You are on page 1of 14

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/282649718

Ensemble-Based Optimization of the WaterAlternating-Gas-Injection Process


ARTICLE in SPE JOURNAL OCTOBER 2015
Impact Factor: 1.31 DOI: 10.2118/173217-PA

READS

102

2 AUTHORS:
Bailian Chen

Albert Reynolds

University of Tulsa

University of Tulsa

5 PUBLICATIONS 20 CITATIONS

167 PUBLICATIONS 2,468 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

SEE PROFILE

Available from: Bailian Chen


Retrieved on: 11 January 2016

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

Stage:

Page: 1

Total Pages: 13

Ensemble-Based Optimization of the


Water-Alternating-Gas-Injection Process
Bailian Chen and Albert C. Reynolds, University of Tulsa

Summary
CO2-water-alternating-gas (CO2-WAG) flooding generally leads to
higher recovery than either continuous CO2 flooding or waterflooding. Although CO2 injection increases microscopic displacement
efficiency, unless complete miscibility is achieved, suboptimal
sweep efficiency may be obtained because of gravity segregation
and the channeling of CO2 through high-permeability zones or by
viscous fingering. Alternating water injection with CO2 injection
results in better mobility control and increases sweep efficiency.
Water injection also increases pressure that promotes miscibility.
However, poorly designed WAG parameters can result in suboptimal WAG performance. In this work, given the number of WAG
cycles and the duration of each WAG cycle, we apply a modification of a standard ensemble-based optimization technique to estimate the optimal well controls that maximize life-cycle net present
value (NPV). By optimizing the well controls, we implicitly optimize the WAG ratio (volume of water injected divided by the volume of gas injected). We apply the optimization methodology to a
synthetic, channelized reservoir. The performances of optimized
WAG flooding, optimized waterflooding, and optimized continuous
CO2 flooding are compared. Because of the similarity between
WAG and surfactant alternating gas (SAG foam), we also optimize
the SAG process and provide a more computationally efficient way
to optimize the SAG process with the optimal well controls
obtained from WAG as the initial guesses for the optimal controls
for SAG.
Introduction
WAG injection is an enhanced-oil-recovery method designed to
improve sweep efficiency during gas injection with the injected
water to control the mobility of gas and to stabilize the gas front
(Christensen et al. 2001). The focus of this paper is CO2-WAG
injection, which is a cyclic method of injecting alternating cycles
of CO2 followed by water and then repeating this process during a
number of cycles. Field applications of WAG were successful for
a variety of rock types, but were mainly applied to high-permeability reservoirs (Christensen et al. 2001) although there exist
simulation studies that suggest WAG can yield on the order of 5%
greater oil recovery than is obtained with continuous gas injection
even for low-permeability highly heterogeneous reservoirs (Liao
et al. 2013).
Poor recovery from the WAG technique can be caused by
inappropriately designed WAG parameters (e.g., the WAG ratio,
the number of cycles, and gas- or water-injection rate). Thus, the
optimization of WAG injection is widely recognized as a viable
technique for controlling mobility of gas and the miscible process
to achieve a better recovery (Chen et al. 2010a; Bahagio 2013).
Life-cycle production optimization is an optimal well-control
problem in which the objective is to determine well-operating
conditions that maximize NPV or cumulative oil production during the presumed reservoir life. The previous studies on production optimization mainly focus on optimizing the reservoir
performance under waterflooding (Brouwer and Jansen 2004;
Sarma et al. 2005; Alhuthali et al. 2007; Kraaijevanger et al.
C 2015 Society of Petroleum Engineers
Copyright V

This paper (SPE 173217) was accepted for presentation at the SPE Reservoir Simulation
Symposium, Houston, 2325 February 2015, and revised for publication. Original manuscript
received for review 25 January 2015. Revised manuscript received for review 25 May 2015.
Paper peer approved 19 August 2015.

2007; Jansen et al. 2009; Wang et al. 2009; Chen et al. 2010,
2012; Oliveira and Reynolds 2014); there are few studies on
WAG optimization. The existing studies on WAG optimization
are based, in most cases, on an experimental approach (Amin
et al. 2012; Srivastava and Mahli 2012; Shafian et al. 2013) or
trial-and-error reservoir-simulation work (Pritchard and Nieman
1992; Johns et al. 2003; Mirkalaei et al. 2011; Zhou et al. 2012;
Ghaderi and Clarkson 2012; Bender and Yilmaz 2014). However,
the well-control parameters determined by an experimental
approach or reservoir-simulation experiments may not be close to
an optimal solution. Thus, it is important to adaptively estimate
the optimal WAG ratio, well-injection rates, and optimal well
controls for producers in an automatic way to achieve a better recovery or NPV.
Esmaiel et al. (2005) and Esmaiel (2007) applied experimental
design and a related response-surface model to optimize the
WAG-injection process under geological uncertainty. The optimized recovery was more than double the base recovery. However, the optimal WAG parameters determined remain constant
throughout the simulation. Chen et al. (2010) presented a hybrid
technique, which integrates the orthogonal array and Tabu technique into the genetic algorithm, to determine the optimum WAG
production/injection parameters. However, the possible values of
the control variables are preselected and discrete, and the authors
did not provide information about how they chose the values of
control variables for each injector and producer or the WAG ratio.
It is difficult to know whether the WAG production/injection parameters obtained with their hybrid technique are optimal.
Bahagio (2013) used an ensemble optimization method to optimize the injection process of CO2-WAG in which the length of
each injection cycle is fixed to 1 year (i.e., each gas-injection period lasts 6 months, and each water-injection period lasts 6
months). Moreover, the WAG ratio for the case the author studied
was fixed to a constant value that is not feasible in practice because
of the reservoir heterogeneity, miscibility conditions, injection
conditions, and well operational parameters. In addition, the
author did not estimate the optimal well controls for production
wells, only for injection wells. Although the NPV obtained by the
optimized strategy was higher than that obtained by the base-case
strategy in the study of Bahagio (2013), we believe that one can
achieve better results if more control variables (e.g., well controls
for production wells) are taken into consideration. In an appendix
to Bahagio (2013), the author considered optimizing bottomhole
pressures (BHPs) at injectors together with the switching times for
a simple five-spot example with four injectors and one producer. A
switching time refers to a time at which we change from CO2
injection to water injection or vice versa. Although the addition of
switching times to the set of design variables seems to be appealing, we conjecture that, for the procedure we use here, the incorporation of switching times into the list of optimization variables
would not lead to a significant increase in the optimal NPV. The
issue of switching times will be revisited after we present computational results. We also note that for polymer flooding, Raniolo
et al. (2013) considered optimizing the polymer concentration and
time length of injection of three polymer slugs. Only a single injection well was considered and the BHP at the injection well was
specified and was not an optimization variable. Thus, it is unclear
whether one could have optimized a higher NPV with both BHPs
and injection times as optimization variables.
Here, we apply an ensemble-based optimization (EnOpt) technique, which was successfully used in waterflooding optimization

2015 SPE Journal

1
ID: jaganm Time: 14:50 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

(Lorentzen et al. 2006; Chen et al. 2009; Stordal et al. 2014; Fonseca et al. 2015a, 2015b), to adaptively determine the optimal
well controls (for both injectors and producers) for each cycle
during the presumed reservoir lifetime, without preselecting any
control variables. The EnOpt formula applied is the one suggested
by Do and Reynolds (2013) which invokes fewer assumptions in
the underlying mathematical development than does the standard
Chen et al. (2009) formulation. By optimizing the well controls,
we implicitly optimize the WAG ratio. Because of the similarity
between WAG and SAG foam, we also can apply this optimization technique to the SAG process. However, the computational
time required to estimate the optimal well controls for SAG injection is computationally expensive if we start the optimization
directly from an initial guess that is far from optimal. The computational time required to run a SAG simulation is much longer
than the time required to run a WAG simulation. The longer computational time is a result of the incorporation of the chemical
reactions necessary to model the generation of foam as well as
foam collapse. We illustrate that it is possible to radically reduce
the computational costs of SAG optimization by setting the initial
guess for the optimal well controls for SAG equal to the optimal
well controls generated for the WAG process.
In this work, we first present the WAG optimization problem,
which is followed by the description of the optimization methodology used in this paper. Then, we present the results of CO2WAG optimization on synthetic channelized reservoir examples
and make a comparison between WAG, continuous CO2, and
waterflooding techniques. Thereafter, we provide a more-efficient
method to optimize the SAG process. Finally, comments and conclusions are provided.
Water-Alternating-Gas-Injection
Optimization Problem
In this work, the optimization problem we consider is how to estimate the well controls that maximize the NPV in a WAG
enhanced-oil-recovery process. NPV is defined by
8
hXN 

XNt < Dtn
P
n
n
r

q

c

q
J u
o
w
tn
o;j
w;j
n1 :
j1
1 b365
)
i
XNI 
 k1 cwi  qnwi;k cgi  qngi;k
;         1
where u is an Nu -dimensional column vector that contains all the
well controls at all wells during the productive lifetime; n denotes
the nth timestep of the reservoir simulation; Nt is the total number
of timesteps; the time at the end of the nth timestep is denoted by
tn ; Dtn is nth timestep size; b is the annual discount rate; Np and
NI denote the number of producers and injectors, respectively; ro
is the oil revenue, in USD/STB; cw , cwi , and cgi , respectively,
denote the disposal cost of produced water, the cost of water
injection, and the cost of gas injection, in units of USD/STB; qno;j
and qnw;j , respectively, denote the average oil-production rate and
the average water-production rate at the jth producer for the nth
timestep, in units of STB/D; qnwi;k and qngi;k , respectively, denote
the average water-injection rate and the average gas-injection rate
at the kth injector for the nth timestep, in units of STB/D. For the
examples considered in this work, we neglect the revenue from
hydrocarbon-gas production and the disposal cost of injected gas
(CO2) that is produced.
and uup
We consider only bound constraints and let ulow
i denote
i
the lower bound and upper bound for the ith control variable,
respectively. Then, we can express the WAG-optimization problem as
maximize J uu2RNu
subject to ulow
 u  uup
i ; i 1; 2; 3; ; Nu :
i

. . . . . 2

We apply a logarithm transformation (Gao and Reynolds


2006) to each element of the control vector. The ith component of
the transformed control vector is given by

Stage:


xi ln

Page: 2

Total Pages: 13


ui  ulow
i
; . . . . . . . . . . . . . . . . . . . . . . . . . 3
up
ui  ui

where ui denotes the ith component of u. Note that xi varies from


1 to 1, and thus, after applying log-transformation to the control variables, the original constrained optimization problem is transformed to an unconstrained problem. Even though the log-transform
converts the problem to an unconstrained problem, following Oliveira and Reynolds (2014), in the log-domain, the variables are still
truncated to the interval [7, 7] in the optimization process. Then,
the optimization is performed in terms of the transformed vector x
[i.e., the optimization problem becomes maximized J x ]. After the
x

optimal x is obtained, we apply the inverse log-transformation to


transform x back to u in the original space; that is,
ui

low
expxi uup
uup expxi ulow
i ui
i
i
: . . . . . . . 4
1 expxi
1 expxi

Optimization Methods Considered


In this paper, we apply a slight modification on the standard
EnOpt method (Lorentzen et al. 2006; Chen et al. 2009) to optimize the CO2-WAG process. In this section, we discuss the standard EnOpt method, the modified EnOpt method, and the
hierarchical multiscale optimization (Hi-MO) method (Oliveira
and Reynolds 2014), which will be used to optimize the continuous CO2 injection and continuous waterflooding processes to
compare the optimal NPVs obtained for these processes with the
optimal NPV for CO2-WAG injection.
EnOpt. Assume that J x is the NPV function that we wish to
maximize, where x is the vector related to the control vector u by
Eqs. 3 and 4. Here, J x is maximized with the optimization algorithm given by


dk
; . . . . . . . . . . . . . . . . . . . . . . 5
xk1 xk ak
kdk k1
for k 0; 1; until convergence, where x0 is the initial guess and
xk is the estimate of the optimal control vector at the kth iteration.
Here, at each iteration k, we use the simple notation x^ xk ; ak is
the step size; the search direction vector dk can be expressed as
 T  T
 T

T
dk dk1 ; dk2 ; ; dkm ; ; dknwell T ; . . . . . . . . 6
where nwell is the total number of wells. The vector dkm represents
the search direction for well m (i.e., dkm is the subvector of dk
that produces changes in the components of xk that correspond to
the controls of well m when Eq. 5 is applied). All vectors and subvectors without a transpose sign refer to column vectors. To enumerate the components of a subvector dkm , we use the notation
 

dkm dkm 1 ; dkm 2 ; ; dkm nc T ; . . . . . . . . . . . . . . . . 7
for m 1; 2; ; nwell , where nc is the number of control steps for
each well. Here, each well has the same number of controls (i.e.,
nc is the same for all the wells). To obtain the search direction dk
at iteration k 1, we first generate Ne samples of the Gaussian
random vector X, where X  N x^; Cx (i.e., the mean of X is equal
to x^ xk and its covariance matrix is CX ). These samples can be
generated from
x^j x^ LZj ; j 1; 2; ; Ne ; . . . . . . . . . . . . . . . . . 8
where L is the lower triangular factor in the Cholesky decomposition of CX and the components of Z j are independent, standard,
random-normal deviates; that is, Zj  N 0; INx . LZj is an Nxdimensional Gaussian random vector with mean equal to the Nxdimensional zero vector and covariance matrix, CX ; that is,
LZj  N 0; CX .

2015 SPE Journal


ID: jaganm Time: 14:50 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

The singly smoothed EnOpt search direction used in this work


is the one considered in Do and Reynolds (2013) that at iteration
k 1 is given by

1 XNe  j
x^  x^ J x^j  J x^: . . . . . . . . . . . . . 9
dk
j1
Ne
Do and Reynolds (2013) show that
dk  CX rx J x^; . . . . . . . . . . . . . . . . . . . . . . . . . . 10
where x^ xk . This search direction of Eq. 9 differs from the one
proposed by Chen et al.
(2009). They
use the mean of the samples
hX
 j i

Ne
J x^ =Ne in place of J x^ J xk .
in place of x^ and J
j1
When optimization is based on a single fixed reservoir model, Do
and Reynolds (2013) show that the difference between results
generated from the two basic EnOpt algorithms is negligible even
though the Do-Reynolds EnOpt formulation of Eq. 9 involves
fewer approximations than does the Chen et al. formula. However, the Chen et al. algorithm was developed for robust optimization in which one wishes to maximize the expected value of the
NPV of life-cycle productions over an ensemble of reservoir models. Recently, Fonseca et al. (2015b) generalized Eq. 9 to the robust optimization case and found that, with their new method, one
can obtain a significantly higher value of life-cycle NPV than is
obtained with the Chen et al. (2009) version of EnOpt in the robust optimization case.
In EnOpt (Chen et al. 2009), additional smoothing is typically
performed by multiplying CX , and if we do the same, the final
modified EnOpt search direction, denoted by gk , is given by
gk CX dk : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
When using steepest ascent with the true gradient replaced by a
stochastic gradient, the gradient may be quite rough. For the optimal well-control problem, this rough gradient can sometimes
result in controls at an individual well that vary rapidly in time and
may not be acceptable in practice. Equally importantly, the use of
a rough gradient at each iteration can result in a suboptimal estimate of the maximum achievable NPV although this is not always
the case (Fonseca et al. 2015a, 2015b). One way to control this is
to smooth the stochastic gradient by multiplying the stochastic gradient by a covariance matrix that provides smoothing. See Do and
Reynolds (2013), Yan and Reynolds (2014), and Fonseca et al.
(2015b) as well as the original work of Chen et al. (2009).
When dk of Eq. 9 is multiplied by CX and the resulting CX dk is
used in place of dk in Eq. 5, the resulting algorithm is referred to
as a modified
the resulting Eq. 5 gives a xk1 such
 k1 EnOpt
 k when
k1
>J x ,x
is accepted as the new approximation
that J x
to an x that maximizes NPV. If not, then the stepsize ak is repeatedly reduced by a factor of 2, and xk1 is recomputed with the
k1
such that J xk1 >
reduced
 k stepsize until Eq. 5 produces an x
J x or the maximum number of stepsize cuts has been reached.
If, with the maximum number of stepsize cuts, we cannot find an
xk1 that increases the NPV, then we generate a new set of perturbed controls to try to find an uphill search direction and, if necessary, repeat the stepsize cuts. If after NRes consecutive
resamplings of the perturbations to obtain a search direction, we
still cannot find an xk1 that increases J x after the maximum
number of allowable stepsize cuts, then the algorithm is terminated. Also, if both the following two equations hold,



J xk1  J xk
 eJ . . . . . . . . . . . . . . . . . . . . . . . 12
maxfJ xk ; 1:0g
and

kxk1  xk k2
 ex ; . . . . . . . . . . . . . . . . . . 13
maxfkxk k2 ; 1:0g

the algorithm is terminated. Following the discussion of Subsection 8.5.1 of Oliver et al. (2008), we use eJ 104 and ex 103 .
If total number of allowable simulation runs, NR, is exceeded, we
terminate the algorithm. Typically, we choose NR between 1,000

Stage:

Page: 3

Total Pages: 13

and 5,000, depending on the computational resources available.


For the examples considered here and similar other computations
not shown, increasing NR from 1,000 to 2,000 results in only a
very small increase in the NPV.
Adaptation of Modified EnOpt to WAG. To simulate WAG
with a standard reservoir simulator, we represent each actual injection well by two injection wells, a gas-injection well and a waterinjection well at the same location. The use of two wells is necessary because CMGs GEM, which is the simulator used to model
WAG, does not allow us to switch a well back and forth from a
water injector to a CO2 injector. Each WAG-injection cycle contains two half-cycles, a first half-cycle and a second half-cycle. Specifically, gas is injected during the first half-cycle and water is
injected during the second half-cycle. However, the water-injection
well is turned off (rate fixed at zero) during the first half-cycle of
each cycle, and the gas-injection well is turned off during the second half-cycle of each cycle. When an injection rate is fixed at zero
during a half-cycle, we do not want to include the rate during this
period as an optimization variable; thus, we fix both the rate and the
corresponding search direction equal to zero. As discussed next, the
use of two injection wells at the same location where each well
alternates from open to closed at the end of each half-cycle requires
only minor complications to our in-house optimization code.
Here, nwell represents the number of wells where each physical
injection well is modeled as two wells in the reservoir simulator, a
water-injection well and a gas-injection well. Each of these two
wells is represented by its own subvector in the form of Eq. 7, which
also can be expressed in terms of the number of WAG cycles as
 T  T  T  T
dkm dkm 1;a ; dkm 1;b ; dkm 2;a ; dkm 2;b ; ;
 m T
 T
dk ncycle;a ; dkm ncycle;b T ;                 14
where ncycle denotes the total number of WAG cycles and a and b
represent the first half-cycle and the second half-cycle of each
injection cycle, respectively. Note that each half-cycle may contain more than one control step. Here, for simplicity, we assume
all the half-cycles contain the same number of control steps, ncs ,
which means each entry of dkm , given in the form of Eq. 14, has
ncs subentries; for example, dkm 1;a can be expressed as
 m
 1  2
 ncs T
dk 1;a dkm 1;a ; dkm 1;a ; ; dkm 1;a
 : . . . . . . . . . . 15
We see that nc , the total number of control steps for each well,
is equal to 2ncs ncycle . To account for the fact that, for all the first
half-cycles, the water-injection rates and their corresponding search
directions are zero, and for all the second half-cycles, the gas-injection rates and their corresponding search directions are zero, we
make the modifications discussed next in Steps 1, 2, and 3.
we set the even entries of dkm , that is,
 mStep
1.
 mFor
gas injectors,
 m
dk 1;b , dk 2;b ,, dk ncycle;b in Eq. 14, equal to zero to eliminate
the entries that do not correspond to the design (optimization) variables. Similarly, for water injectors, we have to set the odd
m
in the form
entries
 m of dk mgiven

of Eq. 14 equal to zero, that is,
dk 1;a dk 2;a    dkm ncycle;a 0.
Step 2. Because we smooth the dk of Eq. 9 by multiplying by
the covariance matrix CX which is a block diagonal matrix to
obtain the search direction, the search direction we use in our
algorithm is first computed as
3
2 1 3 2
C1X dk1
dk
7
6
6
3 d 2 7 6 C2 d 2 7
2 1
7
X k
CX 0
0 6
7
6 k 7 6
7
6 . 7 6
.
7
6 0 C2
.
.
7
7
6
6
0
7
6
X
. 7 6
.
7
6
7
gk CX dk 6
76
7;
.. . .
.. 76
6 ..
m
m
m
7
7
6
6
d
C
d
5
4 .
X k
.
.
.
7
6 k 7 6
7 6
7
.
.
nwell 6
7
6 . 7 6
.
0
0 CX
.
5
4 . 5 4
nwell
nwell nwell
dk
CX dk
                   16

2015 SPE Journal

3
ID: jaganm Time: 14:50 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

where Cm
X , m 1; 2; ; nwell is the covariance matrix used to
force smoothness on the well controls for well m. In this work, we
use a spherical covariance function; that is, denoting the i; j
m
entry of Cm
X by Ci; j , we have
"



 #
3 ji  jj
1 ji  jj 3
m
2
Ci; j rm 1 

; . . . . . . . . 17
2 Nsm
2 Nsm
m
if ji  jj  Nsm , and Cm
i;j 0 if ji  jj > Ns . Here, i and j denote
the control step i and j, respectively; rm refers to the standard
deviation; and Nsm is the number of control steps over which we
wish the control at well m to be correlated. As discussed later, the
choice of Nsm at each refinement level of Hi-MO algorithm (Oliveira and Reynolds 2014) controls the degree of smoothing
imposed on the well controls. Also, see Wang et al. (2009) and
Yan and Reynolds (2014) for the effect that the choice of Nsm has
on the temporal smoothness of the estimated optimal controls.
m m
Then the mth component of gk , that is, gm
k  CX dk can be
expressed as
32 dm 3
2
C11 C12 C1nc
k
 1 7
6 C21 C22 C2n 76
6 dkm 2 7
c 7
6
7
m m
76
6
gm
k CX dk 6 .
..
..
.. 76
.. 7
7
6
5
4 ..
.
.
.
4 . 5
 m
Cnc 1 Cnc 2 Cnc nc
dk n
2
 m
 m
 m c 3
C11 dk 1 C12 dk 2    C1nc dk nc
6


 7
7
6 C21 dm C22 dm    C2n dm
c
k 1
k 2
k nc 7
6
7:
6
7
6
..
7
6
.
4
 m
 m
 m 5
Cnc 1 dk 1 Cnc 2 dk 2    Cnc nc dk nc
 m  m
 T
gk 1 ; gk 2 ; ; gm
              18
k nc  :

But gm
k also can be expressed in terms of the number of WAG
cycles as
gm
k
T

 m T  m T  m T  m T
 m T
 m T
gk 1;a ; gk 1;b ; gk 2;a ; gk 2;b ; ; gk ncycle;a ; gk ncycle;b :
                   19
Note that dkm and gm
k in Eq. 18 are expressed in the form of
Eq. 7, not in the form of Eq. 14, to show the matrix multiplication
by Cm
X more clearly.
Step 3. From Eqs. 16, 18, and 19, we see that a zero entry of
dkm in the form of Eq. 14 does not imply that the corresponding
entry of gk in the form of Eq. 19 is zero. Thus, for gas injectors,
we again set the even entries of the subvector gm
k in the form of
Eq. 19 equal to zero; for water injectors, we set the odd entries of
m
gm
k in the form of Eq. 19 equal to zero. This gives the final gk ,
and hence, the gk , which we use as the modified EnOpt search
direction. After the previous modifications, the EnOpt techniques
that were developed for life-cycle waterflooding optimization
can be adapted to optimize the design variables for the WAGinjection process.
Hi-MO Method. In this work, the performances of continuous
CO2 flooding and waterflooding are studied to compare their performance with that of the CO2-WAG technique. The algorithm
applied to optimize the process of continuous CO2 flooding and
waterflooding is the adaptive Hi-MO algorithm (Oliveira and
Reynolds 2014), which can yield a higher NPV than is obtained
by applying the standard procedure of EnOpt. The objectives of
the Hi-MO algorithm are to reduce the number of control variables used to generate a set of less-ill-conditioned optimization
problems, and to improve computational efficiency without compromising the optimal value of the life-cycle NPV obtained.
The Hi-MO method can be applied with any optimization algo-

Stage:

Page: 4

Total Pages: 13

rithm, but here, we only consider its application with the modified
EnOpt method.
In the multiscale approach, we start with a small number of
control steps and find the optimal controls by maximizing J x;
then, we select the new set of control steps for the next level of
parameterization that is based on two main procedures: merging
and splitting. The control steps are selected to be merged if the
following control-variation criterion holds:
juwi  uwi;ref j
 ulow;w
uup;w
i
i

< e u ; . . . . . . . . . . . . . . . . . . . . . . . . 20

where e u is the control-variation tolerance for merging well controls; uwi is the ith control at well w; uwi;ref is a reference value corresponding to the average of all the consecutive controls at well w
and
previous to the ith control which were merged; and uup;w
i
denote the upper bound and lower bound of ith control
ulow;w
i
at well w, respectively. Although the log-transform is used to handle the bound constraints, the control-variation criterion is checked
in the original control space instead of the log-transformed
control space.
All control steps that are not selected to be merged are split
into nsplit (2 or 4) new control steps of uniform length. Here, h
denotes the parameterization level; hmax represents the number of
refinement levels; xh is related to uh , which is the optimal control
vector at parameterization level h by Eqs. 3 and 4. The basic HiMO iteration loop is shown next:
For parameterization level h 0; 1; 2; ; hmax .
Step 1. Solve the optimization problem (EnOpt method).
Use of xh as the initial guess, ehJ and ehx as the convergence tolas the maximum number of
erances in Eqs. 12 and 13, and nmax;h
s
allowable simulation runs at Step 1 of level h obtains
xh1 argmaxJ x: . . . . . . . . . . . . . . . . . . . . . . . . 21
Step 2. Check convergence.
If both of the following equations hold,


J xh1 J xh
 e J . . . . . . . . . . . . . . . . . . . . . . 22
maxfJ xh ; 1:0g
and

kxh1  xh k2
 e x ; . . . . . . . . . . . . . . . . . . . 23
maxfkxh k2 ; 1:0g

then set xopt xh1 and terminate the algorithm. If the total allowmax
able simulation runs, nmax
s , are reached, that is, ns > ns , then set
xopt xh1 and terminate the optimization algorithm. Here e J and
e x are the final convergence tolerances for the change of objective
function and control variables, respectively; ns is the counter for
the number of reservoir simulation runs.
Step 3. Reparameterization: Redefine the control steps for
each well, w 1; 2; 3; ; nwell .
Perform the splitting and merging procedure: for each control
step i 1; 2; 3; ; nwc of each well, determine whether the ith
control should be either merged or split with Eq. 20, where nwc
denotes the number of control steps for well w.
Step 4. Update the covariance matrix.
Decrease the temporal correlation length used to define CX
between the control variables for the optimization step for the
next level by multiplying the current correlation length, T h , by a
decreasing factor b, where 0 < b < 1; that is,
T h1 maxfbT h ; Tmin g; . . . . . . . . . . . . . . . . . . . . . 24
where Tmin is the minimum correlation length which may be zero.
Then, generate the covariance matrix, Ch1
X , for the next level.
Step 5. Update convergence tolerances.
Tighten the tolerance for the optimization step for the next
level by applying the following equations:
maxfcehJ ; e J g . . . . . . . . . . . . . . . . . . . . . . . . 25
eh1
J

2015 SPE Journal


ID: jaganm Time: 14:50 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

25

Pro1

Pro2

Stage:

Page: 5

Total Pages: 13

Pro3

20
Inj1

Inj2

15
Pro4

Pro5

Pro6

10
Inj3

Inj4

5
Pro7
5

Pro8
10

15

Pro9
20

25

Fig. 1Log-permeability distribution, three-channel case.

eh1
maxfcehx ; e x g; . . . . . . . . . . . . . . . . . . . . . . . 26
x
where c is the tolerance-tightening factor. More total simulation
runs for the next refinement level can be allowed by multiplying
by a user-chosen factor greater than unity.
the total nmax;h
s
Then, set h h 1 and go to Step 1.
End (For)
We wish to start with very few controls and long control steps
because there exist cases in which one can obtain an optimal solution with a small number of control steps [e.g., case in which
bang-bang controls are optimal (Zandvliet et al. 2007)]. Thus, we
choose the two initial control steps of equal length. The choice of
nsplit 2 is based on the computational experiments of Oliveira
(2014), but choosing nsplit 4 is not expected to yield radically
different results. After nsplit is chosen, hmax is chosen so that
Dtmin

T
max
nhsplit

; . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

where T is the total production lifetime and Dtmin is the minimum


allowable length of a control step. For example, if T 2,048 days
and Dtmin 32 days, then Eq. 27 yields hmax 6. It is, of course,
possible for the algorithm to terminate before reaching the sixth
refinement.
Case Study
The reservoir model studied in this work is a three-layer synthetic
reservoir with a 25
25
3 grid with the gridblock dimensions
given by Dx Dy 100 ft and Dz 20 ft. It contains four injectors and nine producers. The horizontal log-permeability distribution of the first layer is shown in Fig. 1 with three highpermeability channels. The other two layers have the same horizontal permeability distribution as the first layer. The porosity is
homogeneous with / 0.2. The compressibility of rock and water
is equal to 6:103
105 and 3:3
106 psi1, respectively. The
initial reservoir pressure is 4,300 psi, and the minimum miscibility pressure is equal to 4,150 psi. The reservoir lifetime is set
equal to 2,048 days. A compositional reservoir simulator, GEM
(CMG 2009) from Computer Modelling Group, is used in reservoir simulation for water-alternating gas (WAG) injection (Chen
et al. 2010). There are seven pesudocomponents used to describe
the oil and gas: CO2, N2, C1, C2C4, C5C6, C7C18, and
C19C43.
The control variables at each gas injector are the injection rates
with an upper bound of 20 MM scf/D and a lower bound of 0 MM
scf/D; the control variables at each water injector are the injection
rates with an upper bound of 4,000 STB/D and a lower bound of 0
STB/D; each producer operates under BHP control with an upper

Table 1Summary, comparison among different WAG cycles.

bound of 4,500 psi and a lower bound of 1,500 psi. The initial
guess for rate controls of each water injector is set equal to 2,000
STB/D; the initial guess for rate controls of each gas injector is
equal to 10 MM scf/D, and the initial guess for BHP controls for
each producer is equal to 3,500 psi.
To optimize the NPV, the oil price is set equal to USD 80.0/
STB; the water-injection cost is USD 5.0/STB; the gas-injection
cost is USD 1.5/Mscf; the cost of disposing produced water is
USD 5.0/STB; the annual discount rate is 0.1. Here, we neglect
the revenue of hydrocarbon-gas production and the cost of disposing of breakthrough injection gas. For the optimization of the
SAG process, the cost of injecting water containing dissolved surfactant is set equal to USD 8.0/STB.
We apply the log-transformation to handle the bound constraints for all the optimization methods considered in this paper.
For the ensemble-based optimization (EnOpt) method, we set the
number of samples for gradient averaging to Ne 10; the maximum number of stepsize cuts is set equal to 5, and the maximum
number of resamples is specified as Nres 5; the initial step size
is 1.0; the time-correlation length is 512 days; the perturbation
size is equal to 1.0; the maximum number of allowable simulation
runs is 1,000. When the Hi-MO method is used to optimize the
processes of continuous CO2 flooding and waterflooding, we specify the maximum number of refinement levels as h 6; the minimum control-step length is 32 days; we use two initial control
steps, and the number of splits is given by nsplit 2; the control
tolerance for lumping steps is 0.01; the initial control-correlation
length is T 1 2; 048 with a decreasing factor of 2 for each optimization level; the tolerance-tightening factor is set equal to 0.5;
the maximum number of allowable simulation runs of Level 1 is
100, and this number increases by a factor
set equal to nmax;1
s
of 2 at each level; the total maximum allowable simulation runs
is 1,000.
Comparison Among Different WAG Cycles. To compare the
performance of WAG with different numbers of cycles, we set up
four cases with 4 cycles, 8 cycles, 16 cycles, and 32 cycles. The
number of control steps for each well, nc , is set equal to 64 regardless of the number of cycles, and the length of each control step is
32 days. When the total number of WAG cycles is equal to 4, we
set the length of each of the eight half-cycles equal to 256 days,
and each of the half-cycles contains 8 control steps. For 8 cycles,
the length of each half-cycle is equal to 128 days, and each of the
half-cycles contains 4 control steps, and so on. Table 1 presents
the summary of the initial NPV, final NPV (estimate of optimal
NPV), and cumulative oil production for the different cases. From
the results in Table 1, the final NPV in all the cases is increased
compared with the initial NPV. We see that the initial NPV, the
estimated optimal (final) NPV, and cumulative oil production
increase as the number of cycles increases (cycle time decreases),
but they are fairly close for the 16-cycle and 32-cycle cases.
Fig. 2 shows the remaining oil saturation as a function of the
number of cycles (or cycle time). The color scale corresponds to
the oil saturation, with the darkest blue representing the lowest oil
saturation and the darkest red color representing the highest oil
saturation. For the same layer (Layer 1 or Layer 3 in Fig. 2), as
the number of cycles increases, the remaining oil saturation
decreases. Note that, regardless of the number of cycles, the upper
layer has less remaining oil than the bottom layer because the
gravity segregation of injected water and gas causes CO2, which
has a higher displacement efficiency, to move to the upper layer

2015 SPE Journal

5
ID: jaganm Time: 14:50 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

Layer 1

Layer 3

Stage:

Page: 6

Total Pages: 13

0.80

0.80

0.80

0.80

0.70

0.70

0.70

0.70

0.60

0.60

0.60

0.60

0.50

0.50

0.50

0.50

0.40

0.40

0.40

0.40

0.30

0.30

0.30

0.30

0.20

0.20

0.20

0.20

0.10

0.10

0.10

0.10

0.00

0.00

0.00

0.00

0.80

0.80

0.80

0.80

0.70

0.70

0.70

0.70

0.60

0.60

0.60

0.60

0.50

0.50

0.50

0.50

0.40

0.40

0.40

0.40

0.30

0.30

0.30

0.30

0.20

0.20

0.20

0.20

0.10

0.10

0.10

0.10

0.00

0.00

0.00

(a) 4 cycles

(b) 8 cycles

0.00

(c) 16 cycles

(d) 32 cycles

Fig. 2Remaining oil-saturation distribution of different cycle cases.

whereas water, which has a lower displacement efficiency, moves


to the bottom layer.
One can obtain the estimated optimal controls for all the production wells and injection wells by applying the modified EnOpt
algorithm. Fig. 3 presents the well controls of producers for the
case with 32 WAG cycles. As one can see in Fig. 1, Producer
Pro1 is separated from injectors by one of the three high-permeability channels, which we refer to here as Channel 1. When the
injected water or gas reaches Channel 1, most of the injected fluid
will move toward Pro2 through this channel. Thus, it is reasonable
to expect Pro1 to operate at close-to-minimum allowable BHP
(1,500 psi) throughout the production lifetime, but Pro2 can operate at close-to-minimum allowable BHP only at the early times

Pro1

when it is still economic but shut in from 500 to 1,700 days, as


shown in Fig. 3a. Pro3, Pro4, Pro5, and Pro6 are either in a channel or close to a channel, and all operate at close-to-minimum
allowable BHP at the early times, and thereafter, they are almost
shut in, as shown in Figs. 3b and 3c. Pro7 is shut in at the early
times and thereafter operates at close-to-minimum allowable
BHP. Pro8 is close to the high-permeability channel so that the
estimated control for this producer is close-to-minimum allowable
BHP at the early control steps, but after injected gas and water
breakthrough at this producer, its pressure increases (rate
decreases), which occurs at approximately 1,000 days. Pro9 is far
from the channel and the injectors, so it operates at close-to-minimum allowable BHP during the whole production lifetime, as

Pro2

Pro3

3,500

2,500

1,500

3,500

2,500

1,500
0

500

1,000

1,500

2,000

500

Time (days)

Pro5

1,000

1,500

2,000

Time (days)

(a) Pro1 and Pro2

(b) Pro3 and Pro4

Pro6

Pro7

4,500

Pro8

Pro9

4,500

BHP (psi)

BHP (psi)

Pro4

4,500

BHP (psi)

BHP (psi)

4,500

3,500

2,500

1,500

3,500

2,500

1,500
0

500

1,000

Time (days)
(c) Pro5 and Pro6

1,500

2,000

500

1,000

1,500

2,000

Time (days)
(d) Pro7, Pro8, and Pro9

Fig. 3Estimated optimal well controls for production wells, 32 WAG cycles.
6

2015 SPE Journal


ID: jaganm Time: 14:50 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

15

3,000

10

2,000

1,000

0
4

12

16

20

24

28

Total Pages: 13

32

Water injection

20

4,000

15

3,000

10

2,000

1,000

0
0

12

3,000

10

2,000

1,000

0
8

12

16

20

24

28

Gas injection

Gas-Injection Rate
(MMscf/D)

15

Water-Injection Rate
(STB/D)

Gas-Injection Rate
(MMscf/D)

Water injection
4,000

24

28

32

(b) Inj 2

20

20

Cycle

Cycle
(a) Inj 1
Gas injection

16

4,000

15

3,000

10

2,000

1,000

0
0

32

Water injection

20

12

16

20

24

28

Water-Injection Rate
(STB/D)

Page: 7

Gas injection
4,000

Gas-Injection Rate
(MMscf/D)

Water injection

20

Water-Injection Rate
(STB/D)

Gas-Injection Rate
(MMscf/D)

Gas injection

Stage:

Water-Injection Rate
(STB/D)

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

32

Cycle

Cycle
(c) Inj 3

(d) Inj 4

Fig. 4Estimated optimal well controls for injection wells, 32 WAG cycles; the rates plotted correspond to the half-cycle rates for
the different cycles.

shown in Fig. 3d. The estimated optimal well controls for injection wells for the case with 32 WAG cycles are shown in Fig. 4.
Note that the rates plotted correspond to the half-cycle rates for
the different cycles. Various factors affect the gas- and waterinjection rates, including the miscibility between CO2 and oil, the
channeling of injected fluid, and the shut-in of some production
wells during the production lifetime, all which lead to the optimal
well-injection rates, as shown in Fig. 4.
Fig. 5 shows the estimated optimal WAG ratio (volume of
water injected divided by the volume of gas injected) for each
cycle for the case with 32 cycles. The WAG ratio is affected by
reservoir heterogeneity, miscibility conditions, injection conditions, and well operational parameters, so it is hard to explain
why the estimated WAG ratio changes with the WAG cycle, as
shown in Fig. 5. However, one can obtain some general ideas
from the results of this figure. First of all, the WAG ratio starts
with a relatively high value at the early injection times to inject
more water into the reservoir to block the high-permeability zone
and to increase gas miscibility. Thereafter, the WAG ratio
decreases to a lower value to have more gas injected to obtain a
better microscopic displacement of oil. Finally, the WAG ratio
increases at the end of the reservoir life. From the previous analysis, we can see that it is inappropriate to fix the WAG ratio for
WAG-injection process in practice although, for most cycles, the
WAG ratio is less than 0.2. Note that we implicitly optimize the
WAG ratio for different cycles by optimizing the well controls.

Inj1

1.2

Inj2

Inj3

Inj4

WAG Ratio

1
0.8
0.6

Comparison of WAG, Continuous CO2 Injection, and


Continuous Waterflooding. To compare the efficiency of CO2WAG, continuous CO2, and continuous waterflooding, we consider the following three subcases, all which pertain to the threelayer reservoir.
1. WAG with 32 cycles by use of modified EnOpt for
optimization
2. Continuous CO2 by use of Hi-MO (two initial control steps;
nsplit 2) for optimization
3. Waterflooding by use of Hi-MO (two initial control steps;
nsplit 2) for optimization
For continuous CO2 and waterflooding, we apply the Hi-MO
algorithm (Oliveira and Reynolds 2013) with two initial control
steps and two splits to optimize the processes and to obtain the
final NPV and cumulative oil produced. Because Hi-MO adaptively selects the duration and length of the timesteps and is
designed to mitigate the effects of ill-conditioning, Hi-MO is generally able to provide a higher NPV than is obtained by fixing the
number and length of the control steps a priori. We have not
developed a version of the Hi-MO algorithm applicable for the
WAG process, but it is likely that such an algorithm would yield
higher optimized values of NPV than the ones given in this paper
for the WAG process.
Table 2 presents the summary of the initial NPV, final NPV,
and cumulative oil produced for the three techniques. From the
results in Table 2, all the final NPVs are improved compared with
the initial NPVs. The optimized WAG process yields significantly
more oil production and a higher NPV than those obtained by
optimizing continuous CO2 or waterflooding.
Figs. 6 and 7 display the CO2 concentration in the bottom
layer (Layer 3), and the pressure distribution for the middle layer
(Layer 2), respectively. These two figures compare results from
WAG flooding with those from continuous CO2 flooding after

0.4
0.2
0
0

12

16

20

24

28

32

Cycle
Fig. 5Estimated optimal WAG ratio of each cycle, 32 WAG
cycles.

Table 2Summary, comparison among three techniques.

2015 SPE Journal

7
ID: jaganm Time: 14:50 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

Stage:

Page: 8

Total Pages: 13

1.00

1.00

0.90

0.90

0.80

0.80

0.70

0.70

0.60

0.60

0.50

0.50

0.40

0.40

0.30

0.30

0.20

0.20

0.10

0.10

0.00

(a) 1,440 days (WAG)

0.00

(b) 2,048 days (WAG)

1.00

1.00

0.90

0.90

0.80

0.80

0.70

0.70

0.60

0.60

0.50

0.50

0.40

0.40

0.30

0.30

0.20

0.20

0.10

0.10

0.00

0.00

(c) 1,440 days (CO2)

(d) 2,048 days (CO2)

Fig. 6CO2 concentration in Layer 3 from WAG flooding and continuous CO2 flooding.

1,440 days of production and 2,048 days of production. In Fig. 6,


the color scale corresponds to CO2 concentration with the darkest
red color representing the highest CO2 concentration (1.00) and
the darkest blue color representing the lowest concentration
(0.00). In Fig. 7, the color scale corresponds to pressure with the
darkest red color representing the highest pressure (5,500 psi) and
the darkest blue color representing the lowest pressure (1,500
psi); the upper two figures show the pressure distribution for
Layer 2 for WAG flooding at 1,440 days and 2,048 days and the
bottom two figures show the pressure distribution for Layer 2 of
continuous CO2 flooding at 1,440 days and 2,048 days.
From Fig. 6, we see that WAG flooding yields a higher CO2concentration distribution in the bottom layer than does continuous CO2 flooding. This is because the water injected in the WAG
process can control the mobility of CO2 by blocking the high-permeability channel or zones and by reducing the effects of gravity
segregation that keeps more injected CO2 in the bottom layer,
which, in turn, improves the sweep efficiency and results in higher
oil production. Fig. 7 shows that WAG flooding yields a higher
pressure distribution than continuous CO2 flooding; this is
because the water injected in the WAG process helps to maintain
a higher reservoir pressure because water is immiscible and has a
relatively low compressibility. The higher pressure also enhances
the miscibility of CO2 and oil and, hence, improves the microscopic displacement efficiency.
Fig. 8 shows the Layer 1 and Layer 3 oil-saturation distribution at 2,048 days obtained for the three processes. The color scale
corresponds to oil saturation. From Fig. 8, we can conclude that,
compared with the other two techniques, WAG flooding yields
the lowest remaining oil saturation both in the upper layer and the
bottom layer. This statement also holds for the second layer
although the results are not shown. For continuous CO2 flooding,
more oil is produced from the upper layer (Layer 1) because of

the segregation of CO2 to the upper layer; however, for waterflooding, the bottom layer achieves a lower oil saturation because
of water segregation to the bottom layer. Comparing the remaining oil-saturation distribution for Layer 1 and Layer 3 of the
WAG case, we also can conclude that CO2 does have better microscopic-displacement efficiency than water.
Optimization of SAG-Injection Process. The SAG-injection
process is very similar to the WAG process. Thus, the modified
ensemble-based algorithm still can be applied to optimize the
SAG-injection process. Generally, the SAG model is simulated
with the STARS simulator (Rossen and Boeije 2013), which is
also from Computer Modelling Group; STARS is CMGs
advanced reservoir simulator that includes options to simulate
processes which GEM does not consider. These processes include
foam injection and chemical flooding (Computer Modeling Group
2009). In this part of the paper, we first compare the performance
of the two simulators (GEM and STARS) and thereafter move forward to the optimization of the SAG-injection process with, as
initial guesses, the optimized well controls obtained by the optimization of WAG injection.
Fig. 9 presents plots of NPV vs. the number of simulation runs
for SAG and WAG injection for the cases with 16 cycles and 32
cycles. Both Fig. 9a and Fig. 9b show that WAG injection implemented by GEM yields a much higher final NPV than SAG injection implemented by STARS. However, the cases are such that
we believe that SAG should achieve a better NPV or oil-recovery
factor than WAG because of additional enhanced-oil-recovery
mechanisms generated by surfactant and foam. To understand this
discrepancy, we perform optimization of the WAG process with
STARS as well as GEM, in which the initial guess for the vector
of well controls is the same for both simulators. For the case with

2015 SPE Journal


ID: jaganm Time: 14:50 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

Stage:

Page: 9

Total Pages: 13

5,500

5,500

5,000

5,000

4,500

4,500

4,000

4,000

3,500

3,500

3,000

3,000

2,500

2,500

2,000

2,000

1,500

1,500

(a) 1,440 days (WAG)

(b) 2,048 days (WAG)


5,500

5,500

5,000

5,000

4,500

4,500

4,000

4,000

3,500

3,500

3,000

3,000

2,500

2,500

2,000

2,000

1,500

1,500

(c) 1,440 days (CO2)

(d) 2,048 days (CO2)

Fig. 7Pressure distribution for Layer 2 of WAG flooding and continuous CO2 flooding.

16 cycles, the final NPV obtained by optimizing WAG with GEM


is 18.28% higher than the NPV obtained by optimizing WAG
with STARS; for the 32-cycles WAG case, the final NPV
obtained by optimizing WAG with GEM is 18.58% greater than
the NPV obtained by optimizing WAG with STARS. Our expla-

Layer 1

Layer 3

nation is that this is because STARS uses k-values to model compositional effects, whereas GEM uses an equation-of-state-based
phase-equilibrium package for multiphase pressure/volume/temperature calculations. However, it is well-known that the liquidgas phase envelope can change dramatically as the CO2

0.80

0.80

0.80

0.70

0.70

0.70

0.60

0.60

0.60

0.50

0.50

0.50

0.40

0.40

0.40

0.30

0.30

0.30

0.20

0.20

0.20

0.10

0.10

0.10

0.00

0.00

0.00

0.80

0.80

0.80

0.70

0.70

0.70

0.60

0.60

0.60

0.50

0.50

0.50

0.40

0.40

0.40

0.30

0.30

0.30

0.20

0.20

0.20

0.10

0.10

0.10
0.00

0.00

0.00

(a) WAG

(b) Continuous CO2

(c) Waterflooding

Fig. 8Remaining oil-saturation distribution for three different techniques.


2015 SPE Journal

9
ID: jaganm Time: 14:50 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

5.00E+08

5.00E+08

4.50E+08

4.50E+08

4.00E+08
SAG-STARS
WAG-STARS
WAG-GEM

3.50E+08

NPV ($)

NPV ($)

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

Stage:

Page: 10

Total Pages: 13

4.00E+08
SAG-STARS
WAG-STARS
WAG-GEM

3.50E+08

3.00E+08

3.00E+08
0

200

400

600

800

1,000

200

Simulation Runs

400

600

800

1,000

Simulation Runs

(a) 16 cycles

(b) 32 cycles

Fig. 9NPV values vs. the number of simulation runs for SAG and WAG injection.
4.40E+08

NPV ($)

4.15E+08

3.90E+08
Original initial guess

3.65E+08

WAG controls-GEM

Table 3Summary, computational time for SAG and WAG with


different simulators.

WAG controls-STARS

3.40E+08
0

200

400

600

800

1,000

Simulation Runs

concentration increases. Thus, it is unfair to compare the performance of the WAG injection implemented by GEM with the performance of the SAG injection implemented by STARS. We have
to choose the same simulator to compare the performances of the
SAG and WAG for the same reservoir case. As seen in Fig. 9,
SAG injection yields a higher final NPV than WAG injection
when both WAG and SAG cases are simulated by STARS. As we
have mentioned previously, GEM cannot model the process of
SAG injection, so we cannot compare the performance of WAG
and SAG with GEM.
Table 3 lists the computational time required for the optimization of SAG and WAG processes with different simulators for the
two different cases. Note that, for the same reservoir case, the
computational time for the WAG optimization with GEM is less
than that of WAG simulation with STARS and SAG simulation
with STARS. For the case with 32 cycles, SAG optimization
implemented by STARS uses the most computational time (241
hours). WAG optimization with GEM requires the least computational time (80 hours), less than one-third the computational time
of SAG optimization with STARS. Thus, if we optimize the SAG
process by directly starting with the original initial guess of the
vector of optimal well controls (initial guess for rate controls of
each water injector is 2,000 STB/D, initial guess for rate controls
of each gas injector is 10 MM scf/D and initial guess for BHP
controls for each producer is 3,500 psi), even though we can
finally obtain optimal controls and NPV by directly optimizing
the SAG process, the optimization could be prohibitively computationally expensive, especially for a real-field case. One possible
way to reduce the computational time for the SAG optimization
would be to optimize the reservoir for WAG and then use the
WAG optimal controls as the initial guess for optimization of the
SAG process. Hopefully, by generating a better initial guess by

Fig. 10Performance of SAG optimization with different initial


guesses, 16 WAG cycles.

doing the WAG optimization first, we can reduce the overall computational time required for the SAG optimization. In a sense, this
means that the WAG process can serve as a proxy for optimization of SAG.
Fig. 10 shows the performance of SAG optimization with
three different initial guesses for the case with 16 cycles. The first
initial guess is the original initial guess that was used in previous
cases; the second initial guess is the WAG optimal-control vector
obtained with GEM; the third initial guess is the WAG optimalcontrol vector obtained with STARS. Table 4 presents the initial
NPV, final NPV, simulation runs, and the computational time for
SAG optimization with the three different initial guesses. From
Fig. 10 and Table 4, we see that much faster convergence is
obtained when the initial guess for the SAG optimization corresponds to the WAG optimal well controls.
The total computational times in Table 4 are calculated as follows: For the original initial-guess case, we run 1,000 simulations
with SAG, and the total computational time is 232 hours. For the
second initial-guess case, we first run 1,000 simulations with
GEM to obtain the optimal WAG controls as the initial guess for
SAG; the computational time required is 75 hours. Thereafter, we
run 411 simulations for SAG optimization requiring a computational time of 80 hours, so the total computational time is 155
hours. For the third initial-guess case, we first run 1,000 simulations with STARS to obtain the optimal WAG controls (this optimization requires 101 hours) and thereafter run 299 SAG
simulations which require a computational time of 48 hours, so
the total computational time is 149 hours.

Table 4SAG optimization with different initial guess, 16 WAG cycles.


10

2015 SPE Journal


ID: jaganm Time: 14:51 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

Stage:

Page: 11

Total Pages: 13

Table 5Final NPVs calculated from different random seeds, cases with 16 cycles and 32 cycles.

Table 6Final NPVs calculated from 32-WAG-cycle case with ensemble size equal to 10 and 30.

From the previous analysis, we can see that optimizing the


WAG process to obtain initial guesses for the optimization of the
SAG process improves computational efficiency.
Effect of Random Seed and Ensemble Size. Because of the
stochastic nature of the gradient estimate, it is important to consider the effect of the random seed and ensemble size on the optimization results. Note that, for all the cases studied previously,
we ran a few random seeds and chose the result that achieves the
highest final NPV. Table 5 lists the final NPV calculated from
five different random seeds for the 16- and 32-cycle cases. Note
that both the mean NPV and highest NPV achieved from the 32cycle case are higher than those obtained from a 16-cycle case.
To quantify the impact of ensemble size on the optimization
results, we compare the final NPVs achieved from an ensemble
size equal to 30 with an ensemble size equal to 10. For both ensemble sizes, we ran with five different seeds. For the ensemble
size equal to 30, the maximum-allowable simulation run is set
equal to 1,500, whereas for the ensemble size equal to 10, the
maximum-allowable simulation run is equal to 1,000 as before.
Table 6 presents the final NPVs achieved from the 32-cycle case
with two different ensemble sizes (10 and 30). As we can see
from Table 6, the mean value of NPV achieved from the ensemble
size equal to 30 is not significantly higher than the NPV achieved
with an ensemble size equal to 10, and the highest value obtained
from the ensemble size equal to 30 is even lower than that
obtained with the ensemble size equal to 10 even though we allow
50% more simulation runs when the ensemble size is 30. As demonstrated by Fonseca et al. (2015b), the quality of gradient can be
improved with a relatively large ensemble size. However, for the
compositional simulation, it is computationaly expensive to use a
large ensemble size with more allowable maximum simulation
runs. As seen in Table 6, at least for this problem, an ensemble
size equal to 10 is sufficient to obtain a good gradient and to
achieve a reasonable final NPV.
Comments
The study suggests that the variety of EnOpt techniques that were
developed for life-cycle waterflooding optimization can be
adapted to optimize the design variables for the WAG and SAG
processes. An EnOpt algorithm can easily be coupled with any
reservoir simulator. Thus, the ideas presented should lead to practical tools that allow a field engineer to decide whether waterflooding, continuous gas injection, WAG, or SAG will lead to the
highest NPV and/or the highest oil recovery and, at the same
time, optimize the recovery process chosen.
Conceptually, it is possible to optimize the length of each halfcycle as well as the well controls. In Bahagio (2013), water and
CO2 injection are alternated at each of four injection wells in a
five-spot pattern and the times at which each injector switches
from gas injector to water injector and vice versa are optimized.
These switching times plus the BHPs of injectors are the optimization variables. The single production well is operated at a fixed
BHP. Bahagio (2013) found that a much higher NPV was
obtained with both BHP and switching times as optimization variables than was obtained with only the BHPs, in which each half-

cycle length is equal to 180 days and each control step length for
BHPs was also equal to 180 days. From the results, it seems clear
that the use of the length of control steps equal to 180 days yields
a suboptimal result. Thus, although the incorporation of switching
times is a very appealing approach, it is possible that the results of
Bahagio (2013) overestimate the gain achieved by adding switching time to the set of optimization variables. Unfortunately, GEM
does not provide a means to incorporate switching times at an
injection well to alternate gas and water injection to further investigate the procedure suggested by Bahagio (2013). Moreover,
because we use injection rates instead of injector BHPs and use
control steps within each cycle equal to 32 days, there are two to
sixteen control steps within each cycle, depending on the number
of cycles used. If we consider, for example, the 16-cycle case in
which each half-cycle is 64 days long (each half-cycle has two
control steps), it is effectively possible to control the length of the
injection time of CO2 and water. For example, one could inject
CO2 at a high rate for the first 32 days and inject nothing for the
following 32 days, and then inject water at a low rate for 64 days,
and so on. Although this preceding scenario is not identical to
switching to water injection at the end of 32 days, it is similar
enough so that we conjecture that the addition of switching times
to the set of optimization variables would not yield a significant
increase in the NPV obtained.
Conclusions
The estimated optimal NPV of CO2-WAG injection is improved
significantly by applying production optimization. The initial
NPV, final (estimated) NPV, and cumulative oil production tend
to increase with decreasing cycle time (increasing the number
of cycles).
For the channelized reservoir example considered, the optimized WAG flood results in significantly higher valves of NPV
and oil production than the values obtained with the well controls
obtained by optimizing waterflooding or continuous CO2 injection.
If we wish to optimize the SAG-injection process, we can first
optimize the reservoir for WAG injection to obtain the optimal
controls for WAG, and then treat the optimal controls for WAG
as the initial guesses of the optimal well controls for the SAG process. This initialization procedure can yield significant reduction
in computational cost.
Acknowlegments
The support of the member companies of the University of Tulsa
Petroleum Reservoir Exploitation Projects (TUPREP) is gratefully
acknowledged. Bailian Chen also wishes to acknowledge all his
colleagues in TUPREP for their help.
References
Alhuthali, A. H., Oyerinde, D., and Datta-Gupta, A. 2007. Optimal Waterflood Management Using Rate Control. SPE Res Eval & Eng 10 (5):
539551, 2007. SPE-102478-PA. http://dx.doi.org/10.2118/02478-PA.
Amin, M. E., Zekri, A. Y., Almehaideb, R. et al. 2012. Optimization
of CO2 WAG Processes in a Selected Carbonate Reservoir: An

2015 SPE Journal

11
ID: jaganm Time: 14:51 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

Experimental Approach. Proc., Abu Dhabi International Petroleum


Exhibition and Conference, Abu Dhabi, 1114 Novermber. SPE161782-MS. http://dx.doi.org/10.2118/161782-MS.
Bahagio, D. N. T. 2013. Ensemble Pptimization of CO2 WAG EOR. MS
thesis, Delft University of Technology, Delft, Netherlands.
Bender, S. and Yilmaz, M. 2014. Full-Field Simulation and Optimization
Study of Mature IWAG Injection in a Heavy Oil Carbonate Reservoir.
Proc., SPE Improved Oil Recovery Symposium, Tulsa, USA, 1216
April. SPE-169117-MS. http://dx.doi.org/10.2118/169117-MS.
Brouwer, D. R. and Jansen, J. D. 2004. Dynamic Optimization of Water
Flooding With Smart Wells Using Optimial Control Theory. SPE J. 9
(4): 391402. SPE-78278-PA. http://dx.doi.org/10.2118/78278-PA.
Chen, Y., Oliver, D. S., and Zhang, D. 2009. Efficient Ensemble-Based
Closed-Loop Production Optimization. SPE J. 14 (4): 634645. SPE112873-PA. http://dx.doi.org/10.2118/112873-PA.
Chen, S., Li, H., Yang, D. et al. 2010a. Optimal Parametric Design for
Water-Alternating-Gas (WAG) Process in a CO2-Miscible Flooding
Reservoir. J Can Pet Technol 49 (10): 7582. SPE-141650-PA. http://
dx.doi.org/10.2118/141650-PA.
Chen, C., Li, G, and Reynolds, A. C. 2010b. Closed-Loop Reservoir Management on the Brugge Test Case. Comput. Geosci. 14 (4): 691703.
http://dx.doi.org/10.1007/s10596-010-9181-7.
Chen, C., Li, G., and Reynolds, A. C. 2012. Robust Constrained Optimization of Short- and Long-Term NPV for Closed-Loop Reservoir Management. SPE J. 17 (3): 849864. SPE-141314-PA. http://dx.doi.org/
10.2118/141314-PA.
Christensen, J. R., Stenby, E. H., and Skauge, A. 2001. Review of WAG
Field Experience. SPE Res Eval & Eng 4 (2): 97106. SPE-71203-PA.
http://dx.doi.org/10.2118/71203-PA.
Computer Modelling Group Ltd. 2009. Version 2009.11, Calgary, Canada.
Do, S. T. and Reynolds, A. C. 2013. Theoretical Connections Between
Optimization Algorithms Based On an Approximate Gradient. Comput.
Geosci. 17: 959973. http://dx.doi.org/10.1007/s10596-013-9368-9.
Esmaiel, T. E. H., Fallah, S., and van Kruijsdijk, C. P. J. W. 2005. Determination of WAG Ratios and Slug Sizes Under Uncertainty in a Smart
Wells Environment. Proc., 14th SPE Middle East Oil and Gas Show
and Conference, Bahrain, 1215 March. SPE-93569-MS. http://
dx.doi.org/10.2118/93569-MS.
Esmaiel, T. E. H. 2007. Optimization of WAG in Smart Wells: An Experimental Design Approach. PhD thesis, Delft University of Technology,
Delft, Netherlands (June 2007).
Fonseca, R. M., Leeuwenbrugh, O., Van den Hof, P. M. J. et al. 2015a.
Improving the Ensemble-Optimization Method Through CovarianceMatrix Adaptation. SPE J. 20 (1): 155168. SPE-163657-PA. http://
dx.doi.org/10.2118/163657-PA.
Fonseca, R. M., Kahrobaei, S. S., Van Gastel, L. J. T. et al. 2015b. Quantification of the Impact of Ensemble Size on the Quality of an Ensemble
Gradient Using Principles of Hypothesis Testing. Proc., SPE Reservoir
Simulation Symposium, Houston, USA, 2325 February. SPE173236-MS. http://dx.doi.org/10.2118/173236-MS.
Gao, G. and Reynolds, A. C. 2006. An Improved Implementation of the
LBFGS Algorithm for Automatic History Matching. SPE J. 11 (1):
517. SPE-90058-PA. http://dx.doi.org/10.2118/90058-PA.
Ghaderi, S. M., and Clarkson, C. R. 2012. Optimization of WAG Process
for Coupled CO2 EOR-Storage in Tight Oil Formations: An Experimental Design Approach. Proc., SPE Canadian Unconventional
Resources Conference, Alberta, Canada, 30 October1 November.
SPE-161884-MS. http://dx.doi.org/10.2118/161884-MS.
Jansen, J. D., Douma, S. D., Brouwer, D. R. et al. 2009. Closed-loop Reservoir Management. Proc., SPE Reservoir Simulation Symposium,
The Woodlands, Texas, USA, 24 February. SPE-119098-MS. http://
dx.doi.org/10.2118/119098-MS.
Johns, R. T., Bermudez, L., and Parakh, H. 2003. WAG Optimization for
Gas Floods Above the MME. Proc., SPE Annual Technical Conference and Exhibition, Denver, USA, 58 October. SPE-84366-MS.
http://dx.doi.org/10.2118/84366-MS.
Kraaijevanger, J. F. B. M., Egberts, P. J. P., Valstar, J. R. et al. 2007. Optimal Waterflood Design Using the Adjoint Method. Proc., SPE Reservoir Simulation Symposium, The Woodlands, Texas, USA, 2628
February. SPE-105764-MS. http://dx.doi.org/10.2118/105764-MS.

Stage:

Page: 12

Total Pages: 13

Liao, C., Liao, X., Zhao, X. et al. 2013. Study on Enhanced Oil Recovery
Technology in Low Permeability Heterogeneous Reservoir by WaterAlternate-Gas of CO2 Flooding. Proc., SPE Asia Pacific Oil and Gas
Conference and Exhibition, Jakarta, 2224 October. SPE-165907-MS.
http://dx.doi.org/10.2118/165907-MS.
Lorentzen, R. J., Berg, A. M., Naevdal, G. et al. 2006. A New Approach
for Dynamic Optimization of Waterflooding Problems. Proc., SPE
Intelligent Energy Conference and Exhibition, Amsterdam, 1113
April. SPE-99690-MS. http://dx.doi.org/10.2118/99690-MS.
Mirkalaei, S. M. M. M., Hosseini, S. J., Masoudi, R. et al. 2011. Investigation of Different I-WAG Schemes Towards Optimization of Displacement Efficiency. Proc., SPE Enhanced Oil Recovery Conference, Kuala
Lumpur, 1921 July. SPE-144891-MS. http://dx.doi.org/10.2118/
144891-MS.
Oliveira, D. F. 2014. Gradient and non-Gradient Based Methods for nonLinear Constrained Production Optimization Problems under Geological Uncertainty. PhD dissertation, University of Tulsa, Tulsa,
Oklahoma.
Oliveira, D. F. and Reynolds, A. C. 2014. An Adaptive Hierarchical Multiscale Algorithm for Estimation of Optimal Well Controls. SPE J. 19
(5): 909930. SPE-163645-PA. http://dx.doi.org/10.2118/163645-PA.
Oliver, D. S., Reynolds, A. C., and Liu, N. 2008. Inverse Theory for Petroleum Reservoir Characterization and History Matching. Cambridge:
Cambridge University Press. ISBN 978-0-521-88151-7.
Pritchard, D. W. L. and Nieman, R. E. 1992. Improving Oil Recovery
Through WAG Cycle Optimization in a Gravity-Overide-Dominiated
Miscible Flood. Proc., SPE/DOE Eighth Symposium on Enhanced Oil
Recovery, Tulsa, USA, 2224 April. SPE-24181-MS. http://
dx.doi.org/10.2118/24181-MS.
Raniolo, S., Dovera, L., Cominelli, A. et al. 2013. History Match and
Polymer Injection Optimization in a Mature Field Using the Ensemble
Kalman Filter. Proc., 17th European Symposium on Improved Oil Recovery, St. Petersburg, Russia, 1618 April. http://dx.doi.org/10.3997/
2214-4609.20142642.
Rossen, W. R. and Boeije, C. S. 2013. Fitting Foam Simulation Model Parameters for SAG Foam Applications. Proc., SPE Enhanced Oil Recovery Conference, Kuala Lumpur, 24 July. SPE-165282-MS. http://
dx.doi.org/10.2118/165282-MS.
Sarma, P., Durlofsky, L. J., and Aziz, K. 2005. Implementation of Adjoint
Solution for Optimal Control of Smart Wells. Proc., SPE Reservoir
Simulation Symposium, Houston, USA, 31 January2 February. SPE92864-MS. http://dx.doi.org/10.2118/92864-MS.
Shafian, S. R. M., Bahrim, R. Z. K., Hamid, P. A. et al. 2013. Enhancing
the Efficiency of Immiscible Water Alternating Gas (WAG) Injection
in a Matured, High Temperature and High CO2 Solution Gas
ReservoirA Laboratory Study. Proc., SPE Enhanced Oil Recovery
Conference, Kuala Lumpur, 24 July. SPE-165303-MS. http://
dx.doi.org/10.2118/165303-MS.
Srivastava, J. P. and Mahli, L. 2012. Water-Alternating-Gas (WAG) Injection: A Novel EOR Technique for Mature Light Oil FieldsA Laboratory Investigation for GS-5C Sand of Gandhar Field. Proc., 9th
Biennial International Conference and Exposition on Petroleum Geophysics, Hyderabad, India.
Stordal, A. S., Szklarz, S. P., and Leeuwenburgh, O. 2014. A Closer Look
at Ensemble-Based Optimization in Reservoir Management. Proc.,
14th European Conference on the Mathematics of Oil Recovery
(ECMOR XIV), Catania, Sicily, Italy, 811 September.
Wang, C., Li, G., and Reynolds, A. C. 2009. Production Optimization in
Closed-Loop Reservoir Management. SPE J. 14 (3): 506523. SPE109805-PA. http://dx.doi.org/10.2118/109805-PA.
Yan, X. and Reynolds, A. C. 2014. Optimization Algorithms Based On
Combining FD Approximations and Stochastic Gradients Compared
With Methods Based Only on a Stochastic Gradient. SPE J. 19 (5):
873890. SPE-163613-PA. http://dx.doi.org/10.2118/163613-PA.
Zandvliet, M., Bosgra, O., Jasen, J. et al. 2007. Bang-Bang Control and
Singular Arcs in Reservoir Flooding. J. Petrol. Sci. & Eng. 58:
186200. http://dx.doi.org/10.1016/j.petrol.2006.12.008.
Zhou, D., Yan, M., and Calvin, W. M. 2012. Optimization of a Mature
CO2 FloodFrom Continuous Injection to WAG. Proc., SPE Improved
Oil Recovery Symposium, Tulsa, USA, 1418 April. SPE-154181MS. http://dx.doi.org/10.2118/154181-MS.

12

2015 SPE Journal


ID: jaganm Time: 14:51 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

J173217 DOI: 10.2118/173217-PA Date: 30-September-15

Bailian Chen is a PhD degree candidate in the McDougall


School of Petroleum Engineering at the University of Tulsa. His
research interests include production optimization, enhanced
oil recovery, and uncertainty quantification of data-acquisition programs (pilot and surveillance). Chen holds BS and MS
degrees in petroleum engineering from the China University of
Petroleum (East China). He is a member of SPE.
Albert C. Reynolds is a professor of petroleum engineering
and mathematics, holder of the McMan Chair in Petroleum
Engineering, and Director of TUPREP at the University of Tulsa,
where he has been a faculty member since 1970. Reynolds
research interests include optimization, scientific computa-

Stage:

Page: 13

Total Pages: 13

tion, assessment of uncertainty, reservoir simulation, history


matching, and well testing; he has coauthored more than
100 refereed papers in these areas. Reynolds holds a BA
degree from the University of New Hampshire, an MS degree
from Case Institute of Technology, and a PhD degree from
Case Western Reserve University, all in mathematics. He has
received the SPE Distinguished Achievement Award for Petroleum Engineering Faculty, Reservoir Description and Dynamics Award, Formation Evaluation Award, and John Franklin
Carll Award; he became an SPE Distinguished Member in
1999 and an SPE Honorary Member in 2014. Reynolds is an
associate editor for SPE Journal and for Computational
Geosciences.

2015 SPE Journal

13
ID: jaganm Time: 14:51 I Path: S:/J###/Vol00000/150089/Comp/APPFile/SA-J###150089

You might also like