Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Comput Geosci (2014) 18:563–577

DOI 10.1007/s10596-014-9412-4

ORIGINAL PAPER

Challenges in adjoint-based optimization of a foam EOR


process
M. Namdar Zanganeh · J. F. B. M. Kraaijevanger ·
H. W. Buurman · J. D. Jansen · W. R. Rossen

Received: 28 February 2013 / Accepted: 13 February 2014 / Published online: 4 May 2014
© Springer International Publishing Switzerland 2014

Abstract We apply adjoint-based optimization to a changes and may present similar challenges to gradient-
surfactant-alternating gas foam process using a linear foam based optimization.
model introducing gradual changes in gas mobility and a
nonlinear foam model giving abrupt changes in gas mobility Keywords Adjoint-based optimization · Simulation ·
as function of oil and water saturations and surfactant con- Enhanced oil recovery · Foam
centration. For the linear foam model, the objective function
is a relatively smooth function of the switching time. For the
nonlinear foam model, the objective function exhibits many 1 Introduction
small-scale fluctuations. As a result, a gradient-based opti-
mization routine could have difficulty finding the optimal A hydrocarbon reservoir should be produced in an effi-
switching time. For the nonlinear foam model, extremely cient and, if possible, optimal manner. Our objective here
small time steps were required in the forward integration is maximizing the cumulative oil production in a foam
to converge to an accurate solution to the semi-discrete surfactant-alternating gas (SAG) process [22] by finding
(discretized in space, continuous in time) problem. The the optimum switching time between the surfactant and gas
semi-discrete solution still had strong oscillations in grid- injection cycles. Foam is a means of improving sweep effi-
block properties associated with the steep front moving ciency that reduces the gas mobility by capturing gas in
through the reservoir. In addition, an extraordinarily tight foam bubbles and hindering its movement [20]. Specifically,
tolerance was required in the backward integration to obtain we investigate the performance of adjoint-based optimiza-
accurate adjoints. We believe the small-scale oscillations tion applied to this enhanced oil recovery (EOR) process.
in the objective function result from the large oscillations This work is the first attempt to apply optimal control theory
in gridblock properties associated with the front moving to foam EOR.
through the reservoir. Other EOR processes, including sur- Adjoint-based schemes are very efficient for gradient cal-
factant EOR and near-miscible flooding, have similar sharp culations and are particularly useful in optimization prob-
lems where a large number of controllable parameters must
be optimized. Here, we use the Shell in-house reservoir sim-
M. Namdar Zanganeh ulator MoReS (hereafter referred to as the simulator). The
Xodus Group, Oranjestraat 4 2514 JB The Hague, adjoint routine in the simulator has been extensively vali-
The Netherlands dated for waterflooding [10, 25] and also recently for poly-
mer flooding [24]. In principle, adjoint-based optimization
J. F. B. M. Kraaijevanger · H. W. Buurman
Shell Global Solutions International, Kessler Park 1, could be applied to many other EOR processes.
2288 GS Rijswijk, The Netherlands In Section 2 of this paper, we first review briefly the
history of the application of optimal control theory in
J. D. Jansen · W. R. Rossen ()
petroleum engineering and EOR in particular. Next, in
Department of Geoscience and Engineering,
Delft University of Technology, Delft, Netherlands Section 3, we define the problem and associated simula-
e-mail: w.r.rossen@tudelft.nl tion models. Section 4 describes and explains oscillations
564 Comput Geosci (2014) 18:563–577

that we observed in the numerical simulation of these mod- sequential or simultaneous) of the governing PDEs, derive
els. The effect of the switching time on cumulative oil the adjoint algebraic equations whereafter the forward
production is investigated in Section 5. In particular, we and adjoint algebraic equations are solved numerically.
discuss the application of the adjoint method for calculating Depending on the details of the discretization methods for
the gradient of the cumulative oil production with respect the forward and adjoint equations, these approaches may
to switching time. In Section 6, we investigate the perfor- lead to differences in numerical performance. Ramirez et
mance of a gradient-based optimization routine in finding al. [5, 19] had difficulty in solving the adjoint equations,
the switching time that maximizes the cumulative oil pro- obtained using the first approach, in surfactant-flooding
duction. We examine the performance in a number of cases optimization. The adjoint PDEs had nonsmooth coefficients
varying in the simulation mode (i.e., 1D or 3D) and the caused by the physical nature of the governing forward
foam model (i.e., linear or nonlinear). We end the paper with PDEs, making it difficult to find a stable numerical adjoint
conclusions and implications in Section 7. scheme. Fathi and Ramirez associated this problem with the
discontinuity in the adjoint equation coefficients caused by
two shocks (a Buckley-Leverett shock and a shock at the
2 Optimal control theory surfactant front). In [5, 6], this problem was circumvented
by replacing the adjoint equations with simpler, quasilinear
Optimal control theory is a mathematical discipline aiming approximations. Later, [7] applied optimal control theory to
at optimal control of time-dependent systems governed by micellar/polymer flooding using both the first and the third
partial or ordinary differential equations. In optimal con- approach. Also here they reported problems with disconti-
trol theory, it is a common practice to use gradient-based nuities in the coefficients of the adjoint PDEs. As a remedy,
optimization methods, where the gradient is obtained with they derived adjoint equations using the third approach.
adjoint techniques involving the solution of adjoint equa- They provided some theoretical comparisons between the
tions backward in time. Originally developed for nonlinear approaches, but did not give details about their comparative
trajectory optimization in aerospace engineering, optimal computational performance.
control theory emerged in petroleum engineering during the For 2D steamflood optimization, [13] reported that find-
1970s for computer-assisted history matching; see [17] for ing a stable adjoint solution (using the third approach, but
an overview. Starting about a decade later, the methodology with different time-step sizes to solve the forward and the
was used for the optimization of tertiary recovery processes adjoint equations) could be a significant problem. We note
such as surfactant, polymer, CO2 , or steam flooding; see, that the use of different step sizes requires interpolation
e.g., [5–7, 12–14, 18, 19]. Applications to waterflood opti- of the states, computed during the forward simulation and
mization followed soon and are continuing today (see [8] for during the backward simulation. This gives it a flavor of
a recent overview), while EOR applications seem to have the second approach. Liu and Ramirez [12] extended the
seen a revival; see [24] and [11] on polymer flooding and [9] analysis of steamflood optimization to 3D, and they found
on miscible gas injection. However, systematic numerical that a too-large time-step size for the adjoint solution can
optimization of EOR processes is not yet common practice cause nonsmooth optimal control strategies. They claimed
in the reservoir simulation community. that refinement of the time step should smooth the control
A possible explanation is that applying optimal con- strategies. They, moreover, argued that the adjoint solution
trol theory can lead to complications. A certain method may not necessarily result in correct gradients when com-
may work well for simple examples, but result in subop- puted with the same time-step sizes as used to compute the
timal solutions in more complex systems. A complicating forward solution. This is contrary to the commonly held
factor in the comparison of earlier papers is that there opinion that a correct adjoint solution should be based on
are three different approaches to derive the adjoint equa- exactly identical spatial and time discretization choices for
tions: (1) Starting from the governing partial differential the forward and backward equations.
equations (PDEs) for the time-dependent system, derive Sudaryanto [23] applied approach three in miscible EOR
the adjoint PDEs whereafter the “forward-in-time” sys- and faced problems in finding the optimum switching time
tem PDEs and “backward-in-time” adjoint PDEs are both (i.e., the optimal moment in time to terminate gas injection
discretized in space and time in order to solve them numer- in one injector and start it in another injector) for a displace-
ically. (2) Starting from the ordinary differential equations ment with variable mobility. He stated that these problems
(ODEs) obtained by spatial discretization of the governing are possibly caused by numerical errors in solving large sys-
PDEs, derive the adjoint ODEs whereafter the “forward” tems of state and adjoint equations, but did not analyze the
and adjoint ODEs are both discretized in time in order problem in depth.
to solve them numerically. (3) Starting from the algebraic We also encountered problems related to coefficient val-
equations obtained by spatial and time discretization (either ues that are strongly nonlinear functions of the states,
Comput Geosci (2014) 18:563–577 565

although of a somewhat different manifestation. The main surfactant adsorption affects the value of the optimal slug
issue was the magnitude of the relative tolerance for the size, but does not fundamentally alter the optimization
adjoint linear solver, as discussed in Section 6. In addition, process. Therefore, for simplicity, we exclude it here.
we encountered nonsmooth adjoint gradients with various We use two foam models, one with abrupt changes (non-
fluctuations, as described also in Section 6. linear model) and one without (linear model). In both cases,
foam is not explicitly modeled as such, but only implic-
itly, by introducing a reduction factor of the gas mobility
3 Problem statement in our oil/water/gas system. In these foammodels, the rel-
ative gas mobility in the absence of foam λrg is rescaled
0
We model a foam displacement process in 1D and 3D.
to the relative gas mobility in the presence of foam (λrg )
Specifically, we seek the period of injection of the surfactant
by multiplying it by a dimensionless reduction factor (Eq.
slug in a single-cycle SAG flood that maximizes cumulative
(A-2) in the Appendix); see also [16, 21]. Our foam mod-
oil production. We represent the duration of the surfactant
els incorporate the effects of water saturation, oil saturation,
slug by the switching time (ts ). Before time ts , there is no gas
and surfactant concentration on the reduction in gas rela-
or foam in the reservoir. At ts , surfactant injection is termi-
tive mobility as illustrated in Fig. 1. In the nonlinear foam
nated, and gas injection starts. Foam forms in the reservoir
model, foam is weakened at low water saturations according
where gas, water, and surfactant are all present in sufficient
to Eq. (A-3). Note that foam strength equals zero at the con-
amounts. The total period of injection is fixed; thus as ts
nate water saturation according to this equation; otherwise,
increases, the period of liquid injection increases and that of
the performance of the SAG flood can be greatly distorted
gas injection decreases.
[16]. In addition, foam is killed at high oil saturations as
The 3D reservoir is box-shaped, 100 m × 100 m × 30 m
prescribed in Eq. (A-4). The foam sensitivity to surfactant
(modeled with 10 × 10 × 10 gridblocks), sealed on all
concentration is defined by Eq. (A-5). In the nonlinear foam
bounding surfaces, with one injection and one production
model, foam strength changes abruptly with small changes
well, located at opposite corners of the reservoir. Both wells
in water saturation, in agreement with laboratory data [4].
are vertical, located in the middle of the corner gridblocks,
It also changes abruptly with changes in surfactant con-
and are perforated in the entire interval. Both wells operate
centration, for reasons discussed in [15]. In fact, laboratory
at a constant prescribed bottom-hole pressure; thus, foam
data suggest an even sharper change in foam strength with
affects not only sweep efficiency but also injection and pro-
water saturation than modeled here [4], but, as we will see,
duction rates. In 1D simulations, the reservoir is 100 m ×
the nonlinear functions we use here are already challeng-
1 m × 1 m (100 × 1 × 1 gridblocks). In the 1D simula-
ing enough for the simulator. These types of functions can
tions, wells are positioned in the middle of the first and last
lead to drastic changes in gas mobility in individual grid-
gridblocks. We use the Peaceman model for injectivity into
blocks as fronts pass [15, 21]. Note that all these functions
and productivity from the well gridblocks. The reservoir is
are differentiable over the range of conditions that may be
homogeneous in permeability and porosity.
encountered in the simulation.
We aim at investigating gas performance in sweeping
In general, optimization of an objective function can
oil. EOR is implemented usually after a period of water-
be carried out by combining an adjoint routine and an
flooding. Therefore, we set the initial oil saturation in the
optimization routine for finding the optimum. The adjoint
reservoir at the residual oil saturation to waterflood (Sorw ).
routine is employed to calculate the analytical gradient of
As a result, oil is neither displaced nor produced before the
the objective function with respect to various static and
switching time. Oil, water, and gas are assumed to be com-
dynamic parameters. Gradients point in the direction of the
pletely immiscible. We assume that gas, though immiscible
maximum increase in the objective function, and a gradient-
with oil, is capable of displacing oil; specifically, we set the
based optimization routine utilizes the gradients to search
residual oil saturation to gas flood (Sorg) lower than Sorw .
for the optimum.
We use the linear isoperm model [2] for three-phase rela-
As mentioned above, our objective is finding the switch-
tive permeabilities. These and other simulation parameters
ing time (ts,opt ) that maximizes the objective function, J ,
are summarized in the Appendix. The gas phase is slightly
which is defined as the cumulative oil production at the end
compressible in the simulations (i.e., cg = 1.68E – 8 Pa−1 ).
of the simulation (time T):
Water and oil are incompressible.  T
For the modeling of surfactant, the simulator repre- J = Qo (t)dt, (1)
sents it as a normal active component in the water phase. 0
As will be discussed below, the surfactant concentration where Qo is the surface volume rate of the oil phase,
plays an important role in our foam model. We do not and the end time T is fixed. We choose a simple case
include adsorption of surfactant in our model. The level of with only one control variable (i.e., the switching time
566 Comput Geosci (2014) 18:563–577

1 1 1

Fw Fo Fs

0 0 0
Sw * 1 S o* 1 Cs*
Sw So Cs
a b c
Fig. 1 Foam models. The linear foam model is illustrated by dashed saturation, nonlinear model: foam is killed at So ≥ S∗o = fmoil
lines, and the nonlinear foam model, by solid lines. A value of Fi = (Eq. (A-4)). c Foam sensitivity to surfactant concentration, nonlin-
0 means that foam has no effect on gas mobility (Eq. (A-2)). a ear model: Fs = 1 for Cs ≥ Cs∗ , where Cs∗ = 0.0012 is half the
Foam sensitivity to water saturation, nonlinear model: foam is weak- injected concentration of 0.0024 (weight fraction) (Eq. (A-5)). The
ened at Sw ≤ S∗w = fmdry (Eq. (A-3)). b Foam sensitivity to oil foam models are described further in the Appendix

between the surfactant and gas injection cycles) for this = 2) only if the gas mobility in the first gridblock has
investigation. With only one control variable, the global become significant. When the water saturation in the first
trend of the objective function J (i.e., the cumulative oil gridblock has dropped below the critical value S∗w = 0.316,
production) can be easily constructed with a reasonable gas mobility strongly rises in that gridblock and allows gas
number of simulation runs. With the global trend of J to flow to gridblock 2, causing a jump in the injection rate
known, ts,opt can be easily identified. This way, we have a Qg (see Fig. 3a) and all gridblock pressures (in particu-
solid reference case against which to evaluate the perfor- lar those downstream of gridblock 1, see Fig. 3b). Similar
mance of the adjoint-based optimization methodology. results are reported by [21]. This behavior repeats itself in
the remaining gridblocks until Sw drops below S∗w in the last
4 Numerical oscillations in the forward solution gridblock and gas breaks through to the production well.
Consequently, we observe as many spikes as the number of
In this section, we discuss the strong oscillations that gridblocks in the gas injection rate profile in 1D (Fig. 2).
we observed in the numerical simulation of the 1D and The first five spikes of the profile with 100 gridblocks in
3D models with the nonlinear foam model. In these sim- Fig. 2 have been magnified in Fig. 3. We have also plot-
ulations, the sudden change in the injection composition ted the saturation profiles in the first five gridblocks. The
from 100 % surfactant solution to 100 % gas at the moment at which the nth spike in Qg occurs corresponds to
switching time produces a period of strong oscillations the moment at which water saturation in the nth gridblock
in the gas injection rate (Qg ) shortly after the switch. drops below S∗w = 0.316. The fluctuation in injection rate
We start with the discussion of the oscillations in the 1D decreases in magnitude as the foam bank advances in Fig. 2
model. The reason why we focus on these oscillations because there is more (slightly) compressible gas between
is that they are the cause of certain problems (discussed the injection well and the front to cushion the injection well
in Section 6) that we encountered with gradient-based from the fluctuating mobility at the front. Note that in Fig. 3,
optimization. the decrease in Sw in each gridblock is nearly identical to
We start with the 1D model with the nonlinear foam that in nearby gridblocks, but the fluctuations in pressure
model. Immediately after the switching time, the oil and decrease as the front advances. The fluctuations are smaller
water saturations throughout the reservoir are equal to So = and briefer with 1,000 gridblocks than with 100 (Fig. 2)
Sorw = 0.3 and Sw = 1 - Sorw = 0.7, and the surfactant because it takes less time for the gridblock at the front to dry
concentration (almost) to Cs = 0.0024 (weight fraction). out and allow mobility to increase.
Referring to the nonlinear foam model (see Fig. 1 and the We recall that we maintain
 a constantbottom-hole pres-
inj
Appendix), these conditions are favorable for foam forma- sure in the injection well Pwf = 250bar . Since Qg is pro-
tion [Sw > S∗w = 0.316, So < S∗o = 0.4 and Cs > C∗s = portional to the pressure difference between the wellbore
0.0012]. Hence, when gas starts being injected in the first and P1 , this pressure difference reflects the fluctuations
gridblock (n = 1), its mobility is very low. Since the simula- in P1 . Although the fluctuations in P1 are small and not
tor uses an upwind scheme for the flux calculations between noticeable in Fig. 3b, once those small pressure differences
gridblocks (i.e., the phase flux between two adjacent grid- are multiplied by the large total mobility in the well grid-
blocks uses the phase mobility of the upstream gridblock), block, the product produces noticeable fluctuations in the
gas can flow at a significant rate to the next gridblock (n gas injection rate.
Comput Geosci (2014) 18:563–577 567

Fig. 2 Period of fluctuating 0.85


gas-injection rate in a 1D 100 gridblocks
displacement with ts = 66 days. 0.8 1000 gridblocks
The injection well operates at a
constant prescribed bottom-hole 0.75
pressure. The number of spikes

Q g (m3/day)
is equal to the number of 0.7

gridblocks. This figure also


0.65
illustrates the effect of grid size
on the fluctuations. Finer grids
0.6
reduce the amplitude of the
spikes and shrink the period of 0.55
large fluctuations to a period
close to the switching time at 0.5
ts = 66 days 66 68 70 72 74 76 78 80
Time (days)

In the remainder of this section, we turn our attention caused by the presence of foam; in this example, ts is
to the 3D model, which has similar oscillations if we use 66 days.
the nonlinear foam model. Figure 4 illustrates the Qg pro- In 3D, relating the spikes to the gridblocks is less
file for a 3D reservoir in this case and the fluctuations straightforward than in 1D because there are many more

Fig. 3 Magnification of the first


five spikes in the gas injection
a
1.3
rate in Fig. 2 (with
0.7
100 gridblocks): relation to
1.2
water saturation (a) and pressure
(b) in each gridblock. The 0.6
moment at which the nth spike 1.1
in Qg occurs is the moment at 0.5
which Sw in the nth gridblock 1
drops below S∗w = 0.316
Q g (m3/day)

Sw
0.4
0.9
Grid 1
0.3
0.8
Grid 2
Grid 3
Grid 4 Grid 5
0.7 0.2
Qg

0.6 0.1

0.5 0
66 66.2 66.4 66.6 66.8 67
Time (days)
b
1.3 255

P1
1.2 250

P2 P3 P4 245
1.1 P5
240
Gridblock Pressure (bar)

1
Q g (m 3 /day)

235
0.9
Grid 1 230
0.8
Grid 2 225
Grid 3
Grid 4 Grid 5
0.7
Qg 220

0.6 215

0.5 210
66 66.2 66.4 66.6 66.8 67
Time (days)
568 Comput Geosci (2014) 18:563–577

3500
500
3000 400
300
2500
200
100
2000

Q g (m3/day)
0
60 70 80 90
1500

1000

500

0
0 100 200 300 400 500
Time (days)

Fig. 4 Gas injection rate profile versus time for a 3D displacement. Simulation ends after 540 days and ts = 66 days. The inset magnifies the
fluctuations right after the switching time. The reservoir consists of 1,000 (10 × 10 × 10) gridblocks

flow paths between injection and production well. As in the 5 Calculation of the switching time gradient
1D model, grid refinement causes the number of spikes to
grow while reducing the amplitude of the spikes and shrink- In this paper, the goal is finding the switching time (ts,opt )
ing the period of large fluctuations to a period close to the that maximizes the cumulative oil production (J ). For
switching time (see Fig. 5). gradient-based optimization, this means that we must be
We conclude this section with two remarks that apply to able to compute the derivative dJ /dts .
both 1D and 3D. The dynamic system of interest is a producing reservoir
First, if a constant Qg is prescribed instead of a constant and its wells, described by a set of coupled algebraic and
inj inj
Pwf , fluctuations occur in the Pwf profile to accommodate differential equations. After discretization in space (based
the prescribed Qg (cf. [21]). on gridblocks) and time (using a time stepping method),
we arrive at a discrete dynamic system described by the
Secondly, the fluctuations do not vanish if we take
following set of algebraic equations:
(much) smaller time-steps, but converge to the shapes
as presented above in the limit t → 0. Hence, these gk (xk−1 , xk , tk−1 , tk ) = 0 (k = 1, ...., N ), (2)
fluctuations are not related to the time-step size, but
are due to the spatial discretization scheme used in the where k is the time-step index, and N is the number of time
simulator. steps. At every time step k, the state vector (xk ) is a column

Fig. 5 Effect of grid size on the 700


fluctuations in the gas injection 10x10x10
profile in a 3D displacement 30x30x10
shown in Fig. 4. Finer grids 600
reduce the amplitude of the
spikes and shrink the fluctuation
500
interval to a region close to the
switching time at ts = 66 days.
In the refined case, each 600
Q g (m3/day)

400
gridblock is nine times smaller
than the original gridblock
300 400

200
200

100
0
65 70 75
0
65 75 85 95 105 115 125
Time (days)
Comput Geosci (2014) 18:563–577 569

t0 = 0 t k-1 tk t k+1 tN = T

Fig. 6 Schematic for calculation of the time gradient dJ /dtk in the simulator. All the times except tk are kept fixed in the computation of the
time gradient with respect to tk

vector that contains the saturations, pressures, and compo- 5.1 Calculation of the time gradients dJ /dtk
sitions of all reservoir gridblocks and well discretization with the adjoint method
nodes. During the forward run (i.e., normal simulation run),
xk is successively calculated by solving Eq. (2), starting For the calculation of the time gradients dJ /dtk (k =
with the initial state vector x0 at t0 = 0. 1,2,. . .,N), we use the adjoint method (see, e.g., [3]). Alter-
Before we turn our attention to the derivation of natively, the calculation of these gradients may also be done
an expression for the desired gradient dJ /dts , we first with the perturbation method or with the forward sensitivity
introduce the closely related gradients dJ /dtk (for k = method (see, e.g., [10]). In the forward sensitivity method
1,2,. . .,N). The latter gradients represent the amount of and the adjoint method, the partial derivatives in Eqs. (5)
change in the objective function when tk slightly varies, but to (8) below must be available in the source code of the
other times do not (Fig. 6). In other words, all the times simulator.
except tk are kept fixed for the computation of the time The first step of the adjoint method is to solve the adjoint
gradient with respect to tk . equations for the adjoint variables λk (see, e.g., [3], [8] or
Large values of dJ /dtk are expected only at the end time [10]),
(T) and at those times at which a change occurs in con- ∂J
trol settings like constraint type, constraint value, injection λk Ak + λk+1 Bk+1 = (k = 1, 2, ..., N − 1) , (5)
∂xk
composition, etc. The value of dJ /dtk for k = N (i.e., at tk
= tN = T) reflects the effect of changing the length of the ∂J
λN AN = , (6)
simulation period. ∂xN
The value dJ /dtk at tk = ts reflects the effect of pro- where the square matrices Ak and Bk are defined as follows:
longing or shortening the surfactant injection period on J .
∂gk ∂gk
If it is positive, prolonging the surfactant injection cycle Ak = , Bk = . (7)
leads to an increased J . Conversely, if it is negative, it is ∂xk ∂xk−1
not beneficial to extend the surfactant injection period. It Note that in this formulation the adjoint variables λk are
seems reasonable to set the (desired) gradient dJ /dts equal defined as row vectors. The adjoint equations are solved
to dJ /dtk , so that backward in time, i.e., after solving (6) for λN , the equations
in (5) are solved consecutively for k = N - 1, N - 2,. . ., 1.
dJ dJ The adjoint equations are linear and solved by an iterative
= (where k is the index with tk = ts ) . (3)
dts dtk method. In the simulator, the convergence criterion is that
the norm of the residual has become smaller than a user-
Note, however, that in actual simulations, a change in specified relative tolerance times the norm of the initial
the switching time ts causes a change in many values tk , residual.
depending on the chosen step-size strategy as discussed in Once the adjoint variables λk are known, the gradients
Section 5.2. For a general step-size strategy, all the values dJ /dtk are computed as follows:
tk are dependent on the switching time ts , leading to the
dJ ∂J ∂g ∂gk+1
following improved formula for dJ /dts : dtk = ∂tk − λk ∂tkk − λk+1 ∂tk k = 1, 2, ..., N − 1,
∂gN
(8)
dJ ∂J
N   = ∂tN − λN ∂tN k = N.
dJ  dJ dtk
dtN
= × . (4)
dts
k=1
dtk dts 5.2 Dependence of dJ /dts on the chosen step-size strategy

In practice, the difference of the estimates obtained with for- The computation of the switching time gradient dJ /dts is
mulas (3) and (4) is usually negligible, because all values based on Eq. (4). In this section, we illustrate this computa-
dJ /dtk (except those at tk = ts and at tk = tN = T) are usu- tion for a given step-size strategy.
ally very small (as they represent time discretization errors), We use the step-size strategy that starts with a given
and our assumption of a fixed end time tN = T implies (small) initial step size tstart and doubles the step size each
dtN /dts = 0. In certain cases, however, the difference can be time step until reaching a given maximum step size tmax .
significant. We will discuss this further in Section 5.2. We continue with this maximum step size until we reach the
570 Comput Geosci (2014) 18:563–577

Δt 0.1 0.2 0.4 0.8 1 1 1 0.2 0.1 0.2 0.4 0.8 1 1 1 0.8

t 0 0.1 0.3 0.7 1.5 2.5 3.5 4.5 4.7 4.8 5 5.4 6.2 7.2 8.2 9.2 10
tj

Fig. 7 Simple example for calculating dJ /dts where ts = 4.7 days points. The time-step strategy is to start at both t = 0 and t = ts with t
and T = 10 days. Small changes in the switching time affect only the = tstart = 0.1 day and double it until reaching tmax = 1 day. Note
time points tk in boldface. The term dtk / dts is equal to unity at these that the steps are not of equal size

switching time (with a last time step possibly smaller than around the times that the spikes (in the injection rate Qg )
tmax ). At the switching time, we start again with the ini- occur that have been described in Section 4 (see, e.g., Fig. 2
tial step size tstart and repeat the same step-size strategy or 4). As was illustrated in Fig. 3 for the 1D problem, there
until we reach the simulation end time T. Figure 7 shows an is a repeating cycle of spikes corresponding to the events
example with ts = 4.7 days and T = 10 days. that the water saturation Sw drops below S∗w in successive
If the sum of the intermediate gradients (i.e., the gra- gridblocks. In Fig. 8, we have combined the injection rate
dients corresponding to all tk with ts < tk <T) is large profile and a plot of the time gradients dJ /dtk into one plot,
compared to the gradients at ts and T, it can be generally clearly showing that the time gradients are (relatively) large
concluded that the time discretization error is too large. In at the times that the spikes occur. Apparently, the numer-
that case, a different step-size strategy should be considered. ical solution of the (spatially discretized) problem, and in
Unlike the gradients at ts and T, these intermediate gradients particular the objective function J , is very sensitive to step-
(and their sum) can be made arbitrarily small by choosing size changes around the times that the spikes occur. As
smaller time steps. mentioned above, we could reduce the size of these rela-
In our problem, the time-step sizes in the vicinity of ts tively large intermediate gradients by taking even smaller
must be refined compared to the rest of the time grid. This step sizes. As discussed in Section 6, these spikes do not
causes a more accurate simulation in the hectic period right influence the accuracy of the adjoint gradients at ts . In other
after the switch. The start of gas injection at ts introduces a words, the spikes are an accurate reflection of the behavior
substantial change in the process due to initiation of foam of the discretized system.
formation and coalescence in the reservoir. Therefore, a fine
time grid is required during that time period to capture the
sudden changes more accurately. In most of our simula- 6 Numerical optimization
tions, we took the step-size strategy corresponding to that
in Table 3 in the Appendix, or a small variation thereof. In We investigate the capability of a gradient-based, stee-
our problem, we encountered large intermediate gradients pest-ascent optimization method applied to our SAG

Fig. 8 Illustration of how 2E-6 1.2


fluctuations in gas injection rate d / dt
correlate with fluctuations in
dJ /dt in the 1D simulation 1E-6 Qg 1.1
(with 100 gridblocks). The
moment at which the nth spike 0E+0 1
in gas injection rate occurs 66 66.25 66.5 66.75 67
d / d t (m3/day)

corresponds to the moment at


which water saturation in the nth -1E-6 Grid 4 0.9
Grid 5
Q g (m3/day)

gridblock drops below S∗w =


0.316. Spikes in the
time-gradient plot also occur at -2E-6 Grid 3 0.8
Grid 2
the same moments. For clarity,
we have displayed the gridblock -3E-6 0.7
number corresponding to each
spike beneath the spike in the
time-gradient curve -4E-6 0.6

Grid 1
-5E-6 0.5
Time (days)
Comput Geosci (2014) 18:563–577 571

foam process. Our objective is maximizing the cumu- 6.1 Optimization with the linear foam model in a 3D
lative oil production J by optimizing the switching reservoir
time ts .
We validated the gradients calculated with the adjoint The results for the linear foam model in the 3D reservoir
routine in the simulator by comparing them to the numerical are shown in Fig. 10. J has a global minimum at ts =
gradients obtained with the perturbation method with suffi- 6 days and a global maximum at ts = 28.5 days (magni-
ciently small perturbation sizes, as discussed further below. fied in the inset of Fig. 10). The sign of the adjoint gradients
We use the central difference scheme for the numerical gra- is consistent with the global trend of J . In this case, as
dients. We discovered that an inappropriate choice of the well as those in Sections 6.2 and 6.3, the perturbation size
relative tolerance for the adjoint linear solver (cf. Section t for estimating gradients numerically is 1 day before the
5.1) is an important source of getting incorrect gradients global optimum and 2 days after the optimum. In Fig. 10, the
from the adjoint routine in our problem. We found that 1E- adjoint gradient passes through zero at both global optima.
12 is an adequate choice for the relative tolerance for our Except for the initial point at ts = 1 day, there is a good
1D simulations and for the linear foam model in 3D, but we agreement between the numerical and analytical (adjoint)
had to tighten the relative tolerance to its smallest possible gradients. Optimizing the switching time in this process by
value of 2.25E-16 to obtain accurate adjoint gradients for the steepest-ascent method leads to the correct optimum
the nonlinear foam model in 3D simulations. Further details switching time at 28.5 days.
are in [15]. With these relative tolerances, we are confident Similar results are obtained for the linear foam model
that the adjoint routine computes accurate gradients in our in a 1D reservoir [15]. The steepest-ascent method can
simulations. successfully find the optimum switching time in both cases.
During our validation efforts of the adjoint gradients, we
encountered a significant number of cases, in which the 6.2 Optimization with the nonlinear foam model in a 1D
adjoint gradient did not match the numerical perturbation reservoir
gradient, suggesting that the adjoint gradients were wrong.
However, a closer examination (mainly by sampling J at The nonlinear foam model produces fluctuations and jumps
a finer scale) revealed in all those cases that the adjoint in the adjoint gradient that is not present with the lin-
gradients are correct, and that the differences are caused ear foam model. Figure 11 shows the results in 1D. The
by small-scale local fluctuations in J superimposed on its numerical gradient accurately represents the global trend
global trend. Figure 9 illustrates schematically a possible of J , and it varies smoothly. However, it differs (both in
difference between global and local trends of J . By global sign and magnitude) from the analytical (adjoint) gradi-
we mean a trend of J that is obtained by coarsely sampling ent at certain switching times ts . This difference originates
of the objective function, by choosing large perturbation from the fact that J shows fluctuations as ts varies on the
sizes. The local trend of J is obtained by choosing much scale of 0.01 day superimposed on the larger trend. These
smaller perturbation sizes than those used for constructing fluctuations affect the adjoint gradients and, when
the global trend. In Fig. 9, the global trend of J around taken with sufficiently small perturbations, the numerical
the two black squares appears to be increasing with increas- gradients.
ing ts ; however, the local trend of J focusing on the region Figure 11a illustrates the global trend of J in which the
between the two black squares reveals details that were not global maximum is at ts,opt = 8.25 days. There is a neg-
noticeable in the global trend. ative adjoint gradient and a positive numerical gradient at
As we will further discuss below, the observed dif- ts = 7 days. The global trend of J appears to be monoton-
ferences in local and global trend (i.e., the small-scale ically increasing before ts,opt . However, the expanded view
local fluctuations) of J will pose a serious challenge in Fig. 11b reveals that there are two local minima and one
for the performance of our adjoint-based steepest-ascent local maximum of J before ts,opt . Also, the magnitude of
optimization routine for finding the optimum switching the numerical and adjoint gradients differ around ts,opt as
time. shown in Fig. 11c, because the chosen perturbation sizes are
In the subsequent sections, we present our findings for not sufficiently small in this interval.
the following three cases: The global and local trends of J agree for ts > 10 days.
Therefore, if one approaches the optimum with an initial
 Linear foam model in a 3D reservoir
guess greater than 10 days, the steepest-ascent method can
 Nonlinear foam model in a 1D reservoir
approach the optimum. However, if the initial guess of ts
 Nonlinear foam model in a 3D reservoir
is less than that of ts,opt , for which the global and local
The case of the linear foam model in a 1D reservoir can trends of J do not agree (Fig. 11b), the correct adjoint
be found in [15]. gradients may not accurately reflect the global trend of J
572 Comput Geosci (2014) 18:563–577

Fig. 9 Schematic of global and Global Trend (Coarse Sampling of ) Local Trend (Fine Sampling of )
local trends of J between the
two black squares. The local
trend of J differs from the
global trend of J and reveals
details that were not noticeable
on the global trend

ts ts

that is necessary for finding ts,opt . Therefore, a gradient- injection rate as the foam front advances further from the
based optimization routine might encounter serious well. For switching times greater than 40 days, foam reaches
problems. the column of gridblocks containing the production well,
and after that point in the flood, there is a large reduction
6.3 Optimization with the nonlinear foam model in a 3D in gas injection rate as foam reduces the productivity index
reservoir of the production well. For switching times between 57 and
66 days, foam is able to sweep some oil from the edges of
We now extend the nonlinear foam model to 3D simula- the formation away from the injection well, without much
tions in the presence of gravity. It is helpful to summarize further reduction in injection or production rate. For switch-
the progress of the flood in 3D. The simulation ends after ing times greater than 66 days, foam is present in the entire
540 days. Therefore, the period of gas injection decreases as formation; increasing the switching time then reduces the
switching time increases. As in all our examples, injection period of gas injection with no further harm to injection rate
and production wells operate at fixed bottom-hole pres- or benefit to sweep efficiency. Details are in [15].
sure. For switching times less than 5 days, foam has a The optimal switching time ts,opt is the switching time
marked negative effect because it reduces injectivity without for which the foam front stops just short of the produc-
much improvement in sweep efficiency. Foam in these cases tion well. For switching times greater than ts,opt , the foam
sweeps only a region near the injection well. For switch- front breaks through to the production well during the pro-
ing times from 5 to 40 days, foam does improve sweep cess, reducing cumulative oil production below its value
efficiency, without too much further negative effect on gas at ts,opt .

Fig. 10 Local and global trends 200 900


of J for the linear foam model Adjoint Gradient
in a 3D reservoir. The total Numerical Gradient
simulation time is 200 days. 800
100
Except for the initial point at
ts = 1 day, the numerical and 700
Time Gradient at ts (m3/day)

analytical (adjoint) gradients are 0


0 10 20 30 40 50
in good agreement. The middle t s (days) 600
part of the plot around the
-100
(m3)

maximum is magnified in the 15 max 870


inset. Both plots share the same 500
axis titles. There is no adjoint 10 850
-200
gradient at ts = 0 day, because 830 400
in that case, we start gas 5
injection from the beginning, -300 810
0 300
and there is no switching time 790
22 27 32 37
-400 -5 770 200

-10 750
-500 sign change 100
min
in gradient
Comput Geosci (2014) 18:563–577 573

Fig. 11 a Global trend of J for


the nonlinear foam model in a a 1.65
0.048
1D reservoir. The total Adjoint Gradient
simulation time is 200 days. b Numerical Gradient
Magnification of the switching 1.6
time interval from 7 to 0.028
7.12 days. A pair of arrows in

Time Gradient at ts (m3/day)


panel a highlights the point ts = 1.55
7 days. c Magnification of the
region around the global 0.008
t s (days)
maximum at ts = 8.25 days. The

(m 3 )
1.5
second pair of arrows in panel a 0 5 10 15 20 25 30 35 40
highlights the point ts = 8 days. -0.012
The adjoint gradient changes
1.45
sign at this point. Panels b and c
share the same axis titles with
panel a; we remove them to -0.032
1.4
avoid clutter. There is no adjoint
gradient at ts = 0 day, because
in that case, we start gas
-0.052 1.35
injection from the beginning,
and there is no switching time
a b
0.01
0.01 1.6254
1.62305
0.005
1.6252
8 8.2 8.4
1.62295 -0.01
0
7 7.04 7.08 7.12 1.625

-0.005 1.62285 -0.03


1.6248

-0.01 1.62275 -0.05 1.6246

In the 3D reservoir with the linear foam model, the 3D reservoir, compared to the 1D reservoir with the nonlin-
gradient-based optimization routine was able to success- ear foam model. We observe a difference in the sign of the
fully find the optimum switching time. However, Fig. 12 two gradients near all the optima, i.e., near 5, 40, 57, and
shows differences between both the sign and the magnitude 66 days. Having major differences between the global and
of the numerical and adjoint gradients at certain switch- local trends of J in the neighborhood of the optimum seri-
ing times. These differences are more pronounced in the ously challenges the optimization routine’s performance.

Fig. 12 Global trend of J for 60 3900


the nonlinear foam model in a Adjoint Gradient
3D reservoir. The total 40 3800
Numerical Gradient
simulation time is 540 days. The
3700
sign of the adjoint and 20
numerical gradients are different
Time Gradient at ts (m3 /day)

ts (days) 3600
near ts = 40 and 66 days. The 0
adjoint gradient and J at these 0 25 50 75 100
3500
two points are indicated by -20
(m3)

arrows. There is no adjoint 3400


gradient at ts = 0 day, because -40
3300
in that case, we start gas
injection from the beginning, -60
3200
and there is no switching time
-80
3100

-100 3000

-120 2900
574 Comput Geosci (2014) 18:563–577

These trends agree only after the last local optimum at slugs large enough to sweep the entire reservoir, (1) there
66 days. We investigate the local trend for two of these are no additional, abrupt changes in gas mobility with
points: the global maximum at ts,opt = 40 days and the local increasing slug size as foam is able to form in another grid-
maximum at ts = 66 days, indicated by arrows in Fig. 12. block, and (2) as it happens, there are no further changes
Figure 13 shows the behavior of J very close to ts = in gridblocks in which foam dries out and collapses with
40 days. Right after this point, the global trend appears to increasing surfactant slug size.
be decreasing, but the adjoint gradient is positive. However,
the local trend around 40 days in Fig. 13a reveals that ts,opt
is at 40.1 days, and not at 40 days. Further, the increasing 7 Summary, conclusions, and recommendations
trend of J just after ts = 40 days (Fig. 13b) is consistent
with the positive sign of the adjoint gradient at this point. In Our objective in this work was to investigate the perfor-
addition, Fig. 13 shows that the numerical gradient is inac- mance of a gradient-based optimization routine applied to
curate at this point; hence, the perturbation size used is not a foam SAG EOR process: specifically, maximizing the
small enough. cumulative oil production by finding the optimum switching
Comparing Figs. 10 and 12 reveals that major differences time (ts,opt ) between the surfactant and gas injection cycles,
between the global and local trends of the objective function in either 1D or 3D, with linear or nonlinear foam models.
and fluctuations in the adjoint gradient are introduced when This work is the first attempt to apply optimal control theory
the linear foam model is replaced with the nonlinear foam to foam EOR processes.
model. We conjecture that in most simulation models, the We show that an inappropriate choice of the relative tol-
fluctuations on the local trend of J and the fluctuations in erance for the adjoint linear solver is a source of incorrect
its gradient are small, so that they remain unnoticed. In our analytical gradients in our problem, but accurate gradients
case with the nonlinear foam model, where small changes were obtained with sufficiently small relative tolerance.
in switching time produce relatively large changes in J , We reach the following conclusions from our analysis of
the effects on the gradients are much more pronounced, to the linear and nonlinear foam models:
an extent that the sign of the gradient can reverse and the
adjoint gradients do not reliably lead to a global optimum.
Linear foam model (3D)
That is, the adjoint method gives correct gradients, but these
correct gradients may not accurately reflect the global trend  The local and global trends of J agree over the entire
of J that is necessary for finding ts,opt . time interval.
Another indication that fluctuations in J and in its gra-  The chosen perturbation sizes (1 day) for constructing
dients derive from the nonlinear foam model is that if the numerical gradients are sufficiently small for accu-
foam covers the entire reservoir during the process (i.e., for rate determination of the gradient of J , and they match
ts ≥ 10 days in Fig. 11 in 1D and for ts ≥ 70 days in the adjoint gradients.
Fig. 12 in 3D), all the fluctuations in J disappear, and the  The gradient-based optimization routine is capable of
local and global trends agree. In other words, for surfactant finding the optimum switching time.

Fig. 13 Trend of J around


ts = 40 days. a J versus the
switching time. There are 17
switching times (column 2 of
panel c) in this figure. However,
since the middle points are very
closely spaced, their trend is not
visible in this figure. The correct
ts,opt is at 40.1 days, and not at
40 days. b J versus time points
identified in column 1 of panel c
in the vicinity of 40 days (i.e.,
time point 9). c Table of
switching times and time points
used to construct panels a and b

a b c
Comput Geosci (2014) 18:563–577 575

Nonlinear foam model (1D and 3D) of the global trend of J . As a result, in the cases studied
in this work, the numerical gradients would be preferred
 Replacing the linear foam model with the nonlinear
over the adjoint gradient to be used in the steepest-
foam model introduced major differences between the
ascent method for finding the optimum switching time
local and global trends of J and fluctuations in the
with the nonlinear foam model.
adjoint gradient at some switching times in both 1D
and 3D simulations. The local and global trends agree,  The appropriate choice of the perturbation size is case-
and the adjoint gradient is free from fluctuations only at dependent.
switching times for which foam sweeps the entire reser-  We believe that both the strong oscillations in the for-
voir within the simulation period. In those cases, the ward solution (as described in Section 4) and the small-
gas mobility reduction factor makes no further abrupt scale fluctuations in the objective function (as described
changes until the end of the simulation once surfactant in Section 6) are caused by the strong variations in gas
concentration exceeds the minimum for foam creation mobility as described in the nonlinear foam model. It
in all the reservoir gridblocks. may well be that the encountered problems can be elim-
 In our 1D case, gradient-based optimization is not suit- inated by the use of a different spatial discretization
able for finding the optimum switching time, unless the scheme in the simulator, which is robust against strong
initial guess is larger than ts,opt . In our 3D case, there variations in gas mobility. However, verifying this and
were major differences between the global and local finding a more robust discretization scheme requires
trends of J in the neighborhood of the optima that further investigation.
would seriously challenge the optimization routine’s
The sort of challenges that we encountered with the non-
performance. As a result, gradient-based optimization
linear foam model in foam EOR processes could be also
is not suitable for finding the optimum switching time present in modeling other EOR processes. Having abrupt
in this case. changes in properties may cause similar symptoms chal-
 We conjecture that in most simulation models, the fluc- lenging gradient-based optimization routines in these appli-
tuations on the gradient of J are small, so that they cations. Obviously, if these symptoms are also observed
remain unnoticed. In our case with the nonlinear foam in those applications, gradient-based optimization routines
model, where small changes in switching time produce might not be suitable, and other options must be considered
relatively large changes in J , the effects on the gradi- for optimizing these processes.
ents are much more pronounced, to an extent that the All the complexities and challenges we identified are
sign of the gradient can reverse and the adjoint gradients
for a case in which only one control variable is involved.
do not reliably lead to a global optimum. That is, the
In cases with multiple control variables, a gradient-based
adjoint method gives correct gradients, but these correct optimization routine might face additional challenges.
gradients may not accurately reflect the global trend of
J that is necessary for finding ts,opt .
Acknowledgments This research was carried out within the context
 The chosen perturbations (i.e., 1 day) for construct- of the Integrated Systems Approach to Petroleum Production (ISAPP)
ing the global trend of J in the fluctuating zones Knowledge Center. ISAPP is a joint project between Delft Univer-
were not sufficiently small to obtain accurate gradients, sity of Technology (TU Delft), Shell International Exploration and
Production (SIEP), and the Dutch Organization for Applied Scientific
except for large switching times, for which the local and Research (TNO). We thank Shell Global Solutions International for
global trends of J were in agreement. Nevertheless, the allowing us to use their in-house reservoir simulator MoReS in this
numerical gradients gave an acceptable representation research.

Table 1 Model parameters.


Foam parameters are those for Swr Sorw Sorg Sgr μw (cp) μo (cp) μg (cp)
the nonlinear foam model 0.1 0.3 0 0 Computed by simulator 5 0.02

Krw Krow Krog Krg


0.3 0.8 0.8 0.94

nw now nog ng
4 2 3 3

fmmob fmdry (S∗w ) epdry fmoil (S∗o ) epoil floil


1,000 0.316 1,000 0.4 1.5 0
576 Comput Geosci (2014) 18:563–577

Table 2 Reservoir and well


parameters for the 3D Length (m) Width (m) Height (m) Depth (m) x (m) y (m) z (m)
simulations 100 100 30 1,600 10 10 3

ϕ kx (mD) ky (mD) kz (mD)


0.2 100 100 10

inj prod
Pref (bar) Pwf (bar) Pwf (bar)
165 250 145

Appendix: Model and simulation parameters where λ0rg is the foam-free relative gas mobility and F w , F o
and F s are given by
Table 1 summarizes the values for the residual saturations, 
arctan epdry(Sw − fmdry)
viscosities, and other parameters in the relative permeability Fw =
and foam models. For the two-phase relative permeability  π
arctan epdry(Swr − fmdry)
of a water/oil system (Sg = 0), we have −
  nw π
Sw − Swr
krw = Krw , (A-1a) ⎧
1 − Swr − Sorw ⎪ 1 So < floil
  ⎨ epoil
1 − Sw − Sorw now Fo = fmoil−So
floil ≤ So < fmoil (A-4)
krow = Krow , ⎪ fmoil−floil
1 − Swr − Sorw ⎩
0 fmoil ≤ So ≤ (1 − Swr )
and for the two-phase relative permeability of a gas/oil
 
system (Sw = Swr ) in the absence of foam, we have Fs = tanh 800 (800Cs )130 (A-5)
 ng
Sg − Sgr
krg = Krg , (A-1b) Equation (A-3) differs from the conventional “dryout”
1 − Swr − Sgr − Sorg function in STARS [4] in that foam is completely destroyed
 
1 − Sg − Sorg − Swr nog at Swr ; for SAG foam processes, this is critical [16]. At other
krog = Krog .
1 − Swr − Sgr − Sorg saturations, Eq. (A-3) is similar to the function in STARS.
The three-phase relative permeabilities are calculated In [16], F w and F o are referred to as the “water-killing”
from the linear isoperm model [2]. That is, krw and krg are and “oil-killing” foam models. Equation (A-5) gives an
defined as above, and oil relative permeability kro is calcu- abrupt reduction in gas mobility at surfactant concentra-
lated using linear interpolation of the two-phase oil relative tion Cs approximately equal to C∗s = 0.0012, which is
permeabilities krow and krog above. half the injected surfactant concentration of 0.0024 (weight
In the presence of foam, gas mobility is altered from the fraction). Unlike the corresponding function in STARS, it
foam-free value. Foam parameters in Table 1 are for the is differentiable at all values of Cs . It is not an accurate
nonlinear foam model (Fig. 1); the simple functions for the description of foam strength as a function of surfactant con-
linear foam model are easily inferred from Fig. 1. The non- centration [1], but it helps to reduce the effects of numerical
linear functions are related to those in the STARS foam dispersion on foam simulations [21].
simulator (Computer Modeling Group, Calgary, Alberta, Table 2 gives the reservoir and well parameters for the
Canada) and used by [16]: 3D simulations. The same parameters and properties apply
to 1D simulations except the length and number of grid-
λ0rg blocks: the 1D reservoir is 100 m × 1 m × 1 m, with 100
λrg = (A-2)
1 + fmmobFw Fo Fs × 1 × 1 gridblocks (i.e., x = y = z = 1 m), unless

Table 3 Time-step strategy for


1D and 3D simulations in Time interval 1D 3D
Section 6
tstart (day) tmax (day) tstart (day) tmax (day)

0 ≤ t ≤ (ts − 0.5) 1E-6 1E-2 1E-6 1E-1


(ts −0.5) < t ≤ (ts + 10) 1E-4 1E-4 5E-3 5E-3
(ts + 10) < t ≤ T 5E-3 5E-2 5E-3 1E-1
Comput Geosci (2014) 18:563–577 577

otherwise is stated. The prescribed injection well pressure SPE 105764 presented at the SPE Reservoir Simulation Sympo-
is 250 bar, and the prescribed production well pressure is sium, Houston, Texas, 26–28 February (2007)
11. Lei, Y., Li, S., Zhang, X., Zhang, Q., Guo, L.: Dynamic opti-
145 bar; the initial reservoir pressure is 165 bar. The well
mization of a polymer flooding process based on implicit discrete
injectivity and productivity indices are calculated using the maximum principle. Math. Problems Eng. Article ID 281567
Peaceman model. (2012). doi:10.1155/2012/281567
We applied slightly different time-step sizes for 1D and 12. Liu, W., Ramirez, W.F.: Optimal control of three-dimensional
steamflooding processes. J. Pet. Sci. Eng. 11(2), 137–154 (1994)
3D simulations: to avoid convergence problems, 1D simu-
13. Liu, W., Ramirez, W.F., Qi, Y.F.: Optimal control of steam-
lations required smaller t because of smaller gridblocks flooding. SPE Adv. Technol. Ser. 1(2), 73–82. SPE-21619-PA
compared to 3D simulations (Table 3). (1993)
14. Mehos, G.J., Ramirez W.F.: Use of optimal control theory to opti-
mize carbon dioxide miscible-flooding enhanced oil recovery. J.
Pet. Sci. Eng. 2(4), 247–260 (1989)
References 15. Namdar Zanganeh, M.: Simulation and optimization of foam EOR
processes. PhD dissertation, Delft University of Technology, Delft
(2011)
1. Apaydin, O., Kovscek, A.R.: Surfactant concentration and end 16. Namdar Zanganeh, M., Kam, S.I., LaForce, T.C., Rossen, W.R.:
effects on foam flow in homogeneous porous media. Transp. The method of characteristics applied to oil displacement by foam.
Porous Media 43(3), 511–536 (2001) SPE J. 16, 8–23. SPE-121580-PA (2011)
2. Baker, L.E.: Three-phase relative permeability correlations. Paper 17. Oliver, D.S., Reynolds, A.C., Liu, N.: Inverse Theory for
SPE 17369 presented at the SPE Enhanced Oil Recovery Sympo- Petroleum Reservoir Characterization and History Matching.
sium, Tulsa, Oklahoma, 16–21 April (1988) Cambridge University Press, Cambridge (2008)
3. Bryson, A.E., Ho, Y.-C.: Applied Optimal Control. Taylor and 18. Ramirez, W.F.: Application of Optimal Control Theory to
Francis (Hemisphere), Levittown (1975) Enhanced Oil Recovery, Vol. 21. Elsevier, Amsterdam (1987)
4. Cheng, L., Reme, A.B., Shan, D., Coombe, D.A., Rossen, W.R.: 19. Ramirez, W.F., Fathi, Z., Cagnol, J.L.: Optimal injection poli-
Simulating foam processes at high and low foam qualities. Paper cies for enhanced oil recovery: part 1—theory and computational
SPE 59287 presented at the 2000 SPE/DOE Symposium on strategies. SPE J. 24(3), 328–332. SPE-11285-PA (1984)
Improved Oil Recovery, Tulsa, OK, 3–5 April (2000) 20. Rossen, W.R.: Foams in Enhanced Oil Recovery. In: Foams: The-
5. Fathi, Z., Ramirez, W.F.: Optimal injection policies for enhanced ory, Measurement, and Applications. Prud’Homme, R.K., Khan,
oil recovery: part 2—surfactant flooding. SPE J. 24(3), 333–341. S. (eds.) Marcel Dekker, New York, pp. 413–464 (1996)
SPE-12814-PA (1984) 21. Rossen, W.R., Zeilinger, S.C., Shi, J.-X., Lim, M.T.: Simplified
6. Fathi, Z., Ramirez, W.F.: Use of optimal control theory for mechanistic simulation of foam processes in porous media. SPE J.
computing optimal injection policies for enhanced oil recovery. 4(3), 279–287 (1999). SPE 57678-PA
Automatica 22(1), 33–42 (1986) 22. Shan, D., Rossen, W.R.: Optimal injection strategies for foam
7. Fathi, Z., Ramirez, W.F.: Optimization of an enhanced oil recov- IOR. SPE J. 9(2), 132–150 (2004). SPE 88811-PA
ery process with boundary controls—a large-scale non-linear 23. Sudaryanto, B.: Optimization of displacement efficiency of oil
maximization. Automatica 23(3), 301–310 (1987) recovery in porous media using optimal control theory. PhD
8. Jansen, J.D.: Adjoint-based optimization of multi-phase flow Dissertation, University of Southern, California, LA (1998)
through porous media—a review. Comput. Fluids 46(1), 40–51 24. Van Doren, J., Douma, S., Wassing, B., Kraaijevanger, H., de
(2011) Zwart, B.-R.: Adjoint-based optimization of polymer flooding.
9. Kourounis, D., Durlofsky, L.J., Jansen, J.D., Aziz, K.: Adjoint Paper SPE 144024 presented at the SPE Enhanced Oil Recovery
formulation and constraint handling for gradient-based optimiza- Conference, Kuala Lumpur 19–21 July (2011)
tion of compositional reservoir flow. Comput. Geosci. (2013). 25. Van Essen, G.M., Jansen, J.-D., Brouwer, D.R., Douma, S.G.,
doi:10.1007/s10596-013-9385-8 Zandvliet, M.J., Rollett, K.I., Harris, D.P.: Optimization of smart
10. Kraaijevanger, J.F.B.M., Egberts, P.J.P., Valstar, J.R., Buurman, wells in the St. Joseph Field. SPE Res. Eval. Eng. 13(4), 588–595.
H.W.: Optimal waterflood design using the adjoint method. Paper SPE-123563-PA (2010)

You might also like