Sunteți pe pagina 1din 40

STRUCTURAL ELEMENTS

3-1

3 STRUCTURAL ELEMENTS
3.1 Introduction
An important aspect of geomechanical analysis and design is the use of structural support to stabilize
a soil or rock mass. The term support describes engineered materials used to restrict displacements
in the immediate vicinity of an opening. In this section, we focus on support provided to reinforce
the soil or rock. The structural element formulations that represent reinforcement support are
described.*
Reinforcement consists of tendons (i.e., cables) or bolts installed in holes drilled in the rock mass.
Reinforcement acts to conserve inherent rock mass strength so that it becomes self-supporting. Two
types of reinforcement model are provided in 3DEC: local and global. A local reinforcement model
considers only the local effect of reinforcement where it passes through existing discontinuities.
A global reinforcement model considers the presence of the reinforcement along its entire length
throughout the rock mass. Reinforcement may also be applied as beams. Beams are structural
elements which are connected to the rock surface by nodes.
The reinforcement types are described separately, below. In all cases, the commands necessary to
define the structure(s) are quite simple, but they invoke a very powerful and flexible structural logic.
This logic is developed with the same finite-difference algorithms as the rest of the code (as opposed
to a matrix-solution approach), allowing the structure to accommodate large displacements, and to
be applied for dynamic as well as static analysis. (See the dynamic analysis option discussed in
Section 2 in Optional Features.)

* Another type of support, surface support, is also provided as an optional feature in 3DEC. Surface
support elements can simulate, for example, tunnel lining; this support type is described in Section 3
in Optional Features.

3DEC Version 5.0

3-2

Theory and Background

3.2 Rock Reinforcement


There are several different types of reinforcement designed to operate effectively in a range of
ground conditions. One type is represented by a reinforcing bar or bolt fully encapsulated in a
strong, stiff resin or grout. This system is characterized by the relatively large axial resistance
to extensions that can be developed over a relatively short length of the shank of the bolt, and
by the high resistance to shear that can be developed by an element penetrating a slipping joint.
A second type of reinforcement system, represented by cement-grouted cables or tendons, offers
little resistance to joint shear, and development of full axial load may require deformation of the
grout over a substantial length of the reinforcing element. These two types of reinforcement are
identified, respectively, as local reinforcement and global (or spatially extensive) reinforcement.
The characteristic behavior of these two rock reinforcement systems has been incorporated into
3DEC.
3.2.1 Local Reinforcement at Joints (STRUCT axial Command)
The local reinforcement formulation considers only the local effect of reinforcement where it passes
through existing discontinuities. The formulation results from observations of laboratory tests of
fully grouted untensioned reinforcement in good quality rocks with one discontinuity, which indicate
that strains in the reinforcement are concentrated across the discontinuity (Bjurstrom 1974 and Pells
1974). This behavior can be achieved in the computational model by calculating, for each element,
the forces generated by displacements across the discontinuity through which the element passes.
The following description of this formulation is taken from Lorig (1985).
This formulation exploits simple force-displacement relations to describe both the shear
and axial behavior of reinforcement across discontinuities. Large shear displacements
are accommodated by considering the simple geometric changes that develop locally in
the reinforcement near a discontinuity. Although the local reinforcement model can be
used with either rigid blocks or deformable blocks, the representation is most applicable
to cases in which deformation of individual rock blocks may be neglected in comparison
with deformation of the reinforcing system. In such cases, attention may be focused
reasonably on the effect of reinforcement near discontinuities.
3.2.1.1 Axial Behavior
Historically, axial testing of rock reinforcement has focused on pull-out tests for two reasons:
(1) ease of experimentation and interpretation of results; and
(2) provision of axial restraint (the main function of reinforcement in the prevailing
conceptual models).
Consequently, a relatively good understanding of axial force-displacement relations has been
achieved. The axial force-displacement relation typically used in the representation of rock reinforcement is shown in Figure 3.1.

3DEC Version 5.0

STRUCTURAL ELEMENTS

3-3

tension

Pult

rupture
Ka

axial
displacement

load reversal

Figure 3.1

Axial behavior of local reinforcement systems

Figure 3.1 indicates an identical response in tension and compression. This may not be the case for
all reinforcing systems.
If pull-test results are not available, the following theoretical expression (given by Gerdeen et al.
1977) may be used to estimate the axial stiffness, Ka , for fully bonded solid reinforcing elements:
Ka = k d1

(3.1)

where:d1 = reinforcement diameter;


k = [ 21 Gg Eb / (d2 / d1 1)]1/2 ;
Gg = grout shear modulus;
Eb = Youngs modulus of reinforcement material; and
d2 = hole diameter.
Comparisons with finite element analyses (Gerdeen et al. 1977) indicate that Eq. (3.1) tends to
slightly overestimate axial stiffnesses.
The ultimate axial capacity of the reinforcement depends on a number of factors, including strength
of the reinforcing element, bond strength, hole roughness, grout strength, rock strength and hole
diameter. In the absence of results of physical tests, empirical relations may be used to estimate the

3DEC Version 5.0

3-4

Theory and Background

ultimate anchorage strength, Pult . One such relation for the design of cement-grouted reinforcement
is given by Littlejohn and Bruce (1975):
Pult = 0.1 c d2 L

(3.2)

where:c = uniaxial compressive strength of massive rocks (100% core recovery) up to a


maximum value of 42 MPa, assuming that the compressive strength of the
cement grout is equal to or greater than 42 MPa; and
L = bond length.
3.2.1.2 Shear Behavior
Recognition that reinforcement also acts to modify the shear stiffness and strength of discontinuities has led to laboratory shear testing of reinforced discontinuities. Experimental results and
theoretical investigations indicate that shearing along a discontinuity induces bending stresses in the
reinforcement that decay very rapidly with distance into the rock from the shear surface. Typically,
the bending stresses are insignificant within one to two reinforcing element diameters.
The shear force-displacement relation typically used to represent shear behavior is shown in Figure 3.2. The figure shows representative responses for reinforcement at various altitudes with
respect to the traversed discontinuity and direction of shear.
If the results of physical tests are not available, the shear stiffness, Ks , may be estimated using the
following expression from Gerdeen et al. (1977):
Ks = Eb I 3
where: = [K / (4 Eb I )]1/4 ;
K = 2Eg / (d2 /d1 1);
I

= second moment of area of the reinforcement element; and

Eg =Youngs modulus of the grout.

3DEC Version 5.0

(3.3)

STRUCTURAL ELEMENTS

3-5

shear
force
0

Ks
1
0= 45

rupture

0= 90
0= 135

Figure 3.2

shear
displacement

Shear behavior of reinforcement system

max , for a reinforcement elEmpirical relations can be used to estimate the maximum shear force, Fs,b
ement at various orientations with respect to a transgressed discontinuity and direction of shear. For
example, Bjurstrom (1974) used the results of shear tests of ungrouted reinforcement perpendicular
to a discontinuity in granite to develop the expression
max
Fs,b
= 0.67 d12 (b c )1/2

(3.4)

where b = yield strength of reinforcement.


In their assessment of maximum shear resistance, St. John and Van Dillen (1983) applied the results
of Azuar et al. (1979). The latter found that the maximum shear force was about half the product
of the uniaxial tensile strength of the reinforcement and its cross-sectional area for reinforcement
perpendicular to the discontinuity. The force increased to 80-90% of that product for reinforcement
inclined with the direction of shear. Shear displacements causing rupture were reported after
approximately two reinforcement diameters for the perpendicular case, and one diameter for the
inclined case. St. John and Van Dillen interpreted differences between strength and amount of
displacement before rupture in terms of the extent of crushing of rock around the reinforcement.

3DEC Version 5.0

3-6

Theory and Background

3.2.1.3 Numerical Formulation


The model in 3DEC assumes that, during shear displacement along a discontinuity, the reinforcement deforms as shown in Figure 3.3. The short length of reinforcement that spans the discontinuity and changes orientation during shear displacement is referred to as the active length. The
assumed geometric changes were originally suggested in a derivation by Haas (1976) for conventional point-anchored reinforcement, and adopted by Fuller and Cox (1978) in considering fully
grouted reinforcement.

Direction of
Shearing

Discontinuity

e
tiv h
c
A gt
n
Le

us
Figure 3.3

(Positive)

Assumed reinforcement geometry after shear displacement, us

It is assumed that the active length changes orientation only as a direct geometric result of shear and
normal displacements at the discontinuity. Methods for estimating the active length are presented in
the next section. The model may be considered to consist of two springs located at the discontinuity
interface and oriented parallel and perpendicular to the reinforcement axis, as shown in Figure 3.4.
Following shear displacement, the axial spring is oriented parallel to the active length, while the
shear spring remains perpendicular to the original orientation, as shown in Figure 3.4. Similar
geometric changes follow displacements normal to the discontinuity.

3DEC Version 5.0

STRUCTURAL ELEMENTS

3-7

Axial Spring

Discontinuity

Shear Spring

Direction of
Shearing

Axial Spring

Discontinuity

Shear Spring

Figure 3.4

Orientation of shear and axial springs representing reinforcement prior to and after shear displacement

The force-displacement models used in 3DEC to represent axial and shear behavior are continuous
linear algorithms written in terms of stiffness (axial or shear) and the ultimate load capacity. Note that
the formulation for local reinforcement in 3DEC is a simplified version of the general formulation
implemented in UDEC (see Itasca 2011).

3DEC Version 5.0

3-8

Theory and Background

The force-displacement relation that describes the axial response is given by


Fa = Ka |ua |

(3.5)

where:Fa is an incremental change in axial force;


ua is an incremental change in axial displacement; and
Ka
is the axial stiffness.
In this formulation, no reduction in stiffness is made to account for crushing of the grout and/or
rock near the discontinuity.
A rupture limit can be specified for the axial force. If the extensional axial strain in the reinforcement
element exceeds a specified strain limit, then the axial force will be set to zero.
The shear force-displacement relation is described in incremental form by the expression
Fs = Ks |us |

(3.6)

where:Fs is an incremental change in shear force;


us is an incremental change in shear displacement; and
Ks
is the shear stiffness.
A rupture limit can also be specified for the shear force. If the relative shear strain in the reinforcement element exceeds a specified strain limit, then the shear force will be set to zero.
The force-displacement relations described above are used to determine forces in the springs arising
from incremental displacements at the endpoints of the active length. The resultant shear and axial
forces are resolved into components parallel and perpendicular to the discontinuity, as shown in
Figure 3.5. Forces are then applied to the neighboring blocks.

3DEC Version 5.0

STRUCTURAL ELEMENTS

3-9

Direction of

Shear Displacement
of Reinforcement

Shearing

Resultant Shear Force

Discontinuity

0
Normal Force
Applied to
Upper Block

Direction of

Shear Force Applied


to Upper Block

Resultant Axial Force

Shearing
Normal Force Applied
to Upper Block

Discontinuity

Shear Force Applied


to Upper Block

Figure 3.5

Resolution of reinforcement shear and axial forces into components parallel and perpendicular to discontinuity

3DEC Version 5.0

3 - 10

Theory and Background

3.2.1.4 Estimation of Active Length


To define the assumed local deformation illustrated in Figure 3.3, an estimate of the active length
is required. It has been shown that the active length extends approximately one to two reinforcing
element diameters on either side of the discontinuity. In the absence of experimental data, results
of theoretical analysis may be used to define the active length. For example, in defining the elastic
shear stiffness, Ks , Gerdeen et al. (1977) also determine a quantity, l, called the load transfer
length, or decay length. If max is the proportion of maximum deflection in the reinforcement,
the relation between it and the load transfer length may be expressed by
e l = max

(3.7)

For example, the point at which the deflection decays to 5% of its maximum value is
e l = 0.05
or
l = 3 / .
This approach was developed for reinforcement oriented perpendicular to the shear plane. Dight
(1982) presents a theoretical analysis for determining the distance from the shear plane to maximum
moment which corresponds with the location of the plastic hinge in the reinforcement element. This
approach places no restrictions on the orientation of the reinforcement with respect to the shear
plane. A significant result of this analysis is that the distance of the plastic hinge from the shear
plane does not appear to vary greatly with shear displacement, especially for displacements greater
than 10 mm (0.4 in) for typical reinforcement systems. This observation is in agreement with the
assumed geometry changes described earlier.

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 11

3.2.1.5 Local Reinforcement Properties


The local reinforcement elements used in 3DEC require the following input parameters:
(1) axial stiffness [force/length];
(2) ultimate axial capacity [force];
(3) 1/2 active length [length];
(4) extensional failure strain (default = infinite) [ ]; and
(5) shear failure strain (default = infinite) [ ].
The axial stiffness and ultimate axial capacity are usually determined to best-fit pull-out tests as
described in Section 3.2.1.1. The axial force-displacement relation follows a constant axial stiffness
until the ultimate axial capacity is reached.
The value for 1/2 the active length can also be back-calculated from experimental testing, as discussed in Sections 3.2.1.2 and 3.2.1.4.
3.2.1.6 Summary of Commands Associated with Local Reinforcement Elements
All of the commands associated with local reinforcement elements are listed in Table 3.1. See
Section 1.3 in the Command Reference for a detailed explanation of these commands.
Table 3.1

Commands associated with local reinforcement elements

STRUCTURE

axial

x1 y1 z1

x2 y2 z2

STRUCTURE

prop np

keyword
rkax
rlen
rsstiffness
rsstrain
rstrain
rult
rushear

value
value
value
value
value
value
value

PLOT

keyword
aforce
bolt

LIST

axial

prop

np

3DEC Version 5.0

3 - 12

Theory and Background

3.2.1.7 Example Application Reinforced Slope


The stability analysis of the rock slope described in the tutorial in Section 2.2 in the Users Guide is
repeated using the local reinforcement elements to represent rock anchors. The slope is stabilized
by adding two horizontal lines of local reinforcement. One line extends from location (30,40,25) to
location (70,40,25), and the other from location (30,20,25) to location (70,20,25). When the joint
friction is reduced, the slope is then stable, as indicated by Figure 3.6. A history of y-displacement
at a point on the slope face also shows that motion is restrained by the reinforcement (see Figure 3.7).
The input commands for this example are listed in Example 3.1.
Local damping (DAMPING local) is applied when the joint friction is reduced in this example. Note
that a slightly different result will be calculated for the z-displacement if DAMPING auto is used.
This can be attributed to the way the different damping schemes apply the damping force to the
velocities in the model (see Section 1.2.3.2). If the friction is reduced in small increments, or the
excavation of the slope is made in stages, then the inertial effects induced by sudden loading changes
can be reduced, and the different damping methods will approach the same solution.
Example 3.1 Reinforced slope
new
poly brick 0,80 -30,80 0,50
plot block
plot reset
plot set dip 70 dd 210
; boundary blocks on sides of slope
jset dip 90 dd 180 org 0,0,0
jset dip 90 dd 180 org 0,50,0
hide range x 0,80 y -30,0 z 0,50
hide range x 0,80 y 50,80 z 0,50
mark region 1
; shallow-dipping fracture planes
jset dip 2.45 dd 235 org 30,0,12.5
jset dip 2.45 dd 315 org 35,0,30
; high angle foliation planes
jset dip 76 dd 270 spac 4 num 5 org 38,0,12.5
hide range x 30,80 y 0,50 z 0,50
jset dip 0 dd 0 org 0,0,10
hide range x 0,80 y -30,80 z 0,10
mark region 2
seek
hide range region 0
hide range x 0,80 y 0,50 z 0,10
hide range x 55,80 y 0,50 z 0,50
hide range x 0,30 y 0,50 z 0,50
; intersecting discontinuities
jset dip 70 dd 200 org 0,35,0

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 13

jset dip 60 dd 330 org 50,15,50


seek
; fix boundary blocks and initialize gravity
fix range x 0,80 y 0,50 z 0,10
fix range x 55,80 y 0,50 z 0,50
fix range region 0
hide range region 0
delete range region 2
gravity 0 0 -10
; assign properties
prop m=1 de=2000 kn=1e9 ks=1e9 f=100.
prop m=2 kn=1e9 ks=1e9 f=0.0
change jmat=2 range plane dip 90 dd 180
; prepare to cycle
hist zvel 30,30,30 ty=1
cyc 500
save slope.sav
; add local reinforcement
struct axial 30,25,40 70,25,40 prop 1
struct axial 30,25,20 70,25,20 prop 1
struct prop 1 rkax = 1e8 rlen = 1.0 rult = 1e8
damp local
prop jmat=1 f=6.0
reset hist time disp
hist zdisp 30,30,30 ty=1
hist unbal
step 2500
title
ROCK SLOPE STABILITY --- WEDGE STABLE WITH LOCAL REINFORCEMENT
plot block disp line color cyan
pause key
hist label 1 Vertical Displacement
plot hist 1 yaxis label Vertical Displacement
save slope_st.sav
ret

3DEC Version 5.0

3 - 14

Theory and Background

E

pGr,00U 
 0
u0 i

0,GGG
7gr/gpGr,0rp:.e:SG0



U :0 
r

0U 
 0
u0 i
  u000y

Figure 3.6

Stabilization of slope by reinforcement

E



pGr,00A 
 0
u0 i

0,GGG
pgr8gpGr,0r5pp5S80





r0  0   

i0



e
  e nm lH











0A 
 0
u0 i
  u000N

Figure 3.7

3DEC Version 5.0

            


e nm H

z-displacement history at (30, 30, 30) on slope face

STRUCTURAL ELEMENTS

3 - 15

3.2.2 Global Reinforcement (STRUCT cable Command)


In assessing the support provided by rock reinforcement, it is often necessary to consider not
only the local restraint provided by reinforcement where it crosses discontinuities, but also the
restraint to intact rock that may experience inelastic deformation in the failed region surrounding
an excavation. Such situations arise in modeling inelastic deformations associated with failed rock
and/or reinforcement systems (e.g., cable bolts) in which the bonding agent (grout) may fail in shear
over some length of the reinforcement.
Cable elements in 3DEC allow the modeling of a shearing resistance along their length, as provided
by the shear resistance (bond) between the grout and either the cable or the host medium. The cable
is assumed to be divided into a number of segments of length, L, with nodal points located at each
segment end. The mass of each segment is lumped at the nodal points, as shown in Figure 3.8.
Shearing resistance is represented by spring/slider connections between the structural nodes and
the block zones in which the nodes are located.
To use the cable logic, deformable blocks must be specified, and the blocks must be made deformable
before the cable elements are installed. Each structural node is associated with a finite difference
zone for calculation of shear forces between the cable and the zones.
reinforcing
element (steel)

grout annulus

Excavation

reinforcement
nodal point
slider (cohesive
strength of grout)
axial stiffness of steel
shear stiffness of grout

Figure 3.8

Conceptual mechanical representation of fully bonded


reinforcement, which accounts for shear behavior of the grout
annulus

3DEC Version 5.0

3 - 16

Theory and Background

3.2.2.1 Axial Behavior


The axial behavior of conventional reinforcement systems may be assumed to be governed entirely
by the reinforcing element itself. The reinforcing element is usually steel, and may be either a bar
or cable. Because the reinforcing element is slender, it offers little bending resistance (particularly
in the case of a cable), and is treated as a one-dimensional member subject to uniaxial tension or
compression. A one-dimensional constitutive model is adequate for describing the axial behavior
of the reinforcing element. In the present formulation, the axial stiffness is described in terms of
the reinforcement cross-sectional area, A (area), and the Youngs modulus, E (E).
The incremental axial force, F t , is calculated from the incremental axial displacement by
F t =

EA
ut
L

(3.8)

where: ut = ui ti


= u1 t1 + u2 t2 + u3 t3
[a]
[b]
[a]
[b]
[a]
= (u[b]
1 u1 )t1 + (u2 u2 )t2 + (u3 u3 )t3
[b]
u[a]
1 , u1 , etc. are the displacements at the cable nodes associated with each cable element. Subscript
1 corresponds to the x-direction, subscript 2 to the z-direction, and subscript 3 to the z-direction.
The superscripts [a], [b] refer to the nodes. The direction cosines t1 , t2 and t3 refer to the tangential
(axial) direction of the cable segment.

A tensile yield force limit (yield) can be assigned to the cable. Accordingly, cable forces that are
greater than the tensile limit (Figure 3.9) cannot develop. If yield is not specified, the cable will
have infinite strength for loading in tension.
In evaluating the axial forces that develop in the reinforcement, displacements are computed at
nodal points along the axis of the reinforcement, as shown in Figure 3.8. Out-of-balance forces at
each nodal point are computed from axial forces in the reinforcement, as well as from shear forces
contributed through shear interaction along the grout annulus. Axial displacements are computed
based on integration of the laws of motion using the computed out-of-balance axial force and a
mass lumped at each nodal point.

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 17

- =HA=

Figure 3.9

Cable material behavior for cable elements

3.2.2.2 Shear Behavior of Grout Annulus


The shear behavior of the grout annulus is represented as a spring-slider system located at the nodal
points shown in Figure 3.8. The shear behavior of the grout annulus, during relative displacement
between the reinforcing/grout interface and the grout/rock interface, is described numerically by
the grout shear stiffness kbond in (Figure 3.10):
Fs
= Kbond (uc um )
L
where:

Fs
Kbond
uc
um
L

(3.9)

= shear force that develops in the grout


(i.e., along the interface between the cable element and the grid);
= grout shear stiffness (kbond);
= axial displacement of the cable;
= axial displacement of the medium (soil or rock); and
= contributing element length.

3DEC Version 5.0

3 - 18

Theory and Background

force/length

Fsmax
L
kbond
1
relative shear
displacement

Fsmax
L

Figure 3.10 Grout material behavior for cable elements


The maximum shear force that can be developed per length of element, Fsmax /L, is limited by
the cohesive strength of the grout (property keyword sbond). The limiting shear-force relation is
depicted by the diagram in Figure 3.10.
Calculation of the relative displacement at the grout/rock interface uses an interpolation scheme to
compute the displacement of the rock in the cable axial direction at the cable node. Each cable
node is assumed to exist within an individual 3DEC tetrahedral zone (hereafter referred to as host
zone). The interpolation scheme uses weighting factors which are based on the distance to each
of the gridpoints of the host zone. The calculation of the weighting factors is based on satisfying
moment equilibrium.
For example, in computing the axial displacement of the grout/rock interface, the following interpolation scheme is used. Consider reinforcement passing through a constant-strain finite difference
tetrahedron making up part of the intact rock, as shown in Figure 3.11(a). The incremental xcomponent of displacement (uxp ) at the nodal point is given by
uxp = W1 ux1 + W2 ux2 + W3 ux3 + W4 ux4
where:ux1 , ux2 , ux3
W1 , W2 , W3 , W4

3DEC Version 5.0

are the incremental gridpoint displacements; and


are weighting factors.

(3.10)

STRUCTURAL ELEMENTS

3 - 19

A similar expression is used for z-component displacements. The weighting factors W1 , W2 , W3 ,


W4 are computed from the position of the nodal point within the tetrahedron:
W1 = V1 / VT

(3.11)

where:VT is the total volume of the finite-difference tetrahedron; and


V1 is the volume of the tetrahedron in Figure 3.11(b).
Incremental x-, y- and z-displacements (Eq. (3.10)) are used at each calculation step to determine the
new local reinforcing orientation. The axial component of displacement of the grout/rock interface
is computed from the current orientation of the reinforcing segment.
Forces generated at the grout/rock interface (Fxp , Fyp , Fzp ) are distributed back to gridpoints
according to the same weighting factors used previously:
Fx1 = W1 Fxp
Fx2 = W2 Fxp

(3.12)

Fx3 = W3 Fxp
Fx4 = W4 Fxp
where Fx1 , Fx2 , Fx3 and Fx4 are forces applied to the gridpoints.

3DEC Version 5.0

3 - 20

Theory and Background

Gridpoint

Constant Strain
Finite Difference
Tetrahedron

"

Reinforcement
Nodal Point

(a) typical reinforcing element passing through a tetrahedral zone

V1

V3
V4


"

V2
!

(b) volumes used in determining weighting factors used to compute


displacement of grout/rock interface
Figure 3.11 Geometry of tetrahedral finite difference zone and transgressing
reinforcement used in distinct element formulation

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 21

3.2.2.3 Normal Behavior at Grout Interface


As explained above, an interpolated estimate of gridpoint velocity is made at each cable node. The
velocity component normal to the average axial cable direction is transferred directly to the node
(i.e., the cable node is slaved to the gridpoint motion in the normal direction). The node exerts
no normal force on the grid if the cable segments on either side of the node are colinear. However,
if the segments make an angle with each other, then a proportion of their axial forces will act in
the mean normal direction. This net force acts both on the gridpoint and on the cable node (in
opposite directions). Thus, an initially straight cable can sustain normal loading if it undergoes
finite deflection.
3.2.2.4 Cable Element Properties
The cable elements used in 3DEC require the following input parameters:
(1) cross-sectional area of cable;
(2) elastic Youngs modulus for cable;
(3) tensile yield strength [force] of the cable;
(4) stiffness of the grout [force/cable length/displacement];
(5) cohesive capacity of the grout [force/cable length]; and
(6) compressive yield strength of the cable [force].
Note that property numbers are assigned to cable elements with the STRUCT cable . . . prop np
command, where np is the material property number for the cable and the grout. Each different
cable can then be assigned geometric and material properties by specifying the STRUCT prop
command with the appropriate property keywords following the cable material property number.
For example,
struct prop=2 kbond = 1e9 sbond = 2e5

assigns a cable bond stiffness value of 109 and a cable bond strength value of 2 105 to property
number 2.
The area, modulus and yield force resistance of the cable are usually readily available from handbooks, manufacturers specifications, etc.
The properties related to the grout are more difficult to estimate. The grout annulus is assumed to
behave as an elastic perfectly plastic solid. As a result of an incremental relative shear displacement,
ut , between the tendon surface and the borehole surface, the incremental shear force, F t ,
mobilized per length of cable is related to the grout stiffness, Kbond :
F t = Kbond ut

(3.13)

3DEC Version 5.0

3 - 22

Theory and Background

Kbond can be estimated from pull-out tests. Alternatively, the stiffness can be calculated from a
numerical estimate for the elastic shear stress, G , obtained from an equation describing the shear
stress at the grout/rock interface (St. John and Van Dillen 1983):
G =
where:

u
G
(D/2 + t) ln(1 + 2t/D)

(3.14)

G = grout shear modulus;


D = reinforcing diameter; and
t = annulus thickness.

Consequently, the grout shear stiffness, Kbond , is simply given by


Kbond =

2 G
ln (1 + 2t/D)

(3.15)

In many cases, the following expression has been found to provide a reasonable estimate of Kbond
for use in 3DEC:
Kbond 

2 G
10 ln (1 + 2t/D)

(3.16)

The one-tenth factor helps to account for the relative shear displacement that occurs between the
host-zone gridpoints and the borehole surface. This relative shear displacement is not accounted
for in the present formulation.
The maximum shear force per cable length in the grout is the bond cohesive strength. The value
for bond cohesive strength can be estimated from the results of pull-out tests conducted at different
confining pressures, or, should such results not be available, the maximum force per length may be
approximated from the peak shear strength (St. John and Van Dillen 1983):
peak = I QB

(3.17)

where I is approximately one-half of the uniaxial compressive strength of the weaker of the rock
and grout, and QB is the quality of the bond between the grout and rock (QB = 1 for perfect
bonding).
Neglecting frictional confinement effects, Sbond may then be obtained from
Sbond = (D + 2t) peak

3DEC Version 5.0

(3.18)

STRUCTURAL ELEMENTS

3 - 23

Failure of reinforcing systems does not always occur at the grout/rock interface. Failure may occur
at the reinforcing/grout interface, as is often true for cable reinforcing. In such cases, the shear stress
should be evaluated at this interface. This means that the expressions (D + 2t) will be replaced
with (D) in Eq. (3.18).
The calculation of cable-element properties is demonstrated by the following example. A 25.4
mm (1 inch)-diameter locked-coil cable was installed at 2.5 m spacing. The reinforcing system is
characterized by the following properties:
cable diameter (D)
hole diameter (D + 2t)
cable modulus (E)
cable ultimate tensile capacity
grout compressive strength
grout shear modulus (G)
cross-sectional area (area)

25.4 mm
38 mm
98.6 GPa
0.548 MN
20 MPa
9 GPa
5 104 m2

Two independent methods are used in evaluating the maximum shear force in the grout. In the
first method, the bond shear strength is assumed to be one-half the uniaxial compressive strength
of the grout. If the grout-material compressive strength is 20 MPa and the grout is weaker than the
surrounding rock, the grout shear strength is then 10 MPa.
In the second method, reported pull-out data are used to estimate the grout shear strength. The
report presents results for 15.9 mm (5/8 inch) diameter steel cables grouted with a 0.15 m (5.9
inch) bond length in holes of varying depths. The testing indicates capacities of roughly 70 kN.
If a surface area of 0.0075 m2 (0.15 m 0.05 m) is assumed for the cables, then the calculated
maximum shear strength of the grout is
70 x 103 N
= 9.33 x 106 N/m2 = 9.33 MPa
0.0075 m2
This value agrees closely with the 10 MPa estimated above, and either value could be used. Assuming failure occurs at the cable/grout interface, the maximum bond force per length is (using
Eq. (3.18) with D + 2t replaced by D)
Sbond = (0.0254 m) (10 MPa) = 800 kN/m
The bond stiffness, Kbond , is estimated from Eq. (3.16). For the assumed values shown above, a
bond stiffness of 1.5 1010 N/m
m is calculated.
Mass scaling is automatically performed to adjust the cable mass to achieve a timestep that coincides
with that calculated for the 3DEC model without cables. If the grout stiffness or axial modulus of
the cable element is very high, it may be necessary to reduce the timestep (using the FRACTION
command) to avoid numerical instability errors.

3DEC Version 5.0

3 - 24

Theory and Background

3.2.2.5 Pretensioning Cable Elements


Cable elements may be pretensioned in 3DEC by assigning the optional value preten with the
STRUCT cable . . . tens command. A positive value for pretensioning assigns an axial force into the
cable-element(s) described by that STRUCT cable command. It is important to note that, initially,
the cable with specified pretension is unlikely to be in equilibrium with other elements of the 3DEC
model to which it is linked. In other words, some displacement of the cable nodes and linked blocks
is probably required to achieve equilibrium. These displacements will likely result in some loss of
the initial pretension.
3.2.2.6 Estimating the Maximum Length for Cable Element Segments
If the segment lengths for a cable element are too long, failure by grout slippage cannot occur
because the cable material yield force will be reached before the shear resistance along the grout
can be mobilized. It is recommended that at least two or three segments be provided along the
development length of a cable in order to account for the possibility of grout slippage. The
development length is the length of bonded reinforcing required to develop the axial capacity of
the cable. The development length, ld , can be estimated from the grout bond strength, Sbond , and
the cable yield force capacity, Fy , using the expression
ld =

Fy

(3.19)

Sbond

3.2.2.7 Summary of Commands Associated with Cable Elements


All of the commands associated with local reinforcement elements are listed in Table 3.2. See
Section 1.3 in the Command Reference for a detailed explanation of these commands.
Table 3.2

Commands associated with cable elements

STRUCT

cable

x1 y1 z1

x2 y2 z2

STRUCT

prop np

keyword
area
e
kbond
sbond
ycomp
yield

value
value
value
value
value
value

PLOT

keyword
cforce
cable

LIST

cable

3DEC Version 5.0

seg ns

prop np

tens preten

STRUCTURAL ELEMENTS

3 - 25

3.2.2.8 Example Application Pull-Test for a Grouted Cable Anchor


The most common way to determine cable bolt properties is to perform pull-out tests on small
segments of grouted cables in the field. Typically, segments from 10 to 50 cm in length are grouted
into boreholes. The ends of these segments are pulled with a jack mounted to the surface of the
tunnel, and connected to the cable via a barrel-and-wedge type anchor. The force applied to the
cable and the deformation of the cable are plotted to produce an axial force/deflection curve. From
this curve, the peak shear strength of the grout bond is normally determined and converted to a
strength in tons/m cable length.
In this example, we simulate a pull-test on a single 15.2 mm diameter cable. The cable material
properties are:
cable area
cable bond length
cable modulus (E)
cable ultimate tensile capacity

181 mm2
17.7 m
98.6 GPa
0.5 MN

We select the following properties for the grout:


grout bond stiffness
grout cohesive strength

1.12 108 N/m/m


1.75 105 N/m

These values are representative of a weak grout (e.g., see Hyett et al. 1992). Note that the cable
length and grout strength are selected for rapid execution of the model. The purpose is to illustrate
the performance of the cable elements.
We apply a load to the cable by gluing a small block to the end of the cable; we can pull the cable
by pulling the block. The cable is divided into 10 segments with one segment located inside the
loading block. The cable is attached to the loading block by assigning a high grout shear strength
to the cable nodes, and a high cable tensile capacity to the cable element embedded in the loading
block. The properties are changed for the cable nodes and element in the loading block by changing
the property number assigned to this cable element. The FISH function change mat changes
the cable element property number for this element. Use the LIST cable command to identify the
address of the cable element located in the loading block.
FISH function pullf is used to monitor the pull force on the cable (pull force per cable length).
The pull force is determined from the sum of reaction forces that develop on the block as the cable
is pulled. Note that the reaction force is set to zero at the end of cycling, so the pull force calculation
in pullf ends one step before cycling stops. Example 3.2 gives the data file for this model.

3DEC Version 5.0

3 - 26

Theory and Background

Example 3.2 Simulation of a pull-test for a grouted cable anchor


new
poly brick -1,1 0,2 -1,1
poly brick -.1,.1 -.27,-.02 -.1,.1
;
loading block
pl bl
pl reset
plot set dip 60 dd 110
gen edge .5 range x -1 1 y 0 2 z -1 1
gen edge .1
prop m=1 bulk 5e9 g 3e9 density 2500
struct cable 0,-.23,0 0,1.77,0 prop 1 seg 10
struct prop 1 area 181e-6 e 98.6e9 yield 5e6
kbond 1.12e8 sbond 1.75e5
; assign mat 2 to cable element in loading block
struct prop 2 area 181e-6 e 98.6e9 yield 1e10 kbond 1.12e8 sbond 1.0e10
; change cable propeties in pull block
change cable 2 range x -.1 .1 y -.3 0 z -.1 .1
;
bound yvel = 0.0
range y -.01 .01
bound yvel = -1.0e-1 range y -.28,-.01
hist ydisp 0,-.27,0
ca boun.fin
def f_block
ib_rock = b_near(0,-.15,0)
end
@f_block
;
def pullf
if cycle < ncycles then
sum = 0.0
y_plus = y_loc +y_tol
ib = block_head
loop while ib # 0
if ib = ib_rock then
ig = b_gp(ib)
loop while ig # 0
y_pos = gp_y(ig)
if y_pos <= y_plus then
igb = gp_bou(ig)
sum = sum - fmem(igb+$kbdfy)
endif
ig = gp_next(ig)
endloop
endif

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 27

ib = b_next(ib)
endloop
last_force = sum / 1.77
endif
pullf = last_force
y_disp = -0.05 * step * tdel
end
set @ncycles = 10000
set @y_loc = -0.001 @y_tol 0.002
hist @pullf
hist @y_disp
hist label 2 Grout Shear Force
hist label 3 Cable Displacement
plot his 2 vs 3 rev yaxis label Grout Shear Force &
xaxis label Cable Displacement
step @ncycles
ret

Figure 3.12 displays the axial force distribution in the cable at the end of the test. A plot of the
total pull force versus displacement resulting from the test is shown in Figure 3.13. As this figure
shows, the peak load is very similar to the input value for grout cohesive strength.
The cable shear bond strength will, in general, increase with increasing effective pressure acting on
the cable. The pressure dependency is not accounted for directly in the present formulation.

E

pGr,007 
 0U
u0 i

0rGGGG
.gppgpGr,0rG5SG5G.0x



7 50m 

 o o

 
50,GGrS4
 50,iG.eSSlG4

07 
 0U
u0 i
  u000x

Figure 3.12 Axial force in cable elements for pull-test

3DEC Version 5.0

3 - 28

Theory and Background

E

pGr,00N 
 0
u0 i

0rGGGG
pgr8gpGr,0rPp,PG,0



p0
00 
i0l,0N 0U   






b b b 2- u












0N 
 0
u0 i
  u000D

           

 b
  b 2- 1G

Figure 3.13 Cable grout shear force in N/m versus cable displacement in meters

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 29

3.3 Beam Structural Elements


3.3.1 Mechanical Behavior
Each beam structural element is defined by its geometric and material properties. A beam is assumed
to be a straight segment of uniform bisymmetrical cross-sectional properties lying between two nodal
points. An arbitrarily curved structural beam can be modeled as a curvilinear structure comprising
a collection of beams. By default, each beam behaves as an isotropic, linearly elastic material with
no failure limit; however, one can specify a limiting axial strength.
Each beam has its own local coordinate system, shown in Figure 3.14. This system is used to
specify both the cross-sectional moments of inertia and applied distributed loading, and to define
the sign convention for force and moment distributions for a single beam (see Figure 3.15). The
beam coordinate system is defined by the locations of its two nodal points, labeled 1 and 2 in
Figure 3.14, and by the vector Y. The beam coordinate system is defined such that:
(1) the centroidal axis coincides with the x-axis;
(2) the x-axis is directed from node-1 to node-2; and
(3) the y-axis is aligned with the projection of Y onto the cross-sectional plane
(i.e., the plane whose normal is directed along the x-axis).
The 12 active degrees-of-freedom of the beam element are shown in Figure 3.14. For each generalized displacement (translation and rotation) shown in the figure, there is a corresponding generalized
force (force and moment).
Forces generated in support elements are applied to the lumped masses which move in response
to unbalanced forces and moments in accordance with the equations of motion. This formulation
has the following desirable characteristics: (1) slip between support and excavation periphery (see
Figure 3.16) is modeled in a manner identical to block interaction along a discontinuity; and (2)
large displacements with non-linear material behavior are readily accommodated.
The S1 axis is given in global coordinates. It is the axis corresponding to the inertial I1. If the tunnel
is in the z-direction, the S1 axis may be given in that direction, and be the same for all elements
around the tunnel. The S2 axis, with inertia I2, will be calculated by the code as normal to both
the S1 and the beam axis, so it will be located in a planar cross-section of the tunnel, varying from
element to element in the radial direction. In this way, all elements share S1, and inertias I1 and
I2 are common to all elements. If it is an I-shaped beam, then the maximum inertia corresponds to
the S1 axis, and the minimum inertia to S2. In Figure 3.14, S1 would be the y-axis shown, and S2
would be the z-axis.

3DEC Version 5.0

3 - 30

Theory and Background

z
z1
x1
w1

u1
1 v1

y1

z2
w2
u2
x2

2
v2

y2

Figure 3.14 Beam coordinate system and 12 active degrees-of-freedom of the


beam element

Figure 3.15 Sign convention for forces and moments at the ends of a beam
(axes show beam coordinate system ends 1 and 2 correspond
with order in nodal connectivity list, and all quantities are drawn
acting in their positive sense)

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 31

Beam properties are easily calculated or obtained from handbooks. For example, typical values for
structural steel are 200 GPa for Youngs modulus, and 0.3 for Poissons ratio. For concrete, typical
values are 25 to 35 GPa for Youngs modulus, 0.15 to 0.2 for Poissons ratio, and 2100 to 2400
kg/m3 for mass density. Composite systems, such as reinforced concrete, should be based on the
transformed section.

Excavation
Periphery

m
1. Lumped Mass

2. Structural Element
m

3. Interface
Lining Interior

Figure 3.16 Lumped mass representation of structure used in explicit formulation

3DEC Version 5.0

3 - 32

Theory and Background

3.3.2 Beam Properties


The beam elements used in 3DEC require the following input properties:
(1) density of material
(2) Youngs modulus
(3) Poissons ratio
(4) cross-sectional area
(5) moment of inertia
(6) tensile yield capacity [force]
(7) compressive yield capacity [force]
All of the commands associated with beam elements are listed in Table 3.3. See Section 1.3 in the
Command Reference for a detailed explanation of these commands.

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 33

Table 3.3 Commands associated with beam elements


STRUCT

beam x1 y1 z1 x2 y2 z2 radial gen seg naxial nradial prop n


<begin xb yb zb end xe ye ze> <connect>

STRUCT

property np

keyword
area
cohesion
density
emod
friction
I
I1
I2
J
kn
ks
nu
S1
tensile
ycomp
yield

PLOT

beam

LIST

beam

keyword
contacts
elements
loads
nodes
property

STRUCT

beam apply

keyword
x
force
free
moment
rx
rfree
rvelocity
velocity

STRUCT

value
value
value
value
value
value
value
value
value
value
value
value
value
value
value
value

ID value
val val val
val val val

val val val


val val val

beam delete

3DEC Version 5.0

3 - 34

Theory and Background

3.3.3 Beam Example


This is a simple example which demonstrates the use of the STRUCT beam command to place a set
of 3 beam support rings inside a tunnel.
The beam material properties are:
area = 0.03 m2
modulus = 200,000 MPa
Poissons ratio = 0.3
density = 2500 kg/m3
moment of inertia = 0.00064 m4
Ks = 10,000 MPa/m
Kn = 10,000 MPa/m
cohesion = 0
tension = 0
friction = 30
Figure 3.17 shows the placement of the beam elements inside the tunnel.

E

pGr,00! 
 0
u0 i

0rGGG
pgr8gpGr,0r5p,5GS0



! 50 
r

 

0! 
 0
u0 i
  u000A

Figure 3.17 Tunnel with beam elements

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 35

Example 3.3 Example of structural beam elements


new
; ------------------------------------------------------------------; structural element example
; tunnel support with beam elements
; ------------------------------------------------------------------;
title
Beam example
;
poly brick -25 25 -25 25 -25 25
plot block
plot reset
;
jset dip 30 dd 90 n 5 sp 10
;
tunnel a -5 -100 0 -3.5 -100 3.5 0 -100 5 3.5 -100 3.5 &
5 -100 0
3.5 -100 -3.5 0 -100 -5 -3.5 -100 -3.5 &
b -5 100 0 -3.5 100 3.5 0 100 5 3.5 100 3.5 &
5 100 0
3.5 100 -3.5 0 100 -5 -3.5 100 -3.5 &
reg 10
;
delete range region 10
;
gen edge 6
;
; E = 50000 MPa, v=0.25
; dens = 2500 kg/m3
prop mat 1 dens 0.0025 bulk 33333 shear 20000
prop mat 1 jkn 10000 jks 4000 coh 1e10 tens 1e1
;
bound yvel 0 range yr -25
bound yvel 0 range yr 25
bound xvel 0 range xr -25
bound xvel 0 range xr 25
bound zvel 0 range zr -25
bound zvel 0 range zr 25
;
insitu stress -5 -5 -5 0 0 0
;
gravity 0 0 -10
;
damp auto
;

3DEC Version 5.0

3 - 36

Theory and Background

struct beam radial_gen 0,-10,0 0,10,0 seg 3 8 prop 1


;
struct prop 1 e 200000 area .03 dens 0.0025
struct prop 1 i 6.4e-4 nu .3
struct prop 1 kn 10000 ks 10000 coh 0 tens 0 fric 30
;
hist xdis 0,0,5 ydis 0,0,5 zdis 0,0,5
;
cy 1000
;
save beam.sav
;
ret

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 37

3.4 Material Properties


Property numbers are assigned to reinforcement elements with the STRUCTURE prop command. It
should be noted that all quantities must be given in an equivalent set of units (see Table 3.4). Note
that for reinforcement elements and cable elements, stiffness has units of [force/displacement] and
strength has units of [force].
Table 3.4

Systems of units structural elements

Property

Unit

SI

Imperial

Area

length2

m2

m2

m2

cm2

ft2

in2

Bond Stiffness

force/length/disp

N/m/m

kN/m/m

MN/m/m

Mdynes/cm/cm

lbf /ft/ft

lbf /in/in

Bond Strength

force/length

N/m

kN/m

mass/volume

kg/m3

103 kg/m3

MN/m
106 kg/m3

Mdynes/cm
106 g/m3

lbf /ft

Density

lbf /in
psi

Elastic Modulus

stress

Pa

kPa

MPa

bar

Stiffness

force/disp

N/m

kN/m

MN/m

Mdynes/cm

Stiffness*

stress/length

Pa/m

kPa/m

MPa/m

bar/m

Yield Strength

force

kN

MN

Mdynes

lbf /ft2
lbf /ft2
lbf /ft
lbf /ft3
lbf

psi
lbf /in
psi/in
lbf

*The beam/rock interface assumes that the beam is in contact along the entire length. The contact
length is used in the calculation of the nodal force. This is different from the stiffness for the liner
elements, which assume point contacts (no length).

3DEC Version 5.0

3 - 38

Theory and Background

3.5 Modeling Considerations


3.5.1 Symmetry Conditions
Structural elements that lie on a plane of symmetry should be assigned full properties for modulus
and stiffness. Property values for cross-sectional area and yield strength should be reduced by 50%
compared to the same property values for elements not on the symmetry plane. Loads applied to
structural elements on symmetry planes should also be reduced by 50% compared to the same loads
applied away from the symmetry plane.
3.5.2 Equilibrium Conditions
The user must decide when the model has reached an equilibrium state. Equilibrium for problems
involving structural elements can be determined by all the usual criteria (e.g., histories, velocity
fields).
3.5.3 Sign Convention
Axial forces in all structural elements are positive in compression. Shear forces follow the opposite
sign convention as that given for zone shear stresses, illustrated in Section 2.7 in the Users Guide.
Axial displacements for cable elements are positive for loading in compression. Axial displacements
for local reinforcement elements are positive for loading in tension.

3DEC Version 5.0

STRUCTURAL ELEMENTS

3 - 39

3.6 References
Bjurstrom, S. Shear Strength on Hard Rock Joints Reinforced by Grouted Untensioned Bolts, in
Proceedings of the 3rd International Congress on Rock Mechanics, Vol. II, Part B, pp. 1194-1199.
Washington, D.C.: National Academy of Sciences (1974).
Dight, P. M. Improvements to the Stability of Rock Walls in Open Pit Mines. Ph.D. Thesis,
Monash University (1982).
Fuller, P. G., and R. H. T. Cox. Rock Reinforcement Design Based on Control of Joint Displacement A New Concept, in Proceedings of the 3rd Australian Tunnelling Conference (Sydney,
Australia, 1978), pp. 28-35. Sydney: Inst. of Engrs., Australia (1978).
Gerdeen, J. C., et al. Design Criteria for Roof Bolting Plans Using Fully Resin-Grouted Nontensioned Bolts to Reinforce Bedded Mine Roof, U.S. Bureau of Mines, OFR 46(4)-80 (1977).
Haas, C. J. Shear Resistance of Rock Bolts, Trans. Soc. Min. Eng. AIME, 260(1), 32-41 (1976).
Hyett, A. J., W. F. Bawden and R. D. Reichert. The Effect of Rock Mass Confinement on the
Bond Strength of Fully Grouted Cable Bolts, Int. J. Rock Mech. Min. Sci. & Geomech. Abstr.,
29(5), 503-524 (1992).
Itasca Consulting Group Inc. UDEC (Universal Distinct Element Code), Version 5.0. Minneapolis:
ICG (2011).
Littlejohn, G. S., and D. A. Bruce. Rock Anchors State of the Art. Part I: Design, Ground
Engineering, 8(3), 25-32 (1975).
Lorig, L. J. A Simple Numerical Representation of Fully Bonded Passive Rock Reinforcement
for Hard Rocks, Computers and Geotechnics, 1, 79-97 (1985).
Par Un Groupe Francais (Azuar, Debreuille, Habib, Londe, Panet and Salembier). Le Renforcement des Massifs Rocheux par Armatures Passives (Rock Mass Reinforcement by Passive Rebars),
in Proceedings of the 4th ISRM Congress (Montreux, Switzerland, September 1979), Vol. 1, pp.
23-30. Rotterdam: A. A. Balkema and The Swiss Society for Soil and Rock Mechanics (1979).
Pells, P. J. N. The Behaviour of Fully Bonded Rock Bolts, in Proceedings of the 3rd International
Congress on Rock Mechanics, Vol. 2, pp. 1212-1217 (1974).
St. John, C. M., and D. E. Van Dillen. Rockbolts: A New Numerical Representation and Its
Application in Tunnel Design, Rock Mechanics Theory - Experiment - Practice (Proceedings
of the 24th U.S. Symposium on Rock Mechanics, Texas A&M University, June, 1983), pp. 13-26.
New York: Association of Engineering Geologists (1983).

3DEC Version 5.0

3 - 40

3DEC Version 5.0

Theory and Background

S-ar putea să vă placă și