Vous êtes sur la page 1sur 8

Modeling arching effects in narrow backfilled stopes with FLAC

L.Li, M.Aubertin & R.Simon

cole Polytechnique de Montral, Qc, Canada

B.Bussire & T.Belem

Universit du Qubec en Abitibi-Tmiscamingue, Qc, Canada

ABSTRACT: Numerical tools can be very useful to investigate the mechanical response of backfilled stopes.
In this paper, the authors show preliminary results from calculations made with FLAC-2D. Its use is
illustrated by showing the influence of stope geometry. The results are compared with analytical solutions
developed to evaluate arching effects in backfill placed in narrow stopes. Some common trends are obtained
with the two approaches, but there are also some differences regarding to magnitude of the stress
redistribution induced by fill yielding.

1 INTRODUCTION
Even if backfill has been placed in underground
stoping areas for many decades, it can be said that
backfilling still remains a growing trend in mining
operations around the world. This is particularly the
case in Canada where significant efforts have been
devoted, over the last twenty five years or so, to
improve our understanding of mining with backfill
(e.g., Nantel 1983, Udd 1989, Hassani & Archibald
1998, Ouellet & Servant 2000, Belem et al. 2000,
2002).
In recent years, the increased use of backfill in
mining has been fueled by environmental
considerations (e.g., Aubertin et al. 2002). Many
regulations now favor (and sometimes require) that
the maximum quantity of wastes from the mine and
the mill be returned to underground workings. This
practice may induce significant advantages, as it can
reduce the environmental impact of surface disposal
and the costs of waste management during mine
operation and upon closure.
The first purpose of mine backfill is nevertheless
to improve ground control conditions around stopes.
Various types of fills can be used to reach this goal,
each with its own characteristics. Backfill is often
required to offer some self support properties, so it
generally includes a significant proportion of binder
such as Portland cement and slag. But even the
strongest backfill is "soft" when compared to the
mechanical properties of the adjacent rock mass.
This difference in stiffness and yielding

characteristics between the two materials can be the


source of a stress redistribution in the backfill and
surrounding walls, as deformation of the backfill
under its own weight may create shear stresses along
the interface. For relatively narrow stopes, the load
transfer to the stiff abutments induce arching effects.
When this phenomenon occurs, the vertical stress
below the main arching area tends to become
smaller than the backfill overburden pressure, as
shown by in situ measurements (e.g., Knutsson
1981, Hustrulid et al. 1989).
The same type of arching behavior is also known
to occur in other types of structural systems, where a
relatively soft material (like soil and grain) is placed
between stiff walls; examples include silos and bins
(Richards 1966, Cowin 1977, Blight 1986), ditches
(Spangler & Handy 1984), and retaining walls (Hunt
1986, Take & Valsangkar 2001).
Arching effects and load redistribution can be
investigated using physical models, in situ
measurements, analytical solutions, and numerical
methods. The latter two approaches are particularly
attractive to identify the main influence factors, and
to evaluate how these may affect the load
distribution in and around backfilled stopes.
In a companion paper, the authors have proposed
simple equations based on the Marston (1930)
theory to evaluate the load distribution in stope
backfill (Aubertin et al. 2003). The results of
analytical calculations have been compared to
numerical modeling performed with a commercially
available finite element code. The calculation results

highlighted some important differences between the


two approaches, for the specific set of assumptions
adopted.
In this paper, the authors use FLAC-2D (Itasca
2002) to further advance our understanding of the
load transfer process in and around narrow
backfilled stopes. In these calculations, some of the
assumptions and input conditions differ from the
previous FEM calculations, including the use of a
somewhat more representative constitutive model
for the backfill. The mining sequence is also taken
into account. It is shown that for specific cases
amenable to analytical solutions, the calculated
results from both approaches are fairly close to each
other.

element; W represents the backfill weight in the unit


thickness layer. At position h, the horizontal layer
element is subjected to a lateral compressive force
C, a shearing force S, and the vertical forces V and
V + dV.
rock mass
Backfill
stope

dh
H

dh

rock mass

S
B

2 ARCHING EFFECTS
Arching conditions can occur in geomaterials such
as soil, jointed rock mass and backfill, when
differential straining mobilizes shear strength while
transferring part of the overburden stress to stiffer
structural components.
Arching typically occurs when portions of a
frictional material yields while the neighboring
material stays in place. As the yielding material
moves between stable zones, the relative movement
within the former is opposed by shear resistance
along the interface with the latter. The shear stress
generated along the contact area tends to retain the
yielding material in its original position. This is
accompanied by a pressure reduction within the
yielding mass and by increased pressure on the
adjacent stiffer material. Above the position of the
main arch, a large fraction of the total overburden
weight in the yielding mass is transferred by
frictional forces to the unyielding ground on both
sides.
Investigations on models and in situ
measurements have shown that the magnitude of the
stress redistribution depends to a large extent on the
deformation of the walls confining the soft material
(e.g., Bjerrum 1972, Hunt 1986).
A few analytical solutions have been developed to
analyze the pressure distribution since the
pioneering work of Janssen (1895) (see Terzaghi
1943 for early geotechnical applications). Among
these, the Marston (1930) theory was proposed to
calculate the loads on conduits in ditches (see also
McCarthy 1988). The authors have used this theory
to develop an analytical solution for arching effects
in narrow backfilled stopes (Aubertin 1999).
Figure 1 shows the loading components for a
vertical stope. On this figure, H is the backfill
height, B the stope width, and dh the size of the layer

layer element

void space

S
V + dV
B

Figure 1. Acting forces on an isolated layer in a vertical stope.

The force equilibrium equations for the layer


element provides an estimate of the stresses acting
across the stope (Aubertin et al. 2003). From these,
the vertical stress can be written as follow:
vh =

B 1 exp(2 Kh / B tan ' )

K
2 tan '

(1)

with
K = hh/vh

(2)

where vh and hh are the vertical and horizontal


stresses at depth h, respectively; represents the unit
weight of the backfill; ' is the effective friction
angle between the wall and backfill, which is often
taken as the friction angle of the backfill, 'bf.
Equations 1 and 2 constitute the Marston theory
solution. In this representation, K is the reaction
coefficient corresponding to the ratio of the
horizontal stress hh to vertical stress vh. This
reaction coefficient depends on the horizontal wall
movement and material properties. When there is no
relative displacement of the walls, the fill is said to
be at rest, and the reaction coefficient is usually
given by (Jaky 1948):
K = K0 = 1 sin'bf

(3)

where 'bf is the friction angle of the backfill. For


typical fill properties ('bf 30 to 35), K0 is much
smaller than unity.
If the walls move outward from the opening, the
horizontal pressure might be relieved, and the
reaction coefficient tends toward the active pressure
coefficient which can be expressed as (Bowles
1988):

with Ka < K0. If an inward movement of the walls


compress the fill, it increases the internal pressure.
Then, the reaction coefficient tends toward the
passive condition, which becomes (Bowles 1988):
K = Kp = tan2(45 + 'bf/2)

(5)

with Kp > 1 > K0.


In the above equations, it is assumed that cohesion
is low in the backfill, so it can be neglected; the fill
then behaves as a granular material. A cohesion
would increase Kp but decrease Ka.
Figure 2 shows values of vh and hh calculated
with Equations 1 and 2 (with K = K0 = 0.5), and
calculated with the overburden pressure (i.e., vh =
h and hh = K0 vh). It can be seen that the
overburden stress represents the upper-bound
condition, which is applicable for low fill thickness
(or for wide stopes). Typically, when H 2 to 3B,
the pressure near the bottom of the stope becomes
more or less independent of the depth of the fill, in
accordance with measurements in bins (Cowin
1977).
v h

0.5

h h

stress (MPa)

0.4

v h
h h

0.3

The dimensions of the opening are H = 45m and B =


6m, with a void of 0.5m left at the top of the stope.
The natural in situ vertical stress v in the rock mass
is obtained by considering the overburden weight
(for an overall depth of 250 m). The natural in situ
horizontal stress h is taken as twice the vertical
stress v, which is a typical situation encountered in
the Canadian Shield. The rock mass is
homogeneous, isotropic and linear elastic, while the
granular backfill obeys a Coulomb criterion. Figure
3b shows the stress-strain relation used with the
Coulomb plasticity model available in FLAC. This
constitutive behavior is different from the one used
for the finite element calculations presented in
Aubertin et al. (2003), which was of the elasticbrittle type. There are no interface elements in the
calculations made with FLAC (see discussion).
0.5m

(4)

void space

E = 300 MPa
v = 0.2
= 1800kg/m3

Marston theory

' = 30

overburden

c = 0 kPa
y
B=6m

for B = 6m

0.2

natural stresses
h = 2v

backfill

H = 45 m

K = Ka = tan2(45 - 'bf/2)

rock mass
(linear elastic)
E = 30 GPa
v = 0.3
= 2700kg/m3

depth = 250 m

rock mass

a)

0.1
0
0

h /B

Figure 2. Overburden pressures are compared to vertical (vh)


and horizontal (hh) stresses calculated with the Marston theory
(Eqs. 1 - 2), with 'bf = 30, = 0.02 MN/m3, and K = K0 = 0.5.

3 NUMERICAL CALCULATIONS
3.1 Vertical stope
In a companion paper, the authors have shown
some calculation results obtained with a finite
element code (Aubertin et al. 2003). Significant
differences have been revealed between the Marston
theory and these numerical calculations, which may
be explained, in part, by different assumptions
associated to the two approaches. In this paper, the
same geometry and material properties (Fig. 3a) are
used for the basic calculations made with FLAC-2D.

b)

Figure 3. a) Narrow stope with backfill (not to scale) used for


modeled with FLAC; the main properties for the rock mass and
backfill are given using classical geomechanical notations; b)
Schematic stress-strain behavior of the backfill (available as a
material model in FLAC).

The mining and filling sequence is considered as


follow in the numerical modeling. The whole stope
is first excavated, and calculations are then
performed with FLAC-2D to an equilibrium state.

Backfill is placed in the mined stope afterward, with


the initial displacement field set to zero when the
calculation is performed. In this manner, wall
convergence before backfilling is not included in the
calculations (this assumption is discussed in Section
4).
Figure 4 shows the vertical stress (Fig. 4a) and
horizontal stress (Fig. 4b) distribution in the
backfilled stope. As can be seen, the vertical and
horizontal stresses show a non uniform distribution.
At a given elevation, both stresses tend to be lower
along the wall than at the center. The stresses along
the central line increase more slowly than the
overburden pressures with depth. This indicates that
arching does take place in this backfilled stope.

and the Marston theory solutions. As expected, the


overburden stress is fairly close to analytical and
numerical results when the backfill depth is small.
At larger depth, arching effects become important
and the vertical and horizontal stresses tend to be
lower than those due to the overburden weight of the
fill. However, the numerical results indicate that the
Marston theory typically overestimates the amount
of stress transfer, hence underestimating the
magnitude of the vertical stress yy and of the
horizontal stress xx along the stope central vertical
line (Fig. 5). Along the walls (Fig. 6), the horizontal
stress is also underestimated by the Marston theory,
while the vertical stress component yy would be
overestimated for the active and at rest cases, with K
= 1/2 or 1/3, respectively (and underestimated with
K = 3, but the passive case is not representative of
this system behavior).
a)

mo d eling w ith F LA C - 2 D

0 .8

yy

(M P a)

o verb urd en stress


0 .6

M arsto n theo ry
K = 1 /3
K = 1 /2
K = 3

0 .4
0 .2
0
0

b)

xx

18

27
h (m)

36

45

modeling w ith F LA C - 2D
overb urden stress
M arston theory

0.3

(M Pa)

a)

0.2
K = 3
0.1
K = 1/2
K = 1/3

0
0

18

27
h (m)

36

45

Figure 5. Comparison of the stresses calculated along the


vertical central line, at different elevations h, with the
analytical and numerical solutions: a) vertical stress yy; b)
horizontal stress xx.
b)
Figure 4. Stress distribution in the backfilled stope calculated
with FLAC-2D: a) vertical stress yy; b) horizontal stress xx.

Figures 5 and 6 present modeling results for


stresses along the full height, with the overburden

Figure 7 shows the stress distribution on the floor


of the stope, as obtained from the numerical and
analytical solutions. It can be seen that the
overburden pressure exceeds the stress magnitudes
given by the Marston theory (with K = 1/2 and 1/3),
which is in fair agreement with the numerical
simulations.

4.1 Influence of mining sequence

yy

(M P a)

o verb urd en stress


M arsto n theo ry

0 .6

K = 1 /3
K = 1 /2
K = 3

0 .4
0 .2
0
0

b) 0.2

18

27
h (m)

36

45

modeling with F LAC - 2D

(M Pa)

overburden stress

xx

M arston theory
K =3

0.1
K = 1/2
K = 1/3
0
0

18

27
36
h (m)

45

Figure 6. Comparison of the stresses on the wall calculated at


different elevations h, with the analytical and numerical
solutions: a) vertical stress yy; b) horizontal stress xx.

In the numerical calculations shown in a


companion paper (Aubertin et al. 2003), the mining
sequence was not taken into account, so the wall
convergence due to elastic straining of the rock mass
was imposed on the fill. This created an increase of
the mean stress in the fill, while vertical and
horizontal stresses locally exceeded the overburden
pressure and the Marston theory solution (near midheight of the stope).
Modeling in this manner implies that the backfill
is placed in the stope before wall displacement takes
place. For a single excavation stope, this is not a
realistic representation (at least for hard rock
masses). However, with a cut-and-fill mining
method where the mining slices (or layers) are
relatively small compared to the whole height of the
stope, filling is usually made quickly after each cut.
In this case, wall convergence after each cut
compress the fill already in place (Knutsson 1981,
Hustrulid et al. 1989). The inward movement of the
walls may then create a condition closer to the
passive pressure case.
a)

3.2 Inclined stope

4 DISCUSSION

0.8
K = 1/3
K = 1/2

K =3

0.4

0
0

b) 0.3

x (m)

modeling with F LAC - 2D

(M P a)

overburden stresses
M arston theory
0.2
K = 1/2

K =3

xx

Mining stopes are rarely vertical. The inclination


of the foot-wall and hanging-wall may have a nonnegligible effect on the load distribution.
Figure 8 shows the geometry of an inclined
backfilled stope modeled with FLAC-2D (a similar
stope was also modeled with the FEM code see
Aubertin et al. 2003). The rock mass and fill
properties as well as the in situ natural stresses are
identical to the previous case (see Fig. 3).
Figure 9 shows numerical calculations and results
based on overburden pressure and on the Marston
theory solution (without modification for
inclination). The horizontal stress calculated with
FLAC-2D along the inclined central line of the stope
is fairly close to the analytical solution (Fig. 9a), but
the vertical stress is underestimated by the Marston
theory (see Fig. 9b). Hence, modifications could be
required to apply such analytical approach to the
case of inclined stopes.

modeling with F LAC - 2D


overburden stresses
M arston theory

1.2

(M Pa)

mo d eling w ith F LA C - 2 D

0 .8

yy

a)

0.1
K = 1/3
0
0

x (m) 4

Figure 7: Stresses calculated at the bottom of the vertical stope,


with the analytical and numerical solutions; a) vertical stress
yy; b) horizontal stress xx.

0.5m

Figs. 5 to 7), at least for preliminary design


calculations.

void space
h

backfill

4.2 Marston theory limitations


H = 45 m

rock mass

B=6m
rock mass
v

stope

y
60

h = 2v

depth = 250 m

Figure 8. The inclined backfilled stope modeled with FLAC2D (properties are given in Fig. 3).

a)

modeling with F LAC - 2D


overb urden
M arston theory

xx

(M P a)

0.2

K = 3
0.1
K = 1/2
K = 1/3

h (m)

0
0

b)

0 .4

18

27

36

45

yy

(M P a)

mo d eling w itho ut F LA C - 2 D
o verb urd en
M arsto n theo ry

0 .2
K = 3
K = 1 /3

Analytical solutions can be useful engineering


tools as they are generally quick, direct and
economic when compared to numerical methods.
However, analytical solutions are only available for
relatively simple situations and may involve strong
simplifying hypotheses. For instance, with the
Marston theory, the shear stress along the interface
between the rock and fill is deduced from the
Coulomb criterion (see details in Aubertin et al.
2003). Its value then corresponds to the maximum
stress sustained by the fill material, as postulated in
the limit analysis approach (e.g., Chen & Liu 1990).
However, numerical simulations indicate that this
assumption is not fully applicable. Figure 10 shows
that for the vertical stope analyzed here the
maximum shear stress is only reached near the
bottom part of the stope. Hence, arching effect and
stress redistribution are thus exaggerated.
Another important assumption in the Marston
theory is that both the horizontal and vertical
stresses are uniformly distributed across the full
width of the stope. Results shown in Figure 11
indicate that this is in accordance with numerical
calculations for the horizontal stress component
(Fig. 11a), but not for the vertical stress which
shows a less uniform distribution (Fig. 11b). Also,
this simplified theory considers that the reaction
coefficient, K, depends exclusively on the fill
property and not on the position in the stope. Results
shown in Figure 12 indicate that this hypothesis is
not too far from the numerical results. Near the
boundary, the value of K would nevertheless be
better described by a K value between Ka and K0.

K = 1 /2

27
36
45
h (m)
Figure 9. Comparison between stresses obtained with
numerical and analytical solutions along the central line of the
inclined stope: a) horizontal stress xx; b) vertical stress yy.

M arston theo ry

18

When a stope is excavated in a single step, wall


convergence essentially takes place before any
backfilling. If the rock mass creep deformation is
negligible, the numerical modeling approach
presented here is more appropriate. In this case, the
Marston theory, with the "at rest" reaction
coefficient (K = K0) can be used to estimate the
induced stresses in a narrow vertical backfill (see

0.02
h (m)

(M P a)

0
- 0.02

18

27

36

45

xy

mo deling w ith F LA C - 2 D

0.04

- 0.04
- 0.06
K = 3

K = 1 /3

K = 1 /2
- 0.08
Figure 10. Comparison of shear stress distribution along the
wall.

Work is underway to modify the analytical


solution to extend the use of the Marston theory to
more general cases.
0.12

at 3/4 H
at 1/2 H
at 1/4H

0.08

xx

(M P a)

0.1

4.4 Interface elements along the walls

0.06
modeling with F LAC - 2D
0.04
0

x (m) 4

b)
at 3/4H
at 1/2H

0.2

at 1/4H

yy

(M Pa)

0.3

0.1
modeling with F LAC - 2D
0
0

x (m)

Figure 11. Distribution of lateral pressure xx a) and vertical


stress yy b) obtained with FLAC-2D across the full width at
different elevations of the vertical stope.

4.3 Constitutive behavior


The reliability of any numerical calculations
depends, to a large extent, on the representativity of
the constitutive models used for the different
materials (and on the corresponding parameters
values). In this paper, a Coulomb plasticity model
(see Fig. 3) was employed for the fill material. This
model is representative of some aspects of the
mechanical behavior of backfill, such as the
nonlinear relationship between the stress and strain
(e.g., Belem et al. 2000, 2002). However, this
simplified model neglects some important
characteristics of the media, including its pressure
dependent behavior under relatively large mean
stresses. More representative models, such as the
modified Cam-Clay model, are built in FLAC (e.g.,
Detournay & Hart, 1999). However, the application
of such model is not straightforward because of the
difficulties in obtaining the relevant material
parameters. The influence of cohesion due to

As was done with a finite element code in a


previous investigation (Aubertin et al. 2003), some
calculations were also performed with interfaces
included in FLAC-2D, to represent the contact
between backfill and rock mass.
Preliminary results (not shown here) indicate that
the presence of interfaces along the walls and floor
of the stope, which allow localized shear
displacements, has relatively little influence on the
stress distribution in the stope and at its boundary.
Some differences between the cases shown here and
models with interfaces nevertheless appear near the
bottom and top of the stope where some stress
reorientation and concentration seem to take place.
This aspect however requires further investigation,
as the representativity of the (Coulomb) strength
criterion and the numerical stability of the
calculations along these elements need more study.
0 .5

at rest

a)

cementation and possible oxidation of the fill


material may also be relevant to include in the
analyses.
An interesting aspect of FLAC is that it allows
user-defined models, which can be introduced with
the language FISH. The authors are now working on
introducing in FLAC a multiaxial, porosity
dependent criterion (Aubertin et al. 2000, Li &
Aubertin 2003) for the yielding and failure
conditions of geomaterials. This aspect will be
presented in upcoming publications.

0 .3

at
at
at
at

f lo o r
1 /2 H
1 /2 H
3 /4 H

active
mod eling w ith F LA C

0 .1
0
2 x (m) 4
6
Figure 12. Reaction coefficient K obtained with analytical and
numerical solutions across the full width of the vertical stope at
different elevations h.

5 CONCLUSION
In this paper, numerical simulations have been
performed with FLAC-2D for a vertical and an
inclined backfilled stope geometry. The results are
compared to the Marston theory solutions. It is
shown that the results obtained with the Marston
theory can be considered as acceptable, especially

for preliminary calculations. Nevertheless, the


numerical results also reveal that the Marston theory
tends to overestimate arching effect, and thus
underestimate the stress magnitude near the bottom
of backfilled stope. Also, the influence of the mining
sequence can not be introduced in the Marston
theory. The numerical works shown here and in a
companion paper (Aubertin et al. 2003) demonstrate
that the filling sequence can significantly influence
the stress distribution in and around filled stopes.
For inclined stopes, the Marston theory is of limited
use to estimate the stress magnitude. Additional
work is underway into both analytical and numerical
solutions to better describe the behavior of
backfilled stope. More work is also needed to study
the rock-fill interface behavior and the actual field
response of backfill in stopes.
Other important issues also remain to be resolved,
including the possible degradation of the arch due to
low pressure (and tensile stresses), the influence of
water flow and distribution in backfilled stopes, the
evolving properties of the fill material (especially
when sulfide materials are present), the dynamic
response of the backfill, and the forces generated on
retaining structures.
ACKNOWLEDGEMENT
Part of this work has been financed through grants
from IRSST and from an NSERC Industrial Chair
(http://www.polymtl.ca/enviro-geremi/).
REFERENCES
Aubertin, M. 1999. Application de la Mcanique des Sols pour
l'Analyse du Comportement des Remblais Miniers
Souterrains. Short Course (unpublished notes), 14e
Colloque en Contrle de Terrain, Val-d'Or, 23-24 mars
1999. Association Minire du Qubec.
Aubertin, M., Bussire, B. & Bernier, L. 2002. Environnement
et Gestion des Rejets Miniers. Manual on CD-ROM, Presses
Internationales Polytechniques.
Aubertin, M., Li, L., Arnoldi, S., Belem, T., Bussire, B.,
Benzaazoua, M. & Simon, R. 2003. Interaction between
backfill and rock mass in narrow stopes. Soil and Rock
America 2003, Cambridge, Mass (in press). Rotterdam:
Balkema.
Aubertin, M., Li, L. & Simon, R. 2000. A multiaxial stress
criterion for short term and long term strength of isotropic
rock media. International Journal of Rock Mechanics and
Mining Sciences, 37: 1169-1193.
Belem, T., Benzaazoua, M. & Bussire, B. 2000. Mechanical
behavior of cemented paste backfill. Proceedings of the 55th
Canadian Geotechnical Conference, 1: 373-380. Canadian
Geotechnical Society.
Belem, T., Benzaazoua, M., Bussire, B. & Dagenais, A.M.
2002. Effects of settlement and drainage on strength
development within mine paste backfill. Tailings and Mine
Waste '02: 139-148. Sets & Zeitlinger.

Blight, G.E. 1986. Pressure exerted by materials stored in silos.


Part I: coarse materials. Gotechnique, 36(1): 33-46.
Bjerrum, L. 1972. Embankments on Soft Ground. Proceedings
of the ASCE Specialty Conference on Performance of Earth
and Earth Supported Structures, 2: 1-54. New York: ASCE
Bowles, J.E. 1988. Foundation Analysis and Design. McGrawHill.
Chen, W.F. & Liu, X.L. 1990. Limit Analysis in Soil
Mechanics. Amsterdam: Elsevier.
Cowin, S.C. 1977. The theory of static loads in bins. Journal of
Applied Mechanics, 44: 409-412.
Detournay, C., & Hart, R. (eds) 1999. FLAC and Numerical
Modeling in Geomechanics - Proceedings of the
International FLAC Symposium, Minneapolis, Minnesota, 13 September 1999. Rotterdam: Balkema.
Hassani, F. & Archibald, J.H. 1998. Mine Backfill. CIM, CDROM.
Hunt, R.E. 1986. Geotechnical Engineering Analysis and
Evaluation. New York: McGraw-Hill Book Company.
Hustrulid, W., Qianyuan Y. & Krauland, N. 1989. Modeling of
cut-and-fill mining systems Nsliden revisited. In F.P.
Hassani, M.J. Scoble & T.R. Yu (eds), Innovation in Mining
Backfill Technology: 147-164. Rotterdam: Balkema.
Itasca 2002. FLAC - Fast Lagrangian Analysis of Continua;
Users Guide. Itasca Consulting Group.
Jaky, J. 1948. Pressure in silos. Proceedings of the 2nd
International Conference on Soil Mechanics and
Foundation Engineering, 1: 103-107. Rotterdam: Balkema.
Janssen, H.A. 1895. Versuche ber Getreidedruck in
Silozellen. Zeitschrift Verein Ingenieure, 39: 1045-1049.
Knutsson, S. 1981. Stresses in the hydraulic backfill from
analytical calculations and in-situ measurements. In O.
Stephansson & M.J. Jones (eds), Proceedings of the
Conference on the Application of Rock Mechanics to Cut
and Fill Mining: 261-268. Institution of Mining and
Metallurgy.
Li, L. & Aubertin, M. 2003. A general relationship between
porosity and uniaxial strength of engineering materials.
Canadian Journal of Civil Engineering (in press).
Marston, A. 1930. The theory of external loads on closed
conduits in the light of latest experiments. Bulletin No. 96,
Iowa Engineering Experiment Station, Ames, Iowa.
McCarthy, D.F. 1988. Essentials of Soil Mechanics and
Foundations: Basic Geotechnics. 4th edition, Prentice Hall.
Nantel, J.H. 1983. A review of the backfill practices in the
mines of the Noranda Group. In S. Granholm (ed), Mining
with Backfill: Proceedings of the International Symposium
on Mining with Backfill: 173-178. Rotterdam: Balkema,
Ouellet, J. & Servant, S. 2000. In-situ mechanical
characterization of a paste backfill with a self-boring
pressuremeter. CIM Bulletin, 93(1042): 110-115.
Richards, J.C. 1966. The Storage and Recovery of Particulate
Solids. Institution of Chemical Engineers, London.
Spangler, M.G. & Handy, R.L. (1984). Soil Engineering.
Harper & Row.
Take, W.A. & Valsangkar, A.J. (2001). Earth pressures on
unyielding retaining walls of narrow backfill width.
Canadian Geotechnical Journal, 38: 1220-1230.
Terzaghi, K. 1943. Theoretical Soil Mechanics. John Wiley &
Sons.
Udd, J.E. 1989. Backfill research in Canadian mines. In F.P.
Hassani, M.J. Scoble & T.R. Yu (eds), Innovation in Mining
Backfill Technology: 3-13. Rotterdam: Balkema.

Vous aimerez peut-être aussi