You are on page 1of 24

Applied Mathematical Modelling 62 (2018) 62–85

Contents lists available at ScienceDirect

Applied Mathematical Modelling


journal homepage: www.elsevier.com/locate/apm

An analytical solution for design of pressure tunnels


considering seepage loads
Mohammad Reza Zareifard
Civil Engineering, Estahban Higher Education Center, Ghaem street, Estahban 133445, Iran

a r t i c l e i n f o a b s t r a c t

Article history: In a pressure tunnel with a permeable lining, development of flow from the tunnel induces
Received 8 November 2017 seepage stresses in the surrounding ground and the lining. In this paper, an analytical so-
Revised 16 May 2018
lution for analysis and design of pressure tunnels is presented, based on a generalized
Accepted 22 May 2018
effective stress principle. The problem is considered as axial-symmetric, and the lining
Available online 29 May 2018
and the rock mass are assumed to be poroelastic, homogeneous and isotropic. The pro-
Keywords: posed method covers various types of pressure tunnels, such as unlined, shotcrete lined,
Pressure tunnel concrete lined, reinforced concrete lined and prestressed concrete lined pressure tunnels.
Hydro-mechanical analysis For unlined and uncracked lined pressure tunnels, the hydraulic analysis is independent of
Seepage forces mechanical responses; thus, closed-form expressions for stresses, strains and pore pressure
Effective stress are derived.
Lining design On the other hand, in cracked reinforced concrete linings, the development of cracks
are considered in the mechanical and hydraulic responses. In this case, the hydraulic and
mechanical analyses are coupled; therefore, the hydraulic and mechanical analyses are car-
ried out by an iterative procedure. The applicability of the proposed solution for the design
of pressure tunnels is illustrated and discussed by using an illustrative example problem.
© 2018 Elsevier Inc. All rights reserved.

1. Introduction

Pressure tunnels are conduits which convey fluids, most often water under pressure. They are either permeable (unlined,
shotcrete lined, concrete lined, reinforced concrete lined or prestressed concrete lined) or impermeable (steel lined, water-
proofing membrane lined or lined with other watertight lining systems). In a permeable pressure tunnel, the seepage forces
are applied to the lining and the surrounding rock mass as body forces, and affect the hydraulic and mechanical responses
of them. However, the analysis and design of pressure tunnels are usually carried out based on the common total stress
approach or the so-called surface load theory, which considers the lining of the pressure tunnel as a impermeable medium
[1,2].
For designing permeable pressure tunnels, the following criteria must be satisfied:

1- Limiting stresses in the lining: the induced stresses in the lining components (concrete and reinforcement) should not
exceed the allowable values.
2- Preventing or controlling cracks in the lining: at low internal water pressures, cracks can be prevented by increasing
thickness of the lining, rock grouting, and can be controlled by reinforcing. However, at high internal pressures, cracks
can only be prevented by prestressing.

E-mail address: zareefard@aut.ac.ir

https://doi.org/10.1016/j.apm.2018.05.032
0307-904X/© 2018 Elsevier Inc. All rights reserved.
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 63

List of Symbols

c concrete cover
Es Elasticity modulus of steel bars
Ec Elasticity modulus of concrete
Er Elasticity modulus of rock mass
FS Factor of safety against hydrojacking
f tc Tensile strength of concrete
f cc Compressive strength of concrete
Hi Water head at internal surface of lining
Ho Water head at internal surface of rock mass
h0 Depth of the tunnel from ground surface
h1 Distance of the tunnel from groundwater surface
r Radial distance from the center of the tunnel
k Permeability.
Km Bulk modulus of matrix material.
Ks Bulk modulus of solid material.
pi Final internal pore-pressure at tunnel wall
pw Induced pore-water pressure
ri Internal radius of the lining
ro External radius of the lining
Rs Radius of influence by seepage
S Average cracks spacing
sb Centre-to-centre spacing of reinforcing bars
ur Induced radial displacement
(x,y) Cartesian coordinates
w Average cracks width;
β Biot constant
εθ Induced circumferential strain
εr Induced radial strain
γw Water specific gravity
μ Dynamic viscosity of water
σs Steel stress
σ Induced Terzaghi’s effective stress
σ  Induced Biot effective stress
σθ Induced circumferential stress
σr Induced radial stress
σ(PRro )
prestressing pressure
φ Bar diameter
ρs Reinforcement ratio
θ Angle measured clockwise from the horizontal direction.
ν Poisson’s ratio
Subscript c Refers to quantities corresponding to the concrete
Subscript g Refers to quantities corresponding to the grouted zone
Subscript l Refers to quantities corresponding to lining.
Subscript r Refers to quantities corresponding to the rock mass.
Subscript s Refers to quantities corresponding to the steel bars.
Superscript BU Refers to the quantities induced by the boundary pressures
Superscript SE Refers to the quantities induced by the seepage forces
Superscript PR Refers to the quantities induced by the prestressing pressures
Subscript (r) Refers to quantities corresponding to the radius r.

3- Limiting water losses: Water losses can be limited by preventing cracks in the concrete lining, and to some extent by
reinforcing, as well as, rock grouting. If these measures are not effective, sealing of the lining is performed.
4- Adequate confinement: Adequate confinement refers to the ability of a rock mass to withstand the internal pressure
from an unlined waterway. If the confinement is not adequate, hydraulic jacking may occur. The jacking pressure
can be defined as the equivalent hydraulic pressure required for increasing the hydraulic conductivity of the jacking
surface significantly. Hydraulic jacking can occur in any unlined pressure tunnel, as well as, in pressure tunnels with
shotcrete and concrete linings.
64 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

Other criteria such as stability of the adjacent slopes and the long-term stability must be controlled, in some cases.
Although, full consideration of the interactions between construction activities, groundwater conditions, rock and lining
characteristics, fracture or crack opening, and transient behavior dictate the use of numerical simulations combined or not
combined with analytical models such as that proposed by Kang et al. [3]; Olumide and Marence [4], and Simanjuntak et al.
[5], the insight into the general nature of the solution (the sensitivity of the solution across a range of possible influencing
variables, etc.) that can be gained from the analytical solutions is an important issue that cannot be ignored. In addition,
a number of assumptions and simplifications are always needed in formulating a preliminary-design analysis, and it is
essential that the design engineer be able to assess a general check on the final numerical analysis. The analytical results
provide a valuable means of making this assessment. In this regard, lack of valid analytical solutions is one of the most
important difficulties in designing pressure tunnels.
In the case of ordinary tunnels, the effect of seepage forces on rock-support interaction calculations has been emphasized
and relevant theoretical studies have been reported by researchers [6–16]. However, compared to other type of tunnels,
limited efforts have been allocated for analysis of pressure tunnels within or above groundwater table under seepage forces
[6,17–24]. Some of solutions have been presented for below groundwater table pressure tunnels [6,19–24] and some of them
have been proposed for above groundwater table [4,5,17,19,20].
In this study, based on the generalized effective stress principle, seepage forces under steady-state flow (due to the inter-
nal water pressure) applied to the ground and the concrete lining of above and below groundwater table pressure tunnels
are estimated quantitatively, and relationships for the induced stresses and strains are derived. It should be noted that, here,
the induced stresses (excess stresses) and displacements (excess displacements) due to the internal service loading beyond
the initial values (due the tunnel excavation) were, only, derived. In this regard, the final concrete lining is generally installed
after the convergence of the tunnel; thus, the excavation loading is not applied to it. Conclusively, the final stresses in the
lining are equal to the induced stresses. In the lining, cracks may be formed, if the induced effective stresses, which is equal
to the final stresses, exceed the tensile strength of the concrete. Cracks can be prevented by prestressing or controlled by
reinforcement. On the other hand, in the rock masses, the final stresses are equal to the sum of the initial excavation values
and the induced values.
The proposed solution can be applied for the stability analyses and design of unlined, shotcrete lined, concrete lined,
reinforced concrete lined and prestressed concrete lined pressure tunnels based on the mentioned design criteria.
For the development of the new solution the following assumptions are made: (1) the rock mass and the lining behave
elastically; (2) A steady state for the flow is considered; (3) Darcy’s law is valid; (4) the tunnel cross-section is circular; (5)
the tunnel is deep enough: h0 > 10ro [23], where h0 is depth of the tunnel and ro is the tunnel radius; and (6) the natural
(for below groundwater pressure tunnels) or artificial (for above groundwater pressure tunnels) water levels are located
at sufficiently great distance from the tunnel: Rs > 10ro [19], where Rs is the radius of influence by seepage and ro is the
tunnel radius. On the basis of assumptions 4 to 6, the problem can be considered to be axial symmetry. In this regard, the
objective of the study is to provide an overview of a practical application of the analytical method for the design of pervious
concrete-lined pressure tunnels.

2. Problem statement

Both the theoretical analysis and the experimental observations have shown that the formation of cracks or fractures is
controlled by the Terzaghi’s effective stress, defined as [25–27]:

σ  = σ − pw , (1)
where σ is the induced total stress; pw is the induced pore water pressure (excess pore pressure); and σ  is the induced
Terzaghi’s effective stress (by convention, compressive stress is considered as positive).
On the other hand, deformations are controlled by Biot’s effective stress [25–27]:

σ  = σ − β pw , (2)
where σ  is the induced Biot’s effective stress; and β is Biot’s constant. The Biot’s constant β depends on the bulk modulus
of the both solid matrix and solid grain material, as described by the following equation [26,28].
Km
β =1− , (3)
Ks
where Km and Ks are the bulk modulus of the matrix material and the solid constituent, respectively. In the proposed
method, the analyses are performed based these effective stress laws in the lining and the rock masses. In a saturated
porous rock, the values of β as low as 0.5 may be expected [27].
The problem to be analyzed is represented in Fig. 1. A section of cylindrical pressure lined tunnel of radius ro is excavated
in a poroelastic rock mass characterized by elasticity modulus Er , Poisson’s ratio ν r , and permeability kr .
In the case of a tunnel in a high groundwater table, the initial pore pressure in the ground is uniform and equal to γ w h1 ,
where h1 > >ro is the distance of the tunnel from the groundwater level. On the other hand, in the case of an above water
table pressure tunnel, both h1 and initial pore pressure are equal to zero. The tunnel is then excavated, lined and a uniform
hydraulic pressure pi is applied to the internal surface of the lining, simultaneously.
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 65

⎛R ⎞
log⎜ s ⎟
pw(r ) = pw ( ro ) ⎝ r ⎠
⎛R ⎞
log⎜⎜ s ⎟⎟
⎝ ro ⎠

pw(ri ) = pi − γ wh1 Zone of influence σr

r σθ
No-flow zone At infinity
Lining σr = 0
pw( r ) = 0
pi
ri
ro
Rs

Fig. 1. Geometry, applied loads and boundary conditions for the tunnel (circular hole in an infinite medium under radial seepage forces).

Axial symmetry condition for geometry and loading will be assumed; thus, the radial flow pattern, in which beyond
a radius of influence r = Rs , the original groundwater condition remains unaltered, will be regarded. On the basis of the
utilized non- uniform pore pressure distribution shown in Fig. 1, the induced pore pressure (excess pore pressure) at the
tunnel radius r = ri , is equal to pw(ri ) = pi − γw h1 , and that at beyond the radius r = Rs is equal to pw(Rs ) = 0.
Therefore, two different zones may be formed around the tunnel: an external no-flow region and an internal region of
influence.
Considering the axial-symmetry condition for the problem of Fig. 1, the stresses induced by seepage forces at a distance
r is defined by the induced total radial stress σ r(r) and the induced total circumferential stress σ θ (r) . Furthermore, the
displacement is defined by the induced radial displacement ur(r) .
As mentioned, the present analysis disregards the loads from the ground due to the initial stress field or excavation
load—i.e., the analysis seeks to determine only the stress state and displacement induced by seepage flow or service load.
However, the final stresses can be obtained by superposition of the excavation and service loads.

3. Governing equations

For the considering axial-symmetric flow, integrating the Darcy’s equation v = ki annular elements, a well-known equa-
tion (with a little change) for the induced pore pressure pw (r) in the zone of influence is obtained [19,21,29]:
r    
γw q log Rrs log rro
pw ( r ) = log + pw ( ro ) = pw ( ro )   + pw ( Rs )  , (4)
2π kr ro log Rr s log Rr s
o o

where q is seepage flow rate (the volume of inflowing water per unit time per unit length of the tunnel), and γ w is water
specific gravity. At radius Rs , the induced pore pressure, i.e. pw(Rs ) ,is equal to zero. In Eq. (4), inward seepage flow rate q is
assumed to be positive.
Furthermore, a radial flow is assumed for estimation of the induced pore pressure in the concrete lining, pw(r) , with
permeability kl . Thus, a similar equation is obtained, for pw(r) in the concrete lining [19,21,29]:
r    
γw q log rro log rri
pw ( r ) = log + pw ( ri ) = pw ( ri )   + pw ( ro )  ro  , (5)
2π kl ri log rro log r i i

where ri is the internal radius of the lining, pw(ri ) = pi − γw h1 is the induced water pressure at internal surface of the lining,
and kl is permeability of the lining.
For concrete linings, the permeability kl is controlled by the annular shrinkage cracks. A permeability magnitude about
1.7 × 10−8 ms can be considered for typical concrete linings with shrinkage cracks [30].
At the radius of influence Rs , pore water pressure is equal to the original ground water pressure and the induced pore
water pressure becomes equal to zero. Thus, the following equation of the seepage flow rate, q for the radial flow can be
derived as follows:
pw ( Rs ) − pi 1
q = 2π  ro   Rs  (6)
γw 1
kc
log ri
+ 1
kr
log ro
66 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

Under axi-symmetrical condition, the induced stress field for each element of the ground and lining has to fulfill the
equilibrium equations as [31]:
dσ  r (r ) σ  θ (r ) − σ  r (r ) d pw ( r )
− +β =0 (7)
dr r dr
dp
In Eq. (7), the term, β dr
w (r )
, is commonly referred as the seepage body force and acts through the pore pressure gradient.
For axial-symmetric condition, the radial strain ε r(r) and circumferential strain ε θ (r) at radius r can be stated in terms of
the radial displacement ur(r) , as follows:

ur ( r ) d ur ( r )
εθ ( r ) = , εr ( r ) = (8)
r dr
The relationships between the induced strains, ε r and ε θ , and induced Biot effective stresses σ   r and σ   θ in each zone
(the rock mass and lining) are given by Hooke’s law for plane-strain condition [31]:
E
σr = [(1 − ν )εr + νεθ ] (9)
( 1 + ν ) ( 1 − 2ν )
E
σθ = [(1 − ν )εθ + νεr ], (10)
( 1 + ν ) ( 1 − 2ν )
where Eand ν are elasticity modulus and Poisson’s ratio, respectively.
Substituting Eqs. (9) and (10) into Eq. (7) and applying Eq. (8) gives the following equation for the unknown induced
radial displacement ur(r) .

ur ( r ) 1 d ur ( r ) d 2 ur ( r ) d pw(r ) (1 + ν )(1 − 2ν )
− − =β , (11)
r2 r dr dr2 dr E (1 − ν )
where ur and pw are the induced radial displacement and the induced pore pressure at the radial distance, r, respectively
and E and ν are Elasticity modulus and Poisson’s ratio of the considering zone, respectively.
In this regard, the components of induced stresses and strains corresponding to the radial seepage forces and the induced
boundary pressures in an annular region with internal and external radii ra and rb are derived in Appendix A.
The solution presented in Appendix A can be utilized for the analyses of the uncracked lining, and the zone of influence
of the rock mass and the no-flow rock mass zone:

1. The uncracked lining has an internal radius ri (at which the induced hydraulic and mechanical pressures pw (ri ) = pi −
γw h1 and σr(r ) = 0 are applied, respectively) and an external radius ro (at which the induced hydraulic and mechanical
i
pressures pw(ro ) and σr(r ) are applied, respectively).
o
2. The zone of influence of the rock mass has an internal radius ro (at which the induced hydraulic and mechanical
pressures pw(ro ) and σr(r ) are applied, respectively) and an external radius Rs (at which the induced hydraulic and
o
mechanical pressures pw(Rs ) = 0and σr(R ) are applied, respectively).
s
3. The no-flow zone of the rock mass has an internal radius Rs (at which the induced hydraulic and mechanical pres-
sures pw(Rs ) = 0 and σr(R ) are applied, respectively) and an external radius r = ∞ (at which the induced hydraulic and
s
mechanical pressures pw(∞) = 0and σ  r(∞) = 0 are applied, respectively).

It is observed that, for calculating stresses and strains in each zone, the influence radius of seepage, and the seepage
flow rate for the equivalent radial flow are only needed.

4. Pore pressure distribution and the radius of influence

Here, two cases of above groundwater and below groundwater pressure tunnels are considered. It should be mentioned
that, using the exact pore pressure distribution in mechanical analysis reduces the simplicity of the solution and so its value.
Furthermore, the derived equations cannot be solved analytically, in the case of above groundwater table tunnels. Thus, in
the proposed solution, axial symmetric condition is used in mechanical analysis and the utilized pore pressure distribution
is independent of the direction and can be derived as a function of the radial distance r. In this regard, seepage flow rate
and an equivalent radius of influence should be obtained.

4.1. Pressure tunnels above water table

As shown in Fig. 2, in an above groundwater level pressure tunnel, seepage flow develops an artificial groundwater level.
The seepage flow in a medium with homogeneous and isotropic permeability can be formulated in terms of a function of
complex variables named characteristic function, in which its real part represents the equipotential lines, and its imaginary
part represents the flow lines.
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 67

Fig. 2. Flow pattern around pressure tunnel above groundwater level- the dashed and the solid curves represent the flow lines and contour lines of pore
pressure, respectively.

In the case of an above groundwater table pressure tunnel, Bouvard and Pinto [17] determined the characteristic function
of the seepage flow, f(z), as:
 q
  3qx  qθ 3qy
f (z ) = log x2 + y2 + +i − , (12)
 4π
4a
 2π
4a
equipotential lines
flow lines

where r and √ θ are the polar coordinates; z = x + iy is a complex variable (x and y are the Cartesian coordinates as shown in
Fig. 2), i = −1 is called the imaginary unit, ro is radius of the tunnel; a = − kq is called the characteristic parameter of the
r
flow, q is seepage flow rate, and kr is permeability of the rock mass.
The value of a is calculated by solving the following equation [17]:
3 a
 a 
hw ( ro ) − ro = log , (13)
4 2π π ro
where hw(ro ) is the hydraulic head at the rock-lining boundary.
The hydraulic head is related to the pore pressure by Bernoulli’s equation:
pw (r,θ )
+ r sin θ = hw (r,θ ) (14)
γw
Furthermore, the following equation can be used for obtaining the radius of influence (here, the reach of the seepage
flow) through the vertical direction Rsy [17]:
a
Rsy = log 2 (15)
π
The pore pressure in the rock mass through the tunnel crown pw(r,90◦ ) is obtained as [17]:
  
−γw q sh π2ar 1 qγw
pw(r,90◦ ) = log  π ro  + ( r − ro ) + pw ( ro ) , (16)
2π kr sh 4 kr
2a

where q is seepage flow rate, ro is external radius of the lining, pw(ro ) is pore pressure at the boundary between the rock
and lining, γ w is the specific gravity of water, and kr is permeability of the rock mass.
The governing equation of curves with constant pore pressure passing through a point yo on y-axis can be obtained using
Eq. (12), Darcy’s equation and Bernoulli’s equation after some manipulations:
  2π y   2π C  
a exp + 1 − 4 exp
x= arccos
a
  a
, (17)
π 2 exp πay
68 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

Fig. 3. Flow pattern around a pressure tunnel with radius ro , beneath a horizontal groundwater table- flow lines and contour lines of hydraulic head.

where:
1 1
 2π y0
  1
πy  1

0
C= Log exp − exp + (18)
4π 4 a 2 a 4
At radius ro , Eq. (16) must be equal to Eq. (5); since, the pore pressure must be continuous over the boundary. The
following equation for the seepage flow rate, q is derived:
γw q r  3 γ q  −q 
pi + log
o
− γw ro + w log =0 (19)
2π kl ri 4kr 2π kr π ro kr
A closed-form expression for q cannot be obtained. Therefore, q must be found numerically by use of suitable root-finding
algorithms such as Newton–Raphson method [32].
It should be noted that, in the mechanical analysis a radial flow pattern is used. In this way, for the utilized symmetrical-
flow regime, an equivalent radius of influence can be calculated by solving Eq. (6) for Rs using the seepage flow rate obtained
from Eq. (19) as:
 r − kkcr  
i −2π kr pi
Rs = ro exp (20)
ro γw q

4.2. Pressure tunnels below water table

In the case of a pressure tunnel with radius ro , at a constant hydraulic head, beneath a horizontal groundwater table in
a homogeneous and isotropic permeable medium (see Fig. 3), the characteristic function of the seepage flow, f(z) is given by
[29]:
r c ( z + cw ir c )
f (z ) = (21)
−cw iz + rc
In this way, the induced pore-water pressure pw at any point around the tunnel is estimated as (Fig 3):
  2 
q c2 r 2 cos2 θ + c r sin θ − c2 ro
pw (r,θ ) = log (22)
2π kr c2 r 2 cos2 θ + (c r sin θ − ro )
2


h1 − h21 − ro2
c= (23)
ro
The seepage flow rate q is calculated from the following equation:
 
2π kr Pwf (ro ) − γw h1
q=  √ 2 2 , (24)
γw log h1 + roh1 −ro
where ro is the tunnel radius, h1 is distance of the tunnel from the groundwater level, (r, θ ) are the polar coordinates,
Pw(r ) = pw(r ) + γw h1 is the final pore pressure at the interface between the lining and the rock mass, and γ w is the specific
f
o o
gravity of water.
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 69

Fig. 4. Variation of bond stress (a), steel stress (b) and tensile concrete stress (c) in the vicinity of a crack according to the CEB-FIP model code [35].

At the interface between the lining and the rock mass, the pore pressure calculated in the lining must be identical to that
in the rock mass, since, the pore pressure must be continuous over the boundary. The following equation for the seepage
flow rate, q is derived from these continuity condition:
pi − γw h1 1
q = 2π  ro   √ , (25)
γw − k1c log + 1
log
h1 + h21 −ro2
ri kr ro

where pi is internal water pressure, and γ w is the specific gravity of water.


For the utilized symmetrical flow regime in the mechanical analysis, an equivalent radius of influence can be computed
by solving Eq. (6) for Rs using the seepage flow rate obtained from Eq. (25) as:

Rs = h1 + h21 − ro2 (26)

5. Cracked lining

5.1. Initiation of cracks in the concrete lining

A reinforced concrete lining is expected to crack longitudinally due to internal pressure loading. In this case, the perme-
ability of the lining will no longer be controlled by annular shrinkage cracks but by the width of the longitudinal cracks.
The lining cannot resist effective stresses more than concrete tensile strength. Thus, the lining will crack, as soon as, the
maximum tensile effective stress (circumferential effective stresses induced in the radius ri of the lining) exceeds the tensile
strength of the concrete. Thus, the condition for formation of the initial cracks is:
σθ (ri ) = ftc , (27)

where f tc is the tensile strength of the concrete.

5.2. Hydraulic and mechanical analyses

In a cracked reinforced concrete lining, the applied seepage body forces corresponds to the hydraulic gradient are de-
pendent on the lining permeability. On the other hand, the permeability of the lining depends upon width of the cracks
developed by the applied seepage loads. This mechanical–hydraulic coupling between seepage and cracks width must be
considered in the mechanical and hydraulic analyses of pressure tunnels with concrete lining.

5.3. Crack width and spacing

Reinforcement is provided in the concrete liner for controlling cracks and for improving the stability of the lining. Rein-
forced concrete linings provide some resistance to the leakage and it should not be considered as a watertight.
70 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

Assuming a laminar flow regime through the longitudinal cracks of aperture (w), evenly distributed around the perimeter
of the lining, the equivalent permeability of the cracked concrete lining can be estimated using the cubic law for viscous
flow between parallel plates [30]:
γw w3
kl = , (28)
12μ S
where w is average cracks width; S is average cracks spacing; μ is dynamic viscosity of water (μ = 10 − 3 Pa.s at 20◦ C).
Based on various test results, Chowdhury and Loo [33] developed an equation for predicting the average spacing, S,
between tensile cracks in reinforced concrete members in terms of the concrete cover, bar spacing, bar diameter and the
reinforcement ratio as follows.
 
φ
S = 0.6 ( c − sb ) + 0.1 , (29)
ρs
where ϕ is bar diameter; ρ s is reinforcement ratio; sb is centre-to-centre spacing of reinforcing bars; and c is concrete cover.
In Eq. (29), it is observed that by keeping the reinforcement ratio unchanged, by decreasing the diameter of reinforcing bars,
the cracks spacing will become larger. As seen in the next sections, decreasing in the spacing of the cracks has the advantage
that it decreases the width of the cracks.
Different researchers have proposed various theoretical and experimental formulas for crack spacing in reinforced con-
crete members. Each of them is applicable for specific geometrical, loading and reinforcement conditions. Thus, in using Eq.
(29), care must be taken. However, in the proposed solution the utilized relationship (29) can be easily replaced by other
relationships. In this regard, parametric and sensibility analyses may be helpful to study the effect of the crack spacing or
the utilized formula on results.

5.4. Stress and strain variations in the reinforcement

The tensile concrete located between two cracks helps to enhance the stiffness of the lining, because the bond strength
between the bars and the concrete is always active in these areas. This is called the “tension stiffening” phenomenon and
is well known to civil engineers. For determining cracks width and calculations of rock –lining interaction, the estimation
of the stress variations in the reinforcement bars and the concrete is essential. However, defining exact stress variations in
the concrete and the reinforcement is practically complicated. Thus, here, two bound models are introduced to estimate the
response of the reinforced concrete lining, here, named upper bound and lower bound.

5.4.1. Upper bound model


In this case, the tension stiffening effect is assumed to be negligible; thus, the bare bar response will be taken into
account and the bond force acting around the steel bars is not incorporated in the calculations. Thus, steel stress is constant
between cracks and is equal to σs = σ̄s (where σ̄s is average stress in the reinforcement); while, concrete stress is equal to
zero.
In this case, the average crack spacing S is multiplied by the lining strain (which is equal to the steel strain) to determine
the average crack width w:
σs
w=− S, (30)
Es
where σ s is steel stress and Es is elasticity modulus of the steel bars.
However, some stress and strain is maintained in concrete surrounding the reinforcement due to the action of bond;
thus, the actual stiffness of the lining is greater than that considered in this model.

5.4.2. Lower bound model


In this case, for evaluating the concrete tension stiffening effect on the reinforced concrete section, a simplified method
proposed by CEB-FIP model code [34] is used.
Between two cracks, the load contribution of the concrete is activated by the bond effect between the reinforcements
and the concrete. The necessary distance for the reinforcements to redevelop the full tensile force in the concrete, lost by
the occurrence of the crack, is the transmission length lsmax . The CEB-FIP [34] simplified model code assumes constant
bond strength along the steel bars. As a consequence, the development of the steel stress and the concrete stress over the
transmission length is linear. In addition, in order to quantify the tension stiffening effect in the cracked member, in the
CEB-FIP model code, the average distance between two cracks is taken as S = 1.5lsmax . It is assumed that, the axial stress
distributes uniformly across the full thickness of the concrete section. Consequently, at mid-distance between two cracks,
the steels work under the stress σ smin , and the concrete works under the constant stress σ cmax = −0.75f tc . The resulted
stresses distribution between two cracks is shown in Fig. 4.
Based on these stress variations, the average stresses in the concrete and reinforcement can be evaluated as [35]:
3 f  tc
σ̄s = σs max + (31)
8ρse
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 71

3
σ̄c = − ftc , (32)
8
where:
As
ρse = (33)
ACe
As is the area of the hoop reinforcement per unit length of the tunnel and, ACe is the contributing area of concrete
affected by the hoop reinforcement, which attains in maximum 7.5 times the diameter of the steel bar outside of the rein-
forcement [20]
In this case, for evaluating crack width, it is necessary to calculate the difference between the average steel deformation
and average concrete deformation within the transmission zone. This difference is approximated by the following [34]:
 
σ̄s σ̄c
w = −S − (34)
Es Ec

5.5. Load taken by the reinforcement

The loading on the reinforcement can be obtained by a compatibility condition. For determining the radial displacement
of reinforcement, the average circumferential stress in the reinforcement is used. Based the axial symmetric condition, the
reinforcement is regarded statically as a steel liner with equivalent thickness, which embedded between two concrete linings
in the same manner as Schleiss [20]. This corresponds to the assumption that, like a steel lining, the reinforcement interacts
with the surrounding internal and external cracked concrete linings.
Here, for the analysis of two cracked concrete linings the Biot constant is taken equal to 1. Since the problem is ax-
p −p
ial symmetric, solving equilibrium equation for the cracked concrete linings under seepage force, ddrp = o roi , the total
r log( r )
i
pressure σrs applied to the steel lining is obtained:
ro
σrs = σr (ro ) − λ1 (35)
rs

( po − pi ) ( ro − ri )
λ1 = − r  (36)
rs log o
ri

Thus, the following equations are obtained for strain and stress in the steel bars.
σrs rs
σs max = (37)
ρs ( r o − r i )

σs max + λ2
ε̄s = (38)
Es

3 f  tc
λ2 = . (39)
8ρse

Where rs , Es and ρ s are radius of reinforcement, elastic modulus of reinforcement, and reinforcement ratio, respectively.

5.6. Boundary pressures

At radius ro , the zone of influence and the lining interact with each other through a mechanical pressure σr(r ) ; and
o
at radius Rs the zone of influence and the no-flow zone of the rock mass interact with each other through a mechanical
pressureσr(R ) . Considering a proper contact at the lining-rock boundary resulted from a thorough contact grouting program,
s
these boundary pressures are obtained for both cases (cracked and uncracked linings) from the compatibility conditions
εθ (r ) = εθ (r ) of influence and εθzone
liner
o
zone
o (R )
of influence = ε no− f low zone :
s θ (R ) s
⎛ SE (zone of influence)    ⎞
εθ ( Rs ) Ec (1 + νr ) ro2 − ri2 R2s − ro2 (1 − 2νr )
⎜ SE (zone of influence)   ⎟
Er ⎜−εθ (Rs ) Er (1 + νc ) R2s − ro2 ro2 (2νc − 1 ) − ri2 ⎟
⎝    ⎠
+ εθ (r( )
SE zone of influence )
− εθ (r( ) ) 2Ec ro2 (1 + νr ) ro2 − ri2
SE liner

σr(Rs ) = − 2     
o o
(40)
2Rs (2νr − 1 ) Er (1 + νc ) ro2 (2νc − 1 ) − ri2 − Ec (1 + νr ) ro2 − ri2
72 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

  
Er Ec ro2 − ri2 εθSE(R(zone
s)
of influence )
− εθ (r( )
SE zone of influence )
+ εθ (r( ) )
SE liner

σr(ro ) = − 2   2 2
o o
(41)
E r ( 1 + νc ) r o ( 2 νc − 1 ) − r i − E c ( 1 + νr ) r o − r i
2

for an uncracked lining and,


⎛    ⎞
−εθ (r( )
SE zone of influence )
2ro2 ρs Es
νr2 − 1 ro − ri
⎜ SE zone of influence)  2 2    ⎟
o

Er ⎝−εθ (R( ) roEr Rs − ro − 2λrs ro2 νr2 − 1 ⎠


s
SE (zone of influence )
  
−εθ (R ) ρs E s ( 1 + ν r ) r o − r i R s − r o ( 2 ν r − 1 )
2 2
σr(Rs ) = − s
 2     (42)
2 R s ν r − 1 ρs E s ( 1 + ν r ) r o − r i + r o E r
2

  SE zone of influence)   
Er Es ρs ro − ri εθ (R( ) − εθ (r( )
SE zone of inf luencet )
− λrs ρs / ro − ri Es
σr(ro ) = −  
s o
(43)
ρs E s ( 1 + ν r ) r o − r i + r o E r
ρs Es (ro−ri )λ1
for a cracked reinforced lining (by neglecting the radial compressibility of the concrete), where, λ = λ2 − rs ,
εθSE(r(oliner
)
)
, is the circumferential strain in SE (zone of influence )
the lining at radius ro corresponding to the seepage forces; εθ (r )
o
is
the circumferential strain in the zone of influence at radius ro corresponding to the seepage forces; and εθSE(R(zone of influence )
s)
is the circumferential strain in the zone of influence at radius Rs corresponding to the seepage forces. These strains are
calculated from formulations presented in Appendix A.
It must be notified that, in the case of an unlined pressure tunnel, the mechanical pressure at the excavation surface
is zero, i.e. σr (ro ) = 0, and the boundary pressure between the zone of influence and the no-flow zone is obtained from a
compatibility condition at radius Rs (εθzone of influence = ε no− f low zone
( Rs ) θ ( Rs ) ) as follows:

εθSE(R(sat
s)
)
σr(Rs ) =    (44)
2R2s Er νr − 1 R2s − ro2
2

6. Effect of prestressing

At high pressure heads, can only be prevented by prestressing. In this regard, both prestressed concrete linings and steel
linings prevent the excessive seepage flow. However, in contrast to the steel lining, the prestressed concrete lining is a more
popular alternative. Because, it shortens the construction period, increases the durability and reduces the costs.
The general principle in the design of a prestressed concrete lining is to place the concrete in compression equivalent to
or greater than the tension which is induced due to internal pressure, in order to minimize cracking, and thereby, provide a
near- watertight lining. The techniques used to apply the prestress pressure include circumferential steel cables, grouting at
the concrete-rock interface, or grouting of a double membrane between two concrete linings [18,36,37].
By using combination of these prestressing techniques, the composite action of the rock and the lining is attained. In
this case, the rock and the lining will remain in contact and a uniform prestressing pressure around the entire surface of
the lining will be distributed. Consequently, the performance of the prestressing is improved [36].
In this paper, the effect of prestressing on the lining is considered by uniform boundary pressure σ(PR ro )
applied to the
external radius of the lining. Using superposition concept, the stresses induced by this prestressing pressure are added to
the seepage induced stresses in the lining.
The stresses induced by the prestressing pressure in the lining are obtained by using Lame’s solution:
 
ri2 ro2
σr(PRr ) = σ(PR
ro ) 1− + σ(PR
ro ) (45)
ro2 − ri2 r2
 
ri2 ro2
σθPR
( r ) = σ ( ro )
PR
1+ + σ(PR
ro ) (46)
ro2 − ri2 r2

7. Calculation procedure

For unlined and uncracked lined tunnels, the hydraulic analysis is independent of mechanical results. In these cases,
after obtaining the hydraulic results, the mechanical results can be calculated, directly. On the other hand, when cracks
are formed in the reinforced concrete sections, the hydraulic analysis is a function of width and space of the cracks. In
this case, the hydraulic and mechanical analyses are dependent and are performed based on the following step-by step
calculation procedure:

(A) Select a trial value for permeability of the cracked lining (kl < kc ).
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 73

y
h0 = 100 m

ro = 4 m
30•

ν r = 0.3, β r = 0.8
x
σ 1′0 = 0.0235(h0 − y )
σ 3′0
= 0.7
σ 1′0
Fig. 5. Schematic representation and properties of the tunnel.

(B) Calculate seepage flow rate, the radius of influence and pore pressure distributions in the lining and the rock mass by
using equations presented in Section 4.
(C) Calculate the mechanical boundary pressures σr(r ) and σr(R ) by using Eqs. (42) and (43).
o s
(D) Calculate stress distributions in the rock mass by using equations presented in Appendix A.
(E) Determine stress distribution in the cracked lining by using equations presented in Section 5.3 and 5.4.
(F) Compute the width and spacing of the cracks in the reinforced concrete lining by using Eqs. (29), (30) and (34).
(G) Calculate the permeability of the cracked lining from Eq. (28).
(H) With the obtained lining permeability repeat the calculation until the width of cracks in the reinforced concrete lining
remains constant.

8. Practical example

An illustrative example is given to demonstrate the performance and potential for practical use of the proposed method
(Fig. 5).
A pressure tunnel of radius ro = 4 m excavated in a rock mass, with Poisson’s ratio, vr = 0.3, Biot–Willis constant β r = 0.8
and permeability , parallel to a deep valley, with slope face of 30◦ to the horizontal, is taken into account. The normal rock
cover, h0 above the tunnel is 100 m. The unit weight of the rock is assumed to be γr = 0.0235 MN m3
, and the ratio of minor to
σ30
major principal in-situ stresses is assumed as: = 0.7. The tunnel is under internal water head of Ho = 350 m.
σ10
In the proposed solution, the radial-flow pattern with an equivalent radius of influence is used. This assumption theoret-
ically is correct for deep tunnels from groundwater table.
In order to examine the effect of the made assumptions, the results for the pressure tunnel (unlined case) excavated
below groundwater table obtained from the proposed solution are compared with those obtained from the exact analytical
solution presented by Fahimifar and Zareifard [23].
In this regard, the induced pore pressure obtained by two methods are shown in Fig. 6 for two cases of h1 = 10 m ( Rros =
4.8: small radius of influence) and h1 = 100 m ( Rros = 50: medium radius of influence). The results show that, as predicted, in
deep tunnels from groundwater level (large radius of influence) the induced pore pressure obtained in different directions
are close to the proposed solution in the vicinity of the tunnel (governing the mechanical analyses), and thus the considered
assumption is correct. However, in shallow tunnels from groundwater level, the induced pore pressures are different for
various directions and the proposed axial symmetric solution varies between the results obtained for different directions.
The comparison between the induced Terzaghi effective radial and circumferential stresses distribution predicted by
Fahimifar and Zareifard’s method [23] and the proposed solution in this paper for case h1 = 100 m are shown in Fig. 7.
The induced effective stresses along the springline obtained by Fahimifar and Zareifard’s method [23] are practically equal
to the corresponding stresses predicted by the proposed solution. For the other directions the results diverges away from
the tunnel surface to some extent. While, near the tunnel, the results obtained by the proposed solution and Fahimifar and
Zareifard’s solution [23] for different directions are identical (which is important in stability analysis), indicating that the
utilized simplification has negligible effect on the induced stresses near the walls of the deep tunnel. In fact, the zone of
interest for the design of pressure tunnels is very close to the opening, where the analytical solution provides appropriate
results.
74 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

Fig. 6. Distribution of the induced pore pressure predicted by the proposed solution and other solutions for cases h1 = 10 m and h1 = 100 m.

The results obtained by Schleiss’s approach [19] assuming the radius of influence equal to the tunnel distance from the
ground-water surface are also shown in these figures. It is observed that Schleiss’s solution [19] underestimates stresses in
the rock mass.
The results of Lame’s solution, which assumes that the lining is impervious and the internal pressure acts as a mechanical
load, are also shown in Fig. 7. As is observed in this figure, in contrast to Lame’s solution, seepage affects a wide zone of
the rock mass surrounding a pervious pressure tunnel.
For the design of the pressure tunnel, three alternatives are considered:
Alternative 1: an unlined tunnel.
Two cases of unlined below water table (the water table is located at the ground surface) and above water table pres-
sure tunnels are considered. The stability of both unlined pressure tunnels under internal water heads of Ho = 100 m and
Ho = 350 m is investigated. It is observed that, totally, 4 cases should be analyzed by the proposed method.
It is clear that for the tunnel below water table, under internal water head Ho = 100 m, at the tunnel surface, the internal
and external water heads are equal. Thus, the induced stresses in the rock mass are equal to zero.
The contour lines of pore pressure (equal pore pressure curves) obtained for the above water table pressure tunnels
(under Ho = 350 m andHo = 100 m) are plotted in Figs. 8 and 9. It is seen that for the case Ho = 350 m the seepage flow
intersects the valley surface; this may lead to instability problems, which should be controlled. The calculated seepage flow
Li Li
rate for the above water table tunnels under Ho = 100 m and Ho = 350 m are −2.21 m.s and −5.76 m.s , respectively. While,
Li
the calculated seepage flow rate, for the below water table tunnels under ho = 350 m, is −4.02 m.s .
The potential for hydrojacking of each plane at any location can be evaluated by the ratio of the minor principal in-situ
stress over the induced stress perpendicular to that plane [21]. In this way, the factor of safety against hydrojacking of a
plane parallel to the valley surface passing through the center of the tunnel (along x-axis) is defined as:

initial
σmin ( h0 )
F S (x ) = , (47)
σ  θ (x )

where σmininitial = σ 0 is the minor principal in-situ stress at the tunnel level; and σ 
( h0 ) 3 θ (x ) is the calculated induced circumfer-
ential stress.
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 75

Fig. 7. Distribution of the induced radial and circumferential stresses predicted by the proposed solution and other solutions.

On the other hand, the factor of safety against hydrojacking of planes parallel to the valley surfaces at any point of the
y-axis is given by:
initial
σmin (y )
F S (y ) = , (48)
σ  r (y )
where σmininitial = σ 0 is the minor principal in-situ stress through the tunnel crown at the radius y; and σ  is the calculated
(y ) 3 (y ) r (y )
induced radial stress at the radius y. The calculated values of FS(x) and FS(y) for the unlined tunnel, are plotted in Figs. 10
and 11. The factor of safety less than or equal to 1 observed for the tunnels under ho = 350 malong x-axis indicates that,
in the vicinity of the tunnel walls, hydrojacking can be initiated from the tunnel perimeter into the surrounding rock mass
along x-axis. However, since the factor of safety increases away from the tunnel, the hydrojacking along this plane will be
arrested a short distance from the tunnel. It is observed that for the tunnel above water table under Ho = 350 m, the factor
of safety of planes normal to y-axis is larger than unity in the vicinity of the tunnel but decrease rapidly away from the
tunnel to a nearly constant value close to one. Therefore, the hydrojacking phenomenon can propagate through the rock
mass away from the tunnel toward the valley walls. Comparison of the results presented in Figs. 10 and 11 shows that, as
expected, the external pore pressure developed by the natural groundwater table increases the stability of pressure tunnels.
It can be concluded that in the early stages of operation of an pressure tunnel above water table, the internal load
should be applied gradually, this increase the external pore pressure by developing a natural ground water table. If possible,
of course, the tunnel should be placed such that the hydraulic grade line is below the water table. The determination of the
permanent ground water table is a very important task and experienced geologists and ground water hydrologists should
be involved in this aspect. If such information is not obtained, a conservative design approach is necessary. In this regard,
parametric studies, which can be carried out with the proposed method, may be beneficial.
76 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

140

120

100
Valley surface
80

60

40

20

0
0 20 40 60 80 100 120 140 160 180 200 220 240 260
-20
y (m)

-40

-60

-80

-100

-120

-140

-160

-180

-200
x (m)
Fig. 8. Contour lines of pore pressure for the unlined pressure tunnel above water table under internal water head ho = 350 m, ( γpww = 50 m ).

If the developed leakage from the tunnel is considered as acceptable, this design option (unlined section) for the pressure
tunnel under Ho = 100 m can be safe and economic. However, for the pressure tunnel under Ho = 350 m, other alternatives
must be taken into account.
It should be noted that, some seepage loss may be allowed depending upon the quantity and value of available water
and the probable effect of seepage on the stability of the terrain and its effect on the environment. A comparison of the
long-term value of the water versus the cost of limiting the seepage can therefore be made.
Alternative 2: a reinforced-concrete lined tunnel is characterized by the following parameters:

ri = 3.7 m, ro = 4.0 m, kc = 10−7 ms , f  cc = 35 MPa, f  tc = −1 MPa


Ec = 25 GPa, vc = 0.3,
ρs = 0.0052, ds = 20 mm, sb = 20 cm, rs = 3.8 m, Es = 200 GPa,
It is assumed that the lining remains in contact with the surrounding rock mass. In this regard, grouting would be carried
out through the concrete liner to improve the elastic interaction between the concrete liner and the surrounding rock.
The pressure tunnel excavated above the water table is investigated under internal water heads of Ho = 100 m and
Ho = 350 m.
For the case Hi = 350 m the calculated maximum stress in the reinforcement for the upper bound and the lower bound
models are 93 MPaand 168.1 MPa, respectively; while for the case Hi = 100 m, the calculated maximum stress in the rein-
forcement for the upper bound and the lower bound models are 26 MPa and 101 MPa, respectively. The results show that
maximum stress in the reinforcement at the cracked sections obtained based the lower bound model is greater than that of
the upper bound model. Because, in lower bound case, the lining is stiffer, due to bond effect; thus, the load contribution
carried by the lining is greater. However, the estimated crack widths obtained based on both the upper bound (neglect-
ing from tension stiffening effect) and the lower bound (considering the tension stiffening effect) models for the cases
Hi = 350 m and Hi = 100 m are 0.151 mm and 0.043 mm, respectively.
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 77

60

40

20

0
0 20 40 60 80 100 120

-20
y (m)

-40

-60

-80

-100

-120
x (m)
Fig. 9. Contour lines of pore pressure for the unlined pressure tunnel above water table under internal water head ho = 100 m, ( γpww = 50 m ).

The contour lines of pore pressure (equal pore pressure curves) obtained for case Hi = 100 m are plotted in Fig. 12. For
cases Hi = 350 m and Hi = 100 m, water heads equal to Ho = 349.3 m and Ho = 89.4 m will be applied to the internal surface
of the rock mass, respectively.
The safety factors against hydrojacking in rock masses surrounding the pressure tunnel with the cracked reinforced con-
crete lining (FS(x) and FS(y) ) is presented in Fig. 13. The results show that the cracked reinforced concrete lining does not
improve the performance of the pressure tunnel significantly (hydraulically and mechanically), and nearly full internal pres-
sures will be applied at the internal surface of the rock mass, especially for tunnels under high internal pressures. Thus,
using reinforcement for controlling successive leakage flow from pressure tunnels under high internal pressure is not effec-
tive measure. However, the reinforced concrete lining contributes in the stability of pressure tunnel; because it can increases
the hydraulic efficiency of the tunnel, and it prevents the progressive erosion of weak rocks, resulting in rockfalls which can
reduce the capacity of the tunnel.
In order to effectively control the leakage, the strain in the reinforcing bars must be kept low by limiting the stress in
the bars to a value considerably lower than that normally allowed. In high internal pressures, prestressed concrete linings
or a steel linings can be considered.
Alternative 3: a prestressed-concrete lined tunnel characterized by the following parameters:

ri = 3.7 m, ro = 4.0 m, kc = 10−7 m


s
, f  cc = 35 MPa, f  tc = −1 MPa
Ec = 25 GPa, vc = 0.3,
σ(PR
ro ) = 1.3 MPa

The pressure tunnel excavated above water table under internal water head of Hi = 350 m is investigated. In this case,
the circumferential stresses in the lining are compressive (maximum compressive stress at the internal surface of the lining
is equal to σθ (r ) = 0.556 MPa < fcc
 ).
i
78 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

10 FS (x-axis, Ho=350 m)

9 FS (y-axis, Ho=350)

8 FS (x-axis, Ho=100 m)

7 FS (y-axis, Ho=100)

6
FS

0
4 9 14 19 24 29 34 39 44 49
r (m)
Fig. 10. Distribution of factor of safety against hydrojacking of planes parallel to the valley surfaces along x and y-axes for the tunnel above water table.

5
FS (x-axis, ho=350 m)

4
FS (y-axis, ho=350)

3
FS

0
4 9 14 19 24 29 34 39 44 49
r (m)
Fig. 11. Distribution of factor of safety against hydrojacking of planes parallel to the valley surfaces along x and y-axes for the tunnel below water table.

The applied water head to the rock surface will be decreased to Ho = 91.9 m. The contour lines of pore pressure (equal
pore pressure curves) and distribution of factor of safety against hydrojacking of planes parallel to the valley surfaces along
x-axis, are plotted in Figs. 14 and 15, respectively. In the vicinity of the tunnel along x-axis the factor of safety is less than
one. It should be emphasized that hydraulic jacking does not necessarily occur in this zone. In this case, the external rock
mass, beyond a disturbed region, withstand the destructive effects of hydraulic jacking phenomenon. In fact, hydrojacking
happens when the induced stresses over a large surface area exceed the initial stresses normal to this surface area. it some-
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 79

120

100

80

60

40

20
y (m)

0
0 20 40 60 80 100

-20

-40

-60

-80

-100

x (m)
Fig. 12. Contour lines of pore pressure for the pressure tunnel above water table with the cracked reinforced concrete lining under internal water head
Hi = 100 m, ( γpww = 10 m ).

times happens that hydraulic fracturing of intact rock blocks is involved [19,38]. It should be noted that, hydrofracturing
is a very usual event in pressure tunnels and produces fractures in a sound rock due to negative effective stresses, while
hydrojacking is the opening of the existing cracks or joints over a wide surface area due to the applied water pressure and
is a very destructive phenomenon [38]. In the case of the present problem, It is suggested that hydrojacking tests [39] be
conducted for the potential hydrojacking zone. On the other hand, for controlling the hydraulic fracturing phenomena in
the vicinity of the tunnel, the final effective stress (sum of the induced and excavation stresses) should be obtained. To
overcome the hydraulic fracturing phenomenon grouting in the problematic zone is suggested.
It is observed that avoiding cracks by prestressing is an appropriate design option for reducing the water losses, and
increasing the stability of pressure tunnels.
80 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

10 FS (x-axis, Hi=350 m)

9 FS (y-axis, Hi=350)

8 FS (x-axis, Hi=200 m)

7 FS (y-axis, Hi=200)

6
FS

0
4 9 14 19 24 29 34 39 44 49
r (m)
Fig. 13. Distribution of factor of safety against hydrojacking of planes parallel to the valley surfaces along the x and y-axes for the pressure tunnel above
water table with the cracked reinforced concrete lining.

9. Comparison with other numerical and analytical solutions

Here, a comparison of the results obtained by the proposed solution with results obtained by other numerical and ana-
lytical solutions for two different cases of pressure tunnel is presented, briefly.

9.1. Simanjuntak et al.’s numerical study [5]

Simanjuntak et al. [5] used finite element model (FEM) analysis to study the response of a circular prestressed concrete-
lined pressure tunnel above water table in a deep homogeneous isotropic rock mass. His study covered the modeling of the
tunnel excavation, installation of support systems, pre-stressing of concrete lining and the loading of internal water pressure.
However, in this section, it is only concentrated on stresses induced by the prestressing and seepage loads. For comparing
the results obtained by the proposed solution and Simanjuntak et al.’s numerical solution, a data used by Simanjuntak et al.
[5] is taken as input data. The essential data for the proposed solution are:
ri = 1.6 m, ro = 2 m, rg = 3 m, Er = Eg = 5.5 GPa, Ec = 30 GPa, νr = νg = νc = 0.25,
kr = 10−6 ms , kg = 8 ∗ 10−8 ms , kc = 10−8 ms , βr = βg = βc = 1, pi = 1.4 MPa, σ(PR
ro ) = 1.0 MPa

where subscript g refers to quantities corresponding to the grouted zone.


Here, there is an annular grouted zone with different parameters around the tunnel, that should be considered. In this
regard, the components of induced stresses and strains corresponding to the radial seepage forces and the induced bound-
ary pressures in this annular region with internal and external radii ro and rg are presented in Appendix A. On the other
hand, in hydraulic analysis, the grouted zone can be considered as a lining-like zone. The continuity of radial stresses, radial
displacements and pore pressures should be satisfied at the boundaries between the lining-the grouted zone, the grouted
zone- the zone of influence and the zone of influence- the no-flow zone. Using the continuity of flow, the equivalent radius
of influence and radius of influence through the tunnel crown are obtained as, Rs = 14.8 mandRsy = 6.2 m, respectively. In this
way, the boundary pore pressures at radii ro and rg are calculated as pw(ro ) = 0.42MPaand pw(rg ) = 0.07MPa, respectively. On
the other hand, the calculated induced boundary stresses at radii ro , rg and Rs are σr(r ) = 0.156 MPa, σr(r ) = 0.23 MPa,
o g
σr(Rs ) = 0.0018 MPa, respectively. Furthermore, the induced circumferential stress at the inner surface of the lining is calcu-
lated as, σθ (r ) = 1.62 MPa.
i
From the numerical analysis [5], the predicted seepage pressures at the internal and external surfaces of the grouted
zone were found to be pw(ro ) = 0.433MPa and pw(rg ) = 0.82MPaand, respectively. When computed induced circumferential
stress at the inner surface of the lining was, σθ (r ) = 1.59 MPa. It is observed that, results obtained using numerical model
i
is in good agreement with the proposed analytical solution.
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 81

120

100

80

60

40

20
y (m)

0
0 20 40 60 80 100

-20

-40

-60

-80

-100

x (m)
Fig. 14. Contour lines of pore pressure for the pressure tunnel above water table with the prestressed concrete lining under internal water head Hi = 350 m,
( γpww = 15 m).

9.2. Schleiss’s analytical solution [20]

Schleiss [20] presented an analytical method for reinforced concrete lined pressure tunnels in elastic homogeneous
isotropic rock masses. In Schleiss’s method [20], the development of cracks in reinforced-concrete linings was taken into
account. In this section, the results obtained by the proposed solution and Schleiss’s analytical solution [20] are compared
for an above water table pressure tunnel with the following parameters (data is taken from [20]):

ri = 1.8 m, rs = 1.9 m, ro = 2.1 m, Er = 4 GPa, Ec = 20 GPa, νr = νc = 0.2, kr = 10−6 m


s
kc = 10−8 ms , βr = βg = βc = 1, ρs = 0.0052, ds = 20 mm, sb = 20 cm, Es = 200 GPa
In Figs. 16 and 17, the width of cracks and the maximum reinforcement stresses in the cracked section obtained by
two methods with increasing the internal water pressure are shown. The differences observed in the result are because
82 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

FS (x-axis, hi=350 m)

10

9
8

6
FS (x-axis, hi=350 m)
FS

4
3

0
4 9 14 19
r (m)
Fig. 15. Distribution of factor of safety against hydrojacking of planes parallel to the valley surfaces along x-axis for the pressure tunnel with prestressed
concrete lining, while in y direction radial stresses are compressive (σ  r(y) > 0).

Fig. 16. Width of cracks as a function of internal pressure.

these solutions have been derived under different assumptions. In this regard, Schleiss [20] utilized different relationships
for width and spacing of cracks. On the other hand, Schleiss [20] assumed that a new secondary crack is formed between
two existing primary cracks when the maximum concrete tensile stress between the cracks reaches the tensile strength
of concrete. Schleiss [20] considered same properties for both types of cracks. However, the secondary cracks may have
different characteristics and performances [40].
The primary cracks form when stresses reach the tensile strength of the concrete and are wider and visible (can be
observed at the surface of the concrete); while, the other type develops at higher loadings and does not progress to the
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 83

Fig. 17. Stress in the steel bars as a function of internal pressure.

concrete surface [40]. In this regard, in contrast to primary cracks, the secondary cracks may not be flow paths and may
not change the permeability of the concrete lining, significantly. However, the secondary cracks reduce the width of primary
cracks as shown by Schleiss [20]. In the proposed solution, the effect of secondary cracks are neglected; thus, it gives results
on the safe side.

10. Conclusions

In this paper, based on a generalized effective stress principle, an analytical solution was introduced, for approximation
of induced stresses and strains in the lining and the surrounding rock mass of circular pressure tunnels. Elastic responses
of the lining and the rock mass are assumed. The solution accounts for the hydraulic and mechanical interactions between
the lining and the surrounding rock mass, as a result of seepage forces. It is relatively simple, easy to use, and can readily
indicate the sensitivity of the chosen solution through a range of possible ground and lining parameters. The proposed
method covers various types of pressure tunnels, such as unlined, shotcrete lined, concrete lined, reinforced concrete lined
and prestressesd concrete lined pressure tunnels. For unlined and uncracked lined pressure tunnels, the hydraulic analysis
is independent of mechanical responses; thus, closed-form expressions for stresses, strains and pore pressure are derived.
On the other hand, in cracked reinforced concrete linings, the development of cracks are considered in the mechani-
cal and hydraulic responses. In this case, the hydraulic and mechanical analyses are coupled; therefore, the hydraulic and
mechanical analyses are carried out by an iterative procedure. Based on the proposed method traditional design criteria in
design of pressure tunnels (stress and crack width limits in the lining, limiting water losses and adequate confinement) can
be controlled.
To demonstrate the performance of the proposed method, practical examples is given for evaluation of different linings
for a pressure tunnel.
It should be notified that, the proposed solution relies on several simplifying assumptions, the most important of which
are the circular geometry of the deep tunnel and uniform pressures at the boundaries of the lining and the rock mass.
However, if the geometry of the tunnel is non-circular, boundary pressures are non-uniform or the tunnel is shallow, the
problem is not axial symmetric and bending moments will be induced in the lining. For situations in which the assump-
tions mentioned above are violated, the proposed analytical method is still useful and the obtained results are still ac-
ceptable for the primary design and checking the final design of pressure tunnels. The comprehensive design of the lining
which would probably require considering the distribution of bending moments and the local loads induced by the non-
uniform boundary pressures can be made on the basis of more rigorous numerical analyses of the rock-lining interaction
problem.
84 M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85

Appendix A. stresses and strains in an annular region with radial flow pattern

A.1. Induced stresses and strains corresponding to the induced pore pressure

An annular region with internal radius ra (in which the mechanical and hydraulic pressures σ  ra and pa are applied,
respectively) and external radius rb (in which the mechanical and hydraulic pressures σ  rb and pb are applied, respectively)
p −p
is considered. Between ra and rb , the radial seepage forces β b rba are applied (Fig A.1).
r log( r )
a
The governing differential Eq. (11) is a linear equation. Thus, the principle of superposition can be applied to solve this
equation.
In this way, the induced stresses and strains in the annular region can be divided to the following components:

a. Components corresponding to the induced pore pressures.


b. Components corresponding to the induced boundary pressures.

In this paper, the superscript BU is used for the quantities corresponding to the boundary pressures and superscript SE
is used for the quantities corresponding to the seepage forces.
Solving differential Eq. (11) and eliminating the integration constants by using boundary conditions σr(SE ra )
= ( 1 − β ) pa
and σr(SE
r )
= ( 1 − β ) p b , σ BU = σ  and σ BU = σ  gives the induced strains corresponding to the seepage forces and bound-
r (r ) ra r (r ) rb
b a b
ary pressures as follows:
α2
εθSE(r ) = α1 + + α3 (2 log (r ) + 1 ) (A1)
r2

1
εrSE(r ) = α1 − α2 + α3 (2 log (r ) − 1 ) (A2)
r2

 1−2ν β p −p r2 −r2 
( ) ( b a )( b a )
( 1 − 2ν ) ( 1 + ν )  rb  + 2β ( pb − pa )rb2
α1 =   log r
a  (A3)
4E (1 − ν ) rb2 − ra2 +4(1 + ν ) pb rb2 − pa ra2

rb2 ra2 (1 + ν )(−3 + 2ν )β ( pb − pa )


α2 =   (A4)
2E (1 − ν ) rb2 − ra2

(1 − 2ν )(1 + ν )β ( pb − pa )
α3 = −   (A5)
4E (1 − ν ) log rrab

 
1+ν   r2 rb2 1+ν 
εθBU(r ) = σ rb − σ  ra 2 a 2 ( 1 − 2ν ) + + σr ( r b ) ( 1 − 2 ν ) (A6)
E rb − ra r2 E

 
1+ν   r2 rb2 1+ν 
ε BU
r (r ) = σ rb − σ  ra 2 a 2 (1 − 2ν )r − + σrb (1 − 2ν ), (A7)
E rb − ra r2 E

where β , E and ν are Biot constant, elasticity modulus and Poisson’s ratio of the annular region, respectively.

Annular region
β , E ,ν

ra rb
pa pb

σ ra ′
σ rb

Fig. A.1. Annular region under radial seepage forces and boundary stresses.
M.R. Zareifard / Applied Mathematical Modelling 62 (2018) 62–85 85

References

[1] Pan, J.Z., Static calculation of water load acting on tunnels, J. Hydraul. Eng. 5 (1981) 73–79.
[2] P. Kumar, B Singh, Design of reinforced concrete lining in pressure tunnels, considering thermal effects and jointed rock mass, Tunnel. Undergr. Space
Technol. 5 (1–2) (1990) 91–101.
[3] B. Kang, X. Ming, C Juntao, Study on coupled seepage and stress fields in the concrete lining of the underground pipe with high water pressure,
Tunnel. Undergr. Space Technol. 24 (3) (2009) 287–295.
[4] Olumide, B.A., Marence, M.A, Design of pressure hydropower plant Ermenek. Int. J. Eng. Tech. 2(4), 2005.
[5] D.Y.F. Simanjuntak, M. Marence, A. Schleiss, A.E. Mynett, Design of pressure tunnels using a finite element model, The Int. J. Hydropower Dams 19 (5)
(2012) 98–105.
[6] E.T. Brown, J.W. Bray, Rock-lining interaction calculations for pressure shafts and tunnels, in: Proceedings of the ISRM Symposium, Aachen, 1982,
pp. 26–28.
[7] I.M. Lee, S.W Nam, The study of seepage forces acting on the tunnel lining and tunnel face in shallow tunnels, Tunnel. Undergr. Space Technol. 16
(2001) 31–40.
[8] S.W. Lee, J.W. Jung, S.W. Nam, I.M Lee, The influence of seepage forces on ground reaction curve of circular opening, Tunnel. Undergr. Space Technol.
22 (1) (2006) 28–38.
[9] C. Carranza-Torres, J Zhao, Analytical and numerical study of the effect of water pressure on the mechanical response of cylindrical lined tunnels in
elastic and elasto-plastic porous media, Rock Mech. Rock Eng. 46 (2009) 531–547.
[10] A. Bobet, Characteristic curves for deep circular tunnels in poroplastic rock, Rock Mech. Rock Eng. 43 (2010) 185–200.
[11] Y.J. Shin, B.M. Kim, J.H. Shin, I.M. Lee, The ground reaction curve of underwater tunnels considering seepage forces, Tunnel. Undergr. Space Technol.
25 (2010) 315–324.
[12] Y.J. Shin, K.L Song, I.M Lee, G.C Cho, Interaction between tunnel supports and ground convergence - Consideration of seepage forces, Int. J. Rock Mech.
Min. Sci. 48 (2011) 394–405.
[13] A. Fahimifar, M.R Zareifard, A theoretical solution for analysis of tunnels below groundwater considering the hydraulic– mechanical coupling, Tunnel.
Undergr. Space Technol. 24 (2009) 634–646.
[14] Fahimifar, A. and Zareifard, M. R. A new elasto-plastic solution for analysis of underwater tunnels considering strain dependent permeability. Struct.
Infrastruct. Eng. Maint. Manag. Life Cycle Des. Perform., 10(11), 2014
[15] M.R. Zareifard, A. Fahimifar, Effect of seepage forces on circular openings excavated in Hoek–Brown rock mass based on a generalised effective stress
principle, Eur. J. Environ. Civ. Eng. 18 (5) (2014).
[16] M.R. Zareifard, A Fahimifar, Elastic–brittle–plastic analysis of circular deep underwater cavities in a Mohr–Coulomb rock mass considering seepage
forces, Int. J. Geomech. 15 (5) (2015).
[17] M. Bouvard, N Pinto, Aménagement Capivari–Cachoeira. Étude du Puits En Charge, 7, La Houille Blanche, Paris, France, 1969, pp. 747–760.
[18] G. Seeber, Power conduits for high-head plants, Water Power Dam Const. (1985).
[19] A.J. Schleiss, Design of pervious pressure tunnels, Int. Water Power Dam Const. 38 (5) (1986) 21–26.
[20] A.J Schleiss, Design of reinforced concrete linings of pressure tunnels and shafts, Hydro. Dams 3 (1997) 88–94.
[21] G. Fernandez, T.A Alvarez, Seepage-induced effective stresses and water pressures around pressure tunnels, J. Geotech. Eng. 120 (1) (1994) 108–128.
[22] A. Bobet, S.W. Nam, Stresses around Pressure Tunnels with Semi-porous Linings, Rock Mech. Rock Eng. 40 (3) (2006) 287–315.
[23] A. Fahimifar, M.R. Zareifard, A new closed-form solution for analysis of unlined pressure tunnels under seepage forces, Int. J. Numer. Anal. Meth.
Geomech. 37 (11) (2013) 1591–1613.
[24] M.R. Zareifard, M.R. Fahimifar, A simplified solution for stresses around lined pressure tunnels considering non-radial symmetrical seepage flow, KSCE
J. Civil Eng. 20 (7) (2016) 2640–2654 2016.
[25] K. Terzaghi, Die berechnung der durchlassigkeitsziffer des tones aus dem verlauf der hydrodynamischen Spannungserscheinungen, Sitznugshrichte,
Akad. Wissen. Wien of Mathem. Naturw. Kl. 132 (1923) 125–138.
[26] M.A. Biot, General theory of three-dimensional consolidation, J. Appl. Phys. 12 (1941) 155–164.
[27] A.W. Skempton, Effective stress in soils, concrete and rock, Pore Pressure and Suction in Soils, Butterworth, London, 1961, pp. 4–16.
[28] A. Nur, J.D. Byerlee, An exact effective stress law for elastic deformation of rocks with fluids, J. Geophys. Res. 76 (1971) 6414–6419.
[29] D. Kolymbas, P. Wagner, Groundwater ingress to tunnels- the exact analytical solution, Tunnel. Undergr. Space Technol (2006) 1–5.
[30] G Fernandez, Behaviour of pressure tunnels and guidelines for liner design, J. Geotech. Eng. 120 (1994).
[31] Timoshenko S.P., Goodier JN.,. Theory of Elasticity, McGraw-Hill, New York,
[32] W.H. Press, S.A. Teukolsky, W.T. Vetterling, B.P. Flannery, Numerical Recipes in Fortran, Cambridge University Press, 1992 1982.
[33] S.H. Chowdhury, Y.C. Loo, A new formula for prediction of crack widths in reinforced and partially prestressed concrete beams, Adv. Struct. Eng. 4 (2)
(2001) 101–109.
[34] CEB/FIP, CEB/FIP model code 1990, CEB Bulletin No. 213/214, Lausanne, Switzerland, 1993.
[35] A. Castel, T. Vidal, R. Francois, Effective tension active cross-section of reinforced concrete beams after cracking, Mater. Struct. 39 (2006) 115–126.
[36] M.P. Thurnherr, I.F. Uherkovich, Prestressed concrete pressure tunnels, Water Power Dam Construct. 12 (1941) 155–164.
[37] R.P. Benson, Design of unlined and lined pressure tunnels, Tunnel. Undergr. Space Technol. 4 (2) (1989) 155–170.
[38] D.U. Deere, G Lombardi, Lining of Pressure Tunnels and Hydrofracturing potential. Victor de Mello Volume, Editora Edgard Blücher Ltda, São Paulo,
Brasil, 1989, pp. 121–128.
[39] Brekke, T.L., Ripley, B.D., Design guidelines for pressure tunnels and shafts. EPRI, Report AP-5273, 1987.
[40] American Concrete Institute, Cracking of concrete members in direct tension, ACI 224.2R-1992, ACI Committee 224, Detroit, Michigan, 1992.

You might also like