Hybrids A-B

download Hybrids A-B

of 21

Transcript of Hybrids A-B

  • 8/8/2019 Hybrids A-B

    1/21

    Acta Materialia 51 (2003) 58015821 www.actamat-journals.com

    Designing hybrid materials

    M.F. Ashby a,, Y.J.M. Brechet b

    a Engineering Department, University of Cambridge, Trumpington Street, CB2 1PZ Cambridge, UKb L.T.P.C.M., Domaine Universitaire de Grenoble, BP75, 38402 Saint Martin dHeres Cedex, France

    Accepted 31 August 2003

    Abstract

    The properties of engineering materials can be mapped, displaying the ranges of mechanical, thermal, electrical andoptical behavior they offer. These maps reveal that there are holes: some areas of property-space are occupied andothers are empty. The holes can sometimes be filled and the occupied areas extended by making hybrids of two ormore materials or of material and space. Particulate and fibrous composites are examples of one type of hybrid, butthere are many others: sandwich structures, foams, lattice structures and more. Here we explore ways of designinghybrid materials, emphasizing the choice of components, their shape and their scale. The new variables expand thedesign 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

    1.1. Extending material-property space

    Fig. 1 is an example of a material-propertychart. It shows the thermal conductivities of some2300 different materials, plotted against their

    Youngs moduli. It is one of many, each a slicethrough material-property space; the assembly ofall the slices can be thought of as a map of thisspace [1,2]. All the charts have one thing in com-mon: parts of them are populated with materials

    Corresponding author. Tel.: +44-01223-332-635; fax: +44-01223-332-662. The Golden Jubilee IssueSelected topics in Materials

    Science and Engineering: Past, Present and Future, edited byS. Suresh.

    1359-6454/$30.00 2003 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

    doi:10.1016/S1359-6454(03)00441-5

    and parts are not. Some parts are inaccessible forfundamental reasons that relate to the size of atomsand the nature of the forces that bind their atomstogether. But other parts are empty even though,in principle, they are accessible. If they wereaccessed, the new materials that are there couldallow novel design possibilities.

    One approach to thisthe traditional oneisthat of developing new metal alloys, new polymerchemistries and new compositions of glass and cer-amics so as to extend the populated areas of theproperty charts, but this can be an expensive anduncertain process. An alternative is to combine twoor more existing materials so as to allow a super-position of their propertiesin short, to create ahybrid (Fig. 2). The spectacular success of carbonand glass-fiber reinforced composites at oneextreme, and of foamed materials at another

  • 8/8/2019 Hybrids A-B

    2/21

    5802 M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    Fig. 1. A material-property chart of thermal conductivity and Youngs modulus for 2300 materials. Each small circle is a plot ofthese 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].)

    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,sandwiches and almost all natural materials. One might imagine

    two further dimension: those of shape and scale.

    (hybrids of material and space) in filling previouslyempty areas of the property charts is encourage-

    ment enough to explore ways in which such

    hybrids might be designed. What is the best wayto go about doing so?

    1.2. What might we hope to achieve?

    Fig. 3 shows schematically the fields occupiedby two families of materials, plotted on a chart

    with properties P1 and P2 as axes. Within each fielda single member of that family is identified(materials M1 and M2). What might be achieved

    by making a hybrid of the two? The figure showsfour 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 ofthe following.

    The best of both scenario (Point A). The ideal,often, is the creation of a hybrid with the best

    properties of both components. There are

  • 8/8/2019 Hybrids A-B

    3/21

    5803M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    Fig. 3. The possibilities of hybridization. The properties of the

    hybrid reflect those of its component materials, combined inone of several possible ways.

    examples, most commonly when a bulk pro-

    perty of one material is combined with the sur-

    face properties of another. Zinc coated steel hasthe strength and toughness of steel with the cor-rosion resistance of zinc. Glazed pottery

    exploits the formability and low cost of clay

    with the impermeability and durability of glass. The rule of mixtures scenario (Point B). When

    bulk properties are combined in a hybrid, as in

    structural composites, the best that can be

    obtained is often the arithmetic average of theproperties of the components, weighted by their

    volume fractions. Thus unidirectional fiber com-posites have an axial modulus (the one parallelto the fibers) that lies close to the rule of mix-tures.

    The weaker link dominates scenario (Point

    C). Sometimes we have to live with a lessercompromise, typified by the stiffness of particu-late composites, in which the hybrid propertiesfall below those of a rule of mixtures, lying

    closer to the harmonic than the arithmetic mean

    of the properties. Although the gains are lessspectacular, they can still be useful.

    The worst of both scenario (point D)notsomething we want.

    These set certain fixed points, but the list is not

    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 filledpolymers, composites or wood are treated as

    materials in their own right, each characterized by

    its own set of material properties. Otherslike gal-vanized steelare seen as one material (steel) towhich 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-wich panels illustrate the duality, sometimesviewed as two sheets of face-material separated by

    a core material, and sometimesto allow compari-son with bulk materialsas a material with theirown density, flexural stiffness and strength. To callany one of these a material and characterize itas such is a useful shorthand, allowing designersto use existing methods when designing with them.

    But if we are to design the hybrid itself, we must

    deconstruct it, and think of it as a combination of

    materials (or of material and space) in a definedgeometry.

    2. The method: A + B + shape + scale

    First, a working definition: a hybrid material isa combination of two or more materials in a prede-termined geometry and scale, optimally serving a

    specific engineering purpose [3], which we para-phrase as A + B + shape + scale. Here we allow

    for the widest possible choice of A and B, includ-ing the possibility that one of them is a gas or sim-

    ply space. These new variables expand the designspace, allowing an optimization of properties that

    is not possible if choice is limited to single, mono-

    lithic materials.The basic idea, illustrated in Fig. 4, is this.

    Monolithic materials offer a certain portfolio of

    properties on which much engineering design isbased. But if the design requirements are excep-

    tionally demanding, no single material may be

  • 8/8/2019 Hybrids A-B

    4/21

    5804 M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

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

    found that can meet them: the requirements lie ina hole in property space. Then the way forward is

    to identify and separate the conflicting require-ments, seeking optimal material solutions for each,and then combine them in ways that retain the

    desirable attributes of both. The best choice is the

    one that ranks most highly when measured by the

    performance metrics that motivate the design: min-imizing mass or cost, or maximizing some aspect

    of performance (the criteria of excellence). Thealternative combinations are examined and

    assessed, using the criteria of excellence to rank

    them. The output is a specification of a hybrid interms of its component materials and geometry.

    Consider as an example the design of a hybrid

    material for long-span power cables. The objec-tives are to minimize the electrical resistance, but

    at the same time to maximize the strength sincethis allows a greater span. In multi-objective opti-

    mization, of which this problem is an example, itis conventional to express each objective such thata minimum is sought; we thus seek materials with

    the lowest values of resistivity, R, and reciprocal

    of tensile strength, 1/sts. Fig. 5 shows the result:materials that best meet the design requirements lie

    near the bottom left. Those with the lowest resist-

    ancecopper, aluminum, and some of theiralloysare not very strong, and the materials thatare strongestdrawn carbon and low-alloy steeldo not conduct very well. Now consider a cable

    made by interweaving strands of copper and steelsuch that each occupies half the cross-section.

    Assuming the steel carries no current and the cop-per no load (the most pessimistic scenario) the per-

    formance of the cable will lie at the point shownon the figureit has twice the resistivity of thecopper and half the strength of the steel. It occupies

    a part of property space that was previously empty,

    offering performance that was not previously poss-ible.

    But while some conflicting requirements can be

    met in this way, others need a more inventiveapproach.1 So the question arises: are there general

    ways in which material hybridization can be

    explored systematically? It is unrealistic to supposethat one method and one tool can solve all such

    problems. Instead we examine examples of hybrid

    design and attempt to extract principles that could

    help tackle other, as yet unformulated problems ofthis class.

    3. A + B: selecting components for composites

    Aircraft engineers, automobile makers, and

    designers of sports equipment all have one thing in

    common: they want materials that are stiff, strong,tough and light. The single-material choices that

    best achieve this are the light alloys: alloys basedon magnesium, aluminum and titanium. Much

    research aims at improving their properties. Butthey are not all that lightpolymers have muchlower densities. Nor are they all that stiffcer-amics are much stiffer and, especially in the form

    of small particles or thin fibers, much stronger.These facts are exploited in the subset of hybrids

    that we usually refer to as particulate and fibrous composites.

    Any two materials can, in principle, be com-

    bined to make a composite, and they can be mixed

    in many geometries (Fig. 6). In this section, we

    1 An interesting example is that of flexible ferromagnets.Monolithic ferromagnetic materials are stiff, metallic or cer-

    amic, solids. Elastomeric ferromagnetic hybrids offer several

    properties that these monolithic solids do not. The hybrids are

    made by mixing up to 30% of sub-micron iron particles into an

    elastomer resin before polymerising it. The result is a compliant

    ferromagnetic material that has the property that it is mag-

    netostrictive, and that its stiffness increases when placed in the

    magnetic field because the magnetic dipoles that are induced inthe particles attract one another. The material has a fast (1 ms)

    response time, making it suitable for vibration damping.

  • 8/8/2019 Hybrids A-B

    5/21

    5805M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    Fig. 5. Designing a hybridhere, one with high strength and high electrical conductivity. The figure shows the resistivity andreciprocal 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].

    Fig. 6. Schematic of hybrids of the composite type: unidirec-

    tional fibrous, laminated fiber, chopped fiber and particulatecomposites. Bounds and limits, described in the text, bracket

    the properties of all of these.

    restrict the discussion to fully dense, strongly

    bonded, composites such that there is no tendency

    for the components to separate at their interfaceswhen 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 scaleone which is large

    compared to that of the componentsa compositebehaves like a homogeneous solid with its own set

    of thermo-mechanical properties. Calculating theseprecisely can be done, but it is difficult. It is mucheasier to bracket them by bounds or limits: upperand lower values between which the properties lie

    [47]. The term bound will be used to describea rigorous boundary, one which the value of theproperty cannotsubject to certain assumptionsexceed or fall below. It is not always possible to

    derive bounds; then the best that can be done is toderive limits outside which it is unlikely that thevalue of the property will lie. The important point

  • 8/8/2019 Hybrids A-B

    6/21

    5806 M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    is that the bounds or limits bracket the properties

    of all arrangements of matrix and reinforcement

    shown in Fig. 6; by using them we avoid the need

    to model individual geometries.

    3.1. Density

    When a volume fraction f of a reinforcement r(density rr) is mixed with a volume fraction (1

    f) of a matrix m (density rm) to form a compositewith no residual porosity, the composite density is

    given exactly by a rule of mixtures (an arithmetic

    mean, weighted by volume fraction)

    r frr (1f)rm. (1)

    The geometry or shape of the reinforcement does

    not matter except in determining the maximumpacking-fraction of reinforcement and thus the

    upper limit for f.

    3.2. Modulus

    The modulus of a composite is bracketed by the

    well-known Voigt and Reuss bounds. The upper

    bound, Eu, is obtained by postulating that on load-ing the two components suffer the same strain; the

    stress is then the volume-average of the localstresses and the composite modulus follows a ruleof mixtures:

    Eu f Er (1f)Em. (2)

    Here Er is the Youngs modulus of the reinforce-ment and Em that of the matrix. The lower bound,

    El, is found by postulating instead that the twocomponents carry the same stress; the strain is the

    volume-average of the local strains and the com-

    posite modulus is

    E1 EmEr

    f Em (1f)Er(3)

    More precise bounds are possible [8,9], but thesimple ones are adequate to illustrate the method.

    3.3. Hybrid design for stiffness at minimumweight

    We need a criterion of excellence to assess the

    merit of any given hybrid. Here our criterion is

    Table 1

    Criteria of excellence for minimum weight design

    Mode of loading and Stiffness at minimum weight

    geometry

    Tensile loading of ties E/rBending of beams E1/ 2/rBending of plates E1/ 3/r

    stiffness per unit mass, measured by the indices

    listed in the table (for derivations see Ref. [2]). If

    a possible hybrid has a value of any one of thesethat exceed those of the light alloys, it achievesour goal.

    Consider, as an illustration of the method, thedesign of a composite for a light, stiff beam of

    fixed section-shape, to be loaded in bending. Theefficiency is measured by the index E1 /2/r shownin Table 1. Imagine, as an example, that the beam

    is at present made of an aluminum alloy. Berylliumis both lighter and stiffer than aluminum; ceramics

    are stiffer, but not all are lighter. What can these

    hybrids offer?Fig. 7 is a small part of the Er property chart.

    Fig. 7. Part of the Er property chart, showing aluminiumalloys, beryllium and alumina (Al203). Bounds for the moduli

    of hybrids made by mixing them are shown. The diagonal con-

    tours plot the criterion of excellence, E1/ 2/r.

  • 8/8/2019 Hybrids A-B

    7/21

    5807M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    Three groups of materials are shown: aluminum

    and its alloys, alumina (Al2O3) and beryllium (Be).

    Composites made by mixing them have densities

    given exactly by Eq. (1) and moduli that are brack-eted by the bounds of Eqs. (2) and (3). Both of

    these moduli depend on volume fraction ofreinforcement, and through this, on density. Upperand lower bounds for the modulusdensityrelationship can thus be plotted onto the Er chartusing volume fraction f as a parameter, as shown

    in Fig. 7. Any composite made by combiningaluminum with alumina will have a modulus con-

    tained in the envelope for AlAl2O3; the same forAlBe. Fibrous reinforcement gives a longitudinal

    modulus (parallel to the fibers) near the upperbound; particulate reinforcement or transverselyloaded fibers give moduli near the lower one.

    Superimposed on Fig. 7 is a grid showing the

    criterion of excellence E1/ 2/r. The bound-envelopefor Alberyllium composites extends almost nor-mal to the grid, while that for AlAl2O3 lies at ashallow angle to it. Beryllium fibers improve per-formance (as measured by E1 /2/r) roughly fourtimes as much as alumina fibers do, for the samevolume fraction. The difference for particulate

    reinforcement is even more dramatic. The lower

    bound for AlBe lies normal to the contours: 30%of particulate beryllium increases E1/2/r by a fac-tor of 1.5. The lower bound for AlAl2O3 is,initially, parallel to the E1 /2/r grid: 30% of par-ticulate Al2O3 gives almost no gain. The underly-ing reason is clear: both beryllium and Al2O3increases the modulus, but only beryllium

    decreases the density; the criterion of excellence ismore sensitive to density than to modulus.

    In Fig. 8 we return to the big picture. It shows

    the moduli and densities of metals and polymers,

    and, encircled by a broken ellipse, those of highperformance carbon, aramid, PE and glass fibers.The construction illustrated in Fig. 7 leads to famil-ies of polymermatrix composites that lie in theshaded ellipse with that name, and to families ofmetalmatrix composites that lie in the ellipseabove it. Both ellipses occupy areas of property

    space that were previously unoccupied by bulk

    materials, and it is an important one, enabling thedesign of new lightweight mechanical structures.

    Similar methods can be used to select materials

    optimum strength, and for tailored values of ther-

    mal conductivity, expansion coefficient and spe-cific heat [7]. The properties of specific composites

    can, of course, be computed in conventional ways.The advantage of this graphical approach is the

    breadth and freedom of conceptual thinking that itallows and the ease of comparison of possible newhybrids with the population of existing materials.

    3.4. Percolation: properties that switch on and

    off

    Fig. 9, a chart of electrical resistivity against

    elastic stiffness (here measured by Youngs

    modulus), has an enormous hole. Materials thatconduct well are stiff; those that are flexible areinsulators. Consider designing materials to fill thehole; to be more specific, consider designing onethat has low modulus, can be molded like a poly-

    mer, and is a good electrical conductor. Suchmaterials find application in anti-static clothing andmats, as pressure sensing elements, even as solder-

    less connections.Metals, carbon and some carbides and intermet-

    allics are good conductors, but they are stiff andcannot be molded. Thermoplastic and thermoset-

    ting elastomers can be molded but do not conduct.How are they to be combined? Metal coating ofpolymers is workable if the product is to be used in

    a protected environment, but the coating is easily

    damaged. If a robust, flexible, product is needed,bulk rather than surface conduction is essential.

    This can be achieved by mixing conducting par-

    ticles into the polymer.To understand how to optimize this we need the

    concept of percolation. Percolation problems are

    easy to define, but not easy to solve. Research

    since 1960 has provided approximate solutions tomost of the percolation problems associated with

    the design of hybrids (see Ref. [10] for a review).Think of mixing conducting and insulating spheres

    of the same size to give a large array. If there are

    too few conducting spheres for them to touch, thearray is obviously an insulator. If each conductingsphere contacts just one other, there is still no con-

    necting path. If, on average, each touched two,there is still no path. Adding more spheres gives

    larger clusters, but they can be large yet still dis-

  • 8/8/2019 Hybrids A-B

    8/21

    5808 M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    Fig. 8. Youngs modulus and density for 1850 polymers, metals and fibers (broken ellipse). Combining these to create polymer andmetal 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].

  • 8/8/2019 Hybrids A-B

    9/21

    5809M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    crete. The array first becomes a conductor when asingle trail of contacts links one surface to the

    other, that is, when the fraction p of conducting

    spheres reaches the percolation threshold, pc. Forsimple cubic packing pc = 0.248, for close packing

    pc = 0.180. For a random array it is somewhere inbetweenapproximately 0.2.2

    Make the spheres smaller and the transition issmeared out. The percolation threshold is still 0.2,

    but the first connecting path is now thin andextremely deviousit is the only one, out of thevast number of almost complete paths, that actually

    connects. Increase the volume fraction and the

    number of conducting paths increases initially as

    (p

    pc)

    2

    , then linearly, reverting to a rule of mix-tures [13]. If the particles are very small, as much

    as 40% may be needed to give good conduction.But a loading of 40% seriously degrades the mold-

    ability and compliance of the polymer.

    Shape gives a way out. If the spheres arereplaced by fibers, they touch more easily and thepercolation threshold falls. If their aspect ratio is

    f = L/d (where L is the fiber length, d thediameter) thento an adequate approximationempirically, the percolation threshold falls from fcto roughly fc/f

    1/2 [1418]. Fig. 9 shows the area

    of the property chart where these hybrids lie. Withsufficient aspect ratio the percolation thresholdfalls to a few percentage points.

    The concept of percolation is a necessary tool

    in designing hybrids. Electrical conductivity worksthat way; so too does the passage of liquids

    through foams or porous mediano connectedpaths, and no fluid flows; just one (out of a millionpossibilities) and there is a leak. Add a few more

    connections and there is a flood. Percolation ideasare particularly important in understanding the

    transport properties of hybrids: properties thatdetermine the flow of electricity or heat, of fluid,or of flow by diffusion, specially when the differ-ences in properties of the components are extreme.

    Most polymers differ from metals in their electricalconductivity by a factor of about 1020. The dif-

    2 These results are for infinite, or at least very large, arrays.Experiments [11,12] generally give values in the range 0.190.22, with some variability because of the finite size of thesamples.

    fusion of water through solids differs from the flowrate of water through channels by a similar factor.

    It is then that single connections really matter.

    Percolation influences mechanical propertiestoo, particularly when mechanical connection is

    important, as in arrays of loose powders or fibers.If there are no bonds between the particles or fib-ers, the array has no tensile stiffness or strength.If each particle is bonded to another, or to several

    forming discrete clusters, there is still no stiffness

    or strength. These only appear when there are con-nected paths running completely through the array.

    The plasticity of 2-phase hybrids, too, can be

    viewed as a percolation problem. Plasticity may

    start in one phase at a low stress, allowing patchesof slip to form, but full plasticity requires that the

    slip patches link to give connected paths throughthe entire cross-section of the sample. Mechanical

    and electrical percolation can be combined,

    exploiting scale. A latex reinforced with a smallvolume fraction of cellulose fibers coated withpolypyrrole to make them conducting, gives a

    material with a shear modulus two orders of mag-nitude higher than the rule of mixture predicts

    combined with good conduction because of thehigh aspect ratio of the fibers [19]. Here we are

    escaping the continuum boundsa topic we returnto later.

    3.5. Creating anisotropy

    The elastic and plastic properties of bulk mono-

    lithic solids are frequently anisotropic, but weakly

    sothe properties do not depend strongly on direc-tion. Hybridization gives a way of creating and

    controlling anisotropy, and it can be large. We

    have already seen an example in Fig. 7, which

    shows the upper and lower bounds for the moduliof composites. The longitudinal properties of

    unidirectional long-fiber composites lie near theupper bound, the transverse properties near the

    lower one. The vertical width of the band in

    between them measures the anisotropy.Consider a second example, that of creating

    hybrids with anisotropic thermal conductivity. A

    saucepan made from a single material when heatedon an open flame, develops hot spots that canlocally burn its contents. That is because the sauce-

  • 8/8/2019 Hybrids A-B

    10/21

    5810 M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    pan is thin, and heat is transmitted through the

    thickness more quickly than it can be spread trans-

    versely to bring the entire pan surface to a uniform

    temperature. The metals of which saucepans areusually madecast iron, or aluminum or copperhave a isotropic thermal conductivities whereaswhat we clearly want is a thermal conductivity thatis higher in the transverse direction than in thethrough-thickness direction. A bi-layer (or multi-

    layer) hybrid can achieve this.

    Heat transmitted transversely in a bi-layer sheethas two parallel paths; the total heat transmitted is

    a sum of that in each of the paths. If it is made of

    a layer of material 1 with thickness t1 and conduc-

    tivity l1, bonded to a layer of material 2 with thick-ness t2 and conductivity l2, the conductivity paral-lel to the layers is

    lII fl1 (1f)l2 (4)

    (a the rule of mixtures), where f= t1/ (t1 + t2). Per-pendicular to the layers the conductivity is

    1

    l

    f

    l1

    (1f)

    l2(5)

    (the harmonic mean). Fig. 10 shows l and l plot-ted against f for a bi-layer of copper (l=390W/m K) and cast iron (l = 30 W/m K). For singlematerials the two are equal; layering them gives

    Fig. 10. Creating anisotropy. The thermal conductivities of

    copper and cast iron are isotropic. Anisotropy is created by

    combining them as a bi-layer.

    trajectories for l and l that separate. Themaximum separation occurs broadly where each

    occupies about half the thickness, where the ratio

    of the conductivities (the anisotropy ratio) is 3.8.Mechanical anisotropy is most easily created

    and managed through shape. This is the topic ofthe next section.

    4. Shape: structures, sandwiches and

    segmented assemblies

    The shape and configuration of components Aand B of a hybrid play a key role in determiningits properties. Shape can be used to enhance or

    diminish stiffness and strength, to impart damagetolerance, andas we have already seentomanipulate the percolation limit.

    4.1. Shape efficiency and shape factors

    Beams with hollow-box or I-sections are stiffer

    and stronger in bending than solid sections of the

    same cross-sectional area; so, too, are panels withribs or waffle stiffeners, or those with an expandedcore to create a sandwich (Fig. 11). These are

    Fig. 11. Making high-efficiency structures. Shape gives thesections a greater flexural stiffness and strength per unit massthan the solid section from which they are made.

  • 8/8/2019 Hybrids A-B

    11/21

    5811M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    examples of the use of shape to increase structural

    efficiency. To characterize this we need a metrica way of measuring the structural efficiency of a

    section shape, independent of the material of whichit is made. An obvious one is that given by the

    ratio j of the stiffness or strength of the shapedsection to that of a neutral reference shape. Fora beam we take the reference shape to be that ofa solid square section with the same cross-sectional

    area, and thus the same mass per unit length, as

    the shaped section. For a panel, we take it to be asolid, plain panel with the same mass per unit area

    as the shaped section, as shown in the figure. Wecall j the shape factor [2,20] and define that for

    stiffness je as

    je Flexural stiffness of shaped section

    Flexural stiffness of reference section(6)

    and that for strength jf as

    jf Flexural strength of shaped section

    Flexural strength of reference section(7)

    Shape can be used to reduce flexural stiffnessand strength as well as increase them. Springs, sus-pensions, flexible cables and other structures that

    must flex yet have high tensile strength, use shapeto give a low bending stiffness. Low shapeefficiency is achieved by forming the material intostrands or leaves, as suggested Fig. 12. Values of

    j for the stiffness of structural sections can be as

    Fig. 12. Making low-efficiency structures. Shape gives thesections a lower flexural stiffness and strength per unit massthan the solid section from which they are made.

    high as 50; for multi-strand or multi-leaf structures

    as low as 0.01.

    Note the origins of efficiency. The flanges of the

    I-section or the faces of the sandwich panels lie farfrom the neutral axis; they stretch when the section

    is loaded in bending. Subdivision, as in Fig. 11,lowers efficiency because the slender strands orleaves bendeasily, but do not stretch when the sec-tion is bent: an n-strand cable is less stiff by a fac-

    tor of 3/n than the solid reference section; an n-leaf panel by a factor 1/n2. There is an underlyingprinciple here: stretch dominated structures have

    high structural efficiency; bending dominated

    structures have low.

    4.2. Shape on a micro-scale

    The sections of Fig. 11 achieve efficiencythrough their macroscopic shape. Structural

    efficiency can be manipulated in another way:through shape on a small scale; microscopic ormicro-structural shape. Wood is an example. Thesolid component of wood (a composite of cellu-lose, lignin and other polymers) is shaped into pris-

    matic cells, each cell like the hollow tube of Fig.11. The effect is to disperse the solid component

    further from the axis of bending or twisting of thebranch or trunk of the tree, increasing its flexuralstiffness and strength. This is not the only possi-

    bility; low efficiency structures give materials withlow stiffness and strength, desirable in cushioningand 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-structurallevel, by bending no matter how it is loaded

    remotely is much less stiff than one that respondsby stretching. Material made by foaming have

    structures that respond in this way.

    Fig. 13 shows an idealized cell of a low-densityfoam. It consists of solid cell walls or edges sur-

    rounding a void space, each cell with an overall

    space-filling shape. Cellular solids are charac-terized by their relative density, which for the

    structure shown here (with t ) is

  • 8/8/2019 Hybrids A-B

    12/21

    5812 M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    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

    2 (8)

    where r is the density of the foam, rs is the den-sity of the solid of which it is made, is the cellsize, and t is the thickness of the cell edges. A

    remote compressive stress s exerts a force F s2 on the cell edges, causing them to bend asshown in the figure, and leading to a bending

    deflection d. For the open-celled structure shownin the figure, the bending deflection is given by

    dFL3

    EsI(9)

    where Es is the modulus of the solid of which thefoam is made and I= t4 /12 is the second momentof area of the cell edge of square cross-section, t t. The compressive strain suffered by the cell asa whole is then e = 2d/. Assembling these resultsgives the modulus E = s/e of the foam as

    E

    Es rrs

    2

    (10)

    (bending dominated behavior)

    Since E = Es when r = rs, we expect the constantof proportionality to be close to unitya specu-lation confirmed both by experiment and bynumerical simulation.

    A similar approach can be used to model non-

    linear properties such as strength. The cell walls

    yield when the force exerted on them exceeds their

    fully plastic moment

    Mf sst

    3

    4(11)

    where ssis the yield strength of the solid of whichthe foam is made. This moment is related to the

    remote stress by M

    FL

    sL3

    . Assembling theseresults gives the failure strength s

    s

    ssrrs

    3/2

    (12)

    (bending dominated behavior)

    This behavior is not confined to open-cell foamswith the structure shown in Fig. 14. Most closed-

    cell foams also follow these scaling laws. At firstsight an unexpected result because the cell faces

    must carry membrane stresses when the foam isloaded, and these should lead to a linear depen-

    dence of both stiffness and strength on relative

    density. The explanation lies in the fact that the

    cell faces are very thin; they buckle or rupture at

    stresses so low that their contribution to stiffness

    and strength is small, leaving the cell edges to

    carry most of the load (for further details, see

    Ref. [21]).

  • 8/8/2019 Hybrids A-B

    13/21

    5813M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

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

    4.4. Ultra-light, high stiffness hybrids

    If conventional foams have low stiffness

    because other configuration of the cell edgesallows them to bend, might it not be possible to

    devise other configurations in which the cell edgeswere made to stretch instead? This thinking leadsto the idea of micro-truss structures [22,23]. To

    understand these we need the Maxwell stability cri-terion.

    The condition that a pin-jointed frame of b strutsand j frictionless joints to be both statically andkinematically determined i.e. just rigid [24,25], in

    2-dimensions, is:

    M b2j 3 0 (13)

    and in 3-dimensions is

    M b3j 6 0 (14)

    If M 0, the frame is a mechanism. It has no

    stiffness or strength, but will collapse if loaded. If

    its joints are locked (instead of pin-jointed) the barsof the frame bend when the structure is loaded. If,

    instead, M 0 the frame ceases to be a mech-anism; its members carry tension or compression

    when the frame is loaded, and it becomes a stretch-

    dominated structure.These criteria give a basis for the design of

    efficient micro-truss structures. For the cellularstructure of Fig. 13 M 0, and it is bending domi-nated. For the structure shown in Fig. 14, however,

    M 0 and it behaves as an almost isotropic,

    stretch-dominated structure. On average one thirdof its bars carry tension when the structure is

    loaded in simple tension, regardless of its direc-

    tion. Thus

    E

    Es

    1

    3rrs (15)

    for isotropic stretch-dominated behavior and

    s

    ss

    1

    3r

    rs (16)for isotropic stretch-dominated behavior.3

    Prismatic structures do even better, provided

    they are loaded parallel to the prism axis. Fig. 15shows four such structures which are common innature. It is helpful to think, as before, of the

    expansion of a solid bar, shown in the center, togive the structures, with no change of massherethe solid black represents solid material, the dottedareas represent low density foam and the open

    areas represent space. The expansion has moved

    material away from the axis of bending, increasingthe second moment of area about any axis con-

    tained in the plane of the cross-section, and

    increasing efficiency in the sense of Eqs. (6) and(7). For these structures the axial modulus and

    strength (measured parallel to the prism axis) fol-

    3 This assumes that failure occurs by the axial yielding of a

    bar. If the bars are slender or have low modulus they may fail

    instead by buckling, giving a lower strength.

  • 8/8/2019 Hybrids A-B

    14/21

    5814 M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    Fig. 15. Four extensive micro-structured materials which are

    mechanically efficient: (a) prismatic cells, (b) fibers embeddedin a foamed matrix, (c) concentric cylindrical shells with foam

    between, and (d) parallel plates separated by foamed spacers.

    lows Eqs. (15) and (16), but with a constant ofproportionality not of 1/3 but of unity:

    E

    Es

    r

    rs (17)for prismatic stretch-dominated behavior and

    s

    ssrrs (18)

    for prismatic stretch-dominated behavior.3

    Loaded transversely, however, they are bendingdominated and follow power laws like those of

    Eqs. (10) and (12); they are thus exceedingly

    anisotropic.

    These results are summarized in Fig. 16, inwhich the modulus E is plotted against the den-

    sity r. Stretch dominated, prismatic microstruc-tures like wood give moduli that scale as r/rs(slope 1); bending dominated, cellular, microstruc-tures like that of polystyrene foam give moduli thatscale as (r/rs)

    2 (slope 2). Given that the density

    can be varied through a wide range, this allows

    great scope for material design. Note how the useof microscopic shape has expanded the occupied

    area of Er space.

    4.5. Ultimate efficiency: the sandwich

    A sandwich panel epitomizes the concept of ahybrid. It combines two materials in a specifiedgeometry and scaleone forming the faces, theother the coreto give a structure of high stiffnessand strength at low weight (Figs. 11 and 17). Theseparation 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, producingan efficient structure for resisting bending andbuckling loads. Sandwiches are found where

    weight-saving is critical: in aircraft, trains, trucksand cars, in portable structures, and in sports

    equipment. Nature, too, makes use of sandwichdesigns: sections through the human skull, thewing of a bird and the stalk and leaves of many

    plants show a low-density foam-like core separat-

    ing solid faces. The faces carry most of the load,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 coreoccupies most of the volume, it must be light, and

    stiff and strong enough to carry the shear stressesnecessary to make the whole panel behave as a

    load bearing unit, but if the core is much thickerthan the faces these stresses are small.

    So far we have spoken of the sandwich as a

    structure: faces of material A supported on a core

    of material B, each with its own density and modu-lus. But we can also think of it as a material with

    its own set of properties, and this is useful because

    it allows comparison with more conventionalmaterials. To do so we must analyze sandwich per-

    formance [21,2629]. We shall use, as a criterionof excellence, the bending stiffness per unit width,Sw divided by the mass per unit area, ma.

    The bending stiffness of the panel per unit

    width, Sw, is given by

    Sw (EI)sand (19)

    112

    (d3 c3)Ef1

    1 BEftc

    2GcL2

    where the dimensions, d, c, t and L are identified

  • 8/8/2019 Hybrids A-B

    15/21

  • 8/8/2019 Hybrids A-B

    16/21

    5816 M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    This, of course, is an idealization. The core

    always shears somewhat, and it does have some

    mass; a more precise analysis copes with this [29].

    It leads to the performance shown in Fig. 18, whichhas been constructed in the same way as Fig. 7. It

    shows the modulus and density of A + B hybrids.The shaded band is bounded by the upper andlower bounds of Eqs. (1)(3), describing particu-late and fibrous composites. The flexural perform-ance of the sandwich is shown as a dashed line; at

    its mid-section it lies a factor of 3 above the upperbound rule of mixtures of Eq. (2). The criterion of

    excellence for minimum weight design with pre-

    scribed bending stiffness, listed in Table 1, is that

    of maximizing E

    1/3

    /r. Contours of this criterionare plotted as diagonal lines on the figure, increas-ing towards the top left. The sandwich out-per-forms all alternative hybrids of A + B.

    4.6. Subdivision as a design variable

    We have already seen how subdivision can

    reduce stiffness (Fig. 12). Here we examineanother way in which it can be used: to impart

    damage tolerance. A glass window, hit by a pro-

    Fig. 18. Sandwich panels (broken line) extend the range of

    flexural modulus per unit mass into areas not occupied bymonolithic materials. Their flexural moduli lie above that pre-dicted by a rule of mixtures by a factor of approximately 3.

    jectile, will shatter. One made of small glass

    bricks, laid as bricks usually are, will lose a brick

    or two but not shatter totally; it is damage-tolerant.

    By sub-dividing and separating the material, acrack in one segment does not penetrate into its

    neighbors, allowing local but not global failure.That is the principle of topological toughening.Builders in stone and brick have exploited the ideafor thousands of years: both materials are almost

    as brittle as glass, but buildings made of themeven those made without cement (dry-stonebuilding)survive ground movement, even earth-quakes, through their ability to deform with some

    local failure, but without total collapse.

    Taking the simplest view, two things are neces-sary for topological damage tolerance: discreteness

    of the structural units, and an interlocking of theunits in such a way that the array as a whole can

    carry load. Brick-like arrangements (Fig. 19a) are

    damage tolerant in compression and shear, but dis-integrate under tension. Strand and layer-like struc-

    tures (shown earlier as Fig. 12) are damage-tolerant

    in tension because if one strand fails the crack doesnot penetrate its neighborsthe principle of multi-strand ropes and cables. The jigsaw puzzle con-figuration (Fig. 19b) carries in-plane tension, com-

    pression, and shear, but at the cost of introducinga 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. [3032] explore aparticular 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 isdone 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-

    tion.The damage tolerance can be understood in the

    following way. We suppose that the units of the

  • 8/8/2019 Hybrids A-B

    17/21

    5817M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    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-planeloads ( sxx, syy, sxy). (c) The units suggested by Dyskin et al., when assembled into a continuous layer and clamped within arigid boundary around its edge, can carry out-of-plane loads and bending moments ( sxz, syz, Mxz, Myz) (Fig. 15(c) derived fromRef. [30]).

    structure are all identical, each with a volume Vs,

    and that they are assembled into a body of volume

    Vt; there are therefore n = Vt/Vs segments. Wedescribe the probability of failure of a segment

    under a uniform tensile stress sby a Weibull prob-ability function:

    Pf(V,s) 1exp VsmVos

    m

    o (24)

    where m, Vo and so are constants [33,34]. If thebody were made of a single monolithic piece ofthe brittle solid, this equation, with V= Vt, woulddescribe the failure probability. To calculate the

    design stress st we set an acceptable value for P,which we call P (say 106, meaning that it is

    acceptable if one in a million fail) and invert theequation to give

    st so VoVt

    ln(1P)1/m (25)Now consider the segmented body. A remote

    stress, if sufficiently large, causes some segmentsto fail. We refer to the fraction that has failed as

    the damage, D. If loaded such that each segmentcarries a uniform stress s, the damage is simplyPf(Vs,s). If some segments fail, the body as awhole remains intact; global failure requires that afraction D , the critical damage, (say, 10%) must

    fail. Inverting Eq. (24) with V= Vs gives the globalfailure stress ss of the segmented body:

    ss so

    Vo

    Vsln(1D)

    1/m

    (26)

    Thus segmentation increases the allowable design

    stress from st to s

    s , factor of

    ssst

    nln(1D)ln(1P)

    1/mnDP

    1/m (27)(expanding the logarithm as a series and retaining

    the first terman acceptable approximation forsmall P and D). Both n and D/P are con-

    siderably greater than 1, so the equation suggests

  • 8/8/2019 Hybrids A-B

    18/21

  • 8/8/2019 Hybrids A-B

    19/21

    5819M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    sm(1f) of the continuum approximation. Fine-grained materials, particularly those with grains of

    a few nanometers, exhibit strengths that exceed

    those of large-grained bulk samples. Scale, then,can have a profound influence on mechanicalproperties.

    Transport properties, too, can be scale depen-dent. The mean-free paths of electrons, phononsand of diffusing atoms and molecules are limited

    by the scale of the microstructure when this is

    smalla fact exploited in micro-cellular foams togive exceptional thermal insulation. Convection,

    acoustic absorption and light scattering, too, are

    directly linked to aspects of structural scale.

    6. Summary and conclusions

    The properties of engineering materials can be

    thought of as defining the axes of a multi-dimen-sional space with each property as a dimension.

    Sections through this space can be mapped. These

    maps reveal that some areas of property-space areoccupied, others are emptythere are holes. Theholes can sometimes be filled by making hybrids:combinations of two (or more) materials, or of

    material in space, in chosen configuration andscale. Here we have surveyed conceptual tools forsuggesting and assessing hybrids to fill specifiedneeds. A useful starting point is the concept of a

    hybrid as A + B + shape + scale. Successfulhybrids, as a rule, exploit the first three of these;with micron and nanometer scale fabrication tech-

    nologies now a reality, it becomes possible to addthe last, opening up wider horizons.

    Continuum bounding methods give tools for

    scanning the possibilities offered by a set of

    hybrids, provided the scale is such that the con-tinuum approximation applies. Shapethe way Aand B are configuredcan extend the populatedareas of property-space in ways that complement

    efforts to create new monolithic materials. Dis-

    criminating choice of shape can enhance or dimin-ish physical, mechanical, thermal and electrical

    properties. Scale introduces a new variable. In

    hybrids with structural units that are sub-micron, anew length scale (basically that of the atom) makes

    itself evident. Here the bounds break down, and

    continuum methods must be replaced by statistical

    or dislocation mechanics.

    Much, in a short paper such as this, has been

    ignored. Fabricating successful hybrids can be dif-ficult and expensive (but so, too, is the alternativeof seeking to develop a new monolithic material).

    Part of the difficulty stems from the multitude ofpossible choices: choice of materials, choice of

    process to combine them, and choice of the internal

    geometry and topology of the constitutive

    materials. Part derives from the need to make these

    choices in such a way as to optimally meet a set

    of design requirements. The hybrid must be both

    feasible and optimal.

    To explore this design space efficiently optimiz-ation tools are needed. The starting point is a sim-

    ple screening of optionsthe obvious way for-ward when the selection space is discrete. Thus it

    is possible to create a database of, say, composites

    by computing the properties of virtual materials

    with 5%, 10%, 15% of reinforcing fibers anddiscrete choices of lay-up; selection proceeds by

    rejecting all the combinations that fail to meet the

    design requirements. This method becomes

    impractical when many material choices and lay-

    ups are allowed. When the optimization variablesare continuous (such as the volume fraction of

    reinforcement in a composite with pre-chosen

    constituents), linear programming or steepest

    descent methods can be efficient. When the poten-tial landscape is very rugged, simulated annealingcan offer a practical alternative. When the variables

    are both discrete (such as the choice of materials)

    and continuous (the thicknesses of face sheets and

    core of a sandwich, for instance), a genetic algor-

    ithm, which allows an unevenly populated space to

    be explored, can be efficient. Refs. [3741]examples of their use in hybrid design.These tools allow promising candidates to be

    identified. They need the back-up of expert toolsadvising on the compatibility of the constituents,

    the practicality of processes to assemble them into

    a hybrid, and the loss of properties (the knock-down factor relating real and ideal hybrids) fora given combination of the materials, architecture

    and process.

  • 8/8/2019 Hybrids A-B

    20/21

    5820 M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    Acknowledgements

    The ideas, methods and tools described here

    have evolved over the past 15 years. Numerouscolleagues in many countries have (sometimes

    unknowingly) stimulated or contributed to thisevolution. Among these we would particularly like

    to recognize Profs. Mick Brown, Chris Calladine,

    Norman Fleck and David Cebon (all of CambridgeUniversity), Dave Embury (McMaster University,

    Canada), Tony Evans (UCSB), John Hutchinson

    (Harvard University) and Haydn Wadley (UVA),and Dr. Luc Salvo (University of Grenoble).

    References

    [1] Ashby MF. On the engineering properties of materials.

    Acta Metall 1989;37:127393.[2] Ashby MF. Materials selection in mechanical design, 2nd

    ed. Oxford: Butterworth Heinemann, 1999.

    [3] Kromm FX, Quenisset JM, Harry R, Lorriot T. An

    example of multimaterial design. Adv Eng Mater

    2002;4:3714.

    [4] Watt JP, Davies GF, OConnell RJ. Reviews of geophys-ics and space physics 1976;14:541.

    [5] Schoutens JE, Zarate DA. Composites 1986;17:188.

    [6] Chamis CC. In: Engineers guide to composite materials.

    Materials Park, Ohio, USA: American Society of Metals,

    1987:3-83-24.[7] Ashby MF. Criteria for selecting the components of com-

    posites. Acta Mater 1993;41:131335.

    [8] Eshelby JD. Proc R Soc Lond 1957;A241:376.

    [9] Hashin Z, Strikman S. J Appl Phys 1962;33:3125.

    [10] Stauffer D, Aharony A. Introduction to percolation theory,

    2nd ed. London, UK: Taylor and Francis, 1994 revised.

    [11] Fitzpatrick JP, Malt RB, Spaepen F. Phys Lett

    1974;A47:207.

    [12] Brown LM. A simple model of a metal non-metal tran-

    sition. Physics Education 1977;July issue:31820.

    [13] Last BJ, Thouless DJ. Phys Rev Lett 1971;27:1719.

    [14] Nielsen LE. Thermal conductivity of particulate-filledpolymers. J Appl Polym Sci 1973;17:3819.

    [15] Nielsen LE. The thermal and electrical conductivity of

    two-phase systems. Ind Eng Chem Fund 1974;13:17.

    [16] Bigg DM. Conductive polymeric compositions. Polym

    Eng Sci 1977;17:892.

    [17] Bigg DM. Mechanical and conductive properties of metal

    fibre-filled polymer composites. Composites 1979;April

    issue:95100.[18] Yi Y-B, Sastry AM. Analytical approximation of the 2-

    dimensional percolation threshold for fiels of over lappingellipses. Phys Rev 2002;E66 066130-1066130-8.

    [19] Brechet Y, Cavaille JY, Chabert E, Chazeau L, Dendievel

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

    effect of fillerfiller and fillermatrix interactions. Adv

    Eng Mater 2001;3(8):571.

    [20] Ashby MF. On material and shape. Acta Mater

    1991;39:1025.

    [21] Gibson LJ, Ashby MF. Cellular solids, structure and

    properties, 2nd ed. Cambridge, UK: Cambridge University

    Press, 1997.

    [22] Deshpande VS, Ashby MF, Fleck NA. Foam topology:

    bending versus stretching dominated architectures. Acta

    Mater 2001;49:103540.

    [23] Deshpande VS, Fleck NA, Ashby MF. Effective properties

    of the octet-truss lattice material. J Mech Phys Sol

    2001;49:174769.

    [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.[26] Allen HG. Analysis and design of structural sandwich

    panels. Oxford, UK: Pergamon Press, 1969.

    [27] Cordon J. Honeycomb structure. In: Dostal CA, editor.

    Engineered materials handbook. Metals Park, Ohio, USA:

    ASM International; 1990:7218.

    [28] Zenkert D. In: An introduction to sandwich construction.

    Solihull, London, UK: Engineering Advisory Services

    Ltd., Chameleon Press Ltd, 1995, ISBN 0 947817778.

    [29] Ashby MF, Evans AG, Fleck NA, Gibson LJ, Hutchinson

    JW, Wadley HNG. Metal foams, a design guide. Oxford,

    UK: Butterworth Heinemann, 2000.

    [30] Dyskin AV, Estrin Y, Kanel-Belov AJ, Pasternak E.

    Toughening by fragmentation: how topology helps. AdvEng Mater 2001;3:8858.

    [31] Dyskin AV, Estrin Y, Kanel-Belov AJ, Pasternak E. Topo-

    logical interlocking of platonic solids: a way to new

    materials and structures. Phil Mag 2003;83:197203.

    [32] Dyskin AV, Estrin Y, Pasternak E, Kohr HC, Kanel-Belov

    AJ. Fracture resistant structures based on topological inter-

    locking with non-planar contacts. Adv Eng Mater

    2003;5(3):1169.

    [33] Weibull W. J Appl Mech 1951;18:293.

    [34] Davidge RW. Mechanical behavior of ceramics. Cam-

    bridge, UK: Cambridge University Press, 1986.

    [35] Curtin WA. Acta Metall et Mater 1993;41:1369.

    [36] Shaw MC. The fracture mode of ceramic/metal multilay-

    ers: role of the interface. Key Eng Mater 1996;116-

    117:26178.

    [37] Salvo L, Brechet Y, Pechambert P, Bassetti D, Jantsen A.

    Logiciels de selection des composites. Materiaux et Tech-

    nique 1998;5:31.

    [38] Landru D, Brechet Y. New design tools for materials and

    process selection. Materiaux et Techniques 2002;4:6.

    [39] Brechet Y, Bassetti D, Landru D, Salvo L. Challenges in

    material and process selection. Prog Mater Sci

    2001;46:407.

    [40] Brechet Y, Ashby MF, Salvo L. Selection des materiaux

    et des procedes de mis en oeuvre. Switzerland: Les Presses

  • 8/8/2019 Hybrids A-B

    21/21

    5821M.F. Ashby, Y.J.M. Brechet / Acta Materialia 51 (2003) 58015821

    Polytechniques et Universitaires Romandes de Lausanne,

    2002.

    [41] Ashby M, Brechet Y, Cebon D, Salvo L. Materials and

    process selection strategies. To appear in Advanced

    Engineering Materials, 2003.

    [42] CES 4. The Cambridge Material Selector, Granta Design,

    Cambridge. 2003;(www.Grantadesign.com)