ArticlePDF Available

Characterizing interwell connectivity in waterflooded reservoirs using data-driven and reduced-physics models: a comparative study

Authors:

Abstract and Figures

Waterflooding is a significantly important process in the life of an oil field to sweep previously unrecovered oil between injection and production wells and maintain reservoir pressure at levels above the bubble-point pressure to prevent gas evolution from the oil phase. This is a critical reservoir management practice for optimum recovery from oil reservoirs. Optimizing water injection volumes and optimizing well locations are both critical reservoir engineering problems to address since water injection capacities may be limited depending on the geographic location and facility limits. Characterization of the reservoir connectivity between injection and production wells can greatly contribute to the optimization process. In this study, it is proposed to use computationally efficient methods to have a better understanding of reservoir flow dynamics in a waterflooding operation by characterizing the reservoir connectivity between injection and production wells. First, as an important class of artificial intelligence methods, artificial neural networks are used as a fully data-driven modeling approach. As an additional powerful method that draws analogy between source/sink terms in oil reservoirs and electrical conductors, capacitance–resistance models are also used as a reduced-physics-driven modeling approach. After understanding each method’s applicability to characterize the interwell connectivity, a comparative study is carried out to determine strengths and weaknesses of each approach in terms of accuracy, data requirements, expertise requirements, training algorithm and processing times.
This content is subject to copyright. Terms and conditions apply.
ORIGINAL ARTICLE
Characterizing interwell connectivity in waterflooded reservoirs
using data-driven and reduced-physics models: a comparative
study
Emre Artun
1
Received: 18 June 2015 / Accepted: 21 December 2015
The Natural Computing Applications Forum 2016
Abstract Waterflooding is a significantly important pro-
cess in the life of an oil field to sweep previously unrecovered
oil between injection and production wells and maintain
reservoir pressure at levels above the bubble-point pressure
to prevent gas evolution from the oil phase. This is a critical
reservoir management practice for optimum recovery from
oil reservoirs. Optimizing water injection volumes and
optimizing well locations are both critical reservoir engi-
neering problems to address since water injection capacities
may be limited depending on the geographic location and
facility limits. Characterization of the reservoir connectivity
between injection and production wells can greatly con-
tribute to the optimization process. In this study, it is pro-
posed to use computationally efficient methods to have a
better understanding of reservoir flow dynamics in a water-
flooding operation by characterizing the reservoir connec-
tivity between injection and production wells. First, as an
important class of artificial intelligence methods, artificial
neural networks are used as a fully data-driven modeling
approach. As an additional powerful method that draws
analogy between source/sink terms in oil reservoirs and
electrical conductors, capacitance–resistance models are
also used as a reduced-physics-driven modeling approach.
After understanding each method’s applicability to charac-
terize the interwell connectivity, a comparative study is
carried out to determine strengths and weaknesses of each
approach in terms of accuracy, data requirements, expertise
requirements, training algorithm and processing times.
Keywords Waterflooding Reservoir characterization
Interwell connectivity Data-driven modeling Artificial
neural networks Reduced-physics modeling
Capacitance–resistance models
1 Introduction
Waterflooding is a secondary oil recovery method that is
applied after the oil has been produced from a reservoir
with its natural energy (known as the primary recovery
period). Based on industry experience, it is known that the
recovery factors (ratio of recovered oil volume to the
original oil volume in the reservoir) with primary recovery
ranges between 15–25 %, while a well-managed water-
flood can increase this to 40–50 % depending on the
reservoir characteristics and efficiency of the waterflooding
operation. Although it can be classified as a common
operation for secondary recovery, varying reservoir char-
acteristics and limited water supplies in some areas make it
critical to have a good understanding of the reservoir and
determining the optimum design schemes of the water-
flooding operation under consideration. Poorly managed
waterflooding operations result in underperforming reser-
voirs with reserves that are left behind, and thus, huge
losses in recovered oil and associated monetary income are
realized. Understanding reservoir flow dynamics to apply
proper reservoir management practices is a complex
problem with well data from isolated localities in an oil
field, data distributed spatially across the whole reservoir
and spanned tens of years of history. Current reservoir
management practices highly depend on numerical flow
simulation models that take months to develop and main-
tain, and cost millions of dollars with both significant
manpower and computational power requirements.
&Emre Artun
artun@metu.edu
1
Petroleum and Natural Gas Engineering Program, Middle
East Technical University, Northern Cyprus Campus,
Mersin 10, Kalkanlı, 99738 Gu
¨zelyurt, Turkey
123
Neural Comput & Applic
DOI 10.1007/s00521-015-2152-0
Although being recognized scientifically as the most
powerful approach, full field-scale models do not easily
allow rapid reservoir analysis that results in right reservoir
management decisions.
Over the last two decades, petroleum industry has been
transformed significantly with the advancement of intelli-
gent field technologies that are mostly based on on-site
instrumentation and automation of field operations. Many
technologies have been adapted to collect significant vol-
umes of data in much shorter time frames. However, the
real challenge has been to convert these data into useful
information to make quick and reliable decisions that
generate value. This challenge can only be overcome by
utilizing proper knowledge management, data assimilation
and data analysis practices. The current paradigm in the
evolution of science also requires advanced data analysis to
synthesize all of the earlier empirical, experimental and
computational findings [1].
An efficient way of managing an hydrocarbon reservoir at
any stage of development is the closed-loop reservoir man-
agement approach [2]. As shown in Fig. 1,thisapproach
requires continuous updating of models after collecting recent
data from the high-resolution and high-frequency sensors in
the oil field that record measurements of time-dependent
(dynamic) properties such as pressure, flow rate and temper-
ature. In this workflow, it is very important to have a model
that can respond to the following primary needs:
A model that can be updated quickly when new data are
available.
A model that is sufficiently accurate and representative
of the actual system (surface or subsurface) so that it
can be used for decision-making purposes.
There are a wide variety of modeling approaches presented
in the reservoir engineering literature. Each modeling
approach represents various complexities, advantages and
disadvantages. In order to find a model that serves to both
of the aforementioned objectives, it would be a better
approach to have access to different modeling options
readily available and choose the right modeling approach
depending on the problem type and scope.
In this study, two conceptually different modeling
approaches are investigated for the purpose of character-
izing interwell connectivity in a waterflooded reservoir:
1. Fully data-driven modeling: artificial neural networks
(ANN)—no functional relationship presumed.
2. Reduced-physics-driven modeling: capacitance–resis-
tance model (CRM)—physics are incorporated with
certain assumptions that simplify the problem.
Both methods have been proven to be potentially efficient
tools for reservoir engineering problems based on studies
presented in the literature. In this study, it is aimed to
investigate the characteristics of each method including
best practices and challenges associated with them to
characterize interwell connectivity between injection and
production wells. These would allow us to compare these
methods with each other from the practical point of view
and develop guidelines for the practicing engineer or asset
team who is responsible for developing an optimum
waterflooding plan. The primary objectives of the study can
be summarized as the following:
1. Utilizing two different modeling approaches: artificial
neural networks (as a data-driven modeling approach)
and capacitance–resistance models (as a reduced-
physics modeling approach) for quantifying interwell
connectivity between injection and production wells in
a waterflooded petroleum reservoir.
2. Assessing and comparing both methods (ANN and
CRM) from different perspectives to determine
strengths and weaknesses of each approach in terms
of accuracy, data requirements, training algorithm,
processing times and expertise requirements.
This study is the first attempt, to the best of our knowledge,
to compare these two modeling approaches for the purpose
of characterizing reservoir connectivity. This comparison
provides with the necessary insight for the practicing
engineer to implement either of these methods for a
waterflooded petroleum reservoir. These tools have great
advantages over other modeling approaches because of
requiring fewer inputs, being much more computationally
efficient and not being dependent on geological uncer-
tainties. Therefore, having the necessary insights would
help to decide which method is more practical for a par-
ticular problem. Based on the analysis performed during
this study, the decision of choice would be affected by the
expertise of the practicing engineer, availability of the data
and the time frame of the study.
System
(reservoir, wells
and facilities)
Sensors
Data
assimilation
algorithms
System models
Optimization
algorithms
Geology,
seismic,
well logs,
well tests,
PVT,
etc.
Measured
output
Predicted
output
Input
Noise
Controllable
output
Output Noise
Fig. 1 Closed-loop reservoir management workflow [2]
Neural Comput & Applic
123
2 Methodology
2.1 Case study: a synthetic reservoir model
In this study, a synthetic streak case study [3] is selected to
implement the aforementioned methods. It is a synthetic
field that consists of 5 vertical injectors, I1 through I5, and
4 vertical producers, P1 through P4 (Fig. 2). The perme-
ability of the reservoir is 5 md, except two high-perme-
ability streaks:
1. Streak-1: 1000 md between I1 and P1 wells.
2. Streak-2: 500 md between I3 and P4 wells.
The porosity is constant and equal to 0.18. Total mobility
of oil and water (koþkw) is 0.45 and is independent of
saturation. Oil, water and rock compressibilities are
5910
-6
psi
-1
,1910
-6
psi
-1
,1910
-6
psi
-1
, respec-
tively. The model is constructed with 1 layer and 31 grid
blocks in each of the xand ydirections with grid sizes of 80
ft 980 ft (Dxand Dy). The thickness of the reservoir is 12
ft (Dz).
This model has been built and run using a commercial,
numerical reservoir simulator [4] that utilizes black-oil
formulation, which is a common formulation used as a
robust approach for waterflooding problems. A variable
water injection rate scenario is implemented, in which
volumetric injection rates are varied significantly over a
period of 100 months (10 years). It is aimed to charac-
terize the connectivity of the system through these rate
fluctuations [3]. The bottom-hole pressure (BHP) for pro-
ducing wells is fixed at 250 psia, and volumetric liquid
production rates are measured (Fig. 3). The volumetric
water injection rates are varied manually, while setting a
limit for maximum bottom-hole pressure of 5000 psia
(Fig. 4). Both injection and production rates are measured
at reservoir conditions (rbbl/day: reservoir barrels per day).
One hundred months of monthly injection/production his-
tory resulted in a sample size of 100 for each well. Other
descriptive statistics of volumetric injection rates for
injector wells, I1 through I5, and volumetric production
rates for producer wells, P1 through P4, are given in
Table 1.
Although the presented synthetic case has rather simple
permeability contrasts, it is a good example of a real
reservoir with high-permeability streaks that must be con-
sidered during a waterflood optimization study. Existence
of such streaks amplifies the importance of characterizing
the connectivity between wells, since they provide a con-
duit in the reservoir to transport the injected water. The fact
that results from a synthetic model are used should not
raise any concern regarding the validity of the methods
presented since both methods are proven to be successful in
real reservoir cases with a number of examples in the lit-
erature. CRMs were successfully applied to a number of
fields [5], and ANNs were successfully utilized for reser-
voir characterization problems with real-field data [69].
Since the main objective in this study is to perform a
comparison of two methods, a synthetic case would be
sufficient. However, it is anticipated that for a more com-
plex reservoir case with more heterogeneities and more
wells, more number of historical observations (more than
10 years of history) of injection and production rates might
be needed for capturing the fluid flow dynamics of the
reservoir.
2.2 Artificial neural networks
Intelligent systems have been applied to many different
types of optimization problems in the petroleum industry.
Most of these problems presented in the literature are based
on development of ANN-based proxy models that can
accurately mimic reservoir models within a reasonable
amount of accuracy and computational efficiency. In some
studies, these models are utilized to construct data-driven
predictive tools and these tools are coupled with evolu-
tionary algorithms to solve the optimization problem effi-
ciently. Several areas of application included reservoir
characterization [69], candidate well selection for
hydraulic fracturing treatments [10], field development
[1113], well placement and trajectory optimization [14
16], scheduling of cyclic steam injection process [17],
screening and optimization of secondary/enhanced oil
recovery[1823], history matching [2426], underground
gas storage management [27], reservoir monitoring and
management [26,28] and modeling of shale-gas reservoirs
[29,30].
1,000
500
5
Permeability (md)
Fig. 2 Synthetic reservoir model used in this study and its
permeability distribution: a reservoir with 2 high-permeability streaks
[3]
Neural Comput & Applic
123
In addition to petroleum engineering and many other
engineering disciplines, artificial neural networks and other
data-driven modeling approaches have been used in many
different kinds of applications such as spatial clustering
[31], cavity-filter optimization [32], electricity load fore-
casting [33], control, pattern recognition, signal processing,
medicine, speech recognition, speech production and
business [34].
The most common training algorithm and also the one
used in this study is the backpropagation algorithm. Also
known as the generalized delta rule, backpropagation
algorithm is a gradient-descent method that minimizes the
total squared error of the output computed by the network.
It played a major role in the re-emergence of neural net-
works in late 1980s. It was introduced as a training method
of multilayer networks to overcome the limitations of
single-layer networks [34]. Backpropagation algorithm is a
supervised training technique (i.e., mapping a given set of
inputs to a specified set of target outputs) and includes
three stages: (1) feedforward of the input training pattern,
(2) calculation and backpropagation of the error and (3)
adjustment of weights. The overall goal is to train the
network such that it can [34]:
Respond correctly to the input patterns that are used for
training (memorization).
Give reasonable responses to similar, but not identical,
input patterns (generalization).
Figure 5shows a multilayer, fully connected network
with one hidden layer. There are ninput neurons in the
input layer, phidden neurons in the hidden layer and m
output neurons in the output layer. There are biases also
shown in this figure whose activation value is constant
during the training (1, in this case). While the number of
inputs and outputs is based on the nature of the problem
studied, the number of hidden neurons is a part of the
network design process and must be optimized by the
designer. A rule-of-thumb formula is presented to calculate
the number of the hidden neurons, which is mostly based
on experience [35]:
NHN ¼NIþNO
2þffiffiffiffiffiffiffi
NTP
pð1Þ
where NIis the number of inputs, NOis the number of
outputs, and NTP is the number of training patterns. It
should be noted that this is not a theoretical formula, and
this number would not necessarily be the best estimate of
the number of hidden neurons. However, it can be used as a
good start for the optimization process. Algorithm 1 shows
a step-by-step explanation of the backpropagation
algorithm.
0
50
100
150
200
250
300
350
400
450
0
500
1000
1500
2000
2500
3000
3500
4000
4500
0 102030405060708090100
Liquid Production Rate (rbbl/d)
P2 and P3
)
d/
lbbr(
e
taR
no
i
t
c
udorP diuqiL
4P dna 1P
Time (months)
P1 P4
P2 P3
Fig. 3 Production history of the
synthetic reservoir model
0
800
1600
2400
3200
4000
0 102030405060708090100
)d/lb
b
r(eta
Rn
oitcejnIret
aW
Time (months)
I1
I2
I3
I4
I5
Fig. 4 Injection history of the synthetic reservoir model
Neural Comput & Applic
123
Table 1 Descriptive statistics
of volumetric injection and
production rates that are output
from the model and used in the
study
Injection well I1 I2 I3 I4 I5
Volumetric water injection rates (rbbl/day)
Mean 1,683.1 999.5 860.4 779.0 931.8
Standard error 55.5 20.9 28.3 21.5 28.1
Median 1,595.2 1,018.6 854.7 792.0 1,020.1
Standard deviation 554.7 208.8 283.4 215.3 280.5
Sample variance 307,638.1 43,595.6 80,313.9 46,345.7 78,686.5
Kurtosis 1.24 0.60 0.37 -0.07 -0.05
Skewness 0.71 -0.74 -0.27 0.04 -0.86
Range 3,258.7 1,053.0 1,288.9 1,034.2 1,128.5
Minimum 305.9 227.0 172.7 272.0 145.0
Maximum 3,564.6 1,280.0 1,461.6 1,306.3 1,273.5
Sum 168,313.9 99,945.2 86,043.1 77,896.8 93,179.6
Count 100 100 100 100 100
Production well P1 P2 P3 P4
Volumetric liquid production rates (rbbl/day)
Mean 2,494.9 157.1 274.3 2,386.1
Standard error 56.8 4.0 3.3 43.8
Median 2,458.5 158.1 272.3 2,354.8
Standard deviation 568.4 39.7 33.2 438.4
Sample variance 323,095.6 1,577.6 1,105.2 192,167.8
Kurtosis 0.34 0.23 0.30 0.41
Skewness 0.27 0.47 -0.33 0.52
Range 3,141.9 193.4 174.5 1,899.1
Minimum 917.5 85.0 182.5 1,509.9
Maximum 4,059.4 278.4 357.0 3,408.9
Sum 249,486.5 15,712.0 27,431.2 238,606.3
Count 100 100 100 100
Neural Comput & Applic
123
The iterative procedure for each training pair shown in
Algorithm 1 is repeated for all training patterns, until a pre-
specified stopping condition is achieved. Processing of
each training data is known as a training event or iteration.
When all training data are processed once, one epoch
(training cycle) is completed. Each training data can be
processed either by random selection, or by rotation. After
each training event, average mean-squared error is calcu-
lated. Achieving minimum mean-squared error of outputs
and maximum number of epochs is among most common
stopping conditions. Once the defined stopping criteria are
satisfied, weights on connection links achieve their opti-
mum states. Provided that the training performance is
satisfactory, the trained network with optimum weights can
be used as a predictive model.
In this study, mapping input–output relationships is
achieved with the inputs of injection rates from the water
injectors and the output of the liquid production rate of a
given producer. By analyzing the weights on connection
links of the trained neural network, interwell connectivity
is quantified. The optimized value of the weight on a given
connection link indicates the degree of influence of the
given input parameter on the output parameter. Therefore,
we propose that once the training is completed, the relative
values of connection links that connect each injector signal
to the producer can be used to quantify the connectivity.
There are individual neural networks for each producer
well in the field. A neural network for a given producer
would provide the connectivity of each injector to that
producer. Once all neural network models are trained, all
connectivities between all injector–producer pairs would
be quantified. This would provide insights about the
waterflood dynamics in the reservoir and help to under-
stand the overall reservoir connectivity to be used for
further optimization studies.
A feedforward artificial neural network is constructed
for each producing well in the reservoir. The training
algorithm used is the Levenberg–Marquardt backpropaga-
tion algorithm [36,37], and due to the low number of total
input/output parameters (5 injectors and 1 producer: 6
parameters), only 1 hidden layer is used with 12 neurons.
The schematic of the architecture of the neural network is
shown in Fig. 6. Eighty percentage of the historical pro-
duction/injection are used for training, 10 % are used for
validation during the training to prevent over-training, and
10 % are used for blind-case testing.
2.3 Capacitance–resistance models
It was suggested that the development of an electrical
model offers the promise of rapid evaluation for non-
mathematical analysis of complex reservoir problems
including understanding of the waterflood performance [3,
38]. An analogy between the flow behaviors of electricity
in electric units and fluid in reservoir units was made
through an experimental study [38]. This analogy implies
that the electrical unit acts as a device which stores the
electrical charge just as reservoir rock is acting as the
storage of reservoir fluids [38]. After considering that the
current may be equivalent to fluid flow, and the pressure is
Fig. 5 Architecture of a
multilayer network
Neural Comput & Applic
123
equivalent to the electrical potential, oil reservoir, as a
porous continuum, is divided into small blocks so that the
material balance can be used assuming the reservoir fluid is
flowing in at one face of the block and out at the opposite
face [38]. Role of such models for rapid estimation of
waterflood performance and optimization was investigated
by calling them capacitance–resistance models (CRMs)
[3]. The base data necessary to run this model are pro-
duction/injection data and well bottom-hole pressure
(BHP) to calibrate the model against a specific reservoir.
CRMs were primarily used for the characterization of
interwell connectivity between injection and production
wells rapidly without needing a geological model [39,40].
After characterizing the connectivity, they are then used to
optimize injection allocation and well locations in water-
flooded reservoirs. An integrated capacitance–resistance
model (ICRM) was presented that uses cumulative water
injection and cumulative total production instead of water
injection rate and total production rate while investigating
the advantages of a linear reservoir model over the non-
linear capacitance–resistance model [41,42].
The main advantage of CRM is that it requires very few
inputs as little as the production/injection history. It is based
on the main assumption that reservoir properties can be
drawn only from production/injection history of wells. Also
it requires that no significant changes in the field are observed
during the analysis period. The primary application area is
for the fields that are observing a secondary recovery period
with water or gas injection. Its applications for primary and
tertiary recovery periods are still being developed.
In these applications, the method enables to quantify
connectivity between injector–producer pairs and aquifer
strength, through history matching the production history
by adjusting model parameters. After the capacitance
models were introduced to understand interwell connec-
tivity [43], CRMs for dynamic evaluation of waterfloods
were presented [5]. Being a simple and user-friendly tool,
the methodology proved to be very powerful in field
applications, especially by quantifying interactions
between injector and producer wells [44]. CRMs for
three different control volumes are presented with semi-
analytical formulations, with each of them having differ-
ent level of complexities [3]:
1. One producer’s control volume,
2. An injector–producer pair’s control volume,
3. A field’s control volume.
In this study, a producer-based control volume is consid-
ered to focus on production wells and how they are con-
nected to different injection wells. Considering Ninumber
of injectors and Npnumber of producers, an in situ volu-
metric balance over the effective pore volume of the pro-
ducer is defined by the following differential equation [45]:
dqjðtÞ
dtþ1
sj
qjðtÞ¼1
sjX
Ni
k¼1
fijiiðtÞJj
dpwf ;j
dtð2Þ
where sjis the time constant for producer jand defined as a
function of total compressibility, ct, pore volume, Vp, and
productivity index, J, of the producer for its effective area:
sj¼ctVp
J

jð3Þ
and, fij is defined as the fraction injection rate of injector, i,
toward producer, j:
fij ¼qijðtÞ
iiðtÞð4Þ
The solution of this equation, including a variation in the
producer’sbottom-hole pressure (BHP), is the following [43]:
qjðtnÞ¼qjðt0Þetnt0
sj

!
zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{
Primary depletion
þet
sj
Zt
to
e
t
sj
1
sjX
Ni
i¼1
fijiiðnÞdt
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Injection input signal
et
sj
Zt
to
e
n
sj
Jj
dpwf;j
dndn
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
BHP variation
ð5Þ
I1
I2
I3
I4
I5
Injection
Rates
Liquid
Production
Rate
12 hidden
neurons
1 output
neuron
5 input
neurons
P
Fig. 6 Schematic of the
architecture of the neural
network constructed with 5
inputs, 1 output and 12 hidden
neurons
Neural Comput & Applic
123
which includes three components for representing the
production rate signal q(t) at any given time on the right-
hand side of the equation:
1. Primary depletion
2. Injection input signal
3. Variation in the producer’s bottom-hole pressure
(BHP)
By applying numerical integration, the integrals in the
above solution were evaluated by proposing two approaches
[45], which includes linear variation of BHP during the
consecutive time intervals, and either stepwise variation in
the injection rate (constant injection rate during a timestep),
or linearly varying injection rate during a timestep. In this
study, injection rates are kept constant during a timestep;
therefore, the former solution is utilized. For the case of
fixed injection rate of iðDtkÞ¼IðkÞ
i, and a linear BHP vari-
ation during time intervals Dtk,ðk¼1;2;...;nÞ,by
assuming a constant productivity index at any given time,
tn, total production rate of producer jcan be written as:
qjðtnÞ¼qjðt0Þetnt0
sj

!
þX
n
k¼1
etntk
sj
1eDtk
sj

!(
X
Ni
i¼1
fijIðkÞ
i
hi
Jjsj
DpðkÞ
wf ;j
Dtk
"#)
ð6Þ
where IðkÞ
iand DpðkÞ
wf ;jrepresent injection rate of injector, i,
and changes in the BHP of the producer, j, during time
interval, tk1to tk, respectively. The stepwise variation in
injection rates is consistent with the discrete nature of field
data that are typically reported in monthly averages [3]. If
the bottom-hole pressure for producing wells does not
change with time, the equation becomes:
qjðtnÞ¼qjðt0Þetnt0
sj

!
þX
n
k¼1
etntk
sj
1eDtk
sj

!
X
Ni
i¼1
fijIðkÞ
i
hi
"#()
ð7Þ
The history-matching process is achieved by inputting
observed production and injection rates for liquids and by
changing the unknown parameters:
Initial production rates, qjðt0Þ,
Time constant for each producer, j,sj,
Fraction injection rate of injector, i, toward producer, j,
fij (i.e., the connectivity between injector, i, and
producer, j.
Through an optimization routine, these parameters are
changed until the average error between observed and
calculated production rates is minimized. This error is
defined as:
MSE ¼PNdata
n¼1ðqobs qestÞ2
Ndata ð8Þ
where Ndata is the number of observations (sample size),
qobs is the observed flow rates, and qest is the flow rate
estimated by the model. The routine is initialized by
assuming values for the time constant and initial flow rates
and calculating the initial fractional flow parameter using
the inverse-distance method [46]:
fij ¼
1
dij
PNpro
j¼1
1
dij
ð9Þ
where dij represents the distance between each injector and
producer. After this initialization, the trust-region reflective
search algorithm [4751] is used to search for the combi-
nation of parameters that provides the lowest range of error
between the observed and estimated rates. After a pre-
specified stopping criteria are met, then the solution is
accepted as the optimum solution. An error tolerance of
1e-07 and a maximum number of function evaluations of
10,000 are used as the convergence criteria. Then, the
system parameters (e.g., fractional flow) are used to char-
acterize the reservoir.
3 Results and discussion
Proposed methods, namely ANNs and CRMs, are applied
to the case study presented in the previous section. The
primary objective was to quantify the connectivity between
injector/producer pairs using both methods. This is
achieved by a contribution parameter derived from the
trained neural network weights, w, in the case of ANNs and
by the fractional flow parameter, f, in the case of CRMs. In
the following subsections, results obtained using these two
methods are presented and discussed.
3.1 Artificial neural networks
The history-matching results are shown in Fig. 7. These
figures show that the training was successful in matching the
historical rates observed. Once the training is completed, it is
expected that the neural network would capture the dynamics
of the reservoir system from observed data. Since no pre-
sumed physical laws are introduced, we call such models
data-driven models. The model, during the training, captures
the physics of the process through the neural network train-
ing, which is an iterative procedure. After the training is
completed by satisfying certain stopping criteria, weights
remain in their optimum state, at which the neural network
can predict the output (production rate) within high levels of
accuracy. In that case, the optimum set of weights would
Neural Comput & Applic
123
represent the contribution of each injection well to the pro-
ducing well’s production, which can be used as a proxy to the
connectivity between injection and production wells. Higher
quantity of weights indicate stronger contribution and thus
stronger connectivity, and lower quantity weights indicate
weaker contributions and weaker connectivity. Contribution
factors calculated in this case are given in Table 2. After
calculating these contribution factors, a connectivity map is
drawn which represents the strength of connection between
each producing and injector well (Fig. 8). This figure is very
similar to the actual reservoir grid in which the permeabili-
ties are shown (Fig. 2). Since the neural network was able to
capture the high-permeability streaks in the reservoir system,
this gives us the confidence that artificial neural networks can
be used to characterize the connectivity of an oil reservoir
system which goes through water injection.
3.2 Capacitance–resistance models
The history-matching results are shown in Fig. 9. These
figures show that the training was successful in matching
the historical rates observed. Based on the CRM formula-
tion, the search process includes searching for the optimum
combination of initial production rates, fraction of flow and
0500 1000 1500 2000 2500 3000
1000
1500
2000
2500
3000
3500
4000
P1
Time, days
q, bbl/d
Actual
ANN
(a) Producer-1
0500 1000 1500 2000 2500 3000
100
150
200
250
300
P2
Time, days
q, bbl/d
Actual
ANN
(b) Producer-2
0500 1000 1500 2000 2500 3000
180
200
220
240
260
280
300
320
340
360
380
P3
Time, days
q, bbl/d
Actual
ANN
(c) Producer-3
0500 1000 1500 2000 2500 3000
1500
2000
2500
3000
3500
P4
Time, days
q, bbl/d
Actual
ANN
(d) Producer-4
Fig. 7 History matching of the producing wells during ANN training
Neural Comput & Applic
123
time constant parameters. The fraction of flow parameter is
used as a proxy to the interwell connectivity between each
injector and producer. The values obtained after the opti-
mization routine is completed are given in Table 3. Using
these values, similar to the ANN case, a connectivity map is
drawn which represents the strength of connection between
each producing and injector well (Fig. 10). As in the case
with ANN, CRM was able to capture both high-perme-
ability streaks in the reservoir model which are shown in
the reservoir grid (Fig. 2). Therefore, we also conclude that
in a similar fashion with artificial neural networks, capac-
itance–resistance models can also be used to characterize
the connectivity of an oil reservoir system which goes
through water injection.
3.3 Comparison of data-driven and reduced-physics
modeling approaches
One of the main objectives of this study is to provide a
comparison of these two methods for the practicing
reservoir engineer or asset team with respect to a number of
aspects. Table 4shows a comparison of these two methods
in different aspects, and a discussion of each aspect is
presented as the following:
Accuracy Prediction capabilities of each method can be
analyzed by comparing each method’s ability to inden-
tify high-connectivity zones in the reservoir. When all of
the 20 interwell connectivities between each of the five
injectors and four producers are ranked, upper 10 values
can be classified in the high-connectivity category, while
lower 10 values can be classified in the low-connectivity
category. One can approximate the connectivity values
for the numerical simulation model, by utilizing the
average permeability between two wells and normaliz-
ing the permeability by the distance between two wells
(if two wells are close to each other and have a high-
permeability streak between them, their connectivity
would be the highest). These values are given in Table 5.
After sorting the connectivity values in the numerical
model together with the predictions of data-driven and
the reduced-physics models, it is seen that the data-
driven and the reduced-physics models were able to
correctly estimate 80 and 70 % of each connectivity
category, respectively. These acceptable accuracy levels
indicate that both methods have similar prediction
capabilities, while the data-driven model has a slightly
better performance.
Data requirements Data-driven modeling approach (ar-
tificial neural networks) is better since it does not have
any limitation regarding the types of input/output data
set used. In this study, only injection and production
rates are used. This can be considered as a minimum
required set of data, since a set of signals would be
needed to be able to relate the connectivity between
wells, and production/injection rates are the most
commonly available data set that can be incorporated.
Meanwhile, the data included can be expanded, by
including well locations, known reservoir properties and/
or pressure data. Artificial neural networks are very
advantageous in terms of flexibility and can be modified
or restructured depending on the reservoir, wells, or any
other known aspects of the field that are studied. On the
other hand, capacitance–resistance models have a certain
formulation and require the data in that formulation
(production/injection rates, bottom-hole flowing pres-
sures, well locations). These are common type of data
available in any oil field; therefore, this does not create a
significant problem in implementing the method. How-
ever, it does not have the flexibility to incorporate more
data, if needed. In that case, the formulation needs to be
modified and redeveloped, which is not practical to do
for a quick application of the method but can be done in
Table 2 Contribution values obtained from weights, w, after ANN
training which is a proxy to the interwell connectivity
fP1 P2 P3 P4
I1 1.000 0.250 0.293 0.204
I2 0.117 0.177 0.521 0.165
I3 0.206 0.205 0.184 0.561
I4 0.239 0.903 0.373 0.438
I5 0.213 0.239 0.402 0.405
0500 1000 1500 2000 2500
−2200
−2000
−1800
−1600
−1400
−1200
−1000
−800
−600
−400
−200
X (m)
Y (m)
Connectivity
Producer
Injector
Fig. 8 Connectivity map of injectors/producers using the contribu-
tion values derived from the trained neural network
Neural Comput & Applic
123
the long term to have a suite of CRMs. This also can be
explained with the modeling approach utilized in each
method. ANNs are purely data-driven, and CRMs are
reduced-physics-driven as explained earlier.
Training algorithm ANNs offer a number of different
training algorithms, and all of them are purely data-
driven algorithms. The choice of the algorithm requires
subject-matter expertise in ANNs. ANNs do not limit the
use of any training algorithm and provide the flexibility
in choosing from the available options. With CRMs, the
0 500 1000 1500 2000 2500 3000
1000
1500
2000
2500
3000
3500
4000
P1
Time, days
q, bbl/d
Actual
CRM
(a) Producer-1
0 500 1000 1500 2000 2500 3000
100
150
200
250
300
P2
Time, days
q, bbl/d
Actual
CRM
(b) Producer-2
0 500 1000 1500 2000 2500 3000
180
200
220
240
260
280
300
320
340
360
380
P3
Time, days
q, bbl/d
Actual
CRM
(c) Producer-3
0500 1000 1500 2000 2500 3000
1500
2000
2500
3000
3500
P4
Time, days
q, bbl/d
Actual
CRM
(d) Producer-4
Fig. 9 History matching of the producing wells during CRM training
Table 3 Fraction of flow, f, values obtained after CRM training
which is a proxy to the interwell connectivity
fP1P2P3P4
I1 1.000 0.002 0.003 0.005
I2 0.548 0.003 0.137 0.303
I3 0.057 0.010 0.033 0.974
I4 0.117 0.178 0.001 0.662
I5 0.126 0.001 0.106 0.766
Neural Comput & Applic
123
training is basically an optimization process, in which
the optimization algorithm and parameters can be
modified. Therefore, CRMs are also can be trained with
a different number of optimization-related options.
Training speed Both methods are promising as the
training of both models took \30 s of CPU time. Even
for more complex reservoir systems, it is fair to expect
training times not more than a few minutes, which is a
reasonable amount of time for decision-making
purposes.
Expertise requirements Both methods are considered to
be moderate in terms of expertise requirements. In
developing a tool that utilizes either of the method, both
methods require significant expertise in certain subjects.
ANNs require to be familiar with the ANN theory and
terms; CRMs require the knowledge of CRM formula-
tion and reservoir engineering concepts. To train a
readily available tool using available data, ANNs still
require to be familiar with the related theory to
determine the neural network architecture and training
parameters. For training of CRMs, although not neces-
sary, knowledge of optimization algorithms and param-
eters would help to find the best optimization approach.
4 Summary and conclusions
In this study, two methods of different modeling approa-
ches, ANNs and CRMs, are studied to quantify the inter-
well connectivity between water-injecting wells and oil-
producing wells for a petroleum reservoir that has gone
through a reasonably long period of waterflooding. The
methods were tested on a synthetic reservoir case, in which
there are two high-permeability streaks (500 and 1000 md)
in a reservoir having a permeability of 5 md, elsewhere.
After calculating the connectivities from the model
parameters, two methods are compared with each other
considering various aspects from the practical points of
view.
Only liquid production and injection rates as well as
well locations are used as the data input to the two meth-
ods. Among these methods, artificial neural networks are
purely data-driven, with no assumption regarding the
governing laws of physics made. The other method,
capacitance–resistance models, can be defined as reduced-
physics-driven. The reason is that the physical laws
included in the formulation require a large number of
assumptions regarding the fluid flow. Both methods require
a training process, in which the developed model learns
from observed data, to capture the actual dynamics in the
reservoir. Therefore, success of both methods depends on
data quality and quantity. It was observed that for a known
reservoir scenario with 2 high-permeability streaks, both
methods were able to capture these streaks within reason-
able ranges of accuracy. Since both methods are practical
and easy to implement, a recommendation can be made
that both methods are applicable for such applications.
The key conclusions obtained from this study can be
summarized as follows:
0500 1000 1500 2000 2500
−2200
−2000
−1800
−1600
−1400
−1200
−1000
−800
−600
−400
−200
X (m)
Y (m)
Connectivity
Producer
Injector
Fig. 10 Connectivity map of injectors/producers using the contribu-
tion values derived from the trained capacitance–resistance model
Table 4 Characteristics of data-driven and reduced-physics modeling approaches in various aspects
Data-driven modeling (ANN) Reduced-physics modeling (CRM)
Accuracy (based on the specific example presented in this study) 80 % 70 %
Data requirements Flexible Fixed
Modeling approach Flexible data-driven model Physics-based, include assumptions
Training algorithm Flexible Flexible
Training speed Fast Fast
Expertise requirements Moderate Moderate
To develop the tool ANN background needed Reservoir eng. background needed
To train ANN background needed Optimization knowledge (for fine-tuning)
Neural Comput & Applic
123
1. Both ANNs and CRMs can be used to quickly estimate
the interwell connectivity between injection wells and
production wells in a reservoir. These tools have great
advantages over numerical modeling because of
requiring fewer inputs and being much more compu-
tationally efficient, while also providing the ability to
utilize available historical data and not being depen-
dent on geological uncertainties.
2. In the example presented here, 10 years of production/
injection history was sufficient for training and
achieving an accurate history matching. For more
complicated reservoir cases (with more number of
wells), higher duration of production and injection
histories might be needed.
3. Both methods are efficient in terms of CPU time
requirements with training times\30 s reported for the
example used in this study. Even for more complex
reservoir systems, this time is not expected to be
greater than a few minutes.
4. In terms of data requirements and modeling approach,
ANNs are more flexible than CRMs, since ANNs are
purely data-driven and do not require any presumed
functional relationship between process variables.
Instead, it derives the relationships through training
of observed data.
5. In both methods, some degree of knowledge is needed
for fine-tuning of results during the training process
which involves optimization of weights or the frac-
tional flow parameter.
Acknowledgments This work is supported by the Middle East
Technical University Northern Cyprus Campus (METU-NCC)—
Campus Research Fund; Project No. BAP-FEN-13-YG-2. The sup-
port is gratefully acknowledged.
Compliance with ethical standards
Conflict of interest We declare that there is no any conflict of
interest to report.
References
1. Hey T, Tansley S, Tole K (2009) The fourth paradigm: data-
intensive scientific discovery. Microsoft Research, Redmond
2. Jansen J, Brouwer R, Douma S (2009) Closed loop reservoir
management. In: SPE reservoir simulation symposium proceed-
ings. SPE 119098. 2–4 February, The Woodlands, Texas
3. Sayarpour M, Zuluaga E, Kabir C, Lake L (2007) The use of
capacitance-resistive models for rapid estimation of waterflood
performance and optimization. In: SPE annual technical confer-
ence and exhibition proceedings. SPE 110081. 11–14 November.
Anaheim, California
4. CMG (2013) CMG IMEX reservoir simulation software, version
2013. Computer Modeling Group, Ltd, Calgary
5. Sayarpour M, Kabir C, Lake L (2009) Field applications of
capacitance–resistance models in waterfloods. SPE Res Eval Eng
12(6):853–864
Table 5 List of connectivity
values from the numerical, data-
driven and reduced-physics
models
Well pair Connectivity
No. Injector Producer Numerical Data-driven Reduced-physics
1 I1 P1 1.000 1.000 1.000
2 I2 P1 0.233 0.548 0.117
3 I3 P1 0.169 0.057 0.206
4 I4 P1 0.104 0.117 0.239
5 I5 P1 0.104 0.126 0.213
6 I1 P2 0.233 0.002 0.250
7 I2 P2 0.104 0.003 0.177
8 I3 P2 0.169 0.010 0.205
9 I4 P2 0.233 0.178 0.903
10 I5 P2 0.104 0.001 0.239
11 I1 P3 0.104 0.003 0.293
12 I2 P3 0.233 0.137 0.521
13 I3 P3 0.169 0.033 0.184
14 I4 P3 0.104 0.001 0.373
15 I5 P3 0.233 0.106 0.402
16 I1 P4 0.104 0.005 0.204
17 I2 P4 0.104 0.303 0.165
18 I3 P4 0.651 0.974 0.561
19 I4 P4 0.233 0.662 0.438
20 I5 P4 0.233 0.766 0.405
Neural Comput & Applic
123
6. Artun E, Mohaghegh S (2011) Intelligent seismic inversion
workflow for high-resolution reservoir characterization. Comput
Geosci 37(2):143–157
7. Raeesi M, Moradzadeh A, Ardejani FD, Rahimi M (2012)
Classification and identification of hydrocarbon reservoir litho-
facies and their heterogeneity using seismic attributes, logs data
and artificial neural networks. J Pet Sci Eng 82–83:151–165
8. Das V, Chatterjee R (2012) Prediction of coal-bed permeability
using artificial neural network. In: Proceedings of 9th biennial
international conference and exposition on petroleum geophysics,
Paper ID: 108, 16–18 February, Hyderabad, India
9. Alizadeh B, Najjari S, Ali KI (2012) Artificial neural network
modelling and cluster analysis for organic facies and burial his-
tory estimation using well log data: a case study of the South Pars
Gas Field, Persian Gulf, Iran. Comput Geosci 45:261–269
10. Mohaghegh S, Balan B, McVey D, Ameri S (1996) A hybrid
neuro-genetic approach to hydraulic fracture treatment design
and optimization. In: SPE annual technical conference and
exhibition proceedings. SPE 36602. 6–9 October. Denver,
Colorado
11. Centilmen A, Ertekin T, Grader A (1999) Applications of neural
networks in multiwell field development. In: SPE annual tech-
nical conference and exhibition proceedings. SPE 56433. 3–6
October. Houston, Texas
12. Doraisamy H, Ertekin T, Grader A (2000) Field development
studies by neuro-simulation: an effective coupling of soft and
hard computing protocols. Comput Geosci 26(8):963–973
13. Mohaghegh S, Modavi A, Hafez H, Haajizadeh M, Kenawy M,
Guruswamy S (2006) Development of surrogate reservoir models
(SRM) for fast-track analysis of complex reservoirs. In: SPE
intelligent energy conference and exhibition proceedings. SPE
99667. 11–13 April. Amsterdam, The Netherlands
14. Johnson V, Rogers L (2001) Applying soft computing methods to
improve the computational tractability of a subsurface simula-
tion-optimization problem. J Pet Sci Eng 29(3–4):153–175
15. Guyaguler B (2002) Optimization of well placement and
assessment of uncertainty. Ph.D. dissertation, Stanford Univer-
sity, Stanford, California
16. Yeten B, Durlofsky L, Aziz K (2003) Optimization of noncon-
ventional well type, location, and trajectory. SPE J 8(3):200–210
17. Patel AN, Davis D, Guthrie CF, Tuk D, Nguyen T, Williams J
(2005) Optimizing cyclic steam oil production with genetic
algorithms. In: SPE western regional meeting proceedings. SPE
93906. 30 March–1 April, Irvine, California
18. Ayala L, Ertekin T (2005) Analysis of gas-cycling performance
in gas/condensate reservoirs using neuro-simulation. In: SPE
annual technical conference and exhibition proceedings. SPE
95655. 9–12 October. Dallas, Texas
19. Artun E, Ertekin T, Watson R, Miller B (2010) Development and
testing of proxy models for screening cyclic pressure pulsing
process in a depleted, naturally fractured reservoir. J Pet Sci Eng
73(1):73–85
20. Artun E, Ertekin T, Watson R, Al-Wadhahi M (2011) Develop-
ment of universal proxy models for screening and optimization of
cyclic pressure pulsing in naturally fractured reservoirs. J Nat Gas
Sci Eng 3(6):667–686
21. Artun E, Ertekin T, Watson R, Miller B (2012) Designing cyclic
pressure pulsing in naturally fractured reservoirs using an
inverse-looking recurrent neural network. Comput Geosci
38(1):68–79
22. Parada CH, Ertekin T (2012) A new screening tool for improved
oil recovery methods using artificial neural networks. In: SPE
western regional meeting proceedings. SPE 153321. 19–23
March, Bakersfield, California
23. Durzman PJ, Leung J, Zanon SD, Amirian E (2013) Data-driven
modeling approach for recovery performance prediction in
SAGD operations. In: Proceedings of the SPE heavy oil confer-
ence. SPE 165557. 11–13 June, Calgary, Alberta, Canada
24. Cullick A, Johnson D, Shi G (2006) Improved and more rapid
history matching with a nonlinear proxy and global optimization.
In: SPE annual technical conference and exhibition proceedings.
SPE 101933. 24–27 September. San Antonio, Texas
25. Silva P, Clio M, Schiozer D (2007) Use of neuro-simulation
techniques as proxies to reservoir simulator: application in pro-
duction history matching. J Pet Sci Eng 57(3–4):273–280
26. Zhao H, Kang Z, Zhang X, Sun H, Cao L, Reynolds A (2015)
INSIM: a data-driven model for history matching and prediction
for waterflooding monitoring and management with a field
application. In: Proceedings of the SPE reservoir simulation
symposium. SPE 173213. 23–25 February, Houston, Texas, USA
27. Zangl G, Giovannoli M, Stundner M (2006) Application of arti-
ficial intelligence in gas storage management. In: SPE Europec/
EAGE annual conference and exhibition proceedings. SPE
100133. 12–15 June. Vienna, Austria
28. Mohaghegh S, Al-Mehairi Y, Gaskari R, Maysami M, Khazaeni
Y, Gashut M, Al-Hammadi AE, Kumar S (2014) Data-driven
reservoir management of a giant mature oilfield in the Middle
East. In: Proceedings of the SPE annual technical conference
and exhibition. SPE 170660. 27–29 October. Amsterdam, The
Netherlands
29. Esmaili S, Mohaghegh S (2015) Full field reservoir modeling of
shale assets using advanced data-driven analytics. Geosci Front.
doi:10.1016/j.gsf.2014.12.006
30. Kalantari-Dhaghi A, Mohaghegh S, Esmaili S (2015) Data-driven
proxy at hydraulic fracture cluster level: a technique for efficient
CO
2
-enhanced gas recovery and storage assessment in shale
reservoir. J Nat Gas Sci Eng. doi:10.1016/j.jngse.2015.06.039
31. Giraldo R, Delicado P, Mateu J (2012) Hierarchical clustering of
spatially correlated functional data. Stat Neerl 66(4):403–421
32. Zhou J, Duan B, Huang J, Hongjun C (2014) Data-driven modeling
and optimization for cavity filters using linear programming sup-
port vector regression. Neural Comput Appl 24(7–8):1771–1783
33. Dong J, Zheng C, Kan G, Zhao M, Wen J, Yu J (2015) Applying
the ensemble artificial neural network-based hybrid data-driven
model to daily total load forecasting. Neural Comput Appl
26(3):603–611
34. Fausett L (1994) Fundamentals of neural networks: architectures,
algorithms, and applications. Prentice-Hall, Englewood Cliffs
35. Neuroshell (1998) Neuroshell 2 tutorial. Ward Systems Inc,
Frederick
36. Demuth H, Beale M (2002) Neural network toolbox for use with
MATLAB. Mathworks Inc, Natick
37. MATLAB (2013) MATLAB neural network toolbox, version
2013a. Mathworks Inc, Natick
38. Bruce W (1943) An electrical device for analyzing oil-reservoir
behavior. Trans AIME 151(1):112–124 SPE-943112-G
39. Weber D. (2008) The use of capacitance–resistance models to
optimize injection allocation and well location in waterfloods.
Ph.D. dissertation, The University of Texas at Austin, Austin,
Texas
40. Biswas D (2011) Techniques enhance reservoir studies. The
American Oil and Gas Reporter 54(7). Derby, Kansas
41. Nguyen A, Kim J, Lake L, Edgar T, Haynes B (2011) Integrated
capacitance-resistive model for reservoir characterization in pri-
mary and secondary recovery. In: SPE annual technical confer-
ence and exhibition proceedings. SPE 147344. 30 October–2
November. Denver, Colorado
42. Kim J, Lake L, Edgar T (2012) Integrated capacitance–resistance
model for characterizing waterflooded reservoirs. In: IFAC
workshop on automatic control in offshore oil and gas production
proceedings. Norwegian University of Science and Technology,
Trondheim, Norway
Neural Comput & Applic
123
43. Yousef A, Jensen J, Lake L (2006) A capacitance model to infer
interwell connectivity from production and injection rate fluctu-
ations. SPE Res Eval Eng 9(6):630–646
44. Sayarpour M, Zuluaga E, Kabir C, Lake L (2009) The use of
capacitance-resistive models for rapid estimation of waterflood
performance and optimization. J Pet Sci Eng 69(3–4):227–238
45. Liang X, Weber B, Edgar T, Lake L, Sayarpour M, Al-Yousef
A(2007) Optimization of oil production based on a capacitance
model of production and injection rates. In: SPE hydrocarbon
economics and evaluation symposium proceedings. SPE 107713.
1–3 April. Dallas, Texas
46. Shephard D (1968) A two-dimensional interpolation for irregu-
larly-spaced data. In: 23rd ACM national conference proceed-
ings, New York, pp 517–524
47. More
´JJ, Sorensen DC (1983) Computing a trust region step.
SIAM J Sci Stat Comput 3:553–572
48. Steihaug T (1983) The conjugate gradient method and trust
regions in large scale optimization. SIAM J Numer Anal
20:626–637
49. Byrd RH, Schnabel RB, Shultz GA (1988) Approximate solution
of the trust region problem by minimization over two-dimen-
sional subspaces. Math Program 40:247–263
50. Branch MA, Coleman TF, Li Y (1999) A subspace, interior, and
conjugate gradient method for large-scale bound-constrained
minimization problems. SIAM J Sci Comput 21(1):1–23
51. MATLAB (2013) MATLAB optimization toolbox, version
2013a. Mathworks Inc, Natick
Neural Comput & Applic
123
... So far in research fields other than nuclear energy, some comparison of model-driven and data-driven approaches has been studied [33][34][35][36]. Generally speaking, the aspects of comparison include accuracy, the data, expertise and algorithm for model building, running speed, etc. ...
... Some comparison of model-driven and data-driven approaches has been studied [33][34][35][36] in research fields other than nuclear energy. Generally speaking, the aspects of comparison include accuracy, the data, expertise and algorithm for model building, running speed, etc. Considering nuclear safety, the interpretability of algorithms and models also needs to be included in the comparison. ...
Article
Nuclear energy plays a crucial role in mitigating climate change as a near carbon-free and clean energy source. Predicting operating parameters is key to the digitalization and intellectualization of nuclear power plants, improving energy efficiency and reducing costs. Parameter prediction methods mainly consist of model-driven and data-driven approaches, and a comparative study is necessary to select the appropriate prediction method or combine them. In this paper, the gated recurrent unit network and the thermal–hydraulic program RELAP5 are chosen as representative data-driven and model-driven approaches to be used for predicting parameters for a specific operating condition to assess the characteristics and capabilities of each. The chosen experiment is a steam generator tube rupture accident which is important to pressurized water reactors. In this paper the modelling process and the prediction results for both approaches are given. The correlation coefficient for dome pressure, dome temperature, and primary outlet temperature of both approaches is greater than 0.993, except that the correlation coefficient of downcomer temperature is 0.83865 using RELAP5 and 0.99392 using gated recurrent unit network. Then the two approaches are compared from six perspectives of practical application, including accuracy (gated recurrent unit network and RELAP5 received scores of 8 and 6, respectively.), data requirements (7 and 5, respectively), model building algorithm (7 and 5, respectively), prediction speed (10 and 6, respectively), expertise requirements (6 and 6, respectively), and interpretability (5 and 9, respectively). It shows that model-driven and data-driven approaches both require specialist knowledge and appropriate data for modelling process. The data-driven gated recurrent unit network exhibits better accuracy and higher speed. The model-driven RELAP5 has better interpretability compared to the black-box gated recurrent unit network. It is suggested that hybrid approaches of model-driven and data-driven approaches could better deal with the prediction problems.
... 20 Reduced-order models have been used to model thermal recovery processes 21 and transient CO 2 and brine leakage, 22 response surface models have been used for purposes such as history matching, 23 and reduced physics models have been used for purposes such as waterflooded reservoirs. 24 The weaknesses of reduced order models can be reduced physics and space-time resolution. In addition, statistics approaches (response surfaces) pose two main problems when they are applied to proxy models, such as numerical reservoir simulation models: the issue of "correlation versus causation" and the assumption of a linear, polynomial, and so on, function for the data being analyzed or modeled. ...
... For example, Panda and Chopra [26] use neural networks to determine well connectivities. Artun [1,2] uses feedforward ANN to quantify well pair connectivities and determine reservoir faults. Cheng et al. [7] use global sensitivity analysis based on long short-term memory (LSTM) to quantify the connectivities between producers and injectors, and other related studies can be found in [18,32]. ...
Article
Full-text available
In mature reservoirs, there are hundreds or thousands of producing and injecting wells operating simultaneously, so it is important to understand the impact of injection wells on producers to maintain pressure and control water production. In this work, we propose a workflow with two strategies, reduced-physics and data-driven modeling, to monitoring producer and injector wells based on interwell connectivity. The monitoring the wells allows to increase oil production, reducing water rate, and avoiding possible fracturing or fault reactivations. Both strategies use production history data only. The inputs in both strategies are injection rates, while output are liquid production rates. The first one, the reduced-physics modeling strategy, is based on the capacitance-resistance modeling for producers (CRMP), which calculates the liquid flowrate of the producing well based on the injection rate, productivity index of producers, time constant, and the connectivity between injectors and producers. The parameters of the CRMP model are obtained by minimizing the error between the observed and calculated liquid flowrates. The optimization algorithm that minimizes the error is the Sequential Quadratic Programming (SQP) and the gradient is obtained by finite differences. The second one, the data-driven modeling strategy is based on artificial neural networks (ANNs), which only use input and output data. The parameters of the artificial neural network, weights, and biases, are adjusted during the training process. Three architectures are proposed to match the outputs based on the inputs: single-layer perceptron, deep learning with multiple layers, and convolutional neural networks. The backpropagation algorithm is used to adjust the weights and biases of the architectures during training. In this study, we propose three alternatives for calculating the connectivities based on the trained model. The first one is based on the optimal weights. The second one is based on the average error after training and shuffling the input data, and the last one is based on the gradient importance. Two synthetic models, Two-phases, and Brush Canyon Outcrop, are used to validate the proposed workflow. The results show that the connectivities calculated by the gradient importance approach are closer to the connectivities obtained by the capacitance-resistance model. On the other hand, the connectivities obtained through the optimal weights and average error strategies show differences of 4% and 5%, respectively.
... Machine-learning (ML)-based approaches are expected to model the complicated injection-production relationships via their powerful learning ability. The multilayer perceptron (MLP) is used in the simulation of waterflooding reservoirs, which can reflect the inter-well connectivity by computing the products of the weight matrices between the input-hidden-layer and hidden-output-layer [17,18]. Furthermore, Mo et al. [19] employ a deep convolutional encoder-decoder network to evaluate the uncertainty of the subsurface flow among the porous media, to tackle the high-dimensional saturation discontinuity problems in the surrogate strategies. ...
Article
Full-text available
Waterflooding is one of the most common reservoir development programs, driving the oil in porous media to the production wells by injecting high-pressure water into the reservoir. In the process of oil development, identifying the underground flow distribution, so as to take measures such as water plugging and profile control for high permeability layers to prevent water channeling, is of great importance for oilfield management. However, influenced by the heterogeneous geophysical properties of porous media, there is strong uncertainty in the underground flow distribution. In this paper, we propose an interpretable recurrent neural network (IRNN) based on the material balance equation, to characterize the flow disequilibrium and to predict the production behaviors. IRNN is constructed using two interpretable modules, where the inflow module aims to compute the total inflow rate from all injectors to each producer, and the drainage module is designed to approximate the fluid change rate among the water drainage volume. On the spatial scale, IRNN takes a self-attention mechanism to focus on the important input signals and to reduce the influence of the redundant information, so as to deal with the mutual interference between the injection–production groups efficiently. On the temporal scale, IRNN employs the recurrent neural network, taking into account the impact of historical injection signals on the current production behavior. In addition, a Gaussian kernel function with boundary constraints is embedded in IRNN to quantitatively characterize the inter-well flow disequilibrium. Through the verification of two synthetic experiments, IRNN outperforms the canonical multilayer perceptron on both the history match and the forecast of productivity, and it effectively reflects the subsurface flow disequilibrium between the injectors and the producers.
Conference Paper
Water-alternating-gas (WAG) injection is an enhanced oil recovery (EOR) technique used to overcome problems related with gas injection, including gravity override, viscous fingering, and channeling. However, the success of WAG injection is influenced by reservoir characteristics, the injector-producer connections, and the injected gas water ratio (GWR). This study proposes a hybrid model that combines the Capacitance Resistance Model (CRM), machine learning model (ML) and oil model to assess and optimize WAG injectors in a carbonate field. Interwell connectivity between injection and production wells is obtained by combining a physics-based reduced order model (CRM) and a machine learning model (ML). The hybrid combination of CRM and ML increases confidence in results, eliminating shortcomings associated with CRM models used alone. Next, the results obtained from injection connectivity are combined with reduced order Power Law oil model to evaluate the impact of injection rate increase on oil production. Finally, injectors are optimized and ranked based on their potential to contribute to oil production. The proposed workflow is applied to a large, complex carbonate field with more than 93 production wells and 47 injection wells. The hybrid combination of CRM and ML used to obtain interwell connectivity compares wells with each other, resulting in higher confidence in the results. Using an ML, results obtained from multiple signals are aggregated to further identify and verify the injector-producer pair connectivity. The further combination of CRM with an oil model helps evaluate additional oil recovery based on the changing gas-water ratio (GWR). A combination of CRM optimization and the ML connectivity results help rank and prioritize 47 injectors quickly and accurately. Five injectors are selected for field testing and results show significant improvement in oil production after implementation of suggested injection schedules. Hybrid models that combine CRM and ML to obtain connectivity address the shortcomings of the individual models. The identification of efficient injectors for WAG injection by the novel use of hybrid models accelerates decision-making. The approach presented can be extended to similar WAG injection fields with many injectors and producers to help optimize the injection strategy. This new approach helps with current digitization strategies in oil companies to accelerate decision making, especially in mature reservoirs where history matching is not available.
Conference Paper
Full-text available
In mature reservoirs, the sweep efficiency of waterflooding is frequently improved by polymer injection. The effectiveness of polymer injection can vary depending on the complexity of the reservoir and chosen injection pattern. As against the traditional reservoir simulation approach that is typically used to evaluate patterns, a novel approach is presented to combine connectivity (obtained from production/injection data history) and remaining oil saturations. This novel workflow evaluates polymer injection patterns and predicts their impact on oil production. This study proposes a hybrid model that combines reduced physics models - Capacitance Resistance Model (CRM), the Buckley Leverett (BL) oil model and machine learning model (ML) model to assess and rank polymer injection patterns for injection optimization strategies. More specifically, interwell connectivity between injection and production wells are obtained by combining physics-based reduced order models (CRM) and machine learning (ML) models. The hybrid combination of CRM and ML helps to provide confidence in results, eliminating shortcomings associated with just using CRM models. Finally, the results obtained from injection connectivity are combined with BL to model oil production rate to evaluate the impact of polymer viscosity on oil production. These first order oil prediction results are further used to rank patterns for polymer injection. The proposed workflow is applied to a complex, large waterflood field in Oman with more than 150 production and 43 injection wells. The hybrid combination of CRM and ML models used to obtain interwell connectivity compares well with tracer data resulting in higher confidence in the results. Using ML model, results obtained from multiple signals are aggregated to further identify and verify the injector-producer pair connectivity. The combination of CRM with BL model helps to evaluate additional oil recovery based on the changing viscosity of the displacing fluid. A combination of remaining oil saturation maps of producers and the impact of displacing fluid viscosity from the model helped rank 43 polymer injection patterns quickly and accurately. Hybrid models that combine CRM and ML to obtain connectivity address shortcomings of individual models such as CRM. The combination of interwell connectivity with remaining oil saturations to rank patterns for polymer injection is unique. The identification of waterflood patterns to choose for injecting polymer by novel use of hybrid models accelerates decision making and is directly applicable as part of Well and Reservoir Management (WRM) for large mature fields in the Middle East with waterflood in patterns and especially, where polymer injection is being considered. The presented approach can accelerate reservoir surveillance and reservoir management decision making as compared to traditional reservoir simulation workflows.
Conference Paper
Water flooding is an established method of secondary recovery to increase oil production in conventional reservoirs. Analytical models such as capacitance resistance models (CRM) have been used to understand the connectivity between injectors and producers to drive optimization. However, these methods are not applicable to waterflood fields at the initial stage of life with limited data (less than 2 years of injection history). In this work, a novel approach is presented that combines analytics and machine learning to process data and hence quantify connectivity for optimization strategies. A combination of statistical (cross-correlation, mutual information) and machine learning (linear regression, random forest) methods are used to understand the relationship between measured injection and production data from wells. This workflow is first validated using synthetic simulation data with known reservoir heterogeneities as well as known connectivity between wells. Each of the four methods is validated by comparing the result with the CRM results, and it was found that each method provides specific insights and has its associated limitations making it necessary to combine these results for a successful interpretation of connectivity. The proposed workflow is applied to a complex offshore Caspian Sea field with 49 production wells and 8 injection wells. It was observed that implementing the diffusivity filter in the models while being computationally expensive, offers additional insights into the transmissibility between injector producer pairs. The machine learning approach addresses injection time delay through feature engineering, and applying a diffusive filter determines effective injection rates as a function of dissipation through the reservoir. Hence, the combined interpretation of connectivity from the different methods resulted in a better understanding of the field. The presented approach can be extended to similar waterflood systems helping companies realize the benefits of digitization, in not just accessing data, but also using data through such novel workflows that can help evaluate and continuously optimize injection processes.
Article
Opportunistic Physics-mining Transfer Mapping Architecture (OPTMA) is a hybrid architecture that combines fast simplified physics models with neural networks in order to provide significantly improved generalizability and explainability compared to pure data-driven machine learning (ML) models. However, training OPTMA remains computationally inefficient due to its dependence on gradient-free solvers or back-propagation with supervised learning over expensively pre-generated labels. This paper presents two extensions of OPTMA that are not only more efficient to train through standard back-propagation but are readily deployable through the state-of-the-art library, PyTorch. The first extension, OPTMA-Net, presents novel manual reprogramming of the simplified physics model, expressing it in Torch tensor compatible form, thus naturally enabling PyTorch's in-built Auto-Differentiation to be used for training. Since manual reprogramming can be tedious for some physics models, a second extension called OPTMA-Dual is presented, where a highly accurate internal neural net is trained apriori on the fast simplified physics model (which can be generously sampled), and integrated with the transfer model. Both new architectures are tested on analytical test problems and the problem of predicting the acoustic field of an unmanned aerial vehicle. The interference of the acoustic pressure waves produced by multiple monopoles form the basis of the simplified physics for this problem statement. An indoor noise monitoring setup in motion capture environment provided the ground truth for target data. Compared to sequential hybrid and pure ML models, OPTMA-Net/Dual demonstrate several fold improvement in performing extrapolation, while providing orders of magnitude faster training times compared to the original OPTMA.