You are on page 1of 31

SPE-173213-MS

INSIM: A Data-Driven Model for History Matching and Prediction for


Waterflooding Monitoring and Management with a Field Application
Hui Zhao, Yangtze University; Zhijiang Kang, China Petroleum and Chemical Corporation; Xiansong Zhang,
CNOOC Research Center; Haitao Sun, and Lin Cao, Yangtze University; Albert C. Reynolds, U. of Tulsa

Copyright 2015, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE Reservoir Simulation Symposium held in Houston, Texas, USA, 2325 February 2015.
This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
We derive and implement an interwell numerical simulation model (INSIM) which can be used as a
calculation tool to approximate the performance of a reservoir under waterflooding. In INSIM, the
reservoir is characterized as a coarse model consisting of a number of interwell control units, where each
unit has two specific parameters: transmissibility and control pore volume. By solving the mass material
balance and front tracking equations for the control units, the interwell fluid rates and saturations are
obtained so that phase producing rates can be predicted. INSIM is applied to perform history matching for
parameter estimation and to infer the interwell connectivity and geological characteristics. INSIM has the
following advantages: (1) the model parameters estimated from history matching provide a relative
characterization of interwell formation properties. The model can handle changes in the flow directions
caused by changing well rates, including shutting in wells or converting producers to injectors, whereas
with the common correlation-based interwell connectivity method, the well interactions are assumed to be
fixed; (2) the previous methods that bear computational similarity to INSIM can only provide the total
liquid production rate whereas, with our procedure, we can calculate the oil and water flow rates and hence
history-match water-cut data; (3) because we can calculate the oil and water flow rates, our method can
be used for waterflooding optimization but with far less computational effort than with the traditional
method based on the use of a reservoir simulator.

Introduction
Production and injection rate data are the most abundant data available in any waterflooding operation.
Analysis of these data can provide useful information about interwell formation characteristics. Rock
property fields, combined with well rates, have a dominant effect on the directions of fluid flow (stream
lines) in a reservoir. History matching with a reservoir simulator is the most common way to condition
the rock property fields to production data. However, production data are never sufficient to resolve the
reservoir properties, e.g., gridblock permeabilities, and few assisted history matching tools exist in
commercial reservoir simulators. Consequently, when a reservoir simulator is used as the forward model
when history matching, the number of reservoir parameters is often reduced to a small number based on
computational experiments and physical insight. Although, the INSIM methodology introduced here does

SPE-173213-MS

limit the number of history matching parameters, the primary objective of INSIM is to provide a fast,
simplified simulation model to calculate flow and transport sufficiently well so that when INSIM is used
for history matching, the resulting history-matched model can be input to INSIM to provide reasonable
future predictions as well as provide information on the flow dynamics of the reservoir. It is hoped that
the model and methodology presented here will prove useful for monitoring and understanding waterflooding operations conducted on a black oil reservoir and that INSIM will ultimately be useful for water
flooding optimization, which will be the subject of a future paper.
Models based on the statistical correlation or the connectivity between injectors and producers
estimated from flow rate data have been used previously to characterize reservoirs for the purpose of
waterflooding management. Specifically, the correlation coefficient from the Spearman rank analysis
method has been used to estimate the interactions between injector-producer well pairs and associate these
interactions with geomechanics (Heffer et al., 1995; Refunjol, 1996; Jansen and Kelkar, 1997; Panda and
Chopra, 1998; Araque-Martinez, 1993; Barros-Griffiths, 1998). Albertoni and Lake (2003) applied
multivariate linear regression and diffusivity filters to estimate connectivity between injectors and
producers where filters were applied to account for the time lag and attenuation of injection rates. The
estimated linear coefficients or weights obtained provide a quantitative estimate of the strength of
communication between a producer and the injectors and provide a qualitative understanding of reservoir
heterogeneity. The Capacitance Mode (CM) developed by Yousef et al. (2006) and further improved by
Kaviani et al. (2008) and Sayarpour (2008) is a more complex model that directly generalizes the
Albertoni and Lake model. The main difference between the CM model and the Albertoni and Lake
(2003) model is that the CM model includes the capacitance (compressibility) as well as resistivity
(transmissibility) effects, whereas the Albertoni and Lake (2003) model neglects compressibility. In the
CM model, each well pair contains two parameters; one parameter is the coefficient that quantifies the
connectivity and another (time constant) quantifies the degree of fluid storage between wells. The CM
model can integrate bottom-hole pressure (BHP) data to improve the quantitative estimates of well
connectivity. Similar to CM, a systemic analysis model was proposed by Hui et al. (2010) in which the
injection-production system is characterized as a first-order linear time-delay system. In the Hui et al.
(2010) model, the number of time-constant parameters is decreased to one-half of the number needed in
the CM but the Hui et al. (2010) model cannot use BHP data. Kaviani et al. (2010) developed a multiwell
productivity index (MPI)-based method to calculate the interwell connectivity from injection-production
data; in this method, a heterogeneity matrix is obtained, and this matrix represents the heterogeneity and
possible anisotropy of the formation. Dinh and Tiab. (2007) and Lee et al. (2009) applied multivariate
linear regression on BHP data measured at producers and water injectors to estimate the interwell
connectivity. In addition, Gherabati et al. (2012) used well locations and flow rate data to develop a
reservoir-scale network model to estimate the formation conductance values between injector and
producers. The conductance value, which is similar to the parameter transmissibility proposed in our
method, was assumed to be constant since they did not consider the fact that the phase mobilities change
with time, while in our method, the total liquid mobility, which is included in the transmissibility, varies
with time.
Compared to a conventional reservoir simulator, the correlation-based models are computationally
inexpensive because they solve for far fewer state variables and also involve far fewer parameters. Thus,
assisted history matching with a correlation-based model is far faster than assisted history matching with
a reservoir simulator regardless of whether gradient-based methods are used (Bissell et al., 1994; Anterion
et al., 1989; Wu et al., 1999; Zhang and Reynolds, 2002; Li et al., 2003; Gao and Reynolds, 2006) or
ensemble-based methods are used (Evensen and van Leeuwen, 2000; Evensen, 2003, 2007; Nvdal et al.,
2005; Gao et al., 2006; Aanonsen et al., 2009; Wang et al., 2010; Emerick and Reynolds, 2012, 2013;
Chen and Oliver, 2013). However, the major deficiencies of correlation-based models are also obvious:
(1) The inversion parameter that represents the connectivity has no precise geological meaning since it

SPE-173213-MS

only reflects the relative relationship or weight between wells and cannot be directly related to an
interwell formation property like the average permeability or transmissibility. This weight is normally viewed as an al1ocation factor of the injection
rate and assumed to be a constant value. This constant value assumption is not true; in particular,
Thiele and Batycky. (2006) have shown that the
al1ocation factor is a dynamic parameter which
changes with flow pattern and injection rates; (2)
The correlation-based models consider the interaction between each injector and producer pair but
neglect the interaction between pairs of producers
Figure 1Illustration of modeling of volume flow units connections
that are in communication. Thus, correlation-based between wells.
models are not applicable when rates change significantly, for example, if a producer is shut in for a long period or one or more producers is converted to
an injector. Therefore, the inversion results are not robust to changes in operational strategy; (3) The
predictive ability of such models is relatively poor as they can output only a total liquid production rate
but cannot predict water cut or individual phase rates. Thus, one cannot use such a model to history match
water-cut data which is a severe handicap because water-cut data contains significant information on
connectivity and reservoir heterogeneity. The ability of INSIM to predict water cut and phase rates is the
most important innovation included in INSIM.
Unlike previous correlation-based models, INSIM is able to effectively predict the water cut and oil
production rate and hence can be used as the forward model for assisted history matching of these data.
Specifically, the model can be used in automatic history matching. Moreover, the model is derived directly
from the correct two-phase flow mass balance equations and thus the transmissibilities derived from
history matching reflect an average transmissibility between wells. In addition, because INSIM is based
on simulation flow equations, it can incorporate large changes in flow rates, flow directions and injector
al1ocation factors, the interaction between pairs of producers and the conversion of producers to injectors.
At the same time, INSIM retains the computational efficiency of previous correlation-based models which
incur far less computational cost than a traditional numerical reservoir simulator. We show that INSIM
has good predictive ability so that it may also be applied for life-cycle production optimization and
closed-loop reservoir management. The application to production optimization will be presented in a
subsequent paper.

Interwell Numeric Simulation Model (INSIM)


In INSIM, we first consider the reservoir as a network model consisting of a series of units connecting
well pairs. Unlike the correlation-based models, INSIM allows not only for injector-producer connections
but also for injector-injector and producer-producer connections. Thus, it is expected that INSIM can
better resolve the flow that occurs in the interwell region between pairs of wells of the same type.
As shown in Fig. 1, well node i is assigned a volume denoted by Vp,i which is depicted by the dashed
red circle in Fig. 1. At present, we only consider fully-penetrating vertical wells and two-dimensional
flow; thus, in Fig. 1, Vp,i is the cylindrical volume enclosed by the red-dashed circle extended over the
reservoir thickness. Flow in the interwell area between well i and its connected well nodes is modeled as
flow within a region characterized by two parameters, transmissibility (Ti,j) and a control pore volume
(Vp,i,j), where the transmissibility controls flow between well j and well i in the interwell volume Vp,i,j. For
the schematic in Fig. 1, j 1,2,3 but there is no limit to how many wells can be connected to well i;
however, the connections have to be defined before any computations can be done. For each single well

SPE-173213-MS

Figure 2Determination of connections between well pairs.

node i, there is a well-controlled pore volume, Vp,i, where the summation of all Vp,is should be equal to
the pore volume of the reservoir. Practically, we can generally use the area enclosed by linking all the
middle points between well i and its connected adjacent wells to define well is controlled area and hence
define Vp,i. We let pi represent the average pressure on Vp,i.
The connections between wells must be specified a priori in INSIM. For correlation-based models,
Kaviani et al. (2010) proposed an approach called windowing to set well connections according to the well
distance. They define a radius for the window of each well where, in their cases, the same radius is used
for all wells. Conceptually, one can define a different connection radius for each well, but such an
approach must be used and coded with extreme caution. In particular, if in defining the connections to well
node i, we introduce a connection between well k and well i, then when defining the connections to well
k, we must include a connection to well k. Here, we simply use the same connection radius for each well.
If the well j is inside the window (connection-radius) of well i, then wells i and j are connected; otherwise
they are not. We can generally take two to three times the reservoir average well spacing or two to three
times the individual smallest well distance as the radius for the window depending on the specific well
configurations. For illustration, consider the five-well case shown in Fig. 2(a). Li,j Lj,i denoted the
distance between well i and well j. The minimum distance between any two connected well pairs is L4,5
and we take 3 times this distance as the window radius. Based on this definition, all pairs of wells are
connected and these connections are represented by the lines connecting well pairs in Fig. 2(a). However,
it is unlikely that all these connections are necessary. Specifically, note that wells 2, 5 and 4 are almost
collinear; thus, we expect that there is a relatively small direct interaction between well 2 and well 4. Thus,
the connection between wells 2 and 4 is deleted. By a similar argument, the direct connection between
wells 1 and 3 is deleted. After windowing, we are left with the connections shown in Fig. 2(b). In the
actual connection determination process implemented in code, we track each triangle formed by the set
of initial connections to the well node (Fig. 2(a)); if two inner angles of such a triangle are both lower than
a specified value (we usually use 15 degrees as the tolerance), the single linear connection between the
wells that is involved with these two inner angles is deleted. With this approach, the connections of Fig.
2(a) are replaced with the connections shown in Fig. 2(b).
s and
s, which
After the connections are determined, the model parameters are defined as the
respectively represent interwell volumes and connection transmissibilities at time zero. To start the history
matching algorithm discussed later, initial guesses for the parameters must be generated based on
whatever geological information is available. Here we give a simple initialization process that requires
estimates of the permeability and porosity fields. If we have the average formation properties of porosity

SPE-173213-MS

Figure 3Fractional flow and its derivative cure.

Figure 4 Saturation distribution under a change in flow direction.

( i), permeability ( i) and net thickness ( i) for each well point, we use the average values at two
connected wells as the initial average parameter for this connection. For example, if wells i and j are
connected, then i,j 0.5( i j); i, j and i,j are defined by the same averaging procedure. The term
as follows:
i,j i,jLi,j is a weight to determine the initial guess for
(1)
for i 1, 2, . . . nw, where nw is the total number of well nodes and VR is the reservoir pore volume.
is calculated by
The initial guess for the
(2)

for i 1, 2, . . . nw and j i 1,i 2, . . . nw; m, m 0, w, is viscosity of phase m; is a units


and
conversion factor where 0.0864 in the system of units used. It is noted that the superscript 0 in
represents the initial time zero as the transmissiblities and volumes change with time in our proposed

SPE-173213-MS

Figure 51D reservoir model.

method. The initial guesses for

and

are adjusted by history matching, i.e, the reservoir variables

history matched are


and
for i 1, 2, . . . nw and j 1, 2, . . . nw.
Material balance equation and pressure calculation
We consider only the two-phase flow of oil and water and set Ti,j as the average total transmissibility
between the ith and jth well. Then, the total mass balance (oil plus water) (neglecting capillary pressure
and gravity) for the ith well is given by
(3)
where nw is the total number of wells; qi(t) in m3/d is the bottom-hole fluid rate of the ith well at time
t (days) and is positive for injection and negative for production; ct,i in units of MPa1 is the total
compressibility for Vp,i. If well j is not connected to well i, we set Ti,j 0 in Eq. 3. Eq. 3 simply represents
the combined pressure equation. Approximating the equation by the implicit finite-difference scheme used
in reservoir simulation gives
(4)
for i 1,2, . . . nw, where tn tn tn1 and t0 0. The transmissibilities ( s), well drainage pore
s) and compressibilities ( s) may all vary with time and are defined below. In our current
volumes (
version of INSIM, similar to the IMPES pressure equation in reservoir simulation, we evaluate nonlinear
terms at the old time level, i.e., terms which depend on pressure and/or water saturation are evaluated at
time tn1 instead of tn. Specifically, in Eq. 4, we use
(5)

(6)
and
(7)
respectively, in place of
,
and
, respectively. Here, So,i and Sw,i are the corresponding oil and
water saturations in the volume of well i; co, cw and cr, respectively, represent oil, water and rock
compressibility; t,i,j is the total mobility, which is calculated using upstream weighting. Specifically, if
, then t,i,j is replaced by the total mobility of well i, t,i. Throughout, we assume oil and water
,
viscosities are constant. Upstream weighting then means that, when
(8)
otherwise,
(9)

SPE-173213-MS

where 0 and w, respectively, denote oil and water viscosity. The


following equation:

s are be updated using the


(10)

Defining
(11)
(12)
and
(13)
the system of equations given by Eq. 4 can be written as
(14)

where
is positive for an injection well and negative for a production well. The system of equations
given by Eq. 14 can be solved for the s assuming the necessary saturations at tn1 are known. After
s, we need to solve fractional flow equations to obtain for the water saturations at
solving for the
tn. Similar to IMPES in reservoir simulation, we alternate solving for pressures and saturations until we
reach the final time of interest.
In Eq. 14, the number of pressure equations is equal to the number of well nodes, whereas in traditional
reservoir simulation, the number of pressure equations is equal to the number of gridblocks. Therefore,
running INSIM requires far less computational time than would be required to run a conventional
black-oil reservoir simulator. After the pressures are calculated, saturations are estimated by the procedure
described in the next section. Then, it is easy to calculate the flow rate between well pairs which can be
used for further diagnosis of well-to-well dynamics. Because the number of well nodes in limited, the
methodology can be expected to give only a rough description of the reservoir fluid distributions.
However, our computational results to date suggest that saturation/pressure distributions calculated from
the history-matched INSIM model are sufficient to give a useful production forecast.
Saturation tracking and well dynamic data prediction
Since the reservoir is modeled as a series of control units between well pairs, the saturation tracking for
each control unit can be viewed as a one-dimensional problem. For oil-water two phase flow, BuckleyLeverett theory provides a relationship between the cumulative water injected and the location of the flood
front. In the traditional way, Buckley-Leverett is used for saturation tracking from injector to producer.
In the computation of the saturation distribution in INSIM, standard Buckley-Leverett is applied if the
flow direction along a connection between a well pair does not change. However, for the case where flow
reversals occur due for example to the conversion of a producer to an injector, we apply the BuckleyLeverett computations in a very ad hoc way to track the flow of injected water. We also use BuckleyLeverett to track flow from producer to producer. This last option is useful when two producers, P-a and
P-b, are connected but only one of these producers, say P-a, is connected to a certain water injector, I-a.
In this scenario, injected water flows along the connection from well I-a to P-a until breakthrough. After
breakthrough, if all of the water cannot be produced by well P-a, in the INSIM model, some of this water

SPE-173213-MS

Figure 6 Water saturation profile by ECLIPSE and INSIM at different times.

will flow along the connection from producer P-a to producer P-b if the pressure at P-b is lower than the
pressure of producer P-a. Such a scenario could occur if the reservoir contains only the five wells depicted
in Fig. 2(b), where we assume that well 2 is a water injection well and all other wells are producers. In
this case, the only way injected water can reach well 4 is through other well nodes. Also, in this case, more
accuracy can be gained by adding a pseudo-well-nodes producing at zero rate and located half way along
connections between the pairs of connected producers. Adding the pseudo-wells allows fluid between two
producers to flow to either active producer via the connection between the two wells that pass through the
inactive pseudo-well between them. Similarly, if two injectors are connected, accuracy is improved by
adding a zero-rate pseudo-well at the midpoint of the line segment connecting the two injectors.
Our first task is to rewrite the standard Buckley-Leverett equation used for computing the water
saturation profile in a form more useful for implementation with INSIM. Throughout,
denoted the
cumulative volume of water injected at reservoir conditions (up to and including time tn) through well
node i. If the cumulative fluid injected at a specific time, t tn, is
, the position xn of a specific water
saturation Sw at tn satisfies
(15)
provided that Sw Swf, where Swf is the saturation at the front and xin is the position at which water
is injected, i.e., the starting inflow coordinate; is porosity; A is the formation cross-section area; fw is
the water fractional flow (water cut), which is shown schematically in Fig. 3 and is given by
(16)

Eq. 15 provides the basic equation for the propagation of water saturation along the node connection
between two well nodes, but to use this equation in INSIM, we will have to rewrite the equation in terms
of volumes involved in the INSIM calculaton; see Fig. 1. Suppose that a point satisfies x so that
is upstream of x. Letting
and
, respectively, denote the saturation and water cut at at time
. From Eq. 15, it follows that
tn, then it follows that
(17)
where the subscript i is introduced to indicate that the water in injected at well node i. Subtracting Eq.
17 from Eq. 15 yields

SPE-173213-MS

Figure 7Water saturation for W3 at different time.

(18)
We let
denote the cumulative volume injected at well i from time t0 0 to tn divided by the
connected pore volume from xu to x, i.e.,
(19)
Substituting

into Eq. 18 and solving for the derivative of the fractional flow at
gives
(20)

More detailed information on the computation of


, then
1 Sor,
is a water injection well at

will be provided later. If xin, i.e., if there


and Eq. 20 corresponds to standard

Buckley-Leverett theory. If
and x correspond to the endpoints of a line segment connecting two
producers, Eq. 20 provides a procedure to obtain the saturation in the well volume centered at x, from its
upstream saturation at xu. However, when x and xu correspond to two producing wells, we will not begin
to apply Eq. 20 until the time t0 when the water saturation at xu exceeds the front water saturation predicted
by Buckley-Leverett theory, i.e., in applying Eq. 19, time zero corresponds to the time t0 when water
begins to flow along the connection between the producer at xu and the producer at x. With these
modifications, Eq. 20 is utilized to track the flow of water between two producing wells.
Eq. 20 indicates that the water-cut derivative (derivative of the water fractional flow function) at a
given point is equal to its value at an upstream point plus the reciprocal of the cumulative number of pore
volumes of fluid that has been based on the pore volume between these two points. As shown later, this
means the derivative of water fractional flow at a given well can be calculated from the water cut
derivatives at all the upstream wells that are connected to it by our model; see Figs. 1 and 2.
,
By solving Eq. 4, all average pressures at time tn for each well volume are determined. For each
we consider all its connected wells and determine which of the connected wells are upstream of the well
. For each well j which is connected to well i (see Fig. 1), well j is upstream of well
corresponding to
n
. According to Eq. 20, we can obtain
, which
i at t if and only if
represents the water-cut derivative at well i at time tn due to flow from well j, from the following equation:

10

SPE-173213-MS

Figure 8 The permeability distribution with four wells.

(21)
where
and
represents the water saturation in the volume associated with
upstream well j. To compute the cumulative pore volume involved in Eq. 21, we first approximate the
flow rate from well j to well i by
(22)
and then replace the pore volume involved in the denominator of the right side of Eq. 19 by the pore
volume between wells i and j to obtain
(23)
so that
is the number of pore volumes of fluid that flows from well j to well i from time t0 to time
t . For well i, if the saturation of upstream well j is lower than Swf, then regardless of the value of
,
.
we set
After calculating the derivative on the left-hand side of Eq. 21, we can directly find a value of Sw,i,j;
see Fig. 3. Then we compute the fractional flow at the well node i as the following flow-weighted average:
(24)
n

where ni is the number of upstream nodes connected to well i. Once


is calculated, we can invert
.
to obtain the updated value of
During the simulation, a well that is upstream of well i may become downstream of well i if (i) well
i is shut in, (ii) infill wells and their connections are added, (iii) a producer is converted to an injection
well, or (iv) producing and/or injections rates are changed substantially. Next, we present an ad hoc way
to track saturations after an operational change. (It is hoped that this ad hoc procedure can eventually be
replaced with a more rigorous front-tracking scheme.) To illustrate the type of situation of interest,

SPE-173213-MS

11

Figure 9 The relative permeability curve.

Figure 10 The rate schedule curve.

consider Fig. 4. In Fig. 4(a), well j and well i, respectively, are located at positions 0 and Li,j, respectively;
we assume that well j is upstream of well i from time 0 to tn= and during this time, water is transported
from well j to well i, either because well j is an injection well or because water flows from well node j
to well i due to an injection well that is connected to well j. At tn=, for some reason, e.g., well i is shut
in, the flow direction changes and well i becomes upstream of well j. The red curve is the saturation
distribution at tn= and the blue curve is the distribution at some time tn after the time tn= at which the flow
reversal occurs. Well js saturation at tn=
is reduced to
at tn due to flow of water from well i to
well j during the time interval (tn=, tn]. Assume
equals the water saturation of position
at tn= (see
red curve), then, according to the Buckley-Leverett equation, we have

12

SPE-173213-MS

Figure 11The initial and final estimated geological parameters, in each set of brackets, the 1st value is transmissibility and the 2nd value is control
pore volume, 104m3.

(25)
where the point
is downstream of well j for the time interval (tn=, tn] so we can calculate the
based on well js saturation at time tn= by Eq. 20, i.e.,
water-cut derivative at
(26)

Substituting Eq. 25 into Eq. 26 and simplifying gives


(27)
Solving for the derivative of the fractional flow at

yields
(28)

where
time tn=;

is the number of pore volumes of fluid that flows from well j to well i from time zero to
is the number of pore volumes of fluid that flows in the opposite direction, i.e., from well

i to well j, from time tn= to tn. If the denominator of Eq. 28 becomes zero or negative, we set
to
a very large number, say 108. As shown later, this derivative is equal to the first term in braces on the right
is set equal
side of Eq. 35 so this first term is set to a very large number which means that
to the second term on the right side of Eq. 35.
At any time qi,j qj,i. At each time step and for each connected well pair, we calculate the number
represents the number of
of flowing fluid pore volumes in both directions in INSIM. Recalling that
pore volumes of fluid that flows from i to j from time zero to time tn, it follows that

SPE-173213-MS

13

Figure 12The objective function value versus iteration number.

(29)
Using Eq. 29, Eq. 28 can be written as
(30)
Because

is larger than

, it follows that

, so

and the saturation at well

j will decrease over the time interval (tn=, tn] as it should be based on the schematic depiction of Fig. 4(a).
becomes less than or equal to 0, we simply set the right side of Eq. 30 to a very large
Again, if
number.
Fig. 4(b) considers the case where well i is changed from a producer to an injector at tn=. In this case,
1 Sor; in this scenario, the true
we assume the water saturation of well i at tn tn= is given by
saturation distribution between the two wells should look qualitatively like the combination of the dashed
and blue curves in Fig. 4(b). According to Eq. 21, based on well i, we also have
(31)
Eq. 31 has meaning only if the calculated value of

corresponds to a value of

the Buckley-Leverett front saturation and 1 Sor. Moreover, if the value of

between

obtained from Eq. 31

is larger than the value obtained from Eq. 30, we assume that water injected through water injection well
i has not reached well j; in this circumstance, we use the saturation value obtained by computing the water
. Othersaturation value corresponding to the fractional flow derivative computed from Eq. 30 as
is the
wise, water injected at well i has broken through at well j, and the saturation at well j at tn,
saturation corresponding to the value
computed with Eq. 31. In short, in INSIM,
is determined
by
(32)
If the direction of flow between wells i and j changes again, we can use a similar procedure.
Specifically, if the flow direction changes at time tn we have

14

SPE-173213-MS

Figure 13Water cut data match and future predictions.

(33)

(34)
and
(35)

Toy Problem
To illustrate the feasibility of the proposed saturation tracking method, we compare results from INSIM
with corresponding results generated from ECLIPSE for a simple 1D simulation model with 100 grid
blocks, where the size of each block is 10 m 10 m 10 m. There are three wells in the model, see Fig.
5. For the first 600 days, water is injected at well W1 at the rate of 15 m3/d and well W2 is produced at
a total liquid rate of 15 m3/d throughout the same time period. At 600 days, W2 changed to an injector
and W1 to a producer. For the time interval (600,1500] days, the water injection rate at W2 is 15 m3/d
and W1 is produced at a total liquid rate of 15 m3/d. W3, which is in the 30th gridblock, is connected to

SPE-173213-MS

15

Figure 14 Water saturation at well nodes.

Table 1The initial transmissibility between wells.


W1 and W2 in the INSIM model, but the rate at W3
W1
W2
W3
W4
is fixed equal to zero, i.e., W3 is a pseudo-well. We
W1
0.00
2.2
2.44
1.64
investigate how accurately INSIM tracks water satW2
2.2
0.00
1.64
2.44
uration at W3. Note that up to 600 days, W3 is
W3
2.44
1.64
0.00
2.2
upstream to W2, but at some time after that, it
W4
1.64
2.44
2.2
0.00
becomes downstream. Fig. 6 shows the water satuTable 2The initial control pore volume( 104m3) between wells.
ration profile at four different times calculated by
ECLIPSE and INSIM. Note that the water saturaW1
W2
W3
W4
W1
0.0000
18.34
16.51
24.68
tion profiles at 750 days and at 990 days are qualW2
18.34
0.0000
24.68
16.51
itatively similar to the profile obtained by connectW3
16.51
24.68
0.0000
18.34
ing the solid black curve and dashed curve in the
W4
24.68
16.51
18.34
0.0000
schematic depiction of Fig. 4(b). The results from
ECLIPSE and INSIM are quite close, except at the
100th grid, which contains a water injector so that
water saturation is 0.8 (1 Sor) in INSIM. The discrepancy at the injection well is due to the fact that
we use the well gridblock saturation from Eclipse and this saturation does not become equal to 1 Sor
until the gridblock becomes fully saturated with water.

16

SPE-173213-MS

Table 3The optimized transmissibility between wells.


Fig. 7 compares the water saturation of W3 as a
W1
W2
W3
W4
function of time obtained from INSIM and
W1
0.00
6.42
0.31
29.26
ECLIPSE. The agreement between the two predicW2
6.42
0.00
2.17
14.24
tions is extremely good. Note that during the first
W3
0.31
2.17
0.00
1.27
600 days, the water saturation of W3 is monotoniW4
29.26
14.24
1.27
0.00
cally increasing due to water injection at W1. After
600 days, the flow direction changes and the satu- Table 4 The optimized control pore volume( 104m3) between wells.
ration of W3 monotonically decreases from 600
W1
W2
W3
W4
W1
0.00
43.13
10.80
37.18
days to 1000 days, at which time the water front due
W2
43.13
0.00
3.81
13.57
to injection at W2 reaches W3 from W2, so that,
W3
10.80
3.81
0.00
10.58
from 1000 days onward, the saturation of W3 inW4
37.18
13.57
10.58
0.00
creases.
Well rate allocation factors
Results from INSIM can be used to calculate the
allocation factor for the fluid rate between each injector-producer well pair using the same basic procedure
used in correlation-based models. However, unlike the result from correlation-based methods, the
allocation factor changes with time and thus reflects the dynamic interactions between wells. As in Gentil
(2005), for an injector-producer pair, the allocation factor of the liquid flow rate from injector i is defined
as the ratio of the liquid flow from injector i to the producer divided by the total injection rate of the
injector. Specifically, for injector i and producer j, the allocation factor
, is given by
(36)

where the sum is over the set of indices of all wells connected to injector i.

History Matching Phase Rate or Water-Cut Data


Predictions with INSIM depend on the sets of the two parameters defined for each connected well pair,
and
, i 1,2, . . . nw, j i 1, i 2, . . . nc. In history matching the phase-rate or water-cut data
at producers, these parameters represent the reservoir parameters (optimization parameters or historymatching parameters). Let Nd denote the number of data to be history matched and dobs denote the
Nd-dimensional column vector populated with these Nd observed data. Letting m denote the column vector
with entries given by optimization parameters,
and
, i 1, 2, . . . nw, j i 1, i 2, . . . nc
inserted in m so that m has the form:
(37)
A specific m that honors a vector of observed data is obtained by minimizing
(38)
such that
(39)
and
(40)
where the constraint of Eq. 40 is applied only if the reservoir pore volume, VR, is assumed to be known.
Note Eq. 39 simply requires that all the parameters be nonegative. In Eq. 40, the matrix CD is the

SPE-173213-MS

17

Figure 15Injection well allocation factors as functions of time.

Figure 16 The parameters Ti,j and Vp,i,j versus time.

covariance matrix for measurement errors. In our results, we assume measurement errors are uncorrelated
so CD is diagonal. The normalized objective function is denoted by ON(m) and is defined as
(41)
By standard arguments, Tarantola (2005), we expect to be able to achieve a history-matched model mc
such that ON (mc) is close to 1. In general, the inverse problem, i.e., the minimization of O(m) subject to
the constraints of Eqs. 39 and 40, is ill-posed and some regularization is required. However, because of
the limited number of connections used to date, we have been able to generate reasonable results without
regularization.
As the computational time required to run INSIM is small, we simply use a gradient-based optimization
algorithm with derivatives computed by the finite-difference method to optimize the objective function
defined in Eq. 38. The algorithm used is an interior-point method from the MATLAB optimization tools.
In the future, an adjoint problem will be solved to obtain the gradient of O(m) at each iteration of the

18

SPE-173213-MS

Figure 17The permeability distribution for the five well example.

optimization algorithm. In the following sections, we investigate the viability of INSIM by considering
two examples. The first is a synthetic example and the second is a field example.

2D Synthetic Model
The reservoir has a uniform grid system of 21 21 with uniform grid spacings given by x y 30
m and z 15 m. The reservoir porosity is uniform with 0.2; we assume that kx ky. The horizontal
permeability distribution with the locations of four wells is shown in Fig. 8. Note that there is a high
permeability channel between W1 and W4 and that W3 is located in a low permeability region. The water
and oil viscosity, respectively, are given by 1.0 and 20.0 mPa.s. The relative permeability curves are
shown in Fig. 9. In units of MPa1, the rock compressibility is 0.00645 and co cw 0.005. All wells
are produced under primary production for the first 60 days. Then well W1 is changed to an injector and
well W4 is closed. At day 1200, W4 is put on production and at 1500 days, this well is converted to a
water injection well. When a well is a water injection well, its injection rate is specified. When a well is
a producer, its flowing bottom-hole pressure (BHP) is specified except for the first 60 days of primary
production when the rates at all wells are specified. When W2 and W3 are producing, the BHP for these
wells is set equal to 10 MPa; when W4 is producing, its BHP is set equal to 30 MPa. The resulting flow
rates are shown in Fig. 10, where a negative rate corresponds to production and a positive rate corresponds
to water injection. The liquid production rate of W3 is approximately one fifth that of W2 during the
whole period, due to the low permeability around W3.
Using the well operating conditions and reservoir model specified in the preceding paragraph, we run
Eclipse 100 for the specified reservoir model to generate water-cut data. Each of these water-cut datum
is perturbed by adding to it a sample from the normal distribution N(0,0.004) in order to obtain data
corrupted by measurement error. These noisy data correspond to the noisy data that represent the entries
of dobs. The data covariance matrix, CD in Eq. 38, is diagonal with all its entries equal to 0.004. The total
number of water-cut data generated for history matching is 116.
With INSIM, we perform history matching for the wells water cut (WWCT) at the first 1680 days and
then give forward predictions and compare the result with those generated with Eclipse 100. Whenever
a well is injecting, its water cut is set equal to 1. In history matching with INSIM, we use a total of only
six connections so there are twelve parameters; see Fig. 11, which is discussed in detail later. Fig. 12

SPE-173213-MS

19

Figure 18 The rate schedule curve.

shows the normalized objective function as a function of the number of iterations of the optimization
algorithm used to minimize Eq. 38. Convergence is obtained after 33 iterations, where the convergence
criteria of the interior-point method of MATLAB is that the relative change in the value of the objective
function is smaller than 108. At convergence, the normalized objective function is reduced to about 2.6
from its initial value of 300. The data match obtained on a well-by-well basis is shown in Fig. 13 in which
red circles denote observed water cut data from Eclipse and the black and dark blue curves, respectively,
represent predictions based on the initial guess and predictions from INSIM based on the history-matched
values of the 12 parameters. Note these results show predictions for both the history-matching period (left
of the vertical dashed lines) and future predictions, which pertain to times that correspond to points to
the right of the vertical dashed lines. The initial guess gives a prediction far from that of Eclipse for wells
W2 and W3, but the predictions from history matching are in good agreement with the observed data
generated using Eclipse by the procedure discussed previously. The future predictions from 1,680 to 2,750
days from INSIM also agree well with those from Eclipse 100. At an injector, the water cut is set equal
to 1 throughout the injection period so that both INSIM and Eclipse match water-cut data during injection.
From Fig. 14, it can be seen that at wells W2 and W3, the water saturations predicted by INSIM using
the history-matched parameter values are in reasonable agreement with the wellblock water saturations
from Eclipse. For well W1, which is an injection well from day 60 onward, the agreement between the
two sets of saturation is not good. However, this is because if a well is injecting water, INSIM sets the
water saturation at the well node equal to 1 Sor, whereas in Eclipse, the wellblock value of water
saturation is not equal to 1 Sor until the wellblock is completely swept with water. A similar comment
pertains to the water saturation of well W4 subsequent to its conversion to an injector at 1,500 days.
The initial guesses for the 12 parameters calculated by Eq. 12 are shown in Tables 1 and 2, and the
corresponding optimized values from history matching with INSIM as the forward model are shown in
Tables 3 and 4. To initialize the 12 history-matching parameters, we assume a homogeneous reservoir, so
the permeability and porosity for each well node are identical with horizontal permeability equal to 1000
mD and net thickness equal to 15m at each well point. Porosity is assumed known and is fixed at 0.2.
The initial and final estimated interwell parameter values are also displayed in Fig. 11. The values of
these estimates indicate that, due to the high permeability channel, the transmissibility between W1 and
W4 is largest with the value of 29.3. The transmissibility and control pore volume of W1 to W2 and W2

20

SPE-173213-MS

Figure 19 The initial and final estimated geological parameters; in the brackets, the 1st value is transmissibility and the 2nd value is control pore
volume, 104m3.

Figure 20 Value of objective function versus number of iterations.

to W4 are both larger than that of W1 to W3 and W4 to W3. According to the well allocation factors,
which are shown in Fig. 15 and discussed below, W2 is mainly influenced by W1, and because W2 is the
major producer, the interwell pore volume between W1 and W2 is the largest interwell pore volume
obtained. Intuitively, we can provide some additional explanation of these results provided we assume that
W2 and W3 are not significantly affected by well W4 during the history matching period, which appears
to be true. From Fig. 15, we see that the allocation of W1 to W2 is about 5 times that of W1 to W3. If
the interwell pore volume from W1 to W2 is also about 5 times that of W1 to W3, then the water-cuts
at W2 and W3 should be very similar. However, the control pore volume from W1 to W2 is less than five
times the interwell pore volume between W1 and W3, so breakthrough occurs at W2 earlier than at W3.
Fig. 15 shows how the well allocation factors between injector-producer pairs vary with time. There
is no water injection prior to 60 days so we simply set all allocation factors to zero during this time period.
Note that no allocation factor is shown between W1 and W4 because these two wells form a injector-

SPE-173213-MS

21

Figure 21Water cut data match and prediction result.

producer pair only for the time period from 1200 to 1500 days. At other times, well W4 is either shut in
or is an injector and well W1 is an injector at all times subsequent to day 60. Note that during the 1200
to 1500 day time period, the two allocation factors of Fig. 15 sum to only about 0.2 which means that
during this time period, about 80% of the water injected at W1 flows along its connection to well W4,
which makes sense because of the high permeability channel connecting these two wells and the high
production rate of W4 during this period. Fig. 16 shows how the transmissibility and control pore volume
of each connected well pair change with time. The control pore volumes increase slowly because the
reservoir pressure is increasing. After W4 becomes an injection well at 1500 days, transmissibility jumps
to a much larger value because the total mobility involved in this transmissibility is evaluated at the
saturation at well W4, which is equal to 1 Sor.
2D Synthetic Model with fifth well
We add a fifth well (W5) at the center of the previous 2D synthetic model to obtain the well configuration
for this example which is shown in Fig. 17. Fig. 18 shows the rate behavior for this example where
negative values indicate production and positive values pertain to water injection. For the first 120 days,
all wells are on primary production. At 120 days, both W1 and W4 are converted to water injector wells.
The center well, well W5, is shut in when its water cut reaches 0.95 at about 900 days, but wells W2 and
W3 remain on production throughout the time period considered which is equal to 2300 days.
Similar to the first example, the water-cut data from time 0 to 1200 days were generated by adding
noise to water-cut values generated from Eclipse with a noise level of 0.02. There are 105 water-cut data

22

SPE-173213-MS

Figure 22Water saturation at well nodes.

Figure 23The injection well allocation factor curves.

SPE-173213-MS

23

Figure 24 The parameter of Ti,j and Vp,i,j changing with time.

Figure 25The relative permeability curves for the Z16 block.

which represent the observed data. In this case there are 8 connections between well pairs and thus there
are 16 parameters to be history matched; see Fig. 19, which is discussed in detail later.
As shown in Fig. 20, starting from an initial value of about 80, the normalized objective function
converges to about 1.6 after 22 iterations. The history-matching result and future predictions generated
with INSIM using the values of the parameters obtained by history matching are shown in Fig. 21. Note
the match and future predictions are even better than in the first example, probably because the additional
connections to well W5 in the interior of the reservoir allow us to resolve flow better, but it could be
simply because we have more parameters to adjust and thus should be able to obtain a better match of
observed water-cut data. Fig. 22 shows that the predictions of well water saturations predicted from
INSIM using the history-matched parameters are close to the corresponding well gridblock saturations
generated from Eclipse.
The initial guess for parameters for history matching are obtained from Eqs. 1 and 2 using the same
permeability, thickness and porosity values used for the first example. The initial and final optimized
interwell parameters are shown in Fig. 19. These results indicate that due to the high permeability channel,

24

SPE-173213-MS

Figure 26 The net thickness distribution of the reservoir (Z16 block).

the optimized and transmissibility values along the connections from well W1 to well W5 and from well
W4 to well W5 are much higher than the transmissibilities along connections between other well pairs.
Fig. 23 shows the well allocation factor values between injector-producer pairs as a function of time.
It can be seen that prior to about 1000 days, the fluid mainly flows to well W5 from the two injectors due
to the high transmissibility between W1 and W5 and W4 and W5, but after W5 is shut in at 1000 days,
the allocation factors between W4 and W2 and W1 and W2 gradually increase and the water from both
W4 and W1 flows mainly to W2.
Fig. 24 shows the transmissibility and control pore volume of two well pairs as a function of time. The
control pore volume for each well pair increases slowly with time due to increasing reservoir pressure. The
transmissibility between wells W1 and W5 jumps at 120 days when W1 becomes an injector because the
total mobility that appears in this transmissibility jumps from its value at Siw to its value at 1 Sor.
Somewhat similarly, the transmissibility between W2 and W5 increases with time because the water
saturation at well W5 increases with time.
Actual Field-Z16 Block
This a single-layer, low permeable reservoir. The oil viscosity is 27.55 mPa.s and the average porosity is
about 0.14. The estimates of total reserves and reservoir pore volume, respectively, are 79.2 104 m3 and
107.8 104m3, respectively. The relative permeability curves are shown in Fig. 25. The reservoir
contains 8 injectors and 15 producers during a waterflooding period of about 2300 days. The dynamic data
to be history matched are the well oil production rate (WOPR), field cumulative oil production (FOPT)
and field water cut (FWCT). All rates are in m3/day. When history matching, the producing wells are
operated at the total liquid rates observed in the field. The standard deviation of measurement errors used
to form the data covariance matrix, CD is set equal to 3% of the observed data. If an observation is equal

SPE-173213-MS

25

Figure 27The initial and final estimated geological parameters; in brackets, the 1st value is transmissibility and the 2nd value is control pore
volume, 104m3.

26

SPE-173213-MS

Figure 28 Objective function value versus number of iterations, field case.

Figure 29 The field cumulative oil production rate and water cut data match result.

to zero, then the corresponding entry of CD is set equal to 0.01. In total, there are 1925 data to be history
matched.
On a well-by-well basis, the connection radius for each individual well is set equal to twice the smallest
well distance to any other well. However, if a connection from well i to well j is determined by considering
the connection radius for well i, that connection remains when we consider the connections to well j using
the connection radius of well j even if well i does not lie within the connection radius of well j, i.e., if j
is connected to i, the well i is automatically connected to well j. Then, the superfluous connections are
removed if the two inner angles of a triangle formed by three connections are less than 15 degrees as we
discussed earlier; see Fig. 2. As this process yields 72 connections between well pairs, there are 144
connection parameters to be estimated (transmissibilities and volumes) by history matching.
To estimate the initial guess of the transmissibility and control volume for each well pair needed to
begin history matching, the average estimates of permeability and net thickness at well points, together
with 0.14, are used in Eqs. 1 and 2. The permeabilities at well points used in 2 are from well logging
with the values provided by the field engineer. The variation in these permeability values is not very large.
A contour plot of the estimated reservoir thickness is shown in Fig. 26. The initial estimates of the values
of connection parameters for the reservoir (Z16 block) are shown in Fig. 27(a).
Fig. 28 presents a plot of the normalized objective function value versus the iteration index. The
optimization algorithm converges after 37 iterations to a value of about 100. That the value at convergence

SPE-173213-MS

27

Figure 30 History match of oil rates.

Figure 31The injection well allocation factor curves.

is higher than the intuitively achievable value of roughly 1.0 suggests that the entries in CD underestimate
the noise level, but increasing the variances of measurement errors by a constant value would not change
the history-match; it would only reduce the values of the normalized objective function we obtain. Fig.
29 and Fig. 30 show the history match of FOPT and FWCT data, as well as predictions from the initial
guess. The red circles are the observed data; the black and dark blue curves, respectively, pertain to the

28

SPE-173213-MS

Figure 32The injection allocation factor map at two different time.

predictions from initial guess and the predictions generated with INSIM using the history-matched
parameters. Note that an excellent match is obtained. Fig. 27 shows the initial and final estimated
parameters (transmissibility and control pore volume) for each connection between each well pair where
the connections colored red pertain to those with higher values of connection transmissibility; many well
pair transmissibilities are smaller than 0.05 as this is a low permeability reservoir.
As noted previously, one of the advantages in INSIM is that it can output the injection allocation fluid
flowing from an injector to producing wells at any time. Fig. 31 presents the injection allocation factor
curves for injectors Z16X-4-3 and Z16X-8-6. For injector Z16X-4-3, most of the injected water mainly
flows to producers Z16X-6-3 and Z16X-5-5. Note that the allocation factors become fairly stable after

SPE-173213-MS

29

1600 days. The injected water from Z16X-8-6 mainly flows to well Z16X-7-6 except in the period from
700 to 1400 days. In this time period, the liquid production rates at wells Z16X-7-5 and Z16X-6-5
gradually increase from about 3.0 m3/day to 11.0 m3/day, while Z16X-7-6 remains at a low production
rate of roughly 3.0 m3/day until 1100 days, and then quickly increases to over 10.0 m3/day.
Fig. 32 shows the allocation factor map between injector and producer pairs at times 1200 days and
1800 days; the orientation of each arrow is from producer to injector and the relative length of the arrows
in Fig. 32 represent the degree of connectivity between an injector and producers. In general, the flow
trends of injected water for wells Z16X-6-4 and Z16X-8-6 at 1800 days are significantly different from
those at 1200 days. Z16X-6-4 mainly affects Z16X-4-31 at the time 1200 days, but mainly influences well
Z16P2 at the time 1800 days. For Z16X-8-6, the major connectivity is changed from well Z16X-6-5 at
time 1200 days to well Z16X-7-6 at 1800 days.

Conclusions
Despite the ad hoc computation of saturation at well nodes after flow direction reversals, the computational results suggest that the interwell numeric simulation model (INSIM) is a simple, computationallyefficient forward model that can replace a reservoir simulator when history-matching phase rate data or
water-cut data obtained during water flooding operations. Although the results presented illustrate the
validity of the INSIM model, it may be possible to significantly improve this model by adding a more
rigorous front-tracking computation and by using the parameters of power law relative permeability
curves as history matching parameters.
For the two synthetic examples, predictions from INSIM based on the history-matched model are in
good agreement with predictions from Eclipse using the true reservoir model. For the field case, we are
able to obtain a very good history match using INSIM as the forward model.
The INSIM simulation process for solving pressure and saturation is fast and stable. History-matched
results obtained by the combination of INSIM with an optimization algorithm allow one to determine the
formation characteristics between well pairs, the rate allocation factors for each injector (the fraction of
injected water that flows to each producer connected to the injector) and the principal directions of the
flow of water from each injection well.
Unlike correlation-based models, INSIM can produce reasonable accurate phase flow rates and
water-cut and handle changes in flow directions due to closing wells or the conversion of producers to
injectors.
With INSIM as the forward model, it is possible to obtain a good history match of field data and
estimate the formation characteristics between well pairs, the rate allocation factors for each injectors
injected water, and the principal directions of the flow of water from each injection well.

Acknowledgements
The support of the member companies of TUPREP is very gratefully acknowledged. Hui Zhao would like
to express his gratitude to the National Natural Science Foundation of China (Grant No.51344003) and
the China Important National Science & Technology Specific Projects (Grant No. 2011ZX05014) for their
generous financial support of the research.

References
Aanonsen, S. I., G. Nvdal, D. S. Oliver, A. C. Reynolds, and B. Valls, Review of ensemble Kalman
filter in petroleum engineering, SPE Journal, 14(3), 393412, 2009.
Albertoni, A. and W. Lake, Inferring connectivity only from well-rate fluctuations in water floods,
SPE, 83381, 2003.

30

SPE-173213-MS

Anterion, F., B. Karcher, and R. Eymard, Use of parameter gradients for reservoir history matching,
in Proceedings of the 10th SPE Reservoir Simulation Symposium, Houston, Texas, 6 8 February,
SPE 18433, 1989.
Araque-Martinez, A., Estimation of Autocorrelation and Its Use in Sweep Efficiency Calculation,
Masters thesis, University of Texas at Austin, 1993.
Barros-Griffiths, I., The Extreme Spearman Rank Correlation Coefficient in the Characterization of
the North Buck Draw field, Masters thesis, University of Texas at Austin, 1998.
Bissell, R., Y. Sharma, and J. E. Killough, History matching using the method of gradients: Two case
studies, in Proceedings of the SPE 69th Annual Technical Conference and Exhibition, SPE 28590,
1994.
Chen, Y. and D. Oliver, Levenberg-Marquardt forms of the iterative ensemble smoother for efficient
history matching and uncertainty quantification, Computational Geosciences, 17(4), 689 703,
2013.
Dinh, A. and D. Tiab, Inferring interwell connectivity from bottomhole pressure fluctuations in
waterfloods, SPE, 106881, 2007.
Emerick, A. A. and A. C. Reynolds, History matching time-lapse seismic data using the ensemble
Kalman filter with multiple data assimilations, COMPGEO, 16(3), 639 659, 2012.
Emerick, A. A. and A. C. Reynolds, Ensemble smoother with multiple data assimilations, Computers
& Geosciences, 2013.
Evensen, G., The ensemble Kalman filter: Theoretical formulation and practical implementation,
Ocean Dynamics, 53, 343367, 2003.
Evensen, G., Data Assimilation: The Ensemble Kalman Filter, Springer, Berlin, 2007.
Evensen, G. and P. J. van Leeuwen, An ensemble Kalman smoother for nonlinear dynamics, MWR,
128(6), 18521867, 2000.
Gao, G. and A. C. Reynolds, An improved implementation of the LBFGS algorithm for automatic
history matching, SPE Journal, 11(1), 517, 2006.
Gao, G., M. Zafari, and A. C. Reynolds, Quantifying uncertainty for the PUNQ-S3 problem in a
Bayesian setting with RML and EnKF, SPE Journal, 11(4), 506 515, 2006.
Gentil, P., The Use of Multilinear Regression Models in Patterned Waterfloods: Physical Meaning of
the Regression Coefficients, PhD. thesis, University of Texas at Austin, 2005.
Gherabati, S., A. Hughes, H. Zhang, and C. While, A large scale network model to obtain interwell
formation characteristics, SPE, 153386, 2012.
Heffer, K., R. Fox, and C. McGill, Novel techniques show links between reservoir flow directionality,
earth stress, fault structure and geomechanical changes in mature waterfloods, SPE, 30711, 1995.
Hui, Z., L. Yang, G. Da, and C. Lin, Study on reservoir interwell dynamic connectivity based on
systemic analysis method, Acta Petrolei Sinica., 31, 633636, 2010.
Jansen, F. and M. Kelkar, Non-stationary estimation of reservoir properties using production data,
SPE, 38729, 1997.
Kaviani, D., J. L. Jensen, L. Lake, and M. Fahes, Estimation of interwerwell connectivity in the case
of fluctuating bottomhole pressures, SPE, 117856, 2008.
Kaviani, D., P. Valk, and J. L. Jensen, Application of the multiwell productivity index-based method
to evaluate interwell connectivity, SPE, 129965, 2010.
Lee, K.-H., A. Ortega, A. M. Nejad, N. Jafroodi, and I. Ershaghi, A novel method for mapping
fracturcs and high permcability channels in waterfloods using injcction and production rate, SPE,
121353, 2009.
Li, R., A. C. Reynolds, and D. S. Oliver, History matching of three-phase flow production data, SPE
Journal, 8(4), 328 340, 2003.

SPE-173213-MS

31

Nvdal, G., L. M. Johnsen, S. I. Aanonsen, and E. H. Vefring, Reservoir monitoring and continuous
model updating using ensemble Kalman filter, SPE Journal, 10(1), 66 74, 2005.
Panda, M. N. and A. K. Chopra, An integrated approach to estimate well interactions, SPE, 39563,
1998.
Refunjol, B., Reservoir characterization of North Buck Draw field based on tracer response and
production/injection analysis, Masters thesis, University of Texas at Austin, 1996.
Sayarpour, M., Development and Application of Capacitance-Resistive Models to Water/C02 Flood,
PhD. thesis, University of Texas at Austin, 2008.
Tarantola, A., Inverse Problem Theory and Methods for Model Parameter Estimation, SIAM,
Philadelphia, USA, 2005.
Thiele, M. R. and R. P. Batycky, Using streamline-derived injection efficiencies for improvcd
waterflood management, SPE Reservoir Evaluation and Engineering, 9, 187196, 2006.
Wang, Y., G. Li, and A. C. Reynolds, Estimation of depths of fluid contacts by history matching using
iterative ensemble-Kalman smoothers, SPE Journal, 15(2), 2010.
Wu, Z., A. C. Reynolds, and D. S. Oliver, Conditioning geostatistical models to two-phase production
data, SPE Journal, 3(2), 142155, 1999.
Yousef, A., J. Jensen, and L. Lake, Analysis and interpretation of interwell connectivity from
production and injection rate fluctuation using a capacitance model, SPE, 99998, 2006.
Zhang, F. and A. C. Reynolds, Optimization algorithms for automatic history matching of production
data, in Proceedings of 8th European Conference on the Mathematics of Oil Recovery, Freiberg,
Germany, 3 6 September, 2002.

You might also like