39
See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/336962719 Analytical solution for stress distribution around deep lined pressure tunnels under the water table Article in International Journal of Rock Mechanics and Mining Sciences · November 2019 DOI: 10.1016/j.ijrmms.2019.104124 CITATIONS 2 READS 324 5 authors, including: Some of the authors of this publication are also working on these related projects: Special Issue on 'Mitigating the Impacts of Mining', International Journal of Minerals, Metallurgy and Materials View project International database collaboration on cemented paste backfill (CPB) View project Xiangjian Dong University of Western Australia 15 PUBLICATIONS 148 CITATIONS SEE PROFILE Ali Karrech University of Western Australia 173 PUBLICATIONS 1,179 CITATIONS SEE PROFILE Chongchong Qi Central South University 72 PUBLICATIONS 1,141 CITATIONS SEE PROFILE Dr. Mohamed Elchalakani University of Western Australia 136 PUBLICATIONS 1,792 CITATIONS SEE PROFILE All content following this page was uploaded by Ali Karrech on 15 November 2019. The user has requested enhancement of the downloaded file.

Analytical solution for stress distribution arou nd deep

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Analytical solution for stress distribution arou nd deep

See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/336962719

Analytical solution for stress distribution around deep lined pressure tunnels

under the water table

Article  in  International Journal of Rock Mechanics and Mining Sciences · November 2019

DOI: 10.1016/j.ijrmms.2019.104124

CITATIONS

2READS

324

5 authors, including:

Some of the authors of this publication are also working on these related projects:

Special Issue on 'Mitigating the Impacts of Mining', International Journal of Minerals, Metallurgy and Materials View project

International database collaboration on cemented paste backfill (CPB) View project

Xiangjian Dong

University of Western Australia

15 PUBLICATIONS   148 CITATIONS   

SEE PROFILE

Ali Karrech

University of Western Australia

173 PUBLICATIONS   1,179 CITATIONS   

SEE PROFILE

Chongchong Qi

Central South University

72 PUBLICATIONS   1,141 CITATIONS   

SEE PROFILE

Dr. Mohamed Elchalakani

University of Western Australia

136 PUBLICATIONS   1,792 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Ali Karrech on 15 November 2019.

The user has requested enhancement of the downloaded file.

Page 2: Analytical solution for stress distribution arou nd deep

Analytical solution for stress distribution around deep lined

pressure tunnels under the water table

Xiangjian Dong, Ali Karrech∗, Chongchong Qi∗, Mohamed Elchalakani, Hakan Basarir

Department of Civil, Environmental and Mining Engineering, The University of Western Australia, 35Stirling Hwy, Crawley, WA 6009, Australia

Abstract

Existing models on the stress distribution around deep buried tunnels pay little attention

to the impact of rock-liner interaction and seepage variation. The conventional solution

for deep lined tunnels has an inherent mathematical restriction since it is limited to thin

liners. In this paper, we solve the poro-mechanical problem of stress and pore pressure

distribution using the complex variable method and the conformal mapping technique. The

proposed solution relaxes the assumptions of thin liner and negligible stresses that limit the

usefulness of existing approaches. To derive the solution, the stress function is expanded

into Laurent series in the complex plane. Analytic expressions for both with and without

gravity conditions are obtained. The proposed solution is compared with the Finite Element

Method; a good agreement has been obtained between the numerical simulations and pro-

posed analytical solution. It is found that with the liner installation, the effective maximum

principal stress (MPS) decreases by about 50 % for the given case condition. In addition,

the pore pressure distribution around the tunnel increases sharply compared to the no liner

condition. The parametric study shows that the effective MPS value is highly affected by

the liner relative thickness, stiffness and permeability. Its location on the tunnel boundary

moves from the horizontal direction to the invert with the increase of the liner thickness

and stiffness. Furthermore, the results show that low liner permeability results in high pore

pressure around the tunnel, which reduces the effective stress within the rock mass.

Keywords: Closed-form solution, Skin effect, Complex variable method, Deep excavation

1

Page 3: Analytical solution for stress distribution arou nd deep

1. Introduction

For deep underground tunnel construction, one of the important issues is the design of

lining with the consideration of groundwater flow into the tunnel [1, 2]. When an underwa-

ter tunnel is excavated, the groundwater may flow in especially when the infrastructure is

located under the water table. The effective stresses around the tunnel are affected by the

consequent seepage forces. Determining the stress distribution around the tunnel is essential

for its influence on the safety of workers and on the daily maintenance, especially when the

underground openings are excavated in areas with strong earthquake potential [3].

Analytical models can provide the stress and displacement fields within the whole domain

through rigorous mathematical derivation that take into account the relevant physical pa-

rameters [e.g. 4]. Analytical solutions are considered more advantageous than other methods

as they allow to explore the governing physics more effectively and to verify the correctness

of the results. However, numerical methods allow to investigate the poromechanics [5–8],

rock instabilities [9, 10] and the non-linear behaviour of geomaterials [11–14] that are dif-

ficult to reach analytically. Given simplified conditions, it is useful for the preliminary

design to obtain an early estimation of the excavation parameters. The complex variable

method has attracted research interest for its high precision and efficiency in solving various

opening problems [15–18]. For deep tunnels, the stress and displacement around circular

and/or non-circular shaped openings with homogeneous elastic media due to excavation are

obtained analytically by using complex variable method [4, 19–24]. However, in these ref-

erences the rock-liner interaction are ignored. The analytical solution proposed by Einstein

and Schwartz [25] was probably the first to consider the liner effect on the rock mass. At a

∗Corresponding authorEmail addresses: [email protected] (Xiangjian Dong),

[email protected] (Ali Karrech), [email protected] (Chongchong Qi)

Preprint submitted to Elsevier June 13, 2019

Page 4: Analytical solution for stress distribution arou nd deep

later stage, Bobet [26] illustrated the stress variation by considering the elastic transversely

anisotropic rock response of a circular lined tunnel. For non-circular lined tunnels, the ana-

lytical solutions were obtained by combining the complex variable method and the conformal

mapping function in elasticity [27–31]. Nevertheless, these solutions are restricted to thin

liners and they neglect the permeability of the liner.

For the pressure tunnel, the support design requires a good understanding of seepage

forces that are induced by interstitial fluid flow. The water inflow into a tunnel has been

substantially studied in recent years [32–37]. In these studies, the water tables were consid-

ered by using conformal mapping, which mapped the half-infinite fully saturated hydraulic

domain onto a circular ring. The complex variable method was used and combined with

specific boundary conditions in terms of constant pore water pressure or hydraulic head on

the tunnel boundary. However, the obtained solutions do not take the properties of liners

(e.g. thickness, stiffness and permeability) into consideration. Zhang et al. [38] consid-

ered the permeability of the liner and proposed a new set of solutions to study its effect

on the ground and tunnel behaviour due to steady water inflow. In addition, Li et al. [1]

investigated and compared the pore pressure expressions obtained from the complex vari-

able method, mirror image method and axisymmetric modelling method by considering the

effect of liner properties. The above mentioned analytic solutions are valid only for shallow

tunnel excavations. Being at depth, the problem we are focusing on here requires different

boundary conditions for stress functions, which correspond to different complex variable rep-

resentations. The problems were decomposed into far-field total stress and far-field uniform

water pressure (no water table assumption) [39–43]. Tran et al. [43] showed that the pore

pressure variation has a great influence on the distribution of stresses and displacements

around the tunnels.

The design of tunnels under high geo-stresses has been studied by many scholars; analyt-

ical and numerical solutions have been derived for steady seepage flow around openings [33–

3

Page 5: Analytical solution for stress distribution arou nd deep

44]. However, limited effort has been dedicated to obtain analytical expressions of effective

stress distribution around deep pressure tunnels considering the seepage effect. Moreover,

most of the existing analytical solutions were obtained either by neglecting the impact of the

lining permeability (without considering the thickness of liner) or by significantly simplify-

ing the boundary conditions (rock-liner interaction). In this paper, we derive an analytical

solution on the basis of superposition of three partial solutions, in which the underwater

lining-induced seepage problem in the half-infinite plane is solved by combining the complex

variable method and conformal mapping technique. The solution is validated by the FEM

modelling. The lining effect on the stress distribution is studied in details and the influ-

encing factors such as liner relative thickness, relative stiffness and relative permeability are

discussed.

2. Definition of the problem

2.1. Problem description

This paper describes the behaviour of deep lined tunnels subjected to far-field initial com-

pression geo-stresses, at a distance h below the water table (as shown in Fig. 1). The liner,

usually made of reinforced concrete, is installed to ensure the stability of the surrounding

rock mass. For an underwater tunnel, if the lining is designed to fully seal the water inflow

achieving a waterproof tunnel, the lining would have to withstand a high water pressure

for a long period, which would result in water leakage and subsequent tunnel lining failure.

If full drainage is implemented in the tunnel construction, a large-scale pumping system is

needed which increases the production cost. Therefore, for economic and safety purposes,

the liner is designed to block limited drainage. Its designed permeability is different from

the rock matrix and varies for different tunnel usages.

The following assumptions are made to obtain the analytical solution for the steady

seepage field of the deep lined tunnel. (1) the lined material and the surrounding rock mass

4

Page 6: Analytical solution for stress distribution arou nd deep

Figure 1: Analytic model for deep lined underwater tunnel.

are homogeneous, isotropic and elastic; (2) the cross-section is circular and suitable for plane

strain condition; (3) a steady-state flow and fully saturated conditions are assumed; (4) the

water table is horizontal and remains unchanged.

2.2. Theoretical framework

Taking water pressure into account, the effective stress is expressed as:

σ′ij = σij − δijαip (1)

where σij is the total stress; δij is the Kronecker symbol which is equal to one and zero when

i = j and i 6= j, respectively; αi represents Biot’s coefficient and p means pore pressure.

For the steady-state flow, the posed problem can be solved by superposing three partial

solutions. The first partial solution is the far-field total stress distribution caused by the

circular opening excavation with finite liner installation (see Fig. 2(a)). The second partial

solution is the pore pressure distribution within the surrounding rock mass below the water

5

Page 7: Analytical solution for stress distribution arou nd deep

table (see Fig. 2(b)). The third partial problem (Fig. 2(c)) is the pore pressure variation

due to liner installation since the permeabilities are different. The superposition principle

applies in this case because the problem is linear both in terms physical responses related

to the solid and fluid phases.

Figure 2: Principle of superposition in solving the poro-mechanical problem.

The solution of any problem in elasticity must satisfy force equilibrium, compatibility of

deformations, and boundary conditions. In particular, to obtain the total stress distribution

induced by tunnel excavation, we use Airy stress function Φ, which fulfils compatibility

equation:

∇2(∇2Φ) = 0 (2)

where ∇2 is the Laplace operator.

The total stresses in the 2D Cartesian coordinates are expressed as:

σx =∂2Φ

∂y2;σy =

∂2Φ

∂x2; τxy = − ∂2Φ

∂x∂y(3)

6

Page 8: Analytical solution for stress distribution arou nd deep

For the steady-state fluid flow, the total head distribution follows Laplace equation:

∇2H = 0 (4)

The pore pressure, a function of total head, is calculated as:

p = γw(H − Z) (5)

where γw is the unit weight of fluid and Z is the elevation head.

It is assumed that the water pressure at inner liner boundary is constant (p), which can

be considered as leakage and/or pumping condition. The total hydraulic head at the the

outer liner boundary is assumed constant (Hr) [1, 39, 40]. Therefore, the pore pressure at

the rock-liner interface is a linear function with respect to the depth py = γw(Hr − y) (as

shown in Fig. 2(c)).

3. Analytical solution

3.1. First partial solution

In this framework, the total stress distribution is solved by using the complex variable

method. The general concepts can be found in [15]. As shown in Fig. 2(a), the liner is

assumed to adhere perfectly to the surrounding rock and its installation is assumed instan-

taneous after the tunnel excavation. For the case of support delay condition (add another

displacement component in the deformation continuity equation), we can use the recent pub-

lication derived by Lu et al. [28] for non-circular openings with thin liner only. However, for

modern excavation processes such as shielded TBM tunnels, it is reasonable to assume that

support systems are installed instantaneously. Therefore, the deformation field continuity

7

Page 9: Analytical solution for stress distribution arou nd deep

at the rock-liner interface can be written as:

Γ[k1ψ1(z)− zψ′1(z)− χ1(z)] = k2ψ2(z)− zψ′2(z)− χ2(z), z = r1eiθ (6)

where the superscripts ′ represents the first order of the derivatives; Γ = G2/G1 and G =

E/(2 + 2ν). To make it clear, the subscripts 1 and 2 represent the rock mass and the liner,

respectively.

On the inner boundary of the liner, the stress relations are expressed as:

σr − iτrθ = 0, z = r2eiθ (7)

The continuity of resultant force at the rock-liner interface results in:

χ1(z) + ψ1(z) + zψ′1(z) = χ2(z) + ψ2(z) + zψ′2(z), z = r1eiθ (8)

For the addition of far-field stress boundary conditions, we consider two different cases,

which are the conditions with and without gravity. 1) in Case one (with gravity), the far-field

stress condition is expressed as:

σv→∞ = σffv + γr(Y − v)

σu→∞ = λ[σffv + γr(Y − v)]

τuv→∞ = 0

(9)

where σffv is the far-field vertical stress, γr is the unit weight of the rock mass (we set the

densities between the rock mass and the liner equal), Y is the relative distance from the

tunnel centre to the upper boundary (as shown in Fig. 2(a)) and λ = ν/(1− ν) represents

the lateral stress coefficient.

8

Page 10: Analytical solution for stress distribution arou nd deep

2) in Case two (without considering gravity), the far-field total stress can be expressed

as: σv→∞ = σffv , σu→∞ = λσffv and τuv→∞ = 0.

The solution procedure is similar to the above mentioned far-field total stress boundary

conditions. We now derive the expressions based on the first case boundary condition. The

stress boundary condition of Eq.(9) for case one is expressed as:

(1 + λ)[σffv + γr(Y − v)] = 4Re[ψ′1(z)], |z| → ∞

(1− λ)[σffv + γr(Y − v)]ei2θ = 2ei2θ[zψ′′1(z) + χ′1(z)] |z| → ∞(10)

We now expand the complex stress functions ψ(z) and χ(z) into Laurent series as [45]:

ψ1(z) =∑∞

k=0[a2kz2k+1 + b2kz

−(2k+1)]

χ1(z) =∑∞

k=0[c2kz2k+1 + d2kz

−(2k+1)]

ψ2(z) =∑∞

k=0[e2kz2k+1 + f2kz

−(2k+1)]

χ2(z) =∑∞

k=0[g2kz2k+1 + h2kz

−(2k+1)]

(11)

where the unknown coefficients a2k, b2k, c2k, d2k, e2k, f2k, g2k and h2k are expressed in

the complex space. Knowing that the opening geometry, liner and far-field stresses are

symmetric to the coordinates, these coefficients are real constants which can be calculated

based on the above mentioned stress and displacement boundary conditions.

By combining Eq.(6) ∼ Eq.(11), and comparing the coefficients of eiθ, ei(2k+1)θ and

e−i(2k+1)θ in each derived equation (see the detailed derivation in Appendix), the following

9

Page 11: Analytical solution for stress distribution arou nd deep

set of equations is obtained:

a0 = 14(1 + λ)[σffv + γr(Y − v)], a2k = 0, k > 1

c0 = 12(1− λ)[σffv + γr(Y − v)], c2k = 0, k > 1

2e0 + h0r−22 = 0, 1©

2a0r1 + d0r−11 − 2e0r1 − h0r−11 = 0, 2©

(k1Γ− Γ + 2)a0r1 − (Γ− 1)d0r−11 − (k2 + 1)e0r1 = 0, 3©∑∞

k=1[2(2k + 1)e2kr2k2 − 2(2k − 1)f2k−2r

−2k2 − 2k(2k + 1)e2kr

2k2

−2k(2k − 1)f2k−2r−2k2 − (2k − 1)g2k−2r

2k−22 + (2k + 1)h2kr

−2k−22 ] = 0, 4©∑∞

k=1[−2k(2k + 1)e2kr2k2 + 2k(2k − 1)f2k−2r

−2k2

−(2k − 1)g2k−2r2k−22 − (2k + 1)h2kr

−2k−22 ] = 0, 5©∑∞

k=1[a2kr2k+11 − (2k − 1)b2k−2r

−2k+11 + d2kr

−2k−11

−e2kr2k+11 + (2k − 1)f2k−2r

−2k+11 − h2kr−2k−11 ] = 0, 6©∑∞

k=1[(k1Γ + 1)a2kr2k+11 + (Γ− 1)(2k − 1)b2k−2r

−2k+11

−(Γ− 1)d2kr−2k−11 − (k2 + 1)e2kr

2k+11 ] = 0, 7©∑∞

k=0[b2kr−2k−11 + (2k + 3)a2k+2r

2k+31 + c2kr

2k+11

−f2kr−2k−11 − (2k + 3)e2k+2r2k+31 − g2kr2k+1

1 ] = 0, 8©∑∞k=0[(k1Γ + 1)b2kr

−2k−11 − (Γ− 1)(2k + 3)a2k+2r

2k+31

−(Γ− 1)c2kr2k+11 − (k2 + 1)f2kr

−2k−11 ] = 0. 9©

(12)

The final equation set contains 3 + 6k unknowns and 3 + 6k independent equations, which

are solved by following matrix:

MX = N (13)

where M is a n × n matrix of coefficients ak, bk, ek, fk, gk and hk; X is a n × 1 vector of

ak, bk, ek, fk, gk and hk; and N contains n constants. Taking these coefficients into the

potential function Eq.(11), the total stresses distribution in both finite thickness liner and

10

Page 12: Analytical solution for stress distribution arou nd deep

surrounding rock mass are therefore obtained analytically.

3.2. Second partial solution

The second partial problem involves water inflow into the lined tunnel. As shown in Fig.

3, the conformal mapping technique [32–36] is applied to deal with this semi-infinite complex

boundary condition. We map the outer liner boundary, water table and the surrounding

rock medium onto a circular ring. The mapping function can be written as:

F = ω(ξ) = −ih1− α2

1 + α2

1 + ξ

1− ξ(14)

where α is defined by the ratio of opening radius (r1) and the depth h.

α =h

r1−√

(h/r1)2 − 1 (15)

The coordinate transformation between the F -plane (F = x + iy) and ξ-plane (ξ =

ρ(x,y)eiθ(x,y)) in Polar coordinates is shown as:

ρ =

√x2 + (y +

√h2 − r21)2

x2 + (y −√h2 − r21)2

θ = arctan(2x

√h2 − r21

x2 + y2 − (h2 − r21))

(16)

In the original F -plane, the x− axis is set at the water table level and the y− axis aligns

with the tunnel centre. After mapping, any point (x, y) below the water table in F -plane can

be conformally mapped into ξ domain with (ρ, θ). The inner circumference in ξ-plane with

radiusα represents the original F -plane with radius r1; the radius of the outer circumference

in ξ-plane is equal to one, which represents the water table level of the original plane. The

mapping result of this mapped function is illustrated in Fig. 3.

The total head expression (Eq.(4)) within the original plane is expressed in the ξ-plane

11

Page 13: Analytical solution for stress distribution arou nd deep

Figure 3: Mapped opening for half-infinite plane.

as:

∂2H

∂ρ2+

1

ρ

∂H

∂ρ= 0 (17)

The general solution to Eq.(17) is:

H = C1 + C2 ln ρ+∞∑n=1

(C3ρn + C4ρ

−n) cosnθ (18)

where ρ and θ represent any point in the ξ domain; C1, C2, C3 and C4 are determined by the

total head boundary conditions at the water table level H(ρ=1) = 0 and rock-liner interface

H(ρ=α) = Hr (constant total head assumption in Section 2.2).

Substituting these boundary conditions into Eq.(18), the total head expression in ξ do-

main can be obtained as:

H =Hr

lnαln ρ (19)

Therefore, the water pressure within the rock mass is denoted as (expressed in original

F region):

pr(x, y) =γwHr

ln(h/r1 −√

(h/r1)2 − 1)ln

√x2 + (y +

√h2 − r21)2

x2 + (y −√h2 − r21)2

− γwy (20)

Eq.(20) shows that the water pressure distribution is a function of the hydraulic head

12

Page 14: Analytical solution for stress distribution arou nd deep

Hr, which is an unknown parameter to be determined. Instead of assuming a value Hr, we

use the equality relationship of water inflow in the rock mass and in the liner to precisely

determine this value.

The water inflow through the rock-liner interface can be obtained as:

Qr = Kr

∫ 2π

0

∂H

∂ρρdθ =

2πKrHr

ln(h/r1 −√

(h/r1)2 − 1)(21)

where Kr is the permeability of the rock mass.

3.3. Third partial solution

The third partial problem (see Fig. 2(c)) is proposed to express the total head at the

outer liner boundary (Hr = py/γw + y). It is firstly solved without taking specific weight

γw and the elevation head y into consideration. The pore pressure distribution within liner

fulfils Laplace equation and the conditions of the conformal mapping of the complex variable

methods. The general solution of pore pressure distribution is written as.

pl = C5 + C6 ln ρ (22)

where C5 and C6 are constants which are determined by the boundary conditions.

When the gravity is neglected, the pore pressure boundary condition at the rock-liner

interface is expressed as p1 = γwHr and the pore pressure at the inner liner p2 = p is constant.

The pore pressure expression is obtained by substituting these boundary conditions into

Eq.(22):

pl = γwHr +p− γwHr

ln(r2/r1)lnρ

r1(23)

The total hydraulic head within the liner is expressed as:

Hl = Hr +p/γw −Hr

ln(r2/r1)ln

√x2 + (y + h)2

r1(24)

13

Page 15: Analytical solution for stress distribution arou nd deep

The above solutions are derived under the condition of negligible gravity. By taking

the gravity and elevation head into consideration, the total hydraulic head expression is

described as follows [35]:

Hl = Hr + A(x, y) +p/γw −Hr +B(x, y)

ln(r2/r1)ln

√x2 + (y + h)2

r1(25)

where A(x, y) and B(x, y) are functions to be determined.

The corresponding boundary conditions are:

Hl = Hr, (x2 + (y + h)2 = r21)

Hl = p/γw + r2 sin θ − h, (x2 + (y + h)2 = r22)

(26)

where θ = arctan y+hx

Substituting Eq.(26) into Eq.(25), the unknowns are solved as A(x, y) = 0 and B(x, y) =

r2 sin θ − h.

The total hydraulic head Hl can be rewritten in Polar coordinates as:

Hl = Hr +p/γw −Hr + r2 sin θ − h

ln(r2/r1)lnρ

r1(27)

The water inflow Ql in the liner is calculated as:

Ql = Kl

∫ 2π

0

∂Hl

∂ρρdθ =

2πKl

ln(r2/r1)(p/γw −Hr − h) (28)

where Kl represents the permeability of the liner.

Combining Eq.(21) and Eq.(28) and setting Qr = Ql, the unknown hydraulic head Hr is

derived as:

Hr =k(p/γw − h) lnα

ln(r2/r1) + k lnα(29)

14

Page 16: Analytical solution for stress distribution arou nd deep

where k = Kl/Kr is the relative permeability between the liner and rock mass.

Substituting Eq.(29) into Eq.(20), the pore pressure distribution within the rock mass is

denoted as:

pr(x, y) =k(p− γwh)

ln(r2/r1) + k ln(h/r1 −√

(h/r1)2 − 1)ln

√x2 + (y +

√h2 − r21)2

x2 + (y −√h2 − r21)2

− γwy (30)

4. Validation of the proposed solution

4.1. Numerical simulation

The general purpose software Abaqus that uses the Finite Element Method (FEM) is used

to validate the analytical solution. The model is developed to simulate a deep buried lined

tunnel. The rock mass and lined material are modelled as solid elements. In this simulation,

as shown in Fig. 1, the domain is discretised with 8-node plane strain quadrilateral element

(CPE8RP). Local fine mesh is applied to capture the displacement/stress component and

to save computational time. We extend the analysed domain to 30 times the opening radius

(100 m from the tunnel centre) to reduce the error caused by the far-field boundary condition

as suggested by Bobet [39].

Table 1 summarises the input mechanical properties. To simulate the far-field condition,

10.0 MPa vertical stress is applied on the top boundary. The horizontal stress is obtained

based on the Poisson’s ratio and the gravity gradient automatically. The vertical and bottom

boundaries are restricted to normal movement. The pore pressure of the inner liner boundary

is 0.2 MPa. The water table is set to 50 m above the tunnel centre and remains unchanged.

The pore pressure at and above the water table is set to zero; it increases linearly with

depth. For this case, the liner thickness is set to 0.5 m, while the solution is not restricted in

terms of liner thickness. The permeability of the liner is selected 100 times smaller than that

of the rock mass, while the stiffness of the liner is relatively higher. For the case without

gravity , the top and bottom boundaries are both equal to 10.0 MPa compressive stresses.

15

Page 17: Analytical solution for stress distribution arou nd deep

Table 1: Mechanical parameters.

Variable Parameter Variable Parameter

σffv (MPa) 10 p (MPa) 0.2Y (m) 100 h (m) 50r1 (m) 3.0 r2 (m) 2.5Kr (m/s) 5× 10−6 Kl (m/s) 5× 10−8

Er (GPa) 20 El (GPa) 200γr (N/m3) 25000 γw (N/m3) 10000νr 0.25 νl 0.25

4.2. Comparison of analytical solutions with numerical results

Fig. 4 shows the pore pressure distribution contour in a half-domain since its distribution

is symmetric to y−axis. It is obvious that the proposed analytical solution matches well with

the FEM modelling, especially in the vicinity of the lined opening. In the area further away

from the opening, the pore pressure obtained using FEM modelling is slightly higher than

that obtained using analytically, because of the mesh effect and the pore pressure constraints

along the edges of the boundary. For the fine mesh and the points further away from the

boundary constraints, the results match well.

Apart from comparing it with the numerical simulation, we compared the derived first

partial solution with Lu et al. [46] and Wand and Li [47]. In this comparison, the following pa-

rameters are taken: r1 = 3.0 m, r2 = 2.5 m, νr = νl = 0.25, El/Er = 10.0, σffv = 10.0 MPa

and σffu = 5.0 MPa. Fig. 5 shows the variation of stress concentration factor (SCF), σ/σffv ,

with the ratio distance (r/r2) in the horizontal direction. Clearly, the numerical result is

close to the analytical solution, with finer mesh the stress value would reach the analytic

result especially at the rock-liner interface. The analytical solution derived by Lu et al. [46]

considered slip and no-slip conditions on the interface. The inner pressure of the tunnel was

considering in both Wand and Li [47] and Lu et al. [46] solutions, but they overlooked the

gravity effect. For comparison purpose of the first partial solution with our solution, the

same results were obtained, since all these solutions are derived based on the same theory.

16

Page 18: Analytical solution for stress distribution arou nd deep

Figure 4: Comparison with FEM modelling in terms of pore pressure distribution.

Figure 5: Comparison of the first partial total stress solution.

17

Page 19: Analytical solution for stress distribution arou nd deep

The derived pore pressure distribution is also compared with the published results of

Ming et al. [35] where lining was ignored. In our solution, the pore pressure distribution

obtained in the absence of lining using equation Eq.(30) can be expressed as:

pr(x, y) =γwHr

ln(h/r1 −√

(h/r1)2 − 1)ln

√x2 + (y +

√h2 − r21)2

x2 + (y −√h2 − r21)2

− γwy (31)

where Hr = p/γw−h, which also matches the constant total head assumption. This equation

is comparable with the solution of Ming et al. [35]. Their solution can be considered as a

specific case of our solution, in which the liner is not considered.

Taking liners effect into consideration, we further compare the pore pressure with Li et

al. [1] solution, in which the mirror image method is used. Fig. 6 shows the comparison

result in horizontal direction along the tunnel axis. The relative permeability between the

rock and liner is set to 10 (Kr/Kl = 10.0) to clearly show the variation of pore pressure with

distance. The rest calculation parameters are shown in Table 1. Obviously, the proposed

solution by using complex variable method coincides with the result derived by Li et al. [1].

We also compare our analytical solution to the numerical simulation in terms of effective

principal stresses with respect to distance from the opening boundaries (see Fig. 7 and

Fig. 8). For the total stress partial solution, we consider both conditions with and without

gravity gradient and compare the results to the corresponding numerical simulations. Both

the effective maximum and minimum principal stresses obtained analytically (see the solid

line) in the three representative directions coincide with the corresponding solutions obtained

using FEM simulation (see the dot-dash line). It is suggested that our solutions are valid to

describe the lining effect of the stress distribution within the rock mass.

The effective maximum principal stresses (MPS) on the tunnel boundary in the vertical

direction (point 2 and 3) show a similar results irrespective of the gravity conditions (Fig.

7 (a) and (b)). However, these values vary a lot when the distance increases from the

18

Page 20: Analytical solution for stress distribution arou nd deep

Figure 6: Comparison of pore pressure.

boundary. In the horizontal direction, the MPS values obtained with gravity are larger than

those obtained without it. The magnitude mainly depends on the distance between the

model boundary and the tunnel centre (Y value). When gravity is neglected, the effective

minimum principal stresses increase sharply and then decreases with the distance away

from the tunnel boundary in the vertical directions (Fig. 8 (a) and (b)). When gravity is

considered, this value decreases with respect to distance away from the tunnel boundary.

Considering gravity has a great influence on the stress distribution and should be taken into

account especially when designing deep pressure tunnels.

5. Discussion

5.1. Lining effect on the stress distribution

As shown in Fig. 9, the installed liner highly affects the pore pressure distribution. We

plot the pore pressure expression Eq.(31) obtained without lining first. The pore pressure

drops rapidly to the value near the inner boundary condition at the vicinity of the tunnel,

19

Page 21: Analytical solution for stress distribution arou nd deep

Figure 7: Comparisons of effective maximum principal stress (MPS) between the proposed analytical solutionand numerical simulation in three representative directions; (a) above the tunnel crown, (b) below the tunnelinvert and (c) at horizontal line.

20

Page 22: Analytical solution for stress distribution arou nd deep

Figure 8: Comparisons of effective minimum principal stress between the proposed analytical solution andnumerical simulation in three representative directions; (a) above the tunnel crown, (b) below the tunnelinvert and (c) at horizontal line.

21

Page 23: Analytical solution for stress distribution arou nd deep

regardless of the total head level. While with the liner (the related parameter is shown in

Table 1), the pore pressure near the tunnel drops slightly.

Figure 9: Pore pressure distribution contour with and without liner conditions.

Installing the liner also alters the total stress distribution which changes the effective

stress within the rock mass. Fig. 10 compares the effective MPS variation with and without

liner installation. When there is no liner (dot-dash line in Fig. 10), the MPS concentrates

at the horizontal direction and decreases gradually further away from the opening. On the

tunnel boundary, the stress increases from the tunnel crown to the horizontal direction.

When the liner is installed (solid line in Fig. 10), the MPS appears on the tunnel boundary

with an angle around -40◦ from the x−axis. It is noticed that the effective MPS reduces

by about 50% as it decreases from 33.1 MPa to 17.3 MPa with the liner effect. It is

worthwhile mentioning that the standard total MPS (calculated without considering the

22

Page 24: Analytical solution for stress distribution arou nd deep

effects of gravity and liner) for the same set of parameters is 25.0 MPa in accordance with

the textbook of Jaeger et al. ([48]). As for the minimum principal effective stress shown

in Fig. 11, it increases from the tunnel corner to a high value away from the opening both

with and without lining conditions. On the tunnel boundary, the stress starts to decrease

slightly at the tunnel crown until it reaches a higher level at a point around 40◦ from the

x−axis; it then increases to its peak value in a horizontal direction for the given lined tunnel

case. The same tendency can be found under no liner condition. It is also noticed that the

effective stress contours are not symmetric with respect to x due to gravity.

Figure 10: Effective MPS contour with and without liner conditions.

For given geological conditions, the properties of the liner are selected to resist stress

concentrations; these properties include essentially the thickness, stiffness and permeability.

Herein, We are not intended to give the best possible design, but rather to cover a range of

typical dimensions. To provide acceptable design, only one parameter should be varied at

a time while all other parameters are kept constant. From the derived equations in Section

3, the relative liner thickness would affect both pore pressure distribution and total stress

status around the opening. The pore pressure distribution is also highly related to the

23

Page 25: Analytical solution for stress distribution arou nd deep

Figure 11: Effective minimum principal stress contour with and without liner conditions.

relative permeability of the liner. The relative stiffness of liner, in turn, affects the total

stress distribution. Although the Poisson’s ratio has an influence on the stress, it does not

vary too much for different liner materials. Therefore, we keep its value equal to 0.25.

5.2. Effect of the relative thickness

Fig. 12 shows the variation of effective MPS with different relative liner thicknesses in

the horizontal direction at line 5-6. The two extreme cases with ∆r/r1 = 1 and ∆r/r1 =

0 represent fully filled and no liner installed conditions, respectively. The effective MPS

increases with the relative liner thickness on the boundary. When the relative liner thickness

is larger than 0.143, the effective MPS decreases first and then increases gradually to a stable

value.

In addition, the effect of the relative thickness on the effective MPS within the domain

and its location around the tunnel boundary are studied in detail (see Fig. 13). The effective

MPS reaches its highest value of 33.1 MPa when there is no liner. With the liner, the value

drops sharply and remains around 17 MPa with different liner thicknesses. However, the

concentration location varies a lot. It moves from the tunnel crown to the tunnel invert as

24

Page 26: Analytical solution for stress distribution arou nd deep

Figure 12: Effective MPS variation at line 5-6 on the effect of relative thickness.

the relative liner thickness increases.

5.3. Effect of the relative stiffness

The total stresses shown in Eq.(12) are determined using the relative Young’s modulus

rather than the Young moduli of the rock or the liner alone. The analysed cases cover El/Er

ratios that change over two orders of magnitude, which are thought to be representative in

design cases. As shown in Fig. 14, if the liner stiffness is smaller than the rock mass stiffness,

the stress increases at the excavation boundary. A relative stiffness equal to 0 is considered

when no liner is installed. Under such condition, the total MPS reaches 33.3 MPa on the

tunnel boundary. When the relative stiffness value is less than 5, the stress within the

rock mass decreases with the increase of the liner relative stiffness. The total MPS decays

gradually with the distance. However, when the relative stiffness value is larger than 10, the

total MPS decreases first and then increases to a stable value. The stress variation zone is

in the range of 3 times the initial opening radius (within 15 m to the opening boundary).

25

Page 27: Analytical solution for stress distribution arou nd deep

Figure 13: Effective MPS within the domain and its location on the effect of relative thickness.

In all cases, the total MPS convergence to 12.5 MPa in the horizontal direction.

Fig. 15 also shows that the effective MPS drops when the liner is installed. When the

liner stiffness is equal to or larger than that of the rock mass (relative stiffness is larger

than 1), the effective MPS value reduces to about 20 MPa. The concentration location of

the effective MPS converges to about -40◦ from x-axis with the increase of the liner relative

stiffness.

5.4. Effect of the relative permeability

The pore pressure distribution is also related to the liner relative permeability. In

petroleum engineering, the skin effect is formed since the permeabilities between the well-

bore and the formation are different. The skin is positive if the permeability in the skin

zone is less than that of the formation (Kl/Kr < 1). In contrast, the skin is negative if

it is more than that of the formation (Kl/Kr > 1). If the two permeabilities are equal

(Kl/Kr = 1) there is no skin effect [49]. Fig. 16 indicates that the pore pressure increases

26

Page 28: Analytical solution for stress distribution arou nd deep

Figure 14: Total MPS variation at line 5-6 on the effect of relative stiffness.

Figure 15: Effective MPS within the domain and its location on the effect of relative stiffness.

27

Page 29: Analytical solution for stress distribution arou nd deep

with the decreases of the liner relative permeability. The case Kl/Kr = 0 with the solid red

line in the figure means the impermeable liner condition, while Kl/Kr = 1.0e5 represents

no liner condition, in which the pore pressure at the inner boundary equals the initial pore

pressure as 0.2 MPa which also confirms of no liner condition. Furthermore, the effective

MPS value increases with the liner relative permeability (see in Fig. 17). Changing the liner

permeability affects the pore pressure distribution only; the concentration location of the

effective MPS remains at -37.1◦ on the tunnel boundary.

Figure 16: Pore pressure variation at line 5-6 on the effect of relative permeability.

6. Conclusion

An analytical solution is obtained in terms of stress distribution in deep lined tunnels.

The problem is solved by using the superposition of three partial solutions. The complex

variable method and the conformal mapping technique are applied to solve this half-infinite

hydraulic domain and far-field geo-stress conditions. The proposed solution is compared

28

Page 30: Analytical solution for stress distribution arou nd deep

Figure 17: Effective MPS within the domain and its location on the effect of relative permeability.

with the results of the Finite Element Method obtained in terms of pore pressure and

effective principal stresses. The lining effect on the effective stress distribution is studied

in detail. Furthermore, a range of liner relative thicknesses, relative stiffnesses and relative

permeabilities are performed to investigate the lining effect on the stress variation. The

main findings of this study are summarised as follows:

In deriving the total stress, we considered both with and without gravity for far-field geo-

stress conditions. Both cases were compared to the numerical modelling and they were found

to match well. For the final superposition solutions, both pore pressure distribution and

effective principal stresses obtained analytically coincide with the FEM modelling results.

In addition, the proposed solution can be generalised to any thickness of the liner.

The liner has great influence on the stress distribution after tunnel excavation. When

there is no liner, the pore pressure drops significantly to a low value. While when the liner is

installed (for the given liner property), the pore pressure changes slightly around the tunnel.

29

Page 31: Analytical solution for stress distribution arou nd deep

In addition, under the lined condition, the effective MPS reduces from 33.1 MPa to 17.3

MPa within the domain.

Parametrization results show that the effective MPS value is the coupling effect of the

liner relative thickness, stiffness and permeability. While its location moves from the hor-

izontal direction to the invert with the increase of the liner thickness and stiffness. The

difference in permeability between the rock and liner acts as a skin effect on the fluid flow.

The lower the liner permeability the higher water pressure would be, which reduces effective

stress in the rock mass. In practice, the selection of liner should also take the cost and the

real geological conditions into consideration.

Acknowledgements

The first author would like to acknowledge the financial support provided by the China

Scholarship Council (201606420056).

Appendix

By substituting Eq.(8) into Eq.(6) and combining Eq.(11), the following equations can

be obtained:

(k1Γ + 1)a0r1eiθ + (k1Γ + 1)

∞∑k=1

a2kr2k+11 ei(2k+1)θ + (k1Γ + 1)

∞∑k=0

b2kr−2k−11 e−i(2k+1)θ

− (Γ− 1)a0r1eiθ − (Γ− 1)

∞∑k=0

(2k + 3)a2k+2r2k+31 e−i(2k+1)θ + (Γ− 1)

∞∑k=1

(2k − 1)b2k−2r−2k+11 ei(2k+1)θ

− (Γ− 1)∞∑k=0

c2kr2k+11 e−i(2k+1)θ − (Γ− 1)d0r

−11 eiθ − (Γ− 1)

∞∑k=1

d2kr−2k−11 ei(2k+1)θ

− (k2 + 1)e0r1eiθ − (k2 + 1)

∞∑k=1

e2kr2k+11 ei(2k+1)θ − (k2 + 1)

∞∑k=0

f2kr−2k−11 e−i(2k+1)θ = 0

(A.1)

30

Page 32: Analytical solution for stress distribution arou nd deep

Comparing the coefficients of eiθ, ei(2k+1)θ and e−i(2k+1)θ in Eq.(A.1) and noting that the

equality is independent of θ, we have:

(k1Γ− Γ + 2)a0r1 − (Γ− 1)d0r−11 − (k2 + 1)e0r1 = 0,∑∞

k=1[(k1Γ + 1)a2kr2k+11 + (Γ− 1)(2k − 1)b2k−2r

−2k+11

−(Γ− 1)d2kr−2k−11 − (k2 + 1)e2kr

2k+11 ] = 0,∑∞

k=0[(k1Γ + 1)b2kr−2k−11 − (Γ− 1)(2k + 3)a2k+2r

2k+31

−(Γ− 1)c2kr2k+11 − (k2 + 1)f2kr

−2k−11 ] = 0

(A.2)

Substituting Eq.(11) into Eq.(10), we have:

σffv + σffu =4[∞∑k=0

(2k + 1)a2kr2k3 e

i2kθ −∞∑k=1

(2k − 1)b2k−2r−2k3 e−i2kθ];

(σffv − σffu )ei2θ =2ei2θ[∞∑k=0

2k(2k + 1)a2kr2k3 e

i(2k−2)θ +∞∑k=1

2k(2k − 1)b2k−2r−2k3 e−i(2k+2)θ

+∞∑k=0

(2k + 1)c2kr2k3 e

i2kθ −∞∑k=1

(2k − 1)d2k−2r−2k3 e−i2kθ]

(A.3)

Since r3 →∞ in Eq.(A.3), we have:

a0 = 14(1 + λ)[σffv + γr(Y − v)], a2k = 0, k > 1

c0 = 12(1− λ)[σffv + γr(Y − v)], c2k = 0, k > 1

(A.4)

Substituting Eq.(11) and Eq.(7) into Airy stress function, the following equations can be

obtained:

31

Page 33: Analytical solution for stress distribution arou nd deep

2e0 + h0r−22 +

∞∑k=1

[2(2k + 1)e2kr2k2 − 2(2k − 1)f2k−2r

−2k2 − 2k(2k + 1)e2kr

2k2

− 2k(2k − 1)f2k−2r−2k2 − (2k − 1)g2k−2r

2k−22 + (2k + 1)h2kr

−2k−22 ] cos 2θ

+∞∑k=1

[−2k(2k + 1)e2kr2k2 + 2k(2k − 1)f2k−2r

−2k2 − (2k − 1)g2k−2r

2k−22

− (2k + 1)h2kr−2k−22 ] sin 2θ = 0

(A.5)

Knowing that Eq.(A.5) is independent of θ shows that:

2e0 + h0r−22 = 0,∑∞

k=1[2(2k + 1)e2kr2k2 − 2(2k − 1)f2k−2r

−2k2 − 2k(2k + 1)e2kr

2k2

−2k(2k − 1)f2k−2r−2k2 − (2k − 1)g2k−2r

2k−22 + (2k + 1)h2kr

−2k−22 ] = 0,∑∞

k=1[−2k(2k + 1)e2kr2k2 + 2k(2k − 1)f2k−2r

−2k2

−(2k − 1)g2k−2r2k−22 − (2k + 1)h2kr

−2k−22 ] = 0

(A.6)

Substitution of Eq.(11) into Eq.(8) leads to:

a0r1eiθ +

∞∑k=1

a2kr2k+11 ei(2k+1)θ +

∞∑k=0

b2kr−2k−11 e−i(2k+1)θ + a0r1e

+∞∑k=0

(2k + 3)a2k+2r2k+31 e−i(2k+1)θ −

∞∑k=1

(2k − 1)b2k−2r−2k+11 ei(2k+1)θ +

∞∑k=0

c2kr2k+11 e−i(2k+1)θ

+ d0r−11 eiθ +

∞∑k=1

d2kr−2k−11 ei(2k+1)θ = e0r1e

iθ + h0r−11 eiθ + e0r1e

iθ +∞∑k=1

e2kr2k+11 ei(2k+1)θ

+∞∑k=0

f2kr−2k−11 e−i(2k+1)θ +

∞∑k=0

(2k + 3)e2k+2r2k+31 e−i(2k+1)θ −

∞∑k=1

(2k − 1)f2k−2r−2k+11 ei(2k+1)θ

+∞∑k=0

g2kr2k+11 e−i(2k+1)θ +

∞∑k=1

h2kr−2k−11 ei(2k+1)θ

(A.7)

32

Page 34: Analytical solution for stress distribution arou nd deep

Comparing the coefficients of eiθ, ei(2k+1)θ and e−i(2k+1)θ in Eq.(A.7) and noting that this

equality is independent of θ, we have:

2a0r1 + d0r−11 − 2e0r1 − h0r−11 = 0,∑∞

k=1[a2kr2k+11 − (2k − 1)b2k−2r

−2k+11 + d2kr

−2k−11

−e2kr2k+11 + (2k − 1)f2k−2r

−2k+11 − h2kr−2k−11 ] = 0,∑∞

k=0[b2kr−2k−11 + (2k + 3)a2k+2r

2k+31 + c2kr

2k+11

−f2kr−2k−11 − (2k + 3)e2k+2r2k+31 − g2kr2k+1

1 ] = 0

(A.8)

Combining Eq.(A.4), Eq.(A.6), Eq.(A.8) and Eq.(A.2), we have the following equation

33

Page 35: Analytical solution for stress distribution arou nd deep

set:

a0 = 14(1 + λ)[σffv + γr(Y − v)], k > 1

c0 = 12(1− λ)[σffv + γr(Y − v)], c2k = 0, k > 1

2e0 + h0r−22 = 0, 1©

2a0r1 + d0r−11 − 2e0r1 − h0r−11 = 0, 2©

(k1Γ− Γ + 2)a0r1 − (Γ− 1)d0r−11 − (k2 + 1)e0r1 = 0, 3©∑∞

k=1[2(2k + 1)e2kr2k2 − 2(2k − 1)f2k−2r

−2k2 − 2k(2k + 1)e2kr

2k2

−2k(2k − 1)f2k−2r−2k2 − (2k − 1)g2k−2r

2k−22 + (2k + 1)h2kr

−2k−22 ] = 0, 4©∑∞

k=1[−2k(2k + 1)e2kr2k2 + 2k(2k − 1)f2k−2r

−2k2

−(2k − 1)g2k−2r2k−22 − (2k + 1)h2kr

−2k−22 ] = 0, 5©∑∞

k=1[a2kr2k+11 − (2k − 1)b2k−2r

−2k+11 + d2kr

−2k−11

−e2kr2k+11 + (2k − 1)f2k−2r

−2k+11 − h2kr−2k−11 ] = 0, 6©∑∞

k=1[(k1Γ + 1)a2kr2k+11 + (Γ− 1)(2k − 1)b2k−2r

−2k+11

−(Γ− 1)d2kr−2k−11 − (k2 + 1)e2kr

2k+11 ] = 0, 7©∑∞

k=0[b2kr−2k−11 + (2k + 3)a2k+2r

2k+31 + c2kr

2k+11

−f2kr−2k−11 − (2k + 3)e2k+2r2k+31 − g2kr2k+1

1 ] = 0, 8©∑∞k=0[(k1Γ + 1)b2kr

−2k−11 − (Γ− 1)(2k + 3)a2k+2r

2k+31

−(Γ− 1)c2kr2k+11 − (k2 + 1)f2kr

−2k−11 ] = 0. 9©

(A.9)

It contains 3 + 6k unknowns and 3 + 6k independent equations, which can be solved linearly

by using Python code as follows:

MX = N (A.10)

where M is a n×n matrix of coefficients ak, bk, ek, fk, gk and hk; X is a n×1 vector of ak, bk,

ek, fk, gk and hk; and N contains n constants. We combine k > 1 in Eq.(A.9) 4©- 7© and k > 0

in 8© and 9©, which consist of 6n equations equal to zero without any constant. This indicates

34

Page 36: Analytical solution for stress distribution arou nd deep

that the unknown coefficients of these equations are zero (e2k = h2k = d2k = 0, k > 1;

f2k = g2k = b2k = 0, k > 0). Therefore, the 9 unknowns with remaining 9 independent

equations are solved.

2e0 + h0r−22 = 0,

2a0r1 + d0r−11 − 2e0r1 − h0r−11 = 0,

(k1Γ− Γ + 2)a0r1 − (Γ− 1)d0r−11 − (k2 + 1)e0r1 = 0,

−4f0r−22 − g0 + 3h2r

−42 = 0,

−6e2r22 + 2f0r

−22 − g0 − 3h2r

−42 = 0,

−b0r−11 + d2r−31 − e2r31 + f0r

−11 − h2r−31 = 0,

(Γ− 1)b0r−11 − (Γ− 1)d2r

−31 − (k2 + 1)e2r

31 = 0,

b0r−11 + c0r1 − f0r−11 − 3e2r

31 − g0r1 = 0,

(k1Γ + 1)b0r−11 − (Γ− 1)c0r1 − (k2 + 1)f0r

−11 = 0.

(A.11)

where for case 1) with gravity a0 = (1+λ)[σffv +γr(Y −v)]/4; c0 = (1−λ)[σffv +γr(Y −v)]/2

and for case 2) without gravity a0 = (1 + λ)σffv /4; c0 = (1− λ)σffv /2.

References

[1] P. Li, F. Wang, Y. Long, X. Zhao, Investigation of steady water inflow into a subsea grouted tunnel,

Tunn. Undergr. Space Technol. 80 (2018) 92–102.

[2] P. Li, H. Liu, Y. Zhao, Z. Li, A bottom-to-up drainage and water pressure reduction system for railway

tunnels, Tunn. Undergr. Space Technol. 81 (2018) 296–305.

[3] H. Yu, J. Chen, A. Bobet, Y. Yuan, Damage observation and assessment of the longxi tunnel during

the wenchuan earthquake, Tunn. Undergr. Space Technol. 54 (2016) 102–116.

[4] X. Dong, A. Karrech, H. Basarir, M. Elchalakani, C. Qi, Analytical solution of energy redistribution in

rectangular openings upon in-situ rock mass alteration, Int. J. Rock Mech. Min. Sci. 106 (2018) 74–83.

[5] A. Karrech, Non-equilibrium thermodynamics for fully coupled thermal hydraulic mechanical chemical

processes, J. Mech. Phys. Solids 61 (3) (2013) 819–837.

35

Page 37: Analytical solution for stress distribution arou nd deep

[6] A. Karrech, C. Schrank, R. Freij-Ayoub, K. Regenauer-Lieb, A multi-scaling approach to predict hy-

draulic damage of poromaterials, Int. J. Mech. Sci. 78 (2014) 1–7.

[7] A. Karrech, O. Beltaief, R. Vincec, T. Poulet, K. Regenauer-Lieb, Coupling of thermal-hydraulic-

mechanical processes for geothermal reservoir modelling, J. Earth Sci. 26 (1) (2015) 47–52.

[8] A. Karrech, T. Poulet, K. Regenauer-Lieb, Poromechanics of saturated media based on the logarithmic

finite strain, Mech. Mater. 51 (2012) 118–136.

[9] A. Karrech, K. Regenauer-Lieb, T. Poulet, A damaged visco-plasticity model for pressure and temper-

ature sensitive geomaterials, Int. J. Eng. Sci. 49 (10) (2011) 1141–1150.

[10] A. Karrech, T. Poulet, K. Regenauer-Lieb, A limit analysis approach to derive a thermodynamic damage

potential for non-linear geomaterials, Philos. Mag. 92 (28-30) (2012) 3439–3450.

[11] A. Karrech, K. Regenauer-Lieb, T. Poulet, Frame indifferent elastoplasticity of frictional materials at

finite strain, Int. J. Solids Struct. 48 (3-4) (2011) 397–407.

[12] A. Karrech, C. Schrank, K. Regenauer-Lieb, A parallel computing tool for large-scale simulation of

massive fluid injection in thermo-poro-mechanical systems, Philos. Mag. 95 (28-30) (2015) 3078–3102.

[13] A. Karrech, F. Abbassi, H. Basarir, M. Attar, Self-consistent fractal damage of natural geo-materials

in finite strain, Mech. Mater. 104 (2017) 107–120.

[14] A. Karrech, A. Seibi, D. Duhamel, Finite element modelling of rate-dependent ratcheting in granular

materials, Comput. Geotech. 38 (2) (2011) 105–112.

[15] N. I. Muskhelishvili, Some basic problems of the mathematical theory of elasticity, Springer Science &

Business Media, 1954.

[16] G. N. Savin, Stress concentration around holes, Pergamon Press, New York, 1961.

[17] T. C. Ekneligoda, R. W. Zimmerman, Shear compliance of two-dimensional pores possessing n-fold axis

of rotational symmetry, Proc. Roy. Soc. Lond. Ser. A 464 (2091) (2008) 759–775.

[18] Z. Zhang, M. Huang, X. Xi, X. Yang, Complex variable solutions for soil and liner deformation due to

tunneling in clays, Int. J. Geomech. 18 (7) (2018) 04018074.

[19] G. Exadaktylos, M. Stavropoulou, A closed-form elastic solution for stresses and displacements around

tunnels, Int. J. Rock Mech. Min. Sci. 39 (7) (2002) 905–916.

[20] H. Huo, A. Bobet, G. Fernandez, J. Ramırez, Analytical solution for deep rectangular structures

subjected to far-field shear stresses, Tunn. Undergr. Space Technol. 21 (6) (2006) 613–625.

[21] Z. Zhang, Y. Sun, Analytical solution for a deep tunnel with arbitrary cross section in a transversely

isotropic rock mass, Int. J. Rock Mech. Min. Sci. 48 (8) (2011) 1359–1363.

36

Page 38: Analytical solution for stress distribution arou nd deep

[22] X. Dong, A. Karrech, H. Basarir, M. Elchalakani, A. Seibi, Energy dissipation and storage in under-

ground mining operations, Rock Mech. Rock Eng. 52 (1) (2019) 229–245.

[23] H. T. Manh, J. Sulem, D. Subrin, A closed-form solution for tunnels with arbitrary cross section

excavated in elastic anisotropic ground, Rock Mech. Rock Eng. 48 (1) (2015) 277–288.

[24] G. Zhao, S. Yang, Analytical solutions for rock stress around square tunnels using complex variable

theory, Int. J. Rock Mech. Min. Sci. 80 (2015) 302–307.

[25] H. H. Einstein, C. W. Schwartz, Simplified analysis for tunnel supports, J. Geotech. Eng. Division.

ASCE 105 (GT4) (1979) 499–518.

[26] A. Bobet, Lined circular tunnels in elastic transversely anisotropic rock at depth, Rock Mech. Rock

Eng. 44 (2) (2011) 149–167.

[27] A. Kargar, R. Rahmannejad, M. Hajabasi, A semi-analytical elastic solution for stress field of lined

non-circular tunnels at great depth using complex variable method, Int. J. Solids Struct. 51 (6) (2014)

1475–1482.

[28] A. Lu, N. Zhang, L. Kuang, Analytic solutions of stress and displacement for a non-circular tunnel at

great depth including support delay, Int. J. Rock Mech. Min. Sci. 70 (2014) 69–81.

[29] A. Kargar, R. Rahmannejad, M. A. Hajabasi, The stress state around lined non-circular hydraulic

tunnels below the water table using complex variable method, Int. J. Rock Mech. Min. Sci. 78 (2015)

207–216.

[30] A. Lu, N. Zhang, S. Wang, X. Zhang, Analytical solution for a lined tunnel with arbitrary cross sections

excavated in orthogonal anisotropic rock mass, Int. J. Geomech. 17 (9) (2017) 04017044.

[31] R. Yang, Q. Wang, L. Yang, A closed-form elastic solution for irregular frozen wall of inclined shaft

considering the interaction with ground, Int. J. Rock Mech. Min. Sci. 100 (2017) 62–72.

[32] M. El Tani, Circular tunnel in a semi-infinite aquifer, Tunn. Undergr. Space Technol. 18 (1) (2003)

49–55.

[33] K. Park, A. Owatsiriwong, J. Lee, Analytical solution for steady-state groundwater inflow into a drained

circular tunnel in a semi-infinite aquifer: a revisit, Tunn. Undergr. Space Technol. 23 (2) (2008) 206–209.

[34] P. Arjnoi, J. Jeong, C. Kim, K. Park, Effect of drainage conditions on porewater pressure distributions

and lining stresses in drained tunnels, Tunn. Undergr. Space Technol. 24 (4) (2009) 376–389.

[35] H. Ming, M. Wang, Z. Tan, X. Wang, Analytical solutions for steady seepage into an underwater

circular tunnel, Tunn. Undergr. Space Technol. 25 (4) (2010) 391–396.

[36] Q. Fang, H. Song, D. Zhang, Complex variable analysis for stress distribution of an underwater tunnel

37

Page 39: Analytical solution for stress distribution arou nd deep

in an elastic half plane, Int. J. Numer. Anal. Methods Geomech. 39 (16) (2015) 1821–1835.

[37] A. N. Hassani, H. Farhadian, H. Katibeh, A comparative study on evaluation of steady-state ground-

water inflow into a circular shallow tunnel, Tunn. Undergr. Space Technol. 73 (2018) 15–25.

[38] D. Zhang, Z. Huang, Z. Yin, L. Ran, H. Huang, Predicting the grouting effect on leakage-induced

tunnels and ground response in saturated soils, Tunn. Undergr. Space Technol. 65 (2017) 76–90.

[39] A. Bobet, Effect of pore water pressure on tunnel support during static and seismic loading, Tunn.

Undergr. Space Technol. 18 (4) (2003) 377–393.

[40] A. Bobet, S. Nam, Stresses around pressure tunnels with semi-permeable liners, Rock Mech. Rock Eng.

40 (3) (2007) 287–315.

[41] A. Bobet, Deep tunnel in transversely anisotropic rock with groundwater flow, Rock Mech. Rock Eng.

49 (12) (2016) 4817–4832.

[42] X. Li, S. Du, B. Chen, Unified analytical solution for deep circular tunnel with consideration of seepage

pressure, grouting and lining, J. Cent. South Univ. 24 (6) (2017) 1483–1493.

[43] N. Tran, D. Do, D. Hoxha, A closed-form hydro-mechanical solution for deep tunnels in elastic

anisotropic rock, Eur. J. Environ. Civ. Eng. (2017) 1–17.

[44] Y. Li, Y. Chen, G. Zhang, Y. Liu, C. Zhou, A numerical procedure for modeling the seepage field of

water-sealed underground oil and gas storage caverns, Tunn. Undergr. Space Technol. 66 (2017) 56–63.

[45] S. Li, M. Wang, Elastic analysis of stress–displacement field for a lined circular tunnel at great depth

due to ground loads and internal pressure, Tunn. Undergr. Space Technol. 23 (6) (2008) 609–617.

[46] A. Lu, L. Zhang, N. Zhang, Analytic stress solutions for a circular pressure tunnel at pressure and

great depth including support delay, Int. J. Rock Mech. Min. Sci. 3 (48) (2011) 514–519.

[47] M. Wang, S. Li, A complex variable solution for stress and displacement field around a lined circular

tunnel at great depth, Int. J. Numer. Anal. Methods Geomech. 33 (7) (2009) 939–951.

[48] J. C. Jaeger, N. G. Cook, R. Zimmerman, Fundamentals of rock mechanics, John Wiley & Sons, 2009.

[49] M. F. Hawkins Jr, et al., A note on the skin effect, J. Pet. Technol. 8 (12) (1956) 65–66.

38

View publication statsView publication stats