Next Article in Journal
Analysis of Dynamic Characteristics of Small-Scale and Low-Stiffness Ring Squeeze Film Damper-Rotor System
Next Article in Special Issue
Centrifuge Model Investigation of Interaction between Successively Constructed Foundation Pits
Previous Article in Journal
Mechanical Behaviour and Failure Mode Analysis of Penetrated Mortise–Tenon Joint with Neighbouring Gaps Based on Full-Scale Experiments
Previous Article in Special Issue
Low-Cycle Reverse Loading Tests of the Continuous Basalt Fiber-Reinforced Polymer Column Filled with Concrete
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bulk and Rayleigh Waves Propagation in Three-Phase Soil with Flow-Independent Viscosity

1
School of Civil Engineering, Southeast University, Nanjing 210096, China
2
Key Laboratory of Concrete and Prestressed Concrete Structure of Ministry of Education, Southeast University, Nanjing 210096, China
*
Author to whom correspondence should be addressed.
Submission received: 10 June 2022 / Revised: 8 July 2022 / Accepted: 13 July 2022 / Published: 16 July 2022
(This article belongs to the Special Issue Recent Progress on Advanced Foundation Engineering)

Abstract

:
The flow-independent viscosity of the soil skeleton has significant influence on the elastic wave propagation in soils. This work studied the bulk and Rayleigh waves propagation in three-phase viscoelastic soil by considering the contribution of the flow-independent viscosity from the soil skeleton. Firstly, the viscoelastic dynamic equations of three-phase unsaturated soil are developed with theoretical derivation. Secondly, the explicit characteristic equations of bulk and Rayleigh waves in three-phase viscoelastic soil are yielded theoretically by implementing Helmholtz resolution for the displacement vectors. Finally, the variations of the motion behavior for bulk and Rayleigh waves with physical parameters such as relaxation time, saturation, frequency, and intrinsic permeability are discussed by utilizing calculation examples and parametric analysis. The results reveal that the influence of soil flow-independent viscosity on the wave speed and attenuation coefficient of bulk and Rayleigh waves is significantly related to physical parameters such as saturation, intrinsic permeability, and frequency.

1. Introduction

The wave motion behavior in natural earth foundation is a fundamental scientific problem for theoretical study (e.g., soil dynamics) and practical engineering (e.g., non-destructive examination of natural and artificial materials in foundation soil). Generally, bulk and surface waves are captured in the natural earth foundation. Bulk waves, including longitudinal and transverse waves, are the seismic waves generated by source vibrations and are propagated in the media. Surface waves, such as Rayleigh waves, are secondary waves derived from bulk waves on the surface or layered surface of media. On the other side, the natural earth foundation can be divided into entirely and partially saturated states. The wave propagation behavior in partially saturated soil is significantly more complex than that in completely saturated soil because the capillary pressure and coupling effect will affect the motion behaviors of elastic waves in unsaturated soil [1,2].
Undoubtedly, the emergence of Biot theory [3,4] and mixture theory [5] not only give the sound basis for the research of wave theory but accelerate the research process. For the bulk and Rayleigh waves propagation in saturated soil, researchers including Jones [6], Plona [7], Berryman [1], Berryman [8,9], Zhou and Ma [10], Straughan et al. [11], Rohan et al. [12], Tung [13], and Wang et al. [14] implemented several theoretical and experimental works. They drew that the longitudinal, transverse, and Rayleigh waves depend on not only the frequency but the soil parameters such as permeability and soil mass types. Additionally, the applied research on the dynamic response of natural foundation also aroused broad concern from researchers [15,16,17,18], which undoubtedly promotes the basic research on wave propagation in completely saturated soils. In recent years, with the improvement of mathematics solving ability and engineering precision requirements, several researchers, including Lo [19], Lo et al. [20], and Liu et al. [21,22,23], have implemented works on the bulk and Rayleigh waves propagation in three-phase partially saturated soil. These works concluded that a kind of longitudinal wave (typically named P3 wave) exists in three-phase soil, and the liquid saturation has a significant effect on the bulk and Rayleigh waves.
Additionally, it is usually recognized that the deformation of the soil, even dry soil, is related to time. That is, the soil viscosity is dependent on the soil skeleton viscosity. Nonetheless, most of the existing works on wave propagation in unsaturated soils only take into account the fluid viscosity but seldom mention the effect of solid skeleton viscosity. Appreciatively, the relationship between soil-structure response and the soil skeleton viscosity has been proposed by researchers [24,25]. Moreover, recently, the dependence of soil skeleton viscosity on the motion behaviors of elastic waves in saturated soils has been studied with an analytical solution [26]. To overcome the lack of previous research on the relationship between the soil skeleton viscosity (flow-independent viscosity) and the elastic wave propagation in unsaturated soil, this work simplifies the unsaturated saturated soils to the multiphase mixture consisting of soil particles, water, and air. The viscoelastic wave equations are established with the Biot model, mixture theory, and the generalized Kelvin–Voigt model. Then, the characteristic formulas of longitudinal, transverse, and Rayleigh waves are yielded with theoretical derivation. Finally, the relationships between the motion behaviors (speed and attenuation) of each wave and the soil parameters are graphed and discussed by utilizing calculation examples and parametric analysis.

2. Viscoelastic Dynamic Model

Generally, the deformation of viscoelastic materials is strictly associated with the loading force and relaxation time of stress and strain. Usually, the damping of three-phase unsaturated soil to the elastic wave propagation mainly comes from two aspects: owne is the viscosity from the solid skeleton, called flow-independent viscosity, and the other is the viscosity from the liquid and gas in pores, called flow-dependent viscosity. To express the flow-independent viscosity in three-phase unsaturated soil, the generalized Kelvin–Voigt model [27,28] is introduced in this study, as illustrated in Figure 1. The dashpot depicted in Figure 1 is utilized to represent the negative effect of the solid skeleton on elastic wave propagation.
In Figure 1, the spring represents the linear-elastic response of the solid skeleton in unsaturated soil under load, while the dashpot represents the damping behaviors of the solid skeleton in unsaturated soil. The dashpot avoids the spring directly reaching the load. The correlation between elastic constants and viscous constants of viscoelastic soils can be expressed as [29]
λ v = t s λ e ,   μ v = t s μ e
where the two sets of symbols λe, μe and λv, μv stand for the elastic and viscosity constants of soil, respectively. t s stands for the relaxation time, representing the normalized viscosity without loss of generality [24,30]. In this article, t s is utilized to evaluate the viscosity of the solid skeleton, that is, the magnitude of flow-independent viscosity.
In this work, the unsaturated soil is assumed to be homogeneous, isotropic, and viscoelastic material with multiple pores, composed of liquid and gas occupying pores as well as solid particles. The liquid (water) saturation and gas (air) saturation are represented by S r and S a , respectively, and the constraint S r + S a = 1 is satisfied in this work.
According to Biot porous medium theory, the motion equation of the partially saturated soil, when ignoring the body forces and dissipation, can be expressed as [31]
σ i j , j = ρ ¯ s u ¨ i s + ρ ¯ l u ¨ i l + ρ ¯ a u ¨ i a
where ρ ¯ s = ( 1 n s ) ρ s , ρ ¯ l = n s S r ρ l , and ρ ¯ a = n s S a ρ a . The nomenclature of all the Roman and Greek symbols is stated in Nomenclature. To save space, the nomenclature (physical meaning) of the symbols in the formulas used below is listed in Nomenclature, and this point will not be repeated.
According to Bishop and Blight [32], the following formula is yielded
σ i j = σ i j + p δ i j
where p = S r p l + ( 1     S r ) p a represents the pore fluid pressure formed by liquid and gas in pores.
The stress–strain relationship of the solid skeleton is described as [33]
σ i j = c i j s ε i j c i j e ε i j p
The strain tensors under the general state and the pore pressure are written, respectively, as [33]
ε i j = 1 2 ( u i . j s + u j . i s )
ε 11 p = ε 22 p = ε 33 p = 1 3 K s p .  
The effective stress tensor of unsaturated soil is yielded through combining Equations (4)–(6), as
σ i j = ( λ e ε v + λ v ε v t ) δ i j + 2 ( μ e ε i j + μ v ε i j t ) + K b K s p δ i j
where ε v = u i , i s , and K b = λ e + 2 μ e / 3 .
The total stress of unsaturated soil is yielded by substituting Equation (7) into Equation (3), as
σ i j = ( λ e ε v + λ v ε v t ) δ i j + 2 ( μ e ε i j + μ v ε i j t ) α e p δ i j
where α e = 1     K b / K s .
The stress component acting on the soil skeleton, denoted by σ i j s , can be assumed to
σ i j s = 1 1 n s [ σ i j + n s S r p l δ i j + n s ( 1 S r ) p a δ i j ]
Substituting Equation (8) into Equation (9) and making algebraic operations, Equation (9) can be re-expressed as
σ i j s = 1 1 n s [ ( λ e ε v + λ v ε v t ) δ i j + 2 ( μ e + μ v t ) ε i j + ( n s α e ) [ S r p l + ( 1 S r ) p a ] δ i j ]
The mass balance equation of each phase is constructed after ignoring phase transitions between components, as
ρ ¯ b / t + ( ρ ¯ b u ˙ i b ) , i = 0 b = s , l , a
Further expansion of Equation (11) yields the following formulas as
n ˙ s + ( 1 n s ) ρ ˙ s ρ s + ( 1 n s ) u ˙ i , i s = 0
n s S ˙ r + n ˙ s S r + n s S r ρ ˙ l ρ l + n s S r u ˙ i , i l = 0
n s S ˙ r + n ˙ s ( 1 S r ) + n s ( 1 S r ) ρ ˙ a ρ a + n s ( 1 S r ) u ˙ i , i a = 0
Following Zhang et al. [33], the derivative of each phase density with time is written as
ρ s t = ρ s σ ˙ i i s 3 K s ,   ρ l t = ρ l p ˙ l K l ,   ρ a t = ρ a p ˙ a K a
According to the relationship between matric suction and saturation of unsaturated soil, established by Van Genuchten [34], the following formula is yielded
S ˙ r = χ m d ( 1 S r e s ) S e m + 1 m ( S e 1 m 1 ) d 1 d ( p ˙ l p ˙ a )
where S e = ( S r S r e s ) / ( 1 S r e s ) .
Substituting the first formula in Equation (15) into Equation (12) yields the derivative of porosity over time as
n ˙ s = ( α e n s ) u ˙ i , i s α v u ¨ i , i s + α e n s K s S r p ˙ l + α e n s K s ( 1 S r ) p ˙ a
where α v = ( 3 λ v + 2 μ v )   /   3 K s .
Substituting Equations (15)–(17) into Equations (13) and (14), the following formulas are yielded
T 11 p l + T 12 p a + T 13 u i , i s + T 14 u i , i l + T 15 u i , i a + T 16 u ˙ i , i s = 0
T 21 p l + T 22 p a + T 23 u i , i s + T 24 u i , i l + T 25 u i , i a + T 26 u ˙ i , i s = 0
in which
T 11 = n s A s + ( S r ) 2 ( α e n s ) / K s + S r n s / K l ,   T 12 = S r ( 1 S r ) ( α e n s ) / K s n s A s ,   T 13 = S r ( α e n s ) ,   T 14 = S r n s ,   T 15 = 0 ,   T 16 = α v S r ,   T 21 = T 12 ,   T 22 = n s A s + ( α e n s ) S a S a / K s + n s S a / K a ,   T 23 = S a ( α e n s ) ,   T 24 = 0 ,   T 25 = n s S a ,   T 26 = α v S a ,   A s = χ m d ( 1 S r e s ) ( S e 1 / m 1 ) ( d 1 ) / d S e ( m + 1 ) / m .
Making some algebraic operations for Equations (18) and (19), the pore-fluid pressures are yielded as
p l = t 11 u i , i s + t 12 u i , i l + t 13 u i , i a + t 14 u ˙ i , i s
p a = t 21 u i , i s + t 22 u i , i l + t 23 u i , i a + t 24 u ˙ i , i s
in which
t 11 = T 13 T 22 T 23 T 12 T 11 T 22 T 21 T 12 ,   t 12 = T 14 T 22 T 11 T 22 T 21 T 12 ,   t 13 = T 25 T 12 T 11 T 22 T 21 T 12 ,   t 14 = T 16 T 22 T 26 T 12 T 11 T 22 T 21 T 12 t 21 = T 13 T 21 T 23 T 11 T 12 T 21 T 22 T 11 ,   t 22 = T 14 T 21 T 12 T 21 T 22 T 11 ,   t 23 = T 25 T 11 T 12 T 21 T 22 T 11 ,   t 24 = T 16 T 21 T 26 T 11 T 12 T 21 T 22 T 11 .
The motion formulas of fluid phases accounting for the effect of pore tortuosity are formulated as [33,35]
p l , i = μ l k r l k n s S r ( u ˙ i l u ˙ i s ) + ρ l u ¨ i l + ( τ l 1 ) ρ l ( u ¨ i l u ¨ i s )
p a , i = μ a k r a k n s S a ( u ˙ i a u ˙ i s ) + ρ a u ¨ i a + ( τ a 1 ) ρ a ( u ¨ i a u ¨ i s )
According to the model established by Mualem [36], the relative permeabilities k r l and k r g are written as
k r l = ( S e ) 0.5 [ 1 ( 1 ( S e ) 1 / m ) m ] 2 ,   k r a = ( 1 S e ) 0.5 [ ( 1 ( S e ) 1 / m ) m ] 2
Combination of Equations (20)–(23) and introduction of the relative displacement of liquid ( u i w = n s S r ( u i l     u i s ) ) and gas ( u i g = n s S a ( u i a     u i s ) ) phases yield the following formulas
μ e u i , j j s + ( C 11 + μ e ) u j , j i s + C 12 u j , j i w + C 13 u j , j i g + μ v u ˙ i , j j s + ( C 14 + μ v ) u ˙ j , j i s = ρ u ¨ i s + ρ l u ¨ i w + ρ a u ¨ i g
C 21 u j , j i s + C 22 u j , j i w + C 23 u j , j i g + C 24 u ˙ j , j i s = ρ l u ¨ i s + ν w u ¨ i w + b w u ˙ i w
C 31 u j , j i s + C 32 u j , j i w + C 33 u j , j i g + C 34 u ˙ j , j i s = ρ a u ¨ i s + ν g u ¨ i g + b g u ˙ i g
with
D 1 = t 11 + t 12 + t 13 ,   D 2 = t 21 + t 22 + t 23 ,   ρ = ( 1 n s ) ρ s + n s S r ρ l + n s S a ρ a ,   C 11 = λ e + S r α e D 1 + ( 1 S r ) D 2 α e ,   C 12 = [ S r t 12 + ( 1 S r ) t 22 ] α e / ( n s S r ) , C 13 = α e ( S r t 13 + S a t 23 ) / ( n s S a ) ,   C 14 = λ v + α e ( S r t 14 + S a t 24 ) ,   C 21 = D 1 ,   C 22 = t 12 / ( n s S r ) ,   C 23 = t 13 / ( n s S a ) ,   C 24 = t 14 ,   C 31 = D 2 ,   C 32 = t 22 / ( n s S r ) ,   C 33 = t 23 / ( n s S a ) ,   C 34 = t 24 ,   ν w = τ l ρ l / ( n s S r ) ,   ν g = τ a ρ a / ( n s S a ) ,   b w = μ l / ( k r l k ) ,   b g = μ a / ( k r a k ) .
Equations (25)–(27) represent the viscoelastic wave equations of triphase partially saturated soil. The advantages of the proposed viscoelastic dynamic model can describe the effect of both the flow-independent and flow-dependent viscosities on the soil dynamic behavior. The flow-independent viscosity is represented by the shear and dilatant constants λv and μv in Equations (25)–(27). According to Equation (1), the soil skeleton viscosity can be represented by the relaxation time t s . Correspondingly, the fluid viscosity is characterized by the coefficients b w and b g in Equations (26) and (27). Apparently, the flow-dependent viscosity can finally be characterized by the intrinsic permeability k .

3. Wavefield Solution for Bulk and Rayleigh Waves

Considering the decomposition law of potential function, the displacement vector of each phase is written as
u p h = φ p h + × ψ p h ,   p h = s ,   w ,   g
Substituting Equation (28) into Equations (25)–(27), the following equations are yielded as
( C 11 + 2 μ e ) 2 φ s + C 12 2 φ w + C 13 2 φ g + ( C 14 + 2 μ v ) 2 φ ˙ s = ρ φ ¨ s + ρ l φ ¨ w + ρ a φ ¨ g C 21 2 φ s + C 22 2 φ w + C 23 2 φ g + C 24 2 φ ˙ s = ρ l φ ¨ s + v w φ ¨ w + b w φ ˙ w C 31 2 φ s + C 32 2 φ w + C 33 2 φ g + C 34 2 φ ˙ s = ρ a φ ¨ s + v g φ ¨ g + b g φ ˙ g }
ρ ψ ¨ s + ρ l ψ ¨ w + ρ a ψ ¨ g = μ e 2 ψ s + μ v 2 ψ ˙ s ρ l ψ ¨ s + v w ψ ¨ w + b w ψ ˙ w = 0 ρ a ψ ¨ s + v g ψ ¨ g + b g ψ ˙ g = 0 }

3.1. Bulk Waves

For the bulk wave propagation in the triphase unsaturated soil, the displacement potentials of solid, liquid, and air phases in Equations (29) and (30) can be expressed as
φ p h = A p h exp [ i ( ω t k p r ) ]
ψ p h = B p h exp [ i ( ω t k s r ) ]
where ω = 2 π f , and i = 1 .
The following formulas are obtained by combining Equations (29)–(32), as
( ρ ω 2 [ ( C 11 + 2 μ e ) + i ω ( C 14 + 2 μ v ) ] k p 2 ρ l ω 2 C 12 k p 2 ρ a ω 2 C 13 k p 2 ρ l ω 2 ( C 21 + C 24 i ω ) k p 2 v w ω 2 b w i ω C 22 k p 2 C 23 k p 2 ρ a ω 2 ( C 31 + C 34 i ω ) k p 2 C 32 k p 2 v g ω 2 b g i ω C 33 k p 2 ) ( A s A w A g ) = ( 0 0 0 )
( ρ ω 2 ( μ e + μ v i ω ) k s 2 ρ l ω 2 ρ a ω 2 ρ l ω 2 v w ω 2 b w i ω 0 ρ a ω 2 0 v g ω 2 b g i ω ) ( B s B w B g ) = ( 0 0 0 )
Finally, the characteristic equations for bulk waves in triphase unsaturated viscoelastic soil are derived from Equations (33) and (34) as
| L 11 L 12 L 13 L 21 L 22 L 23 L 31 L 32 L 33 | = 0
| J 11 J 12 J 13 J 21 J 22 J 23 J 31 J 32 J 33 | = 0
with
L 11 = ρ ω 2 [ ( C 11 + 2 μ e ) + i ω ( C 14 + 2 μ v ) ] k p 2 ,   L 12 = ρ l ω 2 C 12 k p 2 ,   L 13 = ρ a ω 2 C 13 k p 2 ,   L 21 = ρ l ω 2 ( C 21 + C 24 i ω ) k p 2 ,   L 22 = v w ω 2 b w i ω C 22 k p 2 ,   L 23 = C 23 k p 2 ,   L 31 = ρ a ω 2 ( C 31 + C 34 i ω ) k p 2 ,   L 32 = C 32 k p 2 ,   L 33 = v g ω 2 b g i ω C 33 k p 2 ,   J 11 = ρ ω 2 ( μ e + μ v i ω ) k s 2 ,   J 12 = ρ l ω 2 ,   J 13 = ρ a ω 2 ,   J 21 = ρ l ω 2 ,   J 22 = v w ω 2 b w i ω ,   J 23 = J 32 = 0 ,   J 31 = ρ a ω 2 , J 33 = v g ω 2 b g i ω .
For the complex wavenumber of the longitudinal wave, Equation (35) can be solved into three meaningful solutions. This means that three kinds of longitudinal waves exist in unsaturated soil. The three longitudinal waves are usually signed as P1, P2, and P3 waves, and the speed of the P1 wave is the fastest, the P2 wave is in the middle, whereas the P3 wave is the slowest. Similarly, Equation (36) can be solved to one meaningful solution for the complex wavenumber of the transverse wave. That is, only one kind of transverse wave exists in unsaturated soil, usually labeled as S wave. Additionally, the following formulas can be yielded from Equations (33)–(36) as
p w s = A w A s = L 11 L 23 L 21 L 13 L 22 L 13 L 12 L 23 ,   p g s = A g A s = L 11 L 22 L 21 L 12 L 23 L 12 L 13 L 22
s w s = B w B s = J 21 J 22 ,   s g s = B g B s = J 31 J 33
Generally, the wave speed and attenuation coefficient of bulk waves are defined as
v p 1 = ω Re ( k p 1 ) ,   v p 2 = ω Re ( k p 2 ) ,   v p 3 = ω Re ( k p 3 ) ,   v s = ω Re ( k s )
δ p 1 = Im ( k p 1 ) ,   δ p 2 = Im ( k p 2 ) ,   δ p 3 = Im ( k p 3 ) ,   δ s = Im ( k s )

3.2. Rayleigh Wave

As portrayed in Figure 2, the Rayleigh wave is generated by superimposing the longitudinal and transverse waves at the soil boundary z = 0 . Accordingly, the displacement potential in Equations (29) and (30) for the Rayleigh wave can be further rewritten as
φ s = [ A ( 1 ) s e i γ 1 z + A ( 2 ) s e i γ 2 z + A ( 3 ) s e i γ 3 z ] e i ( ω t k R x ) φ w = [ p ( 1 ) w s A ( 1 ) s e i γ 1 z + p ( 2 ) w s A ( 2 ) s e i γ 2 z + p ( 3 ) w s A ( 3 ) s e i γ 3 z ] e i ( ω t k R x ) φ g = [ p ( 1 ) g s A ( 1 ) s e i γ 1 z + p ( 2 ) g s A ( 2 ) s e i γ 2 z + p ( 3 ) g s A ( 3 ) s e i γ 3 z ] e i ( ω t k R x ) }
ψ s = B s e i γ 4 z e i ( ω t k R x ) ψ w = s w s B s e i γ 4 z e i ( ω t k R x ) ψ g = s g s B s e i γ 4 z e i ( ω t k R x ) }
where
γ 1 = k p 1 2 k R 2 ,   γ 2 = k p 2 2 k R 2 ,   γ 3 = k p 3 2 k R 2 ,   γ 4 = k s 2 k R 2
As discussed earlier, the Rayleigh wave in unsaturated soil is a kind of superimposed wave at the soil boundary. In this work, considering the stress-free boundary, the stresses ( σ z z and σ x z ) and fluid pressures ( p l and p a ) disappear at the soil boundary. Accordingly, the following formulas can be obtained as
σ z z = ( C 11 + C 14 t ) ( u x s x + u z s z ) + 2 ( μ e + μ v t ) u z s z + C 12 ( u x w x + u z w z ) + C 13 ( u x g x + u z g z ) = 0
σ x z = ( μ e + μ v t ) ( u x s z + u z s x ) = 0
p l = ( C 21 + C 24 t ) ( u x s x + u z s z ) C 22 ( u x w x + u z w z ) C 23 ( u x g x + u z g z ) = 0
p a = ( C 31 + C 34 t ) ( u x s x + u z s z ) C 32 ( u x w x + u z w z ) C 33 ( u x g x + u z g z ) = 0
Substituting Equations (41) and (42) into Equations (44)–(47) and making algebraic operations, the following formula can be derived as
( 11 12 13 14 21 22 23 24 31 32 33 34 41 42 43 44 ) ( A ( 1 ) s A ( 2 ) s A ( 3 ) s B s ) = ( 0 0 0 0 )
with
11 = 2 ( μ e + μ v i ω ) γ 1 2 + [ C 11 + C 14 i ω + C 12 p ( 1 ) w s + C 13 p ( 1 ) g s ] k p 1 2 ,   12 = 2 ( μ e + μ v i ω ) γ 2 2 + [ C 11 + C 14 i ω + C 12 p ( 2 ) w s + C 13 p ( 2 ) g s ] k p 2 2 ,   13 = 2 ( μ e + μ v i ω ) γ 3 2 + [ C 11 + C 14 i ω + C 12 p ( 3 ) w s + C 13 p ( 3 ) g s ] k p 3 2 ,   14 = 2 ( μ e + μ v i ω ) k R γ 4 ,   21 = 2 ( μ e + μ v i ω ) γ 1 k R ,   22 = 2 ( μ e + μ v i ω ) γ 2 k R ,   23 = 2 ( μ e + μ v i ω ) γ 3 k R ,   24 = ( μ e + μ v i ω ) ( k R 2 γ 4 2 ) ,   31 = [ C 21 + C 24 i ω + C 22 p ( 1 ) w s + C 23 p ( 1 ) g s ] k p 1 2 ,   32 = [ C 21 + C 24 i ω + C 22 p ( 2 ) w s + C 23 p ( 2 ) g s ] k p 2 2 ,   33 = [ C 21 + C 24 i ω + C 22 p ( 3 ) w s + C 23 p ( 3 ) g s ] k p 3 2 ,   41 = [ C 31 + C 34 i ω + C 32 p ( 1 ) w s + C 33 p ( 1 ) g s ] k p 1 2 ,   42 = [ C 31 + C 34 i ω + C 32 p ( 2 ) w s + C 33 p ( 2 ) g s ] k p 2 2 ,   43 = [ C 31 + C 34 i ω + C 32 p ( 3 ) w s + C 33 p ( 3 ) g s ] k p 3 2 ,   34 = 44 = 0
Finally, the characteristic equation of the Rayleigh wave is derived from Equation (48) as
| 11 12 13 14 21 22 23 24 31 32 33 34 41 42 43 44 | = 0
By solving Equation (49), the wave speed ( v R ) and attenuation coefficient ( δ R ) of Rayleigh are solved as
v R = ω / Re ( k R ) ,   δ R = Im ( k R )

4. Numerical Examples and Parametric Analysis

This section will utilize numerical calculation and parametric analysis to analyze the dependence of physical parameters for the soil on the motion behaviors (mainly for the speed and attenuation) of bulk and Rayleigh waves. The value of the soil parameter in the following calculation examples refers to Table 1 [2] unless otherwise specified.
When t s is set to zero, the unsaturated porous viscoelastic model developed in this work can be reduced to the traditional, unsaturated poroelastic model. Thus, the results obtained in this work can be validated through comparing them with the results developed in previous work. The comparison results for the wave velocities of P1, S, and Rayleigh waves between this work and the work developed by Yang [2] are presented in Figure 3a, which shows that the analytical solutions obtained by this work and Yang [2] have a good agreement. Additionally, Murphy [37] carried out acoustic measurements of partial gas saturation in tight sandstones employing a pulse transmission technique, obtaining the wave speeds of P1 and S waves. Accordingly, this work introduces the physical parameters of tight sandstones [37] and calculates the corresponding wave speeds of P1 and S waves. The calculated results from this work are compared with the experimental measurement results observed by Murphy [37], as portrayed in Figure 3b. The comparison results in Figure 3 have good consistency both quantitatively and qualitatively.
Figure 4, Figure 5 and Figure 6 depict the variations for v p 1 ,   v s , v R , δ p 1 , δ s , and δ R under various values of S r and t s , where S r ranges from 10% to 100% and t s is taken to be 0 s, 0.5 × 10−3 s, and 1.0 × 10−3 s, respectively. As depicted in Figure 4, Figure 5 and Figure 6, v p 1 is the largest, followed by v s , and then v R . In contrast, the order of attenuation coefficients runs entirely counter to the order of wave speed. As shown in Figure 4a, Figure 5a and Figure 6a, v p 1 ,   v s , and v R decrease obviously with the increase of S r when the foundation soil is partially saturated. Although, there is a sudden sharp increase in v p 1 and v R but not in v s when the foundation soil reaches complete saturation state. At the same time, as clarified in Figure 4b, Figure 5b and Figure 6b, the variations of δ p 1 , δ s , and δ R with S r run diametrically counter to that of v p 1 ,   v s , and v R with S r . On the other side, the effects of t s on v p 1 ,   v s , v R , δ p 1 , δ s , and δ R are conspicuous from the variations in Figure 4, Figure 5 and Figure 6. For P1, S, and Rayleigh waves, the curves reveal that v p 1 ,   v s , v R , δ p 1 , δ s , and δ R enlarge with the increase of t s value. In practical engineering, P1, S, and Rayleigh waves are the main application object, whereas P2 and P3 waves (not shown graphically in this work) are challenging to observe in engineering practice due to their slow speed and fast attenuation. Therefore, for the works of wave propagation in unsaturated or saturated foundation soil, it is essential to account for the effect of flow-independent viscosity from the solid skeleton. Additionally, the soil–water characteristic curve (SWCC) is depicted in Figure 7 (ϕ represents the matric suction), illustrating that the increase of S r will reduce the matric suction ϕ of soil. When soil is completely saturated (Sr = 100%), ϕ will decrease to 0. Undoubtedly, the effect of ϕ and S r on the motion behaviors (speed and attenuation) of bulk and Rayleigh waves is opposite.
The dependency of v p 1 ,   v s , v R , δ p 1 , δ s , and δ R on t s and f are depicted in Figure 8, Figure 9 and Figure 10, therein f ranges from 0.01 Hz to 150 Hz, t s is taken to be 0 s, 0.5 × 10−3 s, and 1.0 × 10−3 s, respectively, and the value of other physical parameters refers to Table 1. It can be captured from Figure 8, Figure 9 and Figure 10 that each wave presents positive relativity between both the wave speeds and attenuation coefficients and the frequency. In contrast, this positive relativity will be hidden when the soil skeleton viscosity is not considered. This is a vital issue worthy of attention in practical engineering applications. Meanwhile, Figure 8, Figure 9 and Figure 10 show that the increase of v p 1 ,   v s , v R , δ p 1 , δ s , and δ R will accompany the enlargement of t s .
As mentioned in Equations (25)–(27), the flow-dependent viscosities from pore-liquid and pore-air are characterized by the parameters b w and b g in Equations (26) and (27) and finally represented by the intrinsic permeability k . For comparison, Figure 11, Figure 12 and Figure 13 demonstrate the effects of k and t s on v p 1 ,   v s , v R , δ p 1 , δ s , and δ R . In this example, the value range of k is 10−12~10−7 m2, while t s ranges from 0 × 10−3 s to 2.0 × 10−3 s. It can be seen from Figure 11a, Figure 12a and Figure 13a that v p 1 ,   v s , and v R remain unchanged when k is in the range of 10−12~10−9 m2 but increase gradually with the enlargement of k in the high permeability zone (i.e., k = 10−9 m2~10−7 m2). For δ p 1 , δ s , and δ R , the variation trends in Figure 11b, Figure 12b and Figure 13b display an approximately normal distribution in the entire permeability range. Furthermore, as above-mentioned discussion, the increase of t s will make the dependences of v p 1 ,   v s , v R , δ p 1 , δ s , and δ R on k move upward along the positive direction of the vertical axis. Apparently, through comparison, it can be seen that compared to the fluid viscosity, the soil skeleton viscosity has a more prominent contribution to the motion behaviors (speed and attenuation) of elastic waves and is more significant for practical engineering applications.
The increase in the relaxation time will influence the interaction between the microscopic particles of the material, and thus will change the wave speed and attenuation coefficient of the bulk and Rayleigh waves. The effect of saturation on the propagation behavior of bulk and Rayleigh waves can be attributed to the increase of saturation. The volume of liquid in soil pores increases, and thus the flow-dependent viscosity increases. It leads to the reduction of the wave velocity of elastic waves and the acceleration of attenuation. However, when the soil tends to be fully saturated, the mutual coupling of the liquid and gas phase disappears, resulting in the disappearance of the capillary effect and matric suction of the soil, so that the wave speeds of P1 and Rayleigh waves suddenly increase and the attenuation coefficients suddenly decrease. The soil intrinsic permeability represents a quantitative property of porous material, depending solely on the pore structure of porous material. The larger the intrinsic permeability, the smaller the effect of flow viscosity on the propagation of bulk and Rayleigh waves, which will have a greater effect on its wave speed and attenuation coefficient.

5. Conclusions

In this article, considering the influence of both the flow-independent and flow-dependent viscosities for soils on the elastic wave propagation, a modified viscoelastic dynamic model of the three-phase partially saturated soil is established to investigate the motion behaviors of bulk and Rayleigh waves. The characteristics equations of bulk and Rayleigh waves have been yielded analytical in explicit form. The dependences of the motion behaviors (speed and attenuation) for various waves on the physical parameters are graphed through employing several numerical examples. The main conclusions are summarized as follows: (i) Both the wave speed and attenuation coefficient of P1, S, and Rayleigh waves increase obviously with the flow-independent viscosity at both the partially and completely saturated states; (ii) In the high permeability zone, both the flow-independent and flow-dependent viscosities have an apparent influence on the propagation for P1, S, and Rayleigh waves; (iii) In the low and middle permeability zone, only the flow-independent viscosity affects the propagation behavior for P1, S, and Rayleigh waves.

Author Contributions

All authors conceived and designed the study. Conceptualization, methodology, and writing—review and editing: Q.G., H.L. and G.D.; writing—original draft preparation and visualization: Q.G. and H.L.; investigation and resources: Z.L. and Q.G.; supervision: G.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (NSFC Grant Nos. 52078128, 51878160).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

The nomenclature of symbols in this paper
Roman Symbols
A p h scalar potential amplitudes of p h phase
B p h vector potential amplitude of p h phase
c ij e isotropic elastic coefficient matrix of soil skeleton
c ij s isotropic viscoelastic coefficient matrix of soil skeleton
d fitting parameters of unsaturated soil
f conventional frequency
i imaginary unit
k intrinsic permeability of unsaturated soil
k R complex wavenumber of Rayleigh wave
k p complex wavenumber of longitudinal wave
k s complex wavenumber of transverse wave
k r a relative permeability of gas phase
k r l relative permeability of liquid phase
K b bulk modulus of soil skeleton
K a bulk modulus of gas phase
K l bulk modulus of liquid phase
K s compression modulus of soil particles
m fitting parameters of unsaturated soil
n s porosity of unsaturated soil
p averaged pore pressure
p a gas pressure
p l liquid pressure
r position vector
S a gas saturation
S r liquid saturation
S e effective liquid saturation
S res liquid saturation at residual state
t s relaxation time
u i a displacement component of gas phase
u i l displacement component of liquid phase
u i s displacement component of solid phase
u i g relative displacement of gas phase
u i w relative displacement of liquid phase
v p 1 wave speed of P1 wave
v p 2 wave speed of P2 wave
v p 3 wave speed of P3 wave
v s wave speed of S wave.
Greek Symbols
γ 1 wavenumber component of P1 wave in z-direction
γ 2 wavenumber component of P2 wave in z-direction
γ 3 wavenumber component of P3 wave in z-direction
γ 4 wavenumber component of S wave in z-direction
δ ij Kronecker delta
δ p 1 attenuation coefficient of P1 wave
δ p 2 attenuation coefficient of P2 wave
δ p 3 attenuation coefficient of P3 wave
δ s attenuation coefficient of S wave
ε v volumetric strain of soil skeleton
ε ij strain tensor under general state
ε ij p strain tensor under pore pressure
λ e ,   μ e elastic constant of soil
λ v ,   μ v viscosity constants of soil
μ a dynamic viscosity of gas phases
μ l dynamic viscosity of liquid phase
ρ a mass density of gas phase
ρ l mass density of liquid phase
ρ s mass density of solid phase
ρ ¯ a apparent density of gas phase
ρ ¯ l apparent density of liquid phase
ρ ¯ s apparent density of solid phase
σ ij total stress
σ ij effective stress tensor of unsaturated soil
τ a tortuosity of gas phase
τ l tortuosity of liquid phase
φ p h scalar potential of p h phase
ψ p h vector potentials of p h phase
χ fitting parameters of unsaturated soil
ω angular frequency.

References

  1. Berryman, J.G. Confirmation of Biot’s theory. Appl. Phys. Lett. 1980, 37, 382–384. [Google Scholar] [CrossRef]
  2. Yang, J. Rayleigh surface waves in an idealised partially saturated soil. Géotechnique 2005, 55, 409–414. [Google Scholar] [CrossRef]
  3. Biot, M.A. Theory of propagation of elastic waves in a fluid-saturated porous solid. I. Low-frequency range. J. Acoust. Soc. Am. 1956, 28, 168–178. [Google Scholar] [CrossRef]
  4. Biot, M.A. Theory of propagation of elastic waves in a fluid-saturated porous solid. II. Higher frequency range. J. Acoust. Soc. Am. 1956, 28, 179–191. [Google Scholar] [CrossRef]
  5. Bowen, R.M. Incompressible porous media models by use of the theory of mixtures. Int. J. of Eng. Sci. 1980, 18, 1129–1148. [Google Scholar] [CrossRef]
  6. Jones, J.P. Rayleigh waves in a porous, elastic, saturated solid. J. Acoust. Soc. Am. 1961, 33, 959–962. [Google Scholar] [CrossRef]
  7. Plona, T.J. Observation of a second bulk compressional wave in a porous medium at ultrasonic frequencies. Appl. Phys. Lett. 1980, 36, 259–261. [Google Scholar] [CrossRef] [Green Version]
  8. Berryman, J.G.; Thigpen, L.; Chin, R.C. Bulk elastic wave propagation in partially saturated porous solids. J. Acoust. Soc. Am. 1988, 84, 360–373. [Google Scholar] [CrossRef]
  9. Berryman, J.G. Fluid effects on shear waves in finely layered porous media. Geophysics 2005, 70, N1–N15. [Google Scholar] [CrossRef]
  10. Zhou, F.; Ma, Q. Propagation of Rayleigh waves in fluid-saturated non-homogeneous soils with the graded solid skeleton distribution. Int. J. Numer. Anal. Methods Geomech. 2016, 40, 1513–1530. [Google Scholar] [CrossRef]
  11. Straughan, B.; Tibullo, V.; Amendola, A. Nonlinear acceleration wave propagation in the DKM theory. Mech. Res. Commun. 2020, 104, 103482. [Google Scholar] [CrossRef]
  12. Rohan, E.; Nguyen, V.H.; Naili, S. Homogenization approach and Floquet-Bloch theory for wave analysis in fluid-saturated porous media with mesoscopic heterogeneities. Appl. Math. Model. 2021, 91, 1–23. [Google Scholar] [CrossRef]
  13. Tung, D.X. Surface waves in nonlocal transversely isotropic liquid-saturated porous solid. Arch. Appl. Mech. 2021, 91, 2881–2892. [Google Scholar] [CrossRef]
  14. Wang, B.; Zhang, X.; Sun, B. Propagation prediction of body waves in fluid-saturated soils with flow-independent viscosity. Symmetry 2022, 14, 408. [Google Scholar] [CrossRef]
  15. Fattah, M.Y.; Abbas, S.F.; Karim, H.H. A model for coupled dynamic elastic-plastic analysis of soils. J. GeoEngineering 2012, 7, 43–50. [Google Scholar]
  16. Fattah, M.Y.; Al Mosawi, M.J.; Al Ameri, A.F. Dynamic response of saturated soil-foundation system acted upon by vibration. J. Earthq. Eng. 2017, 21, 1158–1188. [Google Scholar] [CrossRef]
  17. Alzabeebee, S. Dynamic response and design of a skirted strip foundation subjected to vertical vibration. Geomech. Eng. 2020, 20, 345–358. [Google Scholar]
  18. Alzabeebee, S. Numerical analysis of the interference of two active machine foundations. Geotech. Geol. Eng. 2020, 38, 5043–5059. [Google Scholar] [CrossRef]
  19. Lo, W.C. Propagation and attenuation of Rayleigh waves in a semi-infinite unsaturated poroelastic medium. Adv. Water Resour. 2008, 31, 1399–1410. [Google Scholar] [CrossRef]
  20. Lo, W.C.; Yeh, C.L.; Lee, J.W. Effect of viscous cross coupling between two immiscible fluids on elastic wave propagation and attenuation in unsaturated porous media. Adv. Water Resour. 2015, 83, 207–222. [Google Scholar] [CrossRef]
  21. Liu, H.; Zhou, F.; Wang, L.; Zhang, R. Propagation of Rayleigh waves in unsaturated porothermoelastic media. Int. J. Numer. Anal. Met. 2020, 44, 1656–1675. [Google Scholar] [CrossRef]
  22. Liu, H.; Dai, G.; Zhou, F.; Mu, Z. Propagation behavior of homogeneous plane-P1-wave at the interface between a thermoelastic solid medium and an unsaturated porothermoelastic medium. Eur. Phys. J. Plus 2021, 136, 1163. [Google Scholar] [CrossRef]
  23. Liu, H.; Dai, G.; Zhou, F.; Cao, X. A mixture theory analysis for reflection phenomenon of homogeneous plane-P1-wave at the boundary of unsaturated porothermoelastic media. Geophys. J. Int. 2022, 228, 1237–1259. [Google Scholar] [CrossRef]
  24. Bardet, J.P. A Viscoelastic Model for the Dynamic Behavior of Saturated Poroelastic Soils. J. Appl. Mech. 1992, 59, 128–135. [Google Scholar] [CrossRef]
  25. Xie, K.; Liu, G.; Shi, Z. Dynamic response of partially sealed circular tunnel in viscoelastic saturated soil. Soil Dyn. Earthq. Eng. 2004, 24, 1003–1011. [Google Scholar] [CrossRef]
  26. Chen, W.; Wang, D.; Mou, Y.; Zhao, K.; Chen, G. Effect of flow-independent viscosity on the propagation of Rayleigh wave in porous media. Soil Dyn. Earthq. Eng. 2021, 142, 106564. [Google Scholar] [CrossRef]
  27. Cheng, Z.; Leong, E.C. Finite element simulations of wave propagation in soils using a Viscoelastic model. Soil Dyn. Earthq. Eng. 2016, 88, 207–214. [Google Scholar] [CrossRef]
  28. Sills, G.C.; Wheeler, S.J.; Thomas, S.D.; Gardner, T.N. Behaviour of offshore soils containing gas bubbles. Géotechnique 1991, 41, 227–241. [Google Scholar] [CrossRef]
  29. Michaels, P. In situ determination of soil stiffness and damping. J. Geotech. Geoenviron. Eng. 1998, 124, 709–719. [Google Scholar] [CrossRef]
  30. Militano, G.; Rajapakse, R. Dynamic response of a pile in a multi-layered soil to transient torsional and axial loading. Géotechnique 1999, 49, 91–109. [Google Scholar] [CrossRef]
  31. Zhou, F.; Liu, H.; Li, S. Propagation of thermoelastic waves in unsaturated porothermoelastic media. J. Therm. Stress. 2019, 42, 1256–1271. [Google Scholar] [CrossRef]
  32. Bishop, A.W.; Blight, G. Some aspects of effective stress in saturated and partly saturated soils. Géotechnique 1963, 13, 177–197. [Google Scholar] [CrossRef]
  33. Zhang, M.; Wang, X.; Yang, G.; Xie, L. Solution of dynamic Green’s function for unsaturated soil under internal excitation. Soil Dyn. Earthq. Eng. 2014, 64, 63–84. [Google Scholar] [CrossRef]
  34. Van Genuchten, M.T. A closed-form equation for predicting the hydraulic conductivity of unsaturated soils. Soil Sci. Soc. Am. J. 1980, 44, 892–898. [Google Scholar] [CrossRef] [Green Version]
  35. Capeillère, J.; Mesgouez, A.; Lefeuve-Mesgouez, G. Axisymmetric wave propagation in multilayered poroelastic grounds due to a transient acoustic point source. Soil Dyn. Earthq. Eng. 2013, 52, 70–76. [Google Scholar] [CrossRef] [Green Version]
  36. Mualem, Y. A new model for predicting the hydraulic conductivity of unsaturated porous media. Water Resour. Res. 1976, 12, 513–522. [Google Scholar] [CrossRef] [Green Version]
  37. Murphy, W.F. Acoustic measures of partial gas saturation in tight sandstones. J. Geophys. Res. 1984, 89, 11549–11559. [Google Scholar] [CrossRef]
Figure 1. Generalized Kelvin–Voigt model [27,28].
Figure 1. Generalized Kelvin–Voigt model [27,28].
Applsci 12 07166 g001
Figure 2. The semi-infinite system of Rayleigh wave propagation.
Figure 2. The semi-infinite system of Rayleigh wave propagation.
Applsci 12 07166 g002
Figure 3. Comparisons of the wave speed between this study and the works done by Yang [2] and Murphy [37]. (a) Comparison with Yang (2005). (b) Comparison with Murphy (1984).
Figure 3. Comparisons of the wave speed between this study and the works done by Yang [2] and Murphy [37]. (a) Comparison with Yang (2005). (b) Comparison with Murphy (1984).
Applsci 12 07166 g003
Figure 4. P1 wave propagation for different saturations and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Figure 4. P1 wave propagation for different saturations and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Applsci 12 07166 g004
Figure 5. S wave propagation for different saturations and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Figure 5. S wave propagation for different saturations and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Applsci 12 07166 g005
Figure 6. Rayleigh wave propagation for different saturations and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Figure 6. Rayleigh wave propagation for different saturations and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Applsci 12 07166 g006
Figure 7. Soil–water characteristic curve (SWCC).
Figure 7. Soil–water characteristic curve (SWCC).
Applsci 12 07166 g007
Figure 8. P1 wave propagation for different frequencies and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Figure 8. P1 wave propagation for different frequencies and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Applsci 12 07166 g008
Figure 9. S wave propagation for different frequencies and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Figure 9. S wave propagation for different frequencies and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Applsci 12 07166 g009
Figure 10. Rayleigh wave propagation for different frequencies and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Figure 10. Rayleigh wave propagation for different frequencies and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Applsci 12 07166 g010
Figure 11. P1 wave propagation for different intrinsic permeabilities and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Figure 11. P1 wave propagation for different intrinsic permeabilities and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Applsci 12 07166 g011
Figure 12. S wave propagation for different intrinsic permeabilities and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Figure 12. S wave propagation for different intrinsic permeabilities and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Applsci 12 07166 g012
Figure 13. Rayleigh wave propagation for different intrinsic permeabilities and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Figure 13. Rayleigh wave propagation for different intrinsic permeabilities and relaxation times. (a) Wave speed. (b) Attenuation coefficient.
Applsci 12 07166 g013
Table 1. Physical parameters of unsaturated soil.
Table 1. Physical parameters of unsaturated soil.
ParameterValueParameterValueParameterValue
ns0.4Ks36 GPaλe120 MPa
Sr0.6Kl2.2 GPaμe120 MPa
Sres0.05Ka0.1 MPaμl0.001 Pa·s
ρs2650 kg·m−3χ0.0001μa1.8 × 10−5 Pa·s
ρl1000 kg·m−3m0.5τl1.0
ρa1.3 kg·m−3d2.0τa1.0
k1.0 × 10−11 m2f100 Hz
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, Q.; Liu, H.; Dai, G.; Li, Z. Bulk and Rayleigh Waves Propagation in Three-Phase Soil with Flow-Independent Viscosity. Appl. Sci. 2022, 12, 7166. https://0-doi-org.brum.beds.ac.uk/10.3390/app12147166

AMA Style

Guo Q, Liu H, Dai G, Li Z. Bulk and Rayleigh Waves Propagation in Three-Phase Soil with Flow-Independent Viscosity. Applied Sciences. 2022; 12(14):7166. https://0-doi-org.brum.beds.ac.uk/10.3390/app12147166

Chicago/Turabian Style

Guo, Qing, Hongbo Liu, Guoliang Dai, and Zhongwei Li. 2022. "Bulk and Rayleigh Waves Propagation in Three-Phase Soil with Flow-Independent Viscosity" Applied Sciences 12, no. 14: 7166. https://0-doi-org.brum.beds.ac.uk/10.3390/app12147166

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop