Você está na página 1de 12

SPE 113394

Effect of Depressurization on Trapped Saturations and Fluid Flow Functions


A.N. Nyre, CIPR and IFT/University of Bergen; S.R. McDougall, Heriot-Watt University; A. Skauge, CIPR/University
of Bergen
Copyright 2008, Society of Petroleum Engineers

This paper was prepared for presentation at the 2008 SPE/DOE Improved Oil Recovery Symposium held in Tulsa, Oklahoma, U.S.A., 1923April2008.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.


Abstract
Production below bubblepoint will generate free gas, first as discontinuous gas up to the critical gas saturation and thereafter
free or mobile gas. The level of critical gas saturation is affected by pressure decline rate, interfacial tensions, pore structure,
etc. The gas relative permeability is strongly reduced when trapped gas is present. Recent experimental studies have proved
that gas relative permeability can be several orders of magnitude lower for an internal gas drive process (gas liberation during
depressurization) than for an external gas drive process (gas injection). The critical gas saturation may indirectly influence gas
breakthrough, gas cut, and also oil production.

Network modeling has been used to investigate physical relations to factors influencing the formation of critical gas saturation
and the corresponding flow functions. The rock matrix composition determines, together with irreducible water saturation,
diffusion paths, and therefore the degree of supersaturation in the medium. This mechanism combined with depletion rates,
describes how trapped and mobile gas saturation evolves and determines factors influencing critical gas saturation.

We observe that bubble generation is strongly dependent on depletion rate, which in turn affects the critical gas saturation. The
gas relative permeability is found to be two orders of magnitude lower than for gas injection even at relatively high gas
saturations. We discuss the importance of including the physics of depressurization and advice on correct implementation of
depressurization in reservoir simulations.

The results show that lower coordination number leads to higher critical gas saturation. The variation of critical gas saturation
with pore structure diminishes at higher depletion rate. The significance to field application of depletion is that near the
production well the pore structure has little influence on the critical gas saturation, while at low depletion rate in the reservoir
(far from the well) the pore structure may be an important factor for the critical gas saturation.

Introduction
Studies of production below bubblepoint have been conducted in a variety of ways, ranging from core depletion experiments
with internal drive gas drive or combined internal and external gas drive. As the mechanisms of gas bubble formation and
expansion of the gas phase are difficult to investigate in detail in a porous medium, the effect of pressure depletion below
bubblepoint has also been studied by network modeling. Depletion below bubblepoint involves many coupled mechanisms,
and therefore there is still a lot to gain by improving the process understanding. Some key elements are the complexity of gas
liberation and kinetic processes such as bubble nucleation, diffusion, supersaturation, and in addition to fluid properties, and
the effects of capillary/gravitational/viscous force balance.

Critical gas saturation (Sgc) is one of the parameters that have been extensively investigated. A number of different definitions
have been given for critical gas saturation in the context of solution gas drive, for example critical gas saturation is defined as
the value at which a sample spanning gas cluster first forms (Yortsos and Parlar, 1989), the saturation at the onset of gas
detection (Firoozabadi et al., 1992), the minimum saturation for continuous gas flow (Kamath and Boyer, 1993) or the one at
which the producing gas-oil ratio exceeds the solution gas-oil ratio (Sahni et al., 2001). However, perhaps the simplest
concept, for network modeling purposes of critical gas saturation, is based upon a percolation theory approach: the gas
saturation at which a gas spanning cluster forms throughout the network signals the appearance of gas flow and therefore the
2 SPE 113394
saturation at which this happens becomes the critical gas saturation (as described by Yortsos and Parlar, 1989). This definition
also coincides with the gas saturation where the gas relative permeability becomes non zero.

There are many factors influencing critical gas saturation. The properties of the fluids, such as viscosity, density, the solution
gas-oil ratio, and interfacial tensions, are all key factors. These properties will have an effect on the nucleation potential and
capillarity. Pressure decline is also expected to give higher phase viscosities and densities. The G/O capillary pressure will
increase, because gas-oil IFT will increase at lower pressure. There will be changes in fluid saturation, among other, because
of shrinkage of oil due to mass exchanges, gas expansion and immobile gas (Sgc). Secondly, wettability and spreading, which
is determined by rock and fluid properties together, have an effect on critical gas saturation. And finally, the rock properties,
ranging from core length to coordination number and pore-size distribution, are key factors in bubble nucleation and the
formation of critical gas saturation in depressurization experiments.

A variety of experimental studies show critical gas saturation ranging from 1% to 30% gas saturation. There is also a
considerable amount of reports in the literature subjected to depressurization on the core scale and also in micro models. Some
examples of investigation of critical gas saturation are mentioned here, such as a study on glass micro-models (Danesh et al.,
1987), core studies of limestones (Stewart et al., 1954), vugular carbonates (Abgrall et al., 1973 and Madaoui et al., 1975) and
both studies of sandstone cores and including a review of experimental studies (Skauge et al., 1999). Many parameters have
been identified as having influence on the critical gas saturation, but the main parameter has been the depletion rate. The
depletion rate affects the bubble nucleation.

Some recent modeling studies using pore network modeling have investigated depressurization mechanisms, e.g. McDougall
et al. (1999) developed scaling groups for estimating Sgc at different pressure decline rates. Nucleation in porous media is
regarded as a heterogeneous and time dependent process where bubble densities are determined by the combination of factors
such as the rock, physico-chemical and depletion rate. The bubble densities increase with decreasing diffusion coefficients or
increasing depletion rate. Network modeling has shown that reduced mobility for liberated gas is coupled to the properties of
the porous medium. Reduced mobility is expected in poorly connected networks with many small pores and a high bubble
density (high depletion rate), and in the case of a connected network with many large pores and at low bubble density (low
depletion rate). The higher the coordination number, the less is the difference between external and internal gas drive (Poulsen
et al., 2001). Bondino et al. (2002) studied the effect of viscous forces, capillarity and gravitation upon critical gas saturation,
and reported an increase in Sgc with increasing bubble density.

More detailed investigation of relations between diffusion paths, supersaturation and bubble nucleation is needed, and may
lead to better understanding of how Sgc is determined. The effect of these mechanisms upon critical gas saturation may often
be masked by macroscopic forces like viscous forces, capillarity and gravitation. Network modeling can be utilized in order to
isolate the pore scale effects from the macroscopic force balance.

An important parameter for 3-phase flow is the spreading coefficient, which is related to the surface free energies of the
system. In a water-wet medium, oil is non-spreading if the spreading coefficient,

( )
ow go gw o
S + = , (1)


is negative. In such cases, gas - water contact surfaces may exist as observed in the micromodel experiments of ren et al.,
1990 and Kalaydjian, 1992. In the case of a positive (or zero) coefficient, water spreads over the solid surface, and oil (the
intermediate wetting phase) spreads between gas and water. In terms of network modeling, the former situation means that oil
clusters stay trapped if surrounded by water or gas-filed pores. In the latter case, oil can always drain from the system (via
film-flow) if a continuous combined pathway of oil and gas-filled pores reaches a production port. The existence of film-flow
mechanisms under spreading oil conditions has been observed in the micromodel studies of Henderson et al., 1991. We have
selected to use a non-spreading system in this study. The spreading coefficient is set to -0.005, which is weakly non-
spreading.

The network model and boundary conditions
The porous medium is modeled using a system of interconnected capillary elements, which generally configure to some
known lattice topology. Although these network structures are somewhat idealized, the capillary radii are assigned randomly
from a realistic pore size distribution in an effort to partially reconstruct the actual porous medium under investigation. Many
network models attempt to distinguish between pores and throats, by building networks consisting of hollow spheres
connected by thin capillary tubes. In such models, all of the liquid volume is contained in the spherical pores with pressure
differences being maintained by the throats (Blunt and King, 1991). The approach taken in this work is somewhat more
straightforward. The porous medium is initially modeled using a three-dimensional cubic network of what will be referred to
as pore elements; unlike many previous studies, no distinction is made here between pores and throats. Less connected
SPE 113394 3
systems can be considered by randomly removing a fraction of the pore elements. This simple lattice has dimensions N
x
x N
y

x N
z
where N
x
, N
y
, N
z
are the number of nodes in the respective directions. For relative permeability calculations, the pressure
gradient is taken to be in the x direction.

Here, the drainage mechanisms associated with the development of the non-wetting gas phase during pressure depletion is
examined by means of a pore scale simulator originally developed by McDougall and Sorbie (1999) and McDougall and
Mackay (1998). In this approach, the simulation of the depressurization process was carried out using an internally-seeded
invasion percolation model whereby, after bubbles had instantaneously nucleated, the evolution of the non-wetting phase (gas)
into the surrounding oil-filled pores depended primarily upon their relative capillary entry criteria. In this work, a similar
approach has been taken and the underlying modeling formulation has been significantly expanded on the basis of new
theoretical and experimental observations. Issues such as progressive nucleation, oil shrinkage and viscous mobilization have
all been considered (Bondino et al., 2005).
If the end of the network was sealed, a gas cap would form at the end. Instead our simulator mimics an excerpt of a core. Gas
bubbles which are produced are re-entered at the bottom of the network to simulate migrating bubbles from further down the
core. This leads to more plausible gas saturation curves (see Bondino et al. 2007). Viscous pressure and depletion rate can be
defined independently of each other. If depletion rates are high we get an apparent instantaneous nucleation profile. For low
depletion rates we have progressive nucleation.

In this study we have used a uniform pore-size distribution. The networks have been assigned water-wet properties. The oil
viscosity has been set to the range of light oils (
o
=0.1 cp). The interfacial tensions given in Table 1 are set to give a weakly
non-spreading scenario, and the oil/gas ratio at bubblepoint pressure is Rs=9.5.


Table 1: Interfacial tensions (IFT) [N/m].
Gas-water IFT Gas-oil IFT Oil-water IFT
0.03 0.01 0.025


All simulations have been run on a 3-D network, but for illustrative purposes a 2-D (80X40) = (nx,ny) pore network model is
schematized in Figure 1 in the same way as it appears during a simulation. Viscous and gravitational force are acting in the
same direction, this means that in viscous dominated regime, the gravitational and viscous effects add to each other (See
Figure 1). The orientation of the z axis is outwards from the plane of the sheet. Periodic flow boundary conditions are assumed
in the y and z direction in order to simulate larger systems and eliminate surface effects.






















Figure 1: Spatial schematization of the network model



x (nx)
y (ny)
INLET OUTLET
Direction of gravity
4 SPE 113394
The model take a positional information approach, whereby the degree of local supersaturation is taken into account when
deciding if an embryonic bubble should form in a potential nucleation site.

A crevice of radius W is assumed to be activated when the local supersaturation exceeds the capillary threshold of the crevice,

( )
W
P KC
l
2

, (2)

where K is the gas solubility constant, C the local dissolved gas concentration, P
l
the local liquid pressure, is the gas/oil
interfacial tension ( assumed zero). When a fluid system becomes supersaturated in a porous medium, the first pores to host a
bubble will be those containing the largest crevices. The critical supersaturation for the onset of nucleation consequently
depends on the radius of the largest crevice, whilst the overall degree of supersaturation during the process a function of many
parameters: depletion rate, diffusivity, and the nucleation characteristics of the medium (i.e. the pore size distribution).
The critical gas saturation, Sgc, denotes the volume fraction of the gas phase at the onset of bulk gas flow during the
depressurization of a supersaturated liquid in a porous medium. In the absence of gradients due to viscous or gravity forces,
Sgc is controlled by nucleation, capillary forces, and the rate of decline of the supersaturation.


Discussion and Results
In this paper, we focus on the connectivity in the porous medium represented by the coordination number and variation in fluid
saturation represented by the initial water saturation, and how these parameters affect bubble nucleation and critical gas
saturation (Sgc), at different pressure depletion rates. Experimental results show that critical gas saturation can vary with pore
size distribution. Very heterogeneous rocks generally show higher Sgc. Differences in pore geometry and clay content and
structure also affect the formation of gas bubbles, and thereby critical gas saturation (Kortekaas and van Poelgeest, 1991). The
coordination number has an influence on critical gas saturation, and with network modeling tools we are able to fade the effect
of other mechanism which also contributes to Sgc. To get gas accumulation, a certain degree of supersaturation is needed.
Bubble nucleation is caused by a liquid being depressurized. When a bubble is formed, gas from the surrounding liquid will
diffuse toward that bubble. This reduces the degree of supersaturation, thus diminishing the probability of formation of new
bubbles close to the existing bubble. Diffusion is slow, and depletion can, especially in a poorly connected matrix, cause
supersaturation elsewhere (Kortekaas et al., 1991). During depressurization, the pressure drawdown will create a viscous force
gradient and change in phase viscosity across the near well region of the reservoir. With increasing depletion rate, the viscous
pressure will increase and, together with the connectivity of the porous medium, it will affect the formation of a continuous
gas phase.

Critical gas saturation and depletion rate
In Figure 2 we see that at constant viscous pressure and coordination number, critical gas saturation increases with
increasing depletion rate. This has also been observed experimentally by e.g. Kortekaas and Poelgest (1991), and with network
models by McDougall and Sorbie (1999). In Figure 3 we see an example of how depletion rates affect gas saturation. Higher
depletion rates lead to more supersaturation, as can been seen from the saturation curve in Figure 3. Hence, more bubbles are
nucleated. Nevertheless, most of the simulations tend to have end-point saturation in the same range. Which mean that for low
depletion rates we have larger bubbles growing by gas diffusion.
0.0001
0.001
0.01
0.1
1
10
0 0.05 0.1 0.15 0.2
Sgc
(
d
P
/
d
t
)
/
P
b

Figure 2: Depletion rate divided by bubblepoint pressure
((dp/dt)/Pb) versus critical gas saturation (Sgc). The
bubblepoint pressure is: Pb=754 psi.

0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0.18
0.2
0 200 400 600 800
P (psi)
S
g
dp/dt=5
dp/dt=40
dp/dt=70
dp/dt=300
dp/dt=1000
dp/dt=1500
dp/dt=2000

Figure 3: Gas saturation (Sg) versus pressure (P). Example of
gas development during depletion at different depletion rates
(dp/dt).
SPE 113394 5




Figure 4: Bubble nucleation (Nbub) versus gas saturation (Sg) (left column), and gas saturation (Sg) versus pressure (P) (right
column) at three different depletion rates and two different coordination numbers. Here, depletion rate and viscous pressure are
simultaneously increasing (from top to bottom in figure).

Bubble nucleation
Depletion leads to gas expansion, supersaturated regions, diffusion due to change in chemical potential, nucleation and
bubble formation. Bubble nucleation can only take place if a liquid is supersaturated. The degree of supersaturation is
therefore important to critical gas saturation. Also, depletion rates are important to bubble growth and bubble nucleation. The
number of bubbles and the rate at which the bubbles grow can vary due to factors mentioned here. First, the number of bubbles
nucleated is determined, to a great extent, by the depletion rate (see Figures 4 and 5, left column). If the pressure decline is
rapid, the bubbles will not migrate to other larger bubbles and coalescent. Instead, small bubbles are nucleated throughout the
network. The diffusion paths will determine the rate of growth of these bubbles. With low depletion rates, gas will migrate to
larger bubbles, and the number of bubbles nucleated will be smaller than for high rate depletions. This is the case for
simulations with both high and low viscous pressures (see Figures 4 and 5). We then get a small number of large bubbles
containing the amount of gas equivalent to the large number of small bubbles with high depletion rates, i.e. both high and low
depletion rates give approximately the same end-point saturation.

Bondino et al. (2002) reported an increase in Sgc with increasing bubble density. We report the same for simulations with
low viscous pressure gradient, but with increasing viscous force linked to depletion rate, we find a conditional dependence on
the connectivity of the network. Critical gas saturation will increase with increasing bubble density in poorly connected
10 psi/day
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0 200 400 600 800
P (psi)
S
g
Z=6, deltaP=30d5
Z=4, deltaP=30d5
100 psi/day
0
0.05
0.1
0.15
0.2
0.25
0.3
0 200 400 600 800
P (psi)
S
g
Z=6, deltaP=72d5
Z=4, deltaP=72d5
300 psi/day
0
0.05
0.1
0.15
0.2
0.25
0.3
0 200 400 600 800
P (psi)
S
g
Z=6, deltaP=10d6
Z=4, deltaP=10d6
0
200
400
600
800
1000
1200
1400
1600
0 0.1 0.2 0.3 0.4
Sg
N
b
u
b
10 psi/d, Z=6
10 psi/d, Z=4
Sgc (Z=6)
Sgc (Z=4)
0
500
1000
1500
2000
2500
3000
3500
0 0.1 0.2 0.3
Sg
N
b
u
b
100 psi/d, Z=6
100 psi/d, Z=4
Sgc (Z=6)
Sgc (Z=4)
0
500
1000
1500
2000
2500
3000
3500
0 0.1 0.2 0.3
Sg
N
b
u
b
300 psi/d, Z=6
300 psi/d, Z=4
Sgc (Z=6)
Sgc (Z=4)
6 SPE 113394
networks, but at high coordination number, the critical gas saturation will decrease with increasing bubble density (see Figure
4, left column).



Figure 5: Bubble nucleation (Nbub) versus gas saturation (Sg) (left column), and gas saturation (Sg) versus pressure (P) (right
column) at three different depletion rates and two different coordination numbers. Here, depletion rate is increasing (from top to
bottom in figure), and viscous pressure is kept constant.

The viscous pressure causes gas bubble movement. Low viscous pressures will therefore lead to a situation where large
bubbles coalescent with other large bubbles to form a spanning gas cluster, and hence reaching critical gas saturation. High
viscous pressures will move the small bubbles into each other, forming larger bubbles and eventually reaching critical gas
saturation. High viscous pressures will cause unsteady state relative permeabilities. Now, critical gas saturation is not reached
until there is steady state relative permeability.

10 psi/day
0
0.05
0.1
0.15
0.2
0.25
0 200 400 600 800
P (psi)
S
g
Z=6, deltaP=10d-3
Z=4, deltaP=10d-3
100 psi/day
0
0.05
0.1
0.15
0.2
0.25
0 200 400 600 800
P (psi)
S
g
Z=6, deltaP=10d-3
Z=4, deltaP=10d-3
300 psi/day
0
0.05
0.1
0.15
0.2
0.25
0 200 400 600 800
P (psi)
S
g
Z=6, deltaP=10d-3
Z=4, deltaP=10d-3
0
20
40
60
80
100
120
0 0.05 0.1 0.15 0.2 0.25
Sg
N
b
u
b
10 psi/d, Z=6
10 psi/d, Z=4
Sgc (Z=6)
Sgc (Z=4)
0
50
100
150
200
250
300
350
400
450
0 0.05 0.1 0.15 0.2 0.25
Sg
N
b
u
b
100 psi/d, Z=6
100 psi/d, Z=4
Sgc (Z=6)
Sgc (Z=4)
0
100
200
300
400
500
600
0 0.05 0.1 0.15 0.2 0.25
Sg
N
b
u
b
300 psi/d, Z=6
300 psi/d, Z=4
Sgc (Z=6)
Sgc (Z=4)
SPE 113394 7
0
50
100
150
200
250
300
350
200 300 400 500 600 700 800
P (psi)
N
b
u
bHigh viscous pressure
Low viscous pressure

Figure 6: Number of bubbles nucleated (cumulative) (Nbub) versus pressure (P), at two different viscous pressures.

The number of bubbles nucleated depends on both depletion rate and viscous pressure. In Figure 6 we have two different
viscous pressures at same depletion rate. The number of bubbles nucleated is larger for high viscous pressures, as also reported
by Bondino et al. (2005) and Poulsen et al. (2001). In Figure 5 we see that by keeping a constant (very low) viscous pressure,
the effect of depletion rate upon bubble nucleation can be isolated. The trend that more bubbles are nucleated with increasing
depletion rate is the same as was observed in simulations with high viscous pressures (Figure 4), but a considerably smaller
number of bubbles are nucleated in the absence of viscous pressure. In detail, depletion rates affect supersaturation and
therefore bubble nucleation, which, as mentioned above, will affect critical gas saturations and continuous gas saturation (time
dependent). The supersaturation consists of two components, one due to capillary forces and another due to dynamic effects.
In low permeability rocks the capillary component of supersaturation may become very significant, while in high permeability
rocks this component is relatively insignificant.

Effect of initial water saturation and coordination number
Kortekaas and van Poelgeest (1991) found systematically higher critical gas saturation for watered-out cores compared to
cores at connate water saturation. They explain this by the fact that more gas bubbles are formed for the watered-out case due
to more restricted diffusion of gas through water than through oil. In addition, they point out that the higher gas/water
interfacial tension (IFT) compared to gas/oil IFT increases the relative importance of capillary forces compared to buoyancy
(gravity) forces for the case with high water saturation. This might retard upward migration of gas and lead to higher critical
gas saturations.

The coordination number is the average number of accessible exits from a pore. The composition of the rock itself settles
the physical coordination number (Z). However, the effective coordination number for the oil phase is lower than the physical
coordination number because of water filled pore and throats due to the initial water saturation (Swi). The volume of this
water, as well as its distribution, will then change the overall connectivity, hence affecting the diffusion paths for the oil and
gas in the system.

In Figure 7 we have the results of simulations of three different water saturations (Swi= 0.5, 0.2 and 0.0). The degree of
supersaturation was found to be higher for networks with low initial water saturation, also illustrated by the slope in Figure 7.
Higher gas saturation was obtained with increase in connectivity in the porous medium, and the difference between high and
low connectivity networks is more obvious at high water saturation. The amount of water in the network is clearly limiting the
amount of accumulated gas in the network, i.e. at high water saturations there is less oil initially (see Figure 7).

Figure 7: Developed gas saturation versus pressure at three different initial water saturations both for coordination number 4 and 6.
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0.18
0 100 200 300 400 500 600 700 800
P(psi)
S
g
Swi=0.5, Z=6
Swi=0.5, Z=4
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0 200 400 600 800
P (psi)
S
g
Swi=0.2, Z=6
Swi=0.2, Z=4
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0 200 400 600 800
P (psi)
S
g
Swi=0.0, Z=6
Swi=0.0, Z=4
8 SPE 113394

Derivatives of saturation profiles
Looking at the derivatives of the saturation profile (Figure 8), it is possible to identify a turning point around critical gas
saturation for simulations without initial water saturation. The derivative of the gas saturation is increasing after a gas spanning
cluster is formed. Now, this phenomenon does not occur without the sufficient supersaturation. Even though we have reached
critical gas saturation, the gas saturation is increasing more rapidly. This can only be explained by many bubbles, not yet
connected to the spanning cluster, have nucleated due to the high degree of supersaturation. As these bubbles are connected to
the cluster, we get a decrease in the derivative of the gas saturation. In Figure 9 we see that the growth of the number of gas
bubbles is higher up to around critical gas saturation. Thus, many bubbles, not yet connected, are nucleated up to Sgc, and
once a spanning cluster is reached bubbles nucleate less frequently and the rate at which the gas saturation is growing is
increasing. Diffusion paths are important after Sgc is reached.


Figure 8:The derivative of gas saturation on pressure versus gas saturation at two different initial water saturations (Swi). The left
figure displays simulations with coordination number (Z) set to 4, and the figure on the right has Z=6. Critical gas saturations (Sgc)
are given as straight lines.


Figure 9: The figure to the left shows number of bubbles (Nbub) (cumulative) versus gas saturation (Sg) for two different coordination
numbers (Z) and two different initial water saturations (Swi).

Critical gas saturation and viscous pressure
As mentioned before, high depletion rates can cause large viscous gradients across the medium. In figures 10 and 11 we
have simulations for three different depletion rates (dp/dt), where the viscous pressure (deltaP) is constant in Figure 10 and
increasing with increasing depletion rate in Figure 11. In accordance with the observations of McDougall et al. (1999) and
Bondino et al. (2007), the number of bubbles nucleated is substantially larger for high depletion rates than for low depletion
rates (see Figure 4, left column). Critical gas saturation is increasing with increasing depletion rates for less connected
networks (Z=4). For highly connected networks the situation is opposite, i.e. critical gas saturation is reduced with increasing
depletion rate. This means that the estimation of critical gas saturation is depending on the connectivity, as well as depletion
rate.
Also from Figure 11, we find that the difference in critical gas saturation between high and low connected networks
decreases as depletion rates decrease. If viscous pressure is high, we get conditional relations between Sgc and depletion rate.
That is, pore geometry and permeability are important for high viscous pressures, as shown in Figure 10. At low coordination
0
50
100
150
200
250
300
350
400
450
0 0.1 0.2 0.3 0.4
Sg
N
b
u
b
Z=6, Swi=0.0
Z=4, Swi=0.0
Z=6, Swi=0.2
Z=4, Swi=0.2
Sgc (Z=6, Swi=0.0)
Sgc (Z=4, Swi=0.0)
Sgc (Z=6, Swi=0.2)
Sgc (Z=4, Swi=0.2)
0
50
100
150
200
250
300
350
400
450
0 100 200 300 400 500 600 700 800
P (psi)
N
b
u
b
Z=6, Swi=0.0
Z=4, Swi=0.0
Z=6, Swi=0.2
Z=4, Swi=0.2
Z=4
0
0.001
0.002
0.003
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Sg
d
S
g
/
d
P
Swi=0.2
Swi=0.0
Sgc (Swi=0.2)
Sgc (Swi=0.0)
Z=6
0
0.001
0.002
0.003
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Sg
d
S
g
/
d
P
Swi=0.2
Swi=0.0
Sgc (Swi=0.2)
Sgc (Swi=0.0)
SPE 113394 9
number we get a higher degree of super saturation in the medium. This is due to the limited diffusion paths at low
connectivity.

0
0.05
0.1
0.15
0.2
0.25
2 3 4 5 6
Z
S
g
c
dp/dt=10psi/day
dp/dt=100psi/day
dp/dt=300psi/day

Figure 10: Critical gas saturation (Sgc) versus coordination
number (Z) at three different depletion rates. Here, we have a
constant (and moderate) viscous pressure.


0
0.05
0.1
0.15
0.2
0.25
0.3
2 3 4 5 6
Z
S
g
c
dp/dt=10psi/day
dp/dt=100psi/day
dp/dt=300psi/day

Figure 11: Critical gas saturation (Sgc) versus coordination
number (Z) at three different depletion rates. Here, depletion
rate and viscous pressure are simultaneously increasing.
Effect of coordination number and depletion rate
As we can see from Figure 4, the depletion rate (with corresponding viscous pressure) affects supersaturation, and
coordination number is more important at higher depletion rates. Generally viscous pressure and depletion rate should be
linked. These to mechanisms are in practice hard to distinguish. It is of interest to isolate one effect from the other. This is
possible in network modeling. By decoupling the viscous pressure and depletion rate, leaving viscous pressure low and
constant, we get the results shown in Figure 5. The coordination number, together with depletion rate, will then control the
degree of supersaturation in the medium.
These results show the opposite trend of what was found in Figure 5. For increasing depletion rate the critical gas
saturation is increasing for both high and low connectivity networks. This study show that coordination number has larger
impact on low depletion rates (see Figure 5).
Critical gas saturation is often found to be lower when the pressure decline rate is reduced (Skauge et al. 1999). We report
similar behavior on simulations with moderate viscous pressure (see Figure 10). We also get higher critical gas saturation for
low connected networks, as expected. Nevertheless, the degree of supersaturation is larger for high depletion rates, as was also
found at high viscous pressures.



Figure 12: Relative permeability for gas (Krg) versus gas saturation (Sg) (left column). Oil and gas relative permeability versus oil
saturation (So) (right column). At the two figures on top, the coordination number (Z) is 6, and at the bottom Z=4.
Z=6
0
0.01
0.02
0.03
0.04
0.05
0 0.1 0.2 0.3
Sg
K
r
g
Z=4
0
0.02
0.04
0.06
0.08
0 0.1 0.2 0.3 0.4
Sg
K
r
g
0
0.2
0.4
0.6
0.8
1
0 0.5 1
So
K
r
i
Kro
Krg
0
0.2
0.4
0.6
0.8
1
0 0.5 1
So
K
r
i
Kro
Krg
10 SPE 113394
Relative permeability
Gas relative permeability results from external drive processes generally exceeds gas relative permeability resulting from
internal drive processes as mentioned before. The definition of critical gas saturation in this paper relates to steady-state
relative permeability. At the moment a spanning gas cluster is formed, the steady-state relative gas permeability has a non-zero
value. In Figure 12 we see that relative permeabilities are little affected at high coordination numbers. Results from Poulsen et
al., 2001, showed that with higher coordination number, the difference between internal and external drive gas relative
permeability becomes less. It was also suggested that the formation of isolated gas clusters away from the main spanning
cluster was not so pronounced when the connectivity of the network is high, as when the connectivity was low.

At high bubble nucleation, the difference between the internal and the external drive gas relative permeability seems to
increase. Pore size distribution is also one on the factors that affects the difference between internal and external drive gas
relative permeability. The uniform pore size distribution used in this study, was expected to increase the difference between
internal and external gas relative permeability, but the rather high gas relative permeability may be governed by the high
coordination numbers. As expected, critical gas saturation (the first non-zero value for relative permeability) is reached at
lower gas saturation for simulations with high coordination number.


Conclusions
The network study of depressurization has shown that higher gas saturation was obtained with increase in connectivity in the
porous medium, and the difference between high and low connectivity networks is larger at high water saturation.

High viscous pressure gradient: Critical gas saturation is increasing with increasing depletion rates for less connected networks
(Z=4). For highly connected networks the situation is the opposite, i.e. critical gas saturation is reduced with increasing
depletion rate.

Low viscous pressure gradient: These results show the opposite trend of what was found for simulations with high viscous
pressures. Critical gas saturation is increasing for both high and low connectivity networks with the increase of depletion rate.
Also, coordination number is more important for low depletion rates in this scenario.

The degree of supersaturation was found to be higher for networks with low initial water saturation

The number of bubbles nucleated depends both on the depletion rate and the viscous pressure gradient. With increasing
depletion rate a larger number of bubbles is nucleated. This is the same for simulations with and without viscous pressure
gradient. In the absence of viscous pressure gradient a smaller amount of bubbles is nucleated compared to results which
includes viscous pressure.


Nomenclature
C Local dissolved gas concentration
dp/dt Depletion rate [psi/day]
deltaP, P Viscous pressure [psi]
IFT Interfacial tension
K Gas solubility constant,
Krg Gas relative permeability [fraction]
Kro Oil relative permeability [fraction]

o
Oil viscosity [cp]
Nbub Number of bubbles nucleated
N
x
, N
y
, N
z
Number of nodes in the x, y and z directions
P
l
Local liquid pressure [psi]
Pb Bubblepoint pressure [psi]
P Pressure [psi]
Rs Oil/gas ratio at bubblepoint pressure
Sgc Critical gas saturation [fraction]

gw
Gas-water interfacial tension [N/m]

go
Gas-oil interfacial tension [N/m]

ow
Oil-water interfacial tension [N/m]
Sg Gas saturation [fraction]
So Spreading coefficient [N/m]
Swi Initial water saturation [fraction]
Z Coordination number
SPE 113394 11
References

Abgrall,E., and Iffly,R.: Physical Studies of Flow by Expansion of Dissolved Gas, Revue de LInstitut Francais du Petrole (1973) Vol 28,
no.5, 667692, 1973

Blunt, M. and King, P.: Relative Permeabilities from Two- and Three- Dimensional Pore-Scale Network Modelling, Transport in Porous
Media, vol. 6, pp. 407-433, 1991.

Bondino, I., McDougall, S.R. and Hamon, G: A Pore-Scale Network Modeling Study of Gravitational Effects During Solution Gas Drive:
Results from Macroscale Simulations, proceedings, presented at the SPE Europec/EAGE Annual Conference, London, UK, SPE
107556, 2007

Bondino, I., McDougall, S.R. and Hamon, G: Pore Network Modelling of Heavy Oil Depressurisation: A Parametric Study of Factors
Affecting Critical Gas Saturation and 3-phase Relative Permeabilities, proceedings, presented at the 2002 SPE International Thermal
Operations and Heavy Oil Symposium and International Horizontal Well Technology Conference, Calgary, Alberta, Canada, SPE
78976, 2002.

Bondino, I., and McDougall, S.R.: Pore Network Modelling of Heavy Oil Depressurisation: A Parametric Study of Factors Affecting
Critical Gas Saturation and 3-Phase Relative Permeabilities, SPE/Petroleum Society of CIM/CHOA 78976, 2005

Bondino, I.: The Application of Pore-Scale Modelling Techniques to Pressure Depletion in Porous Media, PhD thesis, 2005

Danesh, A., Peden, J.M, Krinis, D., and Henderson, G.D.: Pore Level Visual Investigation of Oil Recovery by Solution Gas Drive and Gas
Injection, paper SPE 16956 presented at the 62nd AnnualTechnical Conference and Exhibition of SPE, Dallas, TX (Sep. 1987).

Egermann, P., Vizika, O. (2000). A new Method to Determine Critical Gas saturation and Relative Permeability During Depressurisation in
the Near-Wellbore Region. Soc. Core Analyst. Proceedings, Oct. 18-22, 2000. SCA 36-(1).

Henderson, G.D., Danesh A. and Peden, J.M.: An Experimental Investigation of Waterflooding of Gas Condensate Reservoirs and their
Subsequent Blowdown, proceedings, presented at the 6th European IOR-Symposium, Stavanger, Norway, 1991.

Kalaydjian, F.: The Spreading Coefficient, a Key Parameter Ruling the Efficiency of an Oil Recovery Process by Gas Injection in Three-
phase Flow Conditions, proceedings, presented at the 2nd International Symposium on Evaluation of Reservoir Wettability and its
Effect on Oil Recovery, Edinburgh, Scotland, U.K., 1992.

Kamath, J. and Boyer, R.E.: Critical Gas Saturation and Supersaturation in Low-Permeability Rocks, proceedings, presented at the 68th
Annual Technical Conference and Exhibition of the Society of Petroleum Engineers, Houston, Texas, SPE 26663, 1993.

Kortekaas, T.F.M and van Poelgeest, F.: Liberation of Solution Gas During Pressure Depletion of Virgin and Watered-Out Oil Reservoirs,
SPE, August, 1991

Madaoui, K.: Critical Gas Saturation in Depletion Drive,Reports of International Symp. on HC Exploration, Drilling, and Production
Techniques (Dec. 1975) 203-220.

McDougall, S.R. and Mackay, E.J.: The Impact of Pressure-Dependent Interfacial Tension and Buoyancy Forces upon Pressure Depletion
in Virgin Hydrocarbon Reservoirs, Transactions IchemE, vol. 76, part A, pp. 553-561, 1998.

McDougall, S.R. and Sorbie, K.S.: Estimation of Critical Gas Saturation During Pressure Depletion in Virgin and Waterflooded
Reservoirs, Petroleum Geoscience, vol. 5, pp. 229-233, 1999.

Naylor, P., Mogford, D., and Smith, R. (2000). Development of In-situ Measurements to Determine Reservoir Condition Critical Gas
Saturations During Depressurisation. Soc. Core Analysts, SCA 2000-37-(1).

Poulsen, S., McDougall, S.R., Sorbie, K.S., and Skauge, A.: Network Modeling of Internal and External Gas Drive, reviewed proceeding
paper to be presented at the 2001 International Symposium of the Society of Core Analysts (SCA), Edinburgh, Sept. 2001.

Sahni, A., Gadelle, F., Kumar, M., Kovscek, A.R. and Tomutsa, L.: Experiments and Analysis of Heavy Oil Solution Gas Drive,
proceedings, presented at the 2001 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, SPE 71498, 2001.

Skauge, A. Haaskjold, G.S. Ormehaug, P.A. and Aarra, M.G.: Studies of Production under Bubblepoint, EAGE, 1999

Stewart,C.R., C.R., Hunt, E.B., Schneider, F.N., Geffen, T.M., and Berry, V.J.: The Role of Bubble Formation in Oil Recovery by Solution
Gas Drive in Limestones, Trans AIME 201, (1954) 294--301.

12 SPE 113394
Yortsos, Y. C. and Parlar, M.: Phase Change in Binary Systems in Porous Media: Application to Solution Gas Drive, proceedings,
presented at the 64th Annual Technical Conference and Exhibition of the Society of Petroleum Engineers, San Antonio, Texas, U.S.A.,
SPE 19697, 1989.

ren, P.E., Billiotte, J. and Pinczewski, W.V.: Mobilization of Waterflood Residual Oil by Gas Injection for Water-Wet Conditions,
proceedings, presented at the 7th SPE/DOE Symposium on enhanced oil recovery, Tulsa, Oklahoma, U.S.A., SPE 20185, 1990.

Você também pode gostar