Sei sulla pagina 1di 7

Foamy-Oi l Fl ow Foamy-Oi l Fl ow

Brij B. Maini, SPE, U. of Calgary


Di sti ngui shed Author Seri es
54
Di sti ngui shed Author Seri es
OCTOBER 2001
Summary
Foamy-oil flow is a non-Darcy form of two-phase flow of
gas and oil encountered in many Canadian and Venezue-
lan heavy-oil reservoirs during production under solution-
gas drive. Unlike normal two-phase flow, which requires a
fluid phase to become continuous before it can flow, it
involves flow of dispersed gas bubbles. This paper is aimed
at acquainting the readers with this type of flow and its role
in heavy-oil production.
The paper starts with a discussion of what the term
foamy-oil flow means and how it evolved. Then a brief
review of the Canadian field practices is presented. This is
followed by a discussion of the pore-scale mechanisms
involved and the interplay between capillary and viscous
forces. A discussion of the strengths and weaknesses of var-
ious mathematical models proposed for numerical simula-
tion of this type of flow is also included. The paper ends
with a brief discussion of issues that remain unresolved.
Introduction
The term foamy oil originated from observations of sta-
ble foams in samples collected at the wellhead from many
Canadian and Venezuelan heavy-oil wells that produce
under solution-gas drive. The oil produced from these
wells was in the form of thick oil-continuous foam. It was
noticed that, very often, a sample container that was over-
flowing with oil at the time of collection at the wellhead
was nearly empty; less than 20% of its volume was filled
with oil when opened in the laboratory a few days later, by
which time the foam had collapsed. Many of these reser-
voirs also exhibit anomalous production behavior, both in
terms of higher-than-expected well productivity and
remarkably high primary recovery factors,
1
and this obser-
vation was not the result of metering errors caused by
foam. Over the years, this anomalous production behavior
became closely associated with the foamy nature of the
produced oil, and it was suggested that perhaps the two-
phase flow behavior of this type of oil in a reservoir rock is
different from that of a normal oil/gas mixture. The term
foamy-oil flow was coined to distinguish the two-phase
oil/gas flow in a porous medium of such heavy oils from
the normal two-phase behavior. This overview attempts to
acquaint the reader with recent developments related to
this topic, then provides some insight into the mecha-
nisms involved.
Smith
1
appears to be the first to publish a detailed analy-
sis of such unusual production behavior. He attributed it to
the flow characteristics of heavy oil containing a large vol-
ume fraction of very small gas bubbles. He suggested that
the mobility of a dispersion of very small bubbles in the
heavy oil could be several-fold higher than the single-
phase oil mobility. Maini et al.
2
attempted to verify this
assertion of high dispersion mobility in the laboratory but
found that the presence of freshly nucleated gas bubbles
actually decreased the oil mobility. However, they found
that the dispersed flow of gas was indeed possible under
solution-gas-drive conditions. Since then, the flow behav-
ior of such gas-in-oil dispersions has become a subject of
several investigations
315
and considerable speculation,
but it remains controversial and poorly understood.
Nonetheless, it is well accepted that solution-gas drive in
foamy-oil reservoirs involves some unusual effects. It
should be mentioned that such dispersed flow of gas under
solution-gas-drive conditions in the laboratory was noted
by Handy
16
in 1958 in high decline rate experiments.
However, he concluded that, at the low pressure decline
rates in the field, the dispersed flow would not be a signif-
icant factor.
Although there is still debate on the suitability of the
phrase foamy-oil flow to describe the anomalous flow of
the oil/gas mixtures in cold production of heavy oil, the
expression has become a part of petroleum engineering ter-
minology. To some, the term foamy-oil flow suggests
two-phase flow in the form of oil-continuous foam, and
they find it to be a misnomer because the actual
microstructure may not resemble foam. To others, includ-
ing this author, it only denotes the flow of a gas-in-oil dis-
persion, which is what appears to be involved. However,
the full meaning of the term is still evolving, and for now
it can be treated as a catchall phrase for representing the
contribution of nonequilibrium processes in solution-gas
drive in heavy-oil systems.
There are two types of nonequilibrium processes
involved in solution-gas drive in heavy oils. There is the
nonequilibrium between solution gas and free gas that
leads to a possibility of significant supersaturation of dis-
solved gas in the oil phase. The ramifications are delayed
release of solution gas and an apparent bubblepoint that is
lower than the true thermodynamic bubblepoint. This
nonequilibrium process is affected by the kinetics of bub-
ble nucleation and gas diffusivity. Because bubble nucle-
ation is a stochastic process driven by supersaturation, the
degree of supersaturation required before nucleation
occurs depends on the time available for nucleation.
Therefore, this type of nonequilibrium is likely to be more
significant in laboratory experiments, which are run on a
much smaller time scale compared with the field case.
The other nonequilibrium is related to fluid distribution
in the rock. Traditionally, in two-phase-flow situations in a
reservoir, the ratio of viscous to capillary forces is low, and
the capillary forces govern the fluid distribution. There-
fore, it is permissible to assume that the fluids distribute
themselves in such a way that the surface free energy for
Copyright 2001 Society of Petroleum Engineers
This is paper SPE 68885. Distinguished Author Series articles are general, descriptive
representations that summarize the state of the art in an area of technology by describing
recent developments for readers who are not specialists in the topics discussed. Written by
individuals recognized as experts in the area, these articles provide key references to more
definitive work and present specific details only to illustrate the technology. Purpose: to
inform the general readership of recent advances in various areas of petroleum engineering.
56
OCTOBER 2001
fluid/fluid and solid/fluid interfaces is minimized. One
consequence of this assumption is that the fluid distribu-
tion is the same under static and flowing conditions, and
it is not affected by the local pressure gradient. Another
consequence is that the gas phase must become continu-
ous before it can start flowing, with isolated gas bubbles
remaining trapped by capillary forces. However, this
assumption may not be valid in solution-gas drive in
heavy-oil reservoirs. Because of the high oil viscosity and
high drawdown pressure used in cold production, the
local capillary number (k/
og
) can be high enough to
mobilize isolated bubbles. This leads to dispersed flow.
This type of nonequilibrium is affected by the surface ten-
sion of the oil, the absolute permeability, and the value
of the gradient of flow potential in the vicinity of the iso-
lated bubbles.
Both types of nonequilibrium processes play a role in
foamy solution-gas drive. As stated earlier, the nonequilib-
rium related to nucleation and growth of bubbles becomes
less significant when the time scale moves from a few
hours or a few days in the laboratory to years in the field.
However, the other nonequilibrium process is not directly
affected by the time scale, but depends mostly on the cap-
illary number. Hence, it can be equally important in the
laboratory and the field, provided similar rock/fluid prop-
erties and pressure gradients are involved.
It is believed that the second type of nonequilibrium
process is more important in causing the observed anom-
alous production behavior in the field. The nonequilibri-
um processes related to nucleation and growth of bubbles
play a role and need to be understood, but the key to
understanding foamy-oil flow is in understanding the two-
phase flow at a high capillary number with a very large dif-
ference in viscosity of the two phases.
Field Observations
This section attempts to distill the experience of heavy-oil
operators in Canada into a few important observations
concerning cold production. The most notable field ob-
servation is that some unconsolidated-sand heavy-oil
reservoirs, exploited with vertical wells under primary
depletion conditions, perform better when sand is allowed
to flow freely into the wells. Both the oil production rate
and the oil recovery factor are much higher when the sand
is produced into the wells and transported to sur-
face with the oil. Oil production rates have been
reported to be more than 10 times the flow rate
predicted by Darcys law. It is believed that sand
production increases the fluid mobility in the near-
well zone by increasing the permeability in the
affected zone. Also, conventional wisdom predicts
that the solution-gas-drive recovery factor in these
viscous oil reservoirs should be in the range of 1 to
3% of original oil in place (OOIP). The actual and
projected recovery factors are in the range of 5 to
15% OOIP. Typical reservoir characteristics asso-
ciated with successful applications of cold produc-
tion are listed in Table 1. It should be noted that it
might be possible to apply the cold-production
technology to reservoirs that do not fall within the
range of characteristics in Table 1. At present, the
threshold reservoir characteristics have not been
clearly established.
New cold-production wells go through a startup phase,
during which the sand cut and fluid rate change in a char-
acteristic manner. Immediately upon starting production,
the sand cut is very high. The reported volume fraction of
sand in this phase is 10 to 50%.
1719
As production from
the well continues, the fluid rate increases slowly while the
sand cut declines rapidly. After some time of uninterrupt-
ed production, the sand cut stabilizes at a lower value that
seems to be a function of viscosity of the crude oil. The
reported values of sand cut in this phase are in the range
of 0.1 to 5 vol%.
1719
The oil rate continues to increase for
2 to 5 years, then starts a slow decline as reservoir-deple-
tion effects begin to dominate inflow performance. The
total volume of sand produced from cold-production wells
over their productive life can range from 500 to 1000 m
3
for wells that typically produce 10 to 20 m
3
/d of oil. Pro-
gressing-cavity pumps are used to handle this volume of
sand production. Usually, the wells are operated at or near
atmospheric backpressure (i.e., the drawdown is kept near
the maximum). The oil/water/sand/gas mixture is pro-
duced as a foamy mass, which goes to a stock tank for
gravity segregation.
It is believed that the sand production increases the
fluid mobility in the near-well zone by increasing the
permeability in the affected zone.
20
Continuing sand
production generates a growing zone of enhanced per-
meability that could be considered a growing negative-
skin effect. The actual morphology of the affected sand
is not well understood; two plausible models are net-
works of wormholes and uniformly dilated sand. Worm-
holes have been observed in the laboratory by use of
computer-aided-tomography imaging,
21
and their exis-
tence in the field has been inferred from tracer tests. It is
likely that the continued production of sand more or
less eliminates permeability damage in the near-well
zone that could have occurred by fines migration or
asphaltenes precipitation.
Past field experience shows that similar reservoirs that
were produced earlier with sand control yielded much
lower recovery factors. Without sand production, the
reservoir pressure declined more rapidly, and the gas/oil
ratio (GOR) increased much more rapidly. With sand pro-
duction, the GOR increases very slowly throughout the
depletion process.
Characteristics Value
Reservoir rock Unconsolidated sand
Depth, ft 1,3002,600
Net pay, ft 1380
Reservoir pressure, psi 350850
Reservoir temperature, F 5075
Porosity, % 3034
Permeability, md 50010,000
Oil saturation, % 6787
Oil gravity, API 1116
Oil viscosity (in situ), cp 1000100,000
Solution gas/oil ratio, scf/bbl 2575
Primary drive mechanism Solution gas
TABLE 1RESERVOIR CHARACTERISTICS ASSOCIATED
WITH SUCCESSFUL APPLICATIONS OF COLD PRODUCTION
IN CANADA
58
OCTOBER 2001
Another important observation is that these cold-pro-
duction wells can be damaged by extended shut-in peri-
ods. Very often, when a shut-in well is put back on pro-
duction, the inflow rate is only a fraction of the value
before shut-in, and the GOR is much higher. The sand-cut
also is affected by extended shut-in periods, usually show-
ing a significant decline.
Pore-Scale Mechanisms
The enhancement of fluid mobility resulting from sand
production can explain the improved inflow performance
of cold-production wells, but it does not explain the two-
to three-fold improvement in the primary recovery factors.
To explain the improved recovery factors, it is necessary to
postulate a mechanism that traps a larger fraction of the
released solution gas in the reservoir. Assume that such a
mechanism exists and that it is called foamy-oil flow
without worrying about whether oil-continuous foam is
created in the reservoir. The trapping of a larger fraction of
the released gas in foamy-oil flow can occur by an increase
in the critical gas saturation. However, if the critical gas
saturation is defined as the minimum saturation at which
a continuous gas phase can exist in the porous medium
under capillarity-controlled conditions, the critical gas sat-
uration becomes a fixed property of the rock/fluid system,
independent of the flow conditions. Then, the increased
trapping of the released solution gas must be attributed to
some other mechanism.
A plausible mechanism is the absence of capillarity-con-
trolled fluid distribution caused by large viscous forces. It
can lead to a different form of two-phase flow in which at
least a part of the released gas flows in the form of dis-
persed bubbles. The actual structure of such gas-in-oil dis-
persions and the mathematical description of such flow
conditions still are not well established. Much of the earli-
er discussion of foamy-oil flow in the literature was based
on the concept of microbubbles (i.e., bubbles much small-
er than average pore-throat size and, therefore, free to
move with the oil during flow through the rock
11
). This
type of dispersion can be generated only by nucleation of
a very large number of bubbles (explosive nucleation) and
by the presence of a mechanism that prevents these bub-
bles from growing into much larger bubbles with decrease
in the reservoir pressure.
An alternative hypothesis concerning the nature of
foamy-oil flow is that it involves much larger bubbles
migrating with the oil, and that the dispersion is created by
breakup of bubbles during their migration with the oil.
The main difference between the conventional solution-
gas drive and the foamy solution-gas drive is that the pres-
sure gradient in the latter is strong enough to mobilize
growing gas clusters after they have grown to a certain size.
Table 2 compares the pore-level processes occurring in
conventional solution-gas drives with foamy solution-gas
drives. Both processes start by nucleation of gas bubbles
driven by supersaturation of dissolved gas in the oil. This
nucleation is believed to occur in the roughness of pore
walls. It is likely that a large number of bubbles are formed
in the rough cavities of pore walls, but only a few actually
grow out of these cavities and become detached from the
wall because of a capillary barrier to bubble growth.
22
Up to this point, the two processes are very similar. The
difference is what happens to the growing bubble after it
becomes larger than the pore size. In conventional solu-
tion-gas drive, the bubble remains trapped and continues
to grow without ever leaving the pore where it originated.
Eventually, it will grow large enough to occupy several
pore bodies and will become connected to other bubbles
that started in other pores. Finally a continuous gas phase
is created that can flow to the production well. In foamy
solution-gas drive, the bubble, after growing to a certain
size, starts migrating with the oil. Note that this situation
does not imply that the gas bubble and the oil would have
the same average interstitial velocity. The bubble size
depends on the relative magnitude of capillary trapping
forces and the viscous mobilizing forces. The migrating
bubble continues to grow, but is prone to breakup into
smaller bubbles. The breakup of migrating bubbles into
smaller bubbles maintains dispersed gas flow by counter-
acting the effects of bubble coalescence. Because the gas
remains dispersed in the oil, the produced GOR remains
low, and a higher recovery factor is obtained. It should be
stressed that the two scenarios described above are the two
ends of a continuous spectrum of flow behavior. The tran-
sition between the conventional flow and dispersed flow
does not occur abruptly at a certain critical capillary num-
ber. It is a gradual transition that occurs above a threshold
capillary number.
Experimental evidence in support of this model of
foamy-oil flow was presented by Maini.
8
It explains the
rate dependence of recovery factors noted in several labo-
ratory studies. It is also consistent with visual observations
Conventional Solution-Gas Drive Foamy Solution-Gas Drive
Pressure depletion creates supersaturation. Pressure depletion creates supersaturation.
Bubbles nucleate in rough cavities of pore walls. Bubbles nucleate in rough cavities of pore walls.
Some bubbles detach and start growing in pore bodies. Some bubbles detach and start growing in pore bodies.
Bubbles continue to grow in place without vacating the Bubbles start migrating with the oil after growing to a certain size.
pore in which they originated.
Different bubbles originating in different pores grow large Migrating bubbles keep dividing into smaller bubbles.
enough to contact each other.
Bubbles coalesce to form a continuous gas phase. Dispersed flow is achieved by breakup of large bubbles into
smaller bubbles.
Producing GOR increases rapidly once the gas starts to flow Producing GOR remains low.
as a continuous phase.
Reservoir energy is depleted at a low recovery factor. High recovery factors are obtained.
TABLE 2: COMPARISON OF PORE-LEVEL PROCESSES
60
OCTOBER 2001
in micromodels that show the bubble size to be larger than
the pore size. The conditions required for this type of flow
to occur can be summarized as follows.
Viscous forces acting on growing bubbles should
exceed the capillary trapping forces.
Gravitational forces should not be capable of inducing
rapid gravity segregation of the two phases.
Interfacial chemistry effects that hinder bubble coales-
cence also may be needed.
Whether these requirements are met depends on the
rock/fluid properties and the operating conditions. The
first requirement is fulfilled when the pressure gradient is
high enough to mobilize isolated gas ganglia. The second
requirement is related to the role of gravitational forces in
producing segregated flow. If the pore structure is such
that isolated gas ganglia can move under the influence of
gravity, then the flow is likely to become segregated. The
third point is not well established, but it is likely that the
interfacial chemistry plays a role.
In terms of reservoir characteristics, dispersed flow is
more likely to occur when the permeability is high, the oil
is viscous, and the interfacial tension between the oil and
the released solution gas is low. To generate the required
viscous force, a high drawdown pressure is required. The
viscous trapping force decreases when the pore-body/pore-
throat aspect ratio becomes low. Therefore, the foamy-oil-
type dispersed flow is more likely to occur in well-sorted
unconsolidated sands.
Field Implications
If the premise is accepted that foamy-oil flow is a form of
non-Darcy two-phase flow in which the viscous forces
have become comparable with or stronger than the capil-
lary forces, then the implications in terms of field opera-
tions are not difficult to determine. Promoting this type of
flow requires conditions that generate high pressure gradi-
ents in the reservoir. The required pressure-gradient mag-
nitude depends on the sand characteristics and the interfa-
cial tension between the oil and the released gas. When a
new well is put on production, the initial pressure gradient
would be sufficient in the vicinity of the well in all heavy-
oil reservoirs if a high drawdown pressure were used.
However, as the depletion propagates deeper into the
reservoir, the pressure gradient becomes smaller and may
no longer be sufficient to generate the dispersed flow. The
drainage area that can sustain dispersed flow would
depend on the reservoir properties.
When sand is produced with the oil and increases the
fluid mobility in the zone from which sand was removed,
the reach of high pressure gradients needed for foamy flow
becomes much longer. Because mobility becomes high in
the near-well zone, the zone of high pressure gradient
moves deeper into the reservoir. Thus, sand production is
helpful in sustaining the dispersed flow in reservoirs that
would otherwise exhibit this type of flow only during the
initial production period. It is apparent that even with
sand production, the drainage area that can be produced
with foamy flow is not boundless. The required pressure
gradient would be available only up to a certain radial dis-
tance away from the well, beyond which dispersed flow
would not be generated.
As mentioned above, sand production is not a necessary
condition for the existence of foamy-oil flow. If the pressure
gradient is sufficient to mobilize gas ganglia, dispersed flow
will occur with or without sand production. Sand produc-
tion is needed in Canadian wells primarily to make the pro-
duction rates economically viable. The benefit in terms of
longer maintenance of foamy flow is a bonus.
The foregoing also suggests that the foamy flow is less
likely to occur when the reservoirs are exploited with hor-
izontal wells at low drawdown pressures. In Canadian
reservoirs, the production rates obtained with horizontal
wells (without sand production) are similar to those
obtained with sand-producing vertical wells, even when
high drawdown pressures are used in horizontal wells. The
productivity of these horizontal wells is several times high-
er than the productivity of vertical wells that do not pro-
duce sand. The sand production improves the productivi-
ty of vertical wells and makes it comparable with that of
horizontal wells. Under these conditions, horizontal wells
offer no significant economic advantage over the sand-pro-
ducing vertical wells. The increased operating cost of sand
production is more than compensated by the lower capital
cost of the vertical wells. In reservoirs that do not produce
sand, horizontal wells are generally superior.
Numerical Simulation of Foamy-Oil Flow
Numerical simulation of primary depletion in foamy-oil
reservoirs is still based primarily on empirical adjustments
to the conventional solution-gas-drive models. The two-
phase flow of oil and gas mixtures is described by relative
permeability relationships, with some adjustment to the rel-
ative permeability curves and/or to other fluid properties to
account for the effects of foamy flow. The rock/fluid prop-
erties that have been adjusted in such simulations include
the critical gas saturation, oil/gas relative permeability, fluid
and/or rock compressibility, pressure-dependent oil viscos-
ity, absolute permeability, and the bubblepoint pressure.
23
Some of the more interesting attempts to model the foamy
solution-gas drive are based on assumed mechanistic mod-
els of how the gas comes out of solution and what happens
to the released gas. Such models can be divided into two
broad categories: equilibrium and kinetic models.
Equilibrium Models
The motivation for developing such models comes from
their ease of implementation by use of existing reservoir
simulators that assume complete local equilibrium between
different phases. Most simulators also assume that the
mobility of fluids is independent of the capillary number.
Consequently, such models of foamy flow are inherently
incapable of accounting for the thermodynamically unsta-
ble nature of foamy dispersions and, generally, cannot pre-
dict the effect of operating conditions on solution-gas-drive
performance. However, such models have been successful
in attracting considerable attention in the literature and
continue to be used in reservoir simulation studies. Some
of these are briefly described in the following paragraphs.
Pseudobubblepoint Model
Kraus et al.
24
described a pseudobubblepoint model for
primary depletion in foamy-oil reservoirs. In this model,
the pseudobubblepoint pressure is an adjustable parameter
in the fluid property description. All of the released solu-
tion gas remains entrained in the oil phase until the reser-
voir pressure drops to the pseudobubblepoint pressure.
62
OCTOBER 2001
Below this pseudobubblepoint pressure, only a fraction of
the released gas remains entrained; and the gas fraction
decreases linearly to zero with declining pressures. The
entrained gas is treated as a part of the oleic phase, but its
molar volume and compressibility are evaluated with
those of the free gas. Equilibrium ratios from the conven-
tional pressure/volume/temperature (PVT) data are modi-
fied according to the pseudobubblepoint. An example of
primary depletion in a volumetric reservoir shows that
when foamy-oil fluid properties are used in a reservoir
simulator, the predicted results could show three anom-
alous production characteristics observed for foamy-oil
reservoirs, namely high oil recovery, low producing GOR,
and natural pressure maintenance.
Modified Fractional-Flow Models
Such models attempt to match the production behavior by
modifying the fractional-flow curve or the gas/oil relative
permeability curves. Lebel
25
described a model that
assumes that all released solution gas remains entrained in
the oil phase up to a certain system-dependent limiting-
volume fraction. Consequently, as the gas saturation
increases from zero, the fractional flow of gas increases lin-
early with saturation until the limiting entrained gas satu-
ration is reached. Beyond this limiting volume fraction of
dispersed gas, further increase in gas saturation results in
free gas. The effective viscosity of the foamy oil was
assumed to decrease marginally as the volume fraction of
gas increases. The density of the foamy oil was taken as a
volume-weighted average of the densities of the oil and gas
components. The equilibrium gas-in-oil PVT relationship
was assumed to be applicable. This model is equivalent to
a manipulation of the gas relative permeability curve,
which, up to a certain adjustable gas saturation, becomes a
function of oil viscosity and oil relative permeability.
A similar relative permeability-based approach has been
advocated by Firoozabadi and Pooladi-Darvish.
2627
Their
main thesis is that the improved recovery results primari-
ly from reduced relative permeability of gas, which
decreases as the oil viscosity increases. Assuming that the
relative permeability concept can be applied to dispersed
gas flow, the decrease in gas relative permeability with
increasing oil viscosity is expected when it is the pressure
gradient in the oil that is causing the dispersed bubbles to
move. However, it is unlikely that a predetermined gas/oil
relative permeability curve can describe the field situation
in which the capillary number varies with time and posi-
tion in the reservoir.
Reduced Viscosity Models
Claridge and Prats,
28
in an attempt to explain the higher-
than-expected inflow rates, suggested that the asphaltenes
present in the crude oil adhere to the gas bubbles while the
latter are still very tiny. This coating of asphaltenes on the bub-
ble surfaces stabilizes the bubbles at a small size. The bubbles
continue to flow through the rock pores with the oil. The key
element that differentiates this model from the others dis-
cussed above lies in the net effect of asphaltenes adsorption
onto the bubble surfaces on the viscosity of the crude oil.
They suggest that the oil viscosity decreases dramatically
because of the removal of the dispersed asphaltenes. It is dif-
ficult to see why this transfer of asphaltenes to bubble surfaces
would have a large effect on the dispersion viscosity because
the asphaltenes adsorbed on the bubble surfaces are still a part
of the dispersed phase. Moreover, attempts to verify such
reduction in dispersion viscosity in laboratory tests generally
have shown opposite results.
Shen and Batycky
29
formulated the theory of lubrication
in an attempt to show that the apparent viscosity of oil/gas
mixture at a low gas fraction is decreased because of the
lubrication effect of gases coming out of solution. The gas-
lubrication effect requires existence of microbubbles
attached to the walls of pores, which has not been con-
firmed by direct experimental evidence.
Kinetic Models
It is readily apparent that a dispersion of gas in oil is not a
thermodynamically stable species. Given sufficient time
and a helpful environment, the dispersion will separate
into free gas and oil phases. Although the natural tenden-
cy of the dispersion is to move toward segregation of phas-
es, such segregation can be arrested by imposing flow con-
ditions that favor regeneration of the dispersed bubbles.
Therefore, the flow behavior of foamy oil can be expected
to be a function of time as well as of the imposed flow con-
ditions. Kinetic models attempt to capture the time-depen-
dent changes in the flow behavior of foamy oil.
Coombe and Maini
30
described a model that accounts
for the kinetics of physical changes occurring in the mor-
phology of the gas-in-oil dispersion. It defines three non-
volatile components in the oil phase: dead oil, dissolved
gas, and gas dispersed in the form of microbubbles. The
dissolved gas changes to dispersed gas by means of a rate
process that is driven by the existing local supersaturation.
The dispersed gas changes into free gas by a second rate
process. The model was implemented in the Computer
Modeling Groups Stars
31
simulator by use of existing
chemical-reaction routines. Both rate processes were mod-
eled as chemical reactions with specified stoichiometry
and reaction-rate constants. The rate constants must be
determined by history matching.
Although this model accounts for the time-dependent
changes in dispersion properties, it fails to consider the
influence of time and the position-dependent capillary
number. Therefore, it is not useful for predicting the influ-
ence of operating conditions.
A similar approach was used by Sheng et al.,
12
who mod-
eled the rate of release of solution gas by exponential decay
of the local supersaturation and assumed that the gas
evolved from solution remained initially dispersed in the
oil. The dispersed gas disengages from the oil to become
free gas at a rate that is proportional to the volume fraction
of the dispersed gas in the dispersion. Thus, the kinetics of
the process involved in transfer of the solution gas to the
free-gas phase was described by two sequential-rate
processes with associated rate constants. This model was
used successfully for history matching laboratory solution-
gas-drive experiments. However, it was found that the rate
constants depended on the depletion rate used in the tests.
Both of the kinetic models mentioned above account for
time-dependent changes in the dispersion characteristics
by use of simple rate processes. This method represents an
improvement over the equilibrium models. However, the
rate processes involved in foamy solution-gas drive appear
to be controlled by the rock/fluid properties and by the
capillary number. Therefore, the rate constants inferred by
OCTOBER 2001
63 63
history matching a known depletion scenario are not valid
for predicting the outcome of a new scenario involving dif-
ferent flow conditions.
Other noteworthy models have been presented by Joseph
et al.,
32
who treat the flowing gas-in-oil dispersion as a
pseudosingle-phase fluid, and by Arora and Kovscek,
33
who use a bubble-population balance framework.
A simulation model capable of predicting the perform-
ance under different operating conditions is unavailable.
However, there are two promising routes that can be taken
for developing such a model. The first is an extension of the
kinetic models to include the dependence of the rate con-
stants on flow conditions. The second route is to make the
relative permeability in equilibrium models functions of the
capillary number. Egermann and Vizika
34
have recently
reported on experimental verification of the differences in
relative permeability in the far field and the near-wellbore
region. A model based on capillary-number-dependent rel-
ative permeability would not account for the changes in
dispersion properties with time, but may be a reasonable
approximation for fully developed flow in the field.
Conclusions
The past decade has seen a remarkable increase in research
effort in this area, and considerable progress was achieved
in understanding the mechanisms involved. It is clear that
the pressure gradient, rather than the decline rate of aver-
age reservoir pressure, is the driving force for foamy-oil
flow. However, several fundamental questions remain
unanswered. It still is not clear whether the interfacial
properties of the oil play a large role and what interfacial
properties (other than surface tension) are important. The
role of rock properties, such as pore geometry and pore-
size distribution, also is not completely understood. The
rheological properties of such foamy dispersions also are
not well characterized.
In practical field terms, answers to the following several
basic questions are being sought.
What reservoir characteristics make foamy solution-
gas drive possible?
What operating conditions are needed to maintain
foamy solution-gas drive?
How do we optimize the well spacing in foamy-oil
reservoirs?
Does foamy-oil flow occur in steam stimulation of
gassy heavy-oil reservoirs?
What are the effects of foamy primary production on
subsequent secondary and tertiary recovery processes?
Several research groups are active in this area, and it is
likely that answers to these questions will be forthcoming
in the near future.
Nomenclature
k=permeability, m
2
=gradient of flow potential, Pa/m

og
=surface tension, N/m
References
11. Smith, G.E.: Fluid Flow and Sand Production in Heavy Oil
Reservoirs Under Solution Gas Drive, SPEPE (May 1988) 169.
12. Maini, B.B., Sarma, H.K., and George, A.E.: Significance of
Foamy-Oil Behavior in Primary Production of Heavy Oils,
J. Cdn. Pet. Tech., 32, No. 9, 50.
13. Metwally, M. and Solanki, S.: Heavy Oil Reservoir Mecha-
nism, Lindbergh and Frog Lake Fields, Alberta, Part I: Field
Observations and Reservoir Simulation, presented at the 46th
Annual Technical Meeting of CIM, Banff, Canada, May 1995.
14. Solanki, S. and Metwally, M.: Heavy Oil Reservoir Mechanism,
Lindbergh and Frog Lake Fields, Alberta, Part II: Geomechan-
ical Evaluation, paper SPE 30249 presented at the 1995 Inter-
national Heavy Oil Symposium, Calgary, 1921 June.
15. Huerta, M. et al.: Understanding Foamy Oil Mechanisms for
Heavy Oil Reservoirs During Primary Production, paper SPE
36749 presented at the 1996 SPE Annual Technical Confer-
ence and Exhibition, Denver, Colorado, 69 October.
16. Treinen, R.J. et al.: Hamaca: Solution Gas Drive Recovery in
a Heavy Oil Reservoir, Experimental Results, paper SPE
39031 presented at the Fifth Latin American and Caribbean
Petroleum Engineering Conference, Rio de Janeiro, 30
August3 September 1997.
17. Urgelli, D. et al.: Investigation of Foamy Oil Effect From
Laboratory Experiments, paper SPE 54083 presented at the
1999 SPE International Thermal Operations and Heavy Oil
Symposium, Bakersfield, California, 1719 March 1999.
18. Maini, B.B.: Foamy Oil Flow in Primary Production of
Heavy Oil Under Solution Gas Drive, paper SPE 56541 pre-
sented at the 1999 SPE Annual Technical Conference and
Exhibition, Houston, 36 October.
19. Albartamani, N.S., Farouq Ali, S.M., and Lepski, B.: Investi-
gation of Foamy Oil Phenomena in Heavy Oil Reservoirs,
paper SPE 54084 presented at the 1999 SPE International
Thermal Operations and Heavy Oil Symposium, Bakersfield,
California, 1719 March.
10. Tang, G.-Q. and Firoozabadi, A.: Gas- and Liquid-Phase Rela-
tive Permeabilities for Cold Production From Heavy Oil Reser-
voirs, paper SPE 56540 presented at the 1999 SPE Annual
Technical Conference and Exhibition, Houston, 36 October.
11. Sheng, J.J. et al.: Critical Review of Foamy Oil Flow, Trans-
port in Porous Media (1999)35, 157.
12. Sheng, J.J. et al.: Modeling Foamy Oil Flow in Porous
Media, Transport in Porous Media (1999)35, 227.
13. Du, C. and Yortsos, Y.C.: A Numerical Study of the Critical
Gas Saturation in a Porous Medium, Transport in Porous
Media (1999)35, 205.
14. Ehlig-Economides, C.A., Fernandez, B.G., and Gongora,
C.A.: Global Experience and Practices for Cold Production
of Moderate and Heavy Oil, paper SPE 58773 presented at
the 2000 SPE International Symposium on Formation Dam-
age Control, Lafayette, Louisiana, 2324 February.
15. Bora, R., Maini, B.B., and Chakma, A.: Flow Visualization
Studies of Solution Gas Drive Process in Heavy Oil Reservoirs
Using a Glass Micromodel, SPEREE, (June 2000)3, 224.
16. Handy, L.L.: A Laboratory Study of Oil Recovery by Solution
Gas Drive, SPERE (December 1958)310; Trans., AIME 213.
17. Elkins, L.F., Morton, R., and Blackwell, W.A.: Experimental
Fireflood in a Very Viscous Oil-Unconsolidated Sand Reser-
voir, S.E. Pauls Valley Field, Oklahoma, paper SPE 4086 pre-
sented at the 1972 SPE Annual Meeting, San Antonio, Texas,
811 October.
18. Huang, W.S. et al.: Cold Production of Heavy Oil from Hori-
zontal Wells in the Frog Lake Field, paper SPE 37545 present-
ed at the 1997 SPE International Thermal Operations and
Heavy Oil Symposium, Bakersfield, California, 1012 February.
19. McCaffrey, W.J. and Bowman, R.D.: Recent Success in Pri-
mary Bitumen Production, paper No. 6 presented at the 8th
Annual Calgary U. et al. Heavy Oil & Oil Sands Technical
Symposium, Calgary, 14 March 1991.
JPT JPT
64
OCTOBER 2001
20. Dusseault, M.B. and El-Sayed, S.: Heavy-Oil Production
Enhancement by Encouraging Sand Production, paper SPE
59276 presented at the 2000 SPE/DOE Improved Oil Recov-
ery Symposium, Tulsa, Oklahoma, 35 April.
21. Tremblay, B., Sedgwik, G., and Vu, D.: CT Imaging of Sand
Production in a Horizontal Sand Pack Using Live Oil, paper
CIM 98-78 presented at the 49th Annual Technical Meeting
of the Petroleum Society, Calgary, 810 June 1998.
22. Yortsos, Y.C. and Parlar, M.: Phase Change in Binary Systems
in Porous Media: Application to Solution Gas Drive, work
performed under Contract No. FG18-87BC14126, U.S. Dept.
of Energy, Washington, DC, (October 1989).
23. Loughead, D.J. and Saltuklaroglu, M.: Lloydminster Heavy
Oil Production: Why So Unusual? presented at the Ninth
Annual Heavy Oil and Oil Sands Symposium, Calgary, 11
March 1992.
24. Kraus, W.P., McCaffrey, W.J., and Byod, G.W.: Pseudo-Bubble
Point Model for Foamy Oils, paper CIM 93-45 presented at the
CIM 1993 Annual Technical Conference, Calgary, 912 May.
25. Lebel, J.P.: Performance Implications of Various Reservoir
Access Geometries, presented at the 11th Annual Heavy Oil
& Oil Sands Technical Symposium, Calgary, 2 March 1994.
26. Firoozabadi, A.: Mechanism of Solution Gas Drive in Heavy
Oil Reservoirs, J. Cdn. Pet. Tech., (March 2001), 40, No. 3.
27. Pooladi-Darvish, M and Firoozabadi, A.: Solution Gas Drive
in Heavy Oil Reservoirs, J. Cdn. Pet. Tech., (April 1999), 38,
No. 4.
28. Claridge, E.L. and Prats, M.: A Proposed Model and Mecha-
nism for Anomalous Foamy Heavy Oil Behavior, paper
SPE 29243 presented at the 1995 International Heavy Oil
Symposium, Calgary, 1921 June.
29. Shen, C. and Batycky, J.P.: Some Observations of Mobility
Enhancement of Heavy Oils Flowing through Sand Pack
under Solution Gas Drive, paper CIM 96-27 presented at the
47th Annual Technical Conference of the Petroleum Society
of CIM, Calgary, 1012 June 1996.
30. Coombe, D. and Maini, B.: Modeling Foamy Oil Flow, pre-
sented at the Workshop on Foamy Oil Flow held at the Petro-
leum Recovery Inst., Calgary, 27 April 1994.
31. Stars Users Manual, Computer Modeling Group, Calgary,
October 1995.
32. Joseph, D.D., Kamp, A.M., and Bai, R.: A New Modeling
Approach for Heavy-Oil Flow in Porous Media, paper
SPE 69720 presented at the 2001 SPE International Thermal
Operations and Heavy Oil Symposium, Porlamar, Margarita
Island, Venezuela, 1214 March.
33. Arora, P. and Kovscek, A.R.: Mechanistic Modeling of Solution
Gas Drive in Viscous Oils, paper SPE 69717 presented at the
2001 SPE International Thermal Operations and Heavy Oil Sym-
posium, Porlamar, Margarita Island, Venezuela, 1214 March.
34. Egermann, P. and Vizika, O.: Critical Gas Saturation and Rel-
ative Permeability During Depressurization in the Far Field
and the Near-Wellbore Region, paper SPE 63149 presented
at the 2000 SPE Annual Technical Conference and Exhibi-
tion, Dallas, 14 October.
Brij B. Maini, SPE, is Professor of Chemical and Petroleum
Engineering at the U. of Calgary. Previously, he was group
leader for heavy-oil recovery research at the Petroleum
Recovery Inst. Maini holds a BTech degree in chemical engi-
neering from the Indian Inst. of Technology and a PhD
degree in chemical engineering from the U. of Washington.

Potrebbero piacerti anche