Sunteți pe pagina 1din 36

INTERFACES 3-1

3 INTERFACES

3.1 General Comments

There are several instances in geomechanics when it is desirable to represent planes on which sliding
or separation can occur:
1. joint, fault or bedding planes in a geologic medium;
2. an interface between a foundation and the soil;
3. a contact plane between a bin or chute and the material that it contains; and
4. a contact between two colliding objects.
FLAC provides interfaces that are characterized by Coulomb sliding and/or tensile separation.
Interfaces have the properties of friction, cohesion, dilation, normal and shear stiffness, and tensile
strength. Although there is no restriction on the number of interfaces or the complexity of their
intersections, it is generally not reasonable to model more than a few simple interfaces with FLAC
because it is awkward to specify complicated interface geometry. The program UDEC (Itasca 2011)
is specifically designed to model many interacting bodies; it should be used instead of FLAC for
the more complicated interface problems.
An interface can also be specified between structural elements and a grid, or between two structural
elements. Interface/structural connections are described in Section 3.5.6.
Interfaces may also be used to join regions that have different zone sizes. In general, the ATTACH
command should be used to join sub-grids together. However, in some circumstances it may be
more convenient to use an interface for this purpose. In this case, the interface is prevented from
sliding or opening because it does not correspond to any physical entity.*

* The data files in this section are all created in a text editor. The files are stored in the directory
ITASCA\FLAC700\Datafiles\Theory\3-Intface with the extension .DAT. A project file is also
provided for each example. In order to run an example and compare the results to plots in this section,
open a project file in the GIIC by clicking on the File / Open Project menu item and selecting
the project file name (with extension .PRJ). Click on the Project Options icon at the top of the
Project Tree Record, select Rebuild unsaved states, and the example data file will be run, and plots
created.

FLAC Version 7.0


3-2 Theory and Background

3.2 Formulation

An interface is represented as a normal and shear stiffness between two planes which may contact
each other (see Figure 3.1).

LM Lo Side A
N O
gridpoint S ks T

M kn P

LN Side B
zone

S = slider
T = tensile strength
kn = normal stiffness
ks = shear stiffness
Ln = length associated with gridpoint N
Lm = length associated with gridpoint M
denotes limits for joint segments (placed
halfway between adjacent gridpoints)

Figure 3.1 An interface represented by sides a and b, connected by shear


(ks ) and normal (kn ) stiffness springs

For either side of the interface, FLAC uses a contact logic, which is similar in nature to that employed
in the distinct element method (e.g., Cundall and Hart 1992).
The code keeps a list of the gridpoints (i,j) that lie on each side of any particular surface. Each
point is taken in turn and checked for contact with its closest neighboring point on the opposite
side of the interface. Referring to Figure 3.1, gridpoint N is checked for contact on the segment
between gridpoints M and P. If contact is detected, the normal vector, n, to the contact gridpoint,
N, is computed. A length, L, is also defined for the contact at N along the interface. This length is
equal to half the distance to the nearest gridpoint to the left of N plus half the distance to the nearest
gridpoint to the right, irrespective of whether the neighboring gridpoint is on the same side of the
interface or on the opposite side of N. In this way, the entire interface is divided into contiguous
segments, each controlled by a gridpoint.
During each timestep, the velocity, ui , of each gridpoint is determined. Since the units of velocity
are displacement per timestep, and the calculational timestep has been scaled to unity to speed
convergence (see Section 1.3.5), the incremental displacement for any given timestep is

FLAC Version 7.0


INTERFACES 3-3

ui ui (3.1)

The incremental relative displacement vector at the contact point is resolved into the normal and
shear directions, and total normal and shear forces are determined by

(t+(1/2)t)
Fn(t+t) = Fn(t) kn un L (3.2)

(t+(1/2)t)
Fs(t+t) = Fs(t) ks us L

where the stiffnesses, kn and ks , have the units of [stress/displacement].


The following three options are available for specifying the conditions of the interface that may
require adjustment of the contact forces. The material properties related to these conditions are
discussed in Section 3.4.
1. Glued Interfaces If interfaces are declared glued, no slip or opening is
allowed, but elastic displacement still occurs, according to the given stiffnesses.
2. Coulomb Shear-Strength The Coulomb shear-strength criterion limits the
shear force by the relation

Fsmax = cL + tan Fn (3.3)

where c = cohesion (in stress units) along the interface, L = effective contact
length (Figure 3.1), and = friction angle of interface surfaces.
If the criterion is satisfied (i.e., if |Fs | Fsmax ), then Fs = Fsmax , with the
sign of shear preserved.
In addition, the interface may dilate at the onset of slip (nonelastic sliding).
Dilation is governed in the Coulomb model by a specified dilation angle, .
Dilation is a function of the direction of shearing. Dilation increases if the shear
displacement increment is in the same direction as the total shear displacement,
and decreases if the shear increment is in the opposite direction.
During sliding, shear displacement can cause an increase in the effective nor-
mal stress on the interface, according to the relation

|Fs |o Fsmax
n := n + tan kn (3.4)
Lks

FLAC Version 7.0


3-4 Theory and Background

where |Fs |o is the magnitude of shear force before the above correction is
made.
3. Tension Bond Two conditions are available for a bonded interface:
a. Bonded interface If a (positive) tensile bond strength is specified
for an interface, each segment of the interface acts as if it is glued
(elastic response only), while the magnitude of the tensile normal
stress is below the bond strength. If the magnitude of the tensile nor-
mal stress of a segment exceeds the bond strength (set with tbond),
the bond breaks for that segment, and the segment behaves there-
after as un-bonded (separation and slip allowed, as described above,
in the normal way).
A shear bond strength, as well as the tensile bond strength, can be
specified. The bond breaks if the shear stress exceeds the shear bond
strength, or the tensile effective normal stress exceeds the normal
bond strength. The shear bond strength is set to sbr times the normal
bond strength, using the sbratio= sbr property keyword. The default
shear bond strength is 100 times the tensile bond strength.
b. Slip while bonded There is an optional switch (bslip=on) that
allows slip to occur for a bonded interface segment, even though
separation has not occurred. Shear yield is under the control of the
friction and cohesion parameters, using the absolute value of the
effective normal force. Note that dilation response is suppressed
(i.e., = 0) when bslip=on. By default, bslip=off if not specified.
The corrected forces are then rotated back to the global xy-reference frame (from the normal and
shear directions) and lumped onto the adjacent gridpoints in such a ratio as to preserve moment equi-
librium. These are then summed with all other forces when unbalanced force sums are calculated
for each calculation step.
The influence of the interfaces must be accounted for in determining the critical mechanical timestep,
or instability in the solution may result. The way the fictitious gridpoint mass is increased in
proportion to the interface stiffness at that gridpoint to ensure numerical stability is explained in
Section 1.3.5.
The effect of pore pressure is included in the interface calculation by using effective stress as the
basis for the slip condition. Neither pressure drop normal to the joint nor influence of normal
displacement on pore pressure is calculated. Conduction of fluid along the interface is not modeled.

FLAC Version 7.0


INTERFACES 3-5

3.3 Creation of the Required Geometry

When interfaces are present, the FLAC grid must be divided into two or more separate regions, or
sub-grids. The following steps are recommended:
1. Identify the number (N ) of discrete bodies or regions in the problem. List
their corner coordinates and any embedded structures within the regions.
2. Select the overall FLAC grid (number of rows or columns) and divide it into
N regions, each of which is separated from the others by at least one zone
width (to be null zones in the final grid). For this operation, graph paper may
be used to divide the overall grid into regions; at this stage, the geometry is
immaterial we are simply concerned with the topology. In dividing the grid,
pay attention to the mesh density: we need more zones available in areas of
high stress gradient or where the geometry is complicated.
3. Take each subregion in turn and use the GENERATE or INITIAL command to
map the sub-grid onto the region in space that it will occupy. It is advisable
to plot each sub-grid as it is created and mapped (keeping other regions null).
The interior of the sub-grid may also be adjusted to conform to any internal
structures that may be present (e.g., tunnels). These may be filled with solid
material initially, but excavated later. Note that objects (such as tunnels) that
span an interface must be created in separate sections one for each sub-grid
involved.
4. Install the appropriate material models in all sub-grids, and plot the whole grid
to check the correctness of the geometry.
5. Specify interfaces between all pairs of sub-grids that may interact. Plot out
each interface (PLOT iface) in turn to check on correctness.
Section 2.4 gives an example of the rezoning necessary to accommodate a single interface. As
an example of the recommended methodology for a more complicated geometry, we construct
a grid for the physical system, shown in Figure 3.2. The project file for this example is named
TUNNEL.PRJ.
Stage 1 The coordinates are identified in Figure 3.2: there are two intersecting faults, one of
which also intersects a tunnel (or cavern).
Stage 2 We choose a grid of 50 zones by 30 zones and divide it into the four regions shown in
Figure 3.3.

FLAC Version 7.0


3-6 Theory and Background

x = 70m

y=0

R = 5.5m
y = -20m

y = -26m

y = -35m 50m 61m


y = -45m

y = -70m

x = 0m x = 60m x = 100m

Figure 3.2 Geometry to be modeled

Figure 3.3 Grid is divided into four regions

FLAC Version 7.0


INTERFACES 3-7

The FLAC commands to do this are shown in Example 3.1:

Example 3.1 Divide grid into four regions


grid 50 30
mod elas
mod null j=14 i=1,28
mod null i=29
mod null j=20 i=30,50

Stage 3 We now map each region of the grid to its own region of space by using the commands
in Example 3.2:

Example 3.2 Map regions of grid into regions of space


gen 0,-45 0,0 70,0 65.926,-28.519 i=1,29 j=15,31
gen 65.926,-28.519 70,0 100,0 100,-20 i=30,51 j=21,31
gen 60,-70 65.926,-28.519 100,-20 100,-70 i=30,51 j=1,20
gen 0,-70 0,-45 65.926,-28.519 60,-70 i=1,29 j=1,14

The resulting grid is shown in Figure 3.4:

Figure 3.4 The four grid regions are mapped to their correct positions in
space

FLAC Version 7.0


3-8 Theory and Background

In order to create the geometry corresponding to the excavated region, we first move four gridpoints
so that they lie exactly on the endpoints of the straight-line segments. As explained in Section 2.2,
this helps the GENERATE command to do its job correctly. Then, the line segments are rezoned with
GENERATE line commands, and the circular segment is created with the GENERATE arc command,
as shown in Example 3.3:

Example 3.3 Create tunnel region


ini x=50 y=-35 i=22 j=13
ini x=61 y=-35 i=27 j=12
ini x=50 y=-26 i=22 j=18
ini x=61 y=-26 i=27 j=17
gen line 50,-35 61,-35
gen line 50,-35 50,-26
gen line 61,-35 61,-26
gen arc 55.5,-26 61,-26 180

The geometry is illustrated in Figure 3.5. Note that the gridpoint at the left intersection of the
interface and cavern has been moved down slightly; for a more accurate geometry, this point should
be rezoned back to its correct position. (The effect occurs because GENERATE line should not really
be used across interfaces.)

Figure 3.5 After rezoning for the tunnel

FLAC Version 7.0


INTERFACES 3-9

Stage 4 We now have two choices:


1. leave the tunnel material in place and specify four interfaces, removing the
tunnel material later on; or
2. remove the tunnel material and then specify the interfaces. It is still possible
to prestress the material and the faults.
For this example we use option 2.
Stage 5 The cavern is excavated as follows (two commands are needed because there are two
sub-grids that need to be addressed):

Example 3.4 Excavate cavern


mod null reg=23,13
mod null reg=23,15

Now the interfaces are specified. Because the cavern has been removed, the fault that spans it must
be specified in two parts:

Example 3.5 Create faults


int 1 Aside from 1,14 to 22,14 Bside from 1,15 to 22,15
int 2 Aside from 27,14 to 29,14 Bside from 27,15 to 29,15
int 3 Aside from 30,20 to 51,20 Bside from 30,21 to 51,21
int 4 Aside from 29,1 to 29,14 Bside from 30,1 to 30,20
int 5 Aside from 29,15 to 29,31 Bside from 30,21 to 30,31

Properties must also be specified for the intact material and for the interfaces. The initial stresses may
also be installed in both the intact material and the interfaces by one INITIAL command. Figure 3.6
shows the grid with cavern removed, and with interfaces drawn, with crosses to denote associated
gridpoints.

FLAC Version 7.0


3 - 10 Theory and Background

2 3
4

Figure 3.6 Final grid, with cavern removed and interfaces highlighted

FLAC Version 7.0


INTERFACES 3 - 11

3.4 Choice of Material Properties

The material properties (particularly stiffnesses) assigned to an interface depend on the way the
interface is used. Three possibilities are common. The interface may be:
1. an artificial device to connect two sub-grids together;
2. a real interface that is stiff compared to the surrounding material, but which
can slip and perhaps open in response to the anticipated loading (this case also
encompasses the situation in which stiffnesses are unknown or unimportant,
but where slip and/or separation will occur e.g., flow of frictional material
in a bin); or
3. a real interface that is soft enough to influence the behavior of the system (e.g.,
a joint with soft clay filling or a dike containing heavily fractured material).
These cases are examined in detail.

3.4.1 Interface Used to Join Two Sub-Grids

If possible, sub-grids should be joined with the ATTACH command. It is more computationally
efficient to use ATTACH instead of INTERFACE to join sub-grids. See Section 3.2 in the Users
Guide for a description of and restrictions on the ATTACH command.
Under some circumstances it may be necessary to use an interface to join two sub-grids. This type
of interface is declared as glued on the INTERFACE command, thus preventing any slip or separation;
values of friction, cohesion and tensile strength are not needed, and are ignored if given. However,
shear and normal stiffnesses must be provided. It is tempting (particularly for people familiar
with finite element methods) to give a very high value for these stiffnesses to prevent movement
on the interface. However, FLAC does mass scaling (see Section 1.3.5) based on stiffnesses
the response (and solution convergence) will be very slow if very high stiffnesses are specified.
It is recommended that the lowest stiffness consistent with small interface deformation be used.
A good rule-of-thumb is that kn and ks be set to ten times the equivalent stiffness of the stiffest
neighboring zone. The apparent stiffness (expressed in stress-per-distance units) of a zone in the
normal direction is

 
K + 43 G
max (3.5)
zmin

where:K & G are the bulk and shear moduli, respectively; and
zmin is the smallest width of an adjoining zone in the normal direction see
Figure 3.7.

The max [ ] notation indicates that the maximum value over all zones adjacent to the interface is
to be used (e.g., there may be several materials adjoining the interface).

FLAC Version 7.0


3 - 12 Theory and Background

Z min

Interface

Figure 3.7 Zone dimension used in stiffness calculation

To illustrate the approach, consider Figure 3.8, in which two sub-grids of unequal zoning are joined
by the commands in Example 3.6, and are loaded by a pressure on the left-hand part of the upper
surface.

Example 3.6 Joining two sub-grids


g 20 16
mod elas i=1,10 j=1,5
mod elas i=1,20 j=7,16
gen 0,0 0,5 10,5 10,0 i=1,11 j=1,6
gen 0,5 0,10 10,10 10,5 i=1,21 j=7,17
int 1 Aside from 1,6 to 11,6 Bside from 1,7 to 21,7
int 1 glue kn 2e10 ks 2e10
prop dens 1000 sh 3e8 bu 6e8
fix x y j=1
fix x i=1
fix x i=11 j=1,6
fix x i=21 j=7,17
apply p=1e6 i=1,5 j=17
his yd i 1 j 17
step 2000
scl 1 0 9.5 9 1
plo pen bou yd i=5e-4

The minimum zone size adjacent to the interface is 0.5 m, and the value of (K + 4G/3) is 109 Pa.
Hence, we choose both shear stiffness and normal stiffness to be 10 109 /0.5 (i.e., kn = ks =
2 1010 Pa/m). The resulting contours of y-displacement are shown in Figure 3.9.

FLAC Version 7.0


INTERFACES 3 - 13

Figure 3.8 Two unequal sub-grids joined by an interface

B
C
D
E
F
G
H
I
J

Figure 3.9 Vertical displacement contours two joined grids

FLAC Version 7.0


3 - 14 Theory and Background

To test the accuracy of this approach, we do a similar run, but for a single grid with the constant
mesh size of 0.5 m. The data file for this run is given in Example 3.7:

Example 3.7 A single grid for comparison to two sub-grids


g 20 20
mod elas
gen 0,0 0,10 10,10 10,0
prop dens 1000 sh 3e8 bu 6e8
fix x y j=1
fix x i=1
fix x i=21
apply p=1e6 i=1,5 j=21
his yd i 1 j 21
step 2500
scl 1 0 9.5 9 1
plot pen bou yd i=5e-4
ret

The results from this run are given in Figure 3.10, which is plotted with the same scale and contour
intervals as the previous run. The two plots are almost identical, which indicates that the interface
does not affect the behavior to any great extent.

B
C
D
E
F
G
H
I
J

Figure 3.10 Vertical displacement contours single, uniform grid

FLAC Version 7.0


INTERFACES 3 - 15

The prescription given in Eq. (3.5) is reasonable if the material on the two sides of the interface
are similar and variations of stiffness occur only in the lateral directions. However, if the material
on one side of the interface is much stiffer than that on the other, then Eq. (3.5) should be applied
to the softer side. In this case, the deformability of the whole system is dominated by the soft
side; making the interface stiffness ten times the soft-side stiffness will ensure that the interface has
minimal influence on system compliance.

3.4.2 Real Interface Slip and Separation Only

In this case, we simply need to provide a means for one sub-grid to slide and/or open relative to
another sub-grid. The friction (and perhaps cohesion and tensile strength) is important, but the
elastic stiffness is not. The approach of Section 3.4.1 is also used here to determine kn and ks .
However, the other material properties are given real values (see Section 3.4.3 for advice on choice
of properties). As an example, we can allow slip in a bin-flow problem (as shown in Figure 3.11),
corresponding to the data file in Example 3.8:

Example 3.8 Slip in a bin-flow problem


g 7 10
mod mohr i=1,5
mod elas i=7
gen 0,0 0,5 5,5 3,0 i=1,6 j=1,6
gen 3,0 5,5 6,5 6,0 i=7,8 j=1,6
gen 5,5 5,10 6,10 6,5 i=7,8 j=6,11
fix x y i=7,8
fix x i=1
prop dens=2000 shear=1e8 bulk=2e8 fric=30 i=1,5
prop dens=2000 shear=1e8 bulk=2e8 i=7
int 1 Aside from 6,1 to 6,11 Bside from 7,1 to 7,11
int 1 ks=2e9 kn=2e9 fric=15
set large, grav=10
step 1500
plot pen grid vel
ret

FLAC Version 7.0


3 - 16 Theory and Background

Figure 3.11 Flow of frictional material in a bin

3.4.3 Real Interface All Properties Have Physical Significance

In this case, properties should be derived from tests on real joints* (suitably scaled to account for
size effect), or from published data on materials similar to the material being modeled. However,
the comments of Section 3.4.1, with respect to the maximum stiffnesses that are reasonable to use,
also apply here. If the physical normal and shear stiffnesses are less than ten times the equivalent
stiffnesses of adjacent zones, then there is no problem in using physical values. If the ratio is much
more than ten, the solution time will be significantly longer than for the case in which the ratio is
limited to ten, without much change in the behavior of the system. Serious consideration should
be given to reducing supplied values of normal and shear stiffness to improve solution efficiency.
There may also be problems with interpenetration if the normal stiffness, kn , is very low. A rough
estimate should be made of the joint normal displacement that would result from the application of
typical stresses in the system (u = /kn ). This displacement should be small compared to a typical
zone size. If it is greater than, say, 10% of an adjacent zone size, then there is either an error in
one of the numbers, or the stiffness should be increased if calculations are to be done in large-strain
mode.
Joint properties are conventionally derived from laboratory testing (e.g., triaxial and direct shear
tests). These tests can produce physical properties for joint friction angle, cohesion, dilation angle

* Joint is used here as a generic term.

FLAC Version 7.0


INTERFACES 3 - 17

and tensile strength, as well as joint normal and shear stiffnesses. The joint cohesion and friction
angle correspond to the parameters in the Coulomb strength criterion* described in Section 3.2.
Values for normal and shear stiffnesses for rock joints typically can range from roughly 10 to 100
MPa/m for joints with soft clay in-filling, to over 100 GPa/m for tight joints in granite and basalt.
Published data on stiffness properties for rock joints are limited; summaries of data can be found
in Kulhawy (1975), Rosso (1976) and Bandis et al. (1983).
Approximate stiffness values can be back-calculated from information on the deformability and
joint structure in the jointed rock mass, and the deformability of the intact rock. If the jointed rock
mass is assumed to have the same deformational response as an equivalent elastic continuum, then
relations can be derived between jointed rock properties and equivalent continuum properties. For
uniaxial loading of rock containing a single set of uniformly spaced joints oriented normal to the
direction of loading, the following relation applies:

1 1 1
= +
E Er kn s

or (3.6)

E Er
kn =
s (Er E)

where E = rock mass Youngs modulus;


Er = intact rock Youngs modulus;
kn = joint normal stiffness; and
s = joint spacing.

A similar expression can be derived for joint shear stiffness, ks :

G Gr
ks = (3.7)
s (Gr G)

where G = rock mass shear modulus; and


Gr = intact rock shear modulus.

* The Coulomb yield surface provides a reasonable approximation for joint strength for most engi-
neering calculations. More complex joint models, which include, for example, effects of continuous
yielding and displacement weakening, are available. For analysis with other joint models, the user
is referred to UDEC (Itasca 2011).

FLAC Version 7.0


3 - 18 Theory and Background

The equivalent continuum assumption, when extended to three orthogonal joint sets, produces the
relations

 1
1 1
Ei = + (i = 1, 2, 3) (3.8)
Er si kni
 1
1 1 1
Gij = + + (i, j = 1, 2, 3) (3.9)
Gr si ksi sj ksj

Several expressions have been derived for two- and three-dimensional characterizations and multiple
joint sets. References for these derivations can be found in Singh (1973), Gerrard (1982(a) and (b))
and Fossum (1985).
Published strength properties for joints are more readily available than stiffness properties. Sum-
maries can be found, for example, in Jaeger and Cook (1969), Kulhawy (1975) and Barton (1976).
Friction angles can vary from less than 10 for smooth joints in weak rock such as tuff, to over
50 for rough joints in hard rock such as granite. Joint cohesion can range from zero to values
approaching the compressive strength of the surrounding rock.
It is important to recognize that joint properties measured in the laboratory are typically not rep-
resentative of those for real joints in the field. Scale dependence of joint properties is a major
question in rock mechanics. Often, the only way to guide the choice of appropriate parameters is
by comparison to similar joint properties derived from field tests. However, field test observations
are extremely limited. Some results are reported by Kulhawy (1975).
The following example illustrates an application of the interface logic to simulate the physical
response of a rock joint subjected to normal and shear loading. The model represents a direct shear
test, which consists of a single horizontal joint that is first subjected to a normal confining stress
and then to a unidirectional shear displacement. Figure 3.12 shows the model.

FLAC Version 7.0


INTERFACES 3 - 19

JOB TITLE : DIRECT SHEAR TEST (*10^1)


1.600

FLAC (Version 7.00)

LEGEND
1.200
11-Jan-11 16:05
step 3225
-1.050E+00 <x< 2.205E+01
-6.686E+00 <y< 1.641E+01

0.800
Velocity vectors
max vector = 5.048E-07

0 1E -6
Boundary plot
0.400

0 5E 0

0.000

-0.400

Itasca Consulting Group, Inc.


Minneapolis, MN 55401
0.200 0.600 1.000 1.400 1.800
(*10^1)

Figure 3.12 Direct shear test model

First, a normal stress of 10 MPa that is representative of the confining stress acting on the joint is
applied. A horizontal velocity is then applied to the top sub-grid to produce a shear displacement
along the interface. For demonstration purposes, we only apply a small shear displacement of less
than 1 mm to this model.
The average normal and shear stresses, and normal and shear displacements, along the joint are
measured with a FISH function, av str. With this information we can determine the shear strength
and dilation that are produced. The data file for this test is contained in Example 3.9:

Example 3.9 Direct shear test


g 20 21
model elas
gen 0,0 0,10 21,10 21,0
model null j 11
model null i 1,4 j 12,21
model null i 17,20 j 12,21
ini x add .5 j 12 22
ini y add -.5 j 12 22
int 1 Aside from 1,11 to 21,11 Bside from 5,12 to 17,12
int 1 kn 40000 ks 40000 fric 30 dil 6.0
prop dens 2.6e-3 bu 45000 sh 30000
fix x y j=1

FLAC Version 7.0


3 - 20 Theory and Background

fix x i=1 j 1,11


fix x i=21 j=1,11
apply p=10.0 i=5,17 j=22
his unb
solve
;
; functions to calculate average joint stresses and displacements
call int.fin
def ini_jdisp
njdisp0 = 0.0
sjdisp0 = 0.0
pnt = int_pnt
loop while pnt # 0
pa = imem(pnt+$kicapt)
loop while pa # 0
sjdisp0 = sjdisp0 + fmem(pa+$kidasd)
njdisp0 = njdisp0 + fmem(pa+$kidand)
pa = imem(pa)
end_loop
pa = imem(pnt+$kicbpt)
loop while pa # 0
sjdisp0 = sjdisp0 + fmem(pa+$kidasd)
njdisp0 = njdisp0 + fmem(pa+$kidand)
pa = imem(pa)
end_loop
pnt = imem(pnt)
end_loop
end
ini_jdisp
;
def av_str
whilestepping
sstav = 0.0
nstav = 0.0
njdisp = 0.0
sjdisp = 0.0
ncon = 0
jlen = 0.0
pnt = int_pnt
loop while pnt # 0
pa = imem(pnt+$kicapt)
loop while pa # 0
sstav = sstav + fmem(pa+$kidfs)
nstav = nstav + fmem(pa+$kidfn)
jlen = jlen + fmem(pa+$kidlen)
sjdisp = sjdisp + fmem(pa+$kidasd)

FLAC Version 7.0


INTERFACES 3 - 21

njdisp = njdisp + fmem(pa+$kidand)


pa = imem(pa)
end_loop
pa = imem(pnt+$kicbpt)
loop while pa # 0
ncon = ncon + 1
sstav = sstav + fmem(pa+$kidfs)
nstav = nstav + fmem(pa+$kidfn)
jlen = jlen + fmem(pa+$kidlen)
sjdisp = sjdisp + fmem(pa+$kidasd)
njdisp = njdisp + fmem(pa+$kidand)
pa = imem(pa)
end_loop
pnt = imem(pnt)
end_loop
if ncon # 0
sstav = sstav / jlen
nstav = nstav / jlen
sjdisp = (sjdisp-sjdisp0) / (2.0 * ncon)
njdisp = (njdisp-njdisp0) / (2.0 * ncon)
endif
end
hist sstav nstav sjdisp njdisp
;
ini xvel 5e-7 i= 5,17 j 12,22
fix x i= 5,17 j 12,22
;
hist ns 1
;
ini xdis 0.0 ydis 0.0
step 2000

The average shear stress versus shear displacement along the joint is plotted in Figure 3.13, and
the average normal displacement versus shear displacement is plotted in Figure 3.14. These plots
indicate that joint slip occurs for the prescribed properties and conditions. The loading slope in
Figure 3.13 is initially linear and then becomes nonlinear as interface nodes begin to fail, until a
peak shear strength of approximately 5.8 MPa is reached. As indicated in Figure 3.14, the joint
begins to dilate when the interface nodes begin to fail in shear.

FLAC Version 7.0


3 - 22 Theory and Background

JOB TITLE : DIRECT SHEAR TEST

FLAC (Version 7.00)

LEGEND

11-Jan-11 16:05
step 3225 5.000

HISTORY PLOT
Y-axis :
Rev 2 sstav (FISH) 4.000
X-axis :
4 sjdisp (FISH)

3.000

2.000

1.000

5 10 15 20 25 30 35 40

-05
(10 )
Itasca Consulting Group, Inc.
Minneapolis, MN 55401

Figure 3.13 Average shear stress versus shear displacement

JOB TITLE : DIRECT SHEAR TEST

FLAC (Version 7.00)


-05
(10 )
LEGEND

11-Jan-11 16:05
2.500
step 3225

HISTORY PLOT
Y-axis : 2.000
5 njdisp (FISH)
X-axis :
4 sjdisp (FISH) 1.500

1.000

0.500

0.000

5 10 15 20 25 30 35 40

-05
(10 )
Itasca Consulting Group, Inc.
Minneapolis, MN 55401

Figure 3.14 Average normal displacement versus shear displacement

FLAC Version 7.0


INTERFACES 3 - 23

3.5 Modeling Guidelines

3.5.1 Troubleshooting

After the grid has been generated, the interfaces specified, and the properties supplied, FLAC should
be run for one step, and the command PRINT iface used to display the data for all interfaces. (Note
that certain interface data are not computed until a STEP command is given.) In particular, the
lengths and normal-vector directions should be examined for any obvious anomalies. For example,
zero length will be shown for any interface segments not in contact (see Section 3.5.4 for related
information). After applying stresses and executing a few more steps, the printed output should
be examined again to check for nonzero normal stresses (negative for compression), and values of
printed interface properties should be checked. Note that property values will default to zero if not
given. For example, zero normal stresses may indicate that the normal stiffness was omitted from
the data file.

3.5.2 Initial Stresses

If stresses are initialized in the grid (with the INITIAL command) for a range of zones that encompasses
an interface, then interface stresses will be set to corresponding values, resolved into the plane of the
interface. These shear and normal stresses should be in equilibrium with the grid stresses, but there
may be some slight discrepancies in places where the grid is distorted; a few execution steps should
restore equilibrium. Interface stresses will only be initialized if the appropriate INITIAL command
is given after the interface has been created. Stresses should be initialized in small-strain mode
in order to calculate interface stresses. In large-strain mode, some movement must occur for the
contact to be recognized and interface stresses calculated (see Section 3.5.5).
The action of FLAC, when initializing zone stresses in the range of an interface, is to compute
the normal and shear stresses on each interface segment, resolved in the local direction of each
segment. In order to do this, three stress components are necessary: xx , yy and xy . To allow for
the fact that each stress component may be given on separate INITIAL command lines, the interface
remembers the values of previously given stress components, and combines the latest known
values to give the three components needed. Normally, the user should give all three components
to avoid ambiguity. However, if for example only one component is given with an INITIAL
command, and there are no previous INITIAL commands, the other components will be taken as
zero. This is consistent with the logic stated above, but may lead to puzzling results.

FLAC Version 7.0


3 - 24 Theory and Background

3.5.3 Use of Interfaces with Null Zones

Interfaces may not connect to null zones. This condition is only checked when a STEP command
is given, so zones connected to interfaces may be set to null temporarily, and then restored later to
real material. If an interface is found to be connected to one or more null zones (e.g., if material is
mined up to, or includes, an interface), the interface is automatically reorganized when the STEP
command is given (e.g., parts may be deleted to remove the null-zone connections). In this case,
existing interfaces may be renumbered, and new interfaces may be created with new ID numbers.
For future reference, the user should record the new interface numbers whenever zones next to
interfaces are made null. When a new interface is created automatically from part of an old one, all
existing forces and properties are retained.

3.5.4 Overlapping and Intersecting Interfaces

Any gridpoint may belong to any number of interfaces simultaneously, although only one of the
interfaces will be active (i.e., generating forces) because of the geometric improbability of two or
more bodies touching another at one point. FLAC only recognizes contact between two bodies
when the possibility of such contact is specified in advance with an INTERFACE command. If large
geometric changes are possible, all of the possible interactions should be anticipated in advance. For
the example illustrated in Figure 3.15, interfaces should be declared for interactions A-B, C-D, B-D,
A-C and E-F. Faces B and D are not touching initially, but subsequent movement may bring them
into contact. No forces will be generated if they are not actually in contact (however, some computer
time will be expended, since FLAC keeps checking for contact). The same considerations should
be applied to a mining excavation in which the roof may eventually touch the floor; an interface
that connects the roof and floor should be declared.

E F

B C
A D
FIXED FIXED

Figure 3.15 Case in which future contacts need to be anticipated

FLAC Version 7.0


INTERFACES 3 - 25

3.5.5 Tolerances: Differences between Large- and Small-Strain Operation

When distorted sub-grids are brought into contact (e.g., two grids produced with the GENERATE
circle command), there may be some puzzling effects. In small-strain mode, a tolerance of 104
times the local zone size is used for contact (i.e., if a gridpoint is within the tolerance, it will be
recognized as in contact). In large-strain mode, the tolerance is zero (i.e., some movement must
take place before contact actually occurs). Some gridpoints may not be able to touch the opposing
grid because of the discrete nature of the grid (Figure 3.16). In this case, enough touching points
will be found to prevent one grid from interpenetrating the other, but the local stress distribution
will be rather irregular, owing to the isolated contact points. Some problems with irregular grids
could occur if the solution mode is changed in mid-run from small- to large-strain (owing to the
different tolerance in the two modes).

Sub-grid 1

this gridpoint does


not touch
Sub-grid 2

Figure 3.16 Non-touching points in irregular grid

If the coordinates of the model are very large and the zone sizes are small, it is possible that the
interfaces may not be recognized when the model is run in small-strain mode. If the interfaces are
not recognized, or if initializing stresses do not install interface stresses, then use the command SET
tolint to increase the tolerance to a value greater than 104 . Note that the tolerance should not be
greater than 0.1.

FLAC Version 7.0


3 - 26 Theory and Background

3.5.6 Use of Interfaces with Structural Elements

Interfaces can involve structural elements. Beams may interact with the grid, and beams may
interact with other beams. As with the grid-grid interface, a gap between the potentially contacting
objects may exist initially. The INTERFACE command accepts the keyword phrase from node n1 to
node n2, following either aside or bside. The numbers n1 and n2 correspond to structural nodes.
When connecting structural elements to interfaces, these rules must be followed:
1. Only beam or liner elements may be part of an interface.
2. An unbroken series of beam or liner elements must connect the two given
nodes.
3. No branching structures may exist at the time of specification.
4. The active side of the string of beam elements is on the left of the direction
implied by the from ... to construction (i.e., the contacting body approaches
from the left, when facing along the beam in the direction from to). Please
note that it is important to give the two end nodes in the correct order.
5. For the case where a closed loop of beams interacts with a grid via an interface,
the from and to nodes are the same. Hence, the directionality of the chain of
beams cannot be specified in the usual way (i.e., the direction implied by the
from . . . to sequence). To specify the direction in this case, the from node may
be followed by another node number, nx. This second node is the next node
in the sequence, which conveys the direction required. For example,
int 1 as fr no 4,7 to no 22 bs fr 3,3 to 8,8

Here the starting node on the aside is 4, and the next node to be taken is 7. If
there is no neighbor with the ID of 7, an error is signaled. If the double-node
notation is specified for a chain of beams that is not closed, then only the
direction implied by the pair will be tried. However, the double-node notation
is not needed in that case. The second of the pair is optional, and is normally
omitted.
It is recommended that the beam end nodes be specified in advance with STRUCTURE node com-
mands so that there are known ID numbers for the from . . . to specification. Otherwise, node
numbers will be assigned automatically.
The following examples demonstrate the interaction between a beam and interface, and between
two beams. In Example 3.10, a ball is dropped onto a single, soft beam (e.g., a trampoline). The
initial and equilibrium positions are shown in Figure 3.17.

FLAC Version 7.0


INTERFACES 3 - 27

Example 3.10 Ball dropped onto a trampoline


g 5,5
m e
gen circ 2.5 2.5 2.5
m n
m e reg 3 3
pro d 20000 sh .3e7 bu 5e7
struct node 10 -3,-1 fix x y r
struct node 20 8,-1 fix x y r
struct beam beg node 10 end node 20 seg=15
struct prop 1 e 0.5e10 a .01 i 1e-4
int 1 as from node 10 to node 20 bs from 1,2 to 6,2
int 1 ks 1e8 kn 2e8 fric 2
set grav 10
wind -5 10 -5 10
set large
plot hold noh grid beam white
step 500
plot hold noh grid beam white

(a) initial position

(b) equilibrium position

Figure 3.17 Ball falling onto single beam

FLAC Version 7.0


3 - 28 Theory and Background

In Example 3.11, the ball falls onto two beams. The lower beam has a moment of inertia two orders
of magnitude higher than the upper beam.

Example 3.11 Ball dropped onto two beams


g 5,5
m e
gen circ 2.5 2.5 2.5
m n
m e reg 3 3
pro d 20000 sh .3e7 bu 5e7
struct node 10 -3,-1 fix x y r
struct node 20 8,-1 fix x y r
struct beam beg node 10 end node 20 seg=15
struct prop 1 e 0.5e10 a .01 i 1e-4
int 1 as from node 10 to node 20 bs from 1,2 to 6,2
int 1 ks 1e8 kn 2e8 fric 2
;
struct node 100 -3,-2 fix x y r
struct node 200 8,-2 fix x y r
struct beam beg node 100 end node 200 seg=15 prop=2
struct prop 2 e 0.5e10 a .01 i 1e-2
int 2 as from node 20 to node 10 bs from node 100 to node 200
int 2 ks 1e8 kn 2e8 fric 2
set grav 10
wind -5 10 -9 6
set large
plot pen noh grid beam white
step 1500
plot pen noh grid beam white

FLAC Version 7.0


INTERFACES 3 - 29

The initial and equilibrium results are given in Figure 3.18. Note that a higher deformation develops
in the upper beam (beam 2).

(a) initial position

(b) equilibrium position

Figure 3.18 Ball falling onto two beams

A gap between a tunnel liner and the excavation surface can be simulated by installing an interface
between the grid and structural elements representing the liner. The liner will take load when the
gap closes to zero. Example 3.12 contains the data file for this example:

Example 3.12 A gap between a tunnel liner and the excavation surface
;--- demonstration of tunnel liner with gap ---
def setup ; tunnel and liner sizes
xliner = 15.0 ; center of tunnel and liner
yliner = 7.5 ; "
rtunnel = 5.0 ; radius of tunnel
rliner = 4.9 ; radius of liner
nbeam = 24 ; number of beam elements in liner
; (divisible by 4 for best symmetry)
end
setup
grid 30 15

FLAC Version 7.0


3 - 30 Theory and Background

mod mohr
gen circ xliner yliner rtunnel
prop den 2000 shear 1e9 bulk 2e9 tens 1e20 fric 30
fix y j=1
fix x i=1
fix x i=31
set large, grav=10
ini syy -13e5 var 0 3e5
ini sxx -9e5 var 0 2e5
ini szz -10e5 var 0 2.5e5
apply syy=-10e5 j=16 ; surface stress, for large movement
mod null reg 15 7
def liner ; create ring of beam elements
dang = 2.0 * pi / float(nbeam)
x1 = xliner
y1 = yliner + rliner
ang = pi / 2.0
loop n (1,nbeam)
ang = ang + dang
x2 = rliner * cos(ang) + xliner
y2 = rliner * sin(ang) + yliner
command
struct beam beg x1,y1 end x2,y2
end_command
x1 = x2
y1 = y2
end_loop
end
set echo=off
liner
set echo=on
struct prop 1 e=1e12 height 0.28 width 0.28
int 1 Aside from node 1,nbeam to node 1 Bside long from 16,3 to 16,3
int 1 ks 2e9 kn 2e9 fric 30
hist unbal
hist ydis i 16 j 16
hist ydis i 16 j 13
step 1500
save tun_lin.sav
plot boun struct mom yellow fill hold
plot boun beam white ydisp fill hold
ret

A 0.1 m gap is prescribed between the liner and the grid. After the gap closes, moments build in
the liner, as indicated in Figure 3.19. The vertical displacement of the grid is shown by the contour
plot in Figure 3.20.

FLAC Version 7.0


INTERFACES 3 - 31

JOB TITLE : TUNNEL LINER WITH GAP (*10^1)

FLAC (Version 7.00)

2.000
LEGEND

11-Jan-11 16:05
step 1500
-1.667E+00 <x< 3.167E+01 1.500

-9.182E+00 <y< 2.415E+01

Boundary plot

1.000
0 1E 1
Beam Plot
Moment on
Structure Max. Value
0.500
# 1 (Beam ) 1.371E+06

0.000

-0.500

Itasca Consulting Group, Inc.


Minneapolis, MN 55401
0.250 0.750 1.250 1.750 2.250 2.750
(*10^1)

Figure 3.19 Moment in liner after gap closes

JOB TITLE : TUNNEL LINER WITH GAP (*10^1)

FLAC (Version 7.00)

2.000
LEGEND

11-Jan-11 16:05
step 1500
-1.667E+00 <x< 3.167E+01 1.500

-9.182E+00 <y< 2.415E+01

Y-displacement contours
-1.25E-01
1.000
-1.00E-01
-7.50E-02
-5.00E-02
-2.50E-02
0.00E+00 0.500
2.50E-02
5.00E-02

Contour interval= 2.50E-02


Boundary plot 0.000

0 1E 1

-0.500

Itasca Consulting Group, Inc.


Minneapolis, MN 55401
0.250 0.750 1.250 1.750 2.250 2.750
(*10^1)

Figure 3.20 Vertical displacement of grid around lined tunnel

FLAC Version 7.0


3 - 32 Theory and Background

If you type PRINT iface, you will note that normal stresses develop at almost all structural element
nodes along side A of the interface. However, a normal stress develops at only one gridpoint along
side B. This is because the structural element segments do not coincide with the zone edges along
the excavation (see Figure 3.16).
Stress initiation of an interface between a structure and the grid is performed when the stresses
in the grid are initialized. If stresses are initialized in a region that includes an interface, and
a stress gradient is given, the system may not be in exact equilibrium, because of the different
locations at which the stress components are evaluated. This applies to grid/grid interfaces as well
as structure/grid interfaces. For a model with a structure/grid interface, the additional stiffnesses
from the structural nodes will also produce a slight force imbalance that further necessitates some
stepping to reach equilibrium.
Beams can be deleted, and zones can be removed, if they are connected to an interface. Interface
elements will be re-formed and interface stresses will be deleted when zones adjacent to the interface
are nulled.
The following example illustrates the use of beams and interfaces to simulate an embedded retaining
wall (e.g., a sheetpile wall). Interfaces are attached to both sides of the beam element wall to
represent the soil/wall interface. The connection of the structural element nodes to the interface
nodes is order-dependent; be careful to identify the active side of the beam elements when attaching
each interface (see the INTERFACE command in Section 1.3 in the Command Reference). The
soil/wall properties chosen for this example are for demonstration purposes; actual values for wall
friction and adhesion can be found in the literature (e.g., Clayton et al. 1993). The data file is given
in Example 3.13.
The model is stepped to an equilibrium state with the beam and interfaces embedded within the
grid. Then, zones are nulled to simulate the excavation, and the model is stepped to equilibrium
again. Note that the tensile strength of the soil is set to a high value during these stages. This
minimizes the effect of inertial forces during the excavation stage. The tensile strength is reset to
the actual value of zero, and the model is stepped to equilibrium for the final solution. The resulting
displacements in the soil and moments in the wall are plotted in Figure 3.21.

Example 3.13 Modeling an embedded retaining wall


grid 12 11
mod mohr
prop den 2000 bulk 5e9 shear 1e9 coh 1e4 fric 30 tension 1e10
mod null i 5
ini x add -1 i 6 13
fix y j 1
fix x i 1
fix x i 13
ini syy -2.2e5 var 0 2.2e5
ini sxx -1.32e5 var 0 1.32e5
ini szz -0.88e5 var 0 0.88e5
set grav 10

FLAC Version 7.0


INTERFACES 3 - 33

;
struct beam begin 4.0,11 end 4.0 1 seg 10 prop 10
struct prop 10 e 2e9 a 1 i 0.08333 dens 2000
;
inter 1 as from node 11 to node 1 bs from 5,2 to 5,12
inter 2 as from node 1 to node 11 bs from 6,12 to 6,2
inter 1 kn 1e11 ks 1e11 coh 0 tbond 0 fric 30
inter 2 kn 1e11 ks 1e11 coh 0 tbond 0 fric 30
;
attach aside from 5,1 to 5,2 bside from 6,1 to 6,2
;
hist unbal
set large
solve
model null i 1 4 j 5 11
solve
prop tens 0.0
solve
plot hold xdis fill zero inv disp struct mom

JOB TITLE : TUNNEL LINER WITH GAP (*10^1)

FLAC (Version 7.00) 1.200

LEGEND
1.000
11-Jan-11 16:05
step 76590
-1.721E+00 <x< 1.272E+01
-1.721E+00 <y< 1.272E+01
0.800

X-displacement contours
-1.00E-02
-8.00E-03
-6.00E-03 0.600

-4.00E-03
-2.00E-03

Contour interval= 2.00E-03 0.400


(zero contour omitted)
Displacement vectors
max vector = 1.408E-02

0.200
0 2E -2
Beam Plot
Moment on
Structure Max. Value
0.000
# 1 (Beam ) 1.224E+05

Itasca Consulting Group, Inc.


Minneapolis, MN 55401
0.000 0.200 0.400 0.600 0.800 1.000 1.200
(*10^1)

Figure 3.21 Displacements of the soil behind the wall and moment distribution
in the wall

FLAC Version 7.0


3 - 34 Theory and Background

3.5.7 Interfaces and Groundwater Flow

The complete interaction of groundwater with an interface is not modeled. Fluid may flow across
an interface between two grids in the normal direction without resistance, provided that the two
surfaces are in contact; flow within the interface (in the parallel direction) is not modeled. For
both CONFIG gwow and non-CONFIG gwow modes, the fluid pressure at interface nodes is used
to determine the effective stress for the purpose of computing slip or tensile failure conditions, but
the fluid does not exert any mechanical force on the sides of the interface. There is no coupling
between joint movement and volume changes in the fluid. The program UDEC should be used if
full coupling between fluid and solid in an interface is required.
If one side of an interface is connected to beam elements and the other side to the grid, then
the interface/beam will act as an impermeable member (see Section 1.9.6 in Fluid-Mechanical
Interaction for an example). If it is required that flow occurs without resistance, then the interface
should be connected between the sub-grids, and the beam attached to the grid.
Interface/beam connections develop forces in terms of effective stresses. The pore pressure at a
structural node is obtained by interpolation from gridpoint pore pressures.

3.5.8 Access to Interface Variables

The interface data structure can be accessed by the user via the FISH include file INT.FIN (see
Section 4 in the FISH volume). The direct shear test (Example 3.9) uses INT.FIN to access the
interface stresses and displacements. An example is also given in Section 4.2 in the FISH volume.

FLAC Version 7.0


INTERFACES 3 - 35

3.6 References

Bandis, S. C., A. C. Lumsden and N. R. Barton. Fundamentals of Rock Joint Deformation, Int.
J. Rock Mech. Min. Sci. & Geomech. Abstr., 20(6), 249-268 (1983).
Barton, N. The Shear Strength of Rock and Rock Joints, Int. J. Rock Mech. Min. Sci. & Geotech.
Abstr., 13, 255-279 (1976).
Clayton, C. R. I., J. Milititsky and R. T. Woods. Earth Pressure and Earth-Retaining Structures,
pp. 145-148. London: Blackie Academic & Professional (1993).
Cundall, P. A., and R. D. Hart. Numerical Modeling of Discontinua, Engr. Comp., 9, 101-113
(1992).
Fossum, A. F. Technical Note: Effective Elastic Properties for a Randomly Jointed Rock Mass,
Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 22(6), 467-470 (1985).
Gerrard, C. M. Elastic Models of Rock Masses Having One, Two and Three Sets of Joints, Int.
J. Rock Mech. Min. Sci. & Geomech. Abstr., 19, 15-23 (1982b).
Gerrard, C. M. Equivalent Elastic Moduli of a Rock Mass Consisting of Orthorhombic Layers,
Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 19, 9-14 (1982a).
Itasca Consulting Group, Inc. UDEC (Universal Distinct Element Code), Version 5.0. Minneapo-
lis: ICG (2011).
Jaeger, J. C., and N. G. W. Cook. Fundamentals of Rock Mechanics, 2nd Ed. London: Chapman
and Hall (1969).
Kulhawy, Fred H. Stress Deformation Properties of Rock and Rock Discontinuities, Engineering
Geology, 9, 327-350 (1975).
Rosso, R. S. A Comparison of Joint Stiffness Measurements in Direct Shear, Triaxial Compression,
and In Situ, Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 13, 167-172 (1976).
Singh, B. Continuum Characterization of Jointed Rock Masses: Part I The Constitutive Equa-
tions, Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 10, 311-335 (1973).

FLAC Version 7.0


3 - 36 Theory and Background

FLAC Version 7.0

S-ar putea să vă placă și