Вы находитесь на странице: 1из 23

ARTICLE IN PRESS

Journal of Biomechanics 38 (2005) 377–399


www.elsevier.com/locate/jbiomech
www.JBiomech.com

Biomechanics of cellular solids


Lorna J. Gibson
Department of Materials Science and Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Room 8-135,
Cambridge, MA 02139, USA
Accepted 29 September 2004

Abstract

Materials with a cellular structure are widespread in nature and include wood, cork, plant parenchyma and trabecular bone.
Natural cellular materials are often mechanically efficient: the honeycomb-like microstructure of wood, for instance, gives it an
exceptionally high performance index for resisting bending and buckling. Here we review the mechanics of a wide range of natural
cellular materials and examine their role in lightweight natural sandwich structures (e.g. iris leaves) and natural tubular structures
(e.g. plant stems or animal quills). We also describe two examples of engineered biomaterials with a cellular structure, designed to
replace or regenerate tissue in the body.
r 2004 Elsevier Ltd. All rights reserved.

Keywords: Cellular solids; Wood; Trabecular bone; Plant stems; Titanium foams; Scaffolds; Mechanical behavior

1. Introduction attachment, migration, proliferation and functioning of


cells (Fig. 2b).
In nature, materials with a cellular structure are The microstructural features of cellular solids affect-
widespread. Examples of natural materials with pris- ing their mechanical response are most easily observed
matic, honeycomb-like cells include wood and cork (Fig. in engineering honeycombs and foams (Fig. 3). Honey-
1a,b), while those with polyhedral cells include the inner combs, with their prismatic cells, are referred to as two-
core in plant stems and trabecular bone (Fig. 1c,d). They dimensional cellular solids while foams, with their
also appear as the cores in natural sandwich structures: polyhedral cells, are three-dimensional cellular solids.
in long, narrow plant leaves, such as the iris, and in The relative density is the density of the cellular solid
shell-like bones, such as the skull (Fig. 1e,f). Natural divided by that of the solid it is made from, and is
tubular structures often have a honeycomb-like or equivalent to the volume fraction of solid. Foams may
foam-like core supporting a denser outer cylindrical be either open (with solid only at the edges of the
shell, increasing the resistance of the shell to kinking or polyhedra) or closed (with solid membranes over the
local buckling failure (e.g. plant stems and animal quills, faces of the polyhedra). The properties of the foam
Fig. 1g,h). Engineers design biomaterials with a cellular depend on those of the solid making up the cellular
structure to replace or regenerate tissues in the body. material; in materials such as wood, the cell wall is itself
For example, titanium foams are being considered as a multilayered composite material.
substitute materials for trabecular bone (Fig. 2a). And The stress–strain curve for a cellular solid in
porous scaffolds used in tissue engineering are designed compression is characterized by three regimes (Fig.
to mimic the body’s extracellular matrix, allowing the 4a): a linear elastic regime, corresponding to cell edge
bending or face stretching; a stress plateau, correspond-
ing to progressive cell collapse by elastic buckling,
Tel.: 617 253 7107; fax: 617 258 6275. plastic yielding or brittle crushing, depending on the
E-mail address: ljgibson@mit.edu (L.J. Gibson). nature of the solid from which the material is made; and

0021-9290/$ - see front matter r 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jbiomech.2004.09.027
ARTICLE IN PRESS
378 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

Fig. 1. Examples of cellular solids in nature: (a) balsa wood (b) cork (c) inner core of plant stem in (g) (d) trabecular bone (e) iris leaf (f) skull (g)
plant stem (h) porcupine quill. (Fig. 1a,b,d,f reproduced with permission from Gibson and Ashby, 1997; Fig. 1h reproduced with permission from
Gibson et al., 1995.)
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 379

Fig. 2. Examples of biomaterials with a cellular structure (a) titanium


foam (courtesy Louis-Philippe Lefebvre, National Research Council of
Canada) (b) collagen-glycosaminoglycan scaffold for tissue engineer-
ing. (Reprinted from Progress in Materials Science 46, 273–282
Freyman TM, Yannas IV and Gibson LJ Cellular materials as
scaffolds for tissue engineering, copyright 2000, with permission from
Elsevier.)

densification, corresponding to collapse of the cells


throughout the material and subsequent loading of the
cell edges and faces against one another. Many cellular
materials have low relative densities (10–20%) so that
they can be deformed up to large strains (70–80%) Fig. 3. Examples of engineering cellular solids (a) aluminum
honeycomb (b) open-cell polyurethane foam (c) closed-cell polyethy-
before densification occurs. In tension, the linear elastic lene foam (Reproduced with permission from Gibson and Ashby,
response is the same as in compression, at least at small 1997).
strains (Fig. 4b). As the strain increases, the cells become
more oriented with the loading direction, increasing the
stiffness of the material until tensile failure occurs. hedron (a 14-faced polyhedron) in three dimensions
The mechanical response of cellular solids has been (Fig. 5) (Ko, 1965; Patel and Finnie, 1970; Gibson and
modeled by representing the cellular structure in several Ashby, 1982a; Warren and Kraynik, 1997; Zhu et al.,
ways. Initial models performed a structural analysis of a 1997a,b). The geometry of the unit cell makes the
unit cell such as a hexagon (in two dimensions) or a analysis tractable but may not give an exact representa-
dodecahedron (a 12-faced polyhedron) or tetrakaideca- tion of the real material (e.g. foams). An even simpler
ARTICLE IN PRESS
380 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

Fig. 4. Schematic uniaxial stress–strain curves for (a) elastomeric foam in compression (b) elastic–plastic foam in compression (c) elastomeric foam
in tension (d) elastic–plastic foam in tension. (Reproduced with permission from Gibson and Ashby, 1997.)

approach is to use dimensional analysis to model the then show how the models can be applied to honey-
mechanisms of deformation and failure observed in the comb- and foam-like natural cellular solids (wood and
cellular material without specifying the exact cell trabecular bone). The role that cellular materials play in
geometry (Gibson and Ashby, 1982b, 1997). This increasing the efficiency of natural sandwich structures
approach assumes that the cell geometry is similar in (e.g. iris leaves) and tubular structures (e.g. plant stems,
foams of different relative densities. It gives the animal quills) is also described. Cellular materials are
dependence of the properties on the relative density increasingly used as biomaterials to replace or regener-
and the solid properties, but requires experiments to ate tissue in the body. Here we describe titanium foams
determine the constants related to the cell geometry. A that are being considered for bone replacement and
third approach is to use finite element analysis of either porous scaffolds for the regeneration of a wide variety of
regular or random cellular structures (Silva et al., 1995, tissues, including skin, nerve, liver, bone and cartilage.
van der Burg et al., 1997). Finite element analysis allows
local effects, such as imperfections, to be studied (Silva
and Gibson, 1997a; Chen et al., 2001). It can also be 2. Micromechanical models
used in conjunction with imaging techniques such as
micro-computed tomography to model the exact geo- 2.1. Two-dimensional cellular solids (honeycombs), in-
metry of a particular sample (e.g. trabecular bone), plane loading
although this is computationally intensive (van Reitber-
gen, 1995). When loaded uniaxially in the plane of the hexagonal
In this review, we first summarize the results of cells, the cell walls of a honeycomb initially deform by
micromechanical models developed using dimensional bending (so long as the wall thickness, t, is small
analysis and compare them to selected results from the compared with the wall length, l) (Fig. 6a,b). The
analysis of unit cells and finite element modeling. We Young’s modulus, E* can be related to the relative
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 381

Fig. 6. Hexagonal cell (a) undeformed (b) linear elastic bending (c)
non-linear elastic buckling. (Reproduced with permission from Gibson
and Ashby, 1997.)

where Es is the Young’s modulus of the solid cell wall


material and I is the moment of inertia of the wall cross
section (I ¼ bt3 =12). The strain in the x1 direction is then:
d sin y s1 ðh þ l sin yÞbl 2 sin2 y
1 ¼ ¼ : (3)
l cos y 12E s I cos y
The Young’s modulus parallel to x1 is then:
E 1 t3 cos y
¼ : (4)
Fig. 5. Unit cells (a) hexagon (b) pentagonal dodecahedron (c) Es l ðh þ l sin yÞ sin2 y
tetrakaidecahedron. (Reproduced with permission from Gibson and
The Young’s modulus for loading in the x2 direction and
Ashby, 1997.)
the shear modulus for loading in the x1–x2 plane, can be
found in a similar way (Gibson and Ashby, 1997); they
density, r =rs ; the modulus of the solid, Es, and the cell are listed in Table 1. All depend on the Young’s modulus
geometry (h/l, y) using structural mechanics. A stress s1 of the solid from which the honeycomb is made, the cube
acting in the x1 direction induces a load P on the end of of the relative density (since r =rs is proportional to t/l
the inclined cell wall: (Table 1)) and a factor related to the cell geometry
(through h/l and y). The Poisson’s ratios are found by
P ¼ s1 ðh þ l sin yÞb; (1)
taking the negative ratio of the strains normal to and
where b is the wall thickness. The wall deflects by: parallel to the loading direction (Table 1). To a first
approximation, ignoring shear and axial deformations of
Pl 3 sin y the cell wall, we find that they depend only on the cell
d¼ ; (2)
12E s I geometry.
ARTICLE IN PRESS
382 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

Table 1 compress axially, so that the Young’s modulus simply


Properties of Two-Dimensional Cellular Solids
varies with the volume fraction of solid, or with the
In-plane properties Out-of-plane properties relative density:
r h=lþ2 E 3 r
rs ¼ tl 2 cos yðh=lþsin yÞ
¼ : (7)
E s rs
Elastic moduli
E 1 
t 3 E 3 
If buckling is avoided, the yield strength and brittle
Es ¼ l
cos y
ðhþl sin yÞ sin2 y Es ¼ rr
s
E 2  3
t h=lþsin y
crushing strength for loading in the x3 direction also
Es ¼ l cos3 y vary directly with the relative density. The stresses
n12 ¼  2 ¼ cos2 y n31 ¼ n32 ¼ ns
1 ðh=lþsin yÞ sin y corresponding to elastic and plastic buckling are given in
n21 ¼  1 ¼
ðh=lþsin yÞ sin y n13  n23  0 Table 1 (Gibson and Ashby, 1997).
2 cos2 y
G 12 
t 3 h=lþsin y G13 G t

¼ ¼ G23s ¼ 0:577 (regular
Es l
ðh=l Þ2 ð1þ2h=l Þ cos y Gs l
2.3. Three-dimensional cellular materials (foams)
hexagons)
Collapse stresses Dimensional arguments, based on the mechanisms of

s2el 2 2 t3 sel 
Es ¼ n 24p 1
lh2 cos y
3
 t 3 2 l=hþ2
deformation and failure in foams, can be used to
Es l 1n2s ðh=lþsin yÞ cos y
elastic buckling estimate their mechanical properties. The dimensional
   
spl spl
arguments assume that the cells in foams of different

t 2 1
5=3
sys
1
¼ l 2ðh=lþsin yÞ sin y sys
3
¼ 5:6 tl plastic relative densities are geometrically similar. While here
 
buckling; (regular hexagons) we use a cubic cell to illustrate the argument, the same
spl  result is obtained for any cell geometry, so long as the
2 t 2 1
sys ¼ l 2 cos2 y mode of deformation or failure is the same.
ðscr Þ1 
sfs ¼ t 2
l
1 In the linear elastic regime, under uniaxial stress,
3ðh=lþsin yÞ sin y
ðscr Þ2 
t 2 1
open-cell foams deform primarily by bending of the cell
¼
sfs l 3 cos2 y edges. The Young’s modulus, E* can be estimated as
follows (Fig. 7a,b). Under a transverse load, F, the
bending deflection, d; of a strut of length, l, and cross-
At a sufficiently high load, cells collapse, by elastic sectional area proportional to t2, is given by:
buckling, plastic yielding or brittle crushing, depending
Fl 3
on the properties of the cell wall material. The elastic d/ ;
buckling collapse stress is related to the Euler buckling E s t4
load, Pcrit, of the vertical column (Fig. 6c): where Es is the Young’s modulus of the solid. The stress
Pcrit acting on the cell is proportional to F/l2 and the strain is
sel ¼ ; (5) proportional to d/l, giving:
2lb cos y
2 2
where Pcrit ¼ n ph2E s I and n, is a factor describing the E  t4
/ : (8)
rotational stiffness of the node. We find: Es l

s2el n2 p2 t3 1 The relative density of any open-cell foam, r =rs ; is


¼ : (6) proportional to the square of the ratio of the strut
Es 24 lh2 cos y
thickness to length, t/l, so that
For h=l ¼ 1; n ¼ 0:686 and for h=l ¼ 2; n ¼ 0:806:   2
The plastic collapse stress can be calculated from the E r
¼ C1 : (9)
plastic moment to form plastic hinges in the cell walls Es rs
while the brittle crushing stress can be calculated from The analysis gives the dependence of the Young’s
the modulus of rupture required to fracture the walls modulus on the solid modulus and the relative density,
(Gibson and Ashby, 1997) (Table 1). Cell collapse and combines all of the constants of proportionality
progresses throughout the material at a roughly related to the cell geometry into a single constant, C1. By
constant load to large strains. Once all the cells have fitting Eq. (9) to data, we find that C 1  1: A structural
collapsed, and the cell walls press against one another, analysis of a tetrakaidecaheral unit cell model with cell
the stress rises sharply (densification). edges with Plateau borders finds that C 1 ¼ 0:98 (Warren
and Kraynik, 1997). Finite element analysis of random
2.2. Two-dimensional cellular materials (honeycombs), three-dimensional Voronoi foams with struts of circular
out-of-plane loading cross-section finds that C 1 ¼ 0:8 (Vajjhala et al., 2000).
All three approaches give similar results and all give the
For loading along the prism axis, in the out-of-plane same dependence of modulus on the square of the
direction, the cell walls of a honeycomb initially relative density.
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 383

Fig. 7. Dimensional analysis for an open-cell foam (a) undeformed cell (b) linear elastic strut bending. Cell collapse by (c) elastic buckling (d) plastic
yielding (e) brittle crushing. (Reproduced with permission from Gibson and Ashby, 1997.)

As for honeycombs, the Poisson’s ratio of foams 1/3, the shear modulus is
 
depends only on cell geometry and does not depend G  3 r 2
on relative density (assuming that bending deformations ¼ : (10)
E s 8 rs
dominate). Typical values lie around 1/3, although
there is a wide range in reported values. Re-entrant The elastic collapse stress is proportional to the Euler
foams, with negative Poisson’s ratios, have been buckling load divided by l2 (Fig. 7c):
reported with values of 0:8ono0 (see, for example, Pcrit E s t4
sel / 2 / 4 (11)
Lakes, 1987). l l
Like the Young’s modulus, the shear modulus of
or
open-cell foams is related to the bending stiffness of the   2
cell edges, and depends on the square of the relative r
sel ¼ C 2 : (12)
density. For an isotropic foam with a Poisson’s ratio of rs
ARTICLE IN PRESS
384 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

Fitting Eq. (12) to data gives C 2  0:05: Structural 3. Wood: a two-dimensional cellular solid
analysis of a tetrakaidecahedral unit cell gives C2 
0:120:18; depending on the cross section of the edges, The structure of two species of wood, cedar and oak,
for loading in one direction; loading normal to the are shown in Fig. 8. In both softwoods and hardwoods,
square faces does not produce buckling (Zhu et al., the bulk of the cells are long prismatic cells (tracheids in
1997b). softwoods and fibres in hardwoods). The annual growth
The plastic collapse stress, spl is found by equating the rings are distinguished by alternating bands of thin- and
applied moment, M, on a strut from a transverse force F thick-walled cells, corresponding to spring and summer,
to the plastic moment, Mp, required to form plastic respectively. The cells making up the rays are squatter
hinges (Fig. 7d): and more box-like. Hardwoods also have vessels,
relatively large diameter channels through which fluids
M / Fl / spl l 3 ; (13) are conducted.
Compressive stress–strain curves for loading several
species of wood across and along the grain, are shown in
M p / sys t3 (14) Fig. 9 (across the grain is the tangential direction while
along the grain is the axial direction in the figure). Both
giving stiffness and strength depend on the density of the
  3=2 wood. And wood is highly anisotropic: the stiffness and
spl t3 r strength along the grain are much higher than those
/ ¼ C3 ; (15)
sys l rs across the grain.
To a first approximation, the cellular structure of
where sys is the yield strength of the solid cell wall wood can be modeled as a honeycomb (Easterling et al.,
material. Fitting Eq. (15) to data gives C 3  0:3: 1982). Micrographs of specimens of balsa wood loaded
The brittle crushing strength is found in a similar in a small vice in a scanning electron microscope
manner, with (Fig. 7e): indicate that loading along the grain (analogous to the
  3=2 out-of-plane direction in the honeycomb) compresses
scr r the cell walls axially while loading across the grain
¼ C4 ; (16)
sfs rs (analogous to the in-plane direction in the honeycomb)
bends them: the mechanisms of deformation are the
where sfs is the modulus of rupture of the solid cell wall same as those of a honeycomb. The honeycomb model
material. Fitting Eq. (16) to data gives C 4  0:2: suggests that the Young’s modulus along the grain
In closed cell foams, stretching of the cell faces also varies linearly with relative density (Eq. 7) while that
contributes to the mechanical response. In many across the grain varies with the cube of the relative
polymer foams, the faces are thinner than the cell edges, density (Eq. 4). (The values of the cell wall density and
as polymer is drawn away from the faces and into the modulus rs and Es are roughly constant for different
edges by surface tension forces during processing. The species of wood.) The value of Es, the solid cell wall
contribution of the faces to the overall mechanical modulus, also depends on loading direction, as the cell
response depends, then, on the fraction of the solid in wall is a composite of cellulose fibres in a lignin/
the faces and adds a linear relative density term to the hemicellulose matrix. The axial and transverse cell wall
expressions already obtained for the Young’s modulus Young’s moduli have been measured to be 35 and
and the collapse stresses. The derivation of the proper- 10 GPa, respectively (Cave, 1968). Data for the Young’s
ties of closed-cell foams is given in Gibson and Ashby moduli of woods of widely varying density, along and
(1997). across the grain, are fairly well described by the simple
The fluid within the cells can also contribute to the honeycomb model (Fig. 10a). The modulus across the
mechanical response (Gibson and Ashby, 1997). In grain varies with density to a slightly lower power than
open-cell foams, there is the viscous dissipation of fluid 3, as predicted by the honeycomb model, probably due
moving through the cells. This effect depends on the to stiffening effect of the rays.
fluid viscosity, the strain rate, the cell size and the The compressive strength of wood is determined by
sample size. In closed-cell foams, as the volume of the uniaxial yielding for loading along the grain and by the
cell decreases during compression, the pressure of the formation of plastic hinges in bending for loading across
enclosed gas within the cells increases according to the the grain. The strength along the grain varies linearly
ideal gas law, contributing to the overall stress required with relative density while that across the grain varies
to deform the foam. with density squared (spl1 Table 1). The value of the
The modelling of honeycombs and foams are next strength of the cell wall depends on loading direction;
applied to two common natural cellular solids: wood the axial and transverse yield strengths have been
and trabecular bone. measured to be 350 and 135 MPa, respectively (Cave,
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 385

Fig. 8. Scanning electron micrographs of wood (a) cedar, cross section (b) cedar, longitudinal section, (c) oak, cross-section (d) oak, longitudinal
section. (Reproduced with permission from Gibson and Ashby, 1997.)

1969). The trends in the data for the compressive (Ashby et al., 1995):
strengths of woods of widely varying density, along and  1=2
across the grain, are fairly well described by the simple E 1=2 E 1=2
s rs
¼ : (17)
honeycomb model (Fig. 10b). r rs r
The different density dependencies of the modulus
The performance of wood is better than that of the
1=2solid
and strength for loading along and across the grain arise
from which it is made, by a factor of rs =r ; or
from the different mechanisms of deformation and
about 2 for typical softwoods with a relative density of
failure for the two loading directions. The honeycomb
0.3. Woods perform well, even compared to modern
model explains both the strong dependence of the
engineering composites: spruce, for instance has a value
properties on density and the pronounced anisotropy
of E=r1=2 of 7.1 GPa1/2 (Mg m3)1 while a unidirec-
in wood.
tional carbon fibre composite typically has a value of
The linear dependence of modulus and strength on
about 9 GPa1/2 (Mg m3)1. A similar argument can be
density for loading along the grain gives rise to the
made for the strength of a beam in bending. The
remarkable efficiency of wood in resisting bending
performance index to be maximized is then s2=3 =r; for
and buckling loads. For a beam of a given stiffness
wood, loaded along the grain,
or a column of a given elastic buckling load, the
material that minimizes the mass of the beam or column 2=3  
s2=3 sys rs 1=3
is that which maximizes the performance index E 1=2 =r ¼ : (18)
r rs r
(assuming that the length and cross-section of
the member are fixed) (Ashby, 1999). For wood, Again, the performance of wood is better than that of
loaded along the grain, rearranging Eq. (7) then gives the solid from which it is made. The performance of
ARTICLE IN PRESS
386 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

height, h (McMahon and Bonner, 1983). The Euler


buckling load is given by:
EI
Pcr / :
h2
The weight of the tree is:
W / rgd 2 h:
Equating the weight to the Euler buckling load gives:
Ed 2
¼ C:
rgh3
Rearranging gives:
 
Crg 1=2 3=2
d¼ h ; (19)
E
where E is the Young’s modulus of the wood, I is the
moment of inertia of the tree cross-section, r is the
density of the wood, g is the acceleration from gravity,
and C is a constant. We have already seen that for
loading along the grain, the ratio of E=r is a constant
(Eq. 7), so that tree diameter should increase as the 3/2
power of the height. In practice, this gives a good
description of the data for tree height and diameter
(Fig. 11) (McMahon and Bonner, 1983). The absolute
height of the tallest trees (California redwoods, Sequoia
sempervirens) is limited by constraints on water trans-
port (Koch et al., 2004).

4. Trabecular bone: a three-dimensional cellular solid

Trabecular bone exists at the ends of the long bones,


within the vertebral body, and in the core of shell-like
bones such as the skull. It has a cellular structure, with a
relative density typically between about 0.05 and 0.3.
Bone grows in response to load, so that the density of
trabecular bone depends on the magnitude of the loads
and the orientation of the trabeculae depends on the
Fig. 9. Compressive stress–strain curves for wood (a) loading in the
tangential direction, across the grain (b) loading in the axial direction, direction of the loading (Fig. 12). Low-density trabe-
along the grain. Note the different scales for compressive stress. cular bone resembles an open-cell foam. High-density
(Reproduced with permission from Gibson and Ashby, 1997.) trabecular bone has a more plate-like structure, with
perforations through the plates. The orientation of the
trabeculae can vary from almost equiaxed (e.g. in some
spruce s2=3 =r ¼ 92 MPa2=3 (Mg m3)1, for instance, sites in the proximal femur; Gibson and Ashby, 1997) to
is somewhat better than that of a unidirectional a nearly rectangular grid (e.g. in the vertebral body;
carbon fiber reinforced composite s2=3 =r ¼ 75 MPa2=3 Jensen et al. 1990). Because of the variation in
(Mg m3)1. trabecular architecture at even one density, there is
In trees, it is the bending and buckling performance of more variation in the stiffness and strength of trabecular
wood that is critical. Bending loads are imposed on the bone than in many other types of cellular materials.
branches by their own weight and the wind and on the The Young’s modulus of trabecular bone is plotted
trunk by the wind. The trunk is also loaded in axial against relative density in Fig. 13a. The data lie roughly
compression (supporting the weight of the entire tree on a line of slope 2, but with substantial scatter in the
above any given height). Interestingly, dimensional data. We have already seen that for open-cell foams,
analysis of elastic buckling in trees indicates that the with bending-dominated linear elastic behavior, the
diameter, d, should increase with the 3/2 power of the modulus varies with the square of the density (Eq. 9).
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 387

Fig. 10. (a) The Young’s modulus of wood plotted against density. One pair of axes is normalized by the cell wall Young’s modulus in the axial
direction (35 GPa) and by the cell wall density (1500 kg/m3). The other pair of axes corresponds to the raw data. (Reproduced with permission from
Gibson and Ashby, 1997.) (b) The compressive strength of wood plotted against density. One pair of axes is normalized by the axial strength of the
cell wall (350 MPa) and by the cell wall density (1500 kg/m3). The other pair corresponds to the raw data. Data from: Goodman and Bodig (1970,
1971); Bodig and Goodman (1973); Bodig and Jayne (1982); United States Forest Products Laboratory (1974); Dinwoodie (1981) and Easterling et
al. (1982) (Reproduced with permission from Gibson and Ashby, 1997).

There is evidence both from micro-computed tomogra- there is some variation in the value of C1. A framework
phy imaging of deformed specimens and from detailed for incorporating the dependence of the elastic moduli
finite element analysis of representations of trabecular on architecture, or fabric, has been provided by Cowin
bone architecture that bending is the dominant mode of (1985) and has been applied to trabecular bone by
linear elastic deformation in trabecular bone (van various groups (Turner et al., 1990; van Rietbergen
Rietbergen et al., 1995; Muller et al., 1998; Nazarian et al., 1998). Models incorporating both density and
and Muller, 2004). This is not surprising for low density trabecular architecture account for roughly 70–95% of
trabecular bone resembling an open-cell foam. It is more the variance in the elastic constants.
surprising for higher density trabecular bone with a The compressive strength is plotted against relative
more plate-like structure. We infer that the perforations density in Fig. 13b. The data are again fairly well
prevent truly closed-cell plate behavior, so that the described by a line of slope 2. Cellular solids models
dominant mode of deformation is bending, even in these suggest that this corresponds to failure by elastic
higher density structures. We note that the name buckling (Eq. 12) rather than brittle crushing, which
trabecular bone, is then particularly appropriate, given depends on density raised to the 3/2 power (Eq. 16). This
that trabecula means ‘‘little beam’’ in Latin. is a somewhat surprising result, but there is evidence for
Variations in the trabecular architecture give rise to elastic buckling failure in trabecular bone. For instance,
the scatter in the data, even at one density. The recent time-lapsed micro-computed tomography ima-
dimensional arguments used to develop Eq. (9) assume ging of trabecular bone loaded in uniaxial compression
that the cell geometry in the cellular materials of indicates that the maximum local strains were found in
different densities is similar, so that the constant C1 in rod-like elements that were aligned with the compres-
Eq. (9), representing the dependence of the modulus on sion axis. Some of these elements bend, while others are
cell geometry, has a single value. For many engineering compressed and some of those in compression buckle
cellular materials, such as polymer foams made by a (Muller et al., 1998; Nazarian and Muller, 2004). We
blowing process, this is, to a first approximation, a good expect that buckled struts become inelastic at small
assumption. The architecture of trabecular bone varies strains, as microdamage, in the form of microcracks
substantially from one site to another, however, so that within trabeculae, initiates at global strains of less than
ARTICLE IN PRESS
388 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

bone mass is lost through thinning of the trabeculae,


and then, at some critical trabecular thickness, pre-
sumably related to the osteoblast or osteoclast cell size,
through resorption of trabeculae (Fig. 14). Intuitively,
one might expect that the residual stiffness and strength
are more adversely affected by resorption than by
uniform thinning (consider, for instance, the mechanical
consequences of removing a column from a steel
framework in a skyscraper compared with removing
the same mass of material uniformly from all the
members of the framework).
The relative effects of density reduction from uniform
strut thinning versus random strut removal on the
modulus and compressive strength of cellular solids
have been studied using finite element analysis (Silva
and Gibson, 1997a; Vajjhala et al., 2000; Guo and Kim,
1999, 2002). The initial study analyzed regular, hexago-
nal honeycombs as well as random Voronoi honey-
combs (Fig. 15a) while the later studies applied the same
techniques to regular tetrakaidecahedral cells and
random three-dimensional Voronoi cells. For all the
structures studied, random removal of struts led to
greater reductions in stiffness and strength than uniform
thinning of struts (for a given density reduction) (Fig.
15b). For a given structure, the residual stiffness and
Fig. 11. Diameter at breast height plotted against overall height for strength were similar for a given density reduction by
576 record trees from the Social Register of Big Trees. The relationship one of the two mechanisms. The three-dimensional
d / h3=2 is indicated by the dashed line. (After McMahon and Bonner, structures were less sensitive to density reductions than
1983, with permission.) the two-dimensional structures. For instance, a 10%
reduction in density by strut removal decreased the
modulus of a three-dimensional structure by roughly
1% (Moore and Gibson, 2002). Interestingly, micro- 40% while the same density loss reduced that of a two-
damage was often observed to occur in trabeculae dimensional structure by roughly 70% (Fig. 15c).
oriented parallel to the compressive loading direction, Similar models for the relative effect of bone density
consistent with a buckling failure mode. loss through trabecular thinning and resorption have
The similarity in the density dependencies of the also been applied to vertebral bone, accounting for the
Young’s modulus and compressive strength suggest that more aligned trabecular architecture (Silva and Gibson,
the failure strain is a constant. In particular, for 1997b; Espinoza Ortiz et al., 2002; Rajapakse et al.,
specimens of trabecular bone of similar trabecular 2004). The results for the two-dimensional model are
architecture, the values of the constants relating to cell similar to those for regular and random cellular solids
geometry in the equations for modulus (Eq. 9) and (e.g. 10% reduction in the number of trabeculae leads to
compressive strength (Eq. 12) will be fixed, so that the about a 70% reduction in modulus and strength). The
failure strain in compression will have a constant value. results for the three-dimensional model are slightly less
The strain at failure in trabecular bone has been found sensitive than those for the regular or Voronoi cellular
to be independent of density (Kopperdahl and Keaveny, solids (e.g. 10% reduction in the number of trabeculae
1998; Keaveny et al., 1994), and to be isotropic for leads to about a 30% loss in the perturbed cubic bone
particular specimen (Chang et al., 1999). Compressive model while the same reduction in strut number leads to
yield strains (0.2% offset) are higher than those in about a 40% loss in modulus in the three-dimensional
tension (Kopperdahl and Keaveny, 1998; Keaveny et al., regular or Voronoi models).
1994), and can vary between sites (Kopperdahl and To summarize, in normal bone, the linear elastic
Keaveny, 1998; Morgan and Keaveny, 2001). Measured deformation is bending dominated and the Young’s
values of the compressive yield strain in trabecular bone modulus varies as the density squared. The variation in
are 0.84% for human vertebra and 1.09% for bovine trabecular architecture leads to large scatter in the data,
tibia. even at one density. Differences in trabecular architec-
Up until this point, the discussion has focused on ture can be accounted for by the inclusion of fabric in
normal trabecular bone. In patients with osteoporosis, models. There is evidence that the compressive strength
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 389

Fig. 12. Scanning electron micrographs showing the cellular structure of trabecular bone. (a) Specimen taken from the femoral head, showing low-
density, open-cell, rod-like structure. (b) Specimen taken from the femoral head, showing a higher density, perforated plate-like structure. (c)
Specimen taken from the femoral condyle, of intermediate density, showing an oriented structure, with rods normal to parallel plates. (Fig. 12a,b,c
reprinted from Gibson, 1985 with permission of Elsevier.)

of trabecular bone is related to buckling of trabeculae in nature, in, for instance, the skull and the leaves of
leading to a dependence on density squared. The strain plants such as the iris and cattail (Fig. 16). The skull has
to failure in bone is a constant for a given site with two outer faces of dense compact bone separated by a
consistent trabecular architecture. In osteoporotic bone, core of trabecular bone. Iris and cattail leaves have
bone loss occurs through trabecular thinning as well as outer layers that are like a fiber reinforced composite,
resorption. The modulus and strength are more severely with ribs of dense sclerenchyma running along their
reduced by resorption than by uniform thinning of length, separated by a core of either thin-walled
trabeculae. parenchyma cells (in the iris) or a rib-like structure
(in the cattail). The sandwich structure allows the iris
and cattail leaves to have a large surface area for
5. Natural sandwich structures: plant leaves, skull photosynthesis while providing sufficient structural
rigidity and strength at relatively low mass to maintain
Engineers sometimes make use of cellular solids as the their more or less upright position in the living plant.
cores of sandwich structures: components with two stiff, Mechanically, the sandwich structure acts like an
strong outer faces separated by a lightweight honey- I-beam, with the faces corresponding to the flanges of
comb or foam core. The separation of the faces by the the I-beam and the core corresponding to the web. The
core increases the moment of inertia of the section with equivalent moment of inertia of a sandwich beam is
little increase in weight, making sandwich structures (Allen, 1969):
attractive for resisting bending and buckling. Examples
of engineering sandwich panels include aircraft flooring E f bt3 E c bc3 E f btðc þ tÞ2
ðEI Þeq ¼ þ þ ; (20)
panels and downhill skis. Sandwich structures also occur 6 12 2
ARTICLE IN PRESS
390 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

Fig. 13. (a) The Young’s modulus of trabecular bone plotted against density. One pair of axes is normalized by the strut Young’s modulus (12 GPa)
and by the strut density (2000 kg/m3). The other pair of axes corresponds to the raw data. Data from (a) Carter and Hayes, 1977 (b) Carter et al.,
1980 (c) Bensusan et al., 1983 (d) Hvid et al., 1989 (e) Linde et al., 1991 H=human B=bovine. (Reproduced with permission from Gibson and
Ashby, 1997.) (b) The compressive strength of trabecular bone plotted against density. One pair of axes is normalized by the strut strength (136 MPa)
and by the strut density (2000 kg/m3). The other pair corresponds to the raw data. (Reproduced with permission from Gibson and Ashby, 1997.) J,
Behrens et al. (1974) Femur & Tibia; +, Carter and Hayes (1977) Bovine femoral condyle; ’, Carter and Hayes (1977) Tibial plateau; , Galante et
al. (1970) Vertebrae; &, Hayes and Carter (1976) Bovine femoral condyle; K, Weaver and Chalmers (1966) Vertebrae; W, Weaver and Chalmers
(1966) Calcaneus; m, Carter et al. (1980) Femur; E, Bensusan et al. (1983) B. Femur.

where Ef and Ec are the Young’s moduli of the face and The faces carry most of the normal stress while the
the core, t and c are the thicknesses of the face and the core carries most of the shear stress (Allen, 1969).
core, and b is the width of the beam. In practice, for Again, assuming that E f bE c and that t5c; then
efficient sandwich beams, the face material is much
MyE f M
stiffer than that of the core and the faces are thin relative sf ¼ ¼ (24)
ðEI Þeq btc
to the core, so that the third term is much larger than the
first two terms and and the shear stress in the core is
E f btc2 V
ðEI Þeq ¼ : (21) tc ¼ (25)
2 bc
The compliant core in sandwich beams gives rise to where M and V are the bending moment and shear force
significant shear deflections. The shear rigidity of the at a cross-section.
core is (for a beam with stiff, thin faces): Measurements of the geometry of the iris leaf (e.g. the
thickness and relative density of the face, the thickness
ðAG Þeq ¼ bcGc : (22) and relative density of the core, the diameter, spacing
and volume fraction of solid in the ribs) along with
The overall deflection of a sandwich beam, under a load
estimates of the solid tissue Young’s modulus for the
P, is then:
plant cell wall allow estimates of the bending stiffness of
2Pl 3 Pl the iris leaf to be made, using the sandwich beam
d¼ 2
þ ; (23) analysis (Eq. 23). The bending stiffness of iris leaves has
B1 E f btc B2 bcG c
also been measured by hanging small weights from
where B1 and B2 are constants related to the loading cantilevered leaves and measuring the corresponding
geometry (e.g. for a simply supported beam with a deflections. Agreement between the measured and
central load, B1 ¼ 48 and B2 ¼ 4). calculated bending stiffness is sufficiently good to
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 391

acts as an elastic foundation supporting the outer shell,


increasing its resistance to buckling.
In uniaxial compression, the presence of a compliant
core induces an axisymmetric buckling mode rather than
the diamond-pattern associated with a thin walled
cylindrical tube. The buckling wavelength is reduced
and buckling stress is increased, by an amount that
depends on the Young’s modulus of the core, Ec,
relative to that of the outer shell, Es, and the ratio of the
outer shell radius to thickness, a/t (Brush and Almroth,
1962; Siede, 1962; Yao, 1962; Weingarten and Wang,
1976; Karam and Gibson, 1995a). In bending, the
presence of the compliant core resists kinking failure
(e.g. as in a bent drinking straw). The increase in
moment capacity of a tubular beam with a compliant
core again depends on Ec/Es and a/t (Yabuta, 1980;
Karam and Gibson, 1995a). In both uniaxial compres-
sion and bending, stresses within the core decay fairly
rapidly in the radial direction, into the depth of the core:
they are reduced to less than 5% of the maximum value
at a radial distance of about 3 buckling wavelengths.
Material at a greater depth does not resist any
significant load. We note that plant stems are generally
hollow in the center, consistent with this finding. Animal
quills and bird feather rachis have a completely filled
core, possibly for thermal insulation, rather than
mechanical resistance.
The analysis of Karam and Gibson (1995a) has been
compared with experimental measurements of the
buckling load and bending moment capacity of silicone
rubber cylindrical tubes with foam cores (Karam and
Gibson, 1995b). The analysis gives a good description of
the dependence of the uniaxial compressive buckling
Fig. 14. Radiographs of sections through human lumbar vertebrae load on a/t, and the core depth, for the single value of
(L1) (a) normal bone (b) osteoporotic bone, showing thinning and Ec/Es tested (Fig. 18a). It also gives a good description
resorption of trabeculae. (Reproduced with permission from Vajjhala of the bending moment capacity as a function of a/t
et al., 2000, published by the American Society of Mechanical (again, for the single value of Ec/Es tested (Fig. 18b).
Engineers.)
The analysis confirmed that the critical uniaxial buck-
ling loads and bending moments of shells with foam
cores can be increased significantly above those of
conclude that the iris leaf behaves mechanically like a hollow cylinders of equivalent radius and mass; in the
sandwich beam (Gibson et al., 1988). experiments, the bending moment capacity was in-
creased by a factor of up to 4.8. It appears that the
cellular cores of natural tubular structures increase their
6. Natural tubular structures: plant stems, animal quills buckling resistance in axial compression and bending.

Tubular structures in nature often have a dense outer


cylindrical shell with an inner honeycomb- or foam-like
core. Examples include plant stems, animal quills and 7. Cellular solids as biomaterials
bird feather quills (known as the rachis) (Fig. 17). In
plant stems, this is known as the ‘‘core-rind’’ structure Cellular solids are increasingly used as biomaterials to
(Niklas, 1992). Quills from porcupines, hedgehogs, replace or regenerate damaged or diseased tissue. Here
echidnas and spiny rats all have a similar structure we consider two examples: open-cell titanium foams for
(Vincent and Owers, 1986; Karam and Gibson, 1994). In bone implants and porous scaffolds for tissue engineer-
nature, these structures are loaded primarily in bending ing, for the regeneration of a wide range of tissues (e.g.
and/or compression. In all of these instances, the core skin, nerve, bone, cartilage).
ARTICLE IN PRESS
392 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

Fig. 15. (a) Two-dimensional random Voronoi honeycomb (Reprinted from International Journal of Mechanical Sciences 37 Silva et al. The effects
of non-periodic microstructure on the elastic properties of two-dimensional cellular solids. 1161–1177, copyright 1995 with permission from Elsevier).
(b) Young’s modulus and compressive strength of 2D Voronoi honeycombs plotted against reduction in relative density by either uniform thinning or
random removal of struts. The modulus and strength are normalized by the values for the intact honeycomb. (Reproduced with permission from
Vajjhala et al., 2000, published by the American Society of Mechanical Engineers.) (c) Comparison of normalized Young’s modulus plotted against
reduction in relative density by strut removal for a random, three-dimensional Voronoi structure, a regular tetrakaidecahedral structure, a random
two-dimensional Voronoi structure and a regular two-dimensional hexagonal structure. (Reproduced with permission from Vajjhala et al., 2000
published by the American Society of Mechanical Engineers.)

Recently, there have been a number of novel processes For bone implant applications, open-cell titanium
developed for making metallic foams (Ashby et al., foams are being developed (Fig. 19). Titanium has
2000). Open-cell foams are of the most interest as excellent biocompatibility and is already widely used in
biomaterials, as the interconnectivity of the pores allows orthopaedic implants. It is well known that porous
for tissue ingrowth. Open-cell metallic foams for use in titanium coatings on implants induces bone ingrowth
structural and thermal applications have been made (Head et al., 1995). Titanium foams offer the possibility
using an open-cell polymer foam precursor. The of matching the stiffness of the foam with that of bone,
polymer foam is coated with a ceramic slurry which is through control of the relative density. A number of
then dried. To avoid damage to the ceramic foam different processes have been developed.
replica, the voids are filled with a casting sand. The In one processing route, a polymer binder charged
ceramic is then sintered, burning off the original with metallic particles and a foaming agent is foamed by
polymer foam. Molten metal is then cast into the voids heating to a temperature sufficient to melt the polymer,
left by the polymer foam and the mold materials are causing it to flow around the metal particles, and to
removed, leaving a metallic foam replica of the original decompose the foaming agent, generating a gas that
polymer foam. Open-cell metal foams can also be made produces the foam. The temperature is then further
by depositing metal onto a polymer foam template by increased to decompose the polymer and sinter the metal
chemical vapor deposition (CVD), evaporation, or particles together (Gauthier et al., 2003). The relative
electrodeposition. density of these titanium foams is about 30% and the
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 393

Fig. 16. Natural sandwich structures (a) skull (b) iris leaf (c) cattail leaf. ((a) Reproduced with permission from Gibson and Ashby, 1997 (b)
Reproduced from Gibson et al. copyright 1988 with kind permission of Springer Science and Business Media.)

pore size is roughly 450–750 mm. The mechanical be made by Direct Material Deposition (DMD), a Solid
properties such as the Young’s modulus and compres- Free-Form Fabrication technique (Maddox et al., 2003).
sive strength are similar to those of trabecular bone The porous structure is fabricated by using a laser beam
(E  0:6–2.1 GPa, sc  10–25 MPa). In another proces- to melt titanium powder along a computer-controlled
sing route, titanium foams are made using a fugitive pathway.
phase: titanium and ammonium hydrogen carbonate There is also interest in the development of tantalum
powders are mixed, pressed and heat-treated to burn out foams for bone implants, although tantalum is not yet
the fugitive phase, leaving porous titanium (Wen et al., widely used as an orthopaedic implant material. In one
2002). The relative density of these foams ranges from process the tantalum foam is made by chemical vapor
0.2 to 0.65, the Young’s modulus ranges from 2.9 to infiltration (CVI) of tantalum onto a commercially
10.3 GPa and the compressive strength ranges from 25 available reticulated vitreous carbon foam (Zardiackas
to 478 MPa. Values at the lower densities are similar to et al., 2001). The tantalum foam is 99% tantalum and
those for trabecular bone while the values at the higher 1% vitreous carbon and replicates the carbon foam
densities exceed them. The foaming processes described architecture. The relative density is 15–25% and the
by both Gauthier et al. (2003) and Wen et al. (2002) give pore size is nominally 550 mm. The Young’s modulus
relatively random, foam-like porous materials. Porous and compressive strength are roughly 1.5 GPa and
titanium biomaterials with a more regular structure can 35 MPa, respectively.
ARTICLE IN PRESS
394 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

Fig. 17. Natural tubular structures: (a,b) grassy stem (c,d) porcupine quill (e,f) blue jay feather rachis. (c,d reproduced with permission from Gibson
et al., 1995.)

The goal of tissue engineering is to synthesize or solid phase must be biocompatible and promote cell
regenerate tissues and organs. Today, this is done by adhesion and growth. Over time, as the cells produce
providing a synthetic porous scaffold, or matrix, which their own natural extracellular matrix, the synthetic
mimics the body’s own extracellular matrix, onto which matrix should degrade into non-toxic components that
cells attach, multiply, migrate and function. Matrix can be eliminated from the body. Examples of materials
materials must satisfy a number of requirements. The used in synthetic matrices include: poly L lactic acid
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 395

Fig. 19. Titanium foam for bone replacement (image courtesy Louis-
Philippe Lefebvre, National Research Council of Canada).

sufficient mechanical integrity to resist handling during


implantation and in vivo loading.
Matrix materials can be made by a variety of routes:
freeze-drying (Chen et al., 1995; Yannas et al., 1989),
fiber bonding (Mikos et al., 1993; Lu and Mikos, 1996),
foaming (Mikos et al., 1994; James and Kohn, 1996),
salt leaching (Lhommeau et al., 1998) and solid free-
Fig. 18. (a) Uniaxial buckling stress of the fully filled cylinder of a
form fabrications methods such as three-dimensional
given a/t divided by the buckling stress of a hollow cylinder of the same
a/t plotted against Seide’s dimensionless core stiffening parameter, printing (Park et al., 1998) have all been used (Fig. 20).
3=2 
a=t E c =E s : (b) Ratio of measured local buckling moment of As a result of the requirements for the porous structure,
partially filled cylinders to theoretical predictions for empty cylinders many tissue engineering scaffolds resemble open-cell
of equal weight, plotted against a/t for the partially filled cylinders. foams. Their mechanical response, too, is similar to that
(Fig. 18a,b, reprinted from International Journal of Solids and
of an open-cell foam (Fig. 21).
Structures 32, 1285–1306 Karam GN and Gibson LJ Elastic buckling
of cylindrical shells with elastic cores II: Experiments Copyright (1995) The mechanical interaction between cells and the
with permission from Elsevier.) E Weingarten and Wang (1976), matrix is of importance in understanding wound
FEM predictions; W, Kachman (1959), experimental results; J, contraction that occurs during wound healing. It is
Goree and Nash (1962), experimental results; &, Malyutin et al. believed that contraction is associated with the forma-
(1980), experimental results; m, Malyutin et al. (1980), theoretical
tion of scar tissue and that contraction and scar
predictions; ——, Siede (1962), theoretical predictions; — –, Yao
(1962), theoretical predictions; .........., Karam and Gibson (1994), formation must be inhibited for normal tissue to
theoretical predictions; , Karam and Gibson (1994), experimental regenerate (Yannas, 2001). It has been shown, in vitro,
results. that the contractile force developed by fibroblasts on a
collagen-GAG scaffold is associated with elongation or
(PLLA), polyglycolic acid (PGA) and poly DL lactic co- spreading of the cells (Freyman et al., 2001). As the
glycolic acid (PLGA); all of these are currently used in fibroblasts spread, the focal adhesion sites are main-
resorbable sutures. Collagen-based matrices are also tained at the periphery of the cell. Tensile forces in the
available and are widely used for skin regeneration in actin network within the cell induce a compressive force
burn patients. Mineralized matrices are being developed in the matrix strut to which the cell is attached (via the
for bone tissue engineering. focal adhesion sites). As the cell spreads, and the
The porous structure must also be designed to satisfy distance between the adhesion sites increases, the Euler
several requirements. The porosity must be intercon- buckling load for the matrix strut decreases until
nected to allow cell migration, vascularization and eventually the strut buckles, causing contraction of the
diffusion of nutrients. High porosity is needed for cell matrix.
seeding and migration (typical porosities are greater
than 90%). The pore size must be within a critical range
(usually a few hundred mm): the lower bound is 8. Summary
controlled by the size of the cells while the upper bound
is related to the specific surface area through the The cellular structure of wood resembles that of an
availability of binding sites. And the material must have engineering honeycomb. The stiffness and strength of
ARTICLE IN PRESS
396 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

Fig. 20. Examples of porous scaffolds for tissue engineering made by: (a) freeze-drying (collagen-GAG, image (Reprinted from Progress in Materials
Science 46, 273–282 Freyman TM, Yannas IV and Gibson LJ Cellular materials as scaffolds for tissue engineering, copyright 2000, with permission
from Elsevier). (b) fibre bonding (PGA, image reprinted from Mikos et al., copyright 1993 with permission of John Wiley and Sons, Inc.) (c) foaming
(PLLA, image reprinted from Polymer 35, 1068–1077 Mikos AG, Thorsen AJ, Czerwonka LA, Bao Y, Winslow DN, Vacanti JP, Langer R
Preparation and characterization of poly (L-lactic acid) foams. copyright 1994, with permission of Elsevier) (d) salt leaching (tyrosine-derived
polycarbonate, SEM image of a porous scaffold prepared in the laboratory of Professor Kohn at Rutgers University, as described in Lhommeau et
al., 1998).

different species of woods depend on the density of the


species and the direction of loading: woods are much
stiffer and stronger when loaded along the grain than
across it. Models based on dimensional analysis of the
mechanisms of deformation in honeycomb models
(uniaxial compression of the cell wall for loading along
the grain and bending of the cell wall for loading across
the grain) give a good description of the dependence of
the stiffness and strength of wood on relative density
and direction of loading. In particular, the linear
dependence of the stiffness and strength of wood on
the relative density for loading along the grain gives rise
to wood’s remarkable mechanical efficiency: in bending
and buckling, wood is more efficient than the solid from
Fig. 21. Compressive stress–strain curve for a collagen-GAG scaffold. which it is made. In trees, it is the bending and buckling
(Image courtesy of Brendan Harley.) loads that are critical.
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 397

The architecture of trabecular bone is more complex, trabecular bone, plant leaves, plant stems, animal quills)
as bone grows in response to the loads applied to it and as well as novel biomaterials with highly porous, cellular
the magnitude and the direction of the loads vary from structures.
one site to another in the body. Both micro-computed
tomography imaging of compressed specimens of
trabecular bone and detailed finite element models Acknowledgements
suggest that the trabeculae deform to a large extent by
bending and, at sufficiently high loads, by buckling. I am grateful for a number of stimulating and
Cellular solid models for open-cell foams indicate that productive collaborations with colleagues over many
bending and buckling lead to a squared dependence of years: Prof. MF Ashby of Cambridge University
the Young’s modulus and compressive strength on Engineering Department (cellular solids, including
density. The trend in the data for the Young’s modulus wood); Prof. TA McMahon (deceased), formerly of
and compressive strength of trabecular bone indicate Harvard University Division of Applied Sciences
that both vary with density squared, although there is a (trabecular bone); Prof. WC Hayes, formerly of the
large scatter in the data, as a result of variations in the Orthopaedic Biomechanics Laboratory, Beth Israel
trabecular architecture, even at one density. In patients Deaconess Medical Center, Harvard Medical School
with osteoporosis, bone is lost through trabecular (trabecular bone); and Prof. IV Yannas of the Depart-
thinning and resorption. Finite element models show ment of Mechanical Engineering, Massachusetts Insti-
that resorption of trabeculae reduces stiffness and tute of Technology (tissue engineering scaffolds and cell
strength much more dramatically than uniform thinning mechanics). I would also like to acknowledge the
of trabeculae. outstanding graduate students that I have had the great
In nature, as in engineering, cellular solids are good fortune to work with at MIT. Those who
sometimes combined with dense materials to form contributed to the work described in this paper are:
efficient structural components. Both the skull and Dr. Toby Freyman of Boston Scientific, Prof. Ed Guo,
plant leaves such as the iris and cattail have a sandwich Department of Biomedical Engineering, Columbia
structure with two stiff strong faces separated by a University, Mr. Brendan Harley, of the Department of
lightweight core. In the skull, the faces are dense Mechanical Engineering, MIT, Dr. Gebran Karam of
compact bone while the core is trabecular bone. Plant Kredo, s.a.r.l., Dr. Tara Moore of Exponent, Inc., Prof.
leaves such as the iris and cattail have outer layers of Matt Silva, Department of Orthopaedic Surgery,
dense ribs separated by a low-density core of parench- Washington University, and Ms. Surekha Vajjhala, of
yma cells or inner ribs. The separation of the stiff faces Nanostream, Inc. Financial support for our work on
by a lightweight core increases the moment of inertia of cellular solids in nature and biomaterials has been
sandwich structures, with little increase in weight, provided by the National Science Foundation, the
making them efficient for resisting bending and buckling National Institute of Arthritis and Musculoskeletal
loads. Tubular structures in nature often have a dense and Skin Diseases and the Matoula S. Salapatas
outer cylindrical shell supported by an inner core of Professorship at MIT.
honeycomb- or foam-like cells. Examples include plant
stems, animal quills and bird feather rachis. The core
acts as an elastic foundation supporting the outer shell, References
increasing its resistance to buckling.
Cellular solids are increasingly being used as bioma- Allen, H.G., 1969. Analysis and Design of Structural Sandwich Panels.
terials to replace or regenerate damaged or diseased Pergamon Press, Oxford.
Ashby, M.F., 1999. Materials Selection in Mechanical Design, second
tissue. Two examples described here are open-cell
ed. Butterworth Heinemann, Stoneham, MA.
titanium foams for bone implants and porous tissue Ashby, M.F., Gibson, L.J., Wegst, U., Olive, R., 1995. The mechanical
engineering scaffolds, for the regeneration of a wide properties of natural materials. I Material property charts.
range of tissues (e.g. skin, nerve, bone, cartilage). In Proceedings of the Royal Society of London A450, 123–140.
both applications, the cellular structure is critical. In Ashby, M.F., Evans, A., Fleck, N.A., Gibson, L.J., Hutchinson, J.W.,
titanium foams, it allows bone ingrowth and control of Wadley, H.N.G., 2000. Metal Foams: A Design Guide. Butter-
worth Heinemann, Stoneham, MA.
the mechanical properties of the foam to match those of Behrens, J.C., Walker, P.S., Shoji, H., 1974. Variations in strength and
trabecular bone. In scaffolds for tissue engineering structure of cancellous bone at the knee. Journal of Biomechanics
applications the low relative density, open-cell, foam- 7, 201–207.
like structure allows cells to migrate throughout the Bensusan, J.S., Davy, D.T., Heiple, K.G., Verdin, P.J., 1983. 19th
scaffold. Annual Orthopaedic Research Society Meeting 8, pp. 132.
Bodig, J., Goodman, J.R., 1973. Prediction of elastic parameters of
Models for the mechanical behavior of cellular solids wood. Wood Science 5, 249–264.
provide a framework for understanding the mechanics Bodig, J., Jayne, B.A., 1982. Mechanics of Wood and Wood
of the natural structures described in this paper (wood, Composites. Van Nostrand Reinhold, New York.
ARTICLE IN PRESS
398 L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399

Brush, D.O., Almroth, B.O., 1962. Buckling of core-stabilized Guo, X.E., Kim, C.H., 1999. Effects of age-related bone loss: A 3D
cylinders under axisymmetric external loads. Journal of Aerospace microstructural simulation. Proceedings of the 1999 Bioengineering
Science 29, 1164–1170. Conference American Society of Mechanical Engineers Bioengi-
Carter, D.R., Hayes, W.C., 1977. The compressive behavior of bone as neering Division, vol 42. pp. 327–328.
a two-phase porous structure. Journal of Bone and Joint Surgery Guo, X.E., Kim, C.H., 2002. Mechanical consequence of trabecular
59A, 954–962. bone loss and its treatment: A three-dimensional model simulation.
Carter, D.R., Schwab, G.H., Spengler, D.M., 1980. Acta Orthopae- Bone 30, 404–411.
dica Scandinavica 51, 733. Hayes, W.C., Carter, D.R., 1976. Postyield behavior of subchondral
Cave, I.D., 1968. The anisotropic elasticity of the plant cell wall. Wood trabecular bone. Journal of Biomedical Materials Research
Science and Technology 2, 268–278. Symposium 7, 537–544.
Cave, I.D., 1969. The longitudinal Young’s modulus of Pinus radiata. Head, W.C., Bauk, D.J., Emerson, R.H., 1995. Titanium as the
Wood Science and Technology 3, 40–48. material of choice for cementless femoral components in total hip-
Chang, W.C.W., Christensen, T.M., Pinilla, T.P., Keaveny, T.M., arthroplasty. Clinical Orthopaedics and Related Research 311,
1999. Isotropy of uniaxial yield strains for bovine trabecular bone. 85–90.
Journal of Orthopaedic Research 17, 582–585. Hvid, I., Bentzen, S.M., Linde, F., Mosekilde, L., Pongsoipetch, B.,
Chen, C.S., Yannas, I.V., Spector, M., 1995. Pore strain behavior of 1989. X-ray quantitative computed tomography: the relations to
collagen-glycosaminoglycan analogues of extracellular matrix. physical properties of proximal tibial trabecular bone specimens.
Biomaterials 16, 777–783. Journal of Biomechanics 22, 837–844.
Chen, C., Lu, T.J., Fleck, N.A., 2001. Effect of inclusions and holes on James, K., Kohn, J., 1996. New biomaterials for tissue engineering.
the stiffness and strength of honeycombs. International Journal of Materials Research Society Bulletin 21 (11), 22–26.
Mechanical Sciences 43, 487–504. Jensen, K.S., Mosekilde, L., Mosekilde, L., 1990. A model of vertebral
Cowin, S.C., 1985. The relationship between the elasticity tensor and trabecular bone architecture and its mechanical properties. Bone
the fabric tensor. Mechanics and Materials 4, 137–147. 11, 417–423.
Dinwoodie, J.M., 1981. Timber: Its Nature and Behaviour. Van Kachman, D.R., 1959. Test report on buckling of propellant cylinders
Nostrand Reinhold, New York. under compressive loads. Space Technology Labs, Inc. GM
Easterling, K.E., Harrysson, R., Gibson, L.J., Ashby, M.F., 1982. On 59-7520.6.24, 30 November.
the mechanics of balsa and other woods. Proceedings of the Royal Karam, G.N., Gibson, L.J., 1994. Biomimicking of animal quills and
Society of London A383, 31–41. plant stems: natural cylindrical shells with foam cores. Materials
Espinoza Ortiz, J.S., Rajapakse, C.S., Gunaratne, G.H., 2002. Science and Engineering C2, 113–132.
Strength reduction in electrical and elastic networks. Physical Karam, G.N., Gibson, L.J., 1995a. Elastic buckling of cylindrical
Review B66 (14) (144203), 1–8. shells with elastic cores—I Analysis. International Journal of Solids
Freyman, T.M., Yannas, I.V., Pek, Y.S., Yokoo, R., Gibson, L.J., and Structures 32, 1259–1283.
2001. Micromechanics of fibroblast contraction of a collagen-GAG Karam, G.N., Gibson, L.J., 1995b. Elastic buckling of cylindrical
matrix. Experimental Cell Research 269, 140–153. shells with elastic cores—II Experiments. International Journal of
Galante, J., Rostoker, W., Ray, R.D., 1970. Physical properties of Solids and Structures 32, 1285–1306.
trabecular bone. Calcified Tissue Research 5, 236–246. Keaveny, T.M., Wachtel, E.F., Ford, C.M., Hayes, W.C., 1994.
Gauthier, M., Menini, R., Bureau, M.N., So, S.K.V., Dion, M-J., Differences between the tensile and compressive strengths of bovine
Lefebvre, L.P., 2003. Properties of novel titanium foams intended tibial trabecular bone depend on modulus. Journal of Biomecha-
for biomedical applications. American Society for Metals Materials nics 27, 1137–1146.
and Processes for Medical Devices Conference. Anaheim, CA, 8–10 Ko, W.L., 1965. Deformations of foamed elastomers. Journal of
September 2003. Cellular Plastics 1, 45–50.
Gibson, L.J., 1985. The mechanical behavior of cancellous bone. Koch, G.W., Sillett, S.C., Jennings, G.M., Davis, S.D., 2004. The
Journal of Biomechanics 18, 317–328. limits to tree height. Nature 428, 851–854.
Gibson, L.J., Ashby, M.F., 1982a. The mechanics of two-dimensional Kopperdahl, D.L., Keaveny, T.M., 1998. Yield strain behavior of
cellular materials. Proceedings of the Royal Society of London trabecular bone. Journal of Biomechanics 31, 601–608.
A382, 25–42. Lakes, R., 1987. Foam structures with a negative Poisson’s ratio.
Gibson, L.J., Ashby, M.F., 1982b. The mechanics of three-dimen- Science 235, 1038–1040.
sional cellular materials. Proceedings of the Royal Society of Lhommeau, C., Levene, H., Abramson, S., Kohn, J., 1998. Prepara-
London A382, 43–59. tion of highly interconnected porous tyrosine-derived polycarbo-
Gibson, L.J., Ashby, M.F., 1997. Cellular Solids: Structure and nate scaffolds. Tissue Engineering 4, 468.
Properties, second ed. Cambridge University Press, Cambridge. Linde, F., Norgaard, P., Hvid, I., Odgaard, A., Soballe, K., 1991.
Gibson, L.J., Ashby, M.F., Easterling, K.E., 1988. Structure and Mechanical properties of trabecular bone: dependency on strain
mechanics of the iris leaf. Journal of Materials Science 23, rate. Journal of Biomechanics 24, 803–809.
3041–3048. Lu, L., Mikos, A.G., 1996. The importance of new processing
Gibson, L.J., Ashby, M.F., Karam, G.N., Wegst, U., Shercliff, H.R., techniques in tissue engineering. Materials Research Society
1995. The mechanical properties of natural materials II: Micro- Bulletin 21 (11), 28–31.
structures for mechanical efficiency. Proceedings of the Royal Maddox, R.D., Carroll, J.W., Hollister, S.J., Mazumdar, J., Krebs-
Society of London A450, 141–162. bach, P.H., 2003. Gene/cell delivery from designed titanium
Goodman, J.R., Bodig, J., 1970. Orthotropic elastic properties of scaffolds made by free-form fabrication. Society for Biomaterials
wood. Journal of the Structures Division American Society of Civil 29th Annual Meeting Transactions. Reno, Nevada.
Engineers ST11, pp. 2301–2319. Malyutin, I.S., Pilipenko, P.B., Georgievskii, V.P., Smykov, V.I., 1980.
Goodman, J.R., Bodig, J., 1971. Orthotropic strength of wood in Experimental and theoretical study of the stability in axial
compression. Wood Science 4, 83–94. compression, of cylindrical shells reinforced with an elastic filler.
Goree, W.S., Nash, N.W., 1962. Elastic stability of circular cylindrical Prikladnaya Mekhanika 16, 56–60.
shells stabilized by a soft elastic core. Experimental Mechanics 2, McMahon, T.A., Bonner, J.T., 1983. On Size and Life. Scientific
142–149. American Library.
ARTICLE IN PRESS
L.J. Gibson / Journal of Biomechanics 38 (2005) 377–399 399

Mikos, A.G., Bao, Y., Cima, L.G., Ingber, D.E., Vacanti, J.P., van der Burg, M.W.D., Shulmeister, V., van der Geissen, E., Marissen,
Langer, R., 1993. Preparation of poly (glycolic acid) bonded fiber R., 1997. On the linear elastic properties of regular and random
structures for cell attachment and transplantation. Journal of open-cell foam models. Journal of Cellular Plastics 33, 31–54.
Biomedical Materials Research 27, 183–189. van Rietbergen, B., Weinans, H., Huiskes, R., Odgaard, A., 1995. A
Mikos, A.G., Thorsen, A.J., Czerwonka, L.A., Bao, Y., Winslow, new method to determine trabecular bone elastic properties and
D.N., Vacanti, J.P., Langer, R., 1994. Preparation and character- loading using micromechanical finite-element models. Journal of
ization of poly (L-lactic acid) foams. Polymer 35, 1068–1077. Biomechanics 28, 69–81.
Moore, T.L.A., Gibson, L.J., 2002. Microdamage accumulation in van Rietbergen, B., Odgaard, A., Kabel, J., Huiskes, R., 1998.
bovine trabecular bone in uniaxial compression. Journal of Relationships between bone morphology and bone elastic proper-
Biomechanical Engineering 124, 63–71. ties can be accurately quantified using high resolution computer
Morgan, E.F., Keaveny, T.M., 2001. Dependence of yield strain of reconstructions. Journal of Orthopaedics Research 16, 23–28.
human trabecular bone on anatomic site. Journal of Biomechanics Vincent, J.F.V., Owers, P., 1986. Mechanical design of hedgehog
34, 569–577. spines and porcupine quills. Journal of Zoology A210, 55–75.
Muller, R., Gerber, S.C., Hayes, W.C., 1998. Micro-compression: a Warren, W.E., Kraynik, A.M., 1997. Linear elastic behavior of a low-
novel technique for the nondestructive assessment of local bone density Kelvin foam with open cells. Journal of Applied Mechanics
failure. Technology and Health Care 6, 433–444. 64, 787–794.
Nazarian, A., Muller, R., 2004. Time-lapsed microstructural imaging Weaver, J.K., Chalmers, J., 1966. Cancellous bone: Its strength and
of bone failure behavior. Journal of Biomechanics 37, 55–65. changes with aging and an evaluation of some methods of
measuring its mineral content: I Age changes in cancellous bone.
Niklas, K.J., 1992. Plant Biomechanics: An Engineering Approach to
Journal of Bone and Joint Surgery 48A, 289.
Plant Form and Function. University of Chicago Press, Chicago.
Weingarten, V.I., Wang, Y.S., 1976. Stability of shells attached to
Park, A., Wu, B., Griffith, L.G., 1998. Integration and surface
elastic core. Journal of the Engineering Mechanics Division
modification and 3D fabrication techniques to prepare patterned
American Society of Civil Engineers 102, 839–870.
poly (L-lactide) substrates allowing regionally selective cell adhe-
Wen, C.E., Yamada, Y., Shimojima, K., Chino, Y., Hosokawa, H.,
sion. Journal of Biomaterials Science Polymer Edition 9, 89–110.
Mabuchi, M., 2002. Novel titanium foam for bone tissue
Patel, M.R., Finnie, I., 1970. Structural features and mechanical
engineering. Journal of Materials Research 17, 2633–2639.
properties of rigid cellular plastics. Journal of Materials 5, 909–932.
Yabuta, T., 1980. Effects of elastic supports on the buckling of circular
Rajapakse, C.S., Thomsen, J.S., Espinoza Ortiz, J.S., Wimalawansa,
cylindrical shells under bending. Journal of Applied Mechanics 47,
S.T., Ebbesen, E.N., Mosekilde, L., Gunaratne, G.H., 2004. An 866–870.
expression relating breaking stress and density of trabecular bone. Yannas, I.V., 2001. Tissue and Organ Regeneration in Adults.
Journal of Biomechanics 37, 1241–1249. Springer, Berlin.
Siede, P., 1962. The stability under uniaxial compression and lateral Yannas, I.V., Lee, E., Orgill, D.P., Skrabut, E.M., Murphy, G.F.,
pressure of circular-cylindrical shells with a soft elastic core. 1989. Synthesis and characterization of a model extracellular
Journal of Aerospace Science 29, 851–862. matrix that induces partial regeneration of adult mammalian skin.
Silva, M.J., Gibson, L.J., 1997a. The effects of non-periodic Proceedings of the National Academy of Science USA 86, 933–937.
microstructure and defects on the compressive strength of two- Yao, J.C., 1962. Buckling of axially compressed long cylindrical shell
dimensional cellular solids. International Journal of Mechanical with elastic core. Journal of Applied Mechanics 29, 329–334.
Sciences 39, 549–563. Zardiackas, L.D., Parsell, D.E., Dillon, L.D., Mitchell, D.W.,
Silva, M.J., Gibson, L.J., 1997b. Modeling the mechanical behavior of Nunnery, L.A., Poggie, R., 2001. Structure, metallurgy and
vertebral trabecular bone: effects of age-related changes in mechanical properties of a porous tantalum foam. Journal of
microstructure. Bone 21, 191–199. Biomedical Materials Research 58, 180–187.
Silva, M.J., Hayes, W.C., Gibson, L.J., 1995. The effects of non- Zhu, H.X., Knott, J.F., Mills, N.J., 1997a. Analysis of the elastic
periodic microstructure on the elastic properties of two-dimen- properties of open-cell foams with tetrakaidecahedral cells. Journal
sional cellular solids. International Journal of Mechanical Sciences of the Mechanics and Physics of Solids 45, 319–343.
37, 1161–1177. Zhu, H.X., Mills, N.J., Knott, J.F., 1997b. Analysis of the high strain
Turner, C.H., Cowin, S.C., Rho, J.Y., Ashman, R.B., Rice, J.B., 1990. compression of open-cell foams. Journal of the Mechanics and
The fabric dependence of the orthotropic elastic constants of Physics of Solids 45, 1875–11904.
cancellous bone. Journal of Biomechanics 23, 549–561.
United States Forest Products Laboratory 1974. Wood Handbook:
Wood as an Engineering Material. U.S. Department of Agricul- Further reading
ture, Agricultural Handbook 72.
Vajjhala, S., Kraynik, A.M., Gibson, L.J., 2000. A cellular solid model Freyman, T.M., Yannas, I.V., Gibson, L.J., 2000. Cellular materials as
for modulus reduction due to resorption of trabeculae in bone. porous scaffolds for tissue engineering. Progress in Materials
Journal of Biomechanical Engineering 122, 511–515. Science 46, 273–282.

Вам также может понравиться