Home Dynamic response analysis of fluid-saturated porous rectangular plates
Article Publicly Available

Dynamic response analysis of fluid-saturated porous rectangular plates

  • Wang Lixian EMAIL logo
Published/Copyright: October 21, 2020

Abstract

Based on Biot’s model for saturated porous media, the governing equation of fluid-saturated porous rectangular plates is presented, in which the compressibility of solid particles and fluid and the viscosity of pore fluid is taken into account. A series solution is given for simply-supported fluid-saturated porous plates. The accuracy of the solution is validated by degenerating the fluid-saturated porous rectangular plates into single-phase solid rectangular plates. As a numerical example, the free vibration characteristic and the dynamic response under harmonic loads are analysed. The influence of surface infiltration conditions, porosity, pore fluid permeability coefficient and loading frequency on the free vibration frequency is discussed.

1 Introduction

Since Biot (1956) [1], [2] put forward the basic equations describing the dynamic characteristics of saturated porous media, the theory of porous media has become the foundation for studying the dynamic characteristics and performing dynamic analysis of saturated porous media, and has been widely used in different engineering of a variety of fields.

To date, research on saturated porous media in the fields of geotechnical engineering, earthquake engineering and geophysical has mainly focused on geotechnical materials whose geometric characteristics are associated with the half-space domain or an infinite horizontal layer, including theoretical analysis and numerical simulation of dynamic responses [3], [4], [5], [6], [7], [8], [9], [10] and wave propagation characteristics [11], [12], [13], [14]. However, few studies have been focused on flexible structures, such as porous media beams, plates and shell structures. On the one hand, the behaviour of flexible porous structures is an important issue in biomechanics [15], [16], such as the mechanical analysis of cartilage tissue and the stems of plants. On the other hand, porous materials such as polyurethane foams and fibrous materials have been widely used in the automotive and aerospace industries for damping and sound absorption [17], [18]. Therefore, it is necessary to further study the static and dynamic mechanical behaviours of flexible porous structures.

In terms of fluid-saturated porous plate structures, considering the lateral diffusion of fluid, Taber [19] established the solid-phase and fluid-phase governing equation of the isotropic fluid-saturated porous media based on Biot’s model to analyse the dynamic bending of simply supported rectangular porous plates using the Laplace transformation and the perturbation method. Leclaire et al. [20] analysed the transverse vibration problem of rectangular thin plates for four-edge-clamped porous media using the Rayleigh–Ritz method, which takes the effect of fluid viscosity on energy dissipation into account. Based on the research of Theodorakopoulos and Beskos [21] on the bending vibration of porous elastic plates and Biot’s theory, Anke and Heinz [22] established the dynamic mathematical model of saturated porous elastic Mindlin plates and gave the principle of virtual work using the deflection, angle and pore stress as basic unknown quantities. Based on the classical theory of homogeneous plates and Biot’s stress–strain relations in an isotropic porous medium with a uniform porosity, Feng-xi and Xiao-lin [23] researched the dynamic bending mathematical model of saturated porous elastic plates, and the influence of porosity, tortuosity and permeability on the resonances was studied to determine the condition of maximum damping considering these parameters. Nagler and Schanz [24], using the series approximation method, obtained another plate shear deformation theory of porous material plates. Many studies on the mechanical response of fluid-saturated porous plates used various simplified plate theories, such as classical thin plate theory and Reissner–Mindlin thick plate theory. Rezaei and Saidi [25] presented an exact solution for the free vibration analysis of porous rectangular plates under undrained conditions, and the results show that the effect of coupled fluid–solid deformation may not be disregarded. However, other studies have shown that the mechanical quantity cannot be a polynomial of coordinate variables in the thickness direction. In various simplified theories, incompatibility among the fundamental equations can be found, that is, some mechanical quantities can meet only some and not all of the basic equations, and the error will increase greatly with increasing thickness [26].

Based on Biot’s model of porous media and three-dimensional elastic theory, the dynamic governing equation of fluid-saturated porous rectangular plates is established in this paper. The free vibration and forced vibration responses of simply supported fluid-saturated porous rectangular plates are researched. The influence of the surface infiltration conditions, porosity, pore fluid permeability coefficient and loading frequency parameters on the free vibration frequency, the solid skeleton stress, the pore fluid pressure, the solid skeleton displacement and pore fluid displacement of plates are analysed via numerical examples.

2 The governing equation of fluid-saturated porous rectangular plates

A rectangular plate occupying the region [0, L1] × [0, L2] × [−H/2, H/2] in the unstressed reference configuration is described in rectangular Cartesian coordinates xi (i = 1, 2, 3). Based on Biot’s model of porous media, the basic equations of homogeneous saturated porous media are as follows.

The constitutive equations for a fluid-saturated porous media are

(1)σij=λεkkδij+2μεijαpδij
(2) p=MζαMεkk

The strain tensor for infinitesimal deformations is related to the displacements ui by

(3)εij=12(ui,j+uj,i)

The equilibrium equations in the absence of a body force are

(4)σij,j=ρu¨i+ρfw¨i
(5)p,i=ρfu¨i+mw¨i+bw˙i

where σij and p are the total stress components and pore fluid pressure (i, j = 1, 2, 3), respectively. ui and wi are the displacement components of the solid skeleton and pore fluid, that is e = ui,i, and ζ = −wi,i. λ and μ are the Lame constants and εij is the strain vector of the solid skeleton. b is a parameter accounting for internal friction due to the relative motion between the solid and the pore fluid and b = η/kf, where kf is the dynamic permeability and η is the viscosity of the fluid. α and M are the Biot parameters considering compressibility of the two-phase material, and α = 1 − K/Ks, 1/M = (αn)/Ks + n/Kf, where K, Ks and Kf are the bulk modulus of the solid skeleton, solid particles and pore fluid, respectively, while n is the soil porosity. ρ is the total density of the saturated soil and can be represented by ρ = (1 − n)ρs + nρf, where ρs and ρf are the densities of the solid phase and liquid phase, respectively. m is a parametric representation related to the mass density of the pore fluid and pore geometry features m = ρf/n.

Combining Eqs. (1)–(5), the governing equations of the dynamic response of the saturated media can be obtained as

(6)μ2u+(λ+μ+α2M)eαMζ=ρu¨+ρfw¨
(7)(αMeMζ)=(ρfu¨+mw¨)+bw˙

Considering the simply supported boundary conditions, that is

(8)σ11=0,u2=u3=0,w2=w3=0,atx1=0,L1;σ22=0,u1=u3=0,w1=w3=0,atx2=0,L2.

3 The exact solution for fluid-saturated porous rectangular plates

A solution for the displacement field in the fluid-saturated porous rectangular plate is sought in the form.

(9)u1=m1=0,n1=0U1(x3)cosm1πx1L1sinn1πx2L2eiwt,w1=m1=0,n1=0W1(x3)cosm1πx1L1sinn1πx2L2eiwtu2=m1=0,n1=0U2(x3)sinm1πx1L1cosn1πx2L2eiwt,w2=m1=0,n1=0W2(x3)sinm1πx1L1cosn1πx2L2eiwtu3=m1=0,n1=0U3(x3)sinm1πx1L1sinn1πx2L2eiwt,w3=m1=0,n1=0W3(x3)sinm1πx1L1sinn1πx2L2eiwt

where ω denotes the angular frequency. Assume that ξ1 = m1π/L1, ξ2 = n1π/L2 and that m1 and n1 are positive integers. The chosen displacement field described by Eq. (9) satisfies the homogeneous boundary conditions described by Eq. (8) at simply supported boundaries. Substituting Eqs. (3) and (9) into the constitutive equation of Eq. (1), the stresses can be given in terms of displacements as

(10){σ11=m1,n1[(λ+2μ+α2M)ξ1U1αMξ1W1(λ+α2M)ξ2U2αMξ2W2+(λ+α2M)U3+αMW3]sinξ1x1sinξ2x2eiωtσ22=m1,n1[(λ+α2M)ξ1U1αMξ1W1(λ+2μ+α2M)ξ2U2αMξ2W2+(λ+α2M)U3+αMW3]sinξ1x1sinξ2x2eiωtσ33=m1,n1[(λ+α2M)ξ1U1αMξ1W1(λ+α2M)ξ2U2αMξ2W2+(λ+2μ+α2M)U3+αMW3]sinξ1x1sinξ2x2eiωtσ12=m1,n1μ(ξ2U1+ξ1U2)cosξ1x1cosξ2x2eiωtσ13=m1,n1μ(U1+ξ1U3)cosξ1x1sinξ2x2eiωtσ23=m1,n1μ(U2+ξ2U3)sinξ1x1cosξ2x2eiωtp=m1,n1(αMξ1U1+Mξ1W1+αMξ2U2+Mξ2W2αMU3MW3)sinξ1x1cosξ2x2eiωt

where a prime denotes differentiation with respect to x3. Substituting Eq. (9) into Eqs. (6) and (7), the following coupled system of second-order ordinary differential equations can be obtained:

(11a)[ρω2(ξ12+ξ22)μ(λ+μ+α2M)]ξ12U1+μU1+(ρfω2αMξ12)W1(λ+μ+α2M)ξ1ξ2U2αMξ1ξ2W2+(λ+μ+α2M)ξ1U3+αMξ1W3=0
(11b)(λ+μ+α2M)ξ1ξ2U1αMξ1ξ2W1+[ρw2(ξ12+ξ22)μ(λ+μ+α2M)ξ22]U2+μU2+(ρfw2αMξ22)W2+(λ+μ+α2M)ξ2U3+αMξ2W3=0
(11c)(λ+μ+α2M)ξ1U1αMξ1W1(λ+μ+α2M)ξ2U2αMξ2W2+[ρω2μ(ξ12+ξ22)]U3+(λ+2μ+α2M)U3+ρfω2W3+αMW3=0
(11d)(ρfω2αMξ12)U1+(mω2bωiMξ12)W1αMξ1ξ2U2Mξ1ξ2W2+αMξ1U3+Mξ1W3=0
(11e)αMξ1ξ2U1Mξ1ξ2W1+(ρfω2αMξ22)U2+(mω2bωiMξ22)W2+αMξ2U3+Mξ2W3=0
(11f)αMξ1U1Mξ1W1αMξ2U2Mξ2W2+ρfω2U3+αMU3+(mω2bωi)W3+MW3=0

The power series method is used to obtain the solution for Eq. (11). According to a general method, we assume a power series solution for the displacement functions Ui(x3) and Wi(x3) as

(12)Ui(x3)=β=0U˜i(β)x3β,Wi(x3)=β=0W˜i(β)x3β

Substituting Eq. (12) into Eq. (11) yields the following coupled recurrence algebraic relations:

(13a)W˜1(β+1)=ρfω2mω2bωiU˜1(β+1)+ρfω2ξ1(mω2bωi)(β+1)U˜3(β)+ξ1β+1W˜3(β)
(13b)W˜2(β+1)=ρfω2mω2bωiU˜2(β+1)+ρfω2ξ2(mω2bωi)(β+1)U˜3(β)+ξ2β+1W˜3(β)
(13c)W˜3(β+1)=αM(ξ12+ξ22)ρfω2Mξ1(β+1)U˜1(β)+ρfω2ξ2αξ2(mω2bωi)ρfω2(β+1)W˜2(β)+(αMξ22ρfω2)(mω2bωi)+ρfω2Mξ12ρfω2Mξ1(β+1)W˜1(β)αU˜3(β+1)
(13d)U˜1(β+2)=(αρfρ)ω2+(ξ12+ξ22)(λ+2μ)μ(β+2)(β+1)U˜1(β)(λ+μ)ξ1ξ2(mω2bωi)μρfω2(β+2)(β+1)W˜2(β)+[αρfω2+(λ+μ)ξ22](mω2bωi)(ρfω2)2μρfω2(β+2)(β+1)W˜1(β)(λ+μ)ξ1μ(β+2)U˜3(β+1)
(13e)U˜2(β+2)=ξ2[(λ+2μ)(ξ12+ξ22)+(αρfρ)ω2]μξ1(β+2)(β+1)U˜1(β)(λ+μ)ξ2μ(β+2)U˜3(β+1)+(mω2bωi)ξ2[(αρfρ)ω2+(ξ12+ξ22)μ+(λ+μ)ξ22]μρfω2ξ1(β+2)(β+1)W˜1(β)(mω2bωi)[ρω2+μ(ξ12+ξ22)+(λ+μ)ξ22](ρfω2)2μρfω2(β+2)(β+1)W˜2(β)
(13f)U˜3(β+2)=(λ+μ)ξ1(λ+2μ)(β+2)U˜1(β+1)+(λ+μ)ξ2(λ+2μ)(β+2)U˜2(β+1)+(αρfρ)ω2+μ(ξ12+ξ22)(λ+2μ)(β+2)(β+1)U˜3(β)+α(mω2bωi)ρfω2(λ+2μ)(β+2)(β+1)W˜3(β)

Clearly, the recurrence relations described in Eq. (13) are evaluated successively for β = 0, 1, … , to obtain U˜i(β) and W˜i(β) In terms of the eight arbitrary constants U˜1(0), W˜1(0), W˜2(0), U˜3(0), W˜3(0), U˜1(1), U˜2(1), and U˜3(1) and the angular frequency ω as

(14a)U˜i(β)=Ai1(β)(ω)U˜1(0)+Ai2(β)(ω)W˜1(0)+Ai3(β)(ω)W˜2(0)+Ai4(β)(ω)U˜3(0)+Ai5(β)(ω)W˜3(0)+Ai6(β)(ω)U˜1(1)+Ai7(β)(ω)U˜2(1)+Ai8(β)(ω)U˜3(1)
(14b)W˜i(β)=Bi1(β)(ω)U˜1(0)+Bi2(β)(ω)W˜1(0)+Bi3(β)(ω)W˜2(0)+Bi4(β)(ω)U˜3(0)+Bi5(β)(ω)W˜3(0)+Bi6(β)(ω)U˜1(1)+Bi7(β)(ω)U˜2(1)+Bi8(β)(ω)U˜3(1)

Here, Aij(β)(ω) and Bij(β)(ω) are known polynomials in ω and they are determined by the recurrence formula in Eq. (13).

Substituting Eq. (14) into Eq. (12) gives the rewritten displacement expression as

(15a)Ui(x3)=Ai1(x3)U˜1(0)+Ai2(x3)W˜1(0)+Ai3(x3)W˜2(0)+Ai4(x3)U˜3(0)+Ai5(x3)W˜3(0)+Ai6(x3)U˜1(1)+Ai7(x3)U˜2(1)+Ai8(x3)U˜3(1)
(15b)Wi(x3)=Bi1(x3)U˜1(0)+Bi2(x3)W˜1(0)+Bi3(x3)W˜2(0)+Bi4(x3)U˜3(0)+Bi5(x3)W˜3(0)+Bi6(x3)U˜1(1)+Bi7(x3)U˜2(1)+Bi8(x3)U˜3(1)

where Aij(x3)=β=0Aij(β)x3β, and Bij(x3)=β=0Bij(β)x3β. The degree of each of the polynomials increases as more terms are retained in the series expansion in Eq. (12).

Substituting the expressions of displacement in Eq. (15) into the stress components in Eq. (10) represented by displacement yields the following expression of stress:

(16)S=QijU˜

where S = [σ33σ13σ23p]T, U˜=[U˜1(0)W˜1(0)U˜2(0)W˜2(0)U˜3(0)W˜3(0)U˜1(1)U˜3(1)]T, and the matrix of coefficients Qij is a 4 × 8 matrix for each i, j given in Appendix A.

At this point, a series solution of the displacement and stress components are obtained.

3.1 The free vibration of fluid-saturated porous rectangular plates

If the upper and lower surfaces of the plate are both freely permeable, and the boundary conditions for the free vibration are:

(17)σ33=0,σ13=σ23=0,p=0atx3=H/2σ33=0,σ13=σ23=0,p=0atx3=H/2

Therefore, substituting Eq. (16) into the boundary conditions in Eq. (17) yields the frequency equation in matrix form:

(18)Qij(x3)U˜=0

where Qij(x3) is an 8 × 8 matrix and given in Appendix B.

If the upper and lower surfaces of the plate are impermeable, the boundary conditions for the free vibration change to be:

(19)σ33=0,σ13=σ23=0,w3=0atx3=H/2σ33=0,σ13=σ23=0,w3=0atx3=H/2

Similarly, Eqs. (15b), (16) and (19) can obtain the matrix form of Eq. (20):

(20)Gij(x3)U˜=0

where Gij(x3) is an 8 × 8 matrix and given in Appendix C.

The left coefficient determinant of the above equation needs to be set to zero to ensure that a nontrivial solution of Eq. (18) or Eq. (20) exists. Thus, the free vibration frequency of fluid-saturated porous rectangular plates can be obtained if the upper or lower surfaces are either freely permeable or completely impermeable.

3.2 The forced vibration of fluid-saturated porous rectangular plates

Since an arbitrary load function can be expanded as a double Fourier series in x1 and x2, considering the upper and lower surfaces of the plate to be freely permeable, the boundary conditions for the forced vibration with the harmonic load can be expressed as:

(21)σ33=qeiωtsinξ1x1sinξ2x2,σ13=σ23=0,p=0atx3=H/2σ33=0,σ13=σ23=0,p=0atx3=H/2

where q is the amplitude of the normal loads applied on the top surfaces.

Eqs. (16) and (21) can be combined to obtain

(22)Qij(x3)U˜=P

where P=[q0000000]T is a vector of length 8.

If the upper and lower surfaces of the plate are impermeable, the boundary conditions for the forced vibration are:

(23)σ33=qeiωtsinξ1x1sinξ2x2,σ13=σ23=0,w3=0atx3=H/2σ33=0,σ13=σ23=0,w3=0atx3=H/2

Similarly, Eq. (24) can be obtained:

(24)Gij(x3)U˜=P

After determining the constants U˜1(0), W˜1(0), W˜2(0), U˜3(0), W˜3(0), U˜1(1), U˜2(1), and U˜3(1) with Eq. (22) or Eq. (24), the transient response of fluid-saturated porous rectangular plates with harmonic load can be computed by combining Eqs. (15) and (16).

4 Results and discussion

4.1 Results validation

To verify the validity of the presented solution in this paper, the fluid-saturated porous rectangular plates are degenerated into single-phase solid rectangular plates by ignoring the pore fluid pressure p and setting n = 0, ρf = 0 and η = 0. Then, Eq. (18) transforms into

(25)Qij(x3)U˜=0

where Qij(x3) is a 6 × 6 matrix and given in Appendix D. U˜=[U˜1(0)U˜2(0)U˜3(0)U˜1(1)U˜2(1)U˜3(1)]T.

The parameters selected for the numerical simulation of a fluid-saturated porous rectangular plate are as follows [24]:

E=1.44×1010Pa,ν=0.2,ρs=2.458kg/m3,ρf=1.000kg/m3,Ks=3.6×1010pa,L1=L2=1m,Kf=3.3×1010pa,n=0.19andH=0.1m

where E is Young’s modulus and ν is Poisson’s ratio.

The dimensionless fundamental frequencies ω=ωL12HρE obtained from Eq. (25) are given in Table 1. The listed values in Table 1 indicate that our results are in excellent agreement with those presented by Senthil and Bara [27]. We have also listed the natural frequencies computed from the three plate theories in Table 1 to compare the exact results with those obtained from classical plate theory (CPT) [28], first-order shear deformation theory (FSDT) [29] and third-order shear deformation theory (TSDT) [30]. The comparative results show the validity of the proposed algorithm.

Table 1:

Comparison of exact natural frequencies with natural frequencies from the literature with 10 terms in the series solution.

Theoryω1,1(1)ω1,1(2)ω1,1(3)
L1/H = 10
Present analysis5.776227.54646.498
Senthil S.Vel [27]5.776927.55446.503
CPT [28]5.924827.55446.574
FSDT [29]5.769327.55446.574
TSDT [30]5.731727.55446.574

4.2 Free vibration

To analyse the influence of the surface infiltration conditions on the frequency of the fluid-saturated porous rectangular plates, the above calculation parameters are adopted to calculate the natural frequencies of plates according to Eqs. (18) and (20), and the finite series truncated term is β = 10. The boundary conditions on the upper and lower surfaces are permeable and completely impermeable, respectively. The results are listed in Table 2, showing that the surface infiltration conditions have a great influence on the basic natural frequency of vibration but have little impact on the higher-order frequency.

Table 2:

Natural frequencies of the fluid-saturated porous rectangular plates with permeable and completely impermeable upper and lower surfaces.

L1/H = 10ω1,1(1)ω1,1(2)ω1,1(3)ω1,1(4)ω1,1(5)
Free permeable4.132928.1538192.3431310.9472380.5576
Completely impermeable6.417741.8564192.2777309.8759379.9566

The natural frequencies are calculated according to Eqs. (18) and (20) under different porosities, permeability coefficients and boundary permeability conditions. The results are listed in Tables 3–6. Tables 3 and 4 show that the natural frequencies of the fluid-saturated porous plates increase with the permeability coefficient whether the upper and lower surfaces are permeable or completely impermeable. Furthermore, the permeability coefficient has more of an influence on the natural frequencies if the upper and lower surfaces are freely permeable. Tables 5 and 6 indicate that the natural frequencies decrease with increasing porosity, whether the upper and lower surfaces are permeable or completely impermeable, and that the porosity has a greater influence on the natural frequencies. These phenomena result from the effect of deformation coupling between the solid and fluid.

Table 3:

Variations in the natural frequencies of fluid-saturated porous rectangular plates with kf under freely permeable conditions.

L1/H = 10ω1,1(1)ω1,1(2)ω1,1(3)ω1,1(4)ω1,1(5)
kf = 1 × 10−113.266028.1499192.3364310.9297380.5646
kf = 5 × 10−113.429428.1538192.3369310.9306380.5649
kf = 1 × 10−103.703028.1538192.3374310.9448380.5650
kf = 5 × 10−105.116328.1551192.3376310.9455380.5650
kf = 1 × 10−96.017628.1593192.3378310.9468380.5651
kf = 5 × 10−96.139228.2661192.3462310.9827380.5691
Table 4:

Variations in the natural frequencies of fluid-saturated porous rectangular plates with kf under freely impermeable conditions.

L1/H = 10ω1,1(1)ω1,1(2)ω1,1(3)ω1,1(4)ω1,1(5)
kf = 1 × 10−116.385741.8547192.2755309.8590379.9485
kf = 5 × 10−116.414341.8562192.2759309.8705379.9533
kf = 1 × 10−106.415841.8562192.2768309.8790379.9553
kf = 5 × 10−106.431441.8563192.2769309.8795379.9564
kf = 1 × 10−96.457941.8564192.2772309.8810379.9570
kf = 5 × 10−96.464641.8568192.2860309.9202379.9594
Table 5:

Variations in the natural frequencies of fluid-saturated porous rectangular plates with n under freely permeable conditions.

L1/H = 10ω1,1(1)ω1,1(2)ω1,1(3)ω1,1(4)ω1,1(5)
n = 0.17.545733.0610194.0023311.1301410.9468
n = 0.27.209727.4169192.2516311.3392378.6239
n = 0.36.090025.7622191.2195308.6478361.0776
n = 0.45.411422.7983190.6838307.9684351.5924
n = 0.55.151319.3084190.3310307.1975345.4127
Table 6:

Variations in the natural frequencies of fluid-saturated porous rectangular plates with n under freely impermeable conditions.

L1/H = 10ω1,1(1)ω1,1(2)ω1,1(3)ω1,1(4)ω1,1(5)
n = 0.18.104645.2650193.9892310.7382410.8423
n = 0.26.301941.5898192.1471309.7844377.6833
n = 0.35.547939.6392191.2456308.8960361.2867
n = 0.45.205038.4804190.6969308.0337351.6839
n = 0.55.044037.7265190.3310307.2106345.4780

4.3 Harmonic vibration

The response of the fluid-saturated porous rectangular plates under the dynamic load is analysed with the harmonic load q = 1000 pa. The results for the forced vibration are plotted in Figure 1 with the harmonic load and forcing frequencies ω = 5, 10, 20 and 50. The solid skeleton stress, the pore fluid pressure, the solid skeleton displacement and the pore fluid displacement of the fluid-saturated porous rectangular plates decrease with increasing forcing frequencies whether the upper and lower surfaces are permeable or completely impermeable, but the transverse shear stress σ23 and the transverse normal stress σ33 of the solid skeleton are influenced only slightly. Due to the strong transverse pressure gradients that occur on the beam sections and are undertaken by both the solid skeleton and pore fluid as the beam bends, the transverse bending deformation causes compression in the upper beam for x3 < h/2 and tension in the lower beam for x3 > h/2, which is similar to the stress distribution in a single-phase continuous elastic beam. The relationship between the total pressure and pore pressure leads to positive pore pressures in the compression zone of the upper beam and negative pore pressures in the tension zone of the lower beam, resulting in pore suction, as shown in Figure 1.

Figure 1: The variations in the solid skeleton stress, pore fluid pressure, solid skeleton displacement and pore fluid displacement in the thickness direction with ω‾$&#x203e;{\omega }$.(1), (3), (5), (7), (9), (11), (13) and (15) are under freely permeable conditions, and (2), (4), (6), (8), (10), (12), (14) and (16) are under completely impermeable conditions.
Figure 1: The variations in the solid skeleton stress, pore fluid pressure, solid skeleton displacement and pore fluid displacement in the thickness direction with ω‾$&#x203e;{\omega }$.(1), (3), (5), (7), (9), (11), (13) and (15) are under freely permeable conditions, and (2), (4), (6), (8), (10), (12), (14) and (16) are under completely impermeable conditions.
Figure 1:

The variations in the solid skeleton stress, pore fluid pressure, solid skeleton displacement and pore fluid displacement in the thickness direction with ω.

(1), (3), (5), (7), (9), (11), (13) and (15) are under freely permeable conditions, and (2), (4), (6), (8), (10), (12), (14) and (16) are under completely impermeable conditions.

Considering different values of porosity, the variations in the solid skeleton stress, pore fluid pressure and displacements of the solid skeleton and fluid in the thickness direction are shown in Figure 2 with a frequency of ω = 10. The results indicate that porosity affects the solid skeleton stress, pore fluid pressure and solid skeleton and fluid displacements, although the upper and lower surface infiltration conditions have a more significant effect.

Figure 2: The variations in the solid skeleton stress, pore fluid pressure, solid skeleton displacement and pore fluid displacement in the thickness direction with n.(1), (3), (5), (7), (9), (11), (13) and (15) are under freely permeable conditions, and (2), (4), (6), (8), (10), (12), (14) and (16) are under completely impermeable conditions.
Figure 2: The variations in the solid skeleton stress, pore fluid pressure, solid skeleton displacement and pore fluid displacement in the thickness direction with n.(1), (3), (5), (7), (9), (11), (13) and (15) are under freely permeable conditions, and (2), (4), (6), (8), (10), (12), (14) and (16) are under completely impermeable conditions.
Figure 2:

The variations in the solid skeleton stress, pore fluid pressure, solid skeleton displacement and pore fluid displacement in the thickness direction with n.

(1), (3), (5), (7), (9), (11), (13) and (15) are under freely permeable conditions, and (2), (4), (6), (8), (10), (12), (14) and (16) are under completely impermeable conditions.

To analyse the influence of the permeability coefficient on the dynamic responses of plates, Figure 3 shows the variations in the solid skeleton stress, pore fluid pressure and solid skeleton and fluid displacements in the thickness direction. The results indicate that the permeability coefficient has an impact on the solid skeleton stress, pore fluid pressure and solid skeleton and fluid displacements. Moreover, these factors can be affected significantly by the surface infiltration conditions.

Figure 3: The variations in the solid skeleton stress, pore fluid pressure, solid skeleton displacement and pore fluid displacement in the thickness direction with kf.(1), (3), (5), (7), (9), (11), (13) and (15) are under freely permeable conditions, and (2), (4), (6), (8), (10), (12), (14) and (16) are under completely impermeable conditions.
Figure 3: The variations in the solid skeleton stress, pore fluid pressure, solid skeleton displacement and pore fluid displacement in the thickness direction with kf.(1), (3), (5), (7), (9), (11), (13) and (15) are under freely permeable conditions, and (2), (4), (6), (8), (10), (12), (14) and (16) are under completely impermeable conditions.
Figure 3:

The variations in the solid skeleton stress, pore fluid pressure, solid skeleton displacement and pore fluid displacement in the thickness direction with kf.

(1), (3), (5), (7), (9), (11), (13) and (15) are under freely permeable conditions, and (2), (4), (6), (8), (10), (12), (14) and (16) are under completely impermeable conditions.

5 Conclusions

A series solution is presented for the dynamic responses of a simply supported fluid-saturated porous rectangular plate. Considering the compressibility of solid particles and fluid and the viscosity of pore fluid, the dynamic responses of simply supported fluid-saturated porous rectangular plates and the influence of the surface infiltration conditions, porosity and pore fluid permeability coefficient on the free vibration frequency of porous plates are discussed. Parametric studies indicate that the effect of coupling between a solid and fluid is important for increasing the frequency and must be considered in the case of dynamic responses.


Corresponding author: Wang Lixian, School of Civil Engineering, Lanzhou University of Technology, Langongping Road 287#, Lanzhou730050, People’s Republic of China, E-mail:

Award Identifier / Grant number: 51368038

  1. Author contribution: The author has accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This study was funded by Chinese Natural Science Foundation (Grant No. 51978320).

  3. Conflict of interest statement: The author declares no conflicts of interest regarding this article.

Appendix A
Q11=[(λ+α2M)ξ1A11(x3)(λ+α2M)ξ2A21(x3)+(λ+2μ+α2M)A31(x3)αMξ1B11(x3)αMξ2B21(x3)+αMB31(x3)]
Q12=[(λ+α2M)ξ1A12(x3)(λ+α2M)ξ2A22(x3)+(λ+2μ+α2M)A32(x3)αMξ1B12(x3)αMξ2B22(x3)+αMB32(x3)]
Q13=[(λ+α2M)ξ1A13(x3)(λ+α2M)ξ2A23(x3)+(λ+2μ+α2M)A33(x3)αMξ1B13(x3)αMξ2B23(x3)+αMB33(x3)]
Q14=[(λ+α2M)ξ1A14(x3)(λ+α2M)ξ2A24(x3)+(λ+2μ+α2M)A34(x3)αMξ1B14(x3)αMξ2B24(x3)+αMB34(x3)]
Q15=[(λ+α2M)ξ1A15(x3)(λ+α2M)ξ2A25(x3)+(λ+2μ+α2M)A35(x3)αMξ1B15(x3)αMξ2B25(x3)+αMB35(x3)]
Q16=[(λ+α2M)ξ1A16(x3)(λ+α2M)ξ2A26(x3)+(λ+2μ+α2M)A36(x3)αMξ1B16(x3)αMξ2B26(x3)+αMB36(x3)]
Q17=[(λ+α2M)ξ1A17(x3)(λ+α2M)ξ2A27(x3)+(λ+2μ+α2M)A37(x3)αMξ1B17(x3)αMξ2B27(x3)+αMB37(x3)]
Q18=[(λ+α2M)ξ1A18(x3)(λ+α2M)ξ2A28(x3)+(λ+2μ+α2M)A38(x3)αMξ1B18(x3)αMξ2B28(x3)+αMB38(x3)]
Q21=μA11(x3)+μξ1A31(x3),Q22=μA12(x3)+μξ1A32(x3),Q23=μA13(x3)+μξ1A33(x3)
Q24=μA14(x3)+μξ1A34(x3),Q25=μA15((x3)+μξ1A35(x3),Q26=μA16(x3)+μξ1A36(x3)
Q27=μA17(x3)+μξ1A37(x3),Q28=μA18(x3)+μξ1A38(x3)
Q31=μA21(x3)+μξ2A31(x3),Q32=μA22(x3)+μξ2A32(x3),Q33=μA23(x3)+μξ2A33(x3)
Q34=μA24(x3)+μξ2A34(x3),Q35=μA25(x3)+μξ2A35(x3),Q36=μA26(x3)+μξ2A36(x3)
Q37=μA27(x3)+μξ2A37(x3),Q38=μA28(x3)+μξ2A38(x3)
Q41=αMξ1A11(x3)+αMξ2A21(x3)αMA31(x3)+Mξ1B11(x3)+Mξ2B21(x3)MB31(x3)
Q42=αMξ1A12(x3)+αMξ2A22(x3)αMA32(x3)+Mξ1B12(x3)+Mξ2B22(x3)MB32(x3)
Q43=αMξ1A13(x3)+αMξ2A23(x3)αMA33(x3)+Mξ1B13(x3)+Mξ2B23(x3)MB33(x3)
Q44=αMξ1A14(x3)+αMξ2A24(x3)αMA34(x3)+Mξ1B14(x3)+Mξ2B24(x3)MB34(x3)
Q45=αMξ1A15(x3)+αMξ2A25(x3)αMA35(x3)+Mξ1B15(x3)+Mξ2B25(x3)MB35(x3)
Q46=αMξ1A16(x3)+αMξ2A26(x3)αMA36(x3)+Mξ1B16(x3)+Mξ2B26(x3)MB36(x3)
Q47=αMξ1A17(x3)+αMξ2A27(x3)αMA37(x3)+Mξ1B17(x3)+Mξ2B27(x3)MB37(x3)
Q48=αMξ1A18(x3)+αMξ2A28(x3)αMA38(x3)+Mξ1B18(x3)+Mξ2B28(x3)MB38(x3)
Appendix B
Qij(x3)=[Q11(H/2)Q12(H/2)Q13(H/2)Q14(H/2)Q21(H/2)Q22(H/2)Q23(H/2)Q24(H/2)Q31(H/2)Q32(H/2)Q33(H/2)Q34(H/2)Q41(H/2)Q42(H/2)Q43(H/2)Q44(H/2)Q15(H/2)Q16(H/2)Q17(H/2)Q18(H/2)Q25(H/2)Q26(H/2)Q27(H/2)Q28(H/2)Q35(H/2)Q36(H/2)Q37(H/2)Q38(H/2)Q45(H/2)Q46(H/2)Q47(H/2)Q48(H/2)Q11(H/2)Q12(H/2)Q13(H/2)Q14(H/2)Q21(H/2)Q22(H/2)Q23(H/2)Q24(H/2)Q31(H/2)Q32(H/2)Q33(H/2)Q34(H/2)Q41(H/2)Q42(H/2)Q43(H/2)Q44(H/2)Q15(H/2)Q16(H/2)Q17(H/2)Q18(H/2)Q25(H/2)Q26(H/2)Q27(H/2)Q28(H/2)Q35(H/2)Q36(H/2)Q37(H/2)Q38(H/2)Q45(H/2)Q46(H/2)Q47(H/2)Q48(H/2)]
Appendix C
Gij(x3)=[Q11(H/2)Q12(H/2)Q13(H/2)Q14(H/2)Q21(H/2)Q22(H/2)Q23(H/2)Q24(H/2)Q31(H/2)Q32(H/2)Q33(H/2)Q34(H/2)B31(H/2)B32(H/2)B33(H/2)B34(H/2)Q15(H/2)Q16(H/2)Q17(H/2)Q18(H/2)Q25(H/2)Q26(H/2)Q27(H/2)Q28(H/2)Q35(H/2)Q36(H/2)Q37(H/2)Q38(H/2)B35(H/2)B36(H/2)B37(H/2)B38(H/2)Q11(H/2)Q12(H/2)Q13(H/2)Q14(H/2)Q21(H/2)Q22(H/2)Q23(H/2)Q24(H/2)Q31(H/2)Q32(H/2)Q33(H/2)Q34(H/2)B31(H/2)B32(H/2)B33(H/2)B34(H/2)Q15(H/2)Q16(H/2)Q17(H/2)Q18(H/2)Q25(H/2)Q26(H/2)Q27(H/2)Q28(H/2)Q35(H/2)Q36(H/2)Q37(H/2)Q38(H/2)B35(H/2)B36(H/2)B37(H/2)B38(H/2)]
Appendix D
Qij(x3)=[Q11(H/2)Q12(H/2)Q13(H/2)Q21(H/2)Q22(H/2)Q23(H/2)Q31(H/2)Q32(H/2)Q33(H/2)Q14(H/2)Q15(H/2)Q16(H/2)Q24(H/2)Q25(H/2)Q26(H/2)Q34(H/2)Q35(H/2)Q36(H/2)Q11(H/2)Q12(H/2)Q13(H/2)Q21(H/2)Q22(H/2)Q23(H/2)Q31(H/2)Q32(H/2)Q33(H/2)Q14(H/2)Q15(H/2)Q16(H/2)Q24(H/2)Q25(H/2)Q26(H/2)Q34(H/2)Q35(H/2)Q36(H/2)]

References

[1] M. A. Biot, “Theory of propagation of elastic waves in a fluid-saturated porous solid (Ⅰ.low-frequency range),” J. Acoust. Soc. Am., vol. 28, no. 2, pp. 168–178, 1956, https://doi.org/10.1121/1.1908239.Search in Google Scholar

[2] M. A. Biot, “Theory of propagation of elastic waves in a fluid-saturated porous solid (Ⅱ.high-frequency range),” J. Acoust. Soc. Am., vol. 28, no. 2, pp. 179–191, 1956, https://doi.org/10.1121/1.1908241.Search in Google Scholar

[3] A. Gajo and L. Mongiovi, “An analytical solution for the transient response of saturated linear elastic porous media,” Int. J. Rock Mech. Min. Sci. Geomech. Abstr., vol. 3, pp. 399–413, 1996, https://doi.org/10.1002/nag.1610190603.Search in Google Scholar

[4] X. Yang and Y. Pan, “Axisymmetrical analytical solution for vertical vibration of end-bearing pile in saturated viscoelastic soil layer,” Appl. Math. Mech., vol. 2, pp. 193–204, 2010, https://doi.org/10.1007/s10483-010-0207-7.Search in Google Scholar

[5] B. Bai, “Analytical solutions of thermal consolidation for a hollow cylinder saturated porous medium,” Rock Soil Mech., vol. 32, pp. 2901–2906, 2011.Search in Google Scholar

[6] X. L. Zhou, B. Xu, J. H. Wang, and Y. L. Li, “An analytical solution for wave-induced seabed response in a multi-layered poro-elastic seabed,” Ocean Eng., vol. 38, pp. 119–129, 2011, https://doi.org/10.1016/j.oceaneng.2010.10.003.Search in Google Scholar

[7] O. C. Zienkiewicz and T. Shiomi, “Dynamic behaviour of saturated porous media: the generalized biot form formulation and its numerical solution,” Int. J. Numer. Anal. Methods GeoMech., vol. 8, pp. 71–96, 1984, https://doi.org/10.1002/nag.1610080106.Search in Google Scholar

[8] Y. He and B. Han, “A wavelet finite-difference method for numerical simulation of wave propagation in fluid-saturated porous media,” Appl. Math. Mech., vol. 11, pp. 1495–1504, 2008, https://doi.org/10.1007/s10483-008-1110-y.Search in Google Scholar

[9] M. Schanz, “Poroelastodynamics: linear models, analytical solutions, and numerical methods,” Appl. Mech. Rev., vol. 62, pp. 1–15, 2009, https://doi.org/10.1115/1.3090831.Search in Google Scholar

[10] F. X. Zhou, and Y. M. Lai, “Transient dynamic analysis of gradient fluid-saturated soil,” Chin. J. Theor. Appl. Mech., vol. 44, pp. 943–947, 2012.Search in Google Scholar

[11] F. Zhou and Q. Ma, “Propagation of Rayleigh waves in fluid-saturated non-homogeneous soils with the graded solid skeleton distribution,” Int. J. Numer. Anal. Methods GeoMech., vol. 40, no. 11, pp. 1513–1530, 2016, https://doi.org/10.1002/nag.2491.Search in Google Scholar

[12] M. Sharma, “Wave propagation in thermoelastic saturated porous medium,” J. Earth Syst. Sci., vol. 117, pp. 951–958, 2008, https://doi.org/10.1007/s12040-008-0080-4.Search in Google Scholar

[13] V. S. Polenov and A. V. Chigarev, “Wave propagation in a fluid-saturated inhomogeneous porous medium,” J. Appl. Math. Mech., vol. 74, pp. 198–203, 2010, https://doi.org/10.1016/j.jappmathmech.2010.05.010.Search in Google Scholar

[14] Q. Ma and F. Zhou, “Propagation conditions of Rayleigh waves in nonhomogeneous saturated porous media,” Soil Mech. Found. Eng., vol. 53, no. 4, pp. 268–273, 2016, https://doi.org/10.1007/s11204-016-9397-1.Search in Google Scholar

[15] L. A. Taber, “A theory for transverse deflection of poroelastic plates,” ASME J. Appl. Mech., vol. 59, pp. 628–634, 1992, https://doi.org/10.1115/1.2893770.Search in Google Scholar

[16] L. P. Li, G. Cederbaum, and K. Schulgasser, “Theory of poroelastic plates with in-plane diffusion,” Int. J. Solid Struct., vol. 34, pp. 4515–4530, 1997, https://doi.org/10.1016/s0020-7683(97)00045-0.Search in Google Scholar

[17] P. Leclaire, K. V. Horoshenkov, and A. Cummings, “Transverse vibration of a thin rectangular porous platesaturated by a fluid,” J. Sound Vib., vol. 247, pp. l–18, 2001, https://doi.org/10.1006/jsvi.2001.3656.Search in Google Scholar

[18] P. Leclaire and K. V. Horoshenkov, “The vibrational response of aclamped rectangular porous plate,” J. Sound Vib., vol. 247, pp. 19–31, 2001, https://doi.org/10.1006/jsvi.2000.3657.Search in Google Scholar

[19] L. A. Taber, and A. M. Puleo, “Poroelastic plate and shell theories,” Solid Mech. Appl., vol. 35, pp. 323–338, 1996, https://doi.org/10.1007/978-94-015-8698-6_18.Search in Google Scholar

[20] P. Leclaire, K. V. Horoshenkov, and A. Cummings, “Transverse vibrations of a thin rectangular porous plate saturated by a fluid,” J. Sound Vib., vol. 247, pp. 1–18, 2001, https://doi.org/10.1006/jsvi.2001.3656.Search in Google Scholar

[21] D. D. Theodorakopoulos and D. E. Beskos, “Flexural vibrations of poroelastic plate,” Acta Mech., vol. 103, nos 1-4, pp. 191–203, 1994, https://doi.org/10.1007/bf01180226.Search in Google Scholar

[22] Anke Busse Dipl.Ing, “Martin Schanz, Heinz antes. A poroelastic Mindlin-plate,” Pammatone, vol. 3, no. 1, pp. 260–261, 2003. https://doi.org/10.1002/pamm.200310402.Search in Google Scholar

[23] Z. Feng-xi and C. Xiao-lin, “A transverse dynamic deflection model for thin plate made of saturated porous materials,” Z. Naturforsch., vol. 71, no. 10, pp. 94–948, 2016, https://doi.org/10.1515/zna-2016-0208.Search in Google Scholar

[24] L. Nagler and M. Schanz, “An extendable poroelastic plate formulation in dynamics,” Arch. Appl. Mech., vol. 80, pp. 1177–1195, 2010, https://doi.org/10.1007/s00419-010-0429-4.Search in Google Scholar

[25] A. S. Rezaei and A. R. Saidi, “On the effect of coupled solid-fluid deformation on natural frequencies of fluid saturated porous plates[J],” Eur. J. Mech. Solid., vol. 63, no. (Complete), pp. 99–109, 2017, https://doi.org/10.1016/j.euromechsol.2016.12.006.Search in Google Scholar

[26] J. R. Fan, Exact Theory of Thick Laminated Plate and Shell, Beijing, Science Press, 1996.Search in Google Scholar

[27] S. V. Senthil and R. C. Batra, “Three-dimensional exact solution for the vibration of functionally graded rectangular plates,” J. Sound Vib., vol. 272, pp. 703–730, 2004, https://doi.org/10.1016/S0022-460X(03)00412-7.Search in Google Scholar

[28] N. A. Fleck and V. S. Deshpande, “The resistance of clamped sandwich beams to shock loading,” J. Appl. Mech. Trans. ASME, vol. 71, pp. 386–401, 2004, https://doi.org/10.1115/1.1629109.Search in Google Scholar

[29] M. T. Tilbrook, V. S. Deshpande, and N. A. Fleck, “The impulsive response of sandwich beams: analytical and numerical investigation of regimes of behaviour,” J. Mech. Phys. Solid., vol. 54, pp. 2242–2280, 2006, d, https://doi.org/10.1016/j.jmps.2006.07.001.Search in Google Scholar

[30] R. Rajendran and J. M. Lee, “Blast loaded plates,” Mar. Struct., vol. 22, no. 2, pp. 99–127, 2009, https://doi.org/10.1016/j.marstruc.2008.04.001.Search in Google Scholar

Received: 2020-04-28
Accepted: 2020-09-09
Published Online: 2020-10-21
Published in Print: 2020-11-18

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 14.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/zna-2020-0179/html
Scroll to top button