Skip to main content
Proceedings. Mathematical, Physical, and Engineering Sciences logoLink to Proceedings. Mathematical, Physical, and Engineering Sciences
. 2016 Feb;472(2186):20150800. doi: 10.1098/rspa.2015.0800

Refined boundary conditions on the free surface of an elastic half-space taking into account non-local effects

R Chebakov 1, J Kaplunov 1,, G A Rogerson 1
PMCID: PMC4841666  PMID: 27118902

Abstract

The dynamic response of a homogeneous half-space, with a traction-free surface, is considered within the framework of non-local elasticity. The focus is on the dominant effect of the boundary layer on overall behaviour. A typical wavelength is assumed to considerably exceed the associated internal lengthscale. The leading-order long-wave approximation is shown to coincide formally with the ‘local’ problem for a half-space with a vertical inhomogeneity localized near the surface. Subsequent asymptotic analysis of the inhomogeneity results in an explicit correction to the classical boundary conditions on the surface. The order of the correction is greater than the order of the better-known correction to the governing differential equations. The refined boundary conditions enable us to evaluate the interior solution outside a narrow boundary layer localized near the surface. As an illustration, the effect of non-local elastic phenomena on the Rayleigh wave speed is investigated.

Keywords: non-local, elasticity, asymptotic, boundary conditions, Rayleigh wave, inhomogeneous

1. Introduction

Analysis of non-local elastic phenomena is of major interest for various advanced applications including micro- and nanomechanics (e.g. [13]). Non-local elasticity is a particularly powerful and appropriate theory for investigating properties of solids with impurities, dislocations and granular microstructure. The fundamental concepts underpinning contemporary non-local continuum models were developed in a series of well-known papers by Kroner [4], Eringen [5], Eringen & Edelen [6]; see also [7] and references therein. The state of art has been presented by a number of authors throughout the area's scientific development [813]. The latter paper addresses Piola's important contribution, not widely known for a long time to a broad international audience, see also references to Piola's original papers in Dell'Isola et al. [13].

Among other recent publications on the subject, we mention papers by Di Paola & Zingales [14], Di Paola et al. [15], Zingales [16], Schwartz et al. [17], Benvenuti & Simone [18], Abdollahi & Boroomand [19,20], dealing with various analytical and numerical aspects of non-local elasticity. Here, we also cite publications developing novel micromechanical approaches known as ‘structured deformations’ (e.g. [2123]).

Non-local models (e.g. [2426]) are oriented to the investigation of the distant interaction between small material particles, assuming that the stress at a reference point is dependent upon the entire strain field in the body. The associated constitutive relations are usually expressed through integral operators involving internal sizes which characterize microstructure. As a rule (e.g. [27]), the long-wave limit of the non-local elasticity relations is identical to its classical counterpart. We also remark that a number of non-local elasticity predictions are in good agreement with lattice dynamics, including the regions near the boundaries of the body [28].

In spite of the numerous publications, the fundamental effect of boundaries on the implementation of non-local elasticity concepts has not yet been properly addressed. The key point is that the intervals of integration corresponding to the above-mentioned operators, expressing non-local constitutive relations, are dependent on the distance from a reference point to the boundary [25]. This results in boundary layers corresponding to localized non-homogeneous stress and strain fields. In this paper, we fill the gap in tackling the influence of boundary layers on overall dynamic behaviour. Although several authors emphasized the crucial role of boundary layers (e.g. [29,20]), we are not aware of any related asymptotic developments.

As an example, we consider an elastic half-space governed by the non-local equations given in Eringen [25], see §2. For the sake of definiteness, we assume that the non-local behaviour is modelled by an exponential kernel involving a small internal lengthscale. In §3, we proceed with a long-wave asymptotic scheme, originating from Goldenveizer et al. [30] and later developed by, for example, Dai et al. [31] and Aghalovyan [32]. Within the framework of these studies, the characteristic wavelength is assumed to be much greater than a typical microscale parameter. We begin by reducing the original non-local problem to a formulation which is identical to the classical problem for an elastic half-space with a vertical inhomogeneity localized near the surface. The effect of the inhomogeneity can be reduced to effective boundary conditions imposed at a near-surface interface. In this case, we can only asymptotically evaluate the interval, yielding the location of the interface. A better option seems to be a transformation of the effective conditions to refined boundary conditions along the surface of a homogeneous half-space. This approach is exploited in §4, enabling us to evaluate the interior stress and strain outside the narrow boundary layer. In §5, the refined boundary conditions are applied to calculate the non-local correction to the Rayleigh surface wave. The order of this correction exceeds that of the correction established in Eringen [25] associated with the non-local differential equations of motion.

2. Equations of non-local linear elasticity

In this section, we use as our starting point the equations of non-local elasticity, (e.g. [25]). For a homogeneous isotropic elastic solid, we therefore have (2.1)–(2.5) below:

sαβ,α=ρ2uβt2, 2.1

with uββ=1,2,3, the components of the displacement vector, ρ volume density, t time and

sαβ(x)=VK(|xx|,a)σαβ(x)dv(x), 2.2

where sαβ and σαβ are the non-local and classical stress tensors, respectively, considered at time t, x=(x1,x2,x3) is a reference point, V the domain occupied by the body, K(x,a) the so-called non-local modulus, and a is an internal characteristic length, e.g. lattice parameter or granular distance. Throughout the paper we assume that the internal size a is asymptotically small in comparison with a typical wavelength. This long-wave assumption provides the validity of the adapted non-local model for bounded domains as it follows on, in particular, from lattice dynamics [28]; for further details, see concluding remarks.

The function K in (2.2) is normalized over three-dimensional space, so that

VK(|x|,a)dv(x)=1. 2.3

The two equations (2.1) and (2.2) are accompanied by

σαβ=λeγγδαβ+2μeαβ 2.4

and

eαβ=12(uαxβ+uβxα), 2.5

where eαβ is the linear elastic strain tensor, δαβ the Kronecker's delta, and λ and μ are the Lamé constants.

For the sake of definiteness, we specify the three-dimensional exponential non-local modulus in the same way as Eringen [25], thus

K(|x|,a)=1π3/2a3exp[xxa2], 2.6

where in the case of a half-space <x1<, <x2< and 0x3<, (2.2) becomes

sαβ(x)=1π3/2a30dx3dx1dx2exp[(xx)2a2]σαβ(x). 2.7

Let us now expand the stresses σαβ in Taylor series about the reference point x′=x, assuming as before that the typical wavelength characterizing the classical stress field is much greater than the internal size a. Thus, we establish from (2.7) that

sαβ(x)=1aπ{σαβ(x)0exp[(x3x3)2a2]dx3+σαβ(x)x30(x3x3)exp[(x3x3)2a2]dx3}+, 2.8

which on integration yields

sαβ(x)=σαβ(x)2 erfc (x3a)+a2πσαβ(x)x3exp[x32a2]+, 2.9

where  erfc (x)=(2/π)xet2dt. Here, we keep only a linear term in a which is specific for a half-space. Such term does not appear in the case of three-dimensional space.

The formulae (2.2), taking into account (2.4) and keeping the leading-order term in (2.9), may be presented as

sαβ=λeγγδαβ+2μeαβ, 2.10

where

λ=12 erfc(x3a)λandμ=12 erfc(x3a)μ.} 2.11

We remark that the non-local problem (2.1), (2.10)–(2.11) is thus formally equivalent to the classical (‘local’) problem for a vertically inhomogeneous elastic half-space. This problem may in fact be reduced to that of analysis of a homogeneous elastic substrate coated by a vertically inhomogeneous layer of a certain thickness h (figure 1), where ha. This strong inequality justifies the validity of non-local theory on the scale of layer thickness. As a rule, the asymptotic error of the one-term expansion in (2.9) is O(a/h). It is less than O(a/h) only provided that the associated local field is uniform in x3.

Figure 1.

Figure 1.

A homogeneous substrate coated by a vertically inhomogeneous layer of thickness h; ah≪ℓ.

Along the interface x3=h, to within an exponentially small error, erfc (−h/a)=2 and, consequently,

λ(h)=λandμ(h)=μ, 2.12

and the non-local stresses sαβ tend to their local analogues σαβ in (2.9).

In the case of a thin layer of thickness much smaller than a macroscale wavelength, its effect on the substrate may be incorporated by deriving effective boundary conditions using well-known asymptotic methodology (see for example [31,32] and references therein).

3. Asymptotic analysis of a vertically inhomogeneous thin layer

Let us consider a thin, vertically inhomogeneous layer of thickness h≪ℓ, where ℓ is a typical wavelength with ε=h/ℓ assumed to be a small geometric parameter. Equations (2.1) and (2.10) in the previous section are formally identical to the classical ‘local’ equations. They can be rewritten as

siixi+sijxj+s3ix3=ρ2uit2ands3ixi+s3jxj+s33x3=ρ2u3t2,} 3.1

and

sij=ρc22(x3)(uixj+ujxi),sii=ρc12(x3)uixi+ρ(c12(x3)2c22(x3))(ujxj+u3x3),s3i=σi3=ρc22(x3)(uix3+u3xi)ands33=ρc12(x3)u3x3+ρ(c12(x3)2c22(x3))(uixi+ujxj),} 3.2

where ij=1,2 and Einstein's summation convention is not employed. The variable wave speeds in (3.2), inspired by (2.11), are given by

c1(x3)=λ(x3)+2μ(x3)ρandc2(x3)=μ(x3)ρ. 3.3

The traction-free boundary conditions at the surface of the layer x3=0 are given by

s3n=0at x3=0, 3.4

with continuity of displacement along the interface x3=h requiring that

un=vnat x3=h, 3.5

where vn=vn(x1,x2,t) denotes the prescribed displacements in the substrate, n=1,2,3.

We now adapt the asymptotic approach developed by Goldenveizer et al. [30], Dai et al. [31] and Aghalovyan [32] in order to express the stresses s3n along the interface x3=h in terms of the prescribed substrate displacements vn. To begin, we scale the original variables as follows:

ξi=xi,η=x3handτ=tc2, 3.6

where c2=c2′(h), and also define the dimensionless quantities

un=1Vun,vn=1Vvn

and

sij=μVsij,sii=μVsii,s3n=2μhVs3n, 3.7

where V is the maximum displacement amplitude and all quantities with an asterisk are assumed to be of the same asymptotic order.

The equations of motion (3.1) and constitutive relations (3.2) can now be rewritten as

siiξi+sijξj+s3iη=2uiτ2ands33η+ε(s3iξi+s3jξj)=2u3τ2,} 3.8

and

sij=κ22(uiξj+ujξi),εsii=(κ122κ22)u3η+ε(κ12uiξi+(κ122κ22)ujξj),ε2s3i=κ22(uiη+εu3ξi)andε2s33=κ12u3η+ε(κ122κ22)(uiξi+ujξj),} 3.9

with κm′=cm′(x3)/c2m=1,2.

It is convenient to express u3/η in (3.9)2 from (3.9)4, having

sii=4κ22(1κ22κ12)uiξi+2κ22(12κ22κ12)ujξj+ε(12κ22κ12)s33. 3.10

The boundary conditions (3.4) and (3.5) become

s3n=0at η=0andun=vnat η=1.} 3.11

Next, we expand the displacements and stresses in asymptotic series in terms of the previously specified small parameter ε, and thus introduce

(unsiisijs3is33)=(un(0)sii(0)sij(0)s3i(0)s33(0))+ε(un(1)sii(1)sij(1)s3i(1)s33(1))+ 3.12

Substitution of these expressions into equations (3.8)–(3.10), and boundary conditions (3.11) results, at leading order, in the following equations:

sii(0)ξi+sij(0)ξj+s3i(0)η=2ui(0)τ2ands33(0)η=2u3(0)τ2,} 3.13

and

sij(0)=κ22(ui(0)ξj+uj(0)ξi),sii(0)=4κ22(1κ22κ12)ui(0)ξi+2κ22(12κ22κ12)uj(0)ξjandun(0)η=0,} 3.14

together with

un(0)=vnat η=1ands3n(0)=0at η=0.} 3.15

On integrating (3.13)2 and (3.14)3 with respect to η, and taking into account the appropriate boundary conditions (3.15), we may establish that

un(0)=vn 3.16

and

s33(0)=η2v3τ2. 3.17

Now we obtain from (3.14)2

sii(0)=4κ22(1κ22κ12)viξi+2κ22(12κ22κ12)vjξj. 3.18

We finally integrate (3.13)1, using (3.14)2 and (3.16), and then satisfy (3.15)2, to establish that

s3i(0)=η2viτ22viξj20ηκ22dη42viξi20ηκ22(1κ22κ12)dη2vjξiξj0ηκ22(34κ22κ12)dη. 3.19

In terms of the original variables, the expressions for the stresses s3i and s33 may be obtained from (3.19) and (3.17) in the form

s3i=ρ[x32uit2c222uixj20x3κ22dx3ρc222uixi20x34κ22(1κ22κ12)dx3c222ujxixj0x3κ22(34κ22κ12)dx3]ands33=ρx32u3t2,} 3.20

where now un=Vun(0).

In what follows we also use the formula for other components of the non-local stress tensor, which are given by

sii=2ρc22[2(1c22c12)uixi+(12c22c12)ujxj]andsij=ρc22(uixj+ujxi).} 3.21

The stresses at the interface, x3=h, may be expressed through the substrate displacements, yielding

s3i=ρ[h2vit2c222vixj20hκ22dx3ρc222vixi20h4κ22(1κ22κ12)dx3c222vjxixj0hκ22(34κ22κ12)dx3]ands33=ρh2v3t2,} 3.22

where κm′=cm′/c2m=1,2 as above.

4. Refined boundary conditions

For a vertically inhomogeneous layer with the elastic moduli given by (2.11), we obtain

cm2(x3)=12cm2 erfc (x3a),m=1,2, 4.1

where, see (2.12),

c1=λ+2μρandc2=μρ.

Consequently, the integrals in (3.22), under the assumption ah, to within an exponentially small error may be presented in the forms

0hκ22dx3=h(112πah),0h4κ22(1κ22κ12)dx3=4h(1κ2)(112πah)and0hκ22(34κ22κ12)dx3=h(34κ2)(112πah),

where κ=c2/c1. Thus, the stresses along the interface x3=h may be presented as

s3i=ρh[2vit2c22{2vixj2+4(1κ2)2vixi2+(34κ2)2vjxixj}+c22a2hπ{2vixj2+4(1κ2)2vixi2+(34κ2)2vjxixj}]ands33=ρh2v3t2.} 4.2

Inspection of (4.2)1 shows that taking non-local elastic properties into account results in an asymptotic correction of the relative asymptotic order O(a/h). On the other hand, this correction must be greater than the truncation error O(ε) related to the asymptotic derivation of formulae (3.22) in §3. Thus, we arrive at the double strong inequality

aha, 4.3

underlying equations (4.2). We also need to show that the accuracy of the leading-order approximation (2.10) is consistent with the O(a/h) correction associated with (4.2). To this end, we recall that at leading order, the stresses s3i and s33 are expressed in terms of the stresses sii and the displacements un in §3, see (3.13). In this case, the local stresses σii and σij corresponding to their non-local counterparts sii and sij in formula (3.21), following from the dimensionless formula (3.13)1, are given by

σii=2ρc22[2(1κ2)uixi+(12κ2)ujxj]andσij=ρc22(uixj+ujxi).} 4.4

These stresses are uniform across the thickness. As a result, the expected O(a/h) contribution of the second term in the expansion (2.9) vanishes after differentiation with respect to x3. For the same reason, the inertial terms in (3.13) will also not make a O(a/h) contribution to non-local stresses.

It is clear that outside a narrow near-surface layer, all non-local stresses, to within the asymptotic error less than O(a/h), coincide with their local analogues (see (2.9)–(2.12)). Therefore, over the interior domain x3h, we may proceed with a classical problem with constant coefficients, subject to the boundary conditions

σ3n=s3natx3=h, 4.5

where s3n are given by (4.2).

We are now in a position to formulate an inverse problem for a thin homogeneous elastic layer within the classical framework. A crucial aspect that now needs addressing concerns the boundary conditions to be imposed on the surface x3=0 so that the stresses σ3i and σ33, at the interface x3=h, satisfy the conditions (4.5). Let the boundary conditions at the surface of the layer x3=0 be given by

σ3n=pn, 4.6

where pn are the sought for surface stresses.

The asymptotic solution of the classical elastodynamic equations for a thin homogeneous layer, subject to the boundary conditions (4.5) and (4.6) along the faces x3=0 and x3=h, is presented in Dai et al. [31]. The formulae for the stresses of interest at x3=h may be written as

σ3i=ρh[2vit2c22{2vixj2+4(1κ2)2vixi2+(34κ2)2vjxixj}]+piandσ33=ρh2v3t2+p3.} 4.7

Next, on equating the stresses σ3n in (4.5) and (4.7), we have

pi=ρc22a2π{2vixj2+4(1κ2)2vixi2+(34κ2)2vjxixj}andp3=0.} 4.8

Thus, the refined boundary conditions, at x3=0, become

σ3i=ρc22a2π{2vixj2+4(1κ2)2vixi2+(34κ2)2vjxixj}andσ33=0.} 4.9

Outside a narrow boundary layer (x3a), the half-space motion is governed by the elastodynamic equations with constant moduli λ and μ, subject to the boundary conditions (4.9). The last formulae involve O(a/ℓ) correction, where ℓ is a typical macroscale size as described above. This is greater than the O(a2/ℓ2) correction in the differential equations of non-local elasticity [25].

5. Rayleigh surface wave

As an illustration, we consider the effect of non-local elastic behaviour on surface wave propagation in the case of plane strain, in which ∂/∂x2≡0, um=um(x1,x3), m=1,3, and u2=0. Accordingly, the two boundary conditions, following directly from (4.9), become

σ31=2aπρc22(1κ2)2u1x12andσ33=0,} 5.1

and the equations of motion, in terms of wave potentials φ and ψ, are given by

Δφ1c22φt2=0 5.2

and

Δψ1c22ψt2=0, 5.3

where Δ is the two-dimensional Laplacian.

We first look for travelling wave solutions of the form

φ=Aerx3+ik(x1ct)andψ=Beqx3+ik(x1ct),} 5.4

where c is the phase speed and in which the attenuation coefficients are given by

r=k1c2c12andq=k1c2c22.

The displacements may now be expressed in terms of potentials, thus yielding

u1=φ,1+ψ,3=(ikAerx3qBeqx3)eik(x1ct)andu3=φ,3ψ,1=(rAerx3ikBeqx3)eik(x1ct).} 5.5

Next, on substituting (5.5) into the boundary conditions (5.1), we obtain, after taking into account the plane strain forms of (2.4) and (2.5), that

[2c2c22]A+[2i1c2c22]B=0and[i(2akπ(κ21)+21c2c12)]A+[(2c2c22)+2akπ(κ21)1c2c22]B=0.} 5.6

The condition for existence of a non-trivial solution of (5.6) yields

R(γ)4πθ(κ21)γ21γ2=0, 5.7

where θ=a/ℓ=ak/2π≪1 is a small parameter, γ=c/c2 and R(γ) is the Rayleigh denominator, i.e.

R(γ)=(2γ2)24(1γ2)(1κ2γ2).

We may now expand γ as an asymptotic series in the small parameter θ, with

γ=γ0+θγ1+. 5.8

In this case, the Taylor series expansion of R(γ), about γ=γ0, is given by

R(γ)=R(γ0)+R(γ0)(γγ0)+, 5.9

where γ0 is the normalized classical Rayleigh wave speed, i.e. R(γ0)=0. Then, on substituting (5.8) and (5.9) into (5.7), we readily obtain

γ1=4π(κ21)γ021γ02R(γ0) 5.10

and, consequently,

γ=γ0+θ4π(κ21)γ021γ02R(γ0)+, 5.11

where θ=ak/2π, as before.

We remark that the constructed correction, originating from the refined boundary conditions (5.1), exceeds the correction in Eringen [25], associated with the ‘non-local terms’ within the differential equations of motion.

Numerical results are presented in figure 2. The classical Rayleigh root γ0 and the ‘non-local’ root γ in (5.11) are plotted as function of the small parameter θ for the value of Poisson ratio ν=0.25. For this scenario, the coefficient (5.10) takes the value γ1=−0.37, while its ‘local’ counterpart is γ0=0.92. The effect of non-local phenomena decreases the Rayleigh wave speed due to low values of the Lamé parameters, denoting the stiffness of the system, near the surface, see (2.11).

Figure 2.

Figure 2.

Effect of non-local phenomena on Rayleigh wave speed. (Online version in colour.)

6. Concluding remarks

An asymptotic treatment of the non-local boundary value problem under consideration demonstrates the primary importance of analysing the peculiarities of near-surface behaviour. It has been established that the effect of the associated boundary layer may be incorporated just by refining the boundary conditions in classical elasticity. In particular, the refined boundary conditions (4.9) involve an explicit correction to their classical counterparts; this arises by taking into account non-local phenomena.

The linear elastodynamic equations, subject to the derived boundary conditions on the free surface of a homogeneous half-space, enable us to determine the interior stress and strain fields outside a narrow near-surface layer, with thickness satisfying the asymptotic inequality (4.3). As an illustration, O(a/ℓ) non-local correction to the Rayleigh surface wave speed was calculated. This correction is greater than O(a2/ℓ2) correction associated with the non-local equations of motion in Eringen [25].

We recall that approximate nature of non-local models originates from truncation of homogenization procedures, including asymptotic homogenization for periodic structures (e.g. [33,34]), underlying the associated macroscale relations. In this case, the truncation error for the classical boundary conditions should be of the same order as the deviation from the uniform microscale variation of the sought for solution. The latter might be expected to be negligible in comparison with O(a/ℓ) correction suggested in the paper. In particular, it is O(a2/ℓ2) for a range of periodic lattices [35]. This issue certainly merits a thorough consideration.

We remark that the proposed approach is not merely restricted to the exponential kernel (2.6) studied in this paper. We envisage similar non-local effects for a range of kernels having the same asymptotic behaviour at small internal scales. The results obtained may also readily be extended to non-locally elastic solids with a boundary of arbitrary shape. Investigation of elastic waveguides, including beams, plates and shells, with the boundary conditions of the form (4.9) imposed on the free faces would also be of obvious interest. This would in fact seem to be a natural generalization of the above-mentioned example for the Rayleigh surface wave.

The general asymptotic scheme presented in §3 may also seemingly have potential applications outside the area of non-local elasticity. Firstly, we note applications for solids with localized near-surface inhomogeneities, such as functionally graded structures (see for example the review by Birman & Byrd [36]). There is also the possibility of adapting this scheme for long-wave dynamic analysis of vertically inhomogeneous foundations (see Muravskii [37] and references therein).

Acknowledgements

The authors greatly appreciate valuable discussions with Dr D. A. Prikazchikov.

Authors' contributions

All authors developed the asymptotic approach and wrote the paper. All authors gave final approval for publication.

Competing interests

We have no competing interests.

Funding

PhD studies of R.C. were partly supported by Keele University. The support is gratefully acknowledged.

References

  • 1.Drugan WJ, Willis JR. 1996. A micromechanics-based nonlocal constitutive equation and estimates of representative volume element size for elastic composites. J. Mech. Phys. Solids 44, 497–524. (doi:10.1016/0022-5096(96)00007-5) [Google Scholar]
  • 2.Arash B, Wang Q. 2012. A review on the application of nonlocal elastic models in modeling of carbon nanotubes and graphenes. Comput. Mater. Sci. 51, 303–313. (doi:10.1016/j.commatsci.2011.07.040) [Google Scholar]
  • 3.DalCorso F, Deseri L. 2013. Residual stresses in random elastic composites: nonlocal micromechanics-based models and first estimates of the representative volume element size. Meccanica 48, 1901–1923. (doi:10.1007/s11012-013-9713-z) [Google Scholar]
  • 4.Kroner E. 1967. Elasticity theory of materials with long range cohesive forces. Int. J. Solids Struct. 3, 731–742. (doi:10.1016/0020-7683(67)90049-2) [Google Scholar]
  • 5.Eringen AC. 1972. Linear theory of nonlocal elasticity and dispersion of plane waves. Int. J. Eng. Sci. 10, 425–435. (doi:10.1016/0020-7225(72)90050-X) [Google Scholar]
  • 6.Eringen AC, Edelen DGB. 1972. On nonlocal elasticity. Int. J. Eng. Sci. 10, 233–248. (doi:10.1016/0020-7225(72)90039-0) [Google Scholar]
  • 7.Eringen AC. 2002. Nonlocal continuum field theories. New York, NY: Springer Science & Business Media. [Google Scholar]
  • 8.Krumhansl JA. 1968. Some considerations of the relation between solid state physics and generalized continuum mechanics. In Mechanics of Generalized Continua Proc. IUTAM Symp. on the Generalized Cosserat Continuum and the Continuum Theory of Dislocations with Applications, Freudenstadt and Stuttgart, 1967 (ed. E Kroner), pp. 298–311. Berlin, Germany: Springer.
  • 9.Kunin IA. 1984. On foundations of the theory of elastic media with microstructure. Int. J. Eng. Sci. 22, 969–978. (doi:10.1016/0020-7225(84)90098-3) [Google Scholar]
  • 10.Polizzotto C. 2001. Nonlocal elasticity and related variational principles. Int. J. Solids Struct. 38, 7359–7380. (doi:10.1016/S0020-7683(01)00039-7) [Google Scholar]
  • 11.Peddieson J, Buchanan GR, McNitt RP. 2003. Application of nonlocal continuum models to nanotechnology. Int. J. Eng. Sci. 41, 305–312. (doi:10.1016/S0020-7225(02)00210-0) [Google Scholar]
  • 12.Di Paola M, Failla G, Pirrotta A, Sofi A, Zingales M. 2013. The mechanically based non-local elasticity: an overview of main results and future challenges. Phil. Trans. R. Soc. A 371, 20120433 (doi:10.1098/rsta.2012.0433) [DOI] [PubMed] [Google Scholar]
  • 13.Dell'Isola F, Andreaus U, Placidi L. 2014. At the origins and in the vanguard of peridynamics, non-local and higher-gradient continuum mechanics: an underestimated and still topical contribution of Gabrio Piola. Math. Mech. Solids 20, 887–928. (doi:10.1177/1081286513509811) [Google Scholar]
  • 14.Di Paola M, Zingales M. 2008. Long-range cohesive interactions of non-local continuum faced by fractional calculus. Int. J. Solids Struct. 45, 5642–5659. (doi:10.1016/j.ijsolstr.2008.06.004) [Google Scholar]
  • 15.Di Paola M, Failla G, Zingales M. 2009. Physically-based approach to the mechanics of strong non-local linear elasticity theory. J. Elast. 97, 103–130. (doi:10.1007/s10659-009-9211-7) [Google Scholar]
  • 16.Zingales M. 2011. Wave propagation in 1D elastic solids in presence of long-range central interactions. J. Sound Vib. 330, 3973–3989. (doi:10.1016/j.jsv.2010.10.027) [Google Scholar]
  • 17.Schwartz M, Niane NT, Njiwa RK. 2012. A simple solution method to 3D integral nonlocal elasticity: isotropic-bem coupled with strong form local radial point interpolation. Eng. Anal. Bound. Elem. 36, 606–612. (doi:10.1016/j.enganabound.2011.10.004) [Google Scholar]
  • 18.Benvenuti E, Simone A. 2013. One-dimensional nonlocal and gradient elasticity: closed-form solution and size effect. Mech. Res. Commun. 48, 46–51. (doi:10.1016/j.mechrescom.2012.12.001) [Google Scholar]
  • 19.Abdollahi R, Boroomand B. 2013. Benchmarks in nonlocal elasticity defined by Eringen's integral model. Int. J. Solids Struct. 50, 2758–2771. (doi:10.1016/j.ijsolstr.2013.04.027) [Google Scholar]
  • 20.Abdollahi R, Boroomand B. 2014. Nonlocal elasticity defined by Eringen's integral model: introduction of a boundary layer method. Int. J. Solids Struct. 51, 1758–1780. (doi:10.1016/j.ijsolstr.2014.01.016) [Google Scholar]
  • 21.Owen DR, Paroni R. 2000. Second-order structured deformations. Arch. Ration. Mech. Anal. 155, 215–235. (doi:10.1007/s002050000111) [Google Scholar]
  • 22.DelPiero G, Owen DR. 2004. Multiscale modeling in continuum mechanics and structured deformations. International Centre for Mechanical Sciences, vol. 447. Vienna, Austria: Springer Science & Business Media. [Google Scholar]
  • 23.Owen DR, Paroni R. 2015. Optimal flux densities for linear mappings and the multiscale geometry of structured deformations. Arch. Ration. Mech. Anal. 218, 1633–1652. (doi:10.1007/s00205-015-0890-x) [Google Scholar]
  • 24.Eringen AC. 1966. A unified theory of thermomechanical materials. Int. J. Eng. Sci. 4, 179–202. (doi:10.1016/0020-7225(66)90022-X) [Google Scholar]
  • 25.Eringen AC. 1983. On differential equations of nonlocal elasticity and solutions of screw dislocation and surface waves. J. Appl. Phys. 54, 4703–4710. (doi:10.1063/1.332803) [Google Scholar]
  • 26.Eringen AC. 1987. Theory of nonlocal elasticity and some applications. Res. Mech. 21, 313–342. [Google Scholar]
  • 27.Rogula D. 1982. Introduction to nonlocal theory of material media. In Nonlocal theory of material media (ed. D Rogula). International Centre for Mechanical Sciences, vol. 268, pp. 123–222. Vienna, Austria: Springer.
  • 28.Eringen AC, Kim BS. 1977. Relation between non-local elasticity and lattice dynamics. Cryst. Lattice Defects 7, 51–57. [Google Scholar]
  • 29.Bazant ZP, Le J-L, Hoover CG. 2010. Nonlocal boundary layer (NBL) model: overcoming boundary condition problems in strength statistics and fracture analysis of quasibrittle materials. In Fracture mechanics of concrete and concrete structures–recent advances in fracture mechanics of concrete (ed. B-H Oh), pp. 135–143. Seoul, South Korea: Korea Concrete Institute.
  • 30.Goldenveizer AL, Kaplunov JD, Nolde EV. 1993. On Timoshenko–Reissner type theories of plates and shells. Int. J. Solids Struct. 30, 675–694. (doi:10.1016/0020-7683(93)90029-7) [Google Scholar]
  • 31.Dai H-H, Kaplunov J, Prikazchikov DA. 2010. A long-wave model for the surface elastic wave in a coated half-space. Proc. R. Soc. A 466, 3097–3116. (doi:10.1098/rspa.2010.0125) [Google Scholar]
  • 32.Aghalovyan L. 2014. Asymptotic theory of anisotropic plates and shells. World Scientific Publishing Co. Inc. [Google Scholar]
  • 33.Sanchez-Palencia E. 1980. Non-homogeneous media and vibration theory, Lecture Notes in Physics, vol. 127. Berlin, Heidelberg, Germany: Springer.
  • 34.Panasenko N, Bakhvalov NS. 1989. Homogenization: averaging processes in periodic media: mathematical problems in the mechanics of composite materials, Mathematics and its Applications, vol. 36. Dordrecht, The Netherlands: Springer. [Google Scholar]
  • 35.Craster RV, Kaplunov J, Postnova J. 2010. High-frequency asymptotics, homogenisation and localisation for lattices. Q. J. Mech. Appl. Math. 63, 497–519. (doi:10.1093/qjmam/hbq015) [Google Scholar]
  • 36.Birman V, Byrd LW. 2007. Modeling and analysis of functionally graded materials and structures. Applied Mechanics Reviews 60, 195–216. (doi:10.1115/1.2777164) [Google Scholar]
  • 37.Muravskii GB. 2001. Mechanics of non-homogeneous and anisotropic foundations. Berlin, Germany: Springer. [Google Scholar]

Articles from Proceedings. Mathematical, Physical, and Engineering Sciences / The Royal Society are provided here courtesy of The Royal Society

RESOURCES