You are on page 1of 14

SPE 69714

An Integrated Reservoir Model for Sand Production and Foamy Oil Flow During
Cold Heavy Oil Production
Yarlong Wang, SPE, Carl C. Chen, SPE, Petro-Geotech Inc., Calgary, and Maurice B. Dusseault, SPE, Porous Media
Research Institute, University of Waterloo and PRISM Production Technologies Inc.
Copyright 2001, Society of Petroleum Engineers, Inc.
This paper was prepared for presentation at the 2001 SPE International Thermal
Operations and Heavy Oil Symposium held in Margarita, Venezuela, March 12-14, 2001
This paper was selected for presentation by the SPE Program Committee following review
of information contained in an abstract submitted by the author(s). Contents of the paper
as presented, have not been reviewed by the Society of Petroleum Engineers and are
subject to correction by the author(s). The material, as presented, does not necessarily
reflect any position of the Society of Petroleum Engineers, its officers, or members.
Papers presented at SPE meetings are subject to publication review by the Editorial
Committee of the Society of Petroleum Engineers. Permission to copy is restricted to an
abstract of not more than 300 words. Illustrations may not be copied. The abstract should
contain conspicuous acknowledgment of where and by whom the paper was presented.
Write Librarian, SPE, P.O. 833836, Richardson, TX 75083-3836 U.S.A., fax 01-214-952-
9435.

Abstract
Continuous sand production and foamy oil behavior are
both believed to be key factors for the enhanced non-
thermal fluid production in unconsolidated heavy oil
reservoirs in Canada (Alberta and Saskatchewan). The
same mechanisms are likely to be active in similar heavy oil
strata in Venezuela (Faja del Orinoco), Oman, China (Bohai
Bay), and elsewhere. Field experience indicates that
fundamental understanding of sand production mechanisms,
reservoir fabric alteration, foamy oil behavior, pressure
gradient changes, and stress changes are key to successful
operations involving massive continuous sanding. Inter-
relating these factors requires coupling of geomechanics
and fluid flow processes.
An integrated approach incorporating a three-phase,
three-dimensional black-oil model coupled with a
geomechanics model is introduced in this article. Piping
channels (wormholes) are postulated to develop from
perforations when pressure gradients exceed the residual
cohesion of the sand. An elastoplastic constitutive model is
used to describe the reservoir material before seepage forces
liquefy and suspend the sand particles at the advancing tips
of wormholes. The hemispherical wormhole tip is
postulated to propagate as long as a critical tip pressure
gradient is exceeded. A slurry transport model is used to
describe the flow inside the wormholes. Field data from
Frog Lake, Alberta are used to validate the model, and it
appears that the simulation can match the field data
remarkably well.
Introduction
Enhanced heavy oil production can be achieved by heating,
which reduces oil viscosity and facilitates flow. However,
because CH
4
is commonly used to generate steam, thermal
operating costs are high and rising (CDN$10-15/bbl). It is
less well-known that under the right conditions formation
flow characteristics can be improved through non-thermal
massive sand co-production, referred to as CHOP (Cold
Heavy Oil Production).
CHOP has been widely used for a decade in Alberta and
Saskatchewan,
1,2,3,4,5,6,7,8
and operating costs have now been
cut to CDN$4-7/bbl. However, poor understanding of the
CHOP process, a recovery rate limited to ~12-20% in
appropriately screened reservoirs, and difficulties in well
management (i.e. repeated workovers) have been weak
points for this technology. Improving these aspects,
particularly the understanding of CHOP mechanisms, can
lead to direct economic benefits.
This article introduces a model to address reservoir fluid
mobility changes arising from sanding, and pressure drive
changes arising from foamy-oil flow. Simulations are based
on a general three-dimensional, three-phase, black-oil
production model coupled to a one-dimensional slurry flow
model. The latter represents the wormholes (or slurry
transport zone), and the material balance for solids transport
is established based on a two-dimensional classic
geomechanics model.
In the discussions to follow, comparisons are made
between primary production without sand and production
with sand influx allowed (CHOP).
An example from Frog Lake, Alberta is used to validate

2 Y. Wang, C. Chen, M.B. Dusseault SPE 69714

the model. Table 1 shows the reservoir conditions, also
used as the input parameters for the simulations.
Background
Sand production has long been considered undesirable;
9,10

but it is associated with high drawdown and production
rates, which are required for faster, more profitable oil
production. During field operations, sanding risk
management is required, thus engineers try to estimate sand
production risk before drawdown or production rate is
increased. Many existing mathematical models are attempts
to predict sand onset under various operational
conditions.
11,12,13,14
To reduce risk of sand influx, the cased
wellbore may be perforated with small entry ports and in a
preferred orientation with respect to the stress fields. Also,
various sand exclusion methods may be used (screens,
gravel packs, frac-and-pack, slotted liners, resin
squeezes).
Eliminating sand, however, usually impairs production
rate, particularly from viscous oil deposits in unconsolidated
sands. In such cases, sand ingress is necessary to sustain
economic non-thermal production rates.
15,16
Also, in
conventional oil and gas production, rather than precluding
sand ingress, engineers start to apply strategies to balance
the oil and sand production economically.
17
This implies
proper sand management in terms of completion, lifting,
separation, and disposal. The goal is oil rate maximization
while keeping solids production within economic and safe
limits.
High non-thermal primary heavy oil rates in Canada
have been attributed primarily to sand production and
foamy oil behavior.
18,19,20,21,
To evaluate reservoirs and
forecast future production, it is necessary to understand the
physical mechanisms and develop practical tools for design.
Some Fundamental Issues in Model Development
Consolidated sandstones and unconsolidated oil sand
formations may be viewed respectively as strong and weak
strata, and vastly different critical drawdowns for sanding
will arise with such strength differences. In unconsolidated
strata, grain plucking because of high pressure gradients is
possible, but not in highly cemented strata, unless high
shear stresses have already caused fabric disaggregation.
Thus, rocks in different formations behave differently, and
different geomechanical yielding criteria and constitutive
models are needed.
Rather than steady-state flow and constant-density
assumptions, transient fluid flow and pressure-dependent
density are incorporated to simulate time-dependent oil and
gas production. This is needed because sanding in many
circumstances correlates with changes in pressure gradient,
flow regime, or saturation.
22
For example, in reservoirs
containing mainly oil, free gas release can initiate only after
the bottom-hole pressure is reduced below the bubble point.
In addition to the general increase in effective stress as the
result of drawdown, near well-bore pressure gradients rise
as the fluid compressibility becomes more gas-dominated,
and any resultant sand influx is clearly linked to this
transient behavior.
Sand production in this model is described in terms of
the coupled mechanisms of fluid flow, stress evolution,
large deformation, and fluidized slurry transport. Detailed
descriptions may be found in the references and appendices;
space limitations permit only a brief mathematical treatment
herein.
Solids production has been hypothesized to be the result
of mechanical skeleton disaggregation arising from
cohesion loss, dilation and crushing, combined with viscous
drag forces that dislodge the particles and carry them into a
cavity, i.e., a wormhole, perforation tunnel, or some other
cavity filled with fluid.
23
Cavity stability solutions are thus
needed to quantify the conditions under which sand
initiation is likely to occur. Such solutions are obtained by
specifying boundary and initial conditions, geometry and
constitutive laws, then using the stress equilibrium equation,
mass conservation and Darcys law.
24,25

Sand production initiation is assumed to occur when
effective radial stress equals tensile strength (
r
= T
o
).
26
In
practice, T
o
= 0 may be assumed for simplicity, despite the
fact that most intact reservoirs possess some tensile
strength. This is usually justified, as sand production
cannot occur before shear yield; shearing damages the
formation, causes dilation, and destroys natural cohesion. A
T
o
= 0 assumption, however, may be inadequate when
multiphase flow is involved; then, capillary pressure may
provide an apparent T
o
for the particulate assemblage.
Nevertheless, assuming such a capillary pressure is
negligible in comparing to the drag forces induced, both
critical drawdown and flow rate are predicted based on the
condition
r
= 0 (i.e. T
o
= 0).
Functionally, the critical condition
r
= 0 is calculated
using factors such as
i. permeability,
ii. fluid viscosity,
iii. water and gas contents,
iv. drawdown and production rate,
v. cavity radius,
vi. formation cohesion,
vii. friction angle, and perhaps
viii. other nonlinear geomechanical parameters.
Bottom-hole T and initial in situ stresses may be linked
indirectly to this critical condition, and for horizontal
and inclined wells, perforation tunnel geometry,
orientation, and perforation density can also affect the
sand production potential. Our objective in this article is
SPE 69714 An Integrated Reservoir Model for Sand Production and Foamy-Oil Flow During Cold Heavy-Oil Production 3

to demonstrate how to relate solids flux, slurry transport,
and geomechanical behavior to production
enhancement. One may refer to the published technical
literature for details as to how various other factors may
be incorporated into the analysis.
To simulate sanding initiation and transport, a 3-D,
three-phase flow model coupled to a 2-D geomechanics
model and a 1-D slurry transport model are used. Material
balance is considered in three different zones: the wormhole
slurry transport zone, the hemispherical wormhole tip
propagation zone, and the intact formation. The reservoir is
assumed to deform elastoplastically, that is, as an elastic but
ductile solid material, until the radial seepage force detaches
the sand into the cavity, at which point the solid material
becomes disconnected from other solid material, and is
transported in a particulate form in a fluid as a slurry.
Dissolved gas is assumed to exist inside the oil as a
separate component, but not at early time when pore
pressure is below the bubble point. When gas exsolves, it
increases the solution-gas (foamy oil) compressibility and
reduces the fluid density. Three dimensionally convergent
flow and gas exsolution near the wormhole tip sustain
abnormally large pressure gradients during early production
stages. This pressure drive induced by foamy flow
enhances fluid production and promotes more sand
production. More sand production improves reservoir
mobility and effective reservoir transmissivity.
We will outline model development and demonstrate
that simulations using the proposed integrated model closely
emulate field results for sand cut magnitude, sand
production behavior over time, and the enhanced fluid
production response of the well. The major conclusions of
the analyses and the model presented in the article are stated
here so that subsequent developments are clearer:
Wormhole propagation is controlled by a critical
pressure at the tip. However, the tip pressure will
gradually increase as the wormhole lengthens,
eventually approaching the intact reservoir pressure.
Thus, the pressure gradient at the tip gradually drops
until it is insufficient to destabilize the formation, at
which point wormhole growth ceases.
The maximum fluid flow enhancement effect does not
coincide with the peak sand rate, which is consistent
with field observations.
Foamy oil flow not only provides additional driving
energy for oil production, it also triggers more sand
production, thereby further improving reservoir
mobility.
These findings also suggest that controlling production
rate at different sand production stages can help control
sand cut, reservoir enhancement, and ultimately affect the
overall recovery efficiency of the CHOP process.
Model Description
Many researchers have attacked or are actively attacking
the complicated problem of simulating sand production
from reservoirs, and only some of the work is referenced
here.
27,28,29,30,31
A reservoir model incorporating multiphase
flow is needed to simulate the oil/gas/water flow in the
continuous reservoir formation. Slurry transport is assumed
to dominate the production inside the high-permeability
zone (channel) near a well, which is separated from the
continuous formation by a representative hemispherical
cavity front (channel tip). Material balance is set up at the
tip between the slurry traveling along the wormhole and
solid and fluids into the wormhole through the tip. Thus
three models, which dominate the flow inside wormhole,
wormhole tip region, and inside the continuous reservoir
formation respectively, are presented here.
1. Reservoir model
A 3-D, three-phase black-oil model is used to simulate
fluid flow in the reservoir. In general, three variable groups
are of interest: formation strains (or effective stresses), pore
pressures, and phase saturations. Lewis and Sukirman
32

developed a fully coupled formulation containing these
variables to simulate reservoir compaction. The solid
displacement and pressures were identified as primary
variables, whereas saturations were treated as intermediate
variables. Earlier, Lewis et al.
33
1
also considered a coupled
two-phase flow problem, focusing on cases of no phase
changes and linear elastic rock deformation. Similarly Li et
al.
3433
also studied a two-phase flow problem in which the
primary variables were displacement, wetting phase
pressure, and saturation.
Following Li et al., we use skeleton displacement,
wetting phase pressure, and wetting phase saturation as the
primary unknowns. To simplify calculations, a table is
constructed to relate the saturation of each phase (oil and
gas) to the calculated pressure, which in turn is fully
coupled to the displacement. The computational procedure
may be summarized as follows:
Calculate the displacement and wetting phase pore
pressure with the initial saturation.
Calculate the saturation changes that occur as the pore
pressure changes.
Modify saturation-dependent properties such as relative
permeability.
Calculate the pore pressure and the skeleton
deformation again with the new reservoir properties
and iterate to numerical precision.
Add a time increment (incorporating boundary
condition changes if desired) and return to the first
computational step.



4 Y. Wang, C. Chen, M.B. Dusseault SPE 69714

In this scheme, both boundary conditions changes (i.e.
wellbore pressure or drainage radius) and deformation
changes (i.e. pore volume changes) induce pore pressure
changes. In the simulations, however, the oil and gas
components in the oil phase cannot change before the gas
bubble point is reached. When the pressure drops below the
bubble point, the basic nature of oil-water two-phase flow
remains unchanged, except that the non-water phase now
also contains free solution gas as well as oil, and the
saturation of each component (gas and oil) changes with
pore pressure.
As noted above, the pore pressure may evolve because
of pore volume changes and boundary condition changes.
If the pore pressure evolves, the effective stresses also
changes. This may degrade the cohesion, it may cause
dilation, or it may even strengthen the rock, depending on
the magnitude and nature of the deformation and the stress
path followed. Also, phase saturations may change, which
alter the relative permeabilities, and this may cause pressure
changes, stress changes, fabric, and other factors changes.
It appears that any consistent physics-based model of
sanding in oil and gas reservoirs must incorporate this
intrinsic strongly coupled behavior.
Now that the overall process has been briefly outlined, a
more precise mathematical treatment can follow.
The specific mass discharges (mass discharge per unit
area of a spatially fixed coordinate system) are denoted by
m
s
and m
j
, where s represents the solid phase and j = w, o, g
for water, oil and gas phases, respectively (symbology is
found in the Nomenclature section). These specific mass
discharges are related to the specific volume discharges
(volume discharge per unit area of a coordinate system
fixed in space) q
s
and q
j
by:
34,36
) 1 ( and
j j j s s s
q m q m = =
In an elementary volume V, the pore volume is V and the
solid volume is (1 - )V. Given saturation of phase j as S
j
,
the phase j fluid volume in the elementary volume V is
S
j
V. The specific mass discharges are related to the
displacement velocities as follows:
) 2 ( S and ) 1 (
j j j s s
v q v q = =
The relative density
j
of phase j is the mass of phase j per
unit volume. Since density
j
is defined as the mass of
phase j per unit volume of that phase, the relative densities
are given by:
) 3 ( S and ) 1 (
j j j s s
= =
Of course, specific mass discharges must also satisfy mass
conservation. To simplify the geomechanics/reservoir
coupling, only immiscible flow without chemical reactions
is considered. Hereafter, j = w, o and g denote immiscible
water, oil and gas components, and the conservation of
mass of each component requires that:
) 4 ( 0
t
and 0 q
t
j j
j
s s
s
= + +


= + +


q m m
or
) 5 ( 0 ] S [
t
] S [
and 0 ] ) 1 [(
t
] ) 1 [(
j j j j
j j
s s s
s
= + +


= + +


q v
q v
with
) 6 ( 1 S S S
o w g
= + +
The equation of fluid motion has the following form:
) 7 ( p
kk
) ( S
j
j
rj
s j j j

= = v v u
The phase pressures are related to capillary pressures
through the equations:
) 8 ( p p p and p p p
o g cog w o cwo
= =
In conventional reservoir engineering, where the solid phase
is assumed stationary (v
s
= 0), u
j
corresponds to the Darcy
velocity. For deformable media where v
s
0, Darcy flow
describes fluid flow relative to the medium or solid phase
flow velocity, v
s
, as used in Equation (7). Thus when
deformation (i.e., porosity change) is considered, the fluid
flow terms in the conservation equations also depend on the
deformation through v
s
(and consequently porosity).
Equation (5) can be expanded into a form containing terms
involving the solid velocity, which can be related to the
porosity change. If the deformation is small, the solid
velocity can be neglected and simplified equations may be
obtained.
Consider mass conservation for the solid phase as well
as a single fluid component in phase j; i.e., Equations (5),
where terms (1 - )v
s
and v are equivalent to the solid and
fluid bulk volumetric fluxes, respectively. Mass
conservation for the fluid phase in Equation (5) can be
expressed as:
) 9 ( 0 S
) S ( )
kk
(
t
S (
j s j j
j j s j
rj
j
) j j
= + +
+


q v
v

Dropping the third term in (9), which contains the solid
velocity and the gradient of the combined effect of
saturation, porosity, and fluid density, the above equation
can be simplified as:
) 10 ( 0 ) p
kk
(
t
p
S C
t
S
t
S
j j
rj
j
j
j j
j
v
j
= +


q
where
SPE 69714 An Integrated Reservoir Model for Sand Production and Foamy-Oil Flow During Cold Heavy-Oil Production 5

) 11 (
t dt
] ) 1 [( d
) 1 (
1
v s
s
s


=


= v
Assuming the solid skeleton is stationary, i.e. v
s
0,
Equation (9), in which the porosity change appears
explicitly only in the accumulation term, is the standard
mass conservation equation used in most reservoir
simulators:
35

) 12 ( 0 ) S ( )
kk
(
t
S (
j j s j
rj
j
) j j
= +


v
This effectively decouples porosity changes from the fluid
flow term, facilitating an efficient treatment of the coupling.
The approach to simulation by fully coupling the porosity
change and fluid flow will be addressed elsewhere.
36

In Equation (9), absolute permeability and porosity are
functions of the deformation. The permeability is assumed
to change with porosity based on the Kozeny-Poiseuille
correlation.
37

2. Wormhole Propagation Model
A material balance condition is applied to the moving
fluidized slurry or wormhole tip: the slurry mass which
travels toward the well is balanced by the wormhole
propagation and sand influx from the reservoir matrix.
38

) 13 ( ) )( t ( y A Q A
m s o sl
= & v
Here, A is cross-sectional area, and other terms are
defined in the Nomenclature. This equation states that
slurry production is equivalent to the combined mass of (i)
the displaced sand from the wormhole tip and (ii) the
volume of sand produced due to the wormhole propagation.
Solid influx into a wormhole from the formation is
postulated to occur if the pressure gradient exceeds the
critical strength, 2ccos/(1 sin), where c and are
cohesion and friction angle.
39
A coupled elastoplastic
model is assumed for the reservoir rock, combined with
steady-state flow defined by the pressure boundary
conditions at the cavity tip and the external drainage
boundary. The detailed solutions for deformation and
stresses are described in Appendix B. It is noted that the
elastoplastic model is used here only to determine the
material balance between the slurry mass inside the
formation and that flowing into the wormhole so that a
wormhole length can be calculated. Consequently, the
deformation effect on multiphase flow in the continuous
porous medium is not included in this investigation.
3. Slurry Transport Model
Steady-state slurry transport is assumed inside each
wormhole, and the solid supply is assumed to come from
the wormhole tip only. Solutions for the slurry transport
processes, the elastoplastic deformations in the reservoir,
and wormhole tip propagation and are presented in
Appendix A, B and C. Note that the input data required to
fully characterize wormhole propagation are massive and
may not be available to most users. An alternative is simply
to provide a reasonable guess for the rate of wormhole
propagation, which may be also calculated by a separate
package to simplify the simulation process.
Discussion and Field Implications
1. General Comments:
For the reservoirs considered, only non-economical oil rates
can be achieved with either vertical or horizontal non-
thermal wells where sand influx is kept at zero. Note,
however, that there are some heavy oil reservoirs in Alberta
and Saskatchewan (e.g. Cactus Lake) where non-thermal
horizontal wells have proven economic, but even in these
cases, careful analysis suggests that adjacent vertical CHOP
wells are more profitable.
Heating may provide a brief production enhancement,
but unless high-rate steam injection (fracturing) is used,
conductive heat penetration rate is very slow, and a
sustained high temperature is required for sustained
viscosity reduction. Will local heating combined with
CHOP in the same well have some positive result? The
heating effect will probably have minimum impact on sand
production and wormhole propagation because thermal
energy is extremely unlikely to reach the distant wormhole
tips as they grow outward, and thus can have little effect on
the yielding front, the locus of the wormhole tips (Figure 1).
There may be local thermally enhanced flow, but as the cold
sand and oil slurry moves toward the wellbore, it will
absorb the heat and be produced. A sustained wellbore
heating, however, may increase the effective wellbore
radius, allowing the high-pressure gradients to be displaced
into the formation. This issue of combined and different
technologies clearly needs more investigation.
Horizontal CHOP wells could be quite efficient because
the large drainage surface allows more wormholes and thus
an overall greater (albeit slower) heavy oil seepage rate to
develop than in a vertical well. However, horizontal well
success also depends on reservoir thickness, as the
wormhole penetration depth is severely constrained by the
upper and lower boundaries, and it is not likely that the
relative magnitude of the beneficial effect because of sand
influx observed in vertical wells could be approached.
A series of horizontal wells in the heavy oil Pelican
Lake and Amber Lake fields have been economically
successful despite production rate declines of 30-40%/yr.
These wells are characterized by the influx of small
amounts of sand, perhaps 0.1-0.5%, in contrast to typical
vertical CHOP wells where the sand cut is more typically 1-
6% during steady-state production. Note also that vertical
CHOP wells generally have decline rates far less than
horizontal wells because the continuous sanding generates a
growing high permeability zone fast enough to keep up to
economic production rate needs.
15,16
There are however
other difficulties associated with horizontal CHOP wells,

6 Y. Wang, C. Chen, M.B. Dusseault SPE 69714

related mainly to wellbore blockage by sand. The generally
low flow gradients may be insufficient to remobilize the
sand deposited inside the wellbore, and this would tend to
lead to blockage of much of the well (workovers are
expensive in horizontal wells). Despite some early claims
of economic success,
40
no current operator is now known to
aggressively encourage sand influx in horizontal wells, but
all operators using slotted liners accept small amounts of
sand influx, as attempts to block this influx seem to have
deleterious effects on oil rates.
2. Onset of Sanding in Vertical and Horizontal Wells:
In general, the critical flow rate before large-scale sand
production is considered to be proportional to the formation
cohesion after it has yielded because of elevated shear stress
in the region of the initiation point. In a poorly
consolidated reservoir, plastic yield may take place around a
wellbore during drilling and well completion (completion is
generally by perforating). Consequently, the stresses and
critical flow rate in the near-wellbore region are dominated
by the post-yield shear stress (which, because of the plastic
yield and cohesion loss, deformation, is independent of the
far-field shear stress).
Thus, the difference in sanding risk between a horizontal
and vertical well may be controlled solely by the difference
in surface area. Because of a large surface area, the specific
fluid flux in the horizontal well is usually much smaller than
in the vertical well, despite a much higher production rate in
the horizontal well. Returning to the case for the Amber
Lake and Pelican Lake reservoirs, if these reservoirs were to
be developed with aggressive vertical well CHOP
implementation, sand flux per barrel would be perhaps an
order of magnitude larger because of the higher specific
flux. By the same token, the specific fluid flux (per m
2
)
would be an order of magnitude greater than that in
horizontal wells.
3. Sanding Onset in a Cased and Open Hole wells:
Sand production is assumed to occur on the formation wall
when no casing is installed, whereas it only occurs from the
cavity tip of the perforations when casing is used. It can be
shown that the total flow rate required to cause sanding at
the hemispherical wormhole tip is usually lower than that in
a cylindrical case, depending on the length of the open
section, the perforation density, and radius (wormhole
length). This is linked directly to specific flux, as
mentioned above: for the perforation-dominated flow
assumption, flux is dominated by the tips, whereas for the
open-hole case, a much larger surface area is available,
giving lower gradients by virtue of both the large area and
the 2-D convergent nature of the flow. For example the
critical flow rates for both vertical and horizontal well can
be estimated by:
Q
cc
= 16 k r
w
c cos /(1-sin)/ Cased Hole
Q
cs
= 4 k H c cos /(1-sin)/ Open Hole
Where is friction angle, r
w
the perforation tunnel radius,
and H is open-hole section length. A comparison of the two
critical flow rates suggests that a perforated reservoir is
prone to sanding, as Q
cc
/Q
cs
= 4r
w
/H, which is typically less
than 1.
4. Cold Production in Vertical and Horizontal Wells:
Enhanced horizontal well production is potentially more
significant than in a vertical well because of the larger flow
contact area. However, this advantage can be compromised
by factors such as thickness, depth, and other pay zone
conditions. For example, rate enhancement during CHOP
has often been attributed to improved intrinsic permeability.
In our simulations, the production increase is mainly the
result of growing access to the reservoir because of
continued wormhole propagation. In a vertical CHOP well,
these may grow radially (2-D), but in the case of a
horizontal well in a thin reservoir, upward and downward
wormhole growth are severely limited, and the well length
implies that linear (1-D) lateral wormhole growth
dominates (Figure 2). Because of this, CHOP seems to be
less economical in horizontal wells when additional costs of
drilling, completion, and well workovers are considered.
In addition to the production rate enhancement,
wormhole propagation from vertical wells appears to also
increase oil recovery ratio, compared to a non-CHOP
horizontal well. In practice, in suitable reservoirs, vertical
CHOP wells seem able to attain 12-20% OOIP production,
whereas non-CHOP horizontal wells in the same zones may
achieve only 8-10% recovery because of the reduced access
(no wormholes) and the rapid production declines observed.
Validation of the Model and A Case Study
Sand production is viewed as the combined result of
mechanical skeleton destabilization (shearing, cohesion loss
and dilation) and viscous flow drag forces that carry the
loose sand into a slurry-filled cavity. Accordingly,
solutions have been developed based on stress equilibrium
and Darcy flow, incorporating effects of plasticity,
boundary conditions, changing properties, mass
conservation, continuity, and so on. Clearly, the bottom-
hole pressure (BHP) is identified as the direct parameter to
correlate with sand production.
Figure 3 shows a correlation between the predicted
wormhole radius, sand rate, and the enhanced production.
We attempt to establish a link between sand production and
reservoir enhanced performance so that we may achieve a
certain production by adjusting sand rate. It can be observed
that the enhanced oil rate does not necessarily correlate to
the peak sand rate, as we previously believed. Also
predictions using the proposed model are compared to the
sand cut and an average sand production performance in the
Frog Lake and Lindbergh reservoirs
7,8
in Figure 4-5. Data
SPE 69714 An Integrated Reservoir Model for Sand Production and Foamy-Oil Flow During Cold Heavy-Oil Production 7

for sand production from about 40 wells were collected, and
an average cumulative sand production curve used.
Sand production does not continue to increase, nor does
it remain constant: rather, it shows a temporary increase,
followed by a continuous decline, approaching a quasi-
stable slow sanding rate after about 100 days (Figure 4).
Figures 6 also shows the effects of reservoir production
enhancement. Finally an example of the inputs required for
the two-phase flow is presented in Tables 2 and 3.
Closure
A coupled numerical model designed to evaluate the
enhanced production and cumulative sand production in
CHOP wells was developed. The model incorporates a
slurry transport model with three-phase fluid flow. A
moving boundary representing the wormhole tip is defined
based on material balance between the continuous reservoir
and the wormholes.
The model seems to emulate real behavior reasonably
closely. It appears, in this model, that the continuously
improved access to the reservoir afforded by the wormholes
propagating is responsible for the enhanced oil rates, rather
than the intrinsic permeability change. It is also interesting
to note that the bottom-hole pressure implementation
strategy is deeply linked to several aspects of reservoir
response, which opens the possibility that optimization
using BHP to maximize oil recovery ratio is a real
possibility.
There are several debatable issues regarding the
simulation and understanding of the cold production
process. The characteristics of the wormhole network,
interaction among wormholes, and the bifurcation or
termination of wormholes all remain important issues to
more quantitatively predict the enhanced production during
cold production. A nonlinear diffusion equation for pore
pressure can be developed and coupled to the
geomechanical model so that both the effective stresses and
the radius of the sanding zone can be calculated. In
addition, the non-equilibrium kinetics of pressure increase
during the phase of gas supersaturation of the oil can be
incorporated in the proposed solution (this is available in
the reservoir simulator). The latter process can capture the
effects of both bubble dispersion and gas diffusion.
Our model is an attempt to capture the major
mechanisms in CHOP, based on physical arguments and
physical reasonableness, knowledge of field behavior, and
judgment as to which are first-order processes. Given the
fact that the proposed model seems to emulate field
production quite well, it may provide a reasonable
simulation tool for field prediction. Finally, it appears to be
quite easy to extend its capabilities to include other physical
effects because the use of specious correlations or non-
physical adjustments has been avoided.

APPENDIX A: Mixed Fluid and Solid (Slurry) Flow in a
Linear Cavity (Wormhole)
During production, especially from poorly consolidated
sandstone, sand particles can be continuously removed and
carried into the production well as a slurry. The slurry zone
may initiate near the wellbore and propagate into the
reservoir, but a classical geomechanics formulation is only
useful to describe plastic behavior beyond the fluidized
slurry zone. Within the eroding face and slurry transport
zone, a modified formulation proposed by Vardoulakis et
al.
42,50
may be used:
Continuity equation for solids:
t t
m
v
s

) 1 (
&
(A1)
Continuity equation for the fluidized slurry:
i i
s
), cq (
t
) c ( m
+

&
(A2)
Eroded and generated solid mass equation:
i i
s
q q c ) 1 (
m
=

&
(A3)
Carman-Kozeny Permeability law:
2
3
f s f
o
) 1 ( ) ( c
k

+
= (A4)
Continuity between the slurry zone and the plastic zone can
be used to solve the problem. Slurry zone development is
restricted to occur within the plastic zone and the boundary
lies in the location where a critical displacement or strain is
exceeded. An empirical relationship between cohesion and
porosity is also imposed. The parameter , which takes a
dimension of [1/L], is introduced to characterize the
erosion-related formation damage process.

APPENDIX B: Dilatant Volumetric Deformations (in
Reservoir)
In the plastic zone, the total strain tensor may be
decomposed into:
;
p
r
e
r r
+ =
p e

+ =
(B1)

Substitution of the above equation into the 2-D
compatibility equation leads to:
r dr
d
r dr
d
e
r
e
e
p
r
p
p

=

+

(B2)

To evaluate the plastic deformation, a simple non-associated
flow rule is used here; i.e. a plastic potential that is different
from the plastic yield function is used to characterize the
material plastic deformation:
0 N S Q
'
r d d
'
= =

(B3)

8 Y. Wang, C. Chen, M.B. Dusseault SPE 69714

Substituting the above into a general plastic flow rule, we
may define the dilation rate as:
p p
r
p p
r
d
) sin(


+
=
(B4)
Thus, the above equation may be rearranged into:
0 N
p
d
p
r
= +

(B5)

where
) sin( 1
) sin( 1
N
d
d
d

+
=
(B6)

The dilatant angle
d
can be quite different from the internal
friction angle . To calculate volumetric deformation
during plastic deformation, the internal friction angle has
often been used for the dilatant angle, the so-called
associated flow rule; however, this seriously overestimates
dilation, and such a discrepancy can be profoundly
important for poorly consolidated rocks. Thus, a non-
associated flow rule should be used.
Replacing
r
,

in the plastic yielding zone, then


substituting them into Eq. (B2), we have:
1 N
r
1 N
1 N
p
d d
d
cr dr ) t , r ( f r
r
1
+
+

+ =

(B7)
Defining terms as follows:
[ ]
E
S ) 1 (
;
E
) 1 N )( 1 (
; N ) 1 (
E
1
r
3
r
2 r 1
+
=
+
=
+
=
(B8)

Allows us, after some algebraic management, to express the
plastic strain as:
] ) N 1 ( [
2 d 1
'
r 1
p
+ =

d
d
d
N 1
d
3 N
r
'
r
N 1
r
' c
N 1
dr r
r
1
+ +
+
+

(B9)
At the elastoplastic interface, the tangential plastic strain
can be defined as zero if a perfectly plastic medium is
considered, but a non-zero term for the elastoplastic
material with a sudden strain-softening,
) t , R (
e

. In a
general case, the strain can be defined as:
] )
r
R
)( t , R ( ) t , r ( [
d
N 1 '
r
'
r 1
p +

=
dr r
r
1
] ) N 1 ( [
d
d
N
r
R
'
r
N 1
2 d 1
+
+

) t , R ( )
r
R
( ] )
r
R
( 1 [
N 1
p N 1 N 1
d
3 d d

+ +
+
+

+
(B10)
where

[ ]
p r
2
p
S S
E
1
) t , R (

=

(B11)
The elastic component in the plastic zone may be written as:
[ ] { }
r
'
r r
e
S ) 1 ( N ) 1 (
E
1

+
=

(B12)
Note also that the average fluid-solid flow rate may be
defined as:

dr
dp k
) v v ( n q
s f

= =
(B13)
APPENDIX C: Slurry Transport and Wormhole
Propagation (Wormhole Tip)
Wormhole and Plastic Yielding:
In this article, a wormhole is assumed to initiate from a
perforation. The wormhole will enlarge radially to its
maximum stable cross-sectional area, after which its
diameter is assumed stable. Because the effective radial
stress in the lateral wall of a wormhole is more
compressive with a smaller pressure gradient, a wormhole
is assumed to only propagate forward from the tip, where
pressure gradients are elevated by 3-D convergent flow and
interstitial gas bubble evolution. It is also assumed that
inside the plastic zone in the poorly consolidated reservoir,
only a small residual cohesion exists so that sanding can be
initiated anywhere in the plastic zone where conditions are
appropriate, but not outside of it. The sanding zone inside
the plastic zone can be determined by comparing the
effective radial stress to the residual (after yield) cohesion
arising from remnant cementation and capillary pressure.
Once the wormhole tip pressure falls below the critical
level for sanding, it is assumed that the pressure at the
wormhole tip drops to a level similar to that in the wellbore
(minimal p along the wormhole length), and as long as the
pressure at the tip is below the critical level for sanding, it
propagates further. It is noted that a plastic yielding zone
still exists around the lateral wall of the enlarged perforation
tunnel (wormhole), but the wall is stable simply because the
pore pressure gradient is much smaller there. This is
different to those hypotheses suggesting cavity growth
around wormholes.
5

Analyses demonstrate that stress concentrations based
on linear elastic theory show a maximum tangential stress
concentration (

]
max
) around a cavity, very close to its wall.
In a poorly consolidated formation, however, it is the post-
yield stress regime that dominates massive sand production,
and, because of dilation, cohesion loss and stress
redistribution, maximum stresses do not necessarily
concentrate around the cavity wall. Thus, we postulate that
the seepage force near the cavity wall will control sanding
initiation and dictate sanding rate, and this seems to be
supported by experimental observations.
43,44

We suggest that wormhole propagation will stop only if
the pressure at the tip rises (becomes closer to the reservoir
pore pressure) is high enough so that the pressure gradient
is below the critical level for sanding. Such a pressure rise
near the tip may occur as the result of frictional energy
losses along the wormhole wall (i.e. higher p to maintain
the same flux as the wormhole grows in length). By using a
similar formulation,
45
the free sand at the tip generated from
SPE 69714 An Integrated Reservoir Model for Sand Production and Foamy-Oil Flow During Cold Heavy-Oil Production 9

the reservoir may be calculated, and the cumulative amount
of sand production from wellbore may also be calculated by
considering the mass conservation at the wormhole tip, x =
y(t). Mass balance requires that the difference between the
mass inflow and outflow in a control volume with cross-
sectional area A inside a wormhole is equal to the enlarged
cavity:
) )( (
m s o s
t y A Q A v = & (C1)
where the volumetric oil flux Q
o
=
s
A v
s
. Alternatively,
this mass balance equation can be expressed as:
[ ]
s s m
v t y t y = ) ( ) 1 ( ) ( ) 1 ( & & (C2)
where v, are the velocity and porosity, respectively.
(Subscripts m and s denote the reservoir matrix and the
slurry.) Once free sand can be produced from the wormhole
tip, i.e., after initiation, the amount of sand that enters into
each wormhole is steady. This may explain why large sand
cuts may be observed at the early stage of CHOP; the ratio
between sand volume and fluid volume is large, but will
decline as proportionately more liquid (oil) is produced
from the lateral wormhole wall that is increasing in size,
whereas the proportion of free sand that may be produced
from the tip is fixed for a fixed number of wormholes.
Another important factor for sand cut decline is that the
pressure decline at the wormhole tip due to friction can drop
to a level where it can even result in a cessation of free sand
production entirely. A schematic diagram is presented in
Figure 1 to describe the sanding and production processes.
It is important to recognize that initial plastic yielding
near a wellbore or perforation is critical for sand
production, as only a small residual cohesion and extremely
low effective stresses thereafter exist to oppose sand flow.
No consideration herein is given for wormhole branching,
although this is a definite possibility in reality. (Note that
branching is not necessarily a bifurcation at the remotest tip;
a side branch could initiate anywhere along the wormhole if
the conditions existed and if appropriate pressure changes
took place.)
Fluid Pressure and Velocity along Wormhole:
In general, fluid pressure should decline away from the
wormhole tip toward the wellbore because of flow friction
along the wormhole wall, and also because a dense slurry
can dissipate energy internally.
46
The calculation of
wormhole propagation velocity can be an extremely
complicated process that depends on the geomechanical
characteristics of the reservoir and the erosive forces of the
slurry inside the wormhole. Considering solid mass
conservation in a wormhole with a cylindrical geometry:
( ) [ ]
( )
) , (
] 1 [
1 t x R
t
v
x
s b
s s b
+


(C3)
If the slurry transport inside the wormhole is simplified and
considered as a Darcys flow problem in 1-D:
) x ( R
t
] [
x
p k
x
s s
s
s
s
+

(C4)
where r
w
<x <y(t). R(x) is a source term which may reflect
the fluid inflow from the wormhole lateral wall, which, for a
steady state condition, has been defined as:
47,48,49

) p ) x ( p (
) r / r ln(
k r 2
) x ( R
d s
c d m
m c

= (C5)
Subscripts c and d denote cavity and drainage boundaries.
The slurry mobility, k
s
/
s
, is a function of slurry porosity,
conductivity, and viscosity, and average slurry density

s
(
s
) is linked to slurry porosity as:
s f s b s
+ = ) 1 ( (C6)
where
b
and
f
are the densities for skeleton solid matrix
and liquid, respectively. If the slurry is assumed
incompressible and the fluid influx does not change slurry
concentration, the slurry pressure along the wormhole may
be expressed as:


d d R
k
p x p
x
r s s s
s
w s
w

+ = ) (
) 1 (
) ( (C7a)
where
1
) (
] ) (
) 1 (
[ ) ) ( (

=

t y
r
x
s s s
s
w m
w
d d R
k x
p t p

(C7b)
where slurry porosity n
s
can be a function of x, changing
between the wellbore wall and wormhole tip. In reality, the
slurry porosity inside the wellbore is normally 94-99% (1-
6% sand cut by volume) and 51% near the wormhole tip.
p
w
, p
m
are the wellbore pressure and pressure at the tip,
respectively, and the first two of the following boundary
conditions have been used:
; p ) t , r ( p
w w s
= ) t ), t ( y ( p ) t ), t ( y ( p
m s
= (C8)
t
t t y u
x
t t y p k
m
s
s
s

) ), ( (
) 1 (
) ), ( (

(C9)
r
t t y p k
x
t t y p k
m
m
m s
s
s
s

) ), ( ( ) ), ( (

(C10)
where u is the displacement of the solid skeleton from the
reservoir formation, which can be determined in the next
section. Equation (C9) implies that the solid flux entering
the wormhole from the formation is equal to that in the
slurry at the wormhole tip. Equation (C10) can also be used
to determine wormhole tip pressure, p
m
(t), which is time-

10 Y. Wang, C. Chen, M.B. Dusseault SPE 69714

dependent. Note that the reservoir porosity does not appear
on the right-hand side of (C10), as the equation describes
the fluid flux continuity, in which the pressure in the
wormhole governs both the solid and fluid behavior.
Rearranging the mass conservation equation (C1), we have:
( )
s
m s
s
v t y

=
1
) ( & (C11)
A further postulate is made for the velocity of wormhole
propagation: a plastic zone can develop near the spherical
cavity tip (Figure 1). This cavity can be a perforation
channel or a newly developed wormhole tip. If the wellbore
pressure continues to decline, the wormhole will continue to
propagate. If the wellbore pressure is allowed to stabilize at
a certain level (usually 200-400 kPa in practice in vertical
CHOP wells), the wormhole will propagate continuously
until the wormhole tip pressure approaches the reservoir
drainage pressure because of pressure losses as slurry
travels along the channel. The latter is obviously our choice
for our proposed model, in which the fluid pressure is
described by equation (C7).
If slurry mobility and porosity are assumed to be
constant, the pressure along a wormhole can be
approximated as:
[ ] [ ]
[ ]
w
w m w
r ) t ( y sinh
r x sinh ) t ( p x ) t ( y sinh p
) t , x ( p

+
= (C12)
A similar solution for a fixed wormhole length is developed
as:
48

[ ]
y cosh
) x y ( cosh
p p p ) x ( p
w m m


= (C13)
where
) r / r ln(
k 2
r
1
c d m
m
c

=
in which the pressure gradient at the wormhole tip is
assumed to be zero and is the mobility of the slurry inside
a wormhole.
The slurry viscosity is approximated empirically by the
Thomas relation as:
{ } )] n 1 ( 6 . 16 exp[ 00273 . 0 n 05 . 10 n 6 . 22 55 . 13
s
2
s s f s
+ + =
(C14)
and the Kozeny-Carmen relation for the reservoir
permeability is also proposed (in Darcy units here):
2
3 2
3
) 1 (
10 629 . 5
m
m m
m
d
k

=

(C15)
where d
m
is the mean diameter of the sand grains. Note that
the above equation should not be used in the wormhole
when the porosity is approaching 1.



Nomenclature
a Cavity or wormhole tip radius.
c
p
, c
r

Peak and residual cohesion (M-C yield criteria)
c
e
, c
p

Elastic and plastic zone diffusivities
d
m
Mean grain diameter
E, Elastic Young's modulus and Poisson's ratio
G, Lam elastic constants
k
e
, k
m
Elastic and plastic zone permeabilities
K
s
, K
b

Bulk modulus for the solid particle and skeleton
m Specific mass discharge
M
p

Constant related to the nonlinear M-C criterion

e
,
m
,
s
Initial, yielding, and slurry porosities
N
p
, N
r
, N
d
Constants related to the peak, residual, and
dilatant angles of friction (M-C yield criteria)
p Pore pressure
p
m
(t) Wormhole tip pressure
p
s
Slurry pressure
p
o
,p
w
,p
d
Pressures for reservoir, wellbore, and drainage
q(t,p) Dynamic source and sink term for foamy oil
q Average fluid rate (flux)
Q Plastic potential
q
p

Compressive strength, nonlinear M-C criterion
r Radial distance variable
r
c
,r
w
,r
d

Cavity, wellbore, and drainage radii
S
p
,S
r

Peak and residual apparent strength
R, R(x) Plastic radius and wormhole source term
u Radial displacement
v
o
,v
g
,v
w
,v
s
Oil, gas, water and sand velocities
v
f
, v
m
, v
sl
Average fluid, reservoir solid, slurry velocities
x Distance variable inside the wormhole
y(t) Wormhole length as a function of time
,
p
, Poroelastic and poroelastoplastic parameters

s
Seepage force constant

1
,
3
Maximum and minimum principal stresses

r
,

Radial and tangential stresses

h
In-situ stress (considered isotropic in the plane)

p
,
r
,
d
Peak, residual, and dilatant angles
SPE 69714 An Integrated Reservoir Model for Sand Production and Foamy-Oil Flow During Cold Heavy-Oil Production 11

f
,
s
Fluid and solid skeletons density

sat
Reference fluid density

r
,

Total radial and tangential strains

e
r
,
e

Elastic radial and tangential strains

p
r
,
p

Plastic radial and tangential strains

m
,
s
Fluid and slurry viscosity
, ' Transformed constant
Empirical constant for fluid density
Acknowledgments
This project is partially funded by the National Research
Council Canada. Their financial support is deeply
appreciated. Special thanks go to Dr. D.E. Towson for his
support and technical guideline of the work.


Basic Reservoir Data
Reservoir Length 250 m
Reservoir Width 250 m
Reservoir Thickness 5 m
Initial Temperature 80 C
Initial Pressure 3.6 MPa
Oil Bubble Point Pressure 3.4 MPa
Initial Permeability 6000 mD
Initial Porosity 32 %
Rock Cohesion 0.10 psi
Rock Friction Angle 30 Degree
Initial Oil Viscosity 1600 cp
Initial Oil Saturation 0.8 fraction
Density of Oil 1.011 g/cm^3
Density of Water 0.997 g/cm^3
Density of Gas 0.001 g/cm^3
Wormhole Density at Wellbore 28 1/m^2
Wormhole Radius 2.3 cm
Enhanced Permeability 18000 mD
Slurry Porosity 50 %
Slurry Viscosity 5500 cp
Slurry Conductivity 1000000 mD
Bottomhole Pressure 0.5 MPa
Table 1: Input Data for Simulations






TABLE 2: OIL/GAS RELATIVE PERMEABILITY
S
L
K
rog
K
rg
P
cgo
(kPa)
-------------------------------------
0.2900 0.0000 0.1700 10.342
0.3950 0.0294 0.1120 7.281
0.4330 0.0461 0.1022 6.178
0.5150 0.0883 0.0855 3.785
0.5690 0.1172 0.0761 2.213
0.6140 0.1433 0.0654 0.903
0.6630 0.1764 0.0500 -0.524
0.7190 0.2170 0.0372 -2.158
0.7500 0.2255 0.0285 -3.061
0.8050 0.2919 0.0195 -4.661
0.8500 0.3373 0.0121 -5.971
0.8990 0.5169 0.0026 -7.398
1.000 1.000 0.0000 -10.342




Table 3: Oil/water Two Phase Case
SWT KRW KRO PCOW(MPa)
------------------------------------
0.200 0.0000 1.0000 0
0.250 0.0102 0.7690 0
0.294 0.0168 0.7241 0
0.357 0.0275 0.6206 0
0.414 0.0424 0.5040 0
0.490 0.0665 0.3714 0
0.557 0.0970 0.3029 0
0.630 0.1148 0.1555 0
0.673 0.1259 0.0956 0
0.719 0.1381 0.0576 0
0.789 0.1636 0.0000 0
1.000 0.2500 0.0000 0

12 Y. Wang, C. Chen, M.B. Dusseault SPE 69714

Figure 1: Sanding Mechanism and the
Conceptual Model Description
Plastic Yielding Oil/Sand
Slurry Zone
Active Sanding
Well
bore
Wellbore
Heating
Perforation
(Wormhole)
Figure 2: Possible Wormhole Network in
Vertical and
Horizontal Vertica
0
2
4
6
8
10
12
14
0 20 40 60 80 100 120 140 160 180 200 220 240
Production Time (day)
O
i
l

R
a
t
e

a
n
d

W
o
r
m
h
o
l
e

L
e
n
g
t
h
0.00
0.20
0.40
0.60
0.80
1.00
1.20
Qo-m3
Worm_L-m
Qsand-m3
Figure 3: Correlation between Predicted Sand
Rate, Wormhole Radius, and the Enhanced
Sand Cut
0
5
10
15
20
25
30
35
40
0 20 40 60 80 100 120 140 160 180 200 220 240
Production Time (day)
S
a
n
d

C
u
t
(
%
)
Model Prediction
Field Data
Figure 4: History Match by the Proposed
Model to the Field Data from Frog
Lake, Alberta
7,8

Sand Cumulative Production
0
10
20
30
40
50
60
70
80
0 20 40 60 80 100 120 140 160
Production Time (day)
Q
c
s
a
n
d
(
m
3
)
Model Prediction
Field Data
Figure 5: Sand Rate Prediction and a Field
Validation
7,8

Oil Production Rate (with Sand Production)
0
3
6
9
12
15
0 30 60 90 120 150 180 210 240
Production Time (day)
Q
o

(
m
3
/
d
)
Model Prediction
Field Data
Figure 6: Prediction and Validation of the
Enhanced Oil Rate
7,8

SPE 69714 An Integrated Reservoir Model for Sand Production and Foamy-Oil Flow During Cold Heavy-Oil Production 13

REFERENCES

1
Loughead, D.J., and M. Saltuklaroglu, Lloydminster
heavy oil production: Why so unusual? 9
th
Annual Heavy
Oil and Oil Sands Technol. Symp., Calgary, AB, Mar 1992.
2
Yeung, K.C. and M.F. Adamson, Burnt Lake Project -
Bitumen production from the Cold Lake oil sands deposit
without steam. AOSTRA/CHOA Conf., Fueling the
Future, Calgary, AB, Oct 1992.
3
Yeung, K.C., Cold flow production of crude bitumen at
the Burnt Lake Project, Northeastern Alberta, Canada.
Proc. 6
th
UNITAR Conf. on Heavy Crude and Tar Sands,
Houston, 1995.
4
Dusseault, M.B. and F.J. Santarelli, A conceptual model
for large scale solids production. Proc. Int. Symp.: Rock
Mech. At Great Depth, Vol. 2, Balkema, Rotterdam, 789-
797, 1989.
5
Smith, G.E., Fluid flow and sand production in heavy-oil
reservoirs under solution-gas drive. SPE J. of Production
Engineering, May 1988, 169-180.
6
Geilikman, M.B., M.B. Dusseault, and F.A. Dullien,
Sand production and yield propagation around wellbores.
Proc. CIM 47
th
Ann. Tech. Mtg., Paper
#
94-89, Calgary,
AB, 1994.
7
Metwally, M. and S. Solanki, Heavy oil reservoir
mechanism, Lindbergh and Frog Lake fields, Alberta, Part
I: Field Observations and Reservoir Simulation. Proc. CIM
47
th
Ann. Tech. Mtg., Paper 95-63, Banff, AB, 1995.
8
Solanki, S. and M. Metwally, Heavy oil reservoir
mechanism, Lindbergh and Frog Lake fields, Alberta, Part II:
Geomechanical Evaluation. Int. Heavy Oil Symp., Calgary,
AB, June 19-21, 1995, pp. 87-102, SPE
#
30249.
9
Suman, G.O., R.C. Ellis and R.E. Snyder. Sand Control
Handbook. Gulf Publishing Co., Houston, 89 p, 1985.
10
Kooijman, A.P., P.J. van den Hoek, Ph. de Bree, C.J.
Kenter, Z. Zhang, and M. Khodaverdian, Horizontal
wellbore stability and sand production in weakly
consolidated sandstones. Proc. SPE Ann. Tech. Conf.,
Denver, Colorado, 1996, SPE
#
36419.
11
Weingarten, J.S. and T.K. Perkins, Prediction of sand
production in gas wells: methods and Gulf of Mexico case
studies. SPE
#
24797, 67
th
Annual Tech. Conf., October 4-
7, 1992
12
Veeken, C.A.M., D.R. Davies, C.J. Kenter and A.P.
Kooijman, Sand production prediction review: developing
an integrated approach. SPE
#
22792, 66
th
Annual Tech.
Conf. SPE, Dallas, Texas, 1991
13
Tronvoll, J. and P.M. Halleck, Observations of sand
production and perforation cleanup in a weak sandstone.
Proc. EUROCK 94, Balkema, Rotterdam 1994, 355-360.

14
Wang, Y. and M.B. Dusseault, Sand production
potential near inclined, perforated wellbores. Paper 96-27,
Proc. CIM 47
th
Ann. Tech. Mtg., Calgary, AB, June, 1996.
15
Dusseault, M.B. and S. El-Sayed, CHOP Cold Heavy
Oil Production. Proc. 10
th
European IOR Symposium,
EAGE, Brighton, UK, Aug 1999.
16
Dusseault, M.B. and S. El-Sayed, Heavy oil well
production enhancement by encouraging sand influx.
SPE/DOE IOR Symposium, Tulsa, OK, April 2000, SPE
#
59276.
17
Dusseault, M.B., J. Tronvoll, F. Sanfilippo and F.J.
Santarelli, Skin self-cleaning in high-rate oil wells using
Sand Management. Proc. SPE Int. Conf. on Formation
Damage, Lafayette, LA, Feb 2000, SPE
#
58786.
18
Geilikman, M.B., M.B. Dusseault and F.A.L. Dullien,
Fluid rate in flowing granular medium with moving
boundary. Proc. 4
th
European Conf. on the Mathematics of
Oil Recovery, Rros, Norway, June, 1995, 41-50.
19
Geilikman, M.B., M.B. Dusseault and F.A.L Dullien,
Sand production as viscoplastic granular flow. Proc. SPE
Int. Symp. on Formation Damage Control, Lafayette LA,
Feb. 1995, SPE
#
27343.
20
Dusseault M.B., M.B. Geilikman and T.J.T. Spanos,
Mechanisms of massive sand production in heavy oils.
Proc 7
th
Int. Unitar Conf. on Heavy Oils and Tar Sands, 14
p., Beijing, PRC, Oct 1998.
21
Maini, B.B., H.K. Sarma and A.E. George, Significance
of foamy-oil behaviour in primary production of heavy
oils. Jour. Can. Pet. Tech., Nov. 1993, p. 50.
22
Shen, C. and J. Batycky, Some observations of mobility
enhancement of heavy oils flowing through sand pack under
solution gas drive. Proc CIM 47
th
Ann. Tech. Mtg.,
Calgary, AB, Paper 96-27, June 1996.
23
Chalaturnyk, R.J., B.T. Wagg and M.B. Dusseault. The
mechanism of solids production in unconsolidated heavy-oil
reservoirs. Proc. SPE Int. Symp. on Formation Damage
Control, Lafayette, LA, Feb 1992, 137-150, SPE
#
23780.
24
Wang, Y. and M.B Dusseault, Borehole yield and
hydraulic fracture initiation in poorly consolidated strata -
Part I: Impermeable media. Int. J. Rock Mech. Min. Sci.
and Geomech. Abst., v. 28(4), 1991, 235-246.
25
Wang, Y. and M.B. Dusseault, Borehole yield and
hydraulic fracture initiation in poorly consolidated strata
Part II: Permeable media. Int. J. Rock Mech., Min. Sci.
and Geomech. Abs., v. 28(4), 1991, 247-260.
26
Risnes, R., R.K. Bratli, and P. Horsrud, Sand stresses
around a wellbore. SPEJ, Trans. AIME, Dec 1982, 883-
898.
27
Tremblay, B., G. Sedgewick and K. Forshner, Modelling
of sand production from wells on primary recovery. Proc.

14 Y. Wang, C. Chen, M.B. Dusseault SPE 69714


CIM 47
th
Ann. Tech. Mtg., Calgary, AB, Paper 96-26, June
1996.
28
Geilikman, M.B. and M.B. Dusseault, Sand
Production Caused by Foamy Oil Flow. Transport in
Porous Media, 35: 259-272 1999.
29
Geilikman, M.B., F.A.L. Dullien and M.B. Dusseault,
Erosional creep of fluid-saturated granular medium. J. of
Eng. Mechanics of ASCE, v 123 (7), 653-659, 1997.
30
Geilikman, M.B. and M.B. Dusseault, Fluid-rate
enhancement from massive sand production in heavy oil
reservoirs. J. of Petr. Science & Engineering, 17, 5-18.
Special Issue: Near Wellbore Formation Damage and
Remediation, 1997.
31
Guo, F., R.C.K. Wong, J.S. Weaver, and W.E. Barr,
Heavy oil flow under solution gas drive: Non-
thermodynamic equilibrium. Proc. CIM 48
th
Ann. Tech.
Mtg., Paper 97-127, Calgary, AB, 1997.
32
Lewis, R.W. and Y. Sukirman, Finite element modeling
of three-phase flow in deforming saturated oil reservoirs,
Int. J. Num. Anal. Meth. Geomech., Vol. 17, 1993, pp. 577-
598.
32
Lewis, R.W., P.J. Roberts, and B.A. Schrefler, Finite
Element modeling of two-phase heat and fluid flow in
deforming porous media, Transport in Porous Media, Vol.
4, 1989, pp. 319-334.
33
Lewis, R.W., P.J. Roberts, and B.A. Schrefler, Finite
Element modeling of two-phase heat and fluid flow in
deforming porous media, Transport in Porous Media, Vol.
4, 1989, pp. 319-334.

34
Li, X., O.C. Zienkiewicz, and Y. M. Xie, A numerical
model for immiscible two-phase fluid flow in a porous
medium and its time domain solution, Int. J. Num. Anal.
Meth. Geomech., Vol. 30, 1990, pp. 1195-1212
35
Biot, M.A., General theory of three-dimensional
consolidation, J. Appl. Phys. 12, 155-164 , 1941
36
Rice, J.R. and Cleary, M.P., Some basic stress-diffusion
solutions for fluid saturated elastic porous media with
compressible constituents, Rev. Geophys. Space., 14, 227-
241,1976
37
Wang, Y. A coupled geomechanics-reservoir model and
applications to wellbore stability and sand prediction, SPE
69718, International Thermal Operations and Heavy Oil
Symp. Held in Margarita, Venezuela, 2001
38
Tortike, S. W.: Numerical Simulation of Thermal,
Multiphase Fluid Flow in an Elastoplastic Deforming Oil
Reservoir, Ph.D. Thesis, University of Alberta, 1991.



38
Y. Wang and J.-Y. Yuan, Cold production and
wormhole propagation in poorly consolidated reservoirs,
7
th
UNITAR International Conference, MS No. 41, Beijing,
Oct., 1998
39
Bratli, R.K. and R. Risnes, Stability and failure of sand
arches, SPEJ, Trans. AIME, April 1981, 236-248
40
Huang, W.S., B.E. Marcum, M.R. Chase and C.L. Yu,
Cold production of heavy oil from horizontal wells in the
Frog Lake Field. Proc. SPE Conf. International Thermal
Operations and Heavy Oil Sym., Bakersfield, CA, Feb
1997, SPE
#
37545.
42
Vardoulakis, I., M. Stavropoulou, and P. Papanastasiou,
Hydro-mechanical aspects of the sand production
problem, Transport in porous media, 22, pp.225-244,
1996
43
Tremblay, B., G. Sedgwick and K. Forshner, Imaging of
sand production in horizontal pack by X-ray computed
tomography. Proc. Int. Heavy Oil Symp., Calgary, AB,
June 1995, 71-86. SPE
#
30248.
44
Tremblay, B., G. Sedgwick and K. Forshner, Imaging of
sand production in a horizontal pack by X-ray computed
tomography. SPE Formation Evaluation, 1996, p. 94.
45
Wang, Y., Sand production and foamy oil flow in heavy-
oil reservoirs. Proc. Int. Thermal Operations & Heavy Oil
Symp., Bakersfield, CA, Feb., 1997, SPE
#
37553.
46
Frankel, N.A. and A. Acrivos, On the Viscosity of a
Concentrated Suspension of Solid Spheres. Chem. Eng. 22:
847-853, 1967.
47
Yuan, J.-Y., A. Babchin and B. Tremblay, Sand
transport through a partially filled wormhole. AACI Report
#
9798-13, 1998.
48
Yuan, J.-Y., B. Tremblay and A. Babchin, A wormhole
network model of cold production. ADOE/ARC Core
Industry Research Program, Report #9697-17, 1997.
49
Yuan, J.-Y., A. Babchin and B. Tremblay, Modelling the
wormhole flow in cold production. ADOE/ARC Core
Industry Research Program, Report #9697-5. 1996.
50
Papamichos, E., and M. Stavropoulou, An erosion-
mechanical model for sand production rate prediction, Int.
J. Rock Mech. Min Sci. & Geomech. Abstr. Vol 35(5), 531-
532,1998

You might also like