: An Engineering Approach to Plant
394. Form and Function. University of Chicago Press.
Chen, Q., Pugno, N.M., 2012. In-plane elastic buckling of hierarchical hon- Ostoja-Starzewski, M., 2002. Lattice models in micromechanics. Appl.
eycomb materials. Eur. J. Mech. A/Solids 34, 120–129. Mech. Rev. 55 (6), 35–60.
Civalek, O., 2007. Nonlinear analysis of thin rectangular plates on Pugno, N.M., 2006. Mimicking nacre with super-nanotubes for producing
Winkler–Pasternak elastic foundations by DSC–HDQ methods. Appl. optimized super-composites. Nanotechnology 17, 5480–5484.
Math. Model. 31, 606–624. Sanchis Gritsch, C., Murphy, R.J., 2005. Ultrastructure of fibre and
Davini, C., Ongaro, F., 2011. A homogenized model for honeycomb cellular parenchyma cell walls during early stages of culm development in
materials. J. Elast. 104, 205–226. dendrocalamus asper. Ann. Bot. 95, 619–629.
Dawson, M.A., Gibson, L.J., 2007. Optimization of cylindrical shells with Van Liedekerke, P., Ghysels, P., Tijskens, E., Samaey, G., Smeedts, B.,
compliant cores. Int. J. Solids Struct. 44, 1145–1160. Roose, D., Ramon, H., 2010. A particle-based model to simulate the
D’Mello, R.J., Waas, A.M., 2013. Inplane crush response and energy absorp- micromechanics of single-plant parenchyma cells and aggregates.
tion of circular cell honeycomb filled with elastomer. Compos. Struct. Phys. Biol. 7.
106, 491–501. Wang, X.L., Stronge, W.J., 1999. Micropolar theory of two-dimensional
Dos Reis, F., Ganghoffer, J.F., 2012. Construction of micropolar continua stresses in elastic honeycomb. Proc. R. Soc. Lond. A 455, 2091–2116.
from the asymptotic homogenization of beam lattices. Comput. Struct. Warner, M., Thiel, B.L., Donald, A.M., 20 0 0. The elasticity and failure of
112-113, 354–363. fluid-filled cellular solids: theory and experiment,. PNAS 97 (4), 1370–
Eremeyev, V.A., Lebedev, L.P., Altenbach, H., 2013. Foundations of Microp- 1375.
olar Mechanics. Springer. Warren, W.E., Byskov, E., 2002. Three-fold symmetry restrictions on two-
Fratzl, P., Weinkamer, R., 2007. Natures hierarchical materials,. Prog. Mater. dimensional micropolar materials,. Eur. J. Mech. A/Solids 21, 779–792.
Sci. 52, 1263–1334. Wu, N., Pitts, M.J., 1999. Development and validation of a finite element
Fratzl, P., Weinkamer, R., 2011. Mechanical adaptation of biological materi- model of an apple fruit cell. Postharvest Biol. Tec. 16, 1–8.
als - the examples of bone and wood. Mat. Sci. Eng.: C 31, 1164–1173. Zhu, H.X., Melrose, J.R., 2003. A mechanics model for the compression of
Georget, D.M.R., Smith, A.C., Waldron, K.W., 2003. Modelling of carrot tis- plant and vegetative tissues,. J. theor. Biol. 221, 89–101.
sue as a fluid-filled foam. J. Mater. Sci. 38, 1933–1938.
Gibson, L.J., 1989. Modelling the mechanical behavior of cellular materials.
Mat. Sci. Eng. A110, 1–36.

You might also like