You are on page 1of 31

SPE-175025-MS

Tie-Line Solutions for MMP Calculations By Equations-of-State


Saeid Khorsandi, and Russell T. Johns, The Pennsylvania State University

Copyright 2015, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in Houston, Texas, USA, 28 –30 September 2015.

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
Minimum miscibility pressure (MMP) is one of the most important parameters in the design of successful
gas floods. One technique for calculation of MMP uses the analytical method based on the method of
characteristics (MOC). Although MOC is the fastest MMP calculation method available today, it can
suffer from robustness owing to complex phase behavior observed in some field displacements. The
current MOC method implemented in commercial software assumes that the composition route is
composed of a series of shocks from one key tie line to the next. For some displacements, however, these
key tie lines do not control miscibility and significant errors in MMP can result. One way to eliminate the
error lies in constructing the entire composition route, which is challenging for many complex multi-
component oil displacements.
This paper examines how to determine the unique composition route for all conceivable phase behavior
complexities, making the MOC method for MMP calculation significantly more robust. The approach
relies on solving for all tie lines in the composition route directly, independent of fractional flow. MMP
estimation is based solely on the tie-line route calculation in tie-line space, where both tie-line rarefactions
and shocks may exist. MMPs that result from 1-D displacements of oil by gas are therefore decoupled
from fractional flow, making the calculation significantly easier and more robust. Once tie lines in the
composition route are determined, specific compositions for a given fractional flow can be mapped onto
the tie-line solution, completing the composition route if desired. Compared to current methods, the
formulation in tie-line space has a smaller rank, few or no umbilic points, and as stated is independent of
fractional flow. Example solutions show that MMPs are easily calculated with this approach for complex
displacements that were incorrectly or very difficult to solve before, including multicomponent gas
injection with bifurcating phase behavior.

Introduction
Gas flooding recently became the most widely used and prolific enhanced oil recovery (EOR) technique
(OGJ 2014), and is increasingly being considered for CO2 storage (Li et al. 2014a). Injection of CO2 and
other gases to recover trapped oil is expected to increase, including its use for oil shale (Sheng 2015) or
heavy oil (Okuno and Xu 2014) reservoirs. An increase in gas flooding will require better tools to model
more complex phase behavior encountered. Such complexities can complicate the determination of one
of the most important design parameters in a gas flood, the minimum miscibility pressure (MMP).
2 SPE-175025-MS

Slim-tube experiments are widely accepted as the best experimental procedure to determine the MMP
for miscible gas floods (Jarrell et al. 2002). Slim-tube experiments use real fluids that can capture the
complex interactions between flow and phase behavior in porous media such as those that occur in
condensing and vaporizing (CV) drives (Zick 1986, Stalkup 1987, Johns et al. 1993). Slim-tube
experiments, however, take significant time to conduct, and are expensive. Thus, only a few MMPs can
be obtained this way in practice. Other experimental methods like the rising-bubble apparatus (RBA)
(Christiansen and Haines, 1987) and the vanishing-interfacial tension test (VIT) (Rao 1997) have been
developed to limit slim-tube experiments for MMP calculation but these experiments fail to capture the
interaction of phase behavior and flow that occurs in porous media (Zhou and Orr 1998, Orr and Jessen
2007). The MMPs from RBA and VIT tests are accurate for simple binary displacements, but become less
accurate as the number of components increase (Jessen and Orr 2008) and should not be used in practice.
Computational methods are rapid and convenient ways to complement the otherwise slow and
expensive experimental procedures. There are currently three computational methods to determine MMP:
1-D simulation of slim-tube displacements, analytical methods by the MOC, and multiple mixing-cell
methods. The main limitation in computational methods is that they rely on accurate fluid characteriza-
tions using an equation of state (EOS). Thus, the MMP from an equation-of-state model should agree with
the slim-tube MMP values (Jaubert et al. 2002, Egwuenu et al. 2008). Once a reliable EOS is developed
computational methods can be accurate, fast and robust.
Determination of the MMP by 1-D compositional simulation attempts to mimic the flow in porous
media that occurs in slim-tube experiments (Yellig and Metcalfe 1980). Fine-grid compositional simu-
lations, however, can suffer from numerical-dispersion effects causing the MMP to be in error (Stalkup
1987, Johns et al. 2002). Stalkup (1987) plotted recovery vs. where N is the number of grid blocks.
Use of higher-order methods can reduce, but not eliminate, the effect of dispersion (Mallison et al. 2005).
Yan et al. (2012) developed a parallel algorithm for MMP estimation from 1-D simulations, but used only
one simulation at each pressure without varying the number of grid blocks. 1-D slim-tube simulations are
more cumbersome and time consuming than other computational methods because they require numerous
inputs including relative permeability.
Mixing-cell methods estimate the MMP based on repeated contacts between oil and gas. There are a
variety of published mixing-cell methods, but many do not correctly predict the MMP for CV drives.
Ahmadi and Johns (2011) published a simple but accurate multiple mixing-cell model to estimate the
MMP for any drive mechanism. Li et al. (2014b) extended mixing cell methods to three-phase displace-
ments. The multiple mixing cell method of Ahmadi and Johns (2011) is accurate even for complicated
phase behavior (Mogensen et al. 2009, PennPVT toolkit manual 2013, Rezaveisi et al. 2015, Khorsandi
et al. 2014). Nevertheless, it would still be useful to have other computational methods, such as those
based on MOC, as another check of the MMP. Ideally, for large gas floods the MMP from slim-tube
experiments, 1-D compositional simulation, multiple mixing cell, and MOC should agree before relying
on detailed compositional simulation.
Analytical methods for MMP estimation are based on the analytical solution of dispersion-free 1-D
flow (Buckley and Leveret 1942, Helfferich and Klein 1970, Helfferich 1981, Pope 1980, Dindoruk 1992,
Johns 1992, Dumore et al. 1984, Orr 2007, Lake et al. 2014). Monroe et al. (1990) first examined the
analytical theory for quaternary displacements and showed that there exists a third key tie line in the
displacement route, which they called the crossover tie line. Orr et al. (1993) and Johns et al. (1993)
confirmed the existence of the crossover tie line for CV drives and presented a simple geometric
construction to find the key tie lines (gas, oil, and Nc-3 crossover tie lines) when successive tie lines were
connected by a shock. They demonstrated that the MMP occurs when one of the key tie lines first
intersects a critical point (becomes zero length) as pressure is increased. Johns et al. (1993) further showed
that the crossover tie line controls the development of miscibility in CV drives, and that the estimated
MMP is below the MMP of either a pure condensing or pure vaporizing drive. Johns and Orr (1996) gave
SPE-175025-MS 3

a procedure to calculate the MMP for more than four components, and extended their geometric
construction to calculate the first multicomponent displacement of 10-component oil by CO2.
Current MOC methods for MMP prediction assume that shocks occur from one key tie line to the next
along these surfaces and that there are only Nc-3 key tie lines (Wang and Orr, 1997). We refer to this as
the “shock-jump” MOC method for MMP prediction in this paper. When shocks are assumed from one
key tie line to the next, the MMP is determined when one of these intersecting key tie lines becomes zero
length. The shock only assumption was made because of the general observation that the composition
route traverses a series of nearly planar pseudoternary ruled surfaces (Johns 1992). For many displace-
ments examined, this approach resulted in very small calculation errors of the key tie lines and the
associated MMP (Wang and Orr 1997, Jessen et al. 1998, Yuan and Johns 2005, Ahmadi and Johns 2011).
Jessen et al. (2001) used shock-jump MOC to develop a fast approach to estimate the nontie-line
rarefactions. However, the shock-jump MOC approach can be significantly in error and has other
limitations associated with it because it only solves for a selected few tie lines in composition space
(Ahmadi et al. 2011, Khorsandi et al. 2014).
Yuan and Johns (2005) showed that there are multiple sets of intersecting tie lines that could satisfy
the shock-jump MOC approach. Further, two-phase regions can bifurcate into separate two-phase regions
(Orr and Jensen 1984, Khorsandi et al. 2014) so that multiple critical points can exist between the
intersecting key tie lines. Mogensen et al. (2009) compared the MMPs predicted by various computational
methods for the Al-Shaheen oil displaced by CO2 and noted a significant difference of thousands of psi
between the MMP predicted by the shock-only MOC and other MMP methods for the heavier reservoir
fluids. Ahmadi et al. (2011) explained these differences using a simple pseudoternary diagram and their
mixing cell method. They showed that the two-phase region splits into two separate two-phase regions
(L1-L2 and L1-V regions), and that this bifurcation causes the shock jump MOC method to fail because
the key tie lines no longer control miscibility. Ahmadi et al. (2011) gave an approximate fix for bifurcating
phase behavior by checking the length of the tie lines between each of the key tie lines to identify if a
critical point is present or is forming between them.
Khorsandi et al. (2014) developed the complete MOC composition route for the complex bifurcating
phase behavior displacement of oil by pure CO2 injection using the pseudoternary system from Ahmadi
et al. (2011). The results showed for the first time that a combination of shocks and rarefaction waves
could exist along the nontie-line path. By considering the entire composition path, not just the key tie
lines, the calculated MMP agreed well with those from the mixing-cell method. They also showed that in
this complex ternary displacement the displacement mechanism has features of both a condensing and
vaporizing drive, which was thought to be possible only for gas floods with four or more components.
Further, there was no MMP for some oil compositions considered in the displacements by CO2, although
the displacement efficiency could be very high in those cases.
One approach to correct the MOC limitation is to determine the exact dispersion-free composition route
by avoiding the assumption that shocks exist from one key tie line to the next. This more-accurate MOC
approach could estimate the MMP by constructing the composition routes for varying pressure, as was
done by Johns and Orr (1996) for 10-component oil displaced by CO2. However the MOC solution for
real gas floods with bifurcating phase behavior or a multicomponent injection gas can be more compli-
cated. Thus, in practice a new approach is needed to simplify the construction of the entire composition
route.
Analytical solutions to hyperbolic equations with constant initial and injection conditions are called
Riemann problems. The ultimate goal of solving Riemann problems is to find a complete and automatic
Riemann solver for a specific type of problem. For gas floods, this means finding a Riemann solver for
all possible phase behavior, initial and injection compositions, and number of components. Besides being
able to solve for MMP automatically, the Riemann solver could be used in front tracking methods (Holden
and Risbero 2013) to solve for complex water-alternating-gas (WAG) displacements or other displace-
4 SPE-175025-MS

ments with nonuniform initial and/or injection conditions. Such a solver and front tracking scheme was
developed by Issacson (1989) for polymer floods. Later, Johansen and Winther (1989) included multi-
component adsorption in their polymer model. Juanes and Lie (2008) developed a Riemann solver and
front tracking method for first-contact miscible water alternating gas floods, while Johns (1992) developed
a front tracking algorithm for two-component partially miscible gas floods, where components can
transfer between phases. More complex Riemann solvers for gas floods have not yet been published.
A broad understanding of ternary displacements is the key in finding a Riemann solver for gas floods.
This is because ternary systems are the building blocks for multicomponent displacements. That is,
composition routes follow a series of successive pseudoternary ruled surfaces (Johns and Orr 1996).
Analytical solutions for ternary displacements of various types have been studied by different researchers
(Dindoruk 1992, Johns 1992, LaForce and Johns 2005, Seto and Orr 2009, Khorsandi et al. 2014), but
these solutions are complex to construct and have not led to a complete Riemann solver. A simpler
approach is needed for these more complex displacements.
The idea that fractional flow has no effect on tie-line route and MMP has been developed over the years
by many authors, although initially the view was the opposite. Metacalfe et al. (1973), for example,
believed that the MMP is dependent on fractional flow. Their conclusions were based, however, on coarse
slim-tube simulations that were significantly impacted by a large level of dispersion. Such a high level of
dispersion does not exist in slim-tube experiments. In theory, the MMP should be determined with no
dispersion present (Johns et al. 2002). Stalkup (1987) showed using 1-D simulations corrected for
dispersion that the MMP is not affected by relative permeability for condensing/vaporizing drives. Jaubert
et al. (1998) stated that rock type did not impact experimental MMP measurements for fifty example
fluids. Zhao et al. (2006) compared simulation results to their mixing-cell algorithm, and concluded that
MMP is not affected by fractional flow.
Johns (1992) and Dindoruk (1992) provided evidence that MMPs are likely independent of fractional
flow. Johns (1992) demonstrated that tie lines connected by a shock must intersect at a composition
outside of the two-phase region, and in many cases outside of positive composition space. Dindoruk
(1992) showed that the nontie-line eigenvectors are tangent to the ruled surfaces formed from the
intersection of the tie-line extensions at their envelope curves. Bedrikovetsky and Chumak (1992)
proposed an auxiliary system of gas flood equations similar to the one that Issacson (1998) derived for
polymer flooding. They described the tie-line route for a four-component displacement with constant
K-values using their auxiliary system of equations. Entov (1997) further suggested that potential coor-
dinate transformations could be used such that Nc-2 of the eigenvalues would become independent of
fractional flow, where Nc is the number of components. Pires et al. (2006) expanded on this idea and
developed Lagrangian coordinates to split the equations into two parts; a set of equations dependent only
on phase behavior, and one additional equation based on fractional flow. Dutra et al. (2009) used the
splitting approach to construct a tie-line route for a four-component displacement, but incorrectly showed
an elliptic region in tie-line space. Further, their coordinate transformation has singular points that make
it impossible to solve some displacements (Khorsandi et al. 2015).
In this paper we apply a splitting technique to multicomponent gas displacements to separate the
tie-line solution from fractional flow. Our approach does not suffer from the singularities in present Pires
et al. (2006) and Dutra et al. (2009). The solution in tie-line space is constructed for a variety of fluid
models including pseudoternary displacements with bifurcating phase behavior (Khorsandi et al. 2014),
and four- and five-component displacements. The approach developed offers the potential for finding a
complete Riemann solver for any initial and injection condition. Finally the MMP is calculated for several
fluids using the analytical solution based solely on solving the continuous tie-line problem, where tie-line
rarefactions and shocks can exist in tie-line space. Thus, we eliminate the need for the “shock jump”
assumption in determining the MMP.
SPE-175025-MS 5

Mathematical model
The calculation of MMP relies on the accurate development of the dispersion-free composition route using
the method of characteristics. The multicomponent 1-D dispersion-free displacements can be modeled
with (Helfferich 1981)
(1)

where Ci and Fi are overall volume fraction and overall fractional flow of component i, tD is
dimensionless time or pore volume injected (PVI), xD is dimensionless length, and Nc is the number of
components. Equations (1) can be split into two parts; equations that depend only on tie lines, and an
equation that depends on flow. Splitting is achieved using the following Lagrangian coordinate transfor-
mation (Pires et al. 2006, Khorsandi et al. 2015),
(2)

The signs in Eqs. (2) are determined based on the location of the injection composition relative to the
equal velocity curve (EVC), where the EVC curve is defined as the curve in the two-phase region where
F1 equals C1. The upper sign of Eqs. (2) is used when C1inj⬎C1EVC(⌫1inj), while the lower sign is used
otherwise. ⌫ is a Nc-2 vector that has elements with unique values for each tie line. The condition for
defining the tie-line space by vector ⌫ is described in Appendix A. Substitution of Eqs. (2) and the tie-line
equations in ⌫ space (shown in Eqs. (5) and (6)) gives the following transformed set of equations that
depends solely on phase behavior.
(3)

where ␣i and ␤i are tie-line coefficients, and


(4)

The tie lines and their coefficients are defined by,


(5)

(6)

where is mole fraction of component i in phase 1 and Ki is the K-value of component i. The sign
for the second term of Eqs. (3) is positive if C1inj⬎C1EVC(⌫inj) and negative otherwise. The eigenvectors
are independent of the sign in Eqs. (3), but the sign of the eigenvalues change correspondingly. We
describe the solution for Eqs. (3) with positive sign because the solution for negative sign is very similar.
The final equation that results from the transformation of Eqs.(1) is dependent on fractional flow.
(7)

Ideally, Eq. (7) could be used to calculate the composition route once the tie-line route is determined.
However, this equation is singular in the single-phase region and at the EVC. In addition the flux, ,
is not a single-valued function of the conserved quantity . Thus, Eq. (7) is not used further in this
paper, and instead, Eqs. (3) are first solved for the tie lines using the specified boundary conditions, and
then in a separate step Eqs. (1) are solved by mapping the fractional flow onto the fixed tie lines (similar
to Gimse and Risbero, 1992).
The characteristic equation using Eqs. (3) are
6 SPE-175025-MS

(8)

where R is a Nc-2 by Nc-2 matrix with elements . The elements of matrix R can be calculated
analytically as shown in Appendix B. The eigenvalues of matrix R are related to the envelope curve of
the tie lines. Each tie line is tangent to Nc ⫺ 2 envelope curves (Dindoruk 1992) at the corresponding
composition, , so that for the kth eigenvalue,
(9)

A series of tie lines tangent to the same envelope curve form a ruled surface and the ruled surfaces can
be calculated by integrating the eigenvectors of Eqs. (8), given by . These ruled surfaces are planar
when K-values are constant (independent of composition) as shown in Appendix C.
The eigenvalues of the characteristic equation (Eqs. 8) give the allowable rates for the overall
Lagrangian flux that can be taken to solve the strong form of Eqs. (3), while the eigenvectors give
allowable Nc-2 ruled surfaces on which the composition route must lie as described by Johns (1992) and
Johns and Orr (1996). The eigenvalues for the tie lines in tie-line space can be both positive and negative,
but are always real. The solution consists of following the eigenvector paths (the strong form) from the
initial to the injection condition, where the solution in tie-line space should be single valued. A
single-valued solution means the path taken in tie-line space should have a corresponding monontonic, but
increasing eigenvalue from the injection tie line to the initial tie line. If the tie-line solution becomes
multivalued based solely on the strong form, the weak form (shocks) must be introduced. Shocks in
tie-line space must be taken that avoid multivalued solutions, but also must satisfy the entropy conditions.
The Lax (1957) and Liu (1976) entropy conditions or the vanishing viscosity condition (Bianchini and
Bressan 2005) can be used to check shock admissibility.
The eigenvalues are easily written for simplified systems using Eqs. (8). For three-component gas
floods, there is only one eigenvalue compared with two in the traditional method that solves Eqs. (1). This
makes the solution of three-component problems simpler, where the sole eigenvalue is equal to and its
eigenvector is trivial (arbitrary scalar). For four-component displacements, there are two eigenvalues
given by,
(10)

where the eigenvalues can be calculated as,


(11)

Each term in Eq. (11) can be calculated analytically as shown in Appendix B. Once an eigenvalue is
determined, its corresponding eigenvector can be calculated from Eqs. (10) or more generally for any
number of components from the matrix R given in Eqs. (8).
Shocks must be conservative. Thus, the upstream and downstream values must satisfy the weak
solution (Rankine-Hugoniot condition) for Eqs. (3),
(12)
SPE-175025-MS 7

where ⌳ is the shock rate of change of the overall Lagrangian flux. Although ⌳ is not a velocity, we
use the term shock velocity loosely as shorthand for ⌳ (and for ␭) in this paper. Eqs. (12) are satisfied if
the two tie lines intersect at.
(13)

Johns and Orr (1996) showed a similar result to Eq. (13), where the velocity of the shock was calculated
by a triangular geometric construction in composition space. In tie-line space, however, we do not need
to apply the geometric construction to find the tie-line routes, although, if desired it can be used to find
the upstream and downstream compositions of the shock once the tie-line route is known.
Initial and injection tie-line selection. There is a potential problem with the solution of the tie-line
route that requires additional explanation. Larson (1979) proved that shocks in and out of the two-phase
region must shock along a tie-line extension. However, multiple tie lines can extend through the initial and
injection compositions if they are single-phase compositions, especially for more complicated phase
behavior that includes bifurcation of a two-phase region. Thus, the correct initial and injection tie lines for
use as boundary conditions in Eqs. (3) must be carefully selected. In general, we first assume that the tie
line found from a negative flash calculation that extends through the initial or injection composition is the
correct tie line. This may or may not be true, and a method is needed to check its validity. One possibility
is to first solve the tie-line route based on these boundary conditions and then determine the entire
composition route. If the composition route is physical (solution is single-valued, satisfies the mass
balances and entropy conditions) the correct tie lines as boundary conditions are selected. Alternatively,
a simple test could be done without the need to calculate the entire composition route. We determine that
test next.
We first solve for the entire tie-line route to identify rarefactions and shock, and their corresponding
velocities in tie-line space. Next, we transform the auxiliary relations for the boundary conditions in
xD⫺tD coordinates to the Lagrangian coordinates of Eqs. (2), where only single-phase initial and injection
compositions are considered (a two-phase composition has a unique tie line and no test is needed). The
transformed boundary conditions for a single-phase composition that is the intersection of multiple tie
lines with constant ⌫ gives,
(14)

We then check if the solved tie-line route satisfies the transformed boundary conditions by comparing
the eigenvalues at the initial and injection compositions to the eigenvalue of any rarefaction or shock
associated with the tie line that extends through that composition. For example, the eigenvalue of the tie
line through the initial composition (or the shock velocity from that tie line to another tie line) must be
less than the eigenvalue associated with the initial composition as given by Eq. (14). For the injection
composition, the reverse is true. Equation (14) defines the single-phase eigenvalues (d␺/d␾) associated
with the boundary conditions in Lagrangian space as ⫺ 1/C1, ini and ⫺ 1/C1, inj (slopes of Eq. (14)). For
C1,k⫽0, however, the slope is infinite so that special care must be taken. To consider this singularity, we
relate the rarefaction and shock velocities to their intersection compositions as defined by Eqs. (9) and
(13) and give an equation for the validity of the tie-line selected as,
(15)

where is calculated by Eqs. (9) and (13) using the velocities associated with the tie lines connected
to the initial and injection tie lines. That is, Eq. (15) determines if the proposed tie-line route with the
specified initial or injection compositions will violate the velocity condition (single-valued solution).
Equation (15) is useful as a simple check of the validity of a tie line found by a negative flash.
However, if the tie line is found not to satisfy Eq. (15) one must search for another tie line that extends
8 SPE-175025-MS

through the initial or injection composition. Li and Johns (2007) and later Li et al. (2012) developed a
constant K-value flash to calculate tie lines without calculating saturations. Their method can be used to
calculate all tie lines that pass through a composition. Juanes (2008) also developed a similar method to
find all tie lines that extend through a single-phase composition when K-values are constant, but that
method is not practical as the number of components increase. Both methods are not satisfactory for our
purpose because tie lines that intersect at a single-phase composition (initial or injection composition) can
also have different K-values.
In this paper, we develop a new and more robust approach to determine tie lines with different K-values
that passes through the initial or injection composition Consider the case that the tie-line route using the
initial tie line does not satisfy Eq. (15) and therefore, we should find another tie line, , that passes
through the initial composition. We know that and intersect at the initial composition, therefore
these two tie lines, if hypothetically connected by a nontie-line shock, would have a shock velocity of
. The problem therefore reduces to finding a hypothetical nontie-line shock with the velocity
determined by the intersection point. Thus, we use Eqs. (12) to find . That is, we perform a line
search along each Hugoniot locus given by Eqs. (12) and test if .

Example ternary displacements


We first construct solutions in tie-line space for ternary displacements to illustrate the advantages of the
new MOC approach. The first example is a simple case where K-values are independent of composition.
The second example is complex case with bifurcating phase behavior as shown in Ahmadi et al. (2011)
and Khorsandi et al. (2014). One of the key steps in the solution for ternary displacements in tie-line space
is to show whether the injection and initial tie lines are connected with a shock or a rarefaction wave or
a combination of both. In addition, multiple tie lines may pass through the initial or injection composi-
tions. Thus, we must check if the tie line from a negative flash converged to the appropriate tie line using
Eqs. (15).
The two-phase region with constant K-values (0.05, 1.2, and 2.5) is shown in Figure 1 by the solid lines.
Three tie lines are also shown in Fig. 1, labeled TL1, TL2 and TL3. The composition path for displacements
with this type of phase behavior is well studied (Johns, 1992, Orr 2007, Seto and Orr 2009).

Figure 1—Analytical solution for ternary displacement in composition space. Point a is on the envelope curve and is given by Eq. (9).
SPE-175025-MS 9

First we consider displacement of I1 with J1, where only one tie line extends through the initial or
injection composition. In composition space, the nontie-line path takes the rarefaction wave indicated by
R1 in Figure 1 from the initial tie line (TL3) to the injection tie line (TL1). A rarefaction wave occurs here
because the velocity of the nontie-line eigenvalue given by the eigenvalue problem using Eqs. (1)
increases from the injection tie line to the initial tie line. Details of the composition route construction can
be found in Orr (2007).
In tie-line space, we solve directly for the tie lines using Eqs. (3) without considering composition
space. The tie-line coefficients (Eqs. 6) for all tie lines between TL1 and TL3 are shown in Figure 2.
Because , the eigenvalues in tie-line space are the slope of the curve of -␣ versus ␤.
Therefore, the single eigenvalue in Lagrangian coordinates (-d␣/d␤) increases from TL1 to TL3 and the
two tie lines are connected with a rarefaction wave as they are in composition space. Figure 3 shows
graphically the scalar equation solved, which is similar to Buckley-Leverett (1942) theory. Thus,
displacements in tie-line space for ternary systems are easier than Buckley-Leverett solutions because the
flux function in Fig. 2 is strictly concave. The eigenvalue is given along the x-axis in Fig. 3, where
(negative sign is used in Eq. (3)).

Figure 2—Tie-line coefficients for ternary phase behavior of Figure 1 in tie-line space.

Figure 3—Analytical solution for three-component displacements showing shocks and rarefactions in Lagrangian coordinates.

Other solutions in both composition and tie-line space are given in Figs. 1 - 3. If we displace I2 by J2,
TL3 is connected to TL1 by the shock S1 shown in Figure 1 by the dashed line and in Fig. 3 by the dotted
line. A shock must occur here because the eigenvalue decreases from the injection to initial tie line. This
10 SPE-175025-MS

shock is a tangent shock (like a Welge tangent shock) that satisfies the entropy condition because it does
not cut through the flux function in Fig. 2.
Next, we construct the solution for displacement of I3 with J1, where two tie lines, TL2 and TL3, now
pass through I2. A negative flash calculation yields TL3. Once a tie-line route is constructed we apply the
test of Eq. (15) for that tie line to see if it is correct. Equation (15) is only satisfied, however, for TL2
because the value of is equal to C1 at a as shown in Fig. 1. Thus, TL3 is discarded and we search for
another tie line using the procedure outlined previously. It is easy to see in Fig. 1 that the only remaining
tie line is TL2, so a tie-line route is constructed with that tie line. The correct (physical) tie-line route
exhibits a rarefaction wave from TL2 to TL1 (see R2 in Fig. 1 and Fig. 3).
In the last example with constant K-values, we discuss the displacement of I4 with J3 in Fig. 1, where
now ⬎ and the positive sign is used in Eqs. (3). The single eigenvalue is therefore equal to
the slope of the curve of ␣ vs ␤ in Fig. 2. Thus, the eigenvalue increases from TL2 to TL1, giving a
rarefaction wave in tie-line space (Fig. 3). In composition space, the nontie-line rarefaction wave in Fig.
1 is labelled by R3. The composition route is clearly different from R2, while the tie-line route follows the
same tie lines, but in reverse order.
Now, we construct the solution in tie-line space for three-component displacements with bifurcating
phase behavior to demonstrate the application and simplicity of using MOC in tie-line space (see Figure
4). The corresponding MOC solutions using Eqs. (1) in composition space were complex to construct and
can be found in Khorsandi et al. (2014). The solutions are identical, but in tie-line space the problem is
as simple as the Buckley-Leverett problem with an S-shaped flux function. The tie-line coefficients for
the tie lines in Fig. 4 now give an S-shaped flux function because of the bifurcating behavior (see Figure
5). The initial and injection compositions considered along with component properties for the PREOS are
given in Table 1.

Figure 4 —Injection and initial compositions considered for bifurcating phase behavior (Ahmadiet al. 2011). Three tie lines extend
through composition. Points a and b lie on the envelope curve (see Khorsandi et al. 2014).
SPE-175025-MS 11

Figure 5—Tie-line coefficients for three-component bifurcating phase behavior in tie-line space, where negative sign in Eqs. (3) is used.

Table 1—Component properties and compositions for bifurcating phase behavior developed based on Ahmadi et al. (2011)
Properties Compositions

Tc (°F) Pc (psia) ␻ I1 I2 I3 J1 BIP

C1N2 ⫺116.59 667.8 0.008 0.1 0.75 0.455 0 C1N2 CO2 nC4
CO2 87.89 1,071 0.225 0.1 0.1 0.000 1 0.119 0.000
nC4 305.69 550.7 0.193 0.0 0.0 0.272 0 0.013 0.126 0.000
C26-35 1258.83 244.98 1.090 0.8 0.15 0.273 0 0.000 0.094 0.000

Two displacements for Fig. 4 are examined here with the same injection composition J1. Because
, we use the negative sign in Eqs. (3). The eigenvalue shown in Figure 6 is equal to the
slope of the curve -␣ vs. ␤. For the displacement of I2 with J1, there is only one tie line that passes through
I2. The solution for displacement of I2 by J1 in tie-line space, therefore, is a rarefaction wave to TL3
followed by a tangent shock from TL3 to TL5(see Figs. 5 and 6). (This is a tangent shock owing to the
entropy and velocity conditions). There are three possible tie lines for initial composition I1, however. We
can construct the solution in tie-line space for each one and then check if the solution satisfies Eq. (15).
For TL5, is equal to C1 at point ␣ in Fig. 4, which violates Eq. (15). Similarly, for the tie-line path of
TL4, is equal to C1 at point b, which again violates Eq. (15). For TL2, is outside of the phase diagram
and satisfies Eq. (15). Therefore TL2 is the correct tie line and the rest of the solution route is easily
generated as a rarefaction from TL1 to TL2. Figure 7 gives the tie-line solutions plotted against the
Lagrangian coordinate.

Figure 6 —Eigenvalue for three-component displacements with bifurcating phase behavior. Dashed line is a shock.
12 SPE-175025-MS

Figure 7—Analytical tie-line solution showing shocks and rarefactions in Lagrangian coordinates.

Example four-component displacements


In this section, we give tie-line routes for various four-component displacements. The first example
considered is one where K-values are constant and there are multiple crossover tie lines that satisfy the
conditions for shock-jump MOC (Yuan and Johns 2005). The second case considered has K-values that
change with composition based on the PREOS. This second case is the same phase behavior and
displacement as the one considered in Dutra et al. 2009, but here we show that there is no complex
eigenvalue as they reported. The third example is a four-component displacement with bifurcating phase
behavior. This case is too complex to be solved in composition space using the eigenvalue problem based
on Eqs. (1), but here we show it is not difficult to solve using Eqs. (3).
Johns and Orr (1996) showed that the multicomponent solution consists of Nc-2 pseudo-ternary
displacements in composition space. Each of these pseudo-ternary displacements are along one ruled
surface of the route for one family of nontie-line eigenvalues. In tie-line space, the four-component
solutions have one less eigenvalue and only the eigenvalues corresponding to the ruled surfaces remain.
That is, there are Nc-2 eigenvectors in tie-line space, so for four components there are two dependent
parameters to describe tie-line space. This ternary diagrams can be used to represent the solutions. The
largest eigenvalue in tie-line space is associated with the eigenvector path we term the “fast path.” The
“slow path” is for the smallest eigenvalue. Again, these eigenvalues are not velocities like they are for Eqs.
(1), but for simplicity we keep that terminology. Physical solution routes must give single-valued
solutions (velocity conditions must be satisfied) and any shocks present must satisfy the entropy
conditions.
Yuan and Johns (2005) demonstrated the possibility of multiple solutions for the crossover tie line in
the shock-jump MOC method. They considered relatively simple four-component phase behavior using
K-values independent of composition (see K-values and compositions in Table 2). We consider the same
case here to show that tie-line route is easily found using tie-line space. The initial and injection tie lines,
and multiple crossover tie-line solutions for the shock-only MOC approach are given in Figure 8. Yuan
and Johns (2005) only gave two of these three crossover tie lines, while we found three. Only one tie line
extends through the initial or injection compositions.
SPE-175025-MS 13

Table 2—The compositions and K-values for example case (Yuan et al. 2005)
x1 x2 x3 x4

K-values 5.0 2.0 1.2 0.01


Oil 0.0386 0.3485 0.0874 0.5256
Gas 0.1984 0.000 0.000 0.8016

Figure 8 —Quaternary displacement with three possible crossover solutions based solely on shock-jump MOC (Yuan and Johns, 2005).
The two-phase region for each ternary face is outlined by the blue and purple dashed lines.

The ruled surfaces that a composition route must follow in tie-line space are shown in Figure 9. The
ruled surfaces in tie-line space are not surfaces, but are curves, or in this case lines because K-values are
constant. The shock loci coincide exactly with these lines as well (see Appendix C). Figure 9 shows
several fast and slow paths for two different representations, one using a ternary diagram of ␤i where
, and another a Cartesian plot of the equilibrium phase composition for one of the phases (xi).
Thus, with four components the tie-line route is given only by two parameters. Unlike the composition
route solution using Eqs. (1) where there are 2(Nc-2) umbilic points for each tie line, there is only one
equal eigenvalue point (umbilic point) in the entire tie-line space for this problem. This fact greatly
simplifies the construction in tie-line space since there is only one point where an equal eigenvalue switch
can be made from one path to the other. Figure 10 shows the eigenvalues in tie-line space and the single
umbilic point along the ␤1–␤3 axis. No other umbilic points exist in tie-line space.
14 SPE-175025-MS

Figure 9 —Tie-line space and eigenvector paths (ruled surfaces) for constant K-values. The correct crossover tie line is Solution 2.

Figure 10 —Eigenvalues in tie-line space along the line ␤2 ⴝ 0

The three possible crossover tie lines according to the shock jump MOC method are also plotted in Fig.
9. The only physical route that can be constructed is the one with crossover tie line 2 (solution 2 in Fig.
9). Figure 9 shows that the correct solution consists of two shocks S1 and S2 in tie-line space. Crossover
tie lines 1 and 3 are not possible because they would give a multivalued solution (violate the velocity
condition). For example, for crossover tie-line solution 1, the route traverses from the injection point in
tie-line space along the slow path that corresponds to the ␤1–␤3 axis and then continues along the fast path
portion of the same axis until a path to the initial tie line is found. A switch, however, at the corresponding
slow path through the initial tie line is not allowed (violates the velocity condition). Other combinations
of shocks and rarefaction waves can also be eliminated for the same reason.
The tie-line route does not change if we reverse the initial and injection tie lines. This is easily shown
because the slow paths just become fast paths and vice versa. The composition route, however, will
change as we demonstarted for the ternary displacements in Fig. 1. For the reversed displacement, the sign
changes for the second term in Eqs. (3). This result is important because it may explain why complicated
reservoir flow simulations are bounded by the mixing-cell tie lines (see Rezaveisi et al. 2015).
To construct the tie-line route automatically and calculate the MMP, we need to parametrize the lengths
traversed along each pseudo-ternary ruled surface (shock or rarefaction ruled surface) in tie-line space so
that the initial and injection tie lines are connected. We use li as the eigenvector length to travel along
SPE-175025-MS 15

ruled surface i. The value of li for NC-2 ruled surfaces can be positive or negative depending on the
eigenvector (or shock) direction taken, while li,k defined below is always positive,
(16)

where li,k is introduced to allow for a combination of k rarefactions and shocks along one of the
pseudo-ternary ruled surfaces. For four components, we have only two values of li, but each of these has
n segments.
Johns (1992) showed that a pseudo-ternary displacement has only one nontie-line shock or rarefaction
along it if K-values are strictly ordered (n ⫽1 in Eq. (16)). The ternary displacement with bifurcating
phase behavior, however, can have both shocks and rarefactions along the same ruled surface (Khorsandi
et al., 2014). When both exist along a given ruled surface, the shock velocity is the same as the eigenvalue
at which the rarefaction starts or ends. We assume here that the route along a ruled surface can consist
of a maximum of only two parts (rarefaction and shock) because it is unlikely that for crude oil
displacements the two-phase region will bifurcate into three two-phase regions. Therefore, we define the
parameter li to describe each pseudo-ternary displacement as,
(17)

The sign in Eq. (9) is positive if the ith eigenvalue increases from the left tie line (upstream) to the right
tie line (downstream). Integration of rarefaction paths can be done numerically to the accuracy desired.
For shocks, we move along the shock locus starting from the left tie line and increase the size of the shock
along the shock locus using the eigenvector path as the first guess.
We illustrate the solution procedure for the parameters in Table 2. Figure 11 shows the slow and fast
paths for this phase behavior, along with the location of the injection and initial tie lines. The solution
route in tie-line space then reduces to finding the scalars l1 and l2 so that d ⫽ 0. The vector d is a function
of li so Newton’s method is easily used. For this displacement n ⫽ 1.
16 SPE-175025-MS

We now consider the more complex case where K-values are dependent on composition using PREOS
(see Table 3 and Dutra et al. (2009)). For this case at 2900 psia and 160°F, the tie-line route and critical
locus is given in Figure 12. Tie-line space is not defined outside of the critical locus (no tie lines extend
through it). The tie-line route in Fig. 12 consists of shock S from the injection composition to the crossover
tie line (TL1), followed by the rarefaction wave R. Other routes can be constructed for different pressures,
which give different converged values of l1 and l2 from Newton iteration (see Figure 13). The parameter
l1 is negative because it is a shock. The sign is important because it determines whether that portion of
the tie-line route is a shock or rarefaction wave.

Table 3—Component properties and compositions for displacement by Dutra et al. (2009)
Properties Compositions

Tc (°F) Pc (psia) ␻ I1 J1 BIP

N2 ⫺232.51 492.32 0.040 0.0 0.8 N2 C3 C6


C3 205.97 615.76 0.152 0.0 0.2 0.085 0.000
C6 453.65 430.59 0.296 0.3 0.0 0.150 0.027 0.000
C10 611.16 353.76 0.576 0.7 0.0 0.155 0.020 0.000

Figure 11—Parametrization of tie-line routes quaternary phase behavior (Table 2) for solution by Newton iteration. The scalar
parameters l1 and l2 are determined so that vector d is zero.
SPE-175025-MS 17

Figure 12—Phase diagram in tie-line space and ruled surfaces for four-component displacements in Table 3 at 2900 psia and 160°F.

Figure 13—The tie-line route parameters for the displacement of I1 by J1 in Table 3.

Once the tie-line routes with pressure are found it is easy to estimate the MMP. The MMP is the
pressure at which the smallest tie line in the tie-line route becomes zero length. Because we solve for all
tie lines, not just the key tie lines, the MMP is very accurate and can account for bifurcation of the
two-phase region. For the four-component displacement of Dutra et al. (2009) the MMP is equal to about
7,800 psia as shown in Figure 14. Figure 14 also shows the initial, injection, and crossover tie lines. This
displacement is combined condensing and vaporizing (CV) because the crossover tie line controls the
MMP (is the shortest tie line). Thus, the shock jump MOC method would give about the same MMP
because the K-values are strictly ordered.
18 SPE-175025-MS

Figure 14 —Key tie-line lengths calculated by analytical solution and shortest tie-line length calculated by simulation for the displace-
ment in Table 3.

We compared the analytical solutions to 1-D numerical simulation using the explicit formation in Johns
(1992). 10,000 grid blocks were used to minimize numerical dispersion. The shortest tie-line length at
each time was extrapolated to an infinite number of grid blocks (zero dispersion) using the method of Yan
et al. 2012 for up to 0.5 PVI. The tie lines and MMP values agree well with those from the tie-line MOC
solutions.
Last, we consider a four-component displacement in a system with bifurcating phase behavior that has
never been solved owing to its complexity. The phase behavior was created by adding nC4 to the
three-component system of Ahmadi et al. (2014) (see Table 1 parameters). Figure 15 gives the phase
behavior and MOC tie-line route for displacement of I3 (see Table 1 composition) by pure CO2. The
bifurcation of the two-phase zone is evident in the figure from the hour-glass shape. Figure 16 shows the
fast and slow paths for this phase behavior using the equilibrium mole fractions to describe tie-line space.
The outline of the tie-line space in Fig. 16 is unusual, but easily handled similarly to the last example. The
solution route in tie-line space consists of a slow rarefaction wave (R1) to crossover tie-line a (TL1),
followed by a fast rarefaction wave to tie-line 2 (TL2), where a shock to the initial composition occurs (S1).
Thus, along one ruled surface there is a combination of a rarefaction wave and a shock (n ⫽ 2 in Eq. (16)
on that surface). The shock velocity of S1 is equal to the fast eigenvalue at TL2. Tie-line 2 (TL2) is not a
key tie line, but controls the MMP as pressure is increased. We call that tie-line the “bifurcating” tie line
since it is the shortest tie line along that ruled surface.
SPE-175025-MS 19

Figure 15—Quaternary phase diagram with bifurcating phase behavior generated based on Ahmadi et al. (2011) at 8000 psia and 133°F.

Figure 16 —Four-component displacement of Figure 6 in tie-line space.

The MMP for the bifurcating four-component displacements is about 10,500 psia as shown in Figure
17. Figure 17 also gives the other key tie lines as well as the bifurcating tie line. As shown, the bifurcating
tie line controls miscibility. If the shock jump MOC method of Wang and Orr (1997) were used, the MMP
would be in error by about 4,000 psia (15,000 psia based on the initial tie line compared to 10,500 psia).
The simulation results corrected for dispersion agree well with the tie-line MOC route and MMP of 10,500
psia. The displacement is CV because the shortest tie line is the bifurcating tie line.
20 SPE-175025-MS

Figure 17—Four-component displacements with bifurcating phase behavior. The shock-Jump MOC over predicts MMP by almost 4000
psi

Example five-component displacement


There are three eigenvalues in tie-line space for a five-component displacement, so the tie-line route can
be shown using quaternary diagrams or 3-D plots. Figure 18 gives the tie-line route for the five-component
displacement with compositions and PREOS parameters in Table 4. The injection gas consists of four
components, making this a difficult problem to solve in composition space using Eqs. (1), but not in
tie-line space. There are now three ruled surfaces, and the solution route in tie-line space consists of one
rarefaction and two shocks. There is a shock from the injection tie line to the first cross-over tie line,
followed by a second shock to the second cross-over tie line. The tie line route is completed with a
rarefaction to the initial tie line. The K-values are strictly ordered here, so n ⫽ 1 for all the ruled surfaces.

Figure 18 —Tie-line route in tie-line space for five-component displacement

Table 4 —Input properties for five-component MMP calculations


Properties Compositions

Tc (°F) Pc (psia) ␻ I1 J1 BIP

C1 ⫺116.59 667.8 0.008 0.2067 0.1200 C1 CO2 nC4 C6


CO2 87.89 1,071 0.225 0.1059 0.7590 0.100 0.000
nC4 305.69 550.7 0.193 0.2347 0.0600 0.027 0.126 0.000
C6 453.65 429.8 0.296 0.1100 0.0600 0.422 0.145 0.017
C10 611.16 305.8 0.575 0.3428 0.0000 0.420 0.094 0.000 0.000

Figure 19 shows the key tie-line lengths for displacements at different pressures, along with the
“dispersion-free” shortest tie-line length from fine-grid simulation. Here, the displacement is purely
SPE-175025-MS 21

vaporizing as evidenced by the initial tie line being the shortest tie line at higher pressures. The MMP is
equal to about 1,700 psia when the initial tie line becomes zero length. Figure 20 gives the values of the
parameters li along each ruled surface (either shock or rarefaction) as a function of pressure.

Figure 19 —Key tie-line lengths for five-component displacement. The shortest tie-line length is calculated by simulation.

Figure 20 —The tie-line route parameters at different pressures for five-component displacement.

Conclusions
We showed how to transform the hyperbolic equations to construct tie-line routes and avoid singularities.
The equations are split into two parts: a system of equations that are only a function of phase behavior,
and one equation that is also a function of fractional flow. The main conclusions are:
1. The tie lines in a composition route can be solved more simply using the method of characteristics
(MOC) in tie-line space than in regular composition space. The eigenvectors in tie-line space
describe the ruled surfaces in composition space, but reduce to curves in tie-line space. Further,
there is only one umbilic point in the entire tie-line space.
2. Solution routes in tie-line space are constructed for complex displacements that have not been
solved previously, including a four-component bifurcating displacement and five-component
22 SPE-175025-MS

displacement with a rarefaction wave.


3. The solutions are unique in tie-line space, and solutions previously difficult to rule out (such as
multiple tie-line extensions through the initial or injection composition) are now easily eliminated
using simple conditions.
4. The MMP is more accurately determined when calculated using the tie-line MOC approach
because shorter tie lines between key tie lines (bifurcating tie lines) could control miscibility. The
shock jump method for MMP calculation can have significant error in such cases (cases where the
two-phase region bifurcates).

Acknowledgments
The authors thank BP, Chevron, Denbury, KNPC, Maersk, OMV, and Shell for their financial support of
this research through the Enhanced Oil Recovery JIP in the EMS Energy Institute at The Pennsylvania
State University at University Park, PA. Russell T. Johns is Chair of the undergraduate Petroleum and
Natural Gas Engineering program and holds the Victor and Anna Mae Beghini Faculty Fellowship in
Petroleum and Natural Gas Engineering at The Pennsylvania State University. The authors also thank Dr.
Wen Shen from the Mathematics Department at The Pennsylvania State University for the great
discussions and suggestions.

Nomenclature
C Overall volume fraction
C Overall Composition
EVC Equal velocity curve
F Overall Flux of component
f Funciton
f Fugacity
K K-value
M Derivative Matrix
N Number of moles matrix
n number of moles
Nc Numbe rof components
q Transformation matrix
R Characteristics matrix
t Time
v Composition in tie-line space
x Length
x Phase mole fraction
␣ Tie-line coefficient
␤ Tie-line coefficient
␥ Tie-line space parameter
⌫ Tie line
␭ Eigenvalue
⌳ Shock velocity
␾ Lagrangian coordinate
Ø Fugacity coeffcient
␺ Lagrangian coordinate

Superscripts
SPE-175025-MS 23

1,2 Phase number


e Envelop curve
r Reference tie line

Subscripts
D Dimnesionless
i component

References
Ahmadi, K., Johns, R.T., Mogensen, K., Noman, R., 2011. Limitations of current method-of-
characteristics (MOC) methods using shock-jump approximations to predict MMPs for complex
gas/oil displacements, SPE Journal, pp. 743–750.
Ahmadi, K., and Johns, R.T., 2011. Multiple-mixing-cell method for MMP calculations, SPE Journal,
6(4), pp. 733–742.
Bedrikovetsky, P., Chumak, M., 1992. Riemann problem for two-phase four-and more component
displacement (Ideal Mixtures). In 3rd European Conference on the Mathematics of Oil Recovery.
Bianchini S. and Bressan, A., Vanishing viscosity solutions to nonlinear hyperbolic systems, Annals
of Mathematics, 161, pp. 223–342, 2005.
Buckley, S.E. and Leverett, M.C., 1942. Mechanism of fluid displacement in sands, Trans., AIME,
146, pp. 107–116.
Christiansen, R.L., Haines, H.K., 1987. Rapid measurement of minimum miscibility pressure with the
rising-bubble apparatus. SPE Reservoir Engineering, 2(04), 523–527.
Dindoruk, B., 1992. Analytical theory of multiphase multicomponent displacement in porous media,
Department of petroleum engineering, Stanford, California, Stanford University.
Dutra, T.A., Pires, A.P., Bedrikovetsky, P.G. 2009. A new splitting scheme and existence of elliptic
region for gas flood modeling. SPE Journal, 14(01), 101–111.
Entov, V. M, 1997. Nonlinear waves in physicochemical hydrodynamics of enhanced oil recovery.
Multicomponent flows. International Conference on Porous Media: Physics, Models, Simulation,
Moscow.
Gelfand, I. M., 1963. Some problems in the theory of quasilinear equations. Amer. Math. Soc. Transl,
29(2), 295–381.
Gimse, T., Risebro, N.H., 1992. Solution of the Cauchy problem for a conservation law with
discontinuous flux function. SIAM J. Math. Anal. 23, pp. 635–648.
Helfferich, F.G., 1981. Theory of multicomponent, multiphase displacement in porous media, SPE
Journal, pp. 61–62.
Isaacson, E.L. 1989. Global solution of a Riemann problem for a non-strictly hyperbolic system of
conservation laws arising in enhanced oil recovery. Enhanced Oil Recovery Institute, University
of Wyoming.
Holden, H., Risebro, N.H. 2013. Front tracking for hyperbolic conservation laws (Vol. 152). Springer.
Jarrell, P.M., Fox, C.E., Stein, M.H., Webb, S.L. 2002. Practical aspects of CO2 flooding, SPE
Monograph Series. Society of Petroleum Engineers.
Jaubert, J.N., Avaullee, L., Pierre, C., 2002. Is it still necessary to measure the minimum miscibility
pressure?. Industrial & engineering chemistry research, 41(2), 303–310.
Jaubert, J.N., Wolff, L., Neau, E., Avaullee, L., 1998. A very simple multiple mixing cell calculation
to compute the minimum miscibility pressure whatever the displacement mechanism. Industrial &
engineering chemistry research, 37(12), 4854 –4859.
24 SPE-175025-MS

Jessen, K., Wang, Y., Ermakov, P., Zhu, J., Orr, F.M.Jr, 2001. Fast, approximate solutions for 1D
multicomponent gas-injection problems. SPE Journal, 6(04), 442–451.
Jessen, K., Michelsen, M. L., Stenby, E.H., 1998. Global approach for calculation of minimum
miscibility pressure. Fluid Phase Equilibria, 153(2), 251–263.
Jessen, K., Orr, F.M. 2008. On interfacial-tension measurements to estimate minimum miscibility
pressures. SPE Reservoir Evaluation & Engineering, 11(05), 933–939.
Johns, R.T., 1992. Analytical theory of multicomponent gas drives with two-phase mass transfer, PhD
dissertation, Department of petroleum engineering, Stanford, California, Stanford University.
Johns, R.T., Dindoruk, B., Orr, F.M., 1993. Analytical theory of combined condensing/vaporizing gas
drive, SPE Advanced Technology Series, Vol. 1, No. 2.
Johns, R.T., Orr, F.M.Jr., 1996. Miscible gas displacement of multicomponent oils, SPE Journal, 1
(1), pp. 39 –50.
Johns, R.T., Sah, P., and Solano, R., 2002. Effect of dispersion on local displacement efficiency for
multicomponent enriched-gas floods above the minimum miscibility enrichment. SPE Res Eval &
Eng. 5(1): 4 –10.
Johansen, T., Winther, R., 1989. The Riemann problem for multicomponent polymer flooding. SIAM
Journal on Mathematical Analysis, 20(4), 908 –929.
Johansen, T., Wang, Y., Orr, F.M.Jr, Dindoruk, B., 2005. Four-component gas/oil displacements in
one dimension: part I: global triangular structure. Transport in porous media, 61(1), 59 –76.
Juanes, R., 2008. A robust negative flash based on a parameterization of the tie-line field. Fluid Phase
Equilibria, 267(1), 6 –17.
Juanes, R., Lie, K.A., 2008. Numerical modeling of multiphase first-contact miscible flows. Part2.
Front-tracking/streamline simulation. Transport in Porous Media, 72, 97(120).
Khorsandi, S., Ahmadi, K., Johns, R.T., 2014. Analytical solutions for gas displacements with
bifurcating phase behavior. SPE Journal, 19(05), 943–955.
Khorsandi, S., Shen, W., Johns, R.T., 2015. Global Riemann solver and front tracking approximation
of three-component gas floods, verbal communication.
LaForce, T.C., Johns, R.T., 2005. Effect of quasi-piston-like flow on miscible gas flood recovery. In
SPE Western Regional Meeting. Society of Petroleum Engineers.
Lake, L.W., Johns, R.T., Rossen, W.R., Pope, G.A., 2014. Fundamentals of enhanced oil recovery.
Society of Petroleum Engineers.
Lax, P. D., 1957. Hyperbolic systems of conservation laws II, Comm. Pure Appl. Math. 10, pp.
537–566.
Li, Y., Johns, R.T., 2007. A rapid and robust method to replace Rachford-Rice in flash calculations.
In SPE Reservoir Simulation Symposium. Society of Petroleum Engineers.
Li, Y., Johns, R.T., Ahmadi, K., 2012. A rapid and robust alternative to Rachford–Rice in flash
calculations. Fluid Phase Equilibria, 316, 85–97.
Li, L., Khorsandi, S., Johns, R.T., Dilmore, R., 2014a. Reduced-order model for CO2 enhanced oil
recovery and storage using a gravity-enhanced process. In SPE Annual Technical Conference and
Exhibition. Society of Petroleum Engineers.
Li, L., Khorsandi, S., Johns, R.T., Ahmadi, K. 2014b. Multiple Mixing Cell Method for Three-
Hydrocarbon-Phase Displacements. In SPE Improved Oil Recovery Symposium. Society of
Petroleum Engineers.
Liu, T. P., 1976. The entropy condition and the admissibility of shocks, J. Math. Anal. Appl. 53, pp.
78 –88.
Mallison, B.T., Gerritsen, M.G., Jessen, K., Orr, F.M., 2005. High order upwind schemes for
two-phase, multicomponent flow. SPE Journal, 10(03), 297–311.
SPE-175025-MS 25

Metcalfe, R.S., Fussell, D.D., Shelton, J.L., 1973. A multicell equilibrium separation model for the
study of multiple contact miscibility in rich-gas drives. Society of Petroleum Engineers Journal,
13(03), 147–155.
Mogensen, K., Hood, P., Lindeloff, N., Frank, S., Noman, R., 2009. Minimum miscibility pressure
investigation for a gas-injection EOR project in Al Shaheen field, offshore Qatar, SPE ATCE,
Louisiana.
Monroe, W.W., Silva, M.K., Larsen, L.L., Orr, F.M., 1990. Composition paths in four component
systems: Effect of dissolved methane on 1-D CO2 flood performance, SPERE, pp.423–432.
Oil and Gas Journal, Survey: Miscible CO2 continues to eclipses steam in US EOR production 2014
worldwide EOR survey, 112(4), April, 2014.
Okuno, R., Xu, Z., 2014. Efficient displacement of heavy oil by use of three hydrocarbon phases. SPE
Journal, 19(05), 956 –973.
Orr, F.M. and Jensen, C.M., Interpretation of pressure-composition phase diagrams for CO2/Crude-
Oil systems, SPE Journal, pp. 485–497, 1984.
Orr, F.M., Johns, R.T., Dindoruk, B., 1993. Development of miscibility in four-component CO2
floods, SPE Reservoir Engineering, 8 (2), pp. 135–142.
Orr, F.M., Jessen, K., 2007. An analysis of the vanishing interfacial tension technique for determi-
nation of minimum miscibility pressure. Fluid phase equilibria, 255(2), 99 –109.
Peng, D.Y. and Robinson, D.B., 1976. A new two-constant equation of state, Industrial and engi-
neering chemistry fundamentals, 5 (1), pp. 59 –64.
PennPVT toolkit, Gas Flooding Joint Industry Project, Director: Dr. Russell T. Johns, EMS Energy
Institute, The Pennsylvania State University, University Park, PA, 2013.
Pires, A.P., Bedrikovetsky, P.G., Shapiro, A.A., 2006. A splitting technique for analytical modelling
of two-phase multicomponent flow in porous media. Journal of Petroleum Science and Engineer-
ing, 54(67).
Pope, G.A., 1980. The application of fractional flow theory to enhanced oil recovery. Society of
Petroleum Engineers Journal, 20(03), 191–205.
Rao, D.N., 1997. A new technique of vanishing interfacial tension for miscibility determination. Fluid
phase equilibria, 139(1), 311–324.
Rezaveisi, M., Johns, R.T., Sepehrnoori, K. 2015. Application of multiple-mixing-cell method to
improve speed and robustness of compositional simulation. SPE Journal.
Seto, C.J., Orr, F.M.Jr, 2009. Analytical solutions for multicomponent, two-phase flow in porous
media with double contact discontinuities. Transport in porous media, 78(2), 161–183.
Sheng, J.J., 2015. Enhanced oil recovery in shale reservoirs by gas injection. Journal of Natural Gas
Science and Engineering, 22, 252–259.
Stalkup, F.I., 1987. Displacement behavior of the condensing/vaporizing gas drive process. In SPE
Annual Technical Conference and Exhibition. Society of Petroleum Engineers.
Voskov, D., Tchelepi, H.A., 2008. Compositional space parametrization for miscible displacement
simulation. Transport in Porous Media, 75(1), 111–128.
Wang, Y., Orr, F.M., 1997. Analytical calculation of minimum miscibility pressure. Fluid Phase
Equilibria, 139(1), 101–124.
Yan, W., Michelsen, M.L., Stenby, E. H., 2012. Calculation of minimum miscibility pressure using
fast slimtube simulation. In SPE Improved Oil Recovery Symposium. Society of Petroleum
Engineers.
Yellig, W.F. and Metcalfe, R.S., 1980. Determination and prediction of CO2 minimum miscibility
pressures, JPT, pp. 160 –168.
Yuan, H., Johns, R.T., 2005. Simplified method for calculation of minimum miscibility pressure or
enrichment. SPE Journal, 10(04), 416 –425.
26 SPE-175025-MS

Zhao, G., Adidharma, H., Towler, B.F., Radosz, M., 2006. Minimum miscibility pressure prediction
using statistical associating fluid theory: Two-and three-phase systems. In SPE Annual Technical
Conference and Exhibition. Society of Petroleum Engineers.
Zhou, D., Orr, F.M.Jr, 1998. An analysis of rising bubble experiments to determine minimum
miscibility pressures. SPE Journal, 3(01), 19 –25.
Zick, A. A, 1986. A combined condensing/vaporizing mechanism in the displacement of oil by
enriched gas, SPE ATCE, New Orleans.
SPE-175025-MS 27

Appendix A
Tie-line space

There must be a one-to-one relationship between each tie line in tie-line space and composition space for the transformation
to be useful. First we discuss the thermodynamic definition of the tie-line space. Next we use the inverse function theorem to
define necessary conditions for one-to-one correspondence in tie-line space. Last, we discuss a simple transformation to show
that parameters that describe tie-line space are conservative.
Equilibrium is achieved for equality of the component fugacities at the same temperature and pressure. Fugacity is a
function of temperature, pressure and phase composition. Therefore, at equilibrium the tie lines satisfy,
(A.1)

where the superscript indicates the phase. There are 2Nc unknown compositional variables and Nc⫹ 2 equations in Eqs.
(A.1). The system of equations is completed by adding Nc ⫺ 2 definitions of tie-line space parameters, ␥i,
(A.3)

A simple form of gi can be used such as (Voskov and Tchelepi, 2008), or , or , or ␥i


⫽ ␤i (see Eqs. 6). In a well-defined tie-line space, each tie line is determined by a unique ⌫.
(A.4)

The corresponding tie line for a given ⌫ can be calculated by solving Eqs. (A.1) and (A.3). C1 should also be specified to
calculate the moles of a phase. That is,
(A.5)

Therefore, we can calculate molar properties of the reservoir fluid by specifying the vector v.
(A.6)

Typically in reservoir engineering, overall composition C is used to calculate molar properties of a fluid at constant
temperature and pressure.
(A.7)

We need to show that there is one-to-one relationship between all physical values of C and v and also that ⌫ is constant
along a tie line. If these conditions are true, we can conclude that ⌫ is well defined.
To demonstrate the above conditions, we define an implicit function f that maps composition space to tie-line space. The
exact form of f is not necessary to determine if the transformation is one-to-one, but we just need to show that f is invertible.
Thus, the mapping is expressed by,
(A.8)

A tie-line definition is valid, if (1) f, defined below, is a one-to-one function, (2) d⌫ is zero along each tie line and (3) x1
⫽ y1 for all tie lines. Based on the inverse function theorem, an implicit function is one-to-one if df is locally invertible
everywhere in the domain of the function.
(A.9)

We can remove the first row and column of df because those elements are either 1.0 or 0.0, which means where
is defined as
28 SPE-175025-MS

(A.10)

where A’ and B’ are derivatives of the tie-line coefficients in tie-line space as calculated in Appendix B. The condition
number of can be used to determine if is invertible (condition number must be finite for invertibility). The inverse
function theorem requires checking the condition number of for all tie lines, which is not practical. However, we can check
the criterion at a limited number of tie lines. For example, we can generate a series of tie lines using the mixing cell method
of Ahmadi et al. (2011) with only a few contacts, and find the best choice of ⌫, i.e. the choice for the tie-simplexes in Eqs.
(A-3) that give the smallest average condition number for .
Tie-line space as defined above can be used to study phase behavior. One more condition, however, should be checked
before using the tie-line space definition for analytical solutions using the method of characteristics because the weak solution
of hyperbolic systems can change for a nonlinear change of variables (Gelfan, 1963). We define a conservative tie-line space
such that the weak solution in tie-line space is the same as the weak solution of Eqs. (1). The transformation should satisfy
, where the ␤i are the conserved quantity in Eqs. (3). We define tie-line space in such a way that ␤i provides
a one-to-one mapping from composition space to tie-line space. The ␤ values are unique for each tie line if the envelope curves
of the tie lines do not intersect the hyperplane of . Therefore, we can define a new composition such that the
hyperplane of lies inside the two-phase region. This would ensure that the envelope curves do not intersect the
hyperplane because intersection of tie lines inside the two-phase region is unphysical. We can transfer the composition space
linearly and define a new set of conserved quantities.
(A.11)

where the matrix qij should have an inverse. Then, tie lines can be calculated uniquely in the new composition space.
SPE-175025-MS 29

Appendix B
Derivatives in tie-line space

The analytical derivatives of tie-line coefficients and K-values are necessary to calculate the characteristic matrix R (Eqs. 8)
in tie line space. We also give the simplified derivatives when K-values are independent of compositions.
The differential of component moles in each phase with ⌫ is the key to calculate tie-line derivatives, i.e., . First we need

to rewrite the thermodynamic equilibrium condition in difference form. Two phases at thermodynamic equilibrium should
satisfy the following condition for all components.
(B.1)

where is the fugacity of component i in phase 1. Consider two tie lines a and b, where for each tie line Eqs. (B.1) is
correct. Thus,
(B.2)

where is for phase 1 along tie-line a. As the two tie lines approach each other, we obtain Eqs. (B.3) using the
logarithm of fugacity.
(B.3)

where,
(B.4)

The derivatives of fugacity can be calculated based on derivatives of fugacity coefficients as follows.
(B.5)

with,
(B.6)

where Øi is the fugacity coefficient of component i. The number of moles of each phase do not affect equilibrium
compositions. Therefore we can set the number of moles of phases to be equal to 1.0 and the moles of components in each phase
are the same as the mole fraction of that component in a given phase.
(B.7)

Finally the Nc-2 derivaitves of ⌫ are replaced by.


(B.8)

Therefore we can rewrite Eqs. (B.3, B.7, and B.8) as a linear system of equations with 2Nc component moles for two phases
as,
(B.9)

M is a 2Nc by 2Nc matrix containing the derivatives, , and is a 2Nc vector, whose first Nc⫹2
elements are zero and the last Nc-2 elements are filled with elements of ⌫. We can solve the system of Eqs. (B.9) to determine
using,
(B.10)

where the elements of M-1 are


30 SPE-175025-MS

(B.11)

Similar calculations can be performed for constant K-value by solving Eqs. (B.7) and (B.8) without Eqs. (B.3). can be
replaces by , and can be replaced by . The matrix M is a Nc by Nc which can be used to calculate derivatives of
tie-line coefficients. Next, the derivatives of K-values in tie-line space are derived. K-values are constant along each tie line
so that we can define K-values as a function of ⌫.
(B.12)

The derivatives of K-values with respect to ⌫ should then be calculated by,


(B.13)

Further, the derivatives of the fugacity coefficient can be calculated using the chain rule and matrix M-1. That is,
(B.14)

Finally, the derivatives of tie-line coefficients can be calculated using the chain rule with Eqs. (6).
(B.15)

and,
(B.16)

where and . For constant K-values, Ki,j is zero.


To summarize, the tie-line derivatives can be calculated using the following steps:
1. Calculate fugacity coefficients and their derivatives using the desired equation of state with Eqs. (B.5) and (B.6).
2. Solve Eqs. (B.3), (B.7) and (B.9) simultaneously to calculate M-1.
3. Calculate the derivatives of fugacity coefficients with respect to ⌫.
4. Calculate Ki,j using Eqs. (B.13).
5. A’, B’ can be calculated by Eqs. (B.15) and (B.16), where and .
The matrix R in Eqs. (8) is equal to A’ when ϒi⫽␤i.
The matrixes A’ and B’ can be used to improve initial estimates for flash calculations. Usually the result of a previous flash
is used as an initial estimate. First, the change of ⌫ can be calculated by solving the following equations:
(B.17)

where the superscript r indicates the values for the known tie line or initial estimate. The calculated ⌬⌫ can be used to
update K-values using the derivatives of K-values calculated by Eqs. (B.13).
SPE-175025-MS 31

Appendix C
Constant K-value displacement

We show for any Nc that the ruled surfaces defined by the eigenvectors are planar if K-values are not composition dependent.
Dindoruk (1992) and Johansen et al. (2005) showed the same result for four-component displacements. First we show that
shock loci are planar surfaces in hyperspace, then we demonstrate that ruled surfaces generated from the eigenvalues coincide
with the shock loci and are therefore planar as well. The shock loci are given by (Rankine–Hugoniot) condition.
(C.1)

The term x 1 (x 1 ⫹ ⌬ x 1 ) is common in Eqs. (C.1). Therefore we define the modified “shock velocity”
and rearrange Eqs. (C.1) in a form similar to an eigenvalue problem, where ⌬xi are solved.
(C.2)

The coefficient matrix in Eqs. (C.2) is not a function of ⌬xi, which means that shock loci are straight lines in tie-line space.
In addition, the binodal curves are planes for phase behaviors with constant K-values. The tie lines of a shock locus intersect
two straight lines, the upstream tie line of the shock and the shock locus on the binodal surface. Therefore, the shock loci are
planes in composition space.
The eigenvectors can be calculated by the limit of Eqs. (C.2) as ⌬xi goes to zero. Because shock loci lie in a hyperplane
and an eigenvector is always tangent to the shock locus, the eigenvector is constant along that shock locus. Therefore, the ruled
surfaces given by the eigenvectors are also planar and coincide with the planes of the shock loci.

You might also like