Elastic waves interacting with a thin, prestressed, fiber-reinforced surface film

Elastic waves interacting with a thin, prestressed, fiber-reinforced surface film

International Journal of Engineering Science 48 (2010) 1604–1609 Contents lists available at ScienceDirect International Journal of Engineering Scie...

186KB Sizes 2 Downloads 31 Views

International Journal of Engineering Science 48 (2010) 1604–1609

Contents lists available at ScienceDirect

International Journal of Engineering Science journal homepage: www.elsevier.com/locate/ijengsci

Elastic waves interacting with a thin, prestressed, fiber-reinforced surface film David J. Steigmann Department of Mechanical Engineering, University of California, Berkeley, CA 94720, United States

a r t i c l e

i n f o

Article history: Available online 24 July 2010 To K.R. Rajagopal, for his courage and intellectual honesty, on the occasion of his 60th birthday Keywords: Thin films Surface waves Anisotropy

a b s t r a c t Elastic surface waves propagating at the interface between an isotropic substrate and a thin, transversely isotropic film are analyzed. The transverse isotropy is conferred by fibers lying parallel to the interface. A rigorous leading-order model of the thin-film/substrate interface is derived from the equations of three-dimensional elasticity for prestressed, transversely isotropic films having non- uniform properties. This is used to study Love waves. Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction In this work we derive a rigorous leading-order-in-thickness model for the dynamics of a thin, fiber-reinforced elastic film bonded to an isotropic elastic substrate. The fibers are assumed to lie in the plane of the film and equations of motion are derived for the displacement of the interface. Related work based on a variational argument is given in [1], where references to the pertinent literature may be found. Among these works we direct the reader’s particular attention to [2–4]. Related developments are reported in [5,6]. 2. Basic equations Standard notation is used throughout. Thus, we use bold face for vectors and tensors and indices to denote their components. Latin indices take values in {1, 2, 3}; Greek in {1, 2}. The latter are associated with surface coordinates and associated vector and tensor components. A dot between bold symbols is used to denote the standard inner product. Thus, if A1 and A2   are second-order tensors, then A1  A2 ¼ tr A1 At2 , where tr() is the trace and the superscript t is used to denote the transpose. The notation  identifies the standard tensor product of vectors. If C is a fourth-order tensor, then C[A] is the secondorder tensor with orthogonal components CijklAkl. Finally, we use symbols such as Div and D to denote the three-dimensional divergence and gradient operators, while div and r are reserved for their two-dimensional counterparts. Thus, for example, DivA = Aij,jei and divA = Aia,aei, where {ei} is an orthonormal basis and subscripts preceded by commas are used to denote partial derivatives with respect to Cartesian coordinates. The three-dimensional equation of motion without body force is

€; Div P ¼ qu

ð1Þ

P ¼ S þ HS þ C½H

ð2Þ

where

E-mail address: [email protected] 0020-7225/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.ijengsci.2010.06.032

1605

D.J. Steigmann / International Journal of Engineering Science 48 (2010) 1604–1609

is the linear approximation to the Piola stress, S is the (symmetric) residual stress, H = Du is the gradient of the displacement € is the acceleration, and C is the fourth-order tensor of elastic moduli. The moduli possess the usual minor and field uðx; tÞ; u major symmetries, the latter ensuring that

P ¼ U H;

ð3Þ

1 UðH; xÞ ¼ S  H þ ðHS  H þ H  C½HÞ 2

ð4Þ

where

is the quadratic-order approximation to the strain energy per unit volume of the material region R in which explicit dependence on x 2 R is present if the material is non-uniform. Any such dependence occurs through the residual stress and the moduli. Here we take these to be uniform and thus restrict attention to uniform materials. We suppose that traction data

t ¼ Pn

ð5Þ

are assigned on a part of the boundary @R with exterior unit normal n. We impose the strong-ellipticity condition

ðw  SwÞv  v þ v  w  C½v  w > 0 for all

v  w – 0:

ð6Þ

This is necessary for the undeformed body to be a minimizer of the total strain energy. It is also necessary for minimizers of the potential energy in standard mixed traction/displacement boundary-value problems [7]. 3. Motion of the film/substrate interface The undeformed film/substrate interface is a plane denoted by X. Let k be the unit vector that orients the interface, directed away from the substrate. The reference placement of the film is described by

x ¼ r þ 1k;

ð7Þ

where r 2 X, k is the fixed orientation of the film, and 1 2 [0, h] in the film, where h is the film thickness. The origin of the position r is assumed to lie on X. In the present work the film-substrate combination is a half space that supports a propagating surface wave whose wavelength l (the reciprocal of the wavenumber k) furnishes the only length scale to which h can be compared. Henceforth we regard h as being small in the sense that h/l  1. In the present section it simplifies matters to adopt l as the unit of length (i.e., l = 1, h  1). 0 Let u(r, 1, t) be the function obtained by substituting (7) into u(x, t), and let r() and () , respectively, stand for the (twodimensional) gradient with respect to r at fixed 1 and the derivative @()/@ 1 at fixed r. Further, let

1 ¼ I  k  k;

ð8Þ 0

where I is the identity for three-space; this is the projection onto the translation (vector) space X of X. In [1] it is used to derive

H1 ¼ ru;

Hk ¼ u0

ð9Þ

and the consequent orthogonal decomposition

H ¼ ru þ u0  k:

ð10Þ

Using (8) with P = PI, we also obtain

P ¼ P1 þ Pk  k

ð11Þ

and write (1) in the form

€; div ðP1Þ þ P0 k ¼ qu

ð12Þ +

where div is the two-dimensional divergence on X. This holds at all points of the thin-film, and in the limit 1 &0 in particular, yielding the interfacial equation of motion

€0; div ðP0 1Þ þ P00 k ¼ q0 u

ð13Þ

where, here and henceforth, the subscript ()0 is used to denote the values of functions on the interface X defined by 1 = 0. If t+ is the traction exerted by the environment on the film, then t+ = P+k where P+ is the stress at 1 = h. A Taylor expansion furnishes

tþ ¼ P0 k þ hP00 k þ oðhÞ:

ð14Þ

Accordingly, (13) and (14) combine to yield 1 1 €0: div ðP0 1Þ þ h ðtþ  P0 kÞ þ h oðhÞ ¼ q0 u

ð15Þ

1606

D.J. Steigmann / International Journal of Engineering Science 48 (2010) 1604–1609

For this to furnish a well-defined problem in the limit of small thickness it is necessary that

tþ  P0 k ¼ OðhÞ as h ! 0:

ð16Þ

This is the net traction acting on the lateral surfaces of the film. Let r be the stress in the substrate, assumed to occupy the half-space defined by 1 < 0. Let r0 be the limit of r as 1%0. Then the traction exerted on the substrate by the film at the film-substrate interface X is r0k, whereas that exerted by the substrate on the film is P0k. An elementary pillbox argument yields the exact result

r0 k ¼ P0 k:

ð17Þ

Combining this with (15) and passing to the limit in (15) and (16) yields the leading-order model

€0 div ðhP0 1Þ þ tþ  r0 k ¼ hq0 u

and tþ  P0 k ¼ 0:

ð18Þ

These equations, together with interfacial continuity of displacements, couple the responses of the film and substrate. In particular, if w(x,t) is the displacement field in the substrate, then

w0 ðr; tÞ ¼ u0 ;

ð19Þ

where w0 ¼ wjX . We assume the substrate to be free of residual stress, so that

r ¼ E½Dw;

ð20Þ

where E is the associated tensor of elastic moduli, possessing the usual minor and major symmetries. Thus, in contrast to the film stress P, r is symmetric. This satisfies the equation of motion

€ Div r ¼ qs w;

ð21Þ

where qs is the mass density of the substrate. The foregoing model agrees precisely with the leading-order system derived elsewhere [1] via a variational argument. We demonstrate that (18)2 can be solved uniquely for the derivative u00 . To this end we fix u0 and define

WðaÞ ¼ Uðru0 þ a  kÞ  a  tþ : Let

ð22Þ 

 be a parameter and consider the one-parameter family a(). Let g() = W(a()) and let ()

= d()/d. Then (3) yields

€ þ a_  ðW aa Þa; _ g_ ¼ W a  a_ and g€ ¼ W a  a

ð23Þ

with

W a ¼ P0 k  tþ

and ðW aa Þa_ ¼ ðP0 kÞ ;

ð24Þ

wherein P0 is given by (2) and (10), with a() substituted in place of u00 . From (18)2 we find that W is stationary at a ¼ u00 , while (2) furnishes

a_  ðP0 kÞ ¼ ðk  S0 kÞa_  a_ þ a_  k  C0 ½a_  k:

ð25Þ

€ > 0 on straight-line paths This is strictly positive by virtue of (6), implying that Waa is positive definite. In particular, g defined by a() = (1  )a1 + a2 with a1, a2 fixed and  2 [0, 1]. These paths belong to the domain of W(), the convex set gen_ Þ > gð0Þ _ _ erated by the linear space of 3-vectors. Integrating with respect to  yields gð and gð1Þ  gð0Þ > gð0Þ, proving that W(a) is strictly convex; i.e.,

Wða2 Þ  Wða1 Þ > W a ða1 Þ  ða2  a1 Þ;

ð26Þ

for all unequal pairs a1, a2. Because strictly convex functions have unique stationary points, it follows that

ðru0 ; tþ Þ; u00 ¼ a

ð27Þ

 is the value of a at which Wa vanishes. where a In the present work we impose the condition t+ = 0 for all deformations. Eqs. (2) and (18)2, specialized to the case of zero displacement, then require that

S0 k ¼ 0;

ð28Þ

which in turn may be used to simplify (27) for arbitrary displacements. This yields

u00 ¼ A1 0 ðC0 ½ru0 Þk;

ð29Þ

where A0 is the value on X of the acoustic tensor of the film material, defined, for arbitrary v, by

A0 v ¼ ðC0 ½v  kÞk:

ð30Þ

D.J. Steigmann / International Journal of Engineering Science 48 (2010) 1604–1609

1607

That this is positive definite and hence invertible follows from (6), with (28). Accordingly, (18)1 and (29) furnish a system for the interfacial displacement field u0(r, t). Following Spencer [8] we model the film as a transversely isotropic solid. The axes of transverse isotropy are coincident with the direction fields of the (straight) fiber trajectories. We drop the subscript 0 from all notation denoting interfacial values of variables. The components of C relative to the basis {ei} are [1]

C ijkl ¼ kdij dkl þ lT ðdik djl þ dil djk Þ þ aðdij mk ml þ mi mj dkl Þ þ ðlL  lT Þðmi mk djl þ mi ml djk þ mj mk dil þ mj ml dik Þ þ bmi mj mk ml ;

ð31Þ

where dij is the Kronecker delta; a, b, k, lT and lL are material constants; and the unit vector m, with components mi, is the fiber axis, assumed here to be uniform and lying in the plane of the film. Spencer [8] shows that lT is the shear modulus for shearing in planes transverse to m, whereas lL is the shear modulus for shearing parallel to m. The remaining material constants in (31) may be interpreted in terms of extensional moduli and Poisson ratios [8]. The general form of the residual stress may be derived by enumerating the strain invariants for transverse isotropy that are linear in the (infinitesimal) strain. These are [8] I  H and m  m  H. Comparison with the linear term in (4) then furnishes

S ¼ ST ðI  m  mÞ þ SL m  m;

ð32Þ

where ST is the constant residual stress in the isotropic plane and SL is the constant residual uniaxial stress along m. In this work we consider the film to be a prismatic body formed by the parallel translation of a midplane X in the direction of its unit normal k(=e3). The fibers are assumed to lie parallel to X, so that m3 = 0 and m = maea. Eq. (28) then yields

S0 ¼ S0 m  m;

ð33Þ

where S0 is the residual interfacial stress. The acoustic tensor defined by (30) is

A ¼ kk  k þ lT ðI þ k  kÞ þ ðlL  lT Þm  m;

ð34Þ

with eigenvalues lL, lT and k + 2lT. Inequality (6), with (28), requires that these be strictly positive. Eq. (29) then yields

u00 ¼ ðk þ 2lT Þ1 ½kðdivv Þ þ am  ðrv Þmk  rw;

ð35Þ

where divv = va,a, m  (rv)m = va,bmamb and

v ¼ 1u0

and w ¼ k  u0

ð36Þ

are the in-plane and transverse interfacial displacements, respectively; i.e.,

u0 ¼ v þ wk:

ð37Þ

4. Example: Love waves We are concerned with the acoustic interaction of the film and substrate, the former coinciding with the r1, r2-plane and the latter with the half-space defined by 1 < 0. Accordingly, we study harmonic surface waves whose amplitudes decay with depth in the substrate. For the sake of simplicity we confine attention to Love waves. For waves propagating along the r1direction, these have the form

ui ¼ di2 Fðr 1 ; 1; tÞ;

Fðr 1 ; 1; tÞ ¼ f ðr 1 ; tÞ expðgk1Þ;

ð38Þ

where g and k are positive constants, and

f ðr 1 ; tÞ ¼ A exp½ikðr1  ctÞ;

ð39Þ

in which c is the wavespeed and A is a constant. The induced displacement of the film/substrate interface is

w ¼ u3 ¼ 0;

v a ¼ ua ¼ da2 f ðr1 ; tÞ:

ð40Þ

For uniform isotropic materials, the film-substrate interaction term is [1]

rk ¼ ls gkf ðr1 ; tÞe2 :

ð41Þ

In the substrate, Eq. (21) is satisfied provided that [2,3]

pffiffiffiffiffiffiffiffiffiffiffiffiffi

g ¼ 1  s2 ; where

s ¼ c=cs < 1 and cs ¼

ð42Þ qffiffiffiffiffiffiffiffiffiffiffiffi ls =qs

is the transverse wavespeed in the substrate.

ð43Þ

1608

D.J. Steigmann / International Journal of Engineering Science 48 (2010) 1604–1609

According to (40)2, we have div v = 0 identically at the interface. Using (35) and Eqs. (76) and (82) of [9] we find, after much manipulation, omitted here for the sake of brevity, that (18)1 becomes

hSðrrv Þm  m þ hlT Dv þ hcr½m  ðrv Þm þ hdfm  ½ðrrv Þm  mg þ hðlL  lT Þ½ðm  Dv Þm þ ðrrv Þm  m  rk ¼ hqv tt ;

ð44Þ

where



ka þ a þ lL  lT ; k þ 2lT

d¼b

a2

ð45Þ

k þ 2l T

and where (rrv)m  m = vc,abmambec and Dv = vc,aaec. Substituting (40)2 reduces this to

hf

00

n o   Sm21 e2 þ cm1 m2 e1 þ dm21 m2 m þ ðlL  lT Þ m2 m þ m21 e2 þ lT e2  ls gkf e2 ¼ hq€f e2 ;

ð46Þ

wherein the primes and dots refer to derivatives of f with respect to r1 and t, respectively. This is equivalent to its projections onto e1 and e2, given respectively by

  00 hf m1 m2 c þ dm21 þ lL  lT ¼ 0

ð47Þ

and

hf

00

  Sm21 þ dm21 m22 þ lL  ls gkf ¼ hq€f :

ð48Þ

With f given by (39), Eq. (47) requires that either the expression in parentheses vanishes or that m1m2 vanishes. For typical data on carbon-fiber/epoxy-resin composites [8], we find that the first alternative has no solution for real-valued m1. The remaining alternative yields the two possibilities m = ±e1 and m = ±e2, corresponding to waves propagating parallel and transverse to the fibers, respectively. In the first case; i.e., for waves traveling along the fibers, (48) reduces to 2

hkðS þ lL Þ þ ls g ¼ hqkc ;

ð49Þ

yielding



g ¼ hk rs2 

S þ lL

ls

 ;

ð50Þ

where r = q/qs, and

s ¼ c=cs < 1;

where cs ¼

qffiffiffiffiffiffiffiffiffiffiffiffi ls =qs

ð51Þ

is the transverse shear-wave speed in the substrate. Eliminating g between (49) and (50) yields the dispersion relation

  pffiffiffiffiffiffiffiffiffiffiffiffiffi S þ lL ; 1  s2 ¼  rs2 

ð52Þ

ls

where  = hk. The theory used here to obtain this relation purports to be valid only if ingly, we assume that s2 = 1  2C2 + o(2), solve for C and obtain

s1

  1, so that 1  s2 = O(2). Accord-

 2 1 2 S þ lL  r ; 2 ls

ð53Þ

to leading order. The case of waves traveling in the direction transverse to the fibers leads to the same results, but with S omitted. We observe that the relevant film stiffness in both cases is the longitudinal modulus lL. This is to be expected because the transverse modulus lT pertains to shearing in the isotropic plane orthogonal to m, whereas deformations of the type considered induce shearing in a plane containing m. It is easy to show, from (31), that such shear deformations generate a shear stress equal to 2lL times the shear strain, provided that m is oriented along either of the axes of strain; i.e., along ±e1 or ± e2, as in the foregoing. Acknowledgment The author would like to thank the referees for helpful comments leading to substantial improvement of the article. References [1] D.J. Steigmann, R.W. Ogden, Surface waves supported by thin-film/substrate interactions, IMA J. Appl. Math. 72 (2007) 730–747. [2] J.D. Achenbach, Wave Propagation in Elastic Solids, North-Holland, Amsterdam, 1973. [3] I. Murdoch, The propagation of surface waves in bodies with material boundaries, J. Mech. Phys. Solids 24 (1976) 137–146.

D.J. Steigmann / International Journal of Engineering Science 48 (2010) 1604–1609

1609

[4] A.V. Pichugin, G.A. Rogerson, An asymptotic membrane-like theory for long-wave motion in a pre-stressed elastic plate, Proc. Roy. Soc. Lond. A458 (2002) 1447–1468. [5] A.L. Shuvalov, A.G. Every, On the long-wave onset of dispersion of the surface-wave velocity in coated solids, Wave Motion 45 (2008) 857–863. [6] T.C.T. Ting, Steady waves in an anisotropic elastic layer attached to a half-space or between two half-spaces: a generalization of Love waves and Stoneley waves, Math. Mech. Solids 14 (2009) 52–71. [7] R.J. Knops, L.E. Payne, Uniqueness Theorems in Linear Elasticity, Springer Tracts in Natural Philosophy 19 (1971). [8] A.J.M. Spencer, Constitutive theory for strongly anisotropic solids, in: A.J.M. Spencer (Ed.), Continuum Theory of the Mechanics of Fibre-Reinforced Composites, CISM Courses and Lectures, vol. 282, Springer, Wien, 1984, pp. 1–32. [9] D.J. Steigmann, Linear theory for the bending and extension of a thin, residually stressed, fiber-reinforced lamina, Int. J. Eng. Sci. 47 (2009) 1367–1378.