Open access
Technical Papers
Sep 7, 2022

Seismic Response of Inhomogeneous Soil Deposits with Exponentially Varying Stiffness

Publication: Journal of Geotechnical and Geoenvironmental Engineering
Volume 148, Issue 11

Abstract

The response of an inhomogeneous soil layer with exponentially varying stiffness with depth is explored using one-dimensional viscoelastic wave propagation theory. The governing equation is treated analytically, leading to an exact harmonic solution of the Bessel type. Both positive and negative velocity gradients are examined using a pertinent dimensionless parameter. It is shown that (1) for positive stiffness gradients, strains attenuate with depth faster than displacements that, in turn, attenuate faster than stresses; and (2) close to the soil surface, curvatures are controlled by acceleration, whereas they are controlled by strain at depth. The fundamental natural frequency of the layer compares well against approximations on the basis of the Rayleigh quotient. Novel asymptotic and ad hoc approximate solutions for the base-to-surface transfer function are proposed, providing good alternatives to the complex exact solution at both high and low frequencies. New expressions are derived relating (1) shear strain and peak particle velocity; and (2) curvature and peak ground acceleration close to the soil surface. A full-domain approximation is provided, allowing the practical implementation of the specific velocity model. Numerical examples are presented.

Introduction

Available analytical solutions for the seismic response of continuously inhomogeneous media reveal the possibility of higher amounts of seismic energy reaching the ground surface over media with discontinuous variations in elastic properties with depth (Towhata 1996). Problems involving continuous inhomogeneity have been treated for different velocity models (Ewing et al. 1957; Ambraseys 1959; Toki and Cherry 1972; Schreyer 1977; Gazetas 1982; Dakoulas and Gazetas 1985; Aki 1993; Towhata 1996; Semblat and Pecker 2009; Paraschakis et al. 2010; Rovithis et al. 2011; Mylonakis et al. 2013; Vrettos 2013; Garcia-Suarez 2020) without necessarily acknowledging the importance of continuous versus discontinuous variations in soil material properties with depth. For the common case of a parabolic (power law) velocity model, the special condition of vanishing stiffness at the soil surface has been examined by Dobry et al. (1971), Towhata (1996), Travasarou and Gazetas (2004), Rovithis et al. (2011), and Kausel (2013). More recently, Mylonakis et al. (2013) demonstrated analytically that under viscoelastic conditions and regardless of excitation frequency and material damping, the near-surface shear strain may be either zero or infinite (but never finite) depending solely on rate of inhomogeneity. This result is in contrast to surface displacements that, naturally, are always finite.
The effect of soil inhomogeneity has been studied in multiple dimensions for different types of stress waves. Vrettos (1990) tackled the problem of dispersive horizontally polarized shear surface waves propagating in an inhomogeneous half-space for both bounded and unbounded variations in the shear modulus with depth. A collection of available solutions (not restricted to elastic waves) for wave propagation in inhomogeneous media is available in Manolis and Shaw (1997, 1999, 2000) and Brekhovskikh and Bayer (1976).
The seismic response of inhomogeneous media modeled using “equivalent” homogeneous counterparts, that is, defined by means of an average wave propagation velocity, has been explored against exact solutions (Idriss and Seed 1967; Dobry et al. 1971; Zhao 1997; Rovithis et al. 2011; Garcia-Suarez 2020). Additionally, simple analytical formulae have been proposed for the fundamental dynamic characteristics of inhomogeneous soils at resonance conditions using Rayleigh approximations (Mylonakis et al. 2013) and asymptotic analysis (Garcia-Suarez 2020; Garcia-Suarez et al. 2020).
Despite research on the subject for more than half a century, nearly all analytical solutions for inhomogeneous media are limited to cases in which stiffness and shear wave propagation velocity vary as a power of depth. Although some dependable solutions exist for other variations in stiffness (e.g., the constant minus the exponential function considered in the seminal study by Vrettos 2013), these solutions are semianalytical in nature because they are expressed in the form of an infinite power series and not special functions (e.g., of the Bessel type) associated with a wealth of mathematical properties that facilitate their analytical treatment. Closed-form analytical solutions, including relevant asymptotic expressions and approximations pertaining to low and high frequencies, can shed light on the physics of the problem, which is often not possible with other methods. The gap in knowledge regarding the dynamic behavior of more general velocity profiles provided the motivation for the herein reported work.
In this paper, the seismic response of an inhomogeneous soil layer resting on a rigid base is investigated in the realm of one-dimensional viscoelastic wave propagation theory. An exponential function is adopted, which can describe both positive and negative velocity gradients. The latter may be relevant for cases of large confining stresses under building foundations, stiff overconsolidated surface crusts, or ground improvement near the soil surface, which may result in soil stiffness decreasing with depth (Dobry et al. 1976; Vrettos 2013). For simplicity and as a first approximation, soil mass density and material damping are assumed to be constant with depth. The problem is treated analytically, leading to an exact solution expressed in terms of Bessel functions for the vibrational characteristics of the system, the base-to-surface transfer function, and the variation in ground displacements, shear strains, and curvatures with depth. The solution is compared against Rayleigh approximations for the fundamental frequency of the layer by implementing different shape functions for the corresponding mode shapes. Asymptotic and approximate solutions are proposed for the response in low and high frequencies, which are combined through a smooth transition function to provide a full-domain approximation. The model is validated against a numerical layer transfer-matrix (Haskell-Thomson) formulation for two actual soil profiles. Apart from its intrinsic theoretical interest, the proposed exact solution can be used to assess other relevant solutions for wave propagation and site response analysis.

Problem Statement

The problem considered in this work refers to a continuously inhomogeneous viscoelastic soil layer of thickness H, excited by vertically propagating seismic shear (SH) waves applied at its base (Fig. 1).
Fig. 1. Inhomogeneous soil layer with exponentially varying shear wave velocity with depth over rigid base.
Soil mass density, ρ, and hysteretic damping ratio, ξ, are assumed constant with depth, whereas the shear wave propagation velocity follows the exponential variation
V(z)=V0eαz/zr
(1)
where V0 = shear wave propagation velocity at the ground surface; zr = reference depth; and α = dimensionless inhomogeneity parameter.
Rearranging Eq. (1), parameter α can be defined as
α=ln(V0/VH)1
(2)
where VH(=V0eα) = shear wave propagation velocity at the base (zr=H). Evidently, positive and negative values of α correspond to stiffness increasing (V0<VH) and decreasing (V0>VH) with depth, respectively. For α=0, the model degenerates to a uniform velocity distribution (V0=VH).
The input motion is specified at the base of the system in the form of a harmonic horizontal displacement, u=u0exp(iωt), with ω=(2πf) as the cyclic excitation frequency, thereby generating shear waves propagating vertically up and down (on reflection) in the profile.
Evidently, the increase in the shear wave propagation velocity according to this model is superlinear (when α>0), whereas the corresponding variation in uniform normally consolidated soils is known to be sublinear (Kramer 1996; Towhata 1996). Although this discrepancy poses difficulties in simulating deep uniform soil profiles, the model can successfully describe certain classes of nonuniform soil profiles with discontinuous variations in shear wave velocity with depth via curve fitting. Examples are provided later in this paper. In addition, the model can be used to simulate arbitrary layered velocity profiles (e.g., via a transfer matrix formulation as shown in the Supplemental Materials).

Analytical Investigation

Under harmonic, steady-state oscillations, one-dimensional shear waves in an inhomogeneous soil layer are described using the familiar ordinary differential equation (Ewing et al. 1957; Towhata 1996)
ddz[G(z)dudz]+ρω2u=0
(3)
After introducing the displacement potential Y(z) through the transformation dY/dz=u(z) and integrating with respect to z, the governing equation in Eq. (3) can be rewritten in canonical form
d2dz2Y+k2(z)Y=0
(4)
in which the term associated with the first derivative of the new dependent variable, dY/dz, has disappeared, and the depth-varying shear modulus has escaped differentiation. In the latter equation
k(z)=k0eαz/zr
(5)
stands for the depth-varying wavenumber of the associated stress waves, where k0=(ω/V0) is the value at ground surface. The solution to Eq. (4) is (Wylie and Barret 1986; Trachanas 2004—see Supplemental Materials)
Y(z)=C1J0(k0zrαeαzzr)+C2N0(k0zrαeαzzr)
(6)
where J0() and N0() are the zero-order Bessel functions of the first and the second kind, respectively; and C1, C2 = integration constants to be determined from the boundary conditions.
When differentiating with depth, the general solution for displacements is obtained from Eq. (6) as
u(z)=k0eαzzr[C1J1(k0zrαeαzzr)+C2N1(k0zrαeαzzr)]
(7)
where J1() and N1() denote the corresponding Bessel functions of the first order. The associated shear stresses can be derived by inspection because τ(z)=G(z)γ(z)=G(z)du/dz=G(z)d2Y/dz2=ρω2Y to obtain
τ(z)=ρω2[C1J0(k0zrαeαzzr)+C2N0(k0zrαeαzzr)]
(8)
which reveals that the displacement potential in Eq. (6) is merely a scaled form of shear stress τ(z) on division by the product (ρω2). A relevant discussion is provided in Towhata (1996).
The boundary condition of a traction-free surface, τ(0)=0, yields a proportionality relation between constants C1 and C2 (see Supplemental Materials); the solution for displacements becomes
u(z)=C1k0eαzzr[J1(k0zrαeαzzr)N0(k0zrα)J0(k0zrα)N1(k0zrαeαzzr)]/N0(k0zrα)
(9)
From Eqs. (3) and (9), the corresponding solution for shear strain is
γ(z)=C1k02e2αzzr[J0(k0zrαeαzzr)N0(k0zrα)J0(k0zrα)N0(k0zrαeαzzr)]/N0(k0zrα)
(10)
Finally, operating on Eq. (3), one can easily establish the second derivative d2u/dz2, which expresses the curvature of soil with depth
(1/R)={2[V(z)/V(z)]γ+k2(z)u}
(11)
The last equation, which holds for any smooth velocity model, is particularly useful when shear systems, such as the one at hand, are in contact with embedded flexural structural elements, such as piles or diaphragm walls, to ensure that strains and curvatures are “transferred” from the soil to the structural elements. Relevant discussions on strain and curvature transmissibility pertaining to piles are provided in Mylonakis (2001), Anoyatis et al. (2013), Di Laora and Rovithis (2015), Di Laora et al. (2017) and in a recent review article by Di Laora and Rovithis (2021).
Eq. (11) holds for general variations in shear wave propagation velocity with depth and is not limited to a particular type of soil inhomogeneity. For the velocity model in Eq. (1), one obtains
(1/R)=[(2α/zr)γ+ko2e2αz/zru]
(12)
In light of these analytical developments, the following observations are worth noting. First, the wavenumber [koexp(αz/zr)] appearing in the arguments of the Bessel functions coincides with the variable coefficient k(z) in Eq. (4), which is not the case with other relevant solutions (e.g., Semblat and Pecker 2009; Rovithis et al. 2011; Mylonakis et al. 2013). Second, the multiplier (zr/α) in the arguments of the Bessel functions normalizes the wavenumber to produce a dimensionless argument that coincides with the derivative d/dz[k(z)]=d/dz[koexp(αz/zr)]. Again, this special property of the particular type of inhomogeneity is not encountered in other related solutions. Third, the multipliers exp(2αz/zr) and exp(αz/zr) appear, respectively, in the expressions for strains in Eq. (10) and displacements in Eq. (9), but not in the one for stresses [Eq. (8)]. For the common case of positive α, this implies that strains and displacements attenuate with depth faster than stresses, whereas strains attenuate with depth faster than displacements. This, again, is different from other basic problems in geomechanics, such as the Boussinesq point-load solution, for which stresses attenuate with depth faster than displacements (Poulos and Davis 1974; Davis and Selvadurai 1996). The slower attenuation of stresses relative to strains and displacements in the problem at hand can be interpreted in light of the direct dependence of stresses on dynamic equilibrium with the imposed inertial loads, which forces stresses to exhibit a more “statically determined” behavior than the other response functions. Fourth, for positive values of α, soil curvature (1/R) attenuates rapidly with depth, as evident from the exponential multipliers exp(2αz/zr) and exp(3αz/zr) in the first and second term in the right-hand side of Eq. (12). As the term associated with soil displacement attenuates faster than the one for strain, curvatures at depth tend to be governed by strains. Fifth, the soil curvature (1/R) in an inhomogeneous medium may be larger or smaller than the corresponding one in a homogeneous medium subjected to the same displacement and wavenumber depending on the sign of the first term in the right side of Eq. (11). Sixth, the rate of inhomogeneity, α, does not affect the order of the Bessel functions in the solution. This, again, is different from other relevant problems involving wave propagation in inhomogeneous media (Towhata 1996). Seventh, shear strain at the soil surface, γ(0), is always zero because shear stress is by definition zero, whereas the corresponding shear modulus, G(0), is finite.
For z=0, the displacement at soil surface is obtained as (see Supplemental Materials)
u(0)=C1(2απzr)/N0(k0zrα)
(13)

Natural Frequencies and Mode Shapes

Setting the reference depth zr=H and considering undamped free oscillations, the displacement at the base of the layer must be zero [u(H)=0] at all times. Enforcing this condition to Eq. (9) yields the characteristic equation
J1(knHαeα)N0(knHα)J0(knHα)N1(knHαeα)=0
(14)
the solution to which provides the (infinite) eigenvalues of the layer. Given the transcendental nature of Eq. (14), the eigenvalues need to be established numerically (approximate analytical solutions are subsequently discussed in this paper); these can be written in normalized form as
fnf1H=2πknHeαn=1,2,3,
(15)
where kn(=ωn/V0) is the eigen-wavenumber at the surface; and f1H(=VH/4H) is the fundamental natural frequency of a homogeneous layer of the same thickness as the inhomogeneous one, and shear wave propagation velocity equal to the wave velocity at the base, VH.
Normalizing Eq. (9) by u(0) in Eq. (13) and introducing the dimensionless depth η=z/H yields the dimensionless mode shape
Φ(η)=π2(knHαeαη)[J1(knHαeαη)N0(knHα)J0(knHα)N1(knHαeαη)]
(16)
Figs. 2(a–c) compare the shapes of the first, second, and third modes, respectively, of the inhomogeneous layer computed from Eqs. (14) and (16) for the case in which V0VH, corresponding to positive values of the inhomogeneity parameter α.
Fig. 2. Normalized shapes of first, second, and third natural mode of an inhomogeneous soil layer over rigid rock for V0VH (α>0), graphs (a–c) and V0VH (α<0) graphs (d–f), respectively.
As expected, for values of α close to zero (that is, for V0/VH ratios close to 1), the mode shapes closely resemble those of a homogeneous layer, approaching the ground surface at an angle almost tangent to vertical, that is, at nearly zero shear strain. In contrast, when V0/VH is less than 0.5, modal amplitudes approach the ground surface at a much shallower angle (i.e., at larger strains) and turn vertical very close to the surface. Similar results for V0VH are plotted in Figs. 2(d–f), corresponding to negative values of α. In the latter case, mode shapes approach the ground surface almost tangent to vertical, that is, at nearly zero shear strain. Note the strong amplitudes of the higher modes close to the base when V0/VH is greater than 2.

Steady-State Harmonic Response

Under harmonic base excitation, the base-to-surface transfer function is defined in terms of the familiar amplification factor (Roesset 1977; Kausel and Roesset 1984)
F(ω)u(0)/u(H)
(17)
Taking the ratio of Eqs. (13) and (9) and setting z=H, the amplification function of the inhomogeneous soil layer is readily obtained as
F(ω)=2π{(k0Hαeα)[J1(k0Hαeα)N0(k0Hα)J0(k0Hα)N1(k0Hαeα)]}1
(18)
where k0 = continuous function of excitation frequency ω. Note that soil material damping can be readily accounted for in the this expression by replacing the real wavenumber k0 with its complex counterpart k0*=k0/1+2iξ (Roesset 1977; Kramer 1996).
The analytical base-to-surface transfer function in Eq. (18) is plotted in Fig. 3 for V0VH (α>0), top graph, and V0VH (α<0), bottom graph, with the abscissa normalized by the exact fundamental frequency of the inhomogeneous layer, f1, obtained numerically from Eq. (14).
Fig. 3. Effect of V0/VH ratio on base-to-surface transfer function of the inhomogeneous layer for (a) V0VH; and (b) V0VH. In all plots, ξ=0.05.
All of the plots in Fig. 3 refer to a constant hysteretic damping ratio ξ=0.05. Evidently, the spacing between successive resonant peaks for V0VH gets “compressed” (i.e., the ratios fn/f1 become smaller than in a homogeneous layer) with increasing inhomogeneity. Additionally, the amplification amplitude at a particular resonant peak increases relative to that of a more homogeneous layer. The opposite is true for the case in which V0VH; the natural frequencies tend to become “stretched” relative to f1, and their amplitudes decrease relative to those of a more homogeneous layer. In all cases, the amplitudes of successive resonance peaks decrease with frequency. For the first three resonance frequencies, the distribution with the depth of normalized displacements and their derivatives in an inhomogeneous layer with V0/VH=0.25 is provided in Fig. S1. The observed trends are self-explanatory.

Rayleigh Approach

The fundamental natural frequency of the system can be well approximated by the familiar Rayleigh quotient (Semblat and Pecker 2009; Mylonakis et al. 2013)
ω2=0HG(z)(u(z))2dz/0Hρ(z)u2(z)dz
(19)
This expression is derived in the Supplemental Materials. By replacing u(z) by a unitary dimensionless shape function, ψ(z), representing the fundamental mode shape of the inhomogeneous layer, Φ1(z), considering constant mass density and employing the normalized shear wave propagation velocity V¯s(η)=Vs(η)/VH=e+α(η1), the Rayleigh quotient can be rewritten in the normalized form
f1/f1H=(2/π)[01V¯s2(η)(ψ(η))2dη/01ψ2(η)dη]1/2
(20)
Following Mylonakis et al. (2013), four shape functions are examined (Table 1): linear, parabolic, sinusoidal, and “compatible.” The last function is obtained by imposing a distributed horizontal static load proportional to the soil mass along the height of the column (see Supplemental Materials). The Rayleigh solutions are compared against the numerically evaluated exact solution in Fig. 4 for different values of the inhomogeneity factor α. Specific values of the (f/f1  H) ratio are provided in Table 2, whereas the exact fundamental mode shape in Eq. (16) is compared with these shape functions in Figs. 5(a–c) and (d–f) for Vo/VH ratios lower and higher than unity, respectively.
Table 1. Rayleigh expressions for fundamental natural frequency using different shape functions
Shape functionψ(η)Fundamental frequency (f1/f1H)
Linear1η(2/π)[(3/2α)(1e2α)]1/2
Parabolic1η2(2/π)[(15/8α3)(1+2α22αe2α)]1/2
Sinusoidalcos(π2η)[(π2(1e2α)+8α2)/(2α(4α2+π2))]1/2
Compatible[(2α+1)e2α(2αη+1)e2αη]/[(2α+1)e2α1]2π[32α2(e2α2α22α1)/s]1/2s=32α3+56α2+44α16(2α+1)e2α+5e4α+11
Fig. 4. Normalized fundamental frequency of the inhomogeneous layer as function of inhomogeneity parameter α. Comparison of the exact solution with approximate Rayleigh solutions using Eq. (20); ξ=0.
Table 2. Comparison of specific values of (f/f1H) ratio of inhomogeneous layer between exact solution and approximate Rayleigh quotients
V0/VHαExact solutionRayleigh approximation for different shape functions ψ(η)
CompatibleLinearSinusoidalParabolic
0.12.3050.35880.37110.51130.60390.6596
0.251.3860.58880.60070.64120.71130.7569
0.50.6930.79160.80110.81100.82940.8606
0.750.2870.9140.92160.96150.92120.9396
0.90.1050.96840.97561.04700.96960.9808
101a1.00661.10271.00001.0066
1.50.4051.11881.12411.36901.13631.1212
20.6931.20081.20531.62211.25551.2205
31.0981.31281.31612.10401.46491.3936
51.6091.44641.44923.01091.82191.6863

Note: ψ(η) refers to the shape functions listed in Table 1.

a
A value of 0.9928 was obtained for V0/VH=0.99 (or α=0.01). Due to instabilities associated with the numerical evaluation of the relevant Bessel functions, no valid numerical results could be obtained for V0/VH ratios closer to 1.
Fig. 5. Comparison of exact fundamental mode shape for (a) V0/VH=0.25; (b) V0/VH=0.5; (c) V0/VH=0.75; (d) V0/VH=1.5; (e) V0/VH=2; and (f) V0/VH=3 with shape functions under consideration; ξ=0.
Note that in the special case in which α0, the compatible shape function in Table 1 reduces to the parabolic one. The normalized natural frequency for the linear shape is then obtained as (12)1/2/π1.1027; the parabolic and stiffness compatible shapes yield (10)1/2/π1.0066, whereas the sinusoidal shape yields the exact value of 1 (Kloukinas et al. 2012). These results are provided in Table 2.

Asymptotic and Approximate Solutions

Asymptotic Behavior at High Frequencies

Employing the pertinent asymptotic expressions of the Bessel functions for large arguments (see Supplemental Materials) shows that the amplification function in Eq. (18) simplifies to
Fh(ω)eα/2/cos[k0H(1eα)/α]
(21)
Note that the first term on the right side of this equation, ea/2, is merely the velocity ratio (Vo/VH)1/2. This asymptotic behavior is a general property of wave propagation in inhomogeneous media and is not limited to the specific velocity profile (Aki and Richards 2002; Garcia-Suarez et al. 2020).
This expression can be used to provide reasonable approximations of the higher resonant frequencies of the inhomogeneous stratum through the characteristic equations
cos[k0H(1eα)/a]=0
(22a)
cos[kHH(eα1)/a]=0
(22b)
the solution to which is
(kHH)k(eα1)/a=(2k1)π/2,k=1,2,3,
(23)
where kH(=ω/VH) is the wavenumber at the base of the layer.
In light of Eq. (23), the following equivalent shear wave propagation velocity in the high-frequency range, Veq,h, can be defined
Veq.h=VH  α/(eα1)
(24)
Note that in the limit case in which α approaches zero, this expression duly reduces to that of the homogeneous case (i.e., Veq,h=VH).
The asymptotic solution in Eq. (21) is compared with the exact one [Eq. (18)] in Fig. 6 for different V0/VH ratios. In the abscissa, the excitation frequency is normalized by the exact fundamental frequency, f1, of the inhomogeneous layer. The agreement between the two solutions beyond the first resonance highlights the validity of the asymptotic solution at high frequencies. Evidently, Eq. (21) allows a reasonable approximation of the high-frequency response of the inhomogeneous layer for frequencies beyond approximately one to two times the fundamental natural frequency, f1. Beyond the second resonance (i.e., two to four times f1), the agreement between the results from Eq. (21) and the exact solution is nearly perfect.
Fig. 6. Comparison of base-to-surface amplification function between exact solution [Eq. (18)], high-frequency approximation [Eq. (21)], and low-frequency approximation [Eq. (30)] for V0/VH ratio at (a) 0.25; (b) 0.5; (c) 3; and (d) 5; ξ=0.05, ξeq.l from Eq. (32).
Alternatively, Veq.h in Eq. (24) can be viewed as the shear wave propagation velocity providing equal base-to-surface travel times between the homogeneous and inhomogeneous soil
Veq.h=H[0H1Vs(z)dz]1
(25)
For the velocity profile at hand, Eq. (25) is shown to be equivalent to Eq. (24), which indicates that a homogeneous layer with shear wave velocity Veq.h can indeed model the actual inhomogeneous layer at high frequencies [provided that the multiplier eα/2 in the right side of Eq. (21) is considered]. Similar findings have been reported in Zhao (1997), Rovithis et al. (2011), Mylonakis et al. (2013), and Garcia-Suarez et al. (2020) for different types of inhomogeneity.
In contrast, given the presence of the multiplier eα/2, the amplification function in Eq. (21) does not satisfy the static condition F(0)=1. Evidently, because of the difference in the arguments of Bessel functions of order 0 and 1 in Eq. (18), defining a single “equivalent” wave propagation velocity, applicable to all frequencies, is not possible (a full-domain approximation is subsequently discussed in this paper).
In light of Eq. (23), the spacing between two successive dimensionless natural frequencies of the inhomogeneous layer can be well approximated by
(kHH)k+1(kHH)k=πα/(eα1),k=1,2,3
(26)
This expression is plotted in Fig. 7 against the (V0/VH) ratio. The exact numerical solution for k=1 and 2 is also plotted for comparison. For the homogeneous case (V0=VH), the spacing between any pair of successive dimensionless frequencies equals π. For the inhomogeneous case and positive values of α (i.e., V0/VH<1), the spacing drops below π with decreasing V0/VH, which explains the progressively “compressed” behavior of natural frequencies with increasing inhomogeneity. Conversely, for negative values of α (i.e., V0/VH>1), the spacing rises above unity and increases with increasing V0/VH, which explains the progressively “stretched” behavior of natural frequencies with increasing inhomogeneity. The predictions of Eq. (26) naturally improve at higher resonances (k=3).
Fig. 7. Effect of V0/VH ratio on spacing between two successive dimensionless natural frequencies of the inhomogeneous layer in light of approximate solution in Eq. (26).

Strain-Velocity and Curvature-Acceleration Relations

Using the pertinent asymptotic expansions of the relevant Bessel functions for large arguments (see Supplemental Materials), it is straightforward to show that, at high frequencies, strain amplitude at depth is related to particle velocity at the surface through the equation
γ(z)u˙(0)/V(z)
(27)
which is an extension of the familiar relation for homogeneous soil (Kramer 1996). Likewise, operating on Eq. (12), the curvature at shallow depths can be related to particle acceleration at the surface through
1/R(z)u¨(0)/V2(z)
(28)
Remarkably, the last two equations are related through the expression 1/R(z)=k(z)γ(z) that, again, is a generalized form of available solutions for homogeneous soil (Di Laora et al. 2017; Di Laora and Rovithis 2021).
In contrast, the curvature at depth is related to corresponding shear strain via the equation
1/R(z)2[V(z)V(z)]γ(z)=2γ(z)a/H
(29)
which indicates that the absolute thickness of the layer influences curvatures deep beneath the soil surface. Eq. (29) has no counterpart in available solutions for homogeneous soil.

Low-Frequency Approximation and “Equivalent” Depth

The following approximate relation is proposed for the amplification function [Eq. (18)] at low frequencies
Fl(ω)1/cos(keq.lH)
(30)
which satisfies the limit condition F(0)=1. In this equation, (keq.lH) is a dimensionless wavenumber that reproduces the first resonance. The corresponding wave velocity can be determined using the Rayleigh approach for the compatible shape function (Table 1) by means of the equation
Veq,l=4f1,comH
(31)
where f1,com is the fundamental frequency of the layer. Given that the corresponding resonant peak in an inhomogeneous layer has a different amplitude than in a homogeneous one, the following empirical correction, back-calculated from the numerical data, is proposed for the equivalent damping ratio
ξeq.l(10.08α)ξ
(32)
where ξ is the actual soil material damping. No such correction is needed for higher resonances. The approximate solution in Eq. (30) is plotted in Fig. 6 for comparison. The agreement with the exact solution cannot be overstated. It is stressed that this low-frequency approximation is valid near the first resonance. If Eqs. (30)–(32) are implemented for frequencies past approximately two times the fundamental natural frequency f1, they may induce large errors depending on the characteristics of the layer.
The normalized “equivalent” depth (zeq/H) (Idriss and Seed 1967; Dobry et al. 1976) can be derived in a straightforward manner using Eqs. (1), (24), and (30) to obtain the exact solutions
zeq,h/H=ln[α/(eα1)]/α+1
(33a)
zeq,l/H=ln(f1,com/f1H)/α+1
(33b)
for the high- and low-frequency ranges, respectively.
These expressions are plotted in Fig. 8 against the V0/VH ratio. In the limit case in which α approaches zero, the equivalent depth predicted from Eq. (33) exhibits a singular behavior. However, in this case, the definition of an equivalent depth is of lesser importance because the inhomogeneous layer degenerates to a homogeneous one and, thus, zeq can be designated at any depth. Interestingly, regardless of V0/VH ratio, zeq,h/H is always lower than zeq,l/H, which indicates that a high-frequency response of the inhomogeneous soil is controlled by the soil stiffness close to the surface, whereas a low-frequency response is controlled by the soil stiffness at depth. These observations are in meaningful agreement with recent findings by Garcia-Suarez et al. (2020) and Garcia-Suarez and Asimaki (2020) for a parabolic variation of soil stiffness with depth.
Fig. 8. Normalized equivalent depth against V0/VH ratio in light of Eqs. (33a) and (33b).

Full-Domain Approximation

On the basis of the aforementioned amplification functions, a full-domain approximation is possible according to the connection formula
F(ω)Fl(ω)q(ω)+Fh(ω)[1q(ω)]
(34)
where q(ω) is a smooth transition function. The following parametric function is adopted (Abramowitz and Stegun 1965)
q(ω)1/{1+[k0H/(χ1eχ2αα)]χ3}
(35)
The vector of the fitted parameters (χ1,χ2,χ3) was established numerically at (2.5, 0.25, 20) for V0/VH<1 and (7.5, 1.4, 20) for V0/VH>1, respectively. The full-domain approximation is successfully compared against the exact solution in Fig. 9.
Fig. 9. Comparison of base-to-surface amplification function between exact solution [Eq. (18)] and full-domain approximation in Eq. (34) for V0/VH ratio at (a) 0.25; (b) 0.5; (c) 3; and (d) 5; ξ=0.05.

Case Studies: Analysis of Two Actual Soil Profiles

Two actual profiles with discontinuous variation in shear wave propagation velocity with depth were employed to assess the proposed solutions against a numerical simulation of soil response using a series of a piecewise homogeneous layers, implemented through the classical Haskell-Thompson algorithm that involves a sequence of transfer-matrix multiplications (Thompson 1950; Roesset 1977). To this end, a real soil profile (hereafter referred to as “Profile A”) was selected from the Kiban-Kyoshin Network (KiK-net) database (Okada et al. 2004; NIED 2019) in the Tagahaki site in Ibarakiken Prefecture, where the IBRH13 station of the Kik-net accelerometric network is installed. This site is classified as “Level Ground” (LG), according to the taxonomy proposed in Thompson et al. (2012), which allows for the application of one-dimensional wave propagation theory in a site response analysis. The shear wave propagation velocity profile is plotted in Fig. 10(a) up to an elevation of 34  m, where a rigid base is assumed following the definition of seismic bedrock in EC8.
Fig. 10. Discontinuous Vs profiles employed (a) “Profile A” reported in NIED (2019) for station IBRH13, KiK-net accelerometric network; and (b) San Francisco Bay area deposit–“Profile B,” reported in Gazetas and Dobry (1984).
The second soil profile (hereafter referred to as “Profile B”) is a multilayer deposit typical of the San Francisco Bay Area, resting on bedrock at a depth of approximately 60 m (Gazetas and Dobry 1984), Fig. 10(b). The available data, as reported in these references, are provided in Tables 1 and 2 of the Supplemental Materials. In both case studies, the soil mass density and hysteretic damping ratio were assumed constant with depth, equal to 2  μg/m3 and 5%, respectively. The exponential function velocity model was established by employing a least squares fit. In this manner, the vector (V0, V0/VH) was established at approximately (197  m/s, 0.303) and (134  m/s, 0.265) for Profiles A and B, respectively. Accordingly, the dimensionless inhomogeneity parameter α equals 1.194 and 1.326 in the two cases; the exponential velocity profile is given by (Fig. 10)
V(z)=197e+0.0351z[inm/s]
(36a)
V(z)=134e+0.0221z[inm/s]
(36b)
Worth mentioning is that the Vs data reported in the Kik-net and other databases are typically crude approximations of the actual soil properties. Nevertheless, the intention of this study is to validate the proposed approximate/asymptotic solutions against “exact” solutions for a layered soil rather than evaluate the level of approximation associated with the databases. Studies encompassing uncertainty in site response analysis can be found in Stewart et al. (2017), Stewart and Afshari (2021), and de la Torre et al. (2022).

Frequency-Domain Response

The exact base-to-surface transfer function computed for these exponential models [Eq. (18)] and the prediction of the full-domain approximation [Eq. (34)] are compared with the reference numerical Haskell-Thompson solution in Fig. 11 for each case under consideration. The close agreement of the analytical solutions with the numerical one, which is evident especially in the low-to-intermediate frequency range up to 5 Hz, highlights the applicability of these expressions when the velocity profile of a real soil deposit can be approximated by an exponential function. In further support of this, the first three resonant frequencies and their amplitudes are reported in Table 3, which are separated from the exact values by less than 10%. A code-based approximation, referring to the response of a 30-m thick homogeneous layer with shear wave velocity equal to Vs30, as defined in EC8, is also shown in Fig. 11 and Table 3. The large discrepancies observed between the Vs30 approximation and the more rigorous solutions beyond the first resonance indicate the inability of such a simplified approach to capture the actual response of inhomogeneous soil. Thus, such solutions should be employed with caution, especially for deep soil deposits excited by high-frequency motions.
Fig. 11. Comparison of amplification functions using exact solution for the actual (layered) soil profile, exact solution in Eq. (18) and full-domain approximation in Eq. (34) for exponential model, and solution for homogeneous layer using Vs30 approximation: (a) Profile A; and (b) Profile B; ξ=0.05.
Table 3. Comparison of resonant frequencies and corresponding steady-state amplitudes between Haskell-Thompson technique, exact solution of exponential fit, full-domain approximation, and Vs30 approximation for the real soil profiles “A” and “B” under consideration
ProfileVs profile1st resonance2nd resonance3rd resonance
f1 (Hz)A1f2 (Hz)A2f3 (Hz)A3
AMultilayer (Haskell–Thompson solution)3.1113.838.246.2413.374.17
Exponential fit [Eq. (18)]3.0814.377.736.9012.594.33
Full-domain approximation [Eq. (34)]3.1313.737.457.6912.444.53
Homogeneous profile (Vs30 approximation)2.7912.738.374.2413.952.55
BMulti-layer(Haskell–Thompson solution)1.2314.772.917.054.873.57
Exponential fit [Eq. (18)]1.2714.523.167.225.134.58
Full-domain approximation [Eq. (34)]1.3014.023.028.205.034.84
Homogeneous profile (Vs30 approximation)1.4512.734.364.247.272.55
Worth noting is that application of the compatible shape function in Table 1 to the specific examples yields fundamental natural frequency estimates f1,com equal to 3.13 Hz and 1.29 Hz for Profiles A and B, respectively (Table 4). These values compare well with the exact solutions of 3.08 Hz and 1.27 Hz (a maximum discrepancy of only 1.5%). The associated (f1,com/f1H) ratios highlight the effect of inhomogeneity in reducing the fundamental frequency of the profile relative to a homogeneous layer of equal shear wave propagation velocity at the base.
Table 4. Estimation of first three natural frequencies for the exponential fit of the real soil profiles “A” and “B” using stiffness-compatible shape function (Table 1, f1,com) and high-frequency asymptotic expression [Eq. (26), f2 and f3]
Profilef1H=VH/4H (Hz)f1,com/f1Hf1,com (Hz)f2 (Hz)f3 (Hz)
A4.770.6553.137.4412.4
B2.090.6171.293.015.02
Regarding the higher modes, the application of Eq. (26) yields f2=34α(VH/H(eα1)) and f3=(5/3)f2, which are also reported in Table 4. Both estimates are in meaningful agreement with the exact values in Table 3.
With reference to Profile A, the shear strain and curvature profiles with depth are further explored in Fig. 12 in light of Eqs. (27)–(29). The approximate relations are compared with the exact solution for a unit ground surface acceleration at a frequency f=12.59  Hz corresponding to the third resonance. It is observed that Eq. (27) (plotted for η=z/H>0.4) provides somewhat unconservative prediction of shear strain at depth. With reference to curvature, the general solution in Eq. (11) duly reduces to Eq. (28) at z=0, where the shear strain at the soil surface, γ(0), is zero. Thus, Eq. (28) is applicable regardless of the type of inhomogeneity. As previously mentioned, this property is particularly important when the soil interacts with embedded structural elements, a topic beyond the scope of the paper. Further, Eq. (29) provides a good approximation of curvature at depth, which reflects the dominant contribution of strain in this case. Similar trends are observed for Profile B (not shown here; see the Supplemental Materials).
Fig. 12. Comparison of harmonic shear strains (a) and curvatures (b) with depth between exact and approximate solutions in Eqs. (27)–(29) for Profile A; u¨(0)=1  m/s2, f=f3=12.59  Hz, ξ=0.05.

Time-Domain Response

The predictions of the approximate solutions were further examined in the time domain by implementing these base-to-surface transfer functions F(ω) (Fig. 11) to derive the motion at the ground surface, as(t), for an input motion, ar(t), imposed on bedrock. The transition from the frequency- to the time-domain response is carried out using the familiar operation
as(t)=iFFT[F(ω)×FFT(ar(t))]
(37)
where the symbols FFT and iFFT stand for the Fast Fourier transform and its inverse transform, respectively. To this end, two acceleration records from the M7.6 ChiChi earthquake [Motion #1 in Fig. 13(a)] recorded on August 20, 1999 and the M6.7 Northridge earthquake [Motion #2 in Fig. 13(b)] recorded on October 17, 1994 were selected from the PEER Strong Motion Database (Ancheta et al. 2013) and considered as input motion ar(t) at the base of the soil profiles. Because the amplitude effects and the associated nonlinear response lie beyond the scope of this study (e.g., the elastic soil properties under consideration can be interpreted as either low-strain or strain-compatible quantities), the amplitudes of both motions are scaled to unity. These two records cover a wide range of frequencies for earthquake engineering applications, as reflected in the 5%-damped elastic acceleration spectra normalized by Peak Rock Acceleration (PRA), shown in Fig. 13(c). The seismic response at the ground surface computed using Eq. (37) is shown in Fig. 14, which is plotted in terms of the 5%-damped acceleration spectra. Each plot refers to a different combination of the soil profile and input motion: Figs. 14(a and b) present results for “Profile A” excited by Motions #1 and #2, respectively, whereas Figs. 14(c and d) provide corresponding results for “Profile B.” Evidently, the spectral accelerations derived from the exponential fit using either the exact solution or the full-domain approximation compare well with the numerical solutions for the real multilayer profiles. Lastly, the good performance of the Vs30-based approximation for the shallow profile under low-frequency excitation [Fig. 14(a)] and the poor performance for the other three cases [especially those in Figs. 14(b and c)] are in accordance with the discrepancies observed in the frequency domain.
Fig. 13. Earthquake records selected as input motions at base of the examined soil profiles: (a) acceleration time history from M7.6 ChiChi earthquake (Motion #1); (b) acceleration time history from M6.7 Northridge earthquake (Motion #2); and (c) 5%-damped elastic response spectra.
Fig. 14. Comparison of 5%-damped elastic response spectra computed at ground surface of (a) Profile A under base Motion #1; (b) Profile A under base Motion #2; (c) Profile B under base Motion #1; and (d) Profile B under base Motion #2.

Conclusions

An analytical solution was derived for the dynamic response of a soil layer with stiffness varying exponentially with depth. The main conclusions drawn from the study are as follows:
1.
The dimensionless argument appearing in the wave solution equals the derivative with depth of the wavenumber k(z) in the normal form of the governing equation [Eq. (4)], that is, d/dz[k(z)]=d/dz[koexp(az/zr)]. This is not the case with other relevant solutions for inhomogeneous media and can be viewed as a special property of the specific type of inhomogeneity. Additionally, the rate of inhomogeneity, a, does not affect the order of the Bessel functions that, again, differs from available solutions pertaining to parabolic variations in soil stiffness with depth.
2.
The multipliers exp (2az/zr) and exp(az/zr) of the Bessel functions in the expressions for strains [Eq. (10)] and displacements [Eq. (9)] do not appear in the expression for stresses [Eq. (8)]. This implies that strains and displacements attenuate with depth much faster than stresses, whereas strains attenuate with depth faster than displacements.
3.
The slower attenuation of stresses with depth relative to displacements and strains could be interpreted in light of the direct dependence of stresses on the dynamic equilibrium that forces them to exhibit a more “statically determined” behavior than the other response functions.
4.
The soil curvature (1/R) in an inhomogeneous medium may be larger or smaller than the curvature in a homogeneous medium experiencing the same displacements and under the same wavenumber depending on the difference in signs between the first and second terms on the right side of Eq. (11).
5.
For positive values of the inhomogeneity parameter α, soil curvature (1/R) attenuates fast with depth, as is evident by the exponential multipliers in the first and second terms on the right-hand side of Eq. (12). Because the associated displacement term attenuates with depth faster than that of strain, the curvature at depth tends to be governed by the strain component.
6.
For values of the inhomogeneity parameter α close to zero (or, equivalently, for Vo/VH ratios close to 1), the mode shapes closely resemble those of a homogeneous deposit, approaching the ground surface at a nearly zero shear strain. In contrast, when Vo/VH is less than 0.5, modal amplitudes approach the ground surface at much higher strains that become zero only close to the surface. When VoVH, the mode shapes approach the ground surface almost tangent to vertical, that is, at nearly zero shear strain. The amplitude of the higher modes is particularly pronounced close to the base when Vo/VH is higher than 2.
7.
With reference to the base-to-surface transfer function in Eq. (18), for V0VH (α>0), the frequencies of resonant peaks tend to be “compressed” (i.e., the ratios fn/f1 become smaller than in the homogeneous layer) with increasing inhomogeneity. Additionally, the amplitude of a resonant peak increases relative to the same peak in a more homogeneous layer.
8.
Following up on the previous conclusion, the opposite is true for the case in which V0VH: the frequencies of the resonant peaks tend to get “stretched” relative to the f1 with increasing inhomogeneity. Additionally, their amplitudes tend to decrease relative to that of a less inhomogeneous layer. In all cases, the peak amplitudes at successive resonance peaks are a decreasing function of frequency.
9.
The compatible shape function in Table 1 provides excellent predictions of the fundamental natural frequency of the soil layer using the Rayleigh quotient [Eq. (20)]. In the special case where α0, the compatible shape function collapses into the parabolic one in Table 1.
10.
The asymptotic and approximate solutions in Eqs. (21) and (30) provide good approximations of the high- and low-frequency behavior of the amplification function. To match the response amplitude at first resonance (which is higher than in a homogenous layer for Vo<VH and lower for Vo>VH), an empirical damping adjustment has been proposed in Eq. (32).
11.
Eq. (27) provides realistic yet somewhat unconservative predictions of strain at depth and/or high frequencies. In contrast, Eq. (28) provides exact curvature values at the ground surface. This is true regardless of the specific velocity profile, provided that the shear strain at the surface, γ(0), is zero. This property has important implications in soil-structure interaction, such as in the kinematic bending of piles. Eq. (29) provides a good approximation of the curvature at depth, which reflects the dominant contribution of the soil strain in this case.
12.
The full-domain approximation proposed in Eqs. (34) and (35) provides good predictions of the exact solution in Eq. (18) and allows the exponential velocity model in applications in both frequency and time domains. Lastly, the Vs30 approximation performs well in certain cases [e.g., the shallow profile under the low-frequency excitation in Fig. 14(a)] but poorly in others, such as in deep soil profiles.
As a final remark, it is fair to mention that the increase in wave velocity with depth in the model at hand is superlinear (when α>0), whereas the corresponding variation in uniform, normally consolidated soils is sublinear. This discrepancy inevitably poses difficulties in simulating deep uniform soil profiles; however, the model can successfully describe certain classes of nonuniform soil profiles with discontinuous variation in shear wave velocity with depth via curve fitting (Fig. 10). In addition, the solution at hand can be readily extended to encompass an arbitrary number of inhomogeneous layers or a combination of stacked homogeneous and inhomogeneous layers by means of multiplications of transfer matrices. This is outlined in the Supplemental Materials.

Notation

The following symbols are used in this paper:
C1, C2
integration constants;
f
excitation frequency;
f1, fn
fundamental natural frequency (1st eigenfrequency), nth eigenfrequency;
f1H
fundamental natural frequency of homogeneous layer of thickness H and shear wave propagation velocity VH;
f1,com
fundamental natural frequency based on Rayleigh quotient for stiffness-compatible shape function;
F(ω)
top-to-base amplification function;
Fh(ω)
top-to-base amplification function for high-frequency range;
Fl(ω)
top-to-base amplification function for low-frequency range;
G(z)
shear modulus at depth z;
G0, G(0)
shear modulus at ground surface;
g
acceleration of gravity;
H
thickness of soil layer;
Jν()
Bessel function of first kind and order v;
k(z)
depth-varying wavenumber;
k0
real wavenumber at ground surface;
k0*
complex (damped) wavenumber at ground surface;
kn
nth eigen-wavenumber at ground surface;
kH
real wavenumber at base of layer;
l
step of power series solution;
Nν()
Bessel (Newmann) function of second kind and order v;
q(ω)
transition function;
1/R(z)
soil curvature at depth z;
s
exponent of monomial solution;
u(z)
soil displacement at depth z;
u*(z)
arbitrary (trial) displacement function of depth;
V(z)
shear wave propagation velocity of soil at depth z;
V0
shear wave propagation velocity at ground surface;
VH
shear wave propagation velocity at base of layer;
V¯s(η)
normalized shear wave propagation velocity;
Veq.h
equivalent shear wave propagation velocity for high-frequency range;
Veq,l
equivalent shear wave propagation velocity for low-frequency range;
x
dimensionless depth variable;
Y(z)
displacement potential;
zr
reference depth;
zeq,h
equivalent depth for high-frequency range;
zeq,l
equivalent depth for low-frequency range;
α
dimensionless inhomogeneity parameter;
γ(z)
shear strain at depth z;
η
dimensionless depth;
λ
parameter associated with asymptotic behavior of solution at infinity;
μ
exponent of monomial multiplier in general solution;
ξ
hysteretic damping ratio;
ξeq.l
equivalent hysteretic damping ratio for low-frequency range;
ρ
mass density;
τ(z)
shear stress at depth z;
Φ(η)
dimensionless mode shape;
Φ1(z)
fundamental mode shape;
χ1, χ2, χ3
regression constants;
ψ(η)
dimensionless shape function;
ω
cyclic frequency of excitation; and
ωn
nth eigen cyclic frequency.

Supplemental Materials

File (supplemental materials_gt.1943-5606.0002856_rovithis.pdf)

Data Availability Statement

All data and models generated or used during the study appear in the published article. All code that supports the findings of this study are available from the corresponding author on reasonable request.

Acknowledgments

The authors acknowledge the assistance of Ms. Vasiliki Miha and Ms. Kathleen Schulze, former students of the senior author at the University of Patras and the City University of New York, respectively, during the early stages of this research. Thanks are also due to Dr. Charalambos Paraschakis for fruitful discussions and to Dr. Elia Voyagaki for proofreading the manuscript. Partial funding was received by EU/H2020 under grant agreement number 730900 (SERA) with George Mylonakis as Principal Investigator for University of Bristol.

References

Abramowitz, M., and I. Stegun. 1965. Handbook of mathematical functions with formulas, graphs, and mathematical tables. New York: Dover.
Aki, K. 1993. “Local site effects on weak and strong ground motion.” Tectonophysics 218 (1–3): 93–111. https://doi.org/10.1016/0040-1951(93)90262-I.
Aki, K., and P. G. Richards. 2002. Quantitative seismology. 2nd ed. Sausalito, CA: Univ. Science Bks.
Ambraseys, N. 1959. “A note on the response of an elastic overburden of varying rigidity to an arbitrary ground motion.” Bull. Seismol. Soc. Am. 49 (3): 211–220. https://doi.org/10.1785/BSSA0490030211.
Ancheta, T., et al. 2013. PEER NGA-West2 database. Berkeley, CA: Univ. of California.
Anoyatis, G., R. Di Laora, A. Mandolini, and G. Mylonakis. 2013. “Kinematic response of single piles for different boundary conditions: Analytical solutions and normalization scheme.” Soil Dyn. Earthquake Eng. 44 (Jan): 183–195. https://doi.org/10.1016/j.soildyn.2012.09.011.
Brekhovskikh, L. M., and R. Bayer. 1976. Waves in layered media. New York: Academic Press.
Dakoulas, P., and G. Gazetas. 1985. “A class of inhomogeneous shear modes for seismic response of dams and embankments.” Soil Dyn. Earthquake Eng. 4 (4): 166–182. https://doi.org/10.1016/0261-7277(85)90037-3.
Davis, R. O., and A. P. S. Selvadurai. 1996. Elasticity and geomechanics. Cambridge, UK: Cambridge University Press.
de la Torre, C., B. A. Bradley, C. R. McGann, and J. P. Stewart. 2022. “Can modeling of soil heterogeneity in 2D site response analyses improve predictions at vertical array sites?” Earthquake Spectra.
Di Laora, R., G. Mylonakis, and A. Mandolini. 2017. “Size limitations for piles in seismic regions.” Earthquake Spectra 33 (2): 729–756. https://doi.org/10.1193/032116EQS045M.
Di Laora, R., and E. Rovithis. 2015. “Kinematic bending of fixed-head piles in nonhomogeneous soil.” J. Geotech. Geoenviron. Eng. 141 (4): 04014126. https://doi.org/10.1061/(ASCE)GT.1943-5606.0001270.
Di Laora, R., and E. Rovithis. 2021. Design of piles under seismic loading, analysis of pile foundations subject to static and dynamic loading, 269–300. Boca Raton, FL: CRC Press.
Dobry, R., I. Oweis, and A. Urzua. 1976. “Simplified procedures for estimating the fundamental period of a soil profile.” Bull. Seismol. Soc. Am. 66 (4): 1293–1321.
Dobry, R., R. Whitman, and J. M. Roesset. 1971. Soil properties and the one-dimensional theory of earthquake amplification. M.I.T.. Cambridge, MA: Massachusetts Institute of Technology.
Ewing, W. M., W. S. Jardetzky, and F. Press. 1957. Elastic waves in layered media. New York: McGraw-Hill.
Garcia-Suarez, J. 2020. “Application of path-independent integrals to soil-structure interaction.” Ph.D. dissertation, Div. of Engineering and Applied Science, California Institute of Technology.
Garcia-Suarez, J., and D. Asimaki. 2020. “On the fundamental resonant mode of inhomogeneous soil deposits.” Soil Dyn. Earthquake Eng. 135 (Aug): 106190. https://doi.org/10.1016/j.soildyn.2020.106190.
Garcia-Suarez, J., E. Esmaeilzadeh, and D. Asimaki. 2020. “Seismic harmonic response of inhomogeneous soil: Scaling analysis.” Géotechnique 71 (5): 392–405.https://doi.org/10.1680/jgeot.19.P.042.
Gazetas, G. 1982. “Vibrational characteristics of soil deposits with variable wave velocity.” Int. J. Numer. Anal. Methods Geomech. 6 (1): 1–20. https://doi.org/10.1002/nag.1610060103.
Gazetas, G., and R. Dobry. 1984. “Horizontal response of piles in layered soils.” J. Geotech. Eng. 110 (1): 20–40. https://doi.org/10.1061/(ASCE)0733-9410(1984)110:1(20).
Idriss, I. M., and H. B. Seed. 1967. “Response of horizontal soil layers during earthquakes.” In Research report, soil mechanics and bituminous materials laboratory. Berkeley, CA: Univ. of California.
Kausel, E. 2013. “On the frequencies of inhomogeneous soil strata: Dobry’s paradox.” Soil Dyn. Earthquake Eng. 47 (Apr): 38–40. https://doi.org/10.1016/j.soildyn.2012.09.015.
Kausel, E., and J. M. Roesset. 1984. “Soil amplification: Some refinements.” Soil Dyn. Earthquake Eng. 3 (3): 116–123. https://doi.org/10.1016/0261-7277(84)90041-X.
Kloukinas, P., M. Langousis, and G. Mylonakis. 2012. “Simple wave solution for seismic earth pressures on non-yielding walls.” J. Geotech. Geoenviron. Eng. 138 (12): 1514–1519. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000721.
Kramer, L. S. 1996. Geotechnical earthquake engineering, 653. Englewood Cliff, NJ: Prentice Hall.
Manolis, G. D., and R. P. Shaw. 1997. “Harmonic elastic waves in continuously heterogeneous random layers.” Eng. Anal. Boundary Elem. 19 (3): 181–198. https://doi.org/10.1016/S0955-7997(97)00005-2.
Manolis, G. D., and R. P. Shaw. 1999. “Numerical simulation of transient waves in a heterogeneous soil layer.” Comput. Mech. 23 (1): 75–86. https://doi.org/10.1007/s004660050386.
Manolis, G. D., and R. P. Shaw. 2000. “Fundamental solutions for variable density two-dimensional elastodynamic problems.” Eng. Anal. Boundary Elem. 24 (10): 739–750. https://doi.org/10.1016/S0955-7997(00)00056-4.
Mylonakis, G. 2001. “Simplified model for seismic pile bending at soil layer interfaces.” Soils Found. 41 (4): 47–58. https://doi.org/10.3208/sandf.41.4_47.
Mylonakis, G., E. Rovithis, and C. H. Paraschakis. 2013. “1D harmonic response of layered inhomogeneous soil: Exact and approximate analytical solutions.” In Vol. 2 of Computational methods in earthquake engineering, edited by M. Papadrakakis, M. Fragiadakis, and V. Plevris, 1–32. New York: Springer.
NIED (National Research Institute for Earth Science and Disaster Resilience). 2019. NIED K-NET, KiK-net, National research institute for earth science and disaster resilience. Ibaraki, Japan: NIED. https://doi.org/10.17598/NIED.0004.
Okada, Y., K. Kasahara, S. Hori, K. Obara, S. Sekiguchi, H. Fujiwara, and A. Yamamoto. 2004. “Recent progress of seismic observation networks in Japan–Hi-net, F-net, K-NET and KiK-net.” Earth Planet. Space 56 (8): xv–xxviii. https://doi.org/10.1186/BF03353076.
Paraschakis, C. H., E. Rovithis, and G. Mylonakis. 2010. “1D seismic response of layered inhomogeneous soil: A closed form solution.” In Proc., 9th Int. Congress on Mechanics. Hellenic Society for Theoretical & Applied Mechanics (HSTAM), 639–647. Athens, Greece: Hellenic Society for Theoretical and Applied Mechanics.
Poulos, H. G., and E. H. Davis. 1974. Elastic solution for soil and rock mechanics. New York: John Wiley & Sons.
Roesset, J. M. 1977. “Soil amplification of earthquakes.” In Numerical methods in geotechnical engineering, edited by C. H. Desai, 639–682. New York: McGraw-Hill.
Rovithis, E., C. H. Paraschakis, and G. Mylonakis. 2011. “1D harmonic response of layered inhomogeneous soil: Analytical investigation.” Soil Dyn. Earthquake Eng. 31 (7): 879–890. https://doi.org/10.1016/j.soildyn.2011.01.007.
Schreyer, H. 1977. “One-dimensional elastic waves in inhomogeneous media.” J. Eng. Mech. Div. 103 (5): 979–990. https://doi.org/10.1061/JMCEA3.0002287.
Semblat, J. F., and A. Pecker. 2009. Waves and vibrations in soils: Earthquakes, traffic, shocks, construction works. Pavia, Italy: IUSS Press.
Stewart, J. P., et al. 2017. “Non-ergodic site response in seismic hazard analysis.” Earthquake Spectra 33 (4): 1385. https://doi.org/10.1193/081716eqs135m.
Stewart, J. P., and K. Afshari. 2021. “Epistemic uncertainty in site response as derived from one-dimensional ground response analyses.” J. Geotech. Geoenviron. Eng. 147 (1): 04020146. https://doi.org/10.1061/(ASCE)GT.1943-5606.0002402.
Thompson, E. M., L. G. Baise, Y. Tanaka, and R. E. Kayen. 2012. “A taxonomy of site response complexity.” Soil Dyn. Earthquake Eng. 41 (Oct): 32–43. https://doi.org/10.1016/j.soildyn.2012.04.005.
Thompson, W. P. 1950. “Transmission of elastic waves through a stratified soil strata.” J. Appl. Phys. 21 (2): 89–93. https://doi.org/10.1063/1.1699629.
Toki, K., and S. Cherry. 1972. “Inference of subsurface accelerations and strain from accelerograms recorded at ground surface.” In Proc., 4th European Symp. on Earthquake Engineering. London: European Academy on Earthquake Engineering.
Towhata, I. 1996. “Seismic wave propagation in elastic soil with continuous variation of shear modulus in the vertical direction.” Soils Found. 36 (1): 61–72. https://doi.org/10.3208/sandf.36.61.
Trachanas, S. 2004. Ordinary differential equations. [In Greek.] Crete, Greece: Crete University Press.
Travasarou, T., and G. Gazetas. 2004. “On the linear response of soils with modulus varying as a power of depth-The Maliakos marine clay.” Soils Found. 44 (5): 85–93. https://doi.org/10.3208/sandf.44.5_85.
Vrettos, C. H. 1990. “Dispersive SH-surface waves in soil deposits of variable shear modulus.” Soil Dyn. Earthquake Eng. 9 (5): 255–264. https://doi.org/10.1016/S0267-7261(05)80004-1.
Vrettos, C. H. 2013. “Dynamic response of soils deposits to vertical SH waves for different rigidity depth-gradients.” Soil Dyn. Earthquake Eng. 47: 41–50. https://doi.org/10.1016/j.soildyn.2012.04.003.
Wylie, C. R., and L. C. Barrett. 1986. Advanced engineering mathematics. New York: McGraw-Hill.
Zhao, J. X. 1997. “Modal analysis of soft-soil sites including radiation damping.” Earthquake Eng. Struct. Dyn. 26 (1): 93–113. https://doi.org/10.1002/(SICI)1096-9845(199701)26:1%3C93::AID-EQE625%3E3.0.CO;2-A.

Information & Authors

Information

Published In

Go to Journal of Geotechnical and Geoenvironmental Engineering
Journal of Geotechnical and Geoenvironmental Engineering
Volume 148Issue 11November 2022

History

Received: Jun 8, 2021
Accepted: Apr 26, 2022
Published online: Sep 7, 2022
Published in print: Nov 1, 2022
Discussion open until: Feb 7, 2023

Authors

Affiliations

Emmanouil Rovithis, Ph.D. https://orcid.org/0000-0002-0331-5855
Researcher, Institute for Engineering Seismology and Earthquake Engineering, EPPO-ITSAK, Thessaloniki, Greece. ORCID: https://orcid.org/0000-0002-0331-5855
Professor, Dept. of Civil Infrastructure and Environmental Engineering, Khalifa Univ., Abu Dhabi, UAE; Chair in Geotechnics and Soil-Structure Interaction, Dept. of Civil Engineering, Univ. of Bristol, Bristol, UK; Adjunct Professor, Dept. of Civil and Environmental Engineering, Univ. of California Los Angeles, Los Angeles, CA (corresponding author). ORCID: https://orcid.org/0000-0002-8455-8946. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share