You are on page 1of 21

Acta Materialia 51 (2003) 5801–5821

www.actamat-journals.com

Designing hybrid materials夽


M.F. Ashby a,∗, Y.J.M. Bréchet b
a
Engineering Department, University of Cambridge, Trumpington Street, CB2 1PZ Cambridge, UK
b
L.T.P.C.M., Domaine Universitaire de Grenoble, BP75, 38402 Saint Martin d’Heres Cedex, France

Accepted 31 August 2003

Abstract

The properties of engineering materials can be mapped, displaying the ranges of mechanical, thermal, electrical and
optical behavior they offer. These maps reveal that there are holes: some areas of property-space are occupied and
others are empty. The holes can sometimes be filled and the occupied areas extended by making hybrids of two or
more materials or of material and space. Particulate and fibrous composites are examples of one type of hybrid, but
there are many others: sandwich structures, foams, lattice structures and more. Here we explore ways of designing
hybrid materials, emphasizing the choice of components, their shape and their scale. The new variables expand the
design space, allowing the creation of new “materials” with specific property profiles.
 2003 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Designing materials; Hybrid materials; Composites; Sandwich structures; Foams

1. Introduction: hybrid materials and parts are not. Some parts are inaccessible for
fundamental reasons that relate to the size of atoms
1.1. Extending material-property space and the nature of the forces that bind their atoms
together. But other parts are empty even though,
Fig. 1 is an example of a material-property in principle, they are accessible. If they were
chart. It shows the thermal conductivities of some accessed, the new materials that are there could
2300 different materials, plotted against their allow novel design possibilities.
Young’s moduli. It is one of many, each a slice One approach to this—the traditional one—is
through material-property space; the assembly of that of developing new metal alloys, new polymer
all the slices can be thought of as a map of this chemistries and new compositions of glass and cer-
space [1,2]. All the charts have one thing in com- amics so as to extend the populated areas of the
mon: parts of them are populated with materials property charts, but this can be an expensive and
uncertain process. An alternative is to combine two
or more existing materials so as to allow a super-

Corresponding author. Tel.: +44-01223-332-635; fax: +44- position of their properties—in short, to create a
01223-332-662.

The Golden Jubilee Issue—Selected topics in Materials hybrid (Fig. 2). The spectacular success of carbon
Science and Engineering: Past, Present and Future, edited by and glass-fiber reinforced composites at one
S. Suresh. extreme, and of foamed materials at another

1359-6454/$30.00  2003 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/S1359-6454(03)00441-5
5802 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

Fig. 1. A material-property chart of thermal conductivity and Young’s modulus for 2300 materials. Each small circle is a plot of
these properties for a real material. The large ellipses enclose, approximately, the circles for a given family of materials. A large
area of the chart is empty: there are no materials with high conductivity and low modulus. The challenge is to create hybrids that
fill the hole. (This and the other charts were created using the CES 4 software system, Ref. [42].)

(hybrids of material and space) in filling previously


empty areas of the property charts is encourage-
ment enough to explore ways in which such
hybrids might be designed. What is the best way
to go about doing so?

1.2. What might we hope to achieve?

Fig. 3 shows schematically the fields occupied


by two families of materials, plotted on a chart
with properties P1 and P2 as axes. Within each field
a single member of that family is identified
(materials M1 and M2). What might be achieved
by making a hybrid of the two? The figure shows
four scenarios, each typical of a different class of
hybrid. We consider the case when large values
of P1 and P2 are desirable, low values not. Then,
depending on the shapes of the materials and the
way they are combined, we may find any one of
the following.
Fig. 2. Hybrid materials combine the properties of two (or
more) monolithic materials, or of one material and space. They
include fibrous and particulate composites, foams and lattices, 앫 “The best of both” scenario (Point A). The ideal,
sandwiches and almost all natural materials. One might imagine often, is the creation of a hybrid with the best
two further dimension: those of shape and scale. properties of both components. There are
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5803

exhaustive. Other combinations are possible, some


relying on the physics of percolation, others on
atomistic effects. These will emerge below.

1.3. When is a hybrid a “material”?

There is a certain duality about the way in which


hybrids are viewed and discussed. Some, like filled
polymers, composites or wood are treated as
materials in their own right, each characterized by
its own set of material properties. Others—like gal-
vanized steel—are seen as one material (steel) to
which a coating of a second (zinc) has been
applied, even though this could be regarded as a
new material with the strength of steel but the sur-
face properties of zinc (“stinc”, perhaps?). Sand-
Fig. 3. The possibilities of hybridization. The properties of the
wich panels illustrate the duality, sometimes
hybrid reflect those of its component materials, combined in
one of several possible ways. viewed as two sheets of face-material separated by
a core material, and sometimes—to allow compari-
son with bulk materials—as a “material” with their
examples, most commonly when a bulk pro- own density, flexural stiffness and strength. To call
perty of one material is combined with the sur- any one of these a “material” and characterize it
face properties of another. Zinc coated steel has as such is a useful shorthand, allowing designers
the strength and toughness of steel with the cor- to use existing methods when designing with them.
rosion resistance of zinc. Glazed pottery But if we are to design the hybrid itself, we must
exploits the formability and low cost of clay deconstruct it, and think of it as a combination of
with the impermeability and durability of glass. materials (or of material and space) in a defined
앫 “The rule of mixtures” scenario (Point B). When geometry.
bulk properties are combined in a hybrid, as in
structural composites, the best that can be
obtained is often the arithmetic average of the 2. The method: “A + B + shape + scale”
properties of the components, weighted by their
volume fractions. Thus unidirectional fiber com- First, a working definition: “a hybrid material is
posites have an axial modulus (the one parallel a combination of two or more materials in a prede-
to the fibers) that lies close to the rule of mix- termined geometry and scale, optimally serving a
tures. specific engineering purpose” [3], which we para-
앫 “The weaker link dominates” scenario (Point phrase as “A + B + shape + scale”. Here we allow
C). Sometimes we have to live with a lesser for the widest possible choice of A and B, includ-
compromise, typified by the stiffness of particu- ing the possibility that one of them is a gas or sim-
late composites, in which the hybrid properties ply space. These new variables expand the design
fall below those of a rule of mixtures, lying space, allowing an optimization of properties that
closer to the harmonic than the arithmetic mean is not possible if choice is limited to single, mono-
of the properties. Although the gains are less lithic materials.
spectacular, they can still be useful. The basic idea, illustrated in Fig. 4, is this.
앫 The “worst of both” scenario (point D)—not Monolithic materials offer a certain portfolio of
something we want. properties on which much engineering design is
based. But if the design requirements are excep-
These set certain fixed points, but the list is not tionally demanding, no single material may be
5804 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

Fig. 4. The steps in designing a hybrid to meet given design requirements.

found that can meet them: the requirements lie in met in this way, others need a more inventive
a hole in property space. Then the way forward is approach.1 So the question arises: are there general
to identify and separate the conflicting require- ways in which material hybridization can be
ments, seeking optimal material solutions for each, explored systematically? It is unrealistic to suppose
and then combine them in ways that retain the that one method and one tool can solve all such
desirable attributes of both. The best choice is the problems. Instead we examine examples of hybrid
one that ranks most highly when measured by the design and attempt to extract principles that could
performance metrics that motivate the design: min- help tackle other, as yet unformulated problems of
imizing mass or cost, or maximizing some aspect this class.
of performance (the criteria of excellence). The
alternative combinations are examined and
assessed, using the criteria of excellence to rank 3. A + B: selecting components for composites
them. The output is a specification of a hybrid in
terms of its component materials and geometry. Aircraft engineers, automobile makers, and
Consider as an example the design of a hybrid designers of sports equipment all have one thing in
material for long-span power cables. The objec- common: they want materials that are stiff, strong,
tives are to minimize the electrical resistance, but tough and light. The single-material choices that
at the same time to maximize the strength since best achieve this are the light alloys: alloys based
this allows a greater span. In multi-objective opti- on magnesium, aluminum and titanium. Much
mization, of which this problem is an example, it research aims at improving their properties. But
is conventional to express each objective such that they are not all that light—polymers have much
a minimum is sought; we thus seek materials with lower densities. Nor are they all that stiff—cer-
the lowest values of resistivity, R, and reciprocal amics are much stiffer and, especially in the form
of tensile strength, 1 / sts. Fig. 5 shows the result: of small particles or thin fibers, much stronger.
materials that best meet the design requirements lie These facts are exploited in the subset of hybrids
near the bottom left. Those with the lowest resist- that we usually refer to as particulate and
ance—copper, aluminum, and some of their fibrous composites.
alloys—are not very strong, and the materials that Any two materials can, in principle, be com-
are strongest—drawn carbon and low-alloy steel— bined to make a composite, and they can be mixed
do not conduct very well. Now consider a cable in many geometries (Fig. 6). In this section, we
made by interweaving strands of copper and steel
such that each occupies half the cross-section. 1
Assuming the steel carries no current and the cop- An interesting example is that of flexible ferromagnets.
Monolithic ferromagnetic materials are stiff, metallic or cer-
per no load (the most pessimistic scenario) the per- amic, solids. Elastomeric ferromagnetic hybrids offer several
formance of the cable will lie at the point shown properties that these monolithic solids do not. The hybrids are
on the figure—it has twice the resistivity of the made by mixing up to 30% of sub-micron iron particles into an
copper and half the strength of the steel. It occupies elastomer resin before polymerising it. The result is a compliant
a part of property space that was previously empty, ferromagnetic material that has the property that it is mag-
netostrictive, and that its stiffness increases when placed in the
offering performance that was not previously poss- magnetic field because the magnetic dipoles that are induced in
ible. the particles attract one another. The material has a fast (1 ms)
But while some conflicting requirements can be response time, making it suitable for vibration damping.
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5805

Fig. 5. Designing a hybrid—here, one with high strength and high electrical conductivity. The figure shows the resistivity and
reciprocal of tensile strength for 1700 metals and alloys. We seek materials with the lowest values of both. The construction is for
a hybrid of hard-drawn OFHC copper and drawn low alloy steel, but the figure itself allows many hybrids to be investigated [42].

restrict the discussion to fully dense, strongly


bonded, composites such that there is no tendency
for the components to separate at their interfaces
when the composite is loaded, and to those in
which the scale of the reinforcement is large com-
pared to that of the atom or molecule size and the
dislocation spacing, allowing the use of con-
tinuum methods.
On a macroscopic scale—one which is large
compared to that of the components—a composite
behaves like a homogeneous solid with its own set
of thermo-mechanical properties. Calculating these
precisely can be done, but it is difficult. It is much
easier to bracket them by bounds or limits: upper
and lower values between which the properties lie
[4–7]. The term “bound” will be used to describe
a rigorous boundary, one which the value of the
property cannot—subject to certain assumptions—
Fig. 6. Schematic of hybrids of the composite type: unidirec- exceed or fall below. It is not always possible to
tional fibrous, laminated fiber, chopped fiber and particulate derive bounds; then the best that can be done is to
composites. Bounds and limits, described in the text, bracket derive “limits” outside which it is unlikely that the
the properties of all of these. value of the property will lie. The important point
5806 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

is that the bounds or limits bracket the properties Table 1


of all arrangements of matrix and reinforcement Criteria of excellence for minimum weight design
shown in Fig. 6; by using them we avoid the need
Mode of loading and Stiffness at minimum weight
to model individual geometries. geometry

3.1. Density Tensile loading of ties E/r


Bending of beams E1 / 2 / r
When a volume fraction f of a reinforcement r Bending of plates E1 / 3 / r
(density rr) is mixed with a volume fraction (1⫺
f) of a matrix m (density rm) to form a composite
with no residual porosity, the composite density is stiffness per unit mass, measured by the indices
given exactly by a rule of mixtures (an arithmetic listed in the table (for derivations see Ref. [2]). If
mean, weighted by volume fraction) a possible hybrid has a value of any one of these
that exceed those of the light alloys, it achieves
r ⫽ frr ⫹ (1⫺f)rm. (1) our goal.
The geometry or shape of the reinforcement does Consider, as an illustration of the method, the
not matter except in determining the maximum design of a composite for a light, stiff beam of
packing-fraction of reinforcement and thus the fixed section-shape, to be loaded in bending. The
upper limit for f. efficiency is measured by the index E1 / 2 / r shown
in Table 1. Imagine, as an example, that the beam
3.2. Modulus is at present made of an aluminum alloy. Beryllium
is both lighter and stiffer than aluminum; ceramics
The modulus of a composite is bracketed by the are stiffer, but not all are lighter. What can these
well-known Voigt and Reuss bounds. The upper hybrids offer?
bound, Eu, is obtained by postulating that on load- Fig. 7 is a small part of the E–r property chart.
ing the two components suffer the same strain; the
stress is then the volume-average of the local
stresses and the composite modulus follows a rule
of mixtures:
Eu ⫽ f Er ⫹ (1⫺f)Em. (2)
Here Er is the Young’s modulus of the reinforce-
ment and Em that of the matrix. The lower bound,
El, is found by postulating instead that the two
components carry the same stress; the strain is the
volume-average of the local strains and the com-
posite modulus is
EmEr
E1 ⫽ (3)
f Em ⫹ (1⫺f)Er
More precise bounds are possible [8,9], but the
simple ones are adequate to illustrate the method.

3.3. Hybrid design for stiffness at minimum


weight
Fig. 7. Part of the E–r property chart, showing aluminium
alloys, beryllium and alumina (Al203). Bounds for the moduli
We need a criterion of excellence to assess the of hybrids made by mixing them are shown. The diagonal con-
merit of any given hybrid. Here our criterion is tours plot the criterion of excellence, E1 / 2 /r.
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5807

Three groups of materials are shown: aluminum optimum strength, and for tailored values of ther-
and its alloys, alumina (Al2O3) and beryllium (Be). mal conductivity, expansion coefficient and spe-
Composites made by mixing them have densities cific heat [7]. The properties of specific composites
given exactly by Eq. (1) and moduli that are brack- can, of course, be computed in conventional ways.
eted by the bounds of Eqs. (2) and (3). Both of The advantage of this graphical approach is the
these moduli depend on volume fraction of breadth and freedom of conceptual thinking that it
reinforcement, and through this, on density. Upper allows and the ease of comparison of possible new
and lower bounds for the modulus–density hybrids with the population of existing materials.
relationship can thus be plotted onto the E–r chart
using volume fraction f as a parameter, as shown 3.4. Percolation: properties that switch on and
in Fig. 7. Any composite made by combining off
aluminum with alumina will have a modulus con-
tained in the envelope for Al–Al2O3; the same for Fig. 9, a chart of electrical resistivity against
Al–Be. Fibrous reinforcement gives a longitudinal elastic stiffness (here measured by Young’s
modulus (parallel to the fibers) near the upper modulus), has an enormous hole. Materials that
bound; particulate reinforcement or transversely conduct well are stiff; those that are flexible are
loaded fibers give moduli near the lower one. insulators. Consider designing materials to fill the
Superimposed on Fig. 7 is a grid showing the hole; to be more specific, consider designing one
criterion of excellence E1 / 2 / r. The bound-envelope that has low modulus, can be molded like a poly-
for Al–beryllium composites extends almost nor- mer, and is a good electrical conductor. Such
mal to the grid, while that for Al–Al2O3 lies at a materials find application in anti-static clothing and
shallow angle to it. Beryllium fibers improve per- mats, as pressure sensing elements, even as solder-
formance (as measured by E1 / 2 / r) roughly four less connections.
times as much as alumina fibers do, for the same Metals, carbon and some carbides and intermet-
volume fraction. The difference for particulate allics are good conductors, but they are stiff and
reinforcement is even more dramatic. The lower cannot be molded. Thermoplastic and thermoset-
bound for Al–Be lies normal to the contours: 30% ting elastomers can be molded but do not conduct.
of particulate beryllium increases E1 / 2 / r by a fac- How are they to be combined? Metal coating of
tor of 1.5. The lower bound for Al–Al2O3 is, polymers is workable if the product is to be used in
initially, parallel to the E1 / 2 / r grid: 30% of par- a protected environment, but the coating is easily
ticulate Al2O3 gives almost no gain. The underly- damaged. If a robust, flexible, product is needed,
ing reason is clear: both beryllium and Al2O3 bulk rather than surface conduction is essential.
increases the modulus, but only beryllium This can be achieved by mixing conducting par-
decreases the density; the criterion of excellence is ticles into the polymer.
more sensitive to density than to modulus. To understand how to optimize this we need the
In Fig. 8 we return to the big picture. It shows concept of percolation. Percolation problems are
the moduli and densities of metals and polymers, easy to define, but not easy to solve. Research
and, encircled by a broken ellipse, those of high since 1960 has provided approximate solutions to
performance carbon, aramid, PE and glass fibers. most of the percolation problems associated with
The construction illustrated in Fig. 7 leads to famil- the design of hybrids (see Ref. [10] for a review).
ies of polymer–matrix composites that lie in the Think of mixing conducting and insulating spheres
shaded ellipse with that name, and to families of of the same size to give a large array. If there are
metal–matrix composites that lie in the ellipse too few conducting spheres for them to touch, the
above it. Both ellipses occupy areas of property array is obviously an insulator. If each conducting
space that were previously unoccupied by bulk sphere contacts just one other, there is still no con-
materials, and it is an important one, enabling the necting path. If, on average, each touched two,
design of new lightweight mechanical structures. there is still no path. Adding more spheres gives
Similar methods can be used to select materials larger clusters, but they can be large yet still dis-
5808 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

Fig. 8. Young’s modulus and density for 1850 polymers, metals and fibers (broken ellipse). Combining these to create polymer and
metal matrix composites fills a previously empty hole in material-property space [42].

Fig. 9. When conducting, particles or fibers are mixed into an insulating elastomer, a hole in material-property space is filled.
Carbon-filled butyl rubbers lie in this part of the space [42].
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5809

crete. The array first becomes a conductor when a fusion of water through solids differs from the flow
single trail of contacts links one surface to the rate of water through channels by a similar factor.
other, that is, when the fraction p of conducting It is then that single connections really matter.
spheres reaches the percolation threshold, pc. For Percolation influences mechanical properties
simple cubic packing pc = 0.248, for close packing too, particularly when mechanical connection is
pc = 0.180. For a random array it is somewhere in important, as in arrays of loose powders or fibers.
between—approximately 0.2.2 If there are no bonds between the particles or fib-
Make the spheres smaller and the transition is ers, the array has no tensile stiffness or strength.
smeared out. The percolation threshold is still 0.2, If each particle is bonded to another, or to several
but the first connecting path is now thin and forming discrete clusters, there is still no stiffness
extremely devious—it is the only one, out of the or strength. These only appear when there are con-
vast number of almost complete paths, that actually nected paths running completely through the array.
connects. Increase the volume fraction and the The plasticity of 2-phase hybrids, too, can be
number of conducting paths increases initially as viewed as a percolation problem. Plasticity may
(p⫺pc)2, then linearly, reverting to a rule of mix- start in one phase at a low stress, allowing patches
tures [13]. If the particles are very small, as much of slip to form, but full plasticity requires that the
as 40% may be needed to give good conduction. slip patches link to give connected paths through
But a loading of 40% seriously degrades the mold- the entire cross-section of the sample. Mechanical
ability and compliance of the polymer. and electrical percolation can be combined,
Shape gives a way out. If the spheres are exploiting scale. A latex reinforced with a small
replaced by fibers, they touch more easily and the volume fraction of cellulose fibers coated with
percolation threshold falls. If their aspect ratio is polypyrrole to make them conducting, gives a
f = L / d (where L is the fiber length, d the material with a shear modulus two orders of mag-
diameter) then—to an adequate approximation— nitude higher than the rule of mixture predicts
empirically, the percolation threshold falls from fc combined with good conduction because of the
to roughly fc / f1 / 2 [14–18]. Fig. 9 shows the area high aspect ratio of the fibers [19]. Here we are
of the property chart where these hybrids lie. With escaping the continuum bounds—a topic we return
sufficient aspect ratio the percolation threshold to later.
falls to a few percentage points.
The concept of percolation is a necessary tool 3.5. Creating anisotropy
in designing hybrids. Electrical conductivity works
that way; so too does the passage of liquids The elastic and plastic properties of bulk mono-
through foams or porous media—no connected lithic solids are frequently anisotropic, but weakly
paths, and no fluid flows; just one (out of a million so—the properties do not depend strongly on direc-
possibilities) and there is a leak. Add a few more tion. Hybridization gives a way of creating and
connections and there is a flood. Percolation ideas controlling anisotropy, and it can be large. We
are particularly important in understanding the have already seen an example in Fig. 7, which
transport properties of hybrids: properties that shows the upper and lower bounds for the moduli
determine the flow of electricity or heat, of fluid, of composites. The longitudinal properties of
or of flow by diffusion, specially when the differ- unidirectional long-fiber composites lie near the
ences in properties of the components are extreme. upper bound, the transverse properties near the
Most polymers differ from metals in their electrical lower one. The vertical width of the band in
conductivity by a factor of about 1020. The dif- between them measures the anisotropy.
Consider a second example, that of creating
2
hybrids with anisotropic thermal conductivity. A
These results are for infinite, or at least very large, arrays.
Experiments [11,12] generally give values in the range 0.19– saucepan made from a single material when heated
0.22, with some variability because of the finite size of the on an open flame, develops hot spots that can
samples. locally burn its contents. That is because the sauce-
5810 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

pan is thin, and heat is transmitted through the trajectories for lⲚ and l that separate. The
thickness more quickly than it can be spread trans- maximum separation occurs broadly where each
versely to bring the entire pan surface to a uniform occupies about half the thickness, where the ratio
temperature. The metals of which saucepans are of the conductivities (the anisotropy ratio) is 3.8.
usually made—cast iron, or aluminum or copper— Mechanical anisotropy is most easily created
have a isotropic thermal conductivities whereas and managed through shape. This is the topic of
what we clearly want is a thermal conductivity that the next section.
is higher in the transverse direction than in the
through-thickness direction. A bi-layer (or multi-
layer) hybrid can achieve this. 4. Shape: structures, sandwiches and
Heat transmitted transversely in a bi-layer sheet segmented assemblies
has two parallel paths; the total heat transmitted is
The shape and configuration of components A
a sum of that in each of the paths. If it is made of
and B of a hybrid play a key role in determining
a layer of material 1 with thickness t1 and conduc-
its properties. Shape can be used to enhance or
tivity l1, bonded to a layer of material 2 with thick-
diminish stiffness and strength, to impart damage
ness t2 and conductivity l2, the conductivity paral-
tolerance, and—as we have already seen—to
lel to the layers is
manipulate the percolation limit.
lII ⫽ fl1 ⫹ (1⫺f)l2 (4)
4.1. Shape efficiency and shape factors
(a the rule of mixtures), where f = t1 / (t1 + t2). Per-
pendicular to the layers the conductivity is Beams with hollow-box or I-sections are stiffer
1 f (1⫺f) and stronger in bending than solid sections of the
⫽ ⫹ (5) same cross-sectional area; so, too, are panels with
lⲚ l1 l2 ribs or waffle stiffeners, or those with an expanded
(the harmonic mean). Fig. 10 shows lⲚ and l plot- core to create a sandwich (Fig. 11). These are
ted against f for a bi-layer of copper (l=390
W/m K) and cast iron (l = 30 W / m K). For single
materials the two are equal; layering them gives

Fig. 10. Creating anisotropy. The thermal conductivities of Fig. 11. Making high-efficiency structures. Shape gives the
copper and cast iron are isotropic. Anisotropy is created by sections a greater flexural stiffness and strength per unit mass
combining them as a bi-layer. than the solid section from which they are made.
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5811

examples of the use of shape to increase structural high as 50; for multi-strand or multi-leaf structures
efficiency. To characterize this we need a metric— as low as 0.01.
a way of measuring the structural efficiency of a Note the origins of efficiency. The flanges of the
section shape, independent of the material of which I-section or the faces of the sandwich panels lie far
it is made. An obvious one is that given by the from the neutral axis; they stretch when the section
ratio j of the stiffness or strength of the shaped is loaded in bending. Subdivision, as in Fig. 11,
section to that of a ‘neutral’ reference shape. For lowers efficiency because the slender strands or
a beam we take the reference shape to be that of leaves bend easily, but do not stretch when the sec-
a solid square section with the same cross-sectional tion is bent: an n-strand cable is less stiff by a fac-
area, and thus the same mass per unit length, as tor of 3/πn than the solid reference section; an n-
the shaped section. For a panel, we take it to be a leaf panel by a factor 1/n2. There is an underlying
solid, plain panel with the same mass per unit area principle here: stretch dominated structures have
as the shaped section, as shown in the figure. We high structural efficiency; bending dominated
call j the shape factor [2,20] and define that for structures have low.
stiffness je as
4.2. Shape on a micro-scale
Flexural stiffness of shaped section
je ⫽ (6)
Flexural stiffness of reference section The sections of Fig. 11 achieve efficiency
and that for strength jf as through their macroscopic shape. Structural
efficiency can be manipulated in another way:
Flexural strength of shaped section through shape on a small scale; microscopic or
jf ⫽ (7)
Flexural strength of reference section “micro-structural” shape. Wood is an example. The
Shape can be used to reduce flexural stiffness solid component of wood (a composite of cellu-
and strength as well as increase them. Springs, sus- lose, lignin and other polymers) is shaped into pris-
pensions, flexible cables and other structures that matic cells, each cell like the hollow tube of Fig.
must flex yet have high tensile strength, use shape 11. The effect is to disperse the solid component
to give a low bending stiffness. Low shape further from the axis of bending or twisting of the
efficiency is achieved by forming the material into branch or trunk of the tree, increasing its flexural
strands or leaves, as suggested Fig. 12. Values of stiffness and strength. This is not the only possi-
j for the stiffness of structural sections can be as bility; low efficiency structures give materials with
low stiffness and strength, desirable in cushioning
and packaging.

4.3. Ultra-light, low stiffness hybrids

The point has been made that stretching is a stiff


mode of loading, bending is a compliant one. A
material that responds, at the micro-structural
level, by bending no matter how it is loaded
remotely is much less stiff than one that responds
by stretching. Material made by foaming have
structures that respond in this way.
Fig. 13 shows an idealized cell of a low-density
foam. It consists of solid cell walls or edges sur-
rounding a void space, each cell with an overall
Fig. 12. Making low-efficiency structures. Shape gives the space-filling shape. Cellular solids are charac-
sections a lower flexural stiffness and strength per unit mass terized by their relative density, which for the
than the solid section from which they are made. structure shown here (with t ⬍ ⬍ᐉ) is
5812 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

Fig. 13. A cell in a low density foam. When the foam is loaded, the cell edges bend, giving a low-modulus structure.

r∗
rs

t
ᐉ 冉冊 2
(8)
yield when the force exerted on them exceeds their
fully plastic moment

where r∗ is the density of the foam, rs is the den- sst3


sity of the solid of which it is made, ᐉ is the cell Mf ⫽ (11)
4
size, and t is the thickness of the cell edges. A
remote compressive stress s exerts a force F ⬀
where ssis the yield strength of the solid of which
sᐉ2 on the cell edges, causing them to bend as
the foam is made. This moment is related to the
shown in the figure, and leading to a bending
remote stress by M ⬀ FL ⬀ sL3. Assembling these
deflection d. For the open-celled structure shown
results gives the failure strength s∗
in the figure, the bending deflection is given by

d⬀
FL3
EsI
(9)
s∗ r

s s rs冉冊 3/2
(12)

where Es is the modulus of the solid of which the (bending ⫺ dominated behavior)
foam is made and I = t4 / 12 is the second moment
of area of the cell edge of square cross-section, t This behavior is not confined to open-cell foams
× t. The compressive strain suffered by the cell as with the structure shown in Fig. 14. Most closed-
a whole is then e = 2d / ᐉ. Assembling these results cell foams also follow these scaling laws. At first
gives the modulus E∗ = s / e of the foam as sight an unexpected result because the cell faces
E∗
Es

r
rs 冉冊 2
(10)
must carry membrane stresses when the foam is
loaded, and these should lead to a linear depen-
dence of both stiffness and strength on relative
(bending ⫺ dominated behavior) density. The explanation lies in the fact that the
Since E = Es when r = rs, we expect the constant cell faces are very thin; they buckle or rupture at
of proportionality to be close to unity—a specu- stresses so low that their contribution to stiffness
lation confirmed both by experiment and by and strength is small, leaving the cell edges to
numerical simulation. carry most of the load (for further details, see
A similar approach can be used to model non- Ref. [21]).
linear properties such as strength. The cell walls
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5813

Fig. 14. A micro-truss structure and its unit cell.

4.4. Ultra-light, high stiffness hybrids stretch-dominated structure. On average one third
of its bars carry tension when the structure is
If conventional foams have low stiffness loaded in simple tension, regardless of its direc-
because other configuration of the cell edges tion. Thus

冉冊
allows them to bend, might it not be possible to
E∗ 1 r
devise other configurations in which the cell edges ⬇ (15)
were made to stretch instead? This thinking leads Es 3 rs
to the idea of micro-truss structures [22,23]. To for isotropic stretch-dominated behavior and

冉冊
understand these we need the Maxwell stability cri-
terion. s∗ 1 r
⬇ (16)
The condition that a pin-jointed frame of b struts ss 3 rs
and j frictionless joints to be both statically and
kinematically determined i.e. just rigid [24,25], in for isotropic stretch-dominated behavior.3
2-dimensions, is: Prismatic structures do even better, provided
they are loaded parallel to the prism axis. Fig. 15
M ⫽ b⫺2j ⫹ 3 ⫽ 0 (13) shows four such structures which are common in
and in 3-dimensions is nature. It is helpful to think, as before, of the
expansion of a solid bar, shown in the center, to
M ⫽ b⫺3j ⫹ 6 ⫽ 0 (14) give the structures, with no change of mass—here
If M ⬍ 0, the frame is a mechanism. It has no the solid black represents solid material, the dotted
stiffness or strength, but will collapse if loaded. If areas represent low density foam and the open
its joints are locked (instead of pin-jointed) the bars areas represent space. The expansion has moved
of the frame bend when the structure is loaded. If, material away from the axis of bending, increasing
instead, M ⱖ 0 the frame ceases to be a mech- the second moment of area about any axis con-
anism; its members carry tension or compression tained in the plane of the cross-section, and
when the frame is loaded, and it becomes a stretch- increasing efficiency in the sense of Eqs. (6) and
dominated structure. (7). For these structures the axial modulus and
These criteria give a basis for the design of strength (measured parallel to the prism axis) fol-
efficient micro-truss structures. For the cellular
structure of Fig. 13 M ⬍ 0, and it is bending domi- 3
This assumes that failure occurs by the axial yielding of a
nated. For the structure shown in Fig. 14, however, bar. If the bars are slender or have low modulus they may fail
M ⬎ 0 and it behaves as an almost isotropic, instead by buckling, giving a lower strength.
5814 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

4.5. Ultimate efficiency: the sandwich

A sandwich panel epitomizes the concept of a


hybrid. It combines two materials in a specified
geometry and scale—one forming the faces, the
other the core—to give a structure of high stiffness
and strength at low weight (Figs. 11 and 17). The
separation of the faces by the core increases the
moment of inertia I and the section modulus Z of
the panel with little increase in weight, producing
an efficient structure for resisting bending and
buckling loads. Sandwiches are found where
weight-saving is critical: in aircraft, trains, trucks
and cars, in portable structures, and in sports
equipment. Nature, too, makes use of sandwich
designs: sections through the human skull, the
wing of a bird and the stalk and leaves of many
Fig. 15. Four extensive micro-structured materials which are plants show a low-density foam-like core separat-
mechanically efficient: (a) prismatic cells, (b) fibers embedded ing solid faces. The faces carry most of the load,
in a foamed matrix, (c) concentric cylindrical shells with foam
between, and (d) parallel plates separated by foamed spacers.
so they must be stiff and strong; and they form the
exterior surfaces of the panel so they must tolerate
the environment in which they operate. The core
lows Eqs. (15) and (16), but with a constant of occupies most of the volume, it must be light, and
proportionality not of 1/3 but of unity: stiff and strong enough to carry the shear stresses
E∗ r

Es rs 冉冊 (17)
necessary to make the whole panel behave as a
load bearing unit, but if the core is much thicker
than the faces these stresses are small.
for prismatic stretch-dominated behavior and So far we have spoken of the sandwich as a

冉冊
structure: faces of material A supported on a core
s∗ r
⬇ (18) of material B, each with its own density and modu-
ss rs lus. But we can also think of it as a material with
for prismatic stretch-dominated behavior.3 its own set of properties, and this is useful because
Loaded transversely, however, they are bending it allows comparison with more conventional
dominated and follow power laws like those of materials. To do so we must analyze sandwich per-
Eqs. (10) and (12); they are thus exceedingly formance [21,26–29]. We shall use, as a criterion
anisotropic. of excellence, the bending stiffness per unit width,
These results are summarized in Fig. 16, in Sw divided by the mass per unit area, ma.
which the modulus E∗ is plotted against the den- The bending stiffness of the panel per unit
sity r. Stretch dominated, prismatic microstruc- width, Sw, is given by
tures like wood give moduli that scale as r / rs
Sw ⫽ (EI)sand ⫽ (19)
(slope 1); bending dominated, cellular, microstruc-

冋 册冦 冧
tures like that of polystyrene foam give moduli that 1
scale as (r / rs)2 (slope 2). Given that the density 1 3
(d ⫺ c3)Ef BEftc
can be varied through a wide range, this allows 12 1⫹
great scope for material design. Note how the use 2GcL2
of microscopic shape has expanded the occupied
area of E–r space. where the dimensions, d, c, t and L are identified
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5815

Fig. 16. Foams and micro-truss structures are hybrids of material and space. Their mechanical response depends on their structure.
Foams are usually bending dominated (Eqs. (10) and (12)), and lie along a line of slope 2 on this chart. Micro-truss structures are
stretch-dominated (Eqs. (16) and (17)) and lie on a line of slope 1. Both extend the occupied area of this chart by many decades [42].

(d3/12 is the second moment of area of a homo-


geneous panel of thickness d, and we assume
td). Thus
6t
Ẽ ⫽ E (21)
d f
The panel has a density, r, that is the weighted
Fig. 17. The ultimate light, stiff hybrid: the sandwich panel. mean of its faces and core

in Fig. 16, Ef is Young’s modulus of the face sheets


and Gc is the shear modulus of the core. The
r⫽
2t
d f 冉 冊
2t
r ⫹ 1⫺ rc
d
(22)

numerical constant B depends only on the way the (just Eq. (1) with f replaced by 2t/d). The core den-
panel is loaded. Eq. (19) has two terms. The first, sity is chosen to be low, making the second term
in square brackets, is the stiffness if bending were small compared with the first. Neglecting it, and
the only mode of deformation. The second, in curly combining the last two equations gives

冉冊
brackets, is the knock-down in stiffness caused by
shear in the core. If the core adequately resists r
Ẽ⬇3 E ⫽ 3fEf (23)
shear (and this is easily achieved) the second term rf f
can be neglected. To think of the sandwich as a
material, we define an apparent modulus Ẽ—the Comparing this with Eq. (17) we find a gain in
modulus of a homogeneous material with the same stiffness by a factor of 3 (Es there equates to Ef
bending stiffness as the sandwich, requiring that here). The sandwich in flexure is approximately
three times more efficient than the most efficient
1 3 3 1 d3 fibrous composite, even when the fibers are all
(EI)sand ⫽ (d ⫺c )Ef ⫽ td2Ef ⫽ Ẽ (20)
12 2 12 aligned normal to the axis of bending.
5816 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

This, of course, is an idealization. The core jectile, will shatter. One made of small glass
always shears somewhat, and it does have some bricks, laid as bricks usually are, will lose a brick
mass; a more precise analysis copes with this [29]. or two but not shatter totally; it is damage-tolerant.
It leads to the performance shown in Fig. 18, which By sub-dividing and separating the material, a
has been constructed in the same way as Fig. 7. It crack in one segment does not penetrate into its
shows the modulus and density of A + B hybrids. neighbors, allowing local but not global failure.
The shaded band is bounded by the upper and That is the principle of “topological toughening”.
lower bounds of Eqs. (1)–(3), describing particu- Builders in stone and brick have exploited the idea
late and fibrous composites. The flexural perform- for thousands of years: both materials are almost
ance of the sandwich is shown as a dashed line; at as brittle as glass, but buildings made of them—
its mid-section it lies a factor of 3 above the upper even those made without cement (“dry-stone
bound rule of mixtures of Eq. (2). The criterion of building”)—survive ground movement, even earth-
excellence for minimum weight design with pre- quakes, through their ability to deform with some
scribed bending stiffness, listed in Table 1, is that local failure, but without total collapse.
of maximizing E1 / 3 / r. Contours of this criterion Taking the simplest view, two things are neces-
are plotted as diagonal lines on the figure, increas- sary for topological damage tolerance: discreteness
ing towards the top left. The sandwich out-per- of the structural units, and an interlocking of the
forms all alternative hybrids of A + B. units in such a way that the array as a whole can
carry load. Brick-like arrangements (Fig. 19a) are
4.6. Subdivision as a design variable damage tolerant in compression and shear, but dis-
integrate under tension. Strand and layer-like struc-
We have already seen how subdivision can tures (shown earlier as Fig. 12) are damage-tolerant
reduce stiffness (Fig. 12). Here we examine in tension because if one strand fails the crack does
another way in which it can be used: to impart not penetrate its neighbors—the principle of multi-
damage tolerance. A glass window, hit by a pro- strand ropes and cables. The jigsaw puzzle con-
figuration (Fig. 19b) carries in-plane tension, com-
pression, and shear, but at the cost of introducing
a stress concentration factor of about √R / r, where
R is the approximate radius of a unit and r that
of the interlock. Dyskin et al. [30–32] explore a
particular set of topologies that rely on compress-
ive or rigid boundary conditions to create continu-
ous layers that tolerate out-of-plane forces and
bending moments, illustrated in Fig. 19c. This is
done by creating interlocking units with non-planar
surfaces that have curvature both in the plane of
the surface and normal to it. Provided the array is
constrained at its periphery, the nesting shapes lim-
its the relative motion of the units, locking them
together. The bending stiffness of the array is pro-
portional to the stiffness of the boundary con-
straint, falling to zero as the constraint is relaxed.
Topological interlocking of this sort allows the for-
mation of continuous layers that can be used for
ceramic claddings or linings to give surface protec-
Fig. 18. Sandwich panels (broken line) extend the range of
flexural modulus per unit mass into areas not occupied by tion.
monolithic materials. Their flexural moduli lie above that pre- The damage tolerance can be understood in the
dicted by a rule of mixtures by a factor of approximately 3. following way. We suppose that the units of the
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5817

Fig. 19. Examples of topological interlocking: discrete, unbonded structures that carry load. (a) Brick-like assemblies of rectangular
blocks carry axial compression (syy), but not tension or shear. (b) The 2-dimensional interlocking of a jig-saw puzzle carries in-plane
loads ( ± sxx, ± syy, sxy). (c) The units suggested by Dyskin et al., when assembled into a continuous layer and clamped within a
rigid boundary around its edge, can carry out-of-plane loads and bending moments ( ± sxz, ± syz, Mxz, Myz) (Fig. 15(c) derived from
Ref. [30]).

structure are all identical, each with a volume Vs, stress, if sufficiently large, causes some segments
and that they are assembled into a body of volume to fail. We refer to the fraction that has failed as
Vt; there are therefore n = Vt / Vs segments. We the damage, D. If loaded such that each segment
describe the probability of failure of a segment carries a uniform stress s, the damage is simply
under a uniform tensile stress s by a Weibull prob- Pf(Vs,s). If some segments fail, the body as a
ability function: whole remains intact; global failure requires that a

再 冎
fraction D∗ , the critical damage, (say, 10%) must
Vsm fail. Inverting Eq. (24) with V = Vs gives the global
Pf(V,s) ⫽ 1⫺exp⫺ (24)
Vosmo failure stress s∗s of the segmented body:
where m, Vo and so are constants [33,34]. If the
body were made of a single monolithic piece of
the brittle solid, this equation, with V = Vt, would
ss∗ ⫽ so 再 Vo
Vs
ln(1⫺D∗) 冎 1/m
(26)

describe the failure probability. To calculate the Thus segmentation increases the allowable design
design stress s∗t we set an acceptable value for P, stress from s∗t to s∗s , factor of

再 冎 再 冎
which we call P∗ (say 10⫺6, meaning that it is
ss∗ ln(1⫺D∗) 1/m
D∗ 1/m
acceptable if one in a million fail) and invert the ⫽ n ⬇ n ∗ (27)
equation to give s∗t ln(1⫺P∗) P

s∗t ⫽ so 再 Vo
Vt
ln(1⫺P∗) 冎
1/m
(25)
(expanding the logarithm as a series and retaining
the first term—an acceptable approximation for
small P∗ and D∗). Both n and D∗/P∗ are con-
Now consider the segmented body. A remote siderably greater than 1, so the equation suggests
5818 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

that segmentation always increases the design ries. It is constructed using Kc = 2, and values of
stress. nD∗/P∗ between 104 and 1010. The Weibull modu-
The merit of this argument is its simplicity, illus- lus m of brittle solids typically lies in the range 5
trating how damage tolerance can arise. But it is a (typical of a porous brittle ceramic) and 30 (a value
little too simple. What are the underlying assump- that might describe the variability of a high techni-
tions? First, that the number of segments is large; cal ceramic). When the Weibull modulus is large—
only then can the damage D be reliably equated to more than 25—segmentation offers little gain,
the failure probability Pf. Second, that the stress s indeed there may be a loss of allowable stress. But
is uninfluenced by damage; in reality its average when it is low (as it is in building materials like
value increases by the factor 1 / (1⫺D). Third, that brick and stone, and as it can be for certain cer-
the stress is uniform; in reality both the inter- amics and glasses), there is a gain. Segmentation
locking shapes and the load shed by a failed seg- gives a new design variable. Since the only
ment onto its immediate neighbors increases the requirement is that of interlocking shape, the seg-
stresses locally by a effective stress concentration ments can be made of different materials. Just as
factor Kc. And fourth, that segment failure is builder constructing a wall can insert porous bricks
uncorrelated (“non-associated cracking”), not for ventilation and transparent bricks to admit
linked (“associated cracking”). The first of these is light, the designer of a segmented hybrid can add
simply a pre-condition that must be met for seg- functionality through the choice of material for
mentation to be successful. The second and third the units.
of these can be included approximately by mod-
ifying Eq. (27) to give:

冉 冊再 冎
5. Scale: escaping the continuum bounds
s∗s 1⫺D∗ D∗ 1/m
⬇ n ∗ (28)
s∗t Kc P Continuum methods, on which many of the
The last depends on the values of the Weibull con- arguments of Sections 3 and 4 are based, lead to
stants, and requires numerical modeling of the kind results that depend on shape, but are independent
used by Curtin [35] and Shaw [36] to explore the of scale. Materials, however, are not continua; the
conditions for non-associated cracking. Eq. (28) discreetness at the atomistic level introduces a
however is still useful—it allows for both a gain length scale that, as it is approached (even
and a loss in allowable design stress. remotely), causes properties to behave in ways that
Fig. 20 gives an idea of the messages that it car- differ, sometimes dramatically, from the simple
continuum estimates. The general result is that
mechanical and transport properties become scale-
dependent when the hybrid microstructure is sub-
micron.
This, the bread and butter of the materials scien-
tist, extends beyond the scope of this paper. Here
we simply list some of the origins of scale-depen-
dent properties. We have already seen how seg-
mentation can lead to scale-dependent damage tol-
erance. Small particles and fine fibers are generally
stronger than bulk samples of the same materials
because the probability that they contain a large
defect scales with their volume. Precipitation hard-
ened materials, aged to peak hardness, can
approach a strength of s∗⬇sp√f (where sp is the
Fig. 20. The ratio ss / st for the listed values of nD∗/P∗, with strength of the particles and f their volume
Kc = 2 and D∗ = 0.1. fraction), compared with upper bound s∗⬇spf +
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5819

sm(1⫺f) of the continuum approximation. Fine- continuum methods must be replaced by statistical
grained materials, particularly those with grains of or dislocation mechanics.
a few nanometers, exhibit strengths that exceed Much, in a short paper such as this, has been
those of large-grained bulk samples. Scale, then, ignored. Fabricating successful hybrids can be dif-
can have a profound influence on mechanical ficult and expensive (but so, too, is the alternative
properties. of seeking to develop a new monolithic material).
Transport properties, too, can be scale depen- Part of the difficulty stems from the multitude of
dent. The mean-free paths of electrons, phonons possible choices: choice of materials, choice of
and of diffusing atoms and molecules are limited process to combine them, and choice of the internal
by the scale of the microstructure when this is geometry and topology of the constitutive
small—a fact exploited in micro-cellular foams to materials. Part derives from the need to make these
give exceptional thermal insulation. Convection, choices in such a way as to optimally meet a set
acoustic absorption and light scattering, too, are of design requirements. The hybrid must be both
directly linked to aspects of structural scale. feasible and optimal.
To explore this design space efficiently optimiz-
ation tools are needed. The starting point is a sim-
6. Summary and conclusions ple “screening” of options—the obvious way for-
ward when the selection space is discrete. Thus it
The properties of engineering materials can be
is possible to create a database of, say, composites
thought of as defining the axes of a multi-dimen-
by computing the properties of virtual materials
sional space with each property as a dimension.
with 5%, 10%, 15%… of reinforcing fibers and
Sections through this space can be mapped. These
discrete choices of lay-up; selection proceeds by
maps reveal that some areas of property-space are
rejecting all the combinations that fail to meet the
occupied, others are empty—there are holes. The
holes can sometimes be filled by making hybrids: design requirements. This method becomes
combinations of two (or more) materials, or of impractical when many material choices and lay-
material in space, in chosen configuration and ups are allowed. When the optimization variables
scale. Here we have surveyed conceptual tools for are continuous (such as the volume fraction of
suggesting and assessing hybrids to fill specified reinforcement in a composite with pre-chosen
needs. A useful starting point is the concept of a constituents), linear programming or steepest
hybrid as “A + B + shape + scale”. Successful descent methods can be efficient. When the “poten-
hybrids, as a rule, exploit the first three of these; tial landscape” is very rugged, simulated annealing
with micron and nanometer scale fabrication tech- can offer a practical alternative. When the variables
nologies now a reality, it becomes possible to add are both discrete (such as the choice of materials)
the last, opening up wider horizons. and continuous (the thicknesses of face sheets and
Continuum bounding methods give tools for core of a sandwich, for instance), a genetic algor-
scanning the possibilities offered by a set of ithm, which allows an unevenly populated space to
hybrids, provided the scale is such that the con- be explored, can be efficient. Refs. [37–41]
tinuum approximation applies. Shape—the way A examples of their use in hybrid design.
and B are configured—can extend the populated These tools allow promising candidates to be
areas of property-space in ways that complement identified. They need the back-up of “expert tools”
efforts to create new monolithic materials. Dis- advising on the compatibility of the constituents,
criminating choice of shape can enhance or dimin- the practicality of processes to assemble them into
ish physical, mechanical, thermal and electrical a hybrid, and the loss of properties (the “knock-
properties. Scale introduces a new variable. In down factor” relating real and ideal hybrids) for
hybrids with structural units that are sub-micron, a a given combination of the materials, architecture
new length scale (basically that of the atom) makes and process.
itself evident. Here the bounds break down, and
5820 M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821

Acknowledgements R, Flandin L, Gautier C. Polymer-based nanocompoosite:


effect of filler–filler and filler–matrix interactions. Adv
Eng Mater 2001;3(8):571.
The ideas, methods and tools described here [20] Ashby MF. On material and shape. Acta Mater
have evolved over the past 15 years. Numerous 1991;39:1025.
colleagues in many countries have (sometimes [21] Gibson LJ, Ashby MF. Cellular solids, structure and
unknowingly) stimulated or contributed to this properties, 2nd ed. Cambridge, UK: Cambridge University
evolution. Among these we would particularly like Press, 1997.
[22] Deshpande VS, Ashby MF, Fleck NA. Foam topology:
to recognize Profs. Mick Brown, Chris Calladine,
bending versus stretching dominated architectures. Acta
Norman Fleck and David Cebon (all of Cambridge Mater 2001;49:1035–40.
University), Dave Embury (McMaster University, [23] Deshpande VS, Fleck NA, Ashby MF. Effective properties
Canada), Tony Evans (UCSB), John Hutchinson of the octet-truss lattice material. J Mech Phys Sol
(Harvard University) and Haydn Wadley (UVA), 2001;49:1747–69.
and Dr. Luc Salvo (University of Grenoble). [24] Maxwell JC. On the calculation of the equilibrium and
stiffness for frames. Phil Mag 1864;27:294.
[25] Calladine CR. Theory of shell structures. Cambridge, UK:
Cambridge University Press, 1983.
References [26] Allen HG. Analysis and design of structural sandwich
panels. Oxford, UK: Pergamon Press, 1969.
[1] Ashby MF. On the engineering properties of materials. [27] Cordon J. Honeycomb structure. In: Dostal CA, editor.
Acta Metall 1989;37:1273–93. Engineered materials handbook. Metals Park, Ohio, USA:
[2] Ashby MF. Materials selection in mechanical design, 2nd ASM International; 1990:721–8.
ed. Oxford: Butterworth Heinemann, 1999. [28] Zenkert D. In: An introduction to sandwich construction.
[3] Kromm FX, Quenisset JM, Harry R, Lorriot T. An Solihull, London, UK: Engineering Advisory Services
example of multimaterial design. Adv Eng Mater Ltd., Chameleon Press Ltd, 1995, ISBN 0 947817778.
2002;4:371–4. [29] Ashby MF, Evans AG, Fleck NA, Gibson LJ, Hutchinson
[4] Watt JP, Davies GF, O’Connell RJ. Reviews of geophys- JW, Wadley HNG. Metal foams, a design guide. Oxford,
ics and space physics 1976;14:541. UK: Butterworth Heinemann, 2000.
[5] Schoutens JE, Zarate DA. Composites 1986;17:188. [30] Dyskin AV, Estrin Y, Kanel-Belov AJ, Pasternak E.
[6] Chamis CC. In: Engineers guide to composite materials. Toughening by fragmentation: how topology helps. Adv
Materials Park, Ohio, USA: American Society of Metals, Eng Mater 2001;3:885–8.
1987:3-8–3-24. [31] Dyskin AV, Estrin Y, Kanel-Belov AJ, Pasternak E. Topo-
[7] Ashby MF. Criteria for selecting the components of com- logical interlocking of platonic solids: a way to new
posites. Acta Mater 1993;41:1313–35. materials and structures. Phil Mag 2003;83:197–203.
[8] Eshelby JD. Proc R Soc Lond 1957;A241:376. [32] Dyskin AV, Estrin Y, Pasternak E, Kohr HC, Kanel-Belov
[9] Hashin Z, Strikman S. J Appl Phys 1962;33:3125.
AJ. Fracture resistant structures based on topological inter-
[10] Stauffer D, Aharony A. Introduction to percolation theory,
locking with non-planar contacts. Adv Eng Mater
2nd ed. London, UK: Taylor and Francis, 1994 revised.
2003;5(3):116–9.
[11] Fitzpatrick JP, Malt RB, Spaepen F. Phys Lett
[33] Weibull W. J Appl Mech 1951;18:293.
1974;A47:207.
[34] Davidge RW. Mechanical behavior of ceramics. Cam-
[12] Brown LM. A simple model of a metal non-metal tran-
bridge, UK: Cambridge University Press, 1986.
sition. Physics Education 1977;July issue:318–20.
[13] Last BJ, Thouless DJ. Phys Rev Lett 1971;27:1719. [35] Curtin WA. Acta Metall et Mater 1993;41:1369.
[14] Nielsen LE. Thermal conductivity of particulate-filled [36] Shaw MC. The fracture mode of ceramic/metal multilay-
polymers. J Appl Polym Sci 1973;17:3819. ers: role of the interface. Key Eng Mater 1996;116-
[15] Nielsen LE. The thermal and electrical conductivity of 117:261–78.
two-phase systems. Ind Eng Chem Fund 1974;13:17. [37] Salvo L, Bréchet Y, Pechambert P, Bassetti D, Jantsen A.
[16] Bigg DM. Conductive polymeric compositions. Polym Logiciels de sélection des composites. Matériaux et Tech-
Eng Sci 1977;17:892. nique 1998;5:31.
[17] Bigg DM. Mechanical and conductive properties of metal [38] Landru D, Bréchet Y. New design tools for materials and
fibre-filled polymer composites. Composites 1979;April process selection. Materiaux et Techniques 2002;4:6.
issue:95–100. [39] Bréchet Y, Bassetti D, Landru D, Salvo L. Challenges in
[18] Yi Y-B, Sastry AM. Analytical approximation of the 2- material and process selection. Prog Mater Sci
dimensional percolation threshold for fiels of over lapping 2001;46:407.
ellipses. Phys Rev 2002;E66 066130-1–066130-8. [40] Bréchet Y, Ashby MF, Salvo L. Selection des materiaux
[19] Bréchet Y, Cavaille JY, Chabert E, Chazeau L, Dendievel et des procedes de mis en oeuvre. Switzerland: Les Presses
M.F. Ashby, Y.J.M. Bréchet / Acta Materialia 51 (2003) 5801–5821 5821

Polytechniques et Universitaires Romandes de Lausanne, [42] CES 4. The Cambridge Material Selector, Granta Design,
2002. Cambridge. 2003;(www.Grantadesign.com)
[41] Ashby M, Bréchet Y, Cebon D, Salvo L. Materials and
process selection strategies. To appear in Advanced
Engineering Materials, 2003.

You might also like