You are on page 1of 15

The Application of Fractional Flow Theory

to Enhanced Oil Recovery


Gary A. Pope, SPE, U. of Texas
Abstract
Classical fractional flow theory is generalized,
starting with the Buckley-Leverett theory for
waterflooding. The mathematics are based on the
method of characteristics. The physics and how the
problems are treated are analogous to the
multicomponent adsorption problem of
chromatography; thus, the methods of Helfferich
apply. Enhanced oil-recovery (EOR) cases con-
sidered are polymer flooding, carbonated water-
flooding, hot waterflooding, hydrocarbon miscible
flooding, low-tension flooding, and
micellar/polymer flooding. The alcohol flooding,
enriched gas, and carbon dioxide cases are analogous
in many respects to the micellar/polymer case.
Finally, three-phase flow problems are treated.
Three-phase flow occurs in a variety of EOR
processes.
Introduction
Fractional flow theory has been applied by various
authors to waterflooding, 1-3 polymer
4
carbonated waterflooding,5,6 alcohol flooding, ,8
miscible flooding,
9
steamflooding) 10 and various
types of surfactant flooding.
11
-
fb
Many of the
assumptions made by these authors are the same and
are necessary for obtaining simple analytical or
graphical solutions to the continuity equations.
Typically, the major assumptions, which are
sometimes not stated explicitly, are: (1) one
dimensional flow in a homogeneous, isotropic,
isothermal porous medium, (2) at most, two phases
are flowing, (3) at most, three components are
flowing, (4) local equilibrium exists, (5) the fluids are
incompressible, (6) for sorbing components, the
adsorption isotherm depends only on one component
01977520/80/00067660$00.25
Copyright 1980 Society of Petroleum Engineers
JUNE 1980
and has negative curvature, (7) dispersion is
negligible, (8) gravity and capillarity are negligible,
(9) no fingering occurs, (10) Darcy's law applies, (11)
.the initial distribution of fluids is uniform, and (12) a
continuous injection of constant composition is
injected, starting at time zero.
Several of these assumptions are relaxed easily.
One of the most useful to relax is Assumption 12,
continuous Injection. The principles of
chromatography can be applied to analyze the more
interesting case of injecting one or more slugs. Most
of these processes require slug injection of chemical
or solvent to be economical. In fact, a lower bound
on the slug size necessary to prevent slug breakdown
can be obtained from a simple extension of fractional
flow theory. In this and other extensions the common
new feature is the need to evaluate more than one
characteristic velocity. A second example of this is
the extension of fractional flow theory from
simultaneous immiscible two-phase flow (the
classical Buckley-Leverett water flood problem) to
simultaneous immiscible three-phase flow (the
classical oil/water/gas flow problem). A third
example is the extension to nonisothermal cases.
Here we need to consider the energy balance, mass
balance, and velocity of a front of constant tem-
perature. A fourth example is when one or more
components are partitioning between phases. In all
cases, mathematically, the extension is analogous to
the generalization from the one-component ad-
sorption problems (or two-component ion exchange
problems with a stoichiometric constraint) to
multicomponent sorption problems.
17
The latter
theory has been worked out in a very general way for
many com9onent systems using the concept of
coherence.
1
Pope et al. 18 recently have applied this
theory to reservoir engineering involving sorption
problems.
191
Considering the classical Buckley-Leverett
problem, I the continuity equation for water can be
written
asw + -.!L dfw asw =0.
at . A(j> dS
w
ax
............. (la)
It follows directly that the velocity of a front of
constant saturation, the characteristic velocity, is
( dX) = -.!L dfw . . .............. (lb)
dt SW A(j> dS
w
Direct integration of this single ordinary differential
equation would solve this problem [give
Sw =Sw (X,t)], except that, as is well-known, a
"shock front" may form. This happens if saturation
velocities, V
sw
, upstream are greater than those
downstream. This is true for most oil/water frac-
tional flow curves between certain limits of
saturation, depending on the curvature of f W.
Between such limits, an overall material balance gives
the velocity of the shock, or saturation discontinuity.
192
1- Sorw
Swr
Sw
(a) Fractional Flow Diagram
1- Sorw
o
Xo
(b) Satu ration Profi Ie Before Water BT
Np
I-S
orw
- 5
wr
o
to
(c) Oil Production Curve
Fig. 1-Waterflooding.
q fW2 -f
wl
v LlS
w
= A(j> Sw2 - Swl
................ (2)
At the contact between the shock and the continuous
saturation distribution, these velocities must be equal
and given by
fW2 - fwl
= dfw I _ .............. (3)
dSw SW- sw2
Waterflood
For the typical waterflood case (Fig. 1), Swl is the
known initial, uniform, irreducible water saturation,
SWP and f wi' therefore, is zero. Thus, Eq. 3 yields
the value for Sw2, or Sw2 Ofln be obtained graphically
(Fig. 1) by constructing a tangent line to the frac-
tional flow curve originating at Swr since Eq., 3 is a
straight line with slope df wldSw and intercept
(SwpO). Until breakthrough of the shock front 100070
oil is produced. The breakthrough (BT) time ex-
pressed in pore volumes is just the reciprocal of Eq.
3. After BT, the Welge integration
2
gives the oil
recovery implicity as
at
dfw )-1
tD=( dS
w
' ....................... (5)
In the above case the initial water saturation, SWi'
was taken to be SWP but Eq. 5 holds for any value of
Swi up to 1 - Sorw. N can be expressed explicitly as
a function of tD for lhe case of straight-line relative
permeabiIities, since for this case the f w curve does
not have an inflection point so that no shock front
forms and since Eq. 5 is simply a quadratic that can
be inverted to give and used in
Eq. 4. The velocity of the injected water is given by
the slope of the tangent from the origin. This line
intersects withf w at (s w3,f w3) in Fig. 1.
Polymer Flood
The next step up in complexity is polymer flooding,
where a very small amount by weight of water-
soluble polymer is added to the water to increase its
viscosity and increase efficient displacement of oil.
Polymers also are used in micellar/polymer flooding,
a more complicated process involving surfactants
and typically applied as a tertiary process - i.e., one
intended to mobilize capillary-trapped or residual oil.
Polymer flooding is used to enhance the water-
flooding process and is essential for an economic
water flood when the oil viscosity exceeds a certain
value. Polymer flooding greatly improves areal and
vertical sweep efficiencies and displacement (or
microscopic) efficiency. The improvement in all three
is important and must be considered to evaluate
properly the merit of a given polymer flood.
However, displacement is the only efficiency con-
sidered here, since it is evaluated from fractional
flow theory.
The polymer transports only in the aqueous phase
(no partitioning to oil). Thus, the continuity equation
SOCIETY OF PETROLEUM ENGINEERS JOURNAL
for it or any other component of a similar nature can
be written
a(ct>iswCi) a [ ] q
. at + at (l - ct P sA i + A
a(cJw)
. =0 ............... (6)
ax
Defining an adsorption C
i
in units of amount
adsorbed per unit pore volume,
C. = (l- ctPSAi
I ct> , ..................... (7)
and
a(SwCi) ct> aC
i
q a(cJw)
--"--'-- + - - + - = o. . .. (8)
at ct>i at Act>i ax
The continuity equation for water is
a(swCw) q a(cwfw)
---"------'-'--- + - =0 ........... (9)
at Act> ax
Making the excellent approximation that the
volume fraction of the polymer is negligible com-
pared with the water (setting C
w
= l) and expanding
Eq. 8 leads to this important simplification:
(
Sw+ dCi ) aC
i
+ qfw aC
i
=0.
d
....... (10)
C
i
at Act> ax
Here we also have used Assumption 4 and have taken
ct>i=ct>
Now it again follows directly that
(
dX) = q f w , ........... (lla)
dt c Act> Sw+Di
I
where the general retardation term, D
i
, replaces
dC;ldC
i
, since in the case of a self-sharpening
polymer concentration front a polymer con-
centration shock forms and the polymer con-
centration jumps from zero (its initial value) to its
injection value. Di for such a case is simply
Cio
Di= - .......................... (lIb)
C
io
This is just the amount of polymer adsorption ex-
pressed in pore volumes, and the polymer front lags
by the factor D1 (see Eq. 11) over what it would
otherwise be.
Generally, two saturation shocks form during a
polymer flood (Fig. 2). The first shock forms as the
water saturation increases from its initial value, since
(as in a waterflood) the saturation velocity is in-
creasing going upstream. The second saturation
shock forms at the polymer concentration front-
i.e., where polymer water contacts initial (connate)
water. These are completely miscible fluids. Using
Assumption 6, the displacement between the two
waters will be ideally sharp or piston-like. Therefore,
at this front the velocities of the saturation shock,
water, a front of constant saturation of water, and
polymer must all be equal. Therefore,
fW3 -fW]. _fW]. _ dfw I _ fW3
SW3 -SW3 - SW]. - dS
w
SW=Sw3- SW3 +Dp ,
........................... (12)
JUNE 1980
where f W2 is evaluated from the polymer fractional
flow curve (the one on the right in Fig. 2), which
differs from the water/oil curve only in that the
aqueous phase viscosity is greater than that of water
due to the polymer.
The principle that velocities must be equal at the
contact between a pair of miscible fluids recurs in
subsequent examples. In this case, solution of Eq. 12
for SW.2 and SWJ gives the values of the water
saturatIOn for the upstream side of the first and
second saturation shocks, respectively. The solution
can be obtained graphically by drawing a straight line
from the point (- D p' 0) tangent to the polymer
fractional flow curve (Fig. 2). The oil production is
100070 until BT of the first shock, then at constant cut
f W2 in the connate water/oil bank until BT of Shock
2, and finally sweepout at decreasing oil cut as per
Eqs. 4 and 5 applied to the polymer fractional flow
curve, since the saturation velocities are given by
q dfw
vsw = Act> dS
w
....................... (13)
for Sw > Sw3. For cases where the initial water
fw
(Sw3,f w3l
OIL AND WATER
WITH POLYMER
--..... OIL AND WATER
-Dp 0 Swr
Sw
(al Fractional Flow Diagram
I-S
orw
Sw
I
I Swr
POLYMER WATER..L.. CONNATE WATER
,I ,
o I
Xo
(bl Saturation Profile Before Water BT
Np
I-S
orw
-Swr
POLYMER BT
o
to
(cl Oil Production Curve
Fig. 2 - Polymer flooding.
193
saturation is greater than SwP the above result
requires slight modification. The first shock is from
Swi >Swr to Sw2' and the oil cuts change
correspondingly.
Low-Tension Flood
Consider the application of a low-tension process to
the case where S wi = S wr' This is generally not the
starting point for such a process since it is meant to
mobilize trapped oil after a waterflood or polymer
flood; however, the analysis is a trivial extension of
the polymer flood case, and so will be discussed
briefly. Actually, there are technical and economic
incentives to apply such a process as soon after
primary as possible, but this generally is not done. By
definition, for the purpose at hand, we will give the
low-tension process the following characteristics.
First, it is an aqueous process involving no transfer
of chemical (the chemical is surfactant in practice) to
the oil. Second, it is a low-concentration process
involving negligible volume fraction of chemical in
the aqueous phase. Third, the chemical lowers the
interfacial tension enough to make the capillary
number high enough to detrap oil, although not
necessarily all of it. Fourth, from a practical point of
view, polymer must be added to the chemical
solution for mobility control. (When we consider
slug processes, we will assume also that polymer has
-os 0
I-Sore
Sw
lOW
TENSION
194
Np
So;
SECON DARY TERTIARY
o Swr
Sw
Swr
Sw
(a) Fractional Flow Diagram
I-Sore.;t
I-S
orw
Sw2
Sw
Sw2
Xo
I
lOW TENSION I
W A T E ~ ~ CONNATE WATER
o I
Xo
(b) Saturation Profiles Before Oil Bank BT
Np
Soi
to to
(c) Oil Production Curves
Fig. 3- Lowtension flooding.
been added to the drive water or "mobility buffer.")
Provided only that we have accounted properly for
the resulting change in the chemical solution viscosity
due to the polymer addition and that the amounts of
polymer and surfactant have been selected such that
ideally no chromatographic separation occurs
between them (Ds=Dp), for the moment we can just
consider a single low-tension fractional flow curve.
This curve will be displaced to the right just as the
polymer alone curve was, but now it will terminate at
Sore rather than Sorw (Fig. 3).
By precisely the same analysis as for the polymer
flood case we find that two shocks form, followed by
an oil "tail," but now the final oil saturation ends at
Sore' not Sorw. The graphical construction is shown
in Fig. 3. The secondary case is shown on the left,
and the tertiary case is on the right. The only dif-
ference between the two cases is that the water
saturation jumps from Swr to Sw2 in the secondary
case and from 1- Sorw to Sw2 in the tertiary case.
This results in much more recovery in the secondary
case even though the final oil saturation is the same
(Sore) in both cases. The production curves (Fig. 3)
have been normalized by the initial oil saturation
(Soi)' If the viscosity of the injected chemical
solution is increased, the low-tension fractional flow
curve shifts to the right (Fig. 4). Comparing Figs. 3
and 4 shows that the oil recovery efficiency continues
to improve as the viscosity of the chemical solution is
increased even beyond the point of theoretical
hydrodynamic stability (no fingering). This sur-
prising result becomes very important when con-
sidering the practical slug case, especially in the
tertiary application of the process. Stated another
way, we have an incentive to increase viscosity,
depending on shape of the relative permeability
curves, to higher values than required for unit
mobility ratio, the classical mobility control design
criterion. At some sufficiently high viscosity the
tangent line will intersect the low-tension curve at or
very close to Sw = 1- Sore' and no additional im-
provement will result for still greater viscosities.
Hydrocarbon Miscible Flood
Although its objectives are clearly the same, the
hydrocarbon miscible process' close resemblance to
the above cases with respect to fractional flow theory
requires some explanation. There are important
differences. We assume that a hydrocarbon solvent
capable of fully displacing the oil will be something
like propane; thus, adsorption will not be significant.
More seriously, because such a solvent is not likely to
be as viscous as the crude oil, it will tend to finger
into it, resulting in a drastic reduction in sweep ef-
ficiency. However, Claridge
19
recently has shown
how this can be obviated by proper grading of
viscosity (accomplished in practice by mixing the
solvent and crude in carefully controlled proportions
with a steadily decreasing fraction of the more
viscous crude oil). In any event, we presently ignore
this complication and consider just two idealized
fractional flow curves. This time the curves are for
the immiscible fluid pairs oil/water and
SOCIETY OF PETROLEUM ENGINEERS JOURNAL
solvent/water, both pairs assumed to have the same
relative permeability curves but differ due to the
different viscosities of the oil and solvent. If we plot
the fractional flow of oleic phase, fo' vs. the
saturation of oleic phase, So, rather than f w vs. S w'
and consider the construction for the solvent
displacement, the similarity to previous cases is
striking when the solvent viscosity is greater than the
oil viscosity (Fig. 5). However, this is not usually the
case, as mentioned above, and so we do not treat the f w
miscible case further.
Carbonated Waterflood
This is the last of the isothermal processes involving
only a single characteristic velocity that we will
consider. Although CO
2
flooding commonly is
considered a miscible process, it does not meet the
above criteria due to the appreciable solubility of
CO
2
in both oil and water under reservoir con-
ditions. In fact, its conditional miscibility with crude
oil, if obtainable at all, depends mainly on suf-
ficiently high pressure. In the carbonated water-
flooding process, the water is saturated with CO
2
and injected. The CO
2
partitions into the crude oil,
reducing its viscosity, swelling it, and increasing
injectivity.
-Os 0
Sw
(a) Fractional Flow Diagram
-
I-S
orw
The continuity equation for the CO
2
(letting Cr
and C? be the CO
2
concentrations in water and oil,
respectively) is Sw
Sw2
a( w 0) q a ( w 0)_
at SwCi +SoCi + Act> ax vwCi + foCi -0 .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (14)
Given the equilibrium relationship between the CO
2
dissolved in the water/oil phase, C?= C?< Cr)' and
again (this time less justifiably) neglecting the volume
the CO
2
occupies in the water,
dCC! acl'V q
[sw + (1- Sw) dcr] a/ + Act> [rw +(1-fw)
dCC! acl'V
. a;- =0; ............... (15)
I
thus,
(
dX ) q [ f w + (1 - f w ) (dC? I dCr ) ] N p
dt Cj= Act> Sw+(1-Sw)(dC?ldCn Soi
.......................... (16)
For the self-sharpening CO
2
concentration case, the
result is similar. A jump increase in CO
2
from zero
to its injected value, occurs and has velocity.
q [fw+(1-fw)K ]
v t.Cj' = Act> S w + (1 - S w ) K ,......... (17)
where
o
Xo
(b) Saturation Profile
a
1.0
to
K . ......................... (18)
(c) Oil Production Curve
In either case the velocity of a front of CO
2
-
saturated water and the aqueous saturation velocity
must be equal where the carbonated water and initial
water contact. Taking the simpler, self-sharpening
case first, this means
JUNE 1980
Fig. 4 - Low-tension flooding at high viscosity.
195
Swr
SOLVENT AND
WATER
OIL AND WATER
So
Fractional Flow Diagram
Fig. 5 - Hydrocarbon miscible flood.
o Swr
Sw
(a) Fractional Flow Diagram
Sw
I Swr
CARBONATED WATER+ WATER
o I
Xo
(b) Saturation Profi Ie Before Water BT
Np
I-Sorw - Swr
1.0
to
(c) Oil Production Curve
Fig. 6 - Carbonated waterflooding.
196
[
fw+ (1-fw)K] dfw I
S (1- S )K = _.. .... (19)
w+ w w
This is the equation of a straight line of slope
dfw/dSw passing through the point (-K/I-K,
-K!1-K), as in Fig. 6. Starting at Sw;=SwP first
100070 oil is produced until the first shock BT, then an
oil/water bank free of CO
2
is produced at constant
oil cut f02 until CO
2
BT at the second shock, and
finally oil production at decreasing cuts along the
carbonated water/oil curve until Sorw is reached.
This result is qualitatively just like that for polymer
flooding, but because the CO
2
generally will be
retarded more due to its absorption into the oil than
polymer will due to its adsorption on the rock,
polymer generally will be more efficient.
Hot Waterflood
Coupling the energy and material balances, assuming
heat convection is the dominant form of energy
transport, and going through an analysis similar to
the above (see Appendix A) gives the velocity of a
temperature front for a hot waterflood:
(
dX) q [A A A ]
dt T= Act> PoCo+(PwCw-PoCo)fw
I [Po Co + (PwCw -poCo )Sw
+ p S C R (1 - ct> ) let> ]. .......... (20)
The oil recovery follows directly from this result.
Slugs
For simplicity and definiteness, consider the
hydrocarbon solvent process first. The principles
apply to the other slug processes as well. To be
practical, we must follow the expensive solvent slug
of pore volume Vps with a gas drive under sufficient
pressure for the gas to miscibly displace the solvent.
The velocity of the solvent at S wr is
Vs = -.!L :!so I ..................... (21)
Act> 0 Swr
For convenience, we define a normalized velocity,
Us=vs/(q/Act. Thus, its BT time in pore volumes
is
BT 1
ts = - ............................ (22)
Us
The velocity and BT time of the gas are
1
U
g
= -- .......................... (23)
I-Swr
and
BT 1
tg = - + Vps' ...................... (24)
U
g
Thus, the minimum slug size of solvent required to
prevent deterioration of the solvent slug is
Vps= - =[ :!s: ]:w: +Swr-l. ... (25)
This and similar results are conveniently represented
on a time-distance diagram (Fig.7).
SOCIETY OF PETROLEUM ENGINEERS JOURNAL
In Region I on Fig. 7, the saturations are just the
initial values, So; and Swi. Region A is the oil bank,
Region B is the solvent slug at irreducible water
saturation, and Region C is the gas drive, also at
irreducible water saturation. Between Regions A and
B is a transition zone where both solvent and water
flow. The slope of each line on the time-distance
diagram is the velocity of a particular saturation. The
solvent slug size, Vps is chosen so that the leading
edge of the gas drive just intersects the slowest of
these saturations at the outflow end (X D = 1). If the
fractional flow diagram has zero slope at SwP the
required slug size is infinite. However, for practical
purposes, the slope at a slightly greater water
saturation than Swr can be used to size the solvent
slug. Of course, a small amount of oil will be
trapped. Depending on the shape of the
solvent/water fractional flow curve, very small to
very large solvent slugs may be indicated. In the latter
case, fingering and dispersion actually will be the
limiting factors on economical slug size. The frac-
tional flow result always will be a lower bound.
The same tD -xD diagram can be used to
represent the low-tension slug size requirement
qualitatively if Band C are relabelled chemical slug
and polymer drive and the transition zone is
simultaneous flow of the two phases oil and chemical
solution. However, we can influence the result by
increasing the viscosity of the chemical slug and
polymer drive by increasing their polymer con-
centrations. This shifts the low-tension (surfactant
and polymer in water plus oil) fractional flow
diagram to the right on a (Sw' / w) plot. As
previously discussed, at some high
viscosity the displacement will be piston-like. Then in
the absence of dispersion and adsorption (there will
be no fingering since the slug mobility will be less
than that of the oil bank and the drive mobility equal
to or less than that of the slug), the slug size can
approach zero. Without dispersion, the actual slug
size required will be Ds. With dispersion, the results
will depend on how the tension varies with surfactant
concentration but always will be greater. Never-
theless, a smaller slug size can be bought at the ex-
pense of more polymer. Since the slug cost per barrel
generally will be much greater than the polymer drive
cost per barrel, this will be an attractive tradeoff up
to a certain point.
Three-Phase Flow
Several circumstances involve three phases in
reservoir flooding processes. The classical one in-
volves oil, water, and free gas that has evolved from
solution with the crude oil as the pressure declines
below the crude oil bubble point. At some point
water may invade from an aquifer by edge-water
. drive or be injected as a water flood or polymer flood.
In other cases, inert gas may be injected or natural
gas reinjected. Among the EOR processes, surfactant
and carbon-dioxide flooding are two examples that
sometimes involve three or more phases.
The case of oil/water/gas immiscible flow in
porous media appears below, and Appendix B shows
how other three-phase cases can be treated with the
JUNE 1980
to
Fig. 7 - Time-distance diagram.
same approach. The conservation equations for
water, gas, and oil are, in turn,
as
w
q a/
w
at + AI/> ax =0, ................... (26)
................... (27)
and
as
o
q a/
o
- +-- =0 .................... (28)
at AI/> ax
Since we have
Sw +So +Sg = 1, ...................... (29)
only two of these are independent. Neglecting gravity
and capillarity, the fractional flows are
/
= krw/Jlw
w , (30)
krw/Jlw +ko/Jlo +kg/Jlg
......... (31)
Using the method of characteristics, 20,21 Appendix B
shows that the two characteristic velocities for this
problem are given by
(
dX) q 1 [
u = dt = M"2 /11 + h2
.J (fll + h2)2 - 4(fl tf22 - /IV21 ) ] . . .. (33)
The characteristics in saturation space are given by
dS
w
-/12
= ............... ,. .. . (34)
dS
g
u -h2 .
Since the/ij are functions of saturation only, Eq. 34
can be integrated independently to give saturation
197
Fig. 8 - Schematic saturation paths of three immiscible
fluids.
o
SoO
SwO
SgO
to ......
Fig. 9 - Schematic timedistance diagram for threephase
flow.
z
o
t-

a:::
;:)
t-

en
o
50 B
r-----j
I 5 .
L_o.!...
5 .
Sor _ -- 5 ..... lIL.
______ - ,-__ ...;w;;.;r-j' __ -I
:
.
-------------._._._._---------_ ..
x
Fig. 10 - Saturation profiles for oil/water/gas.
ROUTE "I
INITIAL
SATURATION
Sw So
FI NAL SATU RATION
Fig. 11 - Saturation paths for water displacing oil and gas
at Swr.
198
paths (Fig. 8). The procedure is very similar to that
for either two-component competitive adsorption or
ion exchange with stoichiometry.
Helfferich 2 has shown the equivalence for any
number of invarient phases of arbitrary composition.
One special feature of multiphase flow is the
existence of residual trapped quantities of each phase
at low saturations of the phase trapped. These are
designated SWr> Sgr and Sor and shown in Fig. 8 as
dashed lines parallel to the phase trapped. This
results in a small triangle within the saturation
triangle. Inside the smaller triangle all three phases
flow; outside only one or two phases flow.
In general, once Eq. 34 has been integrated for
particular fractional flow curves, the saturations can
be mapped in distance-time space (Fig. 9). Two
saturation waves develop separated by a plateau of
constant saturation. One or both of the saturation
waves may be self-sharpening, in which case, for step
changes in input, a single discontinuity or shock
front forms and the wave collapses into a single line
on the t D - x D plot. A discontinuity from the plateau
followed by a diffuse zone is also possible. The latter
case is illustrated by the polymer flooding case
(second or upstream wave of Fig. 2b). Now
saturation profiles can be obtained by crossplotting
at fixed time (Fig. 10) and saturation histories ob-
tained by plotting at fixed distance, usually at xD = 1.
Actually, we usually are interested in production
curves rather than saturation histories. To calculate
production, we need the fractional flow of each
fluid. These are calculated using Eqs. 30 through 32
and the known saturations.
Suppose the initial water saturation is at its
residual value and both oil and gas are mobile. Only
water is injected. This problem was solved
numerically by Gottfried et al.
23
The solution takes
on particularly simple form for this case of three-
phase immiscible flow. The two possible saturation
routes are shown in Fig. 11. We must decide which is
physical based on the first wave traveling faster than
the second. Ordinarily, the gas will be much less
viscous than the oil and water, so the first wave will
be an oil/gas wave at residual water. This is followed
by an oil bank at residual gas, a shock to some lower
oil saturation, then slowly decreasing oil saturation
until sweepout to Sw = 1 - Sor - Sgr (see Fig. 10).
This is just what Gottfried et al. found numerically.
Since only gas and oil flow downstream of the first
wave and only water and oil flow upstream of the
first wave, this case can be analyzed by the usual
fractional flow diagrams. These are shown along
with the appropriate graphical construction in Fig.
12; the oil production curve is in Fig. 13.
Three-Component, TwoPhase
Displacements
This case can be applied to several processes.
Examples are alcohol flooding and surfactant
flooding involving only two phases. Also,
displacements with hydrocarbon gas or carbon
dioxide can be approximated provided the crude and
gas can be modeled by three pseudocomponents and
SOCIETY OF PETROLEUM ENGINEERS JOURNAL
the pressure does not vary too much. If the gas and
crude were miscible in all proportions, the previously
considered miscible case would be sufficient.
Generally, however, the phase diagram at fixed
temperature and pressure will involve a two-phase
region (Fig. 14). All the above processes involve
significant interphase mass transfer. In the surfactant
process, for example, except at very low surfactant
concentrations, the aqueous or oleic phase or both
will contain high concentrations of all of the three
components (oil, water, and surfactant) and the low-
tension case previously discussed will not apply.
The differential material balance for any com-
ponent, i, in two-phase flow is
a( ~ ) q a
- c-,S, +C,S2+C +--
at I I I Act> ax
. (C
il
!, + C
i2
h) = o. . .............. (35)
Welge et al.
9
solved this problem for the enriched gas
case. This is the problem that results when a gas of
composition such as Point D in Fig. 14 displaces an
oil of composition such as Point A. There is no
mobile water and no solubility of the hydrocarbons
in the connate water. W achmann
8
solved this
problem for the alcohol case. Larson '5 recently
solved the surfactant case. In the alcohol and sur-
factant cases, only one component is necessary to
represent the oil, but water is mobile and partitioning
between the aqueous and oleic phases is significant
volumetrically. Thus, a phase diagram such as in Fig.
15 is appropriate. The two cases are the same
mathematically, except for the adsorption of the
surfactant and the decreased oleic phase residual
saturation if the interfacial tension is reduced suf-
ficiently, which is usually the case for the surfactant
case. Larson's analysis was restricted to the case of a
self-sharpening surfactant front.
The saturation and surfactant concentration
profiles are shown in Fig. 16. The saturation profile
is qualitatively the same as for the low-tension
surfactant flood. As the oil is banked up, a jump in
oil saturation occurs from Sorw to SoB (assuming, as
is nearly always true, that the velocity of higher
saturations are greater than those of lower
saturations between these limits). Then at the sharp
surfactant front, the surfactant concentration in-
creases from zero downstream to aqueous phase and
oleic phase values upstream corresponding to the
tieline shown in Fig. 15, which passes through the
injected composition, Point D. Without such a tieline
extension, a jump from the oil bank composition,
Point A, to the plait point, Point P, occurs, resulting
in a completely miscible displacement. This is
perhaps the most significant item of Wachmann's
entire analysis. Compositions in that part of the
diagram labeled miscible in Fig. 15 will displace all
the oil miscibly regardless of interfacial tension,
relative permeabilities, etc. This is true regardless of
the sharpening behavior of the surfactant front. For
the small slug case, such displacements will be limited
only by the adsorption and dispersion of the slug and
its hydrodynamic stability.
JUNE 1980
o
Sg = Sgr
5wr o
Sw = Swr
1-5
0r
-5
wr
5 w 5
g
Fig. 12 - Fractional flow curves.
to --...
Fig. 13 - Oil production with initial gas.
C2-
C
S
Fig. 14 - Phase diagram for hydrocarbons.
SURFACTANT OR ALCOHOL
BRINE A B
OIL
Fig. 15 - Phase diagram for surfactant or alcohol.
'99
fw
Sw
Fig. 16 - Surfactant concentration and oil saturation
profiles.
If the injected composition is in the miscible region
of Fig. 15, all original water and oil are displaced,
and with no adsorption the injected fluid has unit
normalized velocity. At contact with the oil/water
bank, the velocities of the water, oil, and injected
fluid are equal, so that
fw = fa = 1
Sw So .
..................... (36)
Graphically, a straight line from the origin to the
point (1,1) (Fig. 17) gives the values for SwB andf wB
or their complements, SoB = 1 - SwB and foB = 1 -
SwB' If the injected surfactant adsorbs Ds pore
volumes per injected pore volume, all velocities are
retarded to 1/(1 + D
s
), and the line should be drawn
from (-Ds' 0).
Now return to the nonmiscible case. Given self-
sharpening, only three regions of differing phase
compositions exist. Far downsteam just oil and water
flow, and far upstream the injected composition
exists. Somewhere in between a surfactant shock
must exist. The velocity of this shock, v Ac is obtained
by making an overall material balance on a moving
front (the shock) across which the concentrations and
saturations discontinuously changed from Cijand SJ
to Cij and SJ. The result is
v = - f') - I'" - 1")
ACi A</> II I 11 I tV I tV I
+Ci2 - Ch J/[ (Cii Si' - Cil Si)-(Ci2Si'
-ChS l)+Cjz-CiI +Cr-C;J. ..... (37)
Since downstream of the surfactant (Component 3)
C31 = C32 = C
3
= 0, .................... (38)
then
200
z:
a
l-
e:(
a::
I-
z:
UJ
u
z:
a
u
I-
z:
e:(
I-
U
e:(
l.&..
a::
::::>
en
o
C32
r-----'
I
I
I C"
I-_M-_-I
I
I SoB
I


Sorw
Sore
r l
Fig. 17 - Completely miscible displacement.
o
I
z:
a
l-
e:(
a::
::::>
l-
e:(
en
...J
a
q fi'+b
VAC3 = A</> Si' +a ' ................. (39a)
where
b= C
32
and
C" c"
31 - 32
C" CAli
...................... (39b)
a= 32 - 3
C
31
- C
32
. . ................... (39c)
The C
3j
are just the phase concentrations at the end
of the tleline shown on Fig. 15. The value for Si' (and
hencefi' from the low-tension fractional flow curve)
can be obtained as follows. The velocity of any
saturation in a two-phase flow region of constant
composition is, from Eq. 35,
q dfl
vs! = A</> dS
I
. . ..................... (40)
At saturation Si', v AC3' and vs! must be equal, so
f'{ +b dfl
S" =-dS I ................. (41)
I +a I s! =sr
The only unknown in this equation is Si'.
Graphically, a line from (-a, -b) tangent to the
low-tension fractional flow curve yields the value for
S'{ (the total composition here is designated C' on
Fig. 15). The oil bank saturation is given by applying
Eq. 37 to a second component, say Component 2,
and then eliminatingfi' and Si'. The result is
l' +d
VAC2 = +c ' ...................... (42a)
where
d=C22 -1-b(c21 -c22)' .............. (42b)
and
c=c22 -1-a(c21 -c22)' ............... (42c)
SOCIETY OF PETROLEUM ENGINEERS JOURNAL
We identify Si as the water saturation in the oil
bank, SwB' and similarlY.!i =fwB' These values can
be obtained graphically by a parallel to the above
tangent passing through (- c, - d). All other up-
stream saturations are obtained in the usual way by
integrating Eq. 4 until the injected composition is
reached.
If the injected composition lies on the binodal
curve, its composition is (Ci'I' C
lI
, C
3I
) and the
final oleic phase saturation will be Sorc' However, if
the injected composition is in the single-phase region
such as Point D (Fig. 15), a second shock front exists
upstream of the first surfactant shock. At this shock,
So goes to zero since only a single phase exists. This
can be thought of as a solubilization front. The
velocity of this shock is given by making a material
balance between the injected composition, say Point
D, and a point in the two-phase region, say Point C.
The result is
q fa c +R
U LlC2 = Act> So c + R ' ................. (43a)
where

R= , .................... (43b)
C
lZ
-ClI
where the phase concentrations at Point C are just
C
lZ
and C
lI
' and is the injected oil composition.
The oleic phase saturation at Point C, sg, is given by
the requirement that the velocity of this saturation
also must be the same as given by Eq. 40. Thus,
, .fb +R dfo I
c =-- .
So +R dS
o

.............. (43c)
In summary, the composition route starts with the
given initial overall composition, A, and going
upstream, develops as follows.
1. A jump to Point B, the oil bank at C
3
= O. Use
Eq.42a.
2. A jump to Point C' on the tieline passing
through Point D. Use Eq. 41.
3. A continuous decrease in So from sg' to sg
occurs at fixed phase compositions, Cij, until Point
C is reached. Use Eq. 40 for continuous change and
Eq. 43c to get Point C.
4. A jump from Point C to Point D, the injected
composition, which is a given known.
The- oil recovery is given as follows. Before oil-
bankBT,
Np =foitD; ......................... (44a)
for t Dl :5, t D :5, t DZ (during oil-bank production),
Np =foitDl + foBt
D
; .................. (44b)
for tD3 <t
D
<tDZ (after surfactant BT),
Np= (Ci'l -Ci'z)(Sw+tWo) +Ci'z -Swi' . (44c)
and
_ ( dfl )-1
tD - dS
I
, ...... (44d)
where
JUNE 1980
tDl = SwB -Swi , ................... (44e)
fwB-fwi
Si'+a
tDZ = fi'+b
...................... (44f)
and
SC+R
tD3 = lo +R' ....................... (44g)
Np above is not corrected for the oil injected,
t D' These equations reduce to the previous
water flood, polymer flood, and low-tension flood
cases when Ci'l = 1 and Ci'z = 0 (no mass transfer of
oil into water and vice versa).
Consider the slug size required for these surfactant
flood cases that involve significant partitioning at
high surfactant concentration. For the miscible case,
the required slug size is affected only by dispersion
and adsorption. Thus the adsorption in pore
volumes, D
s
' is a lower bound on the slug size. With
dispersion, a somewhat larger slug size will be
required. The slug size for the nonmiscible case will
depend on whether we fix as our criterion the
displacement of all oil in the swept volume or the
displacement down to Sorc only. Several con-
siderations enter here. First, if the capillary number
is high enough Sorc will be zero and the required slug
size will be the same for both criteria. However, if
Sorc is not zero" the decision will depend primarily on
the velocity of the solubilization shock, which often
will be very slow, and on the oil content of the
residual oleic phase. If the slug size is to be large
enough to recover all the oil, then the velocity, us' of'
its back side will be given by Eq. 42a and the required
slug size is
1
Vps= - -1. ........................ (45)
Us
If the slug size is to be large enough to reduce So to
Sorc' then the velocity of its back side will be
Us = -.!L dfo I , ................ (46)
Act> dS
o
So = sore
and the required slug size is
dfo -1
Vps=(-dS) +Sorc-1, ............ (47)
o sore
provided Us is greater than U Llc from Eq. 43a. Just as
in the low-tension case, if Uie injected viscosity is
high, a step decrease from SoB to Sorc will occur. Us
will be given by
l+b
Us = , ..................... (48)
I-
S
orc+a
and the required slug size is
I-Sorc +a
Vps = 1 + b +Sorc -1. . ........... (49)
Here, as before, we have assumed that the adsorption
of surfactant is irreversible and the same as that of
polymer if both are expressed in pore volumes. The
physical significance of Eq. 49 is emphasized by
ZOI
defining a partition coefficient, K, and an adsorption
is pore volumes at concentration C
31
' Ds:
C"
K = ~ (50) C
31
' ......................... .
and
C"
D = _3 (51)
s C
31
......................... .
Simplifying Eq. 49 gives
Vps=Ds+KSorc . ..................... (52)
This result implies the slug must be sized to account
for both adsorption on the rock and absorption into
the residual oleic phase. The latter greatly increases
Vps for large K. However, this is an unduly
pessimistic result since K can be reduced by
decreasing the electrolyte concentration in the
polymer drive to K values as low as 0.01 to 0.1,
depending on numerous other factors. Then
dispersion and ion exchange will play an important
role since an electrolyte gradient will be set up which
depends on these phenomena. These con-
siderations
24
,25 are beyond the scope of this paper.
Discussion
Among numerous limitations to the fractional flow
theory presented, the most basic is the uncertainty in
the relative permeabilities upon which the entire
theory of multiphase flow in porous media is based.
This limitation also applies to more general models
of flow in porous media such as the three-
dimensional, three-phase, compressible reservoir
simulators used by reservoir engineers.. This is
because the basic differential equations, which they
solve numerically by finite difference or finite
element, also have the relative permeabilities in them.
The concept of relative permeability has proved
useful for the simultaneous flow of oil and water at
low Reynolds number. 26 This is amply documented
by various laboratory experiments and by the
analysis of field-scale waterfloods. But for more
complicated circumstances the basis for its use is
more tenuous. For example, only a few three-phase
experiments have been done, except those where one
of the phases was immobile (trapped). Dietrich et
al.
27
recently reviewed these experiments and the
attempts to describe three-phase relative per-
meabilities analytically.
Gravity has been accounted for by several
authors. 1,28-31 In some cases gravity is the most
important factor in the entire process. The criteria
for gravity stabilization are well-known and widely
applied.
Another major limitation of the theory is its
restriction to one dimension. However, its extension
to two dimensions through streamline models
32
,33
can and has been done by numerous workers. Patton
et al.
4
did this for polymer flooding, for example.
Additional permeability layers also are added easily,
provided there is no cross flow .34 As the mobility
ratio deviates significantly from one this becomes a
poor approximation, however. It may even be a poor
approximation at unit mobility ratio if transverse
202
dispersion is significant. This may be true for l a r ~ e
permeability contrasts, thin layers, and small slugs. 5
Capillary pressure plays an analogous role for the
immiscible case. Non-Newtonian rheology is another
factor that may be significant in three-dimensional
flow. For example, polymer injectivities are
significantly higher in radial flow than in linear flow
if the polymer solution is shear thinning. These and
similar limitations are best removed by going to a
fully numerical simulator. Even in one space
dimension these can be quite complex. An example
for the surfactant process is the compositional
simulator of Pope and Nelson.
24
See Pope et al.
25
for additional applications of this simulator.
Even the numerical simulators have major
limitations. One of these is the inadequate
representation of fingering. This is particularly
serious for the CO
2
process, which is inherently
unstable from the onset. Fingering is perhaps the
most serious limitation to describing qualitatively the
unstable processes with either fractional flow theory
or numerical simulation. The best approach we have
to the problem is that of Claridge. 19
Particular limitations in addition to the above also
apply to each particular process discussed. A good
discussion of these for the surfactant process is
presented by Larson.
15
The restriction to self-
sharpening systems is an example. The complete
neglect of electrolyte changes and effects is another
example. Even so, the theory sheds a lot of light upon
how the process works qualitatively, which is its
primary purpose in all cases. An extension to include
the three-phase surfactant problem should be useful
in this regard.
In many circumstances CO
2
is only partially
miscible with crude oil, and if mobile water is
present, three-phases will flow. This process is
complicated also by the need to consider the crude oil
as more than one component, the pressure depen-
dence of the phase behavior and fluid properties, and
fingering. Thus, an extension to the tertiary CO
2
process is both one of the most important and dif-
ficult extensions of fractional flow theory yet to be
made. Particularly important are the issues of
retrapping of CO
2
-mobilized oil by water drive, the
required amount of CO
2
, and the efficacy of
alternate water and CO
2
injection.
A novel extension of fractional flow theory to the
determination of the fractional flow function itself
using multiple partitioning tracers was recently
presented by Deans.
36
It seems safe to assume still
other applications will be discovered.
Nomenclature
A
Ai =
area
adsorption of Component per
unit mass of rock
defined by Eqs. 42b and 42c
concentration of Component i in
fluid per unit pore volume
concentration of Component i on
stationary phase (rock) per unit
pore volume of rock
SOCIETY OF PETROLEUM ENGINEERS JOURNAL
concentration of Component i in
Phasej
specific heats of oil, rock, and
water
D; general retardation factor for
Component i, adsorption of
Component i in pore volumes
per pore volume injected eval-
uated at injected concentration
for self-sharpening cases
f = fractional flow, the flux of a given
phase divided by the total flux
f ij the derivatives off with saturation
as defined in Appendix B
k permeability
k r relative permeability
K = partition coefficient defined by
Eq. 18 or Eq. 50
k T = thermal conductivity
_ L = length of porous medium
M = mobility ratio
Np oil production, pore volumes
q flow rate
Sg gas saturation
So oil saturation
So; initial oil saturation
Sw water saturation
Sw; initial water saturation
Sj saturation of Phase j
t time
T=
u=
v
Vps =
X=
X
D
~
P.
P
Ps
(1
e/>
e/>;
non dimensional time or injected
pore volumes of fluid, qt I ALe/>
temperature
normalized frontal velocity, v
divided by qlAe/>
frontal velocity, qlAe/> for single
phase flow in porous media
pore volume of porous media,
ALe/>
slug size, pore volume
distance or space dimension
nondimensional distance, XI L
difference operator
viscosity
density
gain density Of rock (or matrix)
characteristic directions
porosity
effective porosity accessible to
Componenti
Subscripts and Superscripts
JUNE 1980
BT breakthrough time of injected
fluid
e = chemical or carbon dioxide
D nondimensional
g gas
gr gas residual
component number
j phase number
0 oil
oB oil bank
ore oil residual to chemical or sur-
factant
orw oil residual to water
p polymer
r relative value or residual value
s surfactant
w water
wr water residual
References
1. Buckley, S.E. and Leverett, M.C.: "Mechanism of Fluid
Displacement in Sands," Trans., AIME (1942) 146, 107-116.
2. Welge, Henry J.: "A Simplified Method for Computing Oil
Recovery by Gas or Water Drive," Trans., AIME (1952) 195
91-98.
3. Craig, F.F. Jr.: "The Reservoir Engineering Aspects of
Waterflooding," Monograph Series, Society of Petroleum
Engineers, Dallas (1971) 3.
4. Patton, J.T., Coats, K.H., and Colegrove, G.T.: "Prediction
of Polymer Flood Performance," Soc. Pet. Eng. J. (March
1971) 72-84; Trans., AIME, 251.
5. de Nevers, Noel: "A Calculation Method for Carbonated
Waterflooding," Soc. Pet. Eng. J. (March 1964) 9-20; Trans.,
AIME,231.
6. Claridge, E.L. and Bondor, P.L.: "A Graphical Method for
Calculating Linear Displacements With Mass Transfer and
Continuously Changing Mobilities," Soc. Pet. Eng. J. (Dec.
1974) 609-618; Trans., AIME, 257.
7. Taber, J.J., Kamath, I.S.K., and Reed, R.L.: "Mechanism of
Alcohol Displacement of Oil from Porous Media," Soc. Pet.
Eng. J. (Sept. 1961) 195-212.
8. Wachmann, C.: "A Mathematical Theory for the
Displacement of Oil and Water by Alcohol," Soc. Pet. Eng. J.
(Sept. 1964) 250-266; Trans., AIME, 231.
9. Welge, H.J., Johnson, E.F., Erving, S.P., and Brinkman,
F.H.: "The Linear Displacement of Oil From Porous Media
by Enriched Gas," J. Pet. Tech. (Aug. 1961) 787-796; Trans.,
AIME,222.
10. Shutler, N.D. and Boberg, T.C.: "A One-Dimensional
Analytical Technique for Predicting Oil Recovery by
Steamflooding," Soc. Pet. Eng. J. (Dec. 1972) 489-498.
11. Fayers, F.J. and Perrine, R.L.: "Mathematical Description of
Detergent Flooding in Oil Reservoirs," Trans., AIME (1959)
216,277-283.
12. Davis, J.A. and Jones, S.C.: "Displacement Mechanisms of
Micellar Solutions," J. Pet. Tech. (Dec. 1968) 1415-1428;
Trans., AIME, 243.
13. Larson, R.G. and Hirasaki, G.J.: "Analysis of the Physical
Mechanisms in Surfactant Flooding," Soc. Pet. Eng. J. (Feb.
1978) 42-58.
14. Larson, R.G.: "The Influence of Phase Behavior on Sur-
factant Flooding," Soc. Pet. Eng. J. (Dec. 1979) 411-422;
Trans., AIME, 267.
15. Foster, W.R.: "A Low Tension Waterflooding Process
Employing a Petroleum Sulfonate, Inorganic Salts, and a
Biopolymer," J. Pet. Tech. (Feb. 1973) 205-210; Trans.,
AIME,255.
16. Hales, H.B. and Odeh, A.S.: "An Improved Method for
Simulating Ideal Low Tension Flooding Processes," Soc. Pet.
Eng. J. (April 1976) 53-56.
17. Helfferich, F.G. and Klein, G.: "Multicomponent
Chromatography," Marcel Dekker Inc., New York City
(1970).
18. Pope, G.A., Lake, L.W. and Helfferich, F.G.: "Cation
Exchange in Chemical Flooding - Basic Theory Without
Dispersion," Soc. Pet. Eng. J. (Dec. 1978) 418-434.
203
19. Claridge, E.L.: "A Method for Designing Graded Viscosity
Banks," Soc. Pet. Eng. J. (Oct. 1978) 315-324; Trans.,
AIME,265.
20. Aris, R. and Amundson, N.R.: "Mathematical Methods in
Chemical Engineering," First-Order Partial Differential
Equations With Applications, Prentice-Hall Inc., Englewood
Cliffs, NJ (1973) 2.
21. Courant, R. and Hilbert, D.: "Methods of Mathematical
Physics," Partial Differential Equations, Interscience, New
York City (1962) 2.
22. Helfferich, F.G.: "General Theory of Multicomponent,
Multiphase Displacement in Porous Media," paper SPE 8372
presented at the SPE 54th Annual Fall Technical Conference
and Exhibition, Las Vegas, Sept. 23-26, 1979.
23. Gottfried, Byron S., Guilinger, W.H., and Snyder, R.W.:
"Numerical Solutions of the Equations for One-Dimensional
Multi-Phase Flow in Porous Media," Soc. Pet. Eng. J.
(March 1966) 62-72; Trans., AIME, 237.
24. Pope, G.A. and Nelson, R.C.: "A Chemical Flooding
Compositional Simulator," Soc. Pet. Eng. J. (Oct. 1978) 339-
354.
25. Pope, G.A., Wang, Ben, and Tsaur, K.: "A Sensitivity Study
of Micellar/Polymer Flooding," Soc. Pet. Eng. J. (Dec. 1979)
357-368.
26. Jones, S.c. and Roszelle, W.O.: "Graphical Techniques for
Determining Relative Permeability from Displacement Ex-
periments," J. Pet. Tech. (May 1978) 807-817; Trans., AIME,
265.
27. Dietrick, J.K. and Bondor, P.L.: "Three-Phase Oil Relative
Permeability Models," paper SPE 6044 presented at the SPE
51st Annual Fall Technical Conference and Exhibition, New
Orleans, Oct. 3-6, 1976.
28. Claridge, E.L.: "Prediction of Recovery in Unstable Miscible
Flooding," Soc. Pet. Eng. J. (April 1972) 143-155.
29. Hirasaki, G.J.: "Sensitivity Coefficients for History Match-
ing Oil Displacement Processes," Soc. Pet. Eng. J. (Feb.
1975) 39-49.
30. Hawthorne, R.G.: "Two-Phase Flow in Two Dimensional
Systems - Effects of Rate, Viscosity, and Density on Fluid
Displacement in Porous Media," Trans., AIME (1960) 219,
81-93.
31. Hagoort, J.: "Oil Recovery by Gravity Drainage," Soc. Pet.
Eng. J. (June 1980) 139-150.
32. Caudle, Ben H.: Fundamentals of Reservoir Engineering, SPE
videotape course textbooks (1967, 1968) 1,2.
33. Le Blanc, J.L.: "A Streamline Simulation Model for
Predicting The Secondary Recovery of Oil," PhD thesis, U. of
Texas, Austin (1971).
34. Jewett, R.L. and Schurz, G.F.: "Polymer Flooding-A
Current Appraisal," J. Pet. Tech. (June 1970) 675-684.
35. Koonce, K.T. and Blackwell, B.J.: "Idealized Behavior of
Solvent Banks in Stratified Reservoirs," Soc. Pet. Eng. J.
(Dec. 1965) 318-328; Trans., AIME, 234.
36. Deans, H.A.: "Using Chemical Tracers to Measure Fractional
Flow and Saturation In-Situ," paper SPE 7076 presented at
the SPE Fifth Symposium on Improved Methods for Oil
Recovery, Tulsa, April 16-19, 1978.
37. Lake, L.W. and Helfferich, F.G.: "The Effect of Dispersion,
Cation Exchange, and Polymer/Surfactant Adsorption on
Chemical Flood Environment," Soc. Pet. Eng. J. (Dec. 1978)
435-444.
APPENDIX A
Derivation of Energy Balance Equation
In addition to Assumptions 1 through 12 in the text,
assume (1) constant specific heats (Co' C
w
, and
C R ) , (2) no kinetic or potential energy con-
tributions, (3) negligible radiation energy transfer,
(4) no phase changes, (5) no work terms, and (6) no
significant changes in density with temperature. An
energy balance then yields
a [
at (PwSwCw+PoCo (1-Sw)ct>T
204
a [
+ (1-ctp
S
CR T
J
+ ax pwiwCw +PoC
o
(l-fw)T] + :; )=0 ....... (A-I)
Expanding the derivatives and using the continuity
equation for water,
as
w
q dfw as
w
at + Act> dS
w
ax = 0, ............ (A-2)
we get
[PoCo + (PwCw -PoCo )Sw +Ps (1- ct CRIef>]
aT q ]
. at + Act> PoCo + (PwCw -PoCo )fw
aT a aT
.-+-(k
T
-) =0 ........ (A-3)
ax ax ax
At high flow rate we can neglect conduction (k T = 0),
so from Eq. A-3 the velocity of a front of constant
temperature is
(
dX)
dt T=- (:;)
-poCo )fw ]/[PoC
o
+ (pwCw -poCo )Sw
,ps(1-ctC
R
Ief>]. .............. (A-4)
From this we can calculate the position of the hot
water front, after which the fractional flow and,
hence, oil recovery are given by the hot-water/oil
fractional flow curve. The construction is analogous
to the polymer flood case without adsorption or
inaccessible pore volume.
APPENDIXB
Derivation of Characteristic Directions
Eqs. 26 and 27 can be expanded to give space and
time derivatives of saturation only, since
fw =fw (Sw' Sg) andfg=fg(Sw, Sg). The result is
asw -.!L [a
f
w asw afw _ _
at + Act> as
w
ax + aS
g
ax - 0, ... (B 1)
and
-.!L[ af
g
asw _
at + Act> as
w
ax + aS
g
ax -0 .... (B 2)
Using the notation fll =afw1asw, etc., and nor-
malizing t and X,
asw asw _
!l +f11!l +f12 -0, ........ (B 3)
ut
D
uX
D
aX
D
and
asw
at
D
+121 aX
D
+122 aX
D
-0 .......... (B-4)
SOCIETY OF PETROLEUM ENGINEERS JOURNAL
The characteristic directions or velocities, dXDldt
D
for this pair of equations are given by the deter-
minant
fll f12 0
hI h2
0
dX
D
0 dt
D
0
=0. ... (B-5)
0 dX
D
0 dt
D
The solution for dX Dldt
D
is given by Eq. 33.
Next we show how the above solution method for
three-component immiscible flow in porous media
can be used to solve the general three-component
problem. There are three cases to consider: single-,
two-, and three-phase flow. The single-phase case is
trivial unless adsorption and/ or dispersion are
considered; this case is treated elsewhere. 17,18,35,37
For two-phase flow Eq. 35a can be written
ac a
__ I + __ (Citfl + C
i2
h) =0, ....... (B-6)
at
D
aX
D
where i= 1, 2, or 3. Noting that the term
Cilft + C
i2
h can be expressed as a function of any
two of the total component concentrations, say C 1
and C
2
, and expanding the space derivative,
aC
i
a aC
I
a
- +-(Ctfl +C"2h)- +-
at
D
aC
I
1 1 aX
D
aC
2
ac,
. (Cilfl +C
i2
h)-'" =0, ......... (B-7)
aX
D
Formally, this equation is the same as either Eq. B-3
or B-4 if the appropriate substitutions are made, or
the characteristics can be written down directly.
For three-phase flow the equation of continuity is
ac a
-a 1 + -- (Citfl + C
i2
h + C
i3
h) = 0, ... (B-8)
tD aX
D
JUNE 1980
where i= 1,2, or 3.
With three phases all Cij are fixed. Then expanding
the derivatives are
(
aSI afl)
(C
ll
-C
13
) - + -- + (C
12
-C
l3
)
atD aX
D
(
aS2 + ah ) =0, ............... (B-9)
at
D
aX
D
and
(
aSI afl)
(C
21
- C
23
) - + -- + (C
22
-C
23
)
at
D
aX
D
(
aS2 + ah ) =0, .............. (B-lO)
at
D
aX
D
Noting that
afj = Y.J... aSI + Y.J... aS2 ,
aX
D
aS
I
aX
D
aS
2
aX
D
........ (B-ll)
we get
aS
I
afl aS
I
aft aS
2
-+--+--=0, ...... (B-12)
at
D
aS
I
aX
D
aS
2
aX
D
and
aS
2
+ ah aS
I
+ ah aS
2
=0. . ..... (B-13)
at
D
aS
I
aX
D
aS
2
aX
D
Eqs. B-12 and B-13 are identical to Eqs. B-3 and B-4
if Phase 1 is equivalenced with Phase wand Phase 2
equivalenced with Phase g. Thus, the previously
discussed three-phase solution applies directly to any
three-phase case regardless of the nature and
composition of the three phases. Relative per-
meabilities and all phase properties, of course, will
vary among different systems.
SPEJ
Original manuscript received in Society of Petroleum Engineers office May
1978. Paper (SPE 7660) accepted for publication April 6, 1979. Revised
manuscript received Jan. 29,1980.
205

You might also like