Equivalent Constitutive Model For Jointed Rock Masses and Its Application in Large Underground Caverns

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Equivalent Constitutive Model for Jointed Rock Masses

and Its Application in Large Underground Caverns


Yao Yang1; Quan Jiang2; Jian Liu3; Hong Zheng4; Dingping Xu5; and Jun Xiong6

Abstract: Jointed rock masses with different orientations and spacings are widely exposed in rock engineering applications, such as under-
ground caverns, slopes, and tunnels, which significantly influence the rock engineering stability and mechanical behavior of the rock masses.
To better evaluate the mechanical response of the jointed rock masses during the engineering excavation period, an equivalent constitutive model
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

for the jointed rock masses was established based on the joint compliance tensor (JCT) and was adopted to analyze the excavation-induced
mechanical performance of Jinping II hydropower station underground caverns. First, the calculation method of the joint compliance tensor
and equivalent elastic compliance matrix of the rock masses containing multiple joint sets was established considering the deformation charac-
teristics of the joint system comprehensively based on the Oda fracture tensor theory and linear superposition principle. Second, an equivalent
constitutive model for the jointed rock masses was established considering the joint spacing and connectivity. Finally, combined with the
field investigation and monitoring data, the mechanical mechanism of unloading failure of the jointed rock masses in Jinping II hydropower sta-
tion underground caverns was analyzed using the proposed model and numerical simulation. DOI: 10.1061/(ASCE)GM.1943-5622.0002633.
© 2022 American Society of Civil Engineers.
Author keywords: Jointed rock masses; Compliance tensor; Constitutive model; Numerical simulation; Underground caverns.

Introduction significantly influence the mechanical behavior of the rock masses


and the rock engineering stability but also increase the technical
Rock masses with many discontinuities (Zhou et al. 2003; Hauseux difficulty and cost of engineering construction. Specifically, the an-
et al. 2016), like joints and faults, are widely exposed in rock engi- isotropic deformation and failure of the rock masses are intensified
neering, such as underground caverns (Xu et al. 2021; Jiang et al. under the cutting of dominant joints, which induces engineering ac-
2014), slopes (Aladejare and Wang 2018; Kumar et al. 2019), and cidents such as large asymmetric deformation of the tunnel
tunnels (Kaiser and Cai 2012; Zhao et al. 2018; Rehman et al. (Anagnostou 1993; Hoek and Marinos 2000), cracking (Dyskin
2019). Engineering practice has shown that the heterogeneity, an- 1999; Aubertin et al. 2000), and collapse of surrounding rock
isotropy, and discontinuity of the jointed rock masses not only (Kim et al. 2021), resulting in casualties, equipment damage, con-
struction delay, and other irreparable losses, for example, the large
1
deformation of the surrounding rock of more than 200 mm
Ph.D. Candidate, State Key Laboratory of Geomechanics and occurred during the excavation of Jinping I hydropower station
Geotechnical Engineering, Institute of Rock and Soil Mechanics, Chinese
(Wei et al. 2010),and the serious collapse occurred during the ex-
Academy of Sciences, Wuhan, Hubei 430071, China; Univ. of Chinese
Academy of Sciences, Beijing 100049, China. Email: yangyao19@mails
cavation of Jinping Underground Laboratory of China (Li et al.
.ucas.ac.cn 2021). Therefore, to evaluate the mechanical response of the
2
Professor, State Key Laboratory of Geomechanics and Geotechnical jointed rock masses better, it is important to present a constitutive
Engineering, Institute of Rock and Soil Mechanics, Chinese Academy of model that can reflect the basic mechanical properties of the jointed
Sciences, Wuhan, Hubei 430071, China (corresponding author). ORCID: rock masses (orientation, spacing), deformation, and strength an-
https://orcid.org/0000-0001-6039-9429. Email: qjiang@whrsm.ac.cn isotropy under different stress states.
3
Ph.D. Candidate, State Key Laboratory of Geomechanics and At present, the numerical analysis methods for the jointed
Geotechnical Engineering, Institute of Rock and Soil Mechanics, Chinese rock masses are generally divided into an equivalent continuous
Academy of Sciences, Wuhan, Hubei 430071, China; Univ. of Chinese medium method and a discrete medium method, such as the finite-
Academy of Sciences, Beijing 100049, China. Email: liujian19@mails
element method (FEM), finite difference method (FDM), boundary
.ucas.ac.cn
4
Associate Professor, State Key Laboratory of Geomechanics and element method (BEM), and discrete element method (DEM).
Geotechnical Engineering, Institute of Rock and Soil Mechanics, Chinese The equivalent continuum method considers the contribution of
Academy of Sciences, Wuhan, Hubei 430071, China. Email: hzheng@ rock and joints to the deformation and strength of the rock mass
whrsm.ac.cn as a whole and includes the deformation of rock and joints in the
5
Professor, State Key Laboratory of Geomechanics and Geotechnical calculation of rock mass deformation (Yoshida and Horii 2004;
Engineering, Institute of Rock and Soil Mechanics, Chinese Academy of Wang and Huang 2009; Pariseau 2012). This method has high
Sciences, Wuhan, Hubei 430071, China. Email: dpxu@whrsm.ac.cn computational efficiency and can deal with practical engineering
6
Researcher, Dept. of Civil and Environmental Engineering, Univ. of scale problems, but it cannot consider local failure under the control
Alberta, Edmonton, AB T6G 1H9. ORCID: https://orcid.org/0000-0002
of rock discontinuities. The discrete element method is a numerical
-4481-1409. Email: jxiong3@ualberta.ca
Note. This manuscript was submitted on January 21, 2022; approved on explicit solution method proposed and can objectively reflect the
August 8, 2022; published online on November 4, 2022. Discussion period structural characteristics of the jointed rock masses and simulate
open until April 4, 2023; separate discussions must be submitted for indi- the local stress and strain in the rock mass by displaying the joint
vidual papers. This paper is part of the International Journal of Geome- system (Cundall 1971). Later, it was continuously improved by
chanics, © ASCE, ISSN 1532-3641. Lemos et al. (1985), Cundall (1988), and Hart et al. (1988) and

© ASCE 04022259-1 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


gradually became a powerful numerical simulation method for joint systems:
stress and strain analysis of jointed rock masses. At the same ∞ 
πρ
time, theoretical research focuses on the prediction of mechanical F= r3 · n ⊗ n ⊗ · · · ⊗ n · E(n, r)dΩ dr (1)
parameters based on the superposition principle (Goodman et al. 4 0 Ω
1968), elastoplastic damage theory (Chen et al. 2010; Yang et al.
where ρ = crack density, which is defined as m (V )/V; m (V ) = num-
2019), and displacement discontinuity theory (Shen et al. 2016; ber of cracks with center points in a volume V; r = equivalent di-
Do and Wu 2020). In experimental studies, some researchers ob- ameter of the crack; n = normal vector of the crack; E(n,r) =
tained stress–strain relationships through triaxial compression ex- joint probability density function of the crack’s equivalent diameter
periments and compared them with theoretical results (Nova and normal vector; and Ω = whole solid angle (4π) equivalent to a
1980; Tien and Kuo 2001; Singh et al. 2002; Tien et al. 2006; unit sphere.
Chen et al. 2012; Chang et al. 2019). For numerical analysis, If the information of each crack is known, Eq. (1) can be ex-
some scholars have studied the influence of distribution parameters pressed as
(orientation and spacing) and mechanical parameters on the failure
π 
(V )
mode, plastic zone distribution, deformation, and internal force and m
3
failure type of underground caverns support structures (Park and F= (r(k) ) · n(k) ⊗ n(k) ⊗ · · · ⊗ n(k) (2)
4V k=1
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

Adachi 2002; Wang and Huang 2014; Yang et al. 2019; Zhou
et al. 2021). Moreover, many scholars have proposed or improved where the superscript k represents the sequence number of the
the constitutive model of the jointed rock masses (Adhikary and crack.
Dyskin 1998; Sitharam et al. 2001; Wang and Huang 2009; Sains- Considering the combined joint deformation characteristics with
bury and Sainsbury 2017; Zhou et al. 2017; Xu et al. 2017; Das the crack tensor in the form of the deformation stiffness coefficient,
et al. 2019). In addition, other studies have shown that jointed the JCT, which comprehensively considers the geometric charac-
rock masses are prone to induce engineering failure accidents dur- teristics of the joint spatial distribution and deformation properties,
ing the excavation (Chen et al. 2012; Shen et al. 2016; Chang et al. is established, as follows:
2019; Do and Wu 2020). Some researchers have studied the me-  ∞ ∞
chanical behavior of more complex rock masses with multiple πρ 1
JCT = r2 · · n ⊗ n · E(n, r)dr dJF dΩ (3)
joint sets (Chen and Egger 1999; Chen et al. 2007, 2015). Also, 4 Ω 0 0 F
with an increase in the complexity of joint sets, engineering disas-
ters caused by excavation will become more severe. Therefore, where JF = deformation stiffness coefficient of the joint defined by
considering multiple joint sets in the constitutive model is more Eq. (4); and E(n,r,JF) = joint probability density function of the
reasonable to the practical characteristics of the rock mass and crack’s equivalent diameter, azimuth vector, and stiffness
has more significant practical value for the excavation and support coefficient.
design of the jointed rock masses. Kn Ks
In this paper, the joint compliance tensor (JCT) was proposed JF = = (4)
Kn0 Ks0
based on the Oda fracture tensor theory, which comprehensively
considers the deformation characteristics of the joint system and re- where Kn0 and Ks0 = fiducial normal stiffness and shear stiffness,
alizes the single-parameter characterization of joint system space respectively, with the assumption of equal stiffness ratios.
geometry and deformation characteristics. An equivalent constitu- If the geometry and deformation information of each joint in the
tive model was established based on the strain superposition prin- investigation area are known, the JCT can be discretized as follows:
ciple and considered the joint spacing and connectivity index under
the framework of the equivalent continuity method. Based on the 
m(V )
1 1
JCT = · s(k) · · n(k) ⊗ n(k) (5)
failure characteristics of surrounding rock in Jinping II hydropower k=1
V JF
station’s underground caverns during the excavation, combined
with the field investigation and monitoring data, the mechanical where superscript k is the joint sequence number; JF (k) and n (k) =
mechanism of unloading failure of the jointed rock masses was dis- stiffness coefficient and azimuth vector of the kth joint, respec-
cussed based on the proposed model and numerical simulation. tively; and s (k) = area of the kth joint, which is equal to π(r (k))2/4.
Joints are the products of tectonic movement in geological his-
tory. The joints formed in a tectonic process are generally regular
and produced in groups. A combination form is a joint set and sys-
Joint Compliance Tensor tem. The joint set is composed of joints with the similar occurrence
and mechanical properties, whose typical fabric is generally char-
Definition of the Joint Compliance Tensor acterized by joint spacing, connectivity, and orientation. Assuming
that the joints in the research region can be approximately ex-
The spatial deformation anisotropy of the jointed rock masses is pressed by several joint sets, the joint compliance tensor can be ex-
mainly caused by the cutting of the dominant joints, so the quanti- pressed as follows:
tative description of the fabric characteristics of the spatial distribu-
tion for the jointed rock masses is significant in evaluating the 
q
Ni 1
spatial anisotropy deformation characteristics of the jointed rock JCT = · si · · n(i) ⊗ n(i) (6)
i=1
V JF i
masses. Oda (1982) proposed the concept of the fracture tensor,
which combines the tensor product of joint scale, density, and az- where q and i = total number and serial number of the dominant
imuth vector to characterize the geometric fabric characteristics joint sets, respectively; and Ni, si, JFi, and ni = number, average
of the spatial distribution of joint system, as shown in Eq. (1). area, stiffness coefficient, and azimuth vector of joint set i,
This method can be used to characterize the spatial geometric fabric respectively.
characteristics of fractures by a single parameter, which provides a Considering the cylindrical investigation window in Fig. 1,
theoretical basis to evaluate the spatial deformation properties of L and D are the length and diameter of the investigation window,

© ASCE 04022259-2 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


Fig. 1. Joints and statistical parameters in the investigation window.
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

respectively. The volume within the investigation scope can be ex- joint spacing and linear density, Eq. (11) can be simplified as
pressed as follows follows:
πD2 L
V= (7) q
1 1
4 JCT = i
· i · n(i) ⊗ n(i) (12)
i=1
S m JF
The average area of joint set i can be expressed as follows:

π(d i )
2
where Smi = spacing of joint group i.
si = (8)
4
Substituting Eqs. (7) and (8) into Eq. (6), JCT can be expressed Equivalent Elastic Compliance Matrix of Jointed Rock
as follows: Masses

q
Ni π(d i ) 1
2 Based on the statistical data of the joint attitude in the investigation
JCT = · · i · n(i) ⊗ n(i) (9) area, the JCT can be obtained first, and the eigenvalues and eigen-
πD2 L/4 4 JF
i=1
vectors of the JCT can then be calculated. Then, JCT can be ex-
Suppose that the included angle between the normal vector of the pressed as the local compliance matrix (JCT′ ) of the main
joint plane and the survey line isα i, which can be defined as characteristic direction. Based on the deformation superposition
follows: of rock and joints, the global compliance matrix (D′ ) of the rock
mass in the local coordinate system can be deduced (Amadei and
πD2 /4 Goodman 1981), as shown in Eq. (13). T1, T2, and T3 are eigenval-
cosαi = (10)
π(d i )2 /4 ues of the joint compliance tensor, and D′ R is the compliance ma-
trix of the rock. It can be seen from Eq. (13) that the joint system
Therefore, Eq. (9) can be expressed as follows: weakens the stiffness of intact rock and this equation leads to a

q
Ni 1 lower-bound estimate of the equivalent stiffness:
JCT = · · n(i) ⊗ n(i) (11)
L · cosαi JF i
i=1 D′ = D′ R + JCT (13)
In Eq. (11), the first term is the linear density of joint set i in the
investigation window. Therefore, by using the relationship between where

⎡1
T1 ν ν ⎤
+ − − 0 0 0
⎢ E Kn0 E E ⎥
⎢ ⎥
⎢ ν 1 T2 ν ⎥
⎢ − + − 0 0 0 ⎥
⎢ E E Kn0 E ⎥
⎢ ⎥
⎢ ν ν 1 T3 ⎥
⎢ − − + 0 0 0 ⎥
⎢ E E E Kn0 ⎥
⎢   ⎥
D′ = ⎢ 1 1 T2 + T3 ⎥
⎢ + ⎥
⎢ 0 0 0 0 0 ⎥
⎢ 2 G Ks0
  ⎥
⎢ ⎥
⎢ 1 1 T1 + T3 ⎥
⎢ 0 0 0 0 + 0 ⎥
⎢ 2 G Ks0 ⎥
⎢  ⎥
⎣ 1 1 T1 + T2 ⎦
0 0 0 0 0 +
2 G Ks0

© ASCE 04022259-3 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


The local coordinate system constituted by eigenvectors of JCT denoted l3, m3, and n3, respectively; and the corresponding relation-
is oxyz, and the global coordinate system is OXYZ; the stress and ship between the eigenvectors v1, v2, and v3 and the corresponding
strain tensors corresponding to the OXYZ coordinate system are de- strain components e1, e2, and e3 of the global coordinate system
noted σ and ɛ, respectively; the stress and strain tensors corre- should be determined according to the corresponding relationship
sponding to the oxyz coordinate system are denoted σ′ and ɛ′ , between the eigenvalues of the eigenvectors in the equivalent coor-
respectively; the direction cosines of the X-axis and x-, y-, and dinate system. According to the direction cosine calculation, the
z-axes are denoted l1, m1, and n1, respectively; the direction cosines transformation matrix of the stress and strain tensors in the local co-
of the Y-axis and x-, y-, and z-axes are denoted l2, m2 and n2, respec- ordinate system into the stress and strain tensors in the global coor-
tively; the direction cosines of the Z-axis and x-, y-, and z-axes are dinate system can be obtained, as shown in Eq. (14):

⎡ ⎤
l12 m21 n21 2m1 n1 2n1 l1 2l1 m1
⎢ l2 m22 n22 2m2 n2 2n2 l2 2l2 m2 ⎥
⎢ 22 ⎥
⎢ l m23 n23 2m3 n3 2n3 l3 2l3 m3 ⎥
R=⎢ 3 ⎥
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

⎢ l2 l3 (14)
⎢ m2 m3 n2 n3 m2 n3 + n2 m3 n2 l3 + l2 n3 l2 m3 + m2 l3 ⎥

⎣ l3 l1 m3 m1 n3 n1 m3 n1 + n3 m1 n3 l1 + l3 n1 l3 m1 + m3 l1 ⎦
l1 l2 m1 m2 n1 n2 m1 n2 + n1 m2 n1 l2 + l1 n2 l1 m2 + m1 l1

In the Cartesian coordinate system, the transformation law of the Gyz = D−1 −1 −1
44 , Gxz = D55 , Gxy = D66 (22)
strain tensor and stress tensor is consistent, as shown follows:
σ = Rσ ′ (15)
νyz = −D32 · Ey , νxz = −D31 · Ex , νxy = −D21 · Ex (23)
ε = Rε′ (16)
The equivalent elastic compliance matrix is used to represent the
stress–strain relationship, and in the local coordinate system and
the global coordinate system, Eqs. (17) and (18) are
Equivalent Constitutive Model for Jointed Rock
ε′ = D ′ σ ′ (17) Masses

ε = Dσ (18) Equivalent Elastic Compliance Tensor of the Jointed Rock


Therefore, the equivalent elastic compliance matrix of the jointed Masses
rock masses in the global coordinate system is shown as follows: To fully reflect the adverse effects of the dominant joint sets and
′ −1 reasonably simplify the effects of random and shorter fractures
D = RD R (19)
for the rock mass with well-developed joints, these randomly
According to the corresponding relationship between the equiv- shorter fractures and rock are regarded as equivalent rock to-
alent elastic compliance matrix of spatial deformation of the jointed gether, and the mechanical parameters of the jointed rock masses
rock masses and the equivalent elastic deformation parameters of are affected by both joints and rock, as shown in Fig. 2. There-
the rock mass, as shown in Eq. (20) (Lekhnitskii et al. 1964), equiv- fore, to quantify the spatial deformation characteristics of the
alent elastic deformation parameters of the rock mass can be jointed rock masses, a calculation method of the equivalent elas-
obtained to represent the antideformation ability of the rock tic compliance matrix of the jointed rock masses based on the
mass. The equivalent deformation parameters are expressed in JCT was established. In order to use the strain superposition
Eqs. (21)–(23): principle, the following assumptions are made for simplification:
⎡ ⎤ (1) the deformation of rock and joints satisfies Hook’s law; (2)
1 νyx νzx ηx,yz ηx,xz ηx,xy the coupling effect of rock–joints and the interaction among
⎢ Ex − −
⎢ Ey Ez Gyz Gxz Gxy ⎥ ⎥ joints are not considered; and (3) the occurred deformation of
⎢ νxy 1 ν η η η ⎥
⎢− −
zy y,yz y,xz y,xy ⎥ the jointed rock mass is small.
⎢ E ⎥
⎢ x E y E z G yz G xz G xy ⎥ The strain of the jointed rock masses can be composed of equiv-
⎢ ηz,yz ηz,xz ηz,xy ⎥
⎢ νxz νyz 1 ⎥ alent rock’s strain and joint sets’ strain based on the strain superpo-
⎢− − ⎥ sition principle. At the same time, as long as the equivalent
⎢ Ex E E G G G ⎥
D=⎢ ⎥
y z yz xz xy
⎢ ηyz,x ηyz,y ηyz,z μyz,xz μyz,xy ⎥ (20) compliance matrix of rock mass with multiple joint sets is charac-
⎢ 1 ⎥
⎢ E terized, the stress increment can be solved by the deformation in-
⎢ x Ey Ez Gyz Gxz Gxy ⎥ ⎥ crement. The total strain incrementdεof the jointed rock masses
⎢η μxz,xy ⎥
⎢ xz,x ηxz,y ηxz,z μxz,yz 1 ⎥ can be expressed as follows:
⎢ ⎥
⎢ Ex Ey Ez Gyz Gxz Gxy ⎥
⎢ ⎥
⎣ ηxy,x ηxy,y ηxy,z μxy,yz μxy,xz 1 ⎦ dε = dεR + dεJ (24)
Ex Ey Ez Gyz Gxz Gxy
where dεR = strain increment of equivalent rock; and dεJ = strain
Ex = D−1 −1 −1
11 , Ey = D22 , Ez = D33 (21) increment of joint sets.

© ASCE 04022259-4 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Rock mass deformation superposition of the multiple joint set.

The strain increment dεR of equivalent rock can be expressed as


follows:

dεR = DR dσ (25)

where DR = elastic compliance matrix of equivalent rock.


The strain increment dεJ of the joint sets can be expressed as
follows:

dεJ = DJ dσ (26)

where DJ = elastic compliance matrix of the joint sets.


Then, Eq. (24) can be further expressed as follows:

dε = dεR + dεJ = DR dσ + DJ dσ = (DR + DJ )dσ = DEQ dσ (27)

where DEQ = DR + DJ = equivalent elastic compliance matrix of


the jointed rock masses. Therefore, the update of the stress incre-
ment can be completed as long as the DEQ equivalent elastic com-
pliance matrix is obtained.
1. Elastic compliance matrix of equivalent rock DR
The equivalent rock is simplified as an isotropic material, and
DR can be expressed as
⎡ ⎤ Fig. 3. Relation between the global and local coordinate system of the
1 ν ν
− − 0 0 0 joint set.
⎢ E E E ⎥
⎢ ⎥
⎢ ν 1 ν ⎥
⎢− − 0 0 0 ⎥
⎢ E E E ⎥
⎢ ⎥
⎢ ν ν 1 ⎥ I I
⎢− − 0 0 0 ⎥ dN J = T J dσ (29)
R ⎢ E E E ⎥
D =⎢ ⎥ (28)
⎢ 1 ⎥ I
where T J = projection of joint set I on the X–Y plane, X–Z plane,
⎢ 0 0 0 0 0 ⎥
⎢ 2G ⎥ and Y–Z plane of the global coordinate system.
⎢ ⎥
⎢ 1 ⎥ As shown in Fig. 3, φ is the direction of the joint plane, γ is the
⎢ 0 0 0 0 0 ⎥
⎢ 2G ⎥ dip angle of the joint plane, the basis vector of the local coordinate
⎣ 1 ⎦
0 0 0 0 0 system of the joint plane is (v1 , v2 , v3 ), and the basis vector for the
2G global coordinate system XYZ is (e1 , e2 , e3 ).
In the local coordinate system, the basis vector of the joint plane
where E, ν, andG = elastic modulus, Poisson’s ratio, and shear
can be expressed as
modulus, respectively.
2. Elastic compliance matrix of the joint sets DJ v1 = sinφcosγe1 + cosφcosγe2 − sinγe3
For the jointed rock masses per unit volume, when the stress in-
I v2 = −cosφe2 + sinφe3 (30)
creases dσ, the force increment dN J on any joint plane JI can be
expressed as follows: v3 = v1 ⊗ v2 = sinφsinγe1 + cosφsinγe2 + cosγe3

© ASCE 04022259-5 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


Then, the cosine of the angle between (v1 , v2 , v3 ) and (e1 , e2 , e3 ) can be simplified as

lx = cos(v3 , e1 ) = sinφsinγ, ly = cos(v3 , e1 ) = cosφsinγ, lz = cos(v3 , e3 ) = cosγ


mx = cos(v2 , e1 ) = −cosφ, my = cos(v2 , e2 ) = sinφ, mz = cos(v2 , e3 ) = 0 (31)
nx = cos(v1 , e1 ) = sinφcosγ, ny = cos(v1 , e2 ) = cosφcosγ, nz = cos(ν1 , e3 ) = −sinγ
I
Then,T J can be expressed as

⎡ ⎤
cos(v3 , e1 ) 0 0 cos (v3 , e1 ) 0 cos (v3 , e3 )
J ⎢ I ⎥
T =⎣ 0 cos (v3 , e2 ) 0 cos (v3 , e1 ) cos (v3 , e3 ) 0 ⎦
0 0 cos (v3 , e3 ) 0 cos (v3 , e2 ) cos (v3 , e1 )
⎡ ⎤
lx 0 0 ly 0 lz
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

⎢ 0⎥
=⎣ 0 ly 0 lx lz ⎦ (32)
0 0 lz 0 ly lx

I
The stress increment of joint plane dσ J = [σ n , τ s , τ t ]T can be ex- Y-, and Z-directions can be expressed as follows:
pressed as I I
I duJi · PJ
I I I dεJi = , i = x, y, z (39)
dσ J = RJ dN J (33) SJ I
I
where RJ = transformation matrix from the local coordinate sys- The shear deformation can be expressed as
tem of the joint plane to the global coordinate system of the rock I I

mass, which can be expressed as I I


JI
duJi + duJj
dγ Jij = 2dεJij =P · (40)
⎡ ⎤ SJ I
cos (v3 , e1 ) cos (v3 , e2 ) cos (v3 , e3 ) I
I ⎢ ⎥ Then, the deformation of the joint plane dεJ can be expressed as
RJ =⎣ cos (v2 , e1 ) cos (v2 , e2 ) cos (v2 , e3 ) ⎦
I
cos (v1 , e1 ) cos (v1 , e2 ) cos (v1 , e3 ) I PJ I I I I I
⎡ ⎤ dεJ = (RJ T J )T DJ RJ T J dσ (41)
lx ly lz SJ I
⎢ mz ⎥
I I I
=⎣ mx my ⎦ (34) Set W J = RJ T J , and based on the strain superposition princi-
ny ny nz ple, the strain increment of all joint sets can be expressed as

Substitute Eq. (31) into Eq. (34) to form the following equation: 
N
I

N
PJ
I
I I I
dεJ = dεJ = (W J )T DJ (W J )dσ (42)
I=1 I
SJ I
JI JI JI
dσ = R T dσ (35)
where N = number of all joint groups.
Then, the deformation of the joint plane in the local coordinate Therefore, the elastic compliance matrix of all joint sets can be
system can be expressed as expressed as
I I I
dδJ = DJ dσ J (36) 
N
PJ
I
I I I
DJ = (W J )T DJ (W J ) (43)
J I SJ I
where D = elastic compliance matrix of the joint plane. If only the I=1
I
normal and tangential compliance of joints are considered, DJ can
be expressed as follows:
⎡ ⎤ Strength Criterion and Plastic Potential Function for
Dnn 0 0 Jointed Rock Masses
D =⎣ 0
JI
Dss 0 ⎦ (37)
0 0 0 As the Mohr–Coulomb criterion and its parameters are widely ac-
cepted in the field of rock engineering (Hudson et al. 2011; Ulusay
I
then, the joint plane deformation duJ in the global coordinate sys- et al. 2014; Barron and Shen 2018), the Mohr–Coulomb shear yield
tem can be expressed as follows: criterion and tensile strength criterion are adopted for the strength
criterion of the jointed rock masses.
I I I I I I I
duJ = (RJ )T dδJ = (RJ )T DJ RJ T J dσ (38) 1. Equivalent rock
The yield function of the equivalent rock consists of the
For joint group J I, set the joint spacing as S J and the joint connec-
I
shear yield function fms and tensile yield function fmt , expressed
I
tivity as PJ . Under stress increment dσ, the deformation of the joint as Eqs. (44) and (45), respectively:
with unit length in the X-axis direction of the global coordinate sys-
I I I
tem is duJx · PJ · S J ; therefore, the normal deformation in the X-, fms = σ 1 − σ 3 Nϕm + 2cm Nϕm (44)

© ASCE 04022259-6 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


fmt = σ 3 − σ tm (45) gjt = σ 3′ 3′ (53)
where σ1 and σ3 = first principal stress and third principal stress,
respectively; cm and ϕm = cohesion and internal friction angle of where ψj = dilation angle of the joint group.
the equivalent rock, respectively; and Nϕm = (1 + sinϕm)/(1 − As the shear yield occurs to the joint set, in the local coordi-
sinϕm), where σ tm is the tensile strength of the equivalent rock, nate system, the implicit Euler method is used for stress plastic
and its maximum value is σ tm, max = cm / tan ϕm . correction; according to the consistency condition (Itasca
The plastic potential function of the equivalent rock is con- Consulting Group 2011), the corresponding plastic multiplier is
sidered a composite function of shear and tensile strength
(gms , gmt ), and the forms of the shear plastic potential function
j , σ 3′ 3′ )
fjs (τO O
gms and tensile plastic potential function gmt are given as λsj =  (54)
2 (C ′ 266 σ O ′2 O 2 ′
2
1′ 3′ + C 55 σ 2′ 3′ )/τj + C 33 Nψj Nϕj
O
Eqs. (46) and (47), respectively:
gms = σ 1 − σ 3 Nψ m (46)
where C ′ ij = corresponding component of the stiffness matrix of
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

the rock mass under the local coordinate system (Zhou et al.
gmt = σ 3 (47)
2017).
where Nψm = (1 + sinψm)/(1 − sinψm); and ψm = dilation angle The plastic-correction procedure for tension yield is the same
of the equivalent rock. as the UBJ model in FLAC3D, but the elastic constants involved
As the shear yield occurs to the equivalent rock, in the prin- should now beC ′ ij .
cipal stress coordinate system, the implicit Euler method is used
for stress plastic correction. Following the standard procedure of
elastic-guess–plastic-correction implemented in FLAC3D Numerical Program Implementation
(Itasca Consulting Group 2011), the increments of principal Based on the numerical simulation software platform FLACD3D
stresses in one step are as follows: developed by Itasca Consulting Group (2011), the self-defined con-
stitutive model interface provided by the software is used to realize
Δσ 1 = λsm (C11 − C13 Nψ m )
the secondary development of the constitutive mode through the vi-
Δσ 2 = λsm (C21 − C23 Nψ m ) (48) sual studio 2010 software platform and C++ programming lan-
Δσ 3 = λsm (C31 − C33 Nψ m ) guage. Finally, the file was compiled into a dynamic link library
DLL file and embedded into FLACD3D software. The calculation
where Cij = corresponding component of the stiffness matrix of flow of the stress update in the program is shown in Fig. 4.
the rock mass under the principal stress coordinate system
(Zhou et al. 2017); and λsm = plastic multiplier that can be de-
rived using the consistency condition:

1 , σ3 )
fms (σ O O
λsm = (49)
(C11 − C13 Nψ m ) − (C31 − C33 Nψ m )Nϕm
where σ O 1 and σ 3 = maximum and minimum stresses after
O

elastic-guess procedure, respectively.


The plastic-correction procedure for tension yield is the same
as the ubiquitous-joint (UBJ) model in FLAC3D, but the elastic
constants involved should now be Cij.
2. Joint set
The yield function of the joint set consists of a shear yield
function and tensile yield function, which can be expressed as
Eqs. (50) and (51), respectively:
fjs = τj − cj + σ 3′ 3′ tan ϕj (50)

fjt = σ 3′ 3′ − σ tj (51)

where cj and ϕj = cohesion and internal friction angles of the



joint set, respectively; τj = σ 21′ 3′ + σ 223′ ; σ 1′ 3′ , σ 2′ 3′ = shear
stresses along the plane direction;σ 3′ 3′ = tensile stress perpendic-
ular to the plane direction; and σ tj = tensile strength of the joint
set.
The plastic potential function of the joint set is considered a
composite function of shear and tensile strength (gjs , gjt ), and the
forms of the shear plastic potential function gjs and tensile plastic
potential function gjt are expressed as Eqs. (52) and (53),
respectively:
gjs = τj + σ 3′ 3′ tan ψ j (52) Fig. 4. Stress calculation update process.

© ASCE 04022259-7 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


Table 1. Deformation mechanical parameters of the equivalent rock and
joints
Type E (GPa) v cr (MPa) σ tr (MPa) ϕr (°) ψr (°)
Rock 40 0.25 7.2 1 21 15
Note: E = elastic modulus of the equivalent rock; v = Poisson’s ratio of the
equivalent rock; cr = cohesion of the equivalent rock; σ tr = tensile strength
of the equivalent rock; ϕr = internal friction angle of the equivalent rock;
and ψr = dilation angle of the equivalent rock.

Table 2. Deformation mechanical parameters of joints


dd dip Sm Kn Ks cj σ tj ϕj ψj
Type (°) (°) (m) (GPa/m) (GPa/m) (MPa) (MPa) (°) (°)
Fig. 5. River valley landform of the Jinping II hydropower station site.
J-1 300 73 0.1 120 60 2 1 22 18
(Image by Quan Jiang.)
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

J-2 200 80 0.5 100 50 2 1 22 16


J-3 140 30 0.8 72 36 2 1 22 17
Engineering Application Note: dd = dip direction; dip = dip angle; Sm = joint spacing; Kn = normal
stiffness of the joint set; Ks = shear stiffness of the joint set; cj = cohesion
of the joint set; σ tj = tensile strength of the joint set; ϕj = internal friction
Background of the Jinping II Hydropower Project angle of the joint set; and ψj = dilation angle of the joint set.
The Jinping II hydropower station is located on Jinping Dahe Bay
of the Yalong River at the junction of Muli, Yanyuan, and Mian-
ning counties of Liangshan Yi Autonomous Prefecture in Sichuan
Province. The Jinping II hydropower station will take advantage of
the natural drop of the Yalong River’s 150 km river bend, cut and
straighten a 17 km diversion tunnel, and obtain a water head of
310 m. The total installed capacity of the power station is
4,800 MW, the single capacity is 600 MW, the rated head is
288 m, the average annual power generation is 24.23 billion
kW·h, the guaranteed output is 1,972 MW, and the annual utiliza-
tion hour is 5,048 h. It is the largest hydropower station with the
highest head and installed capacity on the Yalong River and an im-
portant hydropower station in the cascade development of the Ya-
Fig. 6. Rose diagram of dominant joints in underground caverns.
long River. The area of the power station site has magnificent
mountains and large terrain ups and downs, with an elevation of
1,330–1,600 m and a slope of 50°–70°, forming a series of steep
walls along the river (Fig. 5). The slope from 1,600 m to approxi-
mately 1,700 m is relatively slow, with a slope of 35°–45°. The lon-
gitudinal axis of the main powerhouse is N35°E, and the total
length of the main powerhouse is 352.44 m. The total height of
the main powerhouse is 72.20 m. The length and width of the
main transformer tunnel are 374.60 and 19.8 m, respectively. Mul-
tiple joint sets are developed in the underground caverns of the
Jinping II hydropower station. The traditional equivalent contin-
uum model, discrete element method, and the layered rock mass
mechanical model cannot accurately characterize the mechanical
Fig. 7. Three-dimensional numerical model.
characteristics of the jointed rock masses. Therefore, the estab-
lished degradation constitutive model of the jointed rock mass
was adopted for the mechanical mechanism of unloading fracture x-direction and 840 wide in the y-direction. The numerical model
of the jointed rock mass in the Jinping II hydropower station under- contains a total of 1,168,000 elements and 235,000 nodes
ground caverns. According to the field investigation statistics and (Fig. 7). The initial geostress condition of the Jinping II hydro-
geological data, the deformation mechanical parameters of equiva- power station site area is obtained by the nonlinear inversion
lent rock and joints are given in Tables 1 and 2, and the structural method (Jiang et al. 2008). In the hub elevation area, the maximum
plane development of the underground caverns is shown in Fig. 6. principal stress with 30°–55° dip angle and S50–75°E azimuth angle
is −14–17 MPa; the intermediate principal stress with the 20°–30°
dip angle and the N20–40°E azimuth angle is −9–11 MPa; the min-
Modeling Results and Analyses
imum principal stress with the 40°–55° dip angle and the N25–40°E
The coordinate origin of the numerical simulation model for Jinp- azimuth angle is −5–8 MPa. Large underground caverns often cross
ing II hydropower station underground caverns is located at the in- different geological units and encounter local bad formations and
tersection of the centerline of unit no. 8 and the central axis of the faults. Therefore, underground plants are usually equipped with
machine hall. The orientation of the three coordinate axes is as fol- multiple system monitoring sections along the axis. The under-
lows: X-axis S55°E, Y-axis N35°E, and the Z-axis coincides with ground caverns of the Jinping II hydropower station are equipped
the geodetic coordinate system. The model is 860 m long in the with five monitoring sections, namely, S1–S5, for subsequent

© ASCE 04022259-8 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


Fig. 8. Systematic monitoring section of underground caverns.

Fig. 10. Layered excavation scheme for underground caverns.


Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

comparative analysis of numerical simulation results and field mea-


surement information, as shown in Fig. 8.

Deformation Characteristics of the Surrounding Rock


Although the multipoint displacement meter installed in the sur-
rounding rock of underground caverns can monitor and obtain de-
formation in real time, it is more comprehensive and clearer to
reveal the overall deformation characteristics of the surrounding
rock through the deformation field of the surrounding rock after ex-
cavation is obtained from numerical simulations. Generally, the de-
formation of the upstream abutment and the upstream sidewall of
the machine hall is large after excavation, with the maximum
(a) value reaching 65–70 mm. The deformation of the surrounding
rock under the busbar tunnel of the downstream sidewall of the ma-
chine hall is the part with the largest deformation on the cross sec-
tion of the whole opening after excavation, and the maximum
deformation can reach approximately 70 mm. In addition, the de-
formation of the surrounding rock of the upstream sidewall of the
transformer chamber is relatively large—up to approximately
60 mm [as shown in Fig. 9(a)]. Considering that the direction of
the steeply dipping stratum at the site has a small included angle
with the axis of the machine hall and the transformer chamber,
the direction of the maximum principal stress is approximately per-
pendicular to the axis of the machine hall. Therefore, the deforma-
tion of the sidewall of underground caverns is bound to be large.
According to the measured deformation of the surrounding rock
(b)
obtained by the multipoint displacement meter at the upper part
of the upstream sidewall of monitoring Sections S1–S4 of the ma-
chine hall, as of early July 2010, the maximum deformation of the
surrounding rock measured at the monitoring point with 1,348 m
elevation of the upstream sidewall of the main power house was
90.18 mm, and the average deformation was 65.5 mm after excava-
tion (as shown in Fig. 10). The measured maximum deformation of
the surrounding rock at the monitoring point with an elevation of
1,355 m is 94.0 mm, and the average deformation is 60.4 mm
after excavation (as shown in Fig. 11). The deformation of the sur-
rounding rock in the middle of the upstream sidewall of the ma-
chine hall obtained by numerical simulation is 65 mm, which is
essentially consistent with the measured deformation magnitude.
In addition, the multipoint displacement meter of the upstream side-
wall of the transformer chamber detected large deformation of the
(c) surrounding rock and multiple monitored anchor cable loads ex-
ceeding the design tonnage. These monitoring results show that
Fig. 9. (a) Deformation field of the typical unit section of underground the deformation of the surroundings of the upstream sidewall of
caverns; and (b and c) overall deformation field of underground the transformer chamber is prominent. At the same time, the unem-
caverns. bedded multipoint displacement meter in the busbar tunnel of the

© ASCE 04022259-9 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


sidewall downstream of the machine hall measured all deformation intuitively shows that large deformation did occur in this part
of the surrounding rock after excavation. However, several annular (as shown in Fig. 12).
cracks with a width of 5–20 mm appeared in the shotcrete and the
surrounding rock near the side entrance of the busbar tunnel, which Stress Characteristics of the Surrounding Rock
The secondary stress adjustment of the surrounding rock occurs
after the excavation of underground caverns. The stress concentra-
tion of caverns is mainly located in the downstream abutment of the
machine hall, the downstream abutment of the transformer cham-
ber, and the foundation pit. The concentrated maximum principal
stress reaches approximately −34 MPa (the compressive stress is
negative and the tensile stress is positive), as shown in Fig. 13.
The stress of the upstream sidewall of the machine hall, the inter-
section between the upstream sidewall of the machine hall and
the high-pressure pipeline, the intersection between the down-
stream sidewall of the machine hall and the busbar tunnel, and
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

the upstream sidewall of the transformer chamber are small after


the excavation of the surrounding rock, and some tensile stress
even appears locally.
Considering the steep slope of the high sidewall of the machine
hall and the transformer chamber, it is very important to implement
anchor bolt, shotcrete, and anchor cable support in time to avoid
Fig. 11. Measured deformation of the upstream sidewall of the ma-
cracking and collapse of the surrounding rock of the high sidewall
chine hall.
caused by adverse secondary stress after excavation. In fact, the

Fig. 12. Annular cracking of surrounding rock and shotcrete near the busbar tunnel.

(a) (b)

Fig. 13. Redistributed stress field of the typical unit section of underground caverns: (a) maximum principal stress; and (b) minimum principal stress.

© ASCE 04022259-10 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


characteristics of redistributed stress in the surrounding rock are reasonability of the stress concentration phenomenon of the down-
controlled by the initial maximum principal stress. The initial max- stream abutment of the machine hall revealed by the preceding nu-
imum principal stress in the rock mass is nearly vertical to the axis merical simulation but also reflect that the numerical calculation
of the machine hall, and the inclination angle is 40°–55°. Under the results can reasonably explain the essential reason for the failure
action of this principal stress, it inevitably leads to stress concentra- of the surrounding rock in this part. On the other hand, the redistri-
tion in the downstream abutment of the cavern, but the rock mass of bution stress field of the surrounding rock obtained by numerical
the upstream sidewall shows an obvious stress relaxation phenom- simulation reveals that the stress relaxation phenomenon of the sur-
enon. Therefore, the characteristics of the geostress field in this area rounding rock of the upstream sidewall of the machine hall and the
should be fully considered in the system support of the whole cav- transformer chamber is obvious, which is consistent with the phe-
erns. During the excavation of Floors III and IV, the shotcrete of the nomenon that the surrounding rock layer and structural plane are
downstream abutment of the machine hall gradually bulged and generally relaxed and open significantly during the excavation pro-
partially dislocated to a certain extent (as shown in Fig. 14). Be- cess of the upstream sidewall of the machine hall and the trans-
cause the strength of the rock mass in the caverns is not enough former chamber (Fig. 15).
and the stress concentration area is gradually formed in this part
during the excavation process, the stress–strength ratio of the sur- Distribution Characteristics of the Plastic Zone
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

rounding rock at the downstream abutment of the machine hall According to the results of the numerical simulation, the plastic
reaches 0.6–0.8 in the redistributed stress field, obviously reaching zone’s depth of the surrounding rock is approximately 2 m; yet,
the threshold value of the crack initiation stress and the stable de- the surrounding rock’s plastic zone reaches approximately 3 m in
velopment stress of the general rock. Therefore, the surrounding depth at the intersection of the caverns and the stress concentration
rock will gradually develop from bulging deformation of concrete part, such as the intersection of the upstream sidewall of the ma-
to cracking and dislocation with the subsequent excavation. The chine hall and the high-pressure pipeline, the downstream abutment
failure characteristics of the surrounding rock not only show the of the transformer chamber, and especially the rock platform above
the intersection of the diffusion area of the foundation pit and the
tailrace tunnel (Fig. 16). Further statistics show that the actual ex-
cavation damage zone depth is consistent with the depth of the plas-
tic zone by numerical simulations (as provided in Table 3). The
surrounding rock of the downstream abutment of the main power-
house and the transformer chamber is generally shear yielded,
while the surrounding rock of the sidewall of the opening is gener-
ally tensile or pull-shear yielded under the control of the stress field
after excavation. The plastic zone extent and yield mode displayed
by numerical simulation results directly reveal the damage degree
and mode of the surrounding rock after excavation. The yield char-
acteristics of the surrounding rock obtained by numerical simula-
tion are also well verified in the actual excavation process of the
caverns; for example, although advanced presupport or lock sup-
port has been done, the surrounding rock of the intersection still ap-
pears to be damaged and collapses during the excavation process of
the intersection of several caverns.
As shown in Fig. 17, the surrounding rock collapse phenome-
non occurs at the intersection of the machine hall and the high-
Fig. 14. Bulging and dislocation of shotcrete at the downstream abut-
pressure pipeline, the transformer chamber, and the busbar tunnel.
ment of the machine hall.
This is in accordance with the numerical simulation, which reveals

(a) (b)

Fig. 15. Surface opening of the sidewall.

© ASCE 04022259-11 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


that the plastic zone extent of the surrounding rock at the intersec- The rock fracture degree is an index proposed by Feng et al.
tion entrance is relatively large. In addition, Fig. 16 shows the shear (2021) for rock mass fracture evaluation, which is defined as
failure characteristics of the surrounding rock of the downstream Eq. (55). Here, we adopted the RFD index to evaluate the fracture
abutment of the machine hall. Fig. 15 shows that tensile failure degree of joined rock masses in underground caverns. The RFD
of the surrounding rock of the sidewall was observed at the site, index of underground caverns after excavation is shown in Fig. 18:
which is also consistent with the failure mode of the surrounding ⎧

⎪ g(θ) Ap2 + Bp + C − q
rock revealed by the numerical simulation in Fig. 16. ⎪1 −
⎨ , F < 0
RFD = g(θ) Ap2 + Bp + C (55)

⎪ εpV γ

⎩1 + p + , F≥0
(εV )| limit (γ )| limit
   
where g(θ) Ap2 + Bp + C − q / g(θ) Ap2 + Bp + C = yield-
ing proximity; and ((εpV )/(εpV )| limit ) + (γ /(γ )| limit ) = measure of in-
elastic failure in the postpeak region, which considers both the
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

influence of the plastic volume change and plastic distortion.g(θ)


is the shape function of the load angle θ in the deviatoric plane,
p is the mean principal stress, and q is the equivalent shear stress.
Following a modified Wiebols–Cook failure criterion, which can
reflect the strain energy accumulation induced by microcracking,
the definitions of coefficients A, B, C, and g(θ)can be found in
Colmenares and Zoback (2002). εpV and γ are the plastic volumetric
strain and equivalent plastic shear strain, respectively.

Fig. 16. Plastic zone distribution of the typical unit section of under-
ground caverns.

Table 3. Measured excavation damage depth of the inspection section of


the main powerhouse
Type S1 S2 S3 S4
Top of the arch 2.8 2.0 1.6 1.4
Upstream EL1364 2.4 2.4 2.8 2.4
Upstream EL1364 1.4 1.0 1.6 2.6
Upstream EL1335 2.8 1.2 3.4 3.2
Upstream EL1335 1.8 1.2 3.4 2.8
Upstream EL1322 2.0 2.0 1.6 2.4
Upstream EL1322 1.2 1.8 1.2 2.0 Fig. 18. RFD of surrounding rock at typical unit sections.

(a) (b)

Fig. 17. Surrounding rock of the typical fork opening collapses: (a) busbar tunnel intersects with the transformer chamber; and (b) high-pressure
pipeline intersects with the machine hall.

© ASCE 04022259-12 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


(a) (b)
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

(c) (d)

(e) (f )

(g) (h)

Fig. 19. Comparison between (a, c, e, and g) the proposed model and (b, d, f, and h) the ubiquitous-joint model in terms of deformation, stress, and RFD.

© ASCE 04022259-13 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


Table 4. Parameters for the UBJ model
K (GPa) G (GPa) c (MPa) ϕ (°) σt (MPa) ψ (°) cj (MPa) ϕj (°) ψj (°) σ ij (MPa) dd (°) dip (°)
27 16 7.2 21 1 15 2 22 18 1 300 73
Note: K = bulk modulus of the equivalent rock; G = shear modulus of the equivalent rock; c = cohesion of the equivalent rock; ϕ = internal friction angle of the
equivalent rock; σ t = tensile strength of the equivalent rock; ψ = dilation angle of the equivalent rock; cj = cohesion of the joint set; ϕj = internal friction angle
of the joint set; ψj = dilation angle of the joint set; σ tj = tensile strength of the joint set; dd = dip-direction; and dip = dip angle.

Table 5. Differences between the proposed constitutive model and UBJ model
Machine Machine Transformer Transformer
Type Location hall hall chamber chamber
Proposed model Deformation of the upstream abutment (mm) 70 65 60 30
Depth of the plastic zone of the upstream sidewall (m) 2.3 1.9 1.9 3.1
Ubiquitous-joint model Deformation of the upstream sidewall (mm) 35 40 30 10
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

Depth of the plastic zone of the downstream abutment (m) 1.1 0.8 0.7 1.3

Correspondingly, (εpV )| limit and (γ )| limit are their limit values. F is Conclusion
the yield function.
In this work, the joint compliance tensor for jointed rock masses is
proposed and an equivalent constitutive model (JCT model) is es-
tablished. Also, the model is adopted for engineering applications
Discussions in Jinping II hydropower station underground caverns. The primary
conclusions are summarized as follows:
A comparison between the proposed model and the ubiquitous- 1. The joint compliance tensor for the jointed rock masses is pro-
joint model built-in FlAC3D is shown in Fig. 19. The parameters posed to characterize the spatial geometry and deformation
for the ubiquitous-joint model are provided in Table 4, and the dif- characteristics of the joint system by a single parameter based
ferences in the deformation and depth of plastic are given in on the Oda fracture tensor theory, which provides a theoretical
Table 5. In general, the deformation and depth of the plastic zone basis to evaluate the deformation resistance and spatial anisot-
calculated by the proposed model are much larger than those calcu- ropy of the jointed rock masses.
lated by the ubiquitous-joint model, and the stress distribution is 2. An equivalent constitutive model for the jointed rock masses is
similar to some extent, but the area of the stress concentration established based on the proposed joint compliance tensor and
area is obviously different in terms of stress characteristics. linear superposition principle under the framework of equiva-
When adopting the proposed model, the deformation of the abut- lent continuity method, and the model can describe the aniso-
ment and sidewall of the upstream of the machine hall is approxi- tropic mechanical characteristics and spatial deformation of
mately 70 and 65 mm, respectively, and the depth of the plastic multisets jointed rock masses.
zone is approximately 2.3 and 1.9 m, respectively; the deformation 3. The combined field investigation monitoring data and applica-
of upstream sidewall and downstream abutment of the transformer tion of the proposed constitutive model to analyze the stability
chamber is approximately 60 and 30 mm, respectively; and the of the jointed rock masses around the caverns at the Jinping II
depth of plastic zone is 1.9 and 3.1 m, respectively; in contrast, hydropower station indicate that they can accurately estimate
when using the ubiquitous-joint model, the deformation of the abut- the locations and depth of the potential collapses. Therefore,
ment and sidewall of the upstream of the machine hall is approxi- they are of significant practical value in the stability evaluation
mately 35 and 40 mm, respectively, and the depth of plastic zone is of the jointed rock masses around underground caverns.
approximately 1.1 and 0.8 m, respectively; the deformation of up-
stream sidewall and downstream abutment of the transformer
chamber is approximately 30 and 10 mm, and the depth of plastic Data Availability Statement
zone is 0.7 and 1.3 m, respectively. According to the comparison of
the calculation results between the proposed JCT model and the Geological data and numerical simulation data used during the
UBJ model, it can be seen that the calculation error is relatively study are available from the corresponding author by request.
large using the UBJ model. That is probably because the UBJ
model can only be applied to the engineering situation of one set
of joints but cannot satisfy the actual demand of multiple sets of Acknowledgments
joints, which does not agree with the real engineering case leading
to the distortion of calculation results. What is more, the UBJ The authors gratefully acknowledge the financial support from the
model also ignores the deterioration law of strength parameters in National Natural Science Foundation of China (No. U1965205).
the process of jointed rock mass, which will also result in a large
error in the simulation effect. Therefore, it is evident that the pro-
posed JCT model can well describe the anisotropic mechanical
References
characteristics and spatial deformation of multisets jointed rock
masses and better evaluate and predict the mechanical response Adhikary, D., and A. Dyskin. 1998. “A continuum model of layered rock
of the jointed rock masses than the traditional UBJ model. That masses with non-associative joint plasticity.” Int. J. Numer. Anal.
is of important practical value in the stability assessment of rock Methods Geomech. 22: 245–261. https://doi.org//10.1002/(SICI)1096
masses with multiple joint sets around underground caverns. -9853(199804)22:4<245::AIDNAG916>3.0.CO;2-R.

© ASCE 04022259-14 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


Aladejare, A. E., and Y. Wang. 2018. “Influence of rock property correla- 637–659. https://doi.org//10.1061/JSFEAQ.000800310.1061/jsfeaq
tion on reliability analysis of rock slope stability: From property char- .0001133.
acterization to reliability analysis.” Geosci. Front. 9 (6): 1639–1648. Hart, R., P. A. Cundall, and J. Lemos. 1988. “Formulation of a three-
https://doi.org/10.1016/j.gsf.2017.10.003. dimensional distinct element model—Part II. Mechanical calculations
Amadei, B., and R. E. Goodman. 1981. “A 3D constitutive relation for frac- for motion and interaction of a system composed of many polyhedral
tured rock masses.” In Proc., Int. Symp. on the Mechanical Behavior of blocks.” Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 25 (3): 117–
Structured Media, 249–268. Ottawa, Canada: Carleton University. 125. https://doi.org/10.1016/0148-9062(88)91215-6.
Anagnostou, G. 1993. “A model for swelling rock in tunnelling.” Rock Hauseux, P., E. Roubin, D. M. Seyedi, and J. B. Colliat. 2016. “FE mod-
Mech. Rock Eng. 26 (4): 307–331. https://doi.org/10.1007/bf01027115. elling with strong discontinuities for 3D tensile and shear fractures:
Aubertin, M., L. Li, and R. Simon. 2000. “A multiaxial stress criterion for Application to underground excavation.” Comput. Methods Appl.
short- and long-term strength of isotropic rock media.” Int. J. Rock Mech. Eng. 309: 269–287. https://doi.org/10.1016/j.cma.2016.05.014.
Mech. Min. Sci. 37 (8): 1169–1193. https://doi.org/10.1016/S1365 Hoek, E., and P. Marinos. 2000. “Predicting tunnel squeezing problems
-1609(00)00047-2. in weak heterogeneous rock mass.” Tunnels Tunnelling Int. 32 (11):
Barron, N., and B. T. Shen. 2018. “Extension strain and rock strength limits 45–51.
for deep tunnels, cliffs, mountain walls and the highest mountains.” Hudson, J. A., J. W. Cosgrove, K. Kemppainen, and E. Johansson. 2011.
Rock Mech. Rock Eng. 51: 3945–3962. https://doi.org/10.1016/j.jrmge “Faults in crystalline rock and estimation of their mechanical properties
.2016.11.004. at the Olkiluoto site, western Finland.” Eng. Geol. 117: 246–258. https://
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

Chang, L., H. Konietzky, and T. Frühwirt. 2019. “Strength anisotropy of doi.org/10.1016/j.enggeo.2010.11.004.


rock with crossing joints: Results of physical and numerical modeling Itasca Consulting Group. 2011. Fast Lagrangian analysis of continua in 3
with gypsum models.” Rock Mech. Rock Eng. 52: 2293–2317. https:// dimensions, user’s guide. Itasca Flac3D V5.0. Minneapolis: Itasca
doi.org//10.1007/s00603-018-1714-8. Consulting Group.
Chen, L., J. F. Shao, and H. W. Huang. 2010. “Coupled elastoplastic dam- Jiang, Q., X. T. Feng, J. L. Chen, C. S. Zhang, and S. L. Huang. 2008.
age modeling of anisotropic rocks.” Comput. Geotech. 37: 187–194. “Nonlinear inversion of 3D initial geostress field in Jinping II hydro-
https://doi.org//10.1016/j.compgeo.2009.09.001. power station region.” Rock Soil Mech. 29 (11): 3003–3010. https://
Chen, S. H., and P. Egger. 1999. “Three dimensional elasto-viscoplastic fi- doi/org/10.10.16285/j.rsm.2008.11.015.
nite element analysis of reinforced rock masses and its application.” Jiang, Q., X. T. Feng, Y. H. Hatzor, X. J. Hao, and S. J. Li. 2014.
Int. J. Numer. Anal. Methods Geomech. 23 (1): 61–78. https://doi.org “Mechanical anisotropy of columnar jointed basalts: An example
/10.1002/(SICI)1096-9853(199901)23:1<61::AID-NAG958>3.0.CO;2 from the Baihetan hydropower station, China.” Eng. Geol. 175: 35–
-V. 45. https://doi.org/10.1016/j.enggeo.2014.03.019.
Chen, X., Z. Liao, and X. Peng. 2012. “Deformability characteristics of Kaiser, P. K., and M. Cai. 2012. “Design of rock support system under rock-
jointed rock masses under uniaxial compression.” Int. J. Min. Sci. burst condition.” J. Rock Mech. Geotech. Eng. 4: 215–227. https://doi.org
Technol. 22: 213–221. https://doi.org//10.1016/j.ijmst.2011.08.012. /10.3724/SP.J.1235.2012.00215.
Chen, Y. F., H. K. Zheng, M. Wang, J. M. Hong, and C. B. Zhou. 2015. Kim, J., C. Kim, G. Kim, I. Kim, Q. Abbas, and J. Lee. 2021. “Probabilistic
“Excavation-induced relaxation effects and hydraulic conductivity var- tunnel collapse risk evaluation model using analytical hierarchy process
iations in the surrounding rocks of a large-scale underground power- (AHP) and Delphi survey technique.” Tunnelling Underground Space
house cavern system.” Tunnelling Underground Space. Technol. 49: Technol. 120: 104262. https://doi.org/10.1016/j.tust.2021.104262.
253–267. https://doi.org/10.1016/j.tust.2015.05.007. Kumar, V., V. Gupta, I. Jamir, and S. L. Chattoraj. 2019. “Evaluation of
Chen, Y. F., C. B. Zhou, and Y. Q. Sheng. 2007. “Formulation of strain- potential landslide damming: Case study of Urni landslide, Kinnaur,
dependent hydraulic conductivity for a fractured rock mass.” Satluj valley, India.” Geosci. Front. 10 (2): 753–767. https://doi.org
Int. J. Rock Mech. Min. Sci. 44: 981–996. https://doi.org/10.1016/j /10.1016/j.gsf.2018.05.004.
.ijrmms.2006.12.004. Lekhnitskii, S. G., P. Fern, and J. J. Brandstatter. 1964. “Theory of elastic-
Colmenares, L. B., and M. D. Zoback. 2002. “A statistical evaluation of in- ity of an anisotropic elastic body.” Phys. Today 17 (1): 84. https://doi
tact rock failure criteria constrained by polyaxial test data for five differ- .org/10.1063/1.3051394.
ent rocks.” Int. J. Rock Mech. Min. Sci. 39 (6): 695–729. https://doi.org Lemos, J. V., R. D. Hart, and P. A. Cundall. 1985. “A generalized distinct
/10.1016/S1365-1609(02)00048-5. element program for modelling jointed rock mass.” In Proc., Symp.
Cundall, P. A. 1971. “A computer model for simulating progressive large- International Society of Rock Mechanics, 335–343. Lapland: Swedish
scale movement in blocky rock systems.” In Vol. 1 of Proc., Int. Symp. Rock Engineering Research Foundation.
on Rock Fracture, 8–11. Nancy, France: School of Engineering Li, S. J., M. Z. Zheng, S. L. Qiu, Z. B. Yao, J. F. Zhou, and P. Z. Pan. 2021.
Geology. “Analysis of disaster characteristics and long-term in-situ mechanical
Cundall, P. A. 1988. “Formulation of a three-dimensional distinct element response of excavation tunnel in Jinping Underground Laboratory of
model—Part I. A scheme to detect and represent contacts in a system China.” J. Tsin. Univ. (Sci. Technol.) 61 (8): 842–852. https://doi.org//
composed of many polyhedral blocks.” Int. J. Rock Mech. Min. Sci. 10.16511/j.cnki.qhdxxb.2021.26.015.
Geomech. Abstr. 25 (3): 107–116. https://doi.org/10.1016/0148 Nova, R. 1980. “The failure of transversely isotropic rocks in triaxial com-
-9062(88)91214-4. pression.” Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 17: 325–332.
Das, A. J., P. K. Mandal, P. S. Paul, R. K. Sinha, and S. Tewari. 2019. https://doi.org//10.1016/0148-9062(80)90515-X.
“Assessment of the strength of inclined coal pillars through numerical Oda, M. 1982. “Fabric tensor for discontinuous geological material.” Soils
modelling based on the ubiquitous joint model.” Rock Mech. Rock Found. 13 (1): 1–9. https://doi.org/10.3208/sandf1972.22.4_96.
Eng. 52: 3691–3717. https://doi.org/10.1007/s00603-019-01826-4. Pariseau, W. G. 2012. “Finite element approach to caving in stratified,
Do, T. N., and J. H. Wu. 2020. “Simulation of the inclined jointed tock jointed rock masses.” Int. J. Rock Mech. Min. Sci. 53: 94–100. https://
mass behaviors in a mountain tunnel excavation using DDA.” doi.org/10.1016/j.ijrmms.2012.05.004.
Comput. Geotech. 117: 103249. https://doi.org//10.1016/j.compgeo Park, S. H., and T. Adachi. 2002. “Laboratory model tests and FE analyses
.2019.103249. on tunneling in the unconsolidated ground with inclined layers.”
Dyskin, A. V. 1999. “On the role of stress fluctuations in brittle fracture.” Tunnelling Underground Space. Technol. 17: 181–193. https://doi.org//
Int. J. Fract. 100: 29–53. https://doi.org/10.1023/A:1018664101433. 10.1016/S0886-7798(02)00003-2.
Feng, X. T., Z. F. Wang, Y. Y. Zhou, C. X. Yang, P. Z. Pan, and R. Kong. Rehman, H., A. M. Naji, J. J. Kim, and H. K. Yoo. 2019. “Extension of
2021. “Modelling three-dimensional stress-dependent failure of hard tunneling quality index and rock mass rating systems for tunnel support
rocks.” Acta Geotech. 16: 1647–1677. https://doi.org/10.1007/s11440 design through back calculations in highly stressed jointed rock mass:
-020-01110-8. An empirical approach based on tunneling data from Himalaya.”
Goodman, R. E., R. L. Taylor, and T. L. Brekke. 1968. “A model for Tunnelling Underground Space Technol. 85: 29–42. https://doi.org/10
the mechanics of jointed rock.” J. Soil Mech. Found. Div. 94 (3): .1016/j.tust.2018.11.050.

© ASCE 04022259-15 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259


Sainsbury, B. L., and D. P. Sainsbury. 2017. “Practical use of the Xu, D. P., X. T. Feng, D. F. Chen, C. Q. Zhang, and Q. X. Fan. 2017.
ubiquitous-joint constitutive model for the simulation of anisotropic “Constitutive representation and damage degree index for the layered
rock masses.” Rock Mech. Rock Eng. 50: 1507–1528. https://doi.org// rock mass excavation response in underground caverns.” Tunnelling
10.1007/s00603-017-1177-3. Underground Space Technol. 64: 133–145. https://doi.org/10.1016/j
Shen, B., J. Shi, M. Rinne, T. Siren, J. Suikkanen, and S. Kwon. 2016. .tust.2017.01.016.
“Two dimensional displacement discontinuity method for transversely Xu, D. P., X. Huang, S. J. Li, H. S. Xu, S. L. Qiu, H. Zheng, and Q. Jiang.
isotropic materials.” Int. J. Rock Mech. Min. Sci. 83: 218–230. https:// 2021. “Predicting the excavation damaged zone within brittle surround-
doi.org//10.1016/j.ijrmms.2016.01.012. ing rock masses of deep underground caverns using a comprehensive
Singh, M., K. S. Rao, and T. Ramamurthy. 2002. “Strength and deforma- approach integrating in situ measurements and numerical analysis.”
tional behaviour of a jointed rock mass.” Rock Mech. Rock Eng. 35: 45– Geosci. Front. 13 (2): 101273. https://doi.org/10.1016/j.gsf.2021
64. https://doi.org//10.1007/s006030200008. .101273.
Sitharam, T. G., J. Sridevi, and N. Shimizu. 2001. “Practical equivalent Yang, W., Q. Zhang, P. G. Ranjith, R. Yu, G. Luo, and C. Huang. 2019. “A
continuum characterization of jointed rock masses.” Int. J. Rock damage mechanical model applied to analysis of mechanical properties
Mech. Min. Sci. 38: 437–448. https://doi.org//10.1016/S1365-1609(01) of jointed rock masses.” Tunnelling Underground Space Technol. 84:
00010-7. 113–128. https://doi.org//10.1016/j.tust.2018.11.004.
Tien, Y. M., and M. C. Kuo. 2001. “A failure criterion for transversely iso- Yoshida, H., and H. Horii. 2004. “Micromechanics-based continuum
tropic rocks.” Int. J. Rock. Mech. Min. Sci. 38: 399–412. https://doi.org//
Downloaded from ascelibrary.org by Quan Jiang on 11/04/22. Copyright ASCE. For personal use only; all rights reserved.

model for a jointed rock mass and excavation analyses of a large-scale


10.1016/S1365-1609(01)00007-7. cavern.” Int. J. Rock Mech. Min. Sci. 41 (1): 119–145. https://doi.org//10
Tien, Y. M., M. C. Kuo, and C. H. Juang. 2006. “An experimental inves-
.1016/s1365-1609(03)00080-7.
tigation of the failure mechanism of simulated transversely isotropic
Zhao, J. S., X. T. Feng, Q. Jiang, and Y. Y. Zhou. 2018. “Microseismicity
rocks.” Int. J. Rock Mech. Min. Sci. 43: 1163–1181. https://doi.org//10
monitoring and failure mechanism analysis of rock masses with weak
.1016/j.ijrmms.2006.03.011.
interlayer zone in underground intersecting chambers: A case study
Ulusay, R., M. Ekmekci, E. Tuncay, and N. Hasancebi. 2014. “Improvement
from the Baihetan hydropower station, China.” Eng. Geol. 245: 44–
of slope stability based on integrated geotechnical evaluations and hydro-
60. https://doi.org/10.1016/j.enggeo.2018.08.006.
geological conceptualisation at a lignite open pit.” Eng. Geol. 181: 261–
Zhou, W. Y., Y. G. Liu, and J. D. Zhao. 2003. “Multi-potential based dis-
280. https://doi.org/10.1016/j.enggeo.2014.08.005.
Wang, T. T., and T. H. Huang. 2009. “A constitutive model for the defor- continuous bifurcation model for jointed rock masses and its applica-
mation of a rock mass containing sets of ubiquitous joints.” Int. J. Rock tion.” Comput. Methods Appl. Mech. Eng. 192: 3569–3584. https://
Mech. Min. Sci. 46: 521–530. https://doi.org//10.1016/j.ijrmms.2008.09 doi.org/10.1016/S0045-7825(03)00356-6.
.011. Zhou, Y. Y., X. T. Feng, D. P. Xu, and Q. X. Fan. 2017. “An enhanced
Wang, T. T., and T. H. Huang. 2014. “Anisotropic deformation of a circular equivalent continuum model for layered rock mass incorporating bed-
tunnel excavated in a rock mass containing sets of ubiquitous joints: ding structure and stress dependence.” Int. J. Rock Mech. Min. Sci.
Theory analysis and numerical modeling.” Rock Mech. Rock Eng. 47: 97: 75–98. https://doi.org//10.1016/j.ijrmms.2017.06.006.
643–657. https://doi.org//10.1007/s00603-013-0405-8. Zhou, Y. Y., D. P. Xu, K. Liu, and D. F. Chen. 2021. “Understanding the
Wei, J. B., J. H. Deng, D. K. Wang, D. W. Cai, and J. Z. Hu. 2010. failure mechanism of a large underground cavern in steeply dipping lay-
“Characterization of deformation and fracture for rock mass in under- ered rock mass using an enhanced ubiquitous-joint model.” Bull. Eng.
ground powerhouse of Jinping I hydropower station.” Chin. J. Rock Geol. Environ. 80: 4621–4638. https://doi.org//10.1007/s10064-021
Mech. Eng. 29 (6): 1198–1205. -02213-6.

© ASCE 04022259-16 Int. J. Geomech.

Int. J. Geomech., 2023, 23(1): 04022259

You might also like