Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Underground Singapore 2016

Discontinuous deformation analysis in rock engineering


modelling
Z. Y. Zhao, H. M. Chen, W. Nie
Nanyang Centre for Underground Space, Nanyang Technological University, Singapore

J. P. Sun
Tritech Consultants Pte Ltd, Singapore

ABSTRACT: The discontinuous deformation analysis (DDA) method can help study both the static
and dynamic behavior of fractured rock masses, in which the fractures can be represented explicitly
with independent physical parameters. In this paper, three application examples are used to
demonstrate the DDA’s capability in rock engineering application. In the first example, the stability
for a classic rock slope is investigated by the DDA, which is compared with the results from the
Sarma’s method. Next, the DDA is applied to study the response of rockbolts under the static pullout
tests. The last example is a coupled hydro-mechanical model to estimate the inflow rate and the
hydraulic pressure distribution for an underground rock cavern considering the grouting effect.

1 INTRODUCTION

The fractured rock mass is a discontinuous medium with full of geological structures, such as
randomly distributed joints and cracks (Fu et al, 2015). These structural planes divide the fractured
rock mass and greatly influence the stability of the underground structures. Thus, it is important to
study the mechanical behaviours for this discontinuous blocky system.

Compared with the experimental study, the numerical analysis is more convenient to address the
discontinuity problems. Different from the continuum methods (FEM), the discontinuous methods are
considered more appropriate to deal with a mesh created by the real rock joints. Among the
discontinuous methods, the discontinuous deformation analysis (DDA) proposed by Shi (1998) is
considered as an attractive numerical tool for geotechnical problems, such as slope engineering,
reinforcement, coupled hydro-mechanical analysis for tunnels/caverns and so on.

In this paper, the DDA will be applied to three engineering applications, including the safety
estimation for a rock slope, response of rockbolts in the static pull-out test and the coupled hydro-
mechanical analysis for the rock caverns. The purpose is to demonstrate the applicability of DDA for
the practical problems.

2 THEORY OF THE DISCONTINUOUS DEFORMATION ANALYSIS (DDA)

The DDA is an implicit method that represents the deformability of the block systems. In this method,
both the static and dynamic analyses are considered. Large displacements, deformations and also the
rotations are allowed. In the DDA model, displacements are the unknowns and represented by six
basic variables in the first-order approximations (three rigid body motion and three constant strain
components) to solve for the global equilibrium equations (Jing, 2003). The displacement  u, v  at
any point  x, y  inside a block can be expressed as follow (Shi, 1988; Hatzor et al, 2004; Bao, 2007;
Sun et al, 2011; Ning et al, 2012):
 u0 
v 
 0
u  1 0 ( y  y0 ) ( x  x0 ) 0 ( y  y0 ) / 2   r0 
    (1)
 v  0 1 ( x  x0 ) 0 ( y  y0 ) ( x  x0 ) / 2    x 
 y 
 
 xy 
where ( x, y) are the coordinates of any point within the rock block; ( x0 , y0 ) are the coordinates of a
point within the rock block, which is usually taken at the block centroid; u0 , v0 are the rigid body
translations at the point  x0 , y0  along the x and y directions; r0 is the rigid body rotation angle in
radian with respect to the point  x0 , y0  ;  x ,  y ,  xy are normal and shear strains of the block. These
six variables are corresponding to the deformation and movement of a general rock block.

The block system can be generated through the contacts among the rock blocks and the displacement
constrains on blocks. If there are n blocks in the block system (Shi, 1988; Sun et al, 2011), the global
equations can be defined by:
 d1   F1 
 K11 K1n   d 2   F2 
  d    F  (2)
  3  3
 K n1 K nn     
   
d n   Fn 
where di  and T Fi  are 6  1 sub-matrices, di  represents the deformation vector
u , v , 
0 0 0 ,  x ,  y ,  xy 
of the block i , while the  Fi  is the loading vector distributed to the six
deformation variables; [ Kij ] is a 6  6 material/contact matrix as each block has six degrees of freedom,
which depends on the material properties of the rock blocks and the contact springs.

3 APPLICATION EXAMPLES

3.1 Safety estimation for the classical rock slope

The DDA can be treated as a tool for the rock slope stability analysis. A classic rock slope from Hoek
(1983) is used for the factors of safety estimation, which will be compared with the results obtained
from the Sarma’s method. Through this simulation, the factors of safety and the reaction force around
the rock blocks can be understood better.

3.1.1 Sarma’s method

Sarma’s method is based on the limit equilibrium analysis that can help estimate the stability of rock
slopes of various shapes (Figure 1). In this method, different shear strengths for the side and the base
of each block can be specified. A fictitious horizontal acceleration K c was proposed to evaluate the
safety of a two-dimensional slope. If Kc  0 , the slope is stable, and if Kc  0 , the slope is at critical
state. When Kc  0 , the slope is unstable.

The limit equilibrium conditions are considered to prevail on both the base and the interfaces of each
rock block at the same time (Sarma, 1979). Therefore, the Mohr-Coulomb failure criterion occurs
along the base and the interfaces of each rock block. Without the external forces, the fictitious
horizontal acceleration Kc can be expressed by (Sun et al, 2011 ):
an  an 1en  ...  a1en en 1...e3e2
Kc  (3)
pn  pn 1en  ...  p1en en 1...e3e2
in which
Wi sin( B ,i   i )  Ri cos  B ,i  
 
 Si 1 sin( B ,i   i   i 1 )  Si sin( B ,i   i   i ) 
ai   (4)
cos( B ,i   i   S ,i 1   i 1 )sec S ,i 1

Wi cos( B ,i   i )
pi  (5)
cos( B ,i   i  S ,i 1   i 1 )secS ,i 1

cos( B ,i   i  S ,i   i )secS ,i
ei  (6)
cos( B ,i   i  S ,i 1   i 1 )secS ,i 1

Ri  cB,i bi seci (7)


Si  cS ,i di (8)
where Wi is the weight of block i ;  i is the angle of the block base with respect to the x axis;  i is the
inclination of the interface measured from the positive y axis to the positive x axis;  B ,i ,  S ,i ,  S ,i 1
are the average friction angles on the base and the sides of the block; cB ,i , cS ,i are the average
cohesions on the base and the side of the block; bi is the horizontal distance of the block base; and d i
is the length of the side of the block.

To obtain the static factor of safety FS, the shear strength values tan  and c are reduced to tan  / F
and c / F until the K c is decreased to zero.

Figure 1. Forces acting on individual rock blocks in Sarma’s method.

3.1.2 Rock slope modeling using the DDA

A fictitious horizontal acceleration K c is applied in the DDA to measure the safety for a two-
dimensional slope. Based on a given factor of safety F, a horizontal volume force KcWi is applied on
each rock block at the same time. The rock block displacement will vary accordingly to different K c .
When the potential failure of the rock slope is going to appear, the critical value of K c can be obtained
(Sun et al, 2011).

In this example, a very large rock slope that excavated in an open pit mine (Hoek, 1983) is used for the
stablility analysis using the DDA (Figure 2). This rock slope has an overall angle of about 30 and the
vertical height is 400m. The Young’s modulus of the rock blocks is 200GPa. The Poisson’s ratio for
each block is set as 0.49 to preserve the area of the blocks in two-dimensional analysis (MacLaughlin
et al, 2001). The contact spring stiffness between blocks is 500GPa. As Sarma’s method is a static
equilibrium analysis, the static computation is executed in the DDA simulations.
Figure 2. Block partition of a rock slope (Hoek, 1983).

The K c is determined in the DDA firstly. Two checked points A and B are located at the top and the
toe of the slope, respectively (Figure 2). The horizontal displacement time histories are monitored for
these two checked points to judge the stability status. If these two rock blocks keep stable during the
calculation, the whole slope is considered stable. The comparison of the F  Kc curves from the two
methods is shown in Figure 3(a). For a given factor of safety F, the value of K c obtained by Sarma’s
method is larger than that from the DDA. And for a given K c , a smaller factor of safety F is obtained
from the DDA method. This is because that the Sarma’s method is based on the initial configuration of
the rock slope, and the rock block only slides along its own base segment which it lies on. But in the
DDA method, the initial configurations and the block kinematics are considered. When a rock block is
going to fail, under the effect of the block gravity and the K c , the small displacement taking place to a
block may carry it to contact with another base segment adjacent to its original base segment. As the
mechanical properties for the adjacent base segment are different, the stability of the corresponding
rock block and even the whole rock slope will change accordingly (Sun et al, 2011).

Figure 3. Simulation results for the rock slope analysis, with (a) factor of safety F versus horizontal acceleration
K c ; (b) shear thrust force (F=3).

The distribution of shear thrust forces between blocks at the verge of the potential failure initiation
with a factor of safety F =3 is shown in Figure 3(b). More regular distributions of shear thrust forces
between blocks are found from Sarma’s method. The reason is that Sarma’s method only considers
that the left block moves upward relative to the right one (Figure 1). In reality, the left block is
possible to move downward relative to the right one. In the DDA, the directions of the block
movement are judged automatically based on the overall deformation and movement of the block
system. During the calculation, all the shear thrust forces are smaller than the corresponding shear
strength provided by the normal thrust forces. Under this condition, the limit equilibrium condition
will not occur along the interfaces between blocks at the initiation of slope failure in this example.
Meanwhile, the ratio between the shear thrust force and the shear strength varies at different interfaces,
thus, the failure will not occur simultaneously among all the interfaces, which is the obvious different
from the assumption in Sarma’s method.

The result from the revised Sarma’s method gets closer to that from the DDA (Figures 3 (b)), although
obvious differences still can be found. However, the revised Sarma’s method can help to assess the
DDA’s effectiveness in the factor of safety evaluation for the rock slope. It shows that the DDA can
provide more realistic interaction forces between blocks with fewer assumptions. Hence, the DDA is a
more reasonable method to analyse rock slope stability.

3.2 Underground reinforcement using rockbolts

Rockbolts are widely used for the rock mass reinforcement in mining and other underground
engineering projects. In this section, rockbolt models are introduced to the DDA. To investigate the
responses of different rockbolts under static pull-out test, the expansion shell anchored bolt, the split
set, the fully grouted rebar and the D-bolt are modelled using the developed rockbolt models.

3.2.1 The rockbolt element in the DDA

To integrate the rockbolt element to the DDA blocky system, the rockbolt element is divided into
segments based on the block boundary. The displacement and the anchor forces along the rockbolt can
be obtained from each inserted rockbolt node in segments (Figure 4(a)). A fixed node, such as the
Node 14 in Figure 4(b), is applied at the place where the external fixture is attached. At the very
beginning, the rock Node i’ and the rockbolt Node i are coincident. As the block begins to move, the
rock Node i’ is updated based on the block movement, while the rockbolt Node i can be calculated
based on the behaviour of the bond related to the rock and the rockbolt. After obtaining the coordinates
of the rockbolt nodes, the shear forces applied on the rock nodes can be calculated and applied as point
loads on the block in the next time step (Nie et al, 2014).

Segement A Segment B Segment C

Element 13 ... Element 4 ... Element 1

block A block B block C 14 13 12 11(10) 9 8 7 6(5) 4 3 2 1 (bolt node)

(Fixed)

14' 13' 12' 11'(10') 9' 8' 7' 6'(5') 4' 3' 2' 1' (rock node)

block A block B block C block A block B block C

(a) (b)

Figure 4. DDA block and rockbolt model, with (a) rockbolt model in rock; (b) Bolt node (i) and the
corresponding rock nodes (i’).

3.2.2 Rockbolts in static pullout tests

The rockbolt material in the tests is modeled by the elastic, linear strain-hardening behavior. The pull-
out tests are carried out to measure the load bearing capacity different types of rockbolts. Figure 5(a)
shows a simple static model of the pull-out test for the current rockbolt models, in which the self-
weight of the rock blocks and rockbolt are ignored. The left block A is fixed in the x-direction at
points M and N, while the right block B is pulled out towards right by two point loads at points P and
Q. The blocks A and B are connected with a 1.8m long rockbolt with positive local axes from left to
right. The distance between two rockbolt nodes is 0.15m, and the space between the two special
anchors of D-bolt is 0.9m (Figure 5(b)). The deformation of rockbolt material follows the stress-strain
curve shown in Figure 5(c) and the parameters for the pull-out tests are listed in Table 1. Some of the
rockbolt properties are obtained from Li (2010) and Li (2012).

(a) (b) (c)

Figure 5. Sketches of the pull-out test and the rockbolt models, with (a) block assembly and boundary; (b)
rockbolt models; and (c) rockbolt material model (Nie et al, 2014).

Table 1. Parameters in the pull-out tests simulation (Nie et al, 2014).

Figure 6(a) is the result for the fully grouted rebar. The maximum axial loads are found at Nodes 7 and
8, in which the element reaches to the plastic stage first. As Nodes 6 and 9 reach to the plastic stage
gradually, the main deformation occurs on these two nodes.

The expansion shell anchor bolt has an equally incremental displacement from the anchored Nodes
1and 2 to the fixed Node 14 (Figure 6c)), and the whole rockbolt shank reaches to the plastic stage
simultaneously. From 25s to 30s, the rockbolt restrains the displacement between the two blocks
(Figure 6(d)). Once the whole shank reaches to the plastic stage (from 30s to 35s), the displacement
increases substantially in the rockbolt with little increment of axial load.

The D-bolt also has evenly axial load distribution along the elements between two special anchored
nodes (Nodes 4 and 11). The deformation mainly occurs in the corresponding elements (Figures 6(c)
and (d)). These elements elongate elastically initially, then after yielding, elongate plastically until the
ultimate strain limit is reached at the joint.

The split set can sustain a large deformation or even larger opening between the rock blocks. But, if
the bond stiffness is low, the blocks cannot be restrained by the low axial loads that the displacement
of rockbolt nodes are evenly distributed along the rockbolt (Figures 6(g) and (h)). Under this condition,
the bolt is easy to be pulled out at a low pullout load level.
200 30
Joint t=25s
t=40s t=25s
t=55s 25 t=40s
150 Yield
t=55s

Displacement (mm)
Axial loading (kN)

20 Joint

100 15

10
50
5

0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Node No. Node No.
(a) (b)

200 60
Joint t=35s Joint
Yield 50 t=30s
t=25s

Displacement (mm)
150
Axial loading (kN)

40

100 30

t=35s 20
50 t=30s
t=25s 10

0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Node No. Node No.
(c) (d)

200 150
t=35s Joint
t=35s
t=28s
t=28s
150 t=20s
Displacement (mm)

t=20s
Axial loading (kN)

100
Yield
100 Joint

50
50

0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Node No. Node No.
(e) (f)
200 200
Joint t=23s Joint t=23s
t=12s
t=12s
150 Yield t=6s 150 t=6s
Axial loading (kN)

Displacement (mm)

100 100

50 50

0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14
1 2 3 4 5 6 7 8 9 10 11 12 13 14
Node No. Node No.
(g) (h)

Figure 6. Axial loads and displacements of bolt nodes under extreme pullout loading:
(a) axial loads along the rebar; (b) displacement along the rebar; (c) axial loads along the expansion shell anchor
bolt; (d) displacement along the expansion shell anchor bolt; (e) axial loads along D-bolt; (f) Displacement along
the D-bolt; (g) axial loads along the split set; and (h) Displacement along the split set (Nie et al, 2014).

The pullout test results show that the D-bolt has higher bearing capacity and deformability. Although
high load-bearing capacity can be found in fully grouted rebar, it is too stiff to deform. Compared with
the rebar, the expansion shell anchor bolt has relatively larger deformability. The split set has the largest
deformation capacity, but it is easy to be pulled out due to the relatively low bearing capacity.
3.3 Coupled hydro-mechanical analysis

A coupled hydro-mechanical model is developed based on the DDA, in which the fractures are
considered as the only pathways to conduct the fluid flow. Using this model, the hydraulic pressure
around the caverns after the excavation can be estimated with considering the grouting effect.

3.3.1 Expression for the hydraulic force in the DDA

The hydraulic pressure is supposed to be linearly distributed over a fracture. The expression of the
hydraulic pressure in the DDA can be defined as follow:

 ( Fx1  Fx 2 ) / 2 
 
(
 y1 F  F y2 ) / 2 
  1 
   Fx 2  1 Fx1   y2  y1   1  Fx 2  Fx1  y1  y0    
  3 6  2  
 
  1 F  1 F   x  x   1 F  F  x  x  
  3 y 2 6 y1  2 1 2  y 2 y1 

1 0  

 
 L   Fx 2  Fx1   x2  x1    Fx 2  Fx1  x1  x0   F (9)
  i
1 1 1
W force
 
3 6 2

 1  
  Fy 2  Fy1   y2  y1    Fy 2  Fy1   y1  y0 
1 1

 3 6  2 
 1  1 1  1  
  Fx 2  Fx1   y2  y1    Fx 2  Fx1  y1  y0    
 2  3 6  2  
 
 1  1 Fy 2  1 Fy1   x2  x1   1  Fy 2  Fy1   x1  x0   
 2  3 

 
  6 2  

where Fx and Fy are the hydraulic pressure in x-direction and y-direction, respectively; ( x1 , y1 ) and
( x2 , y2 ) are the coordinates for two nodal points of the applied fracture; ( x0 , y0 ) is the centre of gravity
of the rock block.

3.3.2 Transient flow expression

To model the transient flow along the fracture network, the equation can be expressed by (Wu, 2009):
1 1
([ A][ K f ]  [ D]){H t }  [ D]{H t t } (10)
t t

where [ A] is the connection relationship sub-matrix for intersection points within the fracture
networks;  K f  is the matrix related to the transmissivity per unit fracture length; [ D] is the storage
matrix; {H t } and {H t t } are the hydraulic head sub-matrices at the time t and t  t , respectively.

3.3.3 Hydraulic pressure distribution considering the grouting zone

As soon as the cavern is excavated, the fluid flow within the domain is not steady initially, but in a
transient state. The water inflow may be quite large at the very beginning. Under this condition,
grouting can be used to reduce the inflow rate. With the grouting zone, the hydraulic head distribution
around the excavation is totally different, which is induced by the different hydraulic conductivity
conditions in the grouting zone and the non-grouting zone. An underground cavern (Figure 7.) is used
here to observe the hydraulic head distribution around the excavation considering the grouting zone.
The whole dimension for the model is 140 167 m2. This cavern has the horse-shoe cross-section with
20 m in width and 27.5 m in height. The water level is located at 0m, and the crown of the cavern is
located at -119m. Parameters used for the simulation are shown in Table 2.
Figure 7. The cross section of the single cavern after excavation for the simulation.

Table 2. Parameters for the coupled DDA hydro-mechanical analysis.

Two grouting zone thicknesses are considered, including 2.5m and 5m, respectively. To model the
grouting effect, the fracture aperture within the grouting area (Figure 8) will be minimized to achieve a
certain hydraulic conductivity. In this case study, the hydraulic conductivity in the grouting zone is
5
1107 m/s, and hence, the aperture is decreased to 7.2  10 m for all the fractures within the grouting
zone. For the none-grouting zone, the hydraulic conductivities are 1106 m/s and 1105 m/s,
4 4
respectively, and the corresponding fracture apertures are 1.6  10 m and 3.4  10 m. The fracture
deformation within the grouting zone is not updated durng the coupled analysis. Only the deformation
of the aperture in the non-grouting is considered.

(a) (b)

Figure 8. The cross section for the grouting zone in the discrete fracture model, with (a) 2.5 m thickness; (b) 5.0
m thickness.

The inflow rate is obtained at 120s when the water flow nearly becomes steady. Figure 9 indicates that
the inflow rate becomes smaller by increasing the thickness of the grouting zone.

The water flow becomes almost steady at 120s, and the hydraulic head around the cavern has
dissipated to a small value. When the hydraulic conductivity in the non-grouting zone is 1.0 106 m/s,
the hydraulic head dissipation is more obvious in the 2.5m grouting zone case (Figures 10(a) and
11(a)). Also, the hydraulic head above the cavern roof drops significantly (Figure 10(a)). When the
conductivity of the none-grouting zone is increased to 1.0 105 m/s, the grouting zone is surrounded
by the high hydraulic head that has less dissipation (Figures 10(b) and 11(b)). In particular, this
hydraulic head distribution is quite similar with the results obtained by the cavern modeling
considering the lining zone. Fernandez (1994) considered that if the conductivity of the non-grouting
zone is relatively large, which is 100 times larger than the one in the grouting zone, the lining zone can
be treated as relative impermeable. In this situation, the hydraulic head lost across the lining zone can
be so large (Fernandez, 1994). Based on his observation, the results from this modeling are reasonable.

Figure 9. The inflow rate for the cavern with considering the grouting zone at 120s.

(a) (b)
6
Figure 10. The hydraulic head distribution at 120s with 2.5m grouting zone, with (a) K ( S )  1  10 m/s; (b)
5
K ( S )  1  10 m/s.
(a) (b)
6
Figure 11. The hydraulic head distribution at 120s with 5m grouting zone, with (a) K ( S )  1  10 m/s; (b)
5
K ( S )  1  10 m/s.

4 CONCLUSIONS

This paper presents the DDA modelling for three engineering applications, and the results verify the
effectiveness of the DDA in solving the practical rock engineering. The special feature of the DDA
method makes it an attractive approach for fractured rock mass modelling, as the discontinuities can be
modelled explicitly.

REFERENCES

Bao H.R., 2007. Discontinuous Deformation Analysis with Nodal Displacements as Unknowns. Ph. D. First year
report. Department of Civil Engineering, Nanyang Technological University.
Fernandez G., 1994. Behavior of Pressure Tunnels and Guidelines for Liner Design. Journal of Geotechnical
Engineering, 120: 1768-1791.
Fu X.D., Sheng Q., Zhang Y.H., Chen J., 2015. Application of the Discontinuous Deformation Analysis Method
to Stress Wave Propagation through a One-dimensional Rock Mass. International Journal of Rock Mechanics &
Mining Sciences, 80:155-170.
Hatzor Y.H., Arzi A.A., Zaslavsky Y., Shapira A., 2004. Dynamic Stability Analysis of Jointed Rock Slopes
Using the DDA Method: King Herod’s Palace, Masada, Israel. International Journal of Rock Mechanics &
Mining Sciences, 41: 813-832.
Hoek E., 1983. Strength of Jointed Rock Masses. Geotechnique, 23: 187-223.
Jing L., 2003. A Review of Techniques, Advances and Outstanding Issues in Numerical Modeling for Rock
Mechanics and Rock Engineering. International Journal of Rock Mechanics & Mining Sciences, 40: 283-353.
Li C.C., 2008. Laboratory Testing and Performance of Rock Bolts. 37. Geomechanick-Kolloquium. Freberg, 47-
48. ISSN 1611-1605.
Li C.C., 2010. A New Energy-absorbing Bolt for Rock Support in High Stress Rock Masses. International
Journal of Rock Mechanics & Mining Sciences, 47: 396-404.
Li C.C., 2012. Performance of D-bolts under Static loading. Rock mechanics and Rock Engineering, 45: 183-192.
MacLaughlin, M.M., Sitar, N., Doolin, D.M. & Abbot, T. 2001. Investigation of Slope-Stability Kinematics
using Discontinuous Deformation Analysis. International Journal of Rock Mechanics & Mining Sciences, 38:
753-762.
Nie W., Zhao Z.Y., Ning Y.J., Guo W., 2014. Numerical Studies on Rockbolts Mechanism using 2D
Discontinuous Deformation Analysis. Tunnelling and Underground Space Technology, 41: 223-233.
Ning Y.J., Zhao Z.Y., Sun J.P., Yuan W.F., 2012. Using the Discontinuous Deformation Analysis to Model
Wave Propagations in Jointed Rock Masses. CEMS, 2329: 1-42.
Shi G.H., 1988. Discontinuous Deformation Analysis: A New Numerical Model for the Static and Dynamics of
Block Systems [Thesis]. Type, UC Berkeley, Berkeley.
Sun J.P., Ning Y.J., Zhao Z.Y., 2011a. Comparative Study of Sarma's Method and the Discontinuous
Deformation Analysis for Rock Slope Stability Analysis. Geomechanics and Geoengineering, 6: 293-302.
Wu Y.Q., 2009. Geohydraulics, 1st ed., Beijing Science and Technology Press. (Chinese)

You might also like