You are on page 1of 6

The Linear Elastic Properties of Open-Cell Foams^

W. E. Warren
Mem. ASME

A. M. Kraynik
Sandia National Laboratories, Albuquerque, NIVI 87185

A theoretical model for the linear elastic properties of three-dimensional open-cell foams is developed. We consider a tetrahedral unit cell, which contains four iden- tical half-struts that join at equal angles, to represent the essential microstructural features of a foam. The effective continuum stress is obtained for an individual tetrahedral element arbitrarily oriented with respect to the principal directions of strain. The effective elastic constants for a foam are determined under the assump- tion that all possible orientations of the unit cell are equally probable in a represen- tative volume element. The elastic constants are expressed as functions of com- pliances for bending and stretching of a strut, whose cross section is permitted to vary with distance from the joint, so the effect of strut morphology on effective elastic properties can be determined. Strut bending is the primary distortional mechanism for low-density foams with tetrahedral microstructure. For uniform strut cross section, the effective Young's modulus is proportional to the volume fraction of solid material squared, and the coefficient of proportionality depends upon the specific strut shape. A similar analysis for cellular materials with cubic microstructure indicates that strut extension is the dominant distortional mechanism and that the effective Young's modulus is linear in volume fraction. Our results em- phasize the essential role of microstructure in determining the linear elastic proper- ties of cellular materials and provide a theoretical framework for investigating nonlinear behavior.

Introduction

In part, the prevalence of cellular solids can be attributed to their mechanical properties. Cellular materials, which we will refer to as foams, encompass biomaterials like bone and lung parenchyma as well as polymeric and other synthetic foams. Here, we address the linear elastic properties of open-cell foams and develop a theoretical framework that can be ex- tended to analyze their large-strain response and ultimate failure. This represents an extension to three-dimensional materials of our previous work on two-dimensional foams (Warren and Kraynik, 1987) in which we analyzed the linear elastic proper- ties of honeycombs with nonuniform film thickness. An im- portant conclusion of that analysis had been reported by Gib- son, et al. (1982): Film bending is the primary film-level defor- mation mechanism for small distortions of low-density foams. = The theories of foam mechanics that originate with the work of Gent and Thomas (1959, 1963) assume that the displace- ; ment of junctions within the microstructure is affine; this

'This work was performed at Sandia National Laboratories and supported by I the U.S. Department of Energy under Contract #DE-AC04-76DP00789. Contributed by the Applied Mechanics Division for presentation at the I Winter Annual Meeting, Chicago, IL, November 28 to December 2, 1988, of [ The American Society of Mechanical Engineers. Discussion on this paper should be addressed to the Editorial Department, lASME, United Engineering Center, 345 East 47th Street, New York, N.Y. pOOn, and will be accepted until two months after final publication of the paper ^KLF in the JOURNAL OF APPLIED MECHANICS. Manuscript received by ASME iApplied Mechanics Division, May 27, 1987; final revision, December 7, 1987. Paper No. 88-WA/APM-20.

disregards requirements of mechanical equilibrium at the junctions. In addition, this suppresses bending and only incor- porates effects associated with stretching of the microstruc- ture. The influence of film-level deformation mechanisms on the elastic properties of a two-dimensional foam is well il- lustrated for the case of uniform film thickness: When bend- ing dominates, the effective Young's modulus is propor- tional to the cube of the solid volume fraction <^; when film stretching dominates, Ey varies linearly with <f>. These conclu- sions do not depend upon the orientation of the principal axes of strain relative to the cell structure since the hexagonal sym- metry of a honeycomb insures isotropy in the plane (Lekhnit- skii, 1963). In summary, the honeycomb is a relatively soft, transversely isotropic material in the linear deformation regime. By contrast, a foam with square network structure is highly anisotropic in the plane and the primary film-level deformation mechanism varies with cell orientation. The work of Ko (1965) suggests that the strong relationship between foam network connectivity, deformation mechanisms, and stiffness, which is observed in two dimen- sions, carries over to three-dimensional materials. Ko relates the network structure of an idealized foam to that of uniform close-packed spheres. He considers two open-cell structures, chooses a particular direction of uniaixial strain, and calculates an effective Young's modulus for each structure using beam theory. His results are consistent with the following: When connecting struts are aligned throughout the material and must stretch to accommodate macroscopic deformation, Ey 0. However, when struts are not aligned and bending at the cell level accommodates macroscopic deformation, E, ~ <j>^.
JUNE1988,Vol.55/341

>ns of the ASME

urnal of Applied Mechanics

Menges and Knipschlld (1975, 1982) studied a mechanically loaded, rigid polyurethane foam under a microscope and observed strut bending. Even though their foams are closed- cell, the films are very thin compared to the struts (Plateau borders), so the response of an open-cell foam is expected. They argue that elastic response is determined by bending and stretching effects and provide a simple analysis for which Ey <l>^ when 4> is small. The choice of microstructural ele- ment-four identical struts meeting at a tetrahedral junc- tion-is a noteworthy feature of their model. Gibson and Ashby (1982) accept bending as the essential small-distortion mechanism for three-dimensional foams, and illustrate their point of view with "cubic" microstructural models for which adjacent struts or films do not align. The microstructure of foams reflects thefr method of preparation, which usually involves a continuous liquid phase that eventually solidifies; therefore, surface tension and related interfacial effects often control foam structure. Two elementary features of liquid foam structure that are required to minimize surface energy are well known (Bikerman, 1973): Three films always meet at equal angles of 120 to form a film junction region called a Plateau border, and four Plateau borders always join at the tetrahedral angle of cos ~'(- 1/3) 109.47. In two dimensions, the hexagonal network structure of a honeycomb satisfies the film-angle constraint for a monodisperse foam and represents a logical model for in- vestigating foam mechanics (Warren and Kraynik, 1987). When considering arbUrary homogeneous deformations of this spatially periodic structure, one can exploit the symmetry involved and reduce the. structural element under considera- tion to a unit cell containing three half-films and their junc- tion. The displacement of the film midpoints is affine-the displacement of the junction is not. A crucial feature of affine film-midpoint displacement is compatibility with both film- level deformation mechanisms, bending and stretching. We note that strict affine film-midpoint displacement is a conse- quence of perfectly ordered structure and cannot be expected for foams that are polydisperse or possess other attributes of disorder. In this work, we assume that essential features of mechanical response for three-dimensional open-cell foams are captured by analyzing the deformation of a representative microstructural element under the assumption that the displacement of strut midpoints is affine. By contrast with our previous analysis' for two-dimensional foams, justification of the basic assumptions is more intuitive. The representative ele- ment under consideration consists of four identical half-struts that meet at equal tetrahedral angles. The struts can be iden- tified with Plateau borders in a precursor liquid foam. While the films in a completely open-cell foam are absent, we only require them to be very thin relative to the transverse dimen- sion of the strut so that they do not affect strut deformation. Our choice of microstructural element is consistent with the two topological features of liquid foam structure discussed previously. One should recognize, however, that no cell in a three-dimensional" foam is a simple polyhedron with planar faces and straight edges, because a- planar polygon cannot have all angles equal to the tetrahedral angle. We neglect cur- vature along the strut axis, which derives from this. The relative orientation of adjacent tetrahedral elements that possess a common strut is assumed to be random. The boun- daries of our unit cell consist of four planes that are perpen- dicular to each strut at its midpoint and form a regular tetrahedron. The corners of the tetrahedral volume element correspond to the centers of adjacent foam'cells. It is impor- tant to recognize that the following analysis explicitly accounts for both the connectivity of struts and mechanical equilibrium at the junctions. Strut deformation due to bending and stretching is also incorporated in a natural way.
342/Vol. 55, JUNE 1988

Analysis

We choose a single tetrahedral unit cell with four half-struts of length L to represent the important microstructural features in an open-cell foam and consider the displacement of and the force acting upon each strut midpoint. The strut geometry and coordinate system for this problem are shown in Fig. 1. We assume that homogeneous deformations of this structure are completely determined by the affine displacement of strut midpoints. The strut-midpoint configuration can be represented by three position vectors, bj, b2, bjt whose com- mon origin coincides with the fourth strut midpoint. Small displacements of each midpoint from its equilibrium position are considered. ' Isolating the tetrahedral element with four identical half- struts as a free body in equilibrium, the displacements of the four strut midpoints with respect to the joint itself are con- sidered first. The force F, at the rth midpoint is represented in terms of components parallel to and normal to the strut centerUne in the form. F,= (F,-e;")e,.+ (e,xF,)Xe,., /=l-4 ' (1) where e, is a unit vector parallel to the rth strut and directed away from the joint. The displacement of the rth midpoint with respect to the joint, A,-, is represented by A,=M(F,-e,)e,--l-N(e, xF,) xe,-+L^xe,-, (2) where M is the axial elastic compliance of the strut, N is the transverse or bending compliance of the strut which is as- sumed to be the same in all directions, and ^ is a rigid-body rotation of the entire unit. The three strut-midpoint position vectors b, can be represented by b, = b^ + S (=1.2,3 (3) where bf is the undeformed position vector and 5, = A, A4. For a given horripgeneous deformation, the affine displacements of the b, determine the 8, and this provides three vector equations in the four unknown forces and rigid- body rotation. The remaining two vector relations necessary to determine the unknown forces and rotation are obtained from equilibrium considerations which provide 4

and
4

(4) De,xF,=0. (=1


z

relative positions of the strut midpoints and the F; represent the forcss at a strut midpoint.

Transactions of the ASME

When the forces at the strut midpoints are known, the effec- tive continuum stress for the tetrahedral unit cell shown in Fig. 2 can be determined. Cutting this volume element on the plane x=0 and considering the effective stress that must act upon the exposed face to maintain equilibrium with the forces on the remaining volume gives 12V2L^((T;i + ff;,^j + ffk) = F2-F3.
l2^f6L^(axyi + Oyyi + = - 3(F2 + Fj) - 2F4,

2yM

( N - M ) f , * ,
(AT-M) 1/3-/4].

7M

2D
yM

(5)
(6)

Similarly, cutting the volume element on the plane^ = 0 gives and on the bottom triangular face in the plane z = ~L, 6V3L2(CTj,^i + a^J + (7k)=-F4, (7) where where i, j, k are unit vectors associated with an (x, y, z) coor- dinate system that is fixed with respect to the unit cell. The system of equations (5)-(7) for determining the six indepen- dent components of the stress tensor can be shown to be con- sistent through the equilibrium relations of equation (4). We now consider two specific homogeneous deformations of the unit cell.
Pure Shear. Consider pure shear with respect to an {x', y', z') coordinate system oriented from the (x, y, z) coor-

~D'
a/,'

{ N - M ) f ^ * ,
(11)

yM

[3(17/^+7M) + (N-M){f,

D= 2 S 8 ^ L M N ( 2 M + N ) .
The functions/i (0, (t>), f2(0,<l>, are given by
'
/I(^.

4>, ^ ) , and/4(0,. <^, ^Z-)

<^>) =4V2(2 sin20-sin40) CQS3(#> -(4


COS26 + 7

cos40),

<t>, i/)=(3 + 4 cos20 7 cos40) cos2^

-l-4V2(3 sin30-sin0) sin3<#> sin2i;' -4V2(2 sin20 + sin40) cos30 cos2t/',

dinate system through the Euler angles 0, 6, and i/- (Goldstein, 1950). The transformation taking (x, y, z) into (x' ,y' ,z')is
x' = (cos^ COS0 -COS0 sin0 sin^)x + (cos\l/ sin0-)-cos0 cos<^ sm\l/)y + smi/ sin0 z, y' = - (sin^ cos<t> + cos6 sin0 cos\p)x

f^(6, <l), \p)=(.2 sin20 + 7 sin40) sin^


4-4V2(COS20-COS40)
COS3^

(12)

sin^

-l-6V2(cos0-cos36) sin30 cos\j/,

(8)

f^(6, 0, i/) =7(2 sin20-sin40) sin3^


+ 4V2(7 cos20-t-cos4e) cos30 sin3^ 4-2V2(7 COS6 + 9 cos30) sin3(/> cos3V'. The/,*(0, <j), are obtained from the/- (6, (j), !/) by replacing terms of the form (smp\p, cospi/) by {cospxp, -sinp^), respectively.
Uniaxial Extension. For uniaxial extension in the 2'-direction, the strut-midpoint displacements are given by

-(sin^ sini^ COS0 cos(j> cosil/)y + cosip sin0 z,


z' = sin0 sin0 xsin0 cos<l> y + cosd z.

The unit vectors i', j', k' of (x', y', z') are given in terms of the unit vectors i, j, k of (x, y, z) by i' =(cos^ cos<j)-cosd sin<^ sin0)i + (cosV' sin</)4-cos6 cos0 sini/')j-Fsin^ sin0 k, j' = -{sm\j/ cos(l> + cos6 sin<^ cos0)i -(sin^ sin<^)-cos6 cosc^ cosi/')j + cos\t sin0 k, k' = sin0 sin</> i sin0 cos<t> j + cosd k. The affine displacements associated with this condition of pure shear are 6, = 7(b?.k')j', /=1,2,3 (10) where 7 is the magnitude of the shearing strain. Evaluating the stresses in (J:, z) and transforming tolx' ,y' ,z') provides (9)

6,=e(b?.k')k',

/= 1,2,3

(13)

where e is the magnitude of the uniaxial strain. Evaluating the stresses in (x, y, z) and transforming to(x' ,y',z') provides

e ( N - M )

D
e(iV-M)
y

[{\6N+nM)+M{f,-m,

d ' ' = - - yM { N - M M ' + f , * ] ,

ID

D
2e

[(16N+llM)+M(/i+/2)], (14)

yM

2D

(N-M) [3/3*

''z'z'

^ 5 Z M N + n M ^ ) - M { N - M ) f { \ , M { N - M ) f 2 * ,

2e o,','=M{N~M)f
2e
~d"

M{N-M)f^*,

Iflg. 2 Schematic of the volume element that contains each tetrahedral


Miement shpwing the coordinate system used to determine the effective tiltess'. In the undeformed state, the volume element Is a regular tetrahedron.

where the functions/, and/* are the same as those for simple shear. By contrast with our previous two-dimensional analysis (Warren and Kraynik, 1987) the stress fields associated with the affine strut-midpoint displacements corresponding to pure shear and uniaxial strain are dependent upon the orientation of the microstructural element with respect to the principal directions of strain.
/

IS of thevASIiiE

6urnal of Applied Mechanics

JUNE1988, Vol.55/343

Effective Elastic Constants


At this point, we enlarge the size scale of our analysis and consider a representative volume of cellular material to con- tain tetrahedral elements of all possible orientations with equal probability. Under these conditions, the effective stress of the representative volume element is the volume average of the stress associated with the individual tetrahedral elements, the average being taken over all possible orientations. The volume-averaged stress a is given by
1

g _

(1W+4M)

2V3L(WN^ + 31NM+4M^) ' {N-M)(5N+4M) (10N^ + 31NM+4M^)


and
1 (22)

IS^LM '
We note that the effective Poisson's ratio satisfies - 1 < ly < 1/2, which insures that the material strain-energy function is positive-definite. Also, the bulk modulus Kf is independent of the bending compliance N. For low-density foams, the ben- ding compliance is much greater than the axial compliance M and our results indicate that the effective Young's modulus Ey and shear modulus (if will be dominated by strut bending. We now show how the effective elastic constants depend upon the volume fraction of solid and the strut shape.

"F

?(<#>.

4')dv,

(15)

where q{<j), 6, ^) represents the stress associated with an in- dividual element of particular orientation. It is clear that the elastic properties of this representative volume element must be isotropic since we have averaged over all possible orienta- tions and there can be no preferred direction within the material which exhibits a different elastic response from any other direction. The volume average implied by equation (15) becomes an average over all possible values of the Euler angles and takes the explicit form

Examples for Specific Strut Morphology


Consider the case where the struts have a uniform cross sec- tion of area A and isotropic moment of inertia Aq^, where q is the radius of gyration of the section. The volume fraction of solid material 4> is assumed to be sufficiently small that the ef- fective elastic length of the strut is equal to the total strut length 2L and <l> is given by

ff = f
where

F Jo

Jo Jo "

a[<!>,6,i/) sind d<l) dd d^,

(16)

2ir p T p 2ir

Jo Jo

sine d<l> dd

=
The axial compliance M and bending compliance N are

Considering the stress components given in equations (11) and (14) for simple shear and uniaxial extension, respectively, the only volume averages required are those of the functions /i, fi/f J./s./3./4. and/4, since all other terms in these expres- sions are independent of the Euler angles. One finds /,=9/5 and /^ =/J =/3 =/J =/, =/; = 0. (17) For an isotropic material the components of stress and strain are related in the usual form ffy = + 2*1/6,y, (18) where Xy and ixj are the effective Lame constants. For our con- dition of pure shear, the only nonzero component of strain is = Vi 7, and from equation (18) the only nonzero stress is given by

M=- L EA

and

"

. . . N=- " 3EAq^ '

(24)

where' is the elastic modulus of the strut material. The com- pliances of equation (24) written in terms of the volume frac- tion of solid are 1 r- 1 i6Eq^L<l>^ = 2V3Z,0 and -= (25)

For low-density foams, where </> ; 1, equations (22) and (25) provide

Ey _ 33\f3q^<t>^ E~ 5A

(26)

^y'z' =t^n-

(19)

Considering specific strut cross sections, the parameter q^/A and the effective Young's moduli are given by Circular: 9V/l = l/47r = 0.080, Triangular: E/=0.91 02,

For our condition of uniaxial extension,, the only nonzero component of strain is e,',' = e, and from equation (18) the nonzero components of stress are

Ox'x' "^y'y'
=(X/+2M/)e.

(20) Plateau Border:

Ef/E=1.10<l>\

(27)

Equations (19) and (20) provide a redundant but consistent set of relations for determining the effective Lame constants. The stress expressions of equations (11) and (14), together with equation (17), provide the result X/=- and (11N-I-4M)

(N-M)(5N+4M) 9Q^LMN0M+N)
(21)

60\f3LN{2M+N) '
where nj- is the effective shear modulus. The more usual elastic constants. Young's rnodulus E/. Poisson's ratio vj-, and bulk modulus Ky, are given by
344/Vol. 55, JUNE 1988

E--/'=1.53 (l>\ ^ 6(2V3- T)^ which indicates the effect of strut shape on the stiffness of low-density foams. The Plateau border cross section referred to in equation (27) corresponds to the space between three identical, mutually tangent circles and represents -the typical shape of a Plateau border in liquid foams when thcfilms are very thin; this shape is representative of liquid foams that solidify with films intact. Film rupture prior to solidification would favor reduction in strut surface area with cifcular cross section representing the extreme. Table 1 'sliows the dependence of effective elastic constants on volume fraction

(20V3-11.)

Transactions'of the'ASME

for equilateral triangular strut sections. We observe that for low-density foams the effective elastic constants E/ and vary as 0^, which indicates the dominance of strut bending. On the other hand, the bulk modulus kj- varies as </>, which in- dicates the absence of strut bending. Thus any deformations of the foam that are pure dilatation will suppress the effect of strut bending, while those deformations that involve distor- tion will be dominated by strut bending. We illustrate the effect of nonuniform strut cross section for the case where the transverse strut dimension varies as

Table 1 Effective elastic constants for uniform struts with triangular cross section
Tetrahedral Structure Cubic structure

(1U40)

f ^ <l> (1+V3)

^
.

(lO310t4^)
(l-0)(6+4)

( 1 -
(4 ^

t{i)=to(l+vk/Lr

(28)

5-

3/3

where tg is the value at the joint, | is a coordinate directed along the strut centerline with its origin at the joint, and ij is a parameter. The compliances M and N are given by

3i/3

"f .

(11*4) 30 (U20)

M=
and

2L 3EAoV
(29)

_ = ^ (U//3) E IB

1= *
E 9 E 9

N= LH4-r,) llEAggl
where the strut has area Ag and moment of inertia Aqq I at the joint, ^ = 0. The volume fraction of solid material is given by
J.

V3Z,2^

^0

[(1+ 7 ,)'^2-1].

(30)

1N+3M miMN
E , = -

'

The compliances expressed in terms of the volume fraction - are 1 3y/3EL<l)v^

2N+3M ALM{4N+M) N-M AN+M '

(34)

M
and
=

2[(1+7J)3^2 _ ! ] [ ( ! +
(31)

36EL<I>^(-^) \ Aa /

\ [(H-r,)'^2 + l]
(4 + jj)

and 1

\2LM
Poisson's ratio satisfies - \ < < 1/4, which insures that the material strain-energy function is positive-definite. The results in equation (34) are in sharp contrast with their counterparts in equation (22) for the tetrahedral element. For low-density foams, where N M, the Young's modulus E^ and shear modulus ixj- are now determined by strut extension rather than bendiiig. For cubic structure, struts of uniform cross section, and small 0

Equations (22) and (31) provide the effective Young's modulus for low-density foams 33V3(/)2 (32) (p' + 3) '

i-i-)

where we have introduced the parameter p = t g / t ( L ) ^ las a measure of strut tapering; p is the ratio of thickness at the joint to that at the midpoint. The case p = 1 corresponds to a strut of uniform thickness as considered previously. The effec- tive Young's modulus of equation (32) is a maximum when p = VJ for all cross-sectional shapes and is
(%) \ / r

(j> = 3A/4L'^.

(35)

'EA

4 4 V 3 ^ gl

(33)

which is 4/3 the value for a uniform strut with the same volume fraction of material.

The compliances M and N are again given by equation (24), which provides the following for struts of uniform triangular cross section and (36)

Foam With Cubic Network Structure


In this section, we present results for an open-cell foam presumed to have cubic network structure in which six iden- tical struts meet at right angles. The analysis parallels that for tetrahedral structure. We first isolate an individual cubic ele- ment as a free body in equilibrium and determine the forces at the strut midpoints resulting from their affine displacements. The effective stress associated with an individual element depends strongly upon the orientation of the joint with respect to the principal strain directions. The effective elastic con- stants for a cellular material composed of cubic joints are ob- tained as before by averaging over all possible orientations, and found to be

Table 1 shows the variation of effective elastic constants with volume fraction for cubic structure with struts of triangular cross section. We note the linear dependence of Young's modulus Ef and shear modulus Vf on 0, which is indicative of the dominance of strut extension and the suppression of strut- bending effects. For small 0, Ej-/E 0/6 and Vf 1/4, as ob- tained from the theory of Gent and Thomas (1959) which does not incorporate bending. When the struts taper according to equation (28), the effec- tive Young's modulusis given by E^ 0(p2 + l)^

'

(37)

Journal of Applied Mechanics

J U N E 1 9 8 8 , Vol'. 5 5 / 3 4 5

This result is independent of the strut cross-sectional shape and attains a maximum value of = 0/6 for the uniform section when p = 1.

References
Bikerman, J. J., 1973, Foams^ Springer-Verlag, New York, pp. 33-64. Gent, A. N., and Thomas, A. G., 1959, "The Deformation of Foamed Elastic Materials," Journal of Applied Polymer Science, Vol. 1, pp. 107-113. Gent, A. N., and Thomas, A. G., 1963, "Mechanics of Foamed Elastic Materials," Rubber Chem, Tech., Vol. 36, pp. 597-610. Gibson, L. J., Ashby, M. F., Schajer, G. S., and Robertson, C. I., 1982, "The Mechanics of Two-Dimensional Cellular Materials," Proceedings of the Royal Society of London Series A, Vol. 382, pp. 25-42. Gibson, L. J., and Ashby, M. F., 1982, "The Mechanics of Three- Dimensional Cellular Materials," Proceedings of the Royal Society of London Series A, Vol. 382, pp. 43-59. Goldstein, H., 1950, Classical Mechanics, Addison-Wesley, Cambridge, MA, pp. 107-109. Ko, W. L., 1965, "Deformations of Foamed Elastomers," Journal of Cellular Plastics, Vol. I, pp. 45-50. Lekhnitskii, S. G., 1963, Theory of Elasticity of an Anisotropic Elastic Body, Holden-Day, San Francisco, pp. 21-25. Menges, G., and Knipschlld, F., 1975, "Estimation of Mechanical Properties for Rigid Polyurethane Foams," Polymer Eng. Sci., Vol. 15, pp. 623-627. Menges, G., and Knipschild, F., 1982, "Stiffness and StrengthRigid Plastic Foams," in Mechanics of Cellular Plastics, ed., Hilyard, N. C., Macmillan, New York, pp. 27-72. Warren, W. E., and Kraynik, A. M., 1987, "Foam Mechanics: The Linear Elastic Response of Two-Dimensional Spatially Periodic Cellular Materials," Mechanics of Materials, Vol. 6, pp. Tl-Vl.

Concluding Remarks
These results illustrate and reinforce the strong connection between elastic properties of foams and their microstructure. The analysis indicates which strut deformation mechanism - bending or stretching determines the various effective elastic constants for particular foam network struc- tures. When network structure is tetrahedral, strut bending provides a square dependence of the distortional moduli upon volume fraction of solid material. This microstructure is favored when surface tension forces are prevalent during foam formation from liquids. For cubic microstructure, which is not consistent with a liquid foam precursor, strut extension results in a linear variation of the moduli with volume frac- tion. Such linear variation is also predicted by theories that do not incorporate strut bending. Previous assertions regarding the importance of a strut bending mechanism depend upon network structure and, therefore, are not universal.

lUTAM Symposium on Elastic Wave Propagation and Ultrasonic Nondestructive Evaluation


The lUTAM symposium of the above title will be held at the University of Colorado, Boulder, during July 30-August 3, 1989. The purpose of the four-day lUTAM symposium is to provide a forum for exchange of ideas among the researchers in elastic-wave propagation, ultrasonic nondestructive evaluation and related areas. We want to assess recent progress and delineate future directions of research. The program will consist of general lectures and selected papers in the areas of wave pro- pagation and scattering in.complex media, thermal waves, impact, dynamic crack growth and radiation, acoustic emission, ultrasonic NDE and fracture mechanics, acoustic microscopy and laser ultrasonics, ultrasonic characterization of material properties, and other related areas. Attendance at the Symposium is by invitation from the Scientific Committee. Con- tributions are invited on any of the topics listed, or on other related topics. Submit three copies of two to three double-spaced page summary of the proposed paper by November 30, 1988, to; Professor S. K. Datta CIRES, Campus Box 449 University of Colorado, Boulder, CO 80309-0449

346/Vol. 55, JUNE 1988

Transactions of the ASME

You might also like