Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

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

net/publication/307942129

Use of Transient Tests to Monitor Progress of Flooding in IOR / EOR Operations

Conference Paper · September 2016


DOI: 10.2118/181473-MS

CITATIONS READS
3 326

3 authors, including:

Medhat Kamal Chuan Tian

44 PUBLICATIONS   372 CITATIONS   
Stanford University
8 PUBLICATIONS   117 CITATIONS   
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Well Test Analysis/Pressure Transient Analysis With Feature-Based Machine Learning View project

Pressure Transient Analysis of Polymer Flooding With Coexistence of Non-Newtonian and Newtonian Fluids View project

All content following this page was uploaded by Chuan Tian on 05 August 2018.

The user has requested enhancement of the downloaded file.


SPE-181473-MS

Transient Tests of Polymer Flooding With Coexistence of Non-Newtonian


and Newtonian Fluids
Medhat Kamal, Chevron; Chuan Tian, Stanford University; and Suleen, Chevron

Copyright 2016, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in Dubai, UAE, 26–28 September 2016.

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
Improved and enhanced oil recovery methods (IOR and EOR, respectively) are used to increase recovery
from proven reserves mainly following waterflooding. Monitoring and managing the progress of flood in
IOR and EOR operations is currently a challenge to the oil industry especially in situations with large well
spacing and cost prohibitive measures like drilling observation wells (e.g., offshore and deep water
applications). Falloff tests have been proven successful under waterflooding operations to determine the
reservoir properties in various banks around injection wells and the location of flood fronts. In this paper
we present a new development that extends transient testing and analysis technology to IOR and EOR
operations during polymer flooding. With the expanded use of Permanent Downhole Pressure Gauges
(PDHG), the new developed technique can be used without additional testing cost or interruption of field
operations.

In this paper, the effects of polymer are described by shear rate dependent viscosity (non-Newtonian flow).
We developed an analytical solution of wellbore pressure by combining the non-Newtonian fluids and the
multi-composite reservoir models. The solution does not only address the polymer region where the fluids
follow either the power-law or the Meter’s model, but also the Newtonian flow in the oil or water region
ahead of the polymer with varying Newtonian and non-Newtonian fluids saturations in both regions. The
developed solution was validated by analyzing synthetic data generated using a commercial numerical
reservoir simulator. In secondary recovery operations, the Newtonian fluid ahead of the polymer bank is
usually oil and in tertiary recovery operations the Newtonian fluid is usually the water used in
waterflooding. The solution provides a deeper understanding about the physics behind the transient
pressure behaviors during polymer flooding, and can be applied to guide a better implementation of well
tests. Interpretation method for falloff tests using the new solution and the conventional Bourdet derivative
and Horner plots is presented indicating that existing commercial well testing software are sufficient to
analyze data with the recent development. The new solution allows us to obtain the reservoir properties
such as fluid mobilities in various banks and the location of the flood front.

The developed solution was applied to field data. The pressure behavior expected from the new solution
was observed in the field data validating our developed technique and yielding the characterization of
reservoir parameters in various banks. Field application results are shown in the paper.

The novelty of this method of characterizing the dynamic properties of the various banks during injection
2 SPE-181473-MS

of non-Newtonian fluids and the location of the flood fronts is that an analytical solution of pressure
transient behavior in two phase flow of non-Newtonian fluids and Newtonian fluids was developed,
validated and used to analyze field data. This is the first analytical solution published to address this
situation.

Introduction
Pressure transient analysis is an important way of reservoir characterization, and it provides valuable
information about permeability, reservoir boundary, wellbore storage, etc. Falloff tests have been proven
successful under waterflooding operations to determine the reservoir properties in various banks around
injection wells and the location of flood fronts (Abbaszadeh and Kamal 1989, Yeh and Agarwal 1992,
and Kamal 2009). However, there have been limited studies about pressure transients during polymer
flooding, mainly due to the complexity in the coexistence of Newtonian fluids (oil and water) and non-
Newtonian fluids (polymer). Some studies investigated the analytical and numerical pressure transient
solutions in the polymer phase, without consideration of the oil region: Van Poollen and Jargon (1969),
Ikoku and Ramey (1978), Odeh and Yang (1979), Vongvuthipornchai and Raghavan (1987), Katime-
Meindl and Tiab (2001), Mahani et al. (2011) and Li and Delshad (2014). Studies of transient solutions
with oil and Newtonian displacing fluids (i.e., water) were reported by Abbaszadeh and Kamal (1989).
Other studies on the two-phase flow of non-Newtonian and Newtonian fluids used numerical methods,
e.g. finite difference method (Lund and Ikoku 1981, Ertekin et al. 1987, and van den Hoek et al. 2012).
As far as we know, there is still no study of the pressure transient behavior of non-Newtonian and
Newtonian fluids using the analytical approach. There are several ongoing polymer flooding projects
where well testing data may be collected. The objective of this study was to develop an analytical solution
that models the pressure transients during polymer flooding, in order to better characterize the reservoir.

In this project, the effects of polymer were described by the shear rate dependent viscosity (non-Newtonian
flow). We developed an analytical solution of wellbore pressure by combining the non-Newtonian fluids
model and the multicomposite reservoir model. The developed solution was validated by comparing it
with numerical simulation results. The solution provides deeper understanding of the physics behind the
pressure behaviors during polymer flooding, and can be applied to guide a better implementation of well
tests. Besides, we propose an interpretation method using the conventional Bourdet derivative and Horner
plots. Using the developed solution and interpretation method, we are able to obtain the reservoir
properties such as fluid mobilities in different regions and the location of the flood front. Finally, the
developed solution was applied to field data. The characteristics of the developed solution were observed
in the field data. Matching our developed solution to the field data provided information about different
reservoir parameters.

The Newtonian fluid that is being displaced by polymer may be oil in secondary recovery operations or
water following waterflooding in tertiary recovery operations. For sake of brevity and clarity in the rest
of this text, we consider the Newtonian fluid to be oil. Results are equally applicable if the Newtonian
fluid is water.

Background Information
Polymer has several different effects: shear rate dependent viscosity (non-Newtonian flow), adsorption,
permeability reduction, etc. In this study, we were only focused on the most important one: shear rate
dependent viscosity. There are many models proposed to describe the correlation between shear rate and
viscosity. One of the most widely used models is the Ostwald-de Waele power-law (Ostwald 1929), which
is given by:
SPE-181473-MS 3

  K n1 , (1)
where  is the viscosity, K is the Ostwald-de Waele power-law coefficient and  is the shear rate.
Different values of n represent different types of fluids. n  1 represents Newtonian fluids, whose
viscosity is independent of shear rate (   K ). For 0  n  1 , the viscosity decreases as shear rate
increases. Such pseudoplastic behavior is observed in many EOR polymers, e.g. xanthan polymer, and
partially hydrolyzed polyacrylamide (HPAM) below a critical shear rate (Chauveteau 1981; Sorbie and
Roberts 1984). This study is only focused on pseudoplastic fluids.

One favorable property of the power-law is the linear correlation between viscosity and shear rate on a
log-log plot, as shown in Figure 1. The slope of the straight line is n  1 , and its intercept with line   1
sec-1 is given by K . This also provides a way to characterize the power-law fluids (values of n and K )
given information on viscosity and shear rate. For pseudoplastic fluids, the straight line shows a negative
slope, which indicates high viscosity at low shear rate and low viscosity at high shear rate.

10
Viscosity (cp)

0.1
0.1 1 10 100
Shear rate (sec-1)
Figure 1: Viscosity versus shear rate of pseudoplastic fluids with K=1.5 and n=0.3.

As polymer being injected into the reservoir, a polymer region is formed close to wellbore, and it pushes
the oil in place. A conceptual visualization of the radial flow model is shown in Figure 2. The flow in the
polymer region (red) honors the power-law. Because shear rate varies at different radial distance, the
viscosity is also different. However, the viscosity in the oil region (yellow) remains constant (   o ) as
the oil region is not affected by the polymer. In this study, we are interested in not only the polymer region
where the power-law model applies, but also the oil region with Newtonian fluid flow. Therefore, the
objective is to develop an approach that combines the multicomposite reservoir model with the power-
law model. Two methods, namely Method A and Method B, are proposed in the next section.

Figure 2: Conceptual model of radial fluid flow in polymer flooding.


4 SPE-181473-MS

Methodologies
To model the flow of non-Newtonian and Newtonian fluids in porous media, the following assumptions
are made: (1) the flow is radial and isothermal, (2) non-Newtonian fluids (e.g. polymer solutions) displace
single-phase Newtonian fluids (e.g. oil), (3) the formation thickness is constant, (4) the reservoir is
homogeneous and isotropic, (5) the fluids are slightly compressible, (6) gravitational effects are negligible,
(7) the non-Newtonian fluids are pseudoplastic. In addition, Method A assumes the non-Newtonian fluids
obey the Ostwald-de Waele power-law, and Method B assumes the non-Newtonian fluids obey the
Meter’s viscosity model (Meter and Bird 1964) which is introduced later in this paper. The pressure
behaviors of rheo-thickening fluids (Bagassi et al. 1989; Dupas et al. 2013) are not discussed in this paper.

Given those assumptions, there are certain limitations when applying the developed solution to model the
full complexities of polymer flooding, e.g. adsorption, degradation (the effects of mechanical degradation
were discussed in detail by Seright 1983, Sorbie and Roberts 1984). In that case, a full-physics numerical
model can be applied to compare against the developed analytical solution. Nevertheless, the model
developed in this study captures the main physics in polymer flooding—the flow of non-Newtonian and
Newtonian fluids in porous media, with an efficient analytical approach.

Method A

One intuitive thought is to discretize the power-law model into multiple banks. With different n values
in each bank, the discretized power-law model is able to model both non-Newtonian and Newtonian fluids.
The procedure to develop the analytical solution in such discretized power-law model is summarized as
follows (details in Appendix A):
a. Start from the governing equation of the original power-law model (Ikoku and Ramey 1978;
Vongvuthipornchai and Raghavan 1987):
n 1
 2 p n p p  qB  c
  r1 n Ff   . (2)
r r r t  2 h 
2
k
b. Discretization:
n j 1
2 p j n j p j 1 n j p j  qB   c 
 r Ffj     . (3)
r r r t  2 h   k j
2

c. Non-dimensionalization:
p jD  c / k  j Ffj  qB /  2 hrw  
n 1
 2 p jD n j p jD
j
1 n j
 r . (4)
rD2 rD rD t D  c / k 1 Ff 1  qB /  2 hrw   n1 1
D
 
d. Laplace transform:
 1


 

p jD  rD j  Aj K vj rD j zF j /  j  B j I vj rD j zF j /  j  .
   (5)

e. Solving for dimensionless pressure in Laplace space:

 z  z   z  z 
K v1    1 I v1    s z  K1v1    1 I1v1  
1  1   1    1   1  
pwD   ,
z  z   z  z    z   z  z   
zK1v1    zCD  K v1    s zK1v1     1  zI1v1    zCD  I v1    s zI1v1   
 1    1   1     1    1   1   
(6)
SPE-181473-MS 5

where  1  (1  n1 ) / (3  n1 ) , 1  (3  n1 ) / 2 . n1 is the Ostwald-de Waele power-law index of the


bank closest to the wellbore.
f. Laplace inversion to obtain pwD using Stehfest algorithm.
g. Dimensionlization to obtain pw .

The discretization in step (b) follows the same procedure proposed by Abbaszadeh and Kamal (1989).
Both the polymer region and the Newtonian fluid region are discretized into a series of radial banks based
on the Buckley-Leverett saturation profile, with denser discretization towards high S p (in this study, we
used 500 banks with S p  0.8 out of 700 banks in total). Next the fractional flow f p and its first order
derivative f p are calculated, and the shock front is identified (an example is discussed later in details in
the paper). Finally the location of bank interfaces is determined by the frontal advance equation
qti f pj
rj  .
 h

n j 1
 qB 
The viscosity at those interfaces is given by  j  F fj 
 2 hr 
, and the calculated viscosity can be
 j 
further used to obtain 1 in Equation 6 in order to solve for the wellbore pressure (details in Appendix
A).

Method B

Another approach is to keep the multicomposite reservoir formula (Abbaszadeh and Kamal 1989), but
calculate the viscosities in each bank beforehand based on some viscosity model, and simply integrate
those calculated viscosities into parameters of  j and  j in the multicomposite reservoir model. Because
the Partial Differential Equation (PDE) formula is not dependent on the power-law anymore, we are free
to choose different viscosity models to calculate the viscosities in each bank. Here, we changed the
viscosity model from power-law to Meter’s model (1964).

In Meter’s viscosity model, the polymer viscosity  p  , C p , r  is modeled as a function of the exponential
coefficient  , polymer concentration (weight%) C p , and radial distance r :
 0p   w
 p  w   1
, (7)
  
1  
  1/2 
where the viscosity at zero shear rate is given by:
 0p  w 1  a1C p  a2C p2  a3C 3p  . (8)
The shear rate is given by the modified Blake-Kozeny capillary bundle equation (Lin 1981; Sorbie 1991):
3.97 E q
 , (9)
kkrp S p 2 hr
where E is the shear rate coefficient used to explain deviation of the porous media from the ideal
capillary-tube-bundle model.
6 SPE-181473-MS

Figure 3 shows a comparison between the power-law model and Meter’s model. For shear rate between
0.3 to 4, the Meter’s model also shows an approximately linear correlation on the log-log plot. However,
the viscosity from Meter’s model stabilizes to constant value as shear rate goes to infinity or zero. This
can be easily understood from Equation 7:  p  w as    ; and  p   p0 as   0 . The bounded
 p values enable the Meter’s model to be applied to a wider range of shear rate compared to the power-
law model, where  p   as   0 .

10
Viscosity (cp)

Meter's
0.1
Power-law

0.01
0.1 1 10 100
Shear rate (sec-1)
Figure 3: Comparison of Meter’s and power-law viscosity models. The Meter’s model has up0 =34 cp.

To apply Method B, the multicomposite reservoir model is discretized based on the same procedure
discussed earlier in Method A. After the location of bank interfaces is determined, the polymer viscosity
at each bank  pj can be obtained from the Meter’s model shown in Equations 7-9. Then  pj is used to
calculate the fluid mobility  j and diffusivity  j , which are needed as model parameters in the
multicomposite reservoir model.

Solutions
Method A

Figure 4 shows the viscosity profiles in the polymer region. For Newtonian fluids ( n  1 ), viscosity is
independent of shear rate and remains constant at different distances. For non-Newtonian fluids
( n  0.2 in this example), viscosity increases (shear rate decreases) as fluids move away from the
wellbore.

krp /  p0
Figure 5 shows the fractional flow curve, which is given by f p  for power-law fluids
krp /  0p  kro / o
with n  1 and n  0.2 . The straight lines in red represent the shock. We can clearly see the injected
polymer leads to a more piston-like displacement, which improves the sweep efficiency.
SPE-181473-MS 7

n=1

Figure 4: Viscosity versus radial distance for power-law fluids with n=1 (left) and n=0.2 (right).

Figure 5: Fractional flow versus saturation for power-law fluids with n=1 (left) and n=0.2 (right).

The pressure behavior during a falloff test generated using Method A is shown in Figure 6. The falloff
pressure was generated using superposition p falloff  p  ti  t   p  t  , where ti is the injection time
prior to shut-in, and t is the shut-in time during a falloff test.

The pressure derivatives on the Bourdet plot clearly show fluid banks with different viscosities. For
instance, the pressure derivative for case n  0.2 (green) can be divided into three parts by line t  3 hr
and t  100 hr. The part at t  3 hr represents the polymer region, where we can observe a continuously
increasing derivative, indicating increasing viscosity away from the wellbore. The part at t  100 hr as
a flat line represents the oil region, where the viscosity remains constant ( o ). The part in 3 hr  t  100
hr is the transition zone from polymer region to oil region. Such transition can be also easily detected from
the Horner plot. On the Horner plot, time increases as we move from right to left. We first observe curves
on the right of the plot representing polymer regions, next is the straight line representing the oil region.
8 SPE-181473-MS

Figure 6: Falloff pressure with different fluids on Bourdet plot (left) and Horner plot (right).

In Figure 6, the transition can be easily identified from both plots in terms of time. From an engineering
point of view, we are also interested in the location of the polymer front in terms of space, in other words,
how far the polymer bank is away from the wellbore. By matching the developed solution to the field data,
we obtain a set of input parameters that represent the true reservoir parameters. Then the front location at
time ti is given by the frontal advance equation (Abbaszadeh and Kamal 1989):
qti f p
r , (10)
 h
where f p is the slope of the shock line on the fractional flow plot (the red line in Figure 5).

Another interesting observation comes from the comparison of cases with different n values. On the
Bourdet plot, the pressure derivatives of three cases overlap with each other in the oil region, indicating
the viscosity is the same in the oil region for all cases. But the derivatives in the vicinity of the wellbore
region are different. This is because different n values lead to different viscosities at high shear rate. Note
that the point where three derivatives intersect corresponds to   1 sec-1.

Method B

The falloff pressure calculated using Method B and Meter’s viscosity model is shown in Figure 7. All the
model parameters are the same as used in Method A (Figure 6), except for polymer viscosity. The results
from both methods (Figure 6 and Figure 7) look similar, because the viscosity models for shear rates that
are typically encountered in the bulk of the polymer bank are similar as shown in Figure 3. Again we
observe the distinction between the polymer and oil regions from both the Bourdet plot and Horner plot,
and the overlapped pressure derivatives in the oil region. One significant difference, however, is the initial
part of the pressure derivatives, corresponding to the vicinity of the wellbore. The derivatives of different
cases converge together. The reason for that is the bounded viscosity at high shear rate in Meter’s model.

One useful application of the pressure derivative curve is to estimate the mobility ratio between different
banks. For instance, the mobility ratio between the fluid bank closest to wellbore and polymer front was
calculated by definition to be about 10 for the case C p  0.2 . Alternatively, one can examine the ratio of
the pressure derivative values at t  3 104 hr and t  3 hr (green curve in Figure 7), which is also
approximately 10.
SPE-181473-MS 9

Figure 7: Falloff pressure of different fluids on Bourdet plot (left) and Horner plot (right).

Comparison of Methods A and B

If we compare both methods based on the pressure results only, the results are quite similar. Especially
when the differences in early transients are hidden by wellbore storage, the results from both methods can
look essentially the same. However, Method B requires less mathematical effort and is more
straightforward to migrate from the multicomposite reservoir model. Besides, Method B has the flexibility
to adapt to different viscosity models. Consequently, Method B is usually preferred over Method A for
practical applications.

Validation of the Analytical Solution


The pressure generated from the analytical solution was compared with results from Chevron’s in-house
reservoir simulator (CHEARS). CHEARS is a 3-D multiphase finite difference simulator where Meter’s
viscosity model is used as default (Chien et al. 1985). A radial model was used in CHEARS and all the
parameters were the same as in the analytical solution. The comparison between falloff pressures
generated from analytical model and CHEARS is shown in Figure 8. A reasonable match between the
analytical results and CHEARS simulation is evident.

Figure 8: Comparison of pressure results generated from analytical model and CHEARS simulation.
10 SPE-181473-MS

Field Example
Myfield E is one of the largest oil fields in Southeast Asia. The reservoir is early Miocene sandstones of
deltaic deposits at an average depth of 2,000 ft. subsea and a maximum vertical oil column of 480 ft.
Average porosity of the reservoir is 26 % and the oil is 36o API. The initial bubble point pressure of the
fluid is 235 psig at a reservoir temperature of 207o F. The field has produced more than 50% of the original
oil in place (OOIP) under primary and secondary recovery operations. Polymer injection is one of the
methods being evaluated to produce the remaining oil and extend the field’s economic life. Data in this
field example are from secondary recovery operation (polymer pushing oil) in a part of the field where
waterflooding was not used.

Method B was chosen to analyze the field data, although either Method A or Method B could have been
used given the comparison of the two methods as shown before. In fact, the difference between the two
methods becomes negligible when early transient behaviors are hidden by wellbore storage. In all the
previous examples, wellbore storage was set to zero to capture the viscosity close to wellbore. However,
zero wellbore storage is not the case for field applications. As the first step to analyze the field data, we
used non-zero wellbore storage as the input for the developed model, which resulted in some interesting
characteristics of the pressure derivatives (shown in Figure 9).

For case C p  0.2 , we first see a straight line with unit slope on the pressure difference and pressure
derivative, indicating wellbore storage effect. Around t  3 hr, the polymer front is represented as a
hump, which followed by the transition zone and the oil region. The polymer effect should not be ignored
on the derivative, because the hump region caused by storage together with polymer is much wider
compared with pure storage effect (blue). Such characteristics were also observed from the field data.

Figure 9: Falloff pressure with C=0.005 STB/psi on Bourdet plot (left) and Horner plot (right).

The field data were obtained from a well in Myfield E. Figure 10 shows the pressure and flow rate history
of the data. A falloff test was implemented after around nine hours of polymer injection. There were other
polymer injections prior to the period shown in Figure 10, unfortunately those testing data were not
available for our analysis. The log-log plot of the original pressure falloff data is shown in Figure 11
(left). To clean up the data, we removed the repeated measurements and used trend line for smoothing.
The data after cleanup are shown on the right of Figure 11. Here, we observe the similar characteristics
as shown in the developed solution with non-zero storage. The first hump indicating wellbore storage,
followed by a small flat zone representing polymer front. The late time pressure was interpreted as the oil
SPE-181473-MS 11

region.

Figure 10: Pressure and flow rate histories of Myfield E, with falloff period marked in yellow.
Pressure [psi]

100

10

1
1E-3 0.01 0.1 1 10
Time [hr]
Figure 11: The original data (left) and clean data (right) from falloff test in Myfield E.

Next we matched the developed model to the field data by tuning input parameters (Figure 12). Because
the polymer front is reached early (about t  1 hr), even a small wellbore storage may hide the polymer
effect on the pressure derivative. That is why it does not show a distinct hump representing polymer front
on the green curve with C  0.03 STB/psi . Figure 12 also shows the results using zero wellbore storage
(blue), to capture the polymer characteristics that not affected by storage. Other parameters are listed in
Table 1. The injection time was estimated to be 1300 hours, which confirmed there were other polymer
injections earlier than the provided data. We used the injection rate immediately before falloff of 2000
STB/d, as a reasonable way to simplify the analysis (Earlougher 1977.) Although the average rate is a
simplified representation of the unknown rates in earlier injections, the analysis method and procedure
would be the same regardless of the used rates, and we do not expect the changes in parameter estimates
to be significant. The C p was estimated to be 0.2, which resulted in  p  15 cp at   7.3 sec-1 that
honored the polymer viscosity measured in the field. Using  p  22 cp at polymer front, the mobility
ratio between oil bank and polymer front was calculated as 28, which is consistent with the pressure
derivative ratio from the log-log plot. With the estimated parameters, the polymer front location was
calculated to be 44 ft. Numerical simulation with the reservoir simulator used in verifying the analytical
solution showed the polymer front to be at 43.35 ft which is in good agreement with the analytical analysis.
12 SPE-181473-MS

Figure 12: Matching Myfield E data (crosses and circles) using C=0 STB/psi (blue) and C=0.03 STB/psi (green).

Table 1: Reservoir and injection data for Myfield E.


k , mD 110
 0.26
h , ft 480

rw , ft 0.3

w , cp 0.2955

o , cp 0.8041

cf , psi -1
3e-6

c p , psi-1 1e-6

co , psi-1 1e-6

s 10
B , STB/RB 1.0
q , STB/d 2000

ti , hours 1300

C p , weight% 0.2

We recognize that the quality of the field data is less than ideal. However, the solution and analysis method
described in this paper, are not less applicable because to-date we do not have extremely good field data.
Actually, it is our hope that publishing this work would encourage operators to collect better data as they
now know an analytical solution exists that can be used with regression methods to allow for practical and
efficient analysis of filed data in reasonable time and provide valuable characterization of the performance
of polymer floods.

Conclusions
1. An analytical solution for pressure transient behavior in two phase flow of a non-Newtonian fluid
and a Newtonian fluid was developed and validated. This is the first analytical solution developed
for this type of problem.
2. Interpretation workflow for falloff tests following polymer injection using the developed solution
was demonstrated with Bourdet derivative and Horner plots.
SPE-181473-MS 13

3. Interpretation of falloff tests using the developed solution can yield information about the mobility
of various (polymer and oil) banks, wellbore conditions and the location of the flood front. This is
a cost effective technology to obtain information needed to manage and optimize polymer flood
operations.
4. The developed solution was used to analyze falloff test data from Myfield E. The pressure transient
characteristics proposed in the solution were indicated from the field data.
5. Using the developed solution, a series of falloff tests may be designed to be run at specific intervals
starting early in the flood operation to characterize the progress of the process, introduce needed
modifications to optimize the flood, and monitor flood front location versus time.

Nomenclature
a1 , a2 , a3 = temperature dependent constants (weight%)-1
B = formation volume factor, RB/STB
c = compressibility, psi-1
c f = formation compressibility, psi-1
co = oil compressibility, psi-1
c p = polymer compressibility, psi-1
C = wellbore storage, STB/psi
CD = dimensionless wellbore storage
C p = polymer concentration, weight%
E = shear rate coefficient
f p = fractional flow, fraction
f p = derivative of fractional flow
Ff = bed factor of the formation, cp-(ft/sec)1-n
F j = diffusivity ratio in multibank system, 1 /  j
h = formation thickness, ft
I v = modified Bessel functions of first kind
k = permeability, md
kro = oil relative permeability
krp = polymer relative permeability
K = Ostwald-de Waele power-law coefficient, cp-secn-1
Kv = modified Bessel functions of the second kind
n = Ostwald-de Waele power-law index
p = pressure, psi
pD = dimensionless pressure
p D = dimensionless pressure in Laplace space
p falloff = falloff pressure, psi
pw = wellbore pressure, psi
pwD = dimensionless wellbore pressure
pwD = dimensionless wellbore pressure in Laplace space
14 SPE-181473-MS

q = injection rate, STB/d


r = radial distance, ft
rD = dimensionless radial distance, r / rw
rw = wellbore radius, ft
s = skin
S = saturation, fraction
t = time, hours
t D = dimensionless time
ti = injection time, hours
z = Laplace transform variable
 = exponent coefficient in Meter’s viscosity model
 = (3  n) / 2
 = shear rate, sec-1
t = falloff time, hours
 = diffusivity, k /  c
 = fluid mobility, k / 
 = viscosity, cp
o = oil viscosity, cp
 p = polymer viscosity, cp
 0p = polymer viscosity at zero shear rate, cp
w = water viscosity, cp
 = (1  n) / (3  n)
 = porosity, fraction

Subscripts
D = dimensionless
j = jth bank
o = oil
p = polymer
w = water

Acknowledgment
The authors thank Chevron for permission to publish this work. Aysegul Dastan and Yan Pan reviewed
the manuscript and made valuable suggestions.

References
Abbaszadeh, M. and Kamal, M. 1989. “Pressure-Transient Testing of Water-Injection Wells,” SPERE
(February 1989) pp 115-24.
Bagassi, M., Chauveteau, G., Lecourtier, J., Englert, J., and Tirrell, M. 1989. “Behavior of Adsorbed
Polymer Layers in Shear and Elongational Flows”, Macromolecules (January 1989): 262-266.
Chauveteau, G. 1981. “Molecular Interpretation of Several Different Properties of Flow of Coiled Polymer
Solutions Through Porous Media in Oil Recovery Conditions,” paper SPE 10060 presented at SPE
ATCE, San Antonio, Texas, USA, 5-7 October.
SPE-181473-MS 15

Chien, M.C.H., Lee, S.T., and Chen, W.H. 1985. “A New Fully Implicit Compositional Simulator,” paper
SPE 13385 presented at the SPE Reservoir Simulation Symposium, Dallas, 10-13 February. doi:
10.2118/13385-MS.
Dupas, A., Henaut, I., Rousseau, D., Poulain, P., Tabary, R., Argillier, J.F., and Aubry, T. 2013. “Impact
of Polymer Mechanical Degradation on Shear and Extensional Viscosities: Toward Better Injectivity
Forecasts in Polymer Flooding Operations”, paper SPE 164083 presented at the SPE International
Symposium on Oilfield Chemistry, Woodlands, Texas, 8-10 April.
Earlougher, R. C. 1977. Advances in Well Test Analysis, SPE Monograph.
Ertekin, T., Cicek, O., Adequmi, M. A., and Daud, M. E. 1987. “Pressure Transient Behavior of Non-
Newtonian/Newtonian Fluid Composite Systems in Porous Media With a Finite Conductivity Vertical
Fracture,” paper SPE 17053 presented at the SPE Eastern Regional Meeting, Pittsburgh, Pennsylvania,
21-23 October.
Ikoku, C. U. and Ramey Jr, H. J. 1978. “Transient Flow of Non-Newtonian Power-Law Fluids in Porous
Media,” SPEJ (June 1979), pp 164-74.
Kamal, M. M. 2009. Transient Well Testing, SPE Monograph Vol. 23, pp 672-87.
Katime-Meindl, I. and Tiab, D. 2001. “Analysis of Pressure Transient Test of Non-Newtonian Fluids in
Infinite Reservoir and in the Presence of a Single Linear Boundary by the Direct Synthesis Technique,”
paper SPE 71587 presented at SPE ATCE in New Orleans, LA, USA 30 September – 3 October.
Li, Z. and Delshad, M. 2014. “Development of an analytical Injectivity Model for Non-Newtonian
Polymer Solutions,” SPEJ (June 2014) pp 381-9.
Lin, E. 1981. “A Study of Micellar/Polymer Flooding Using a Compositional Simulator," Ph.D.
dissertation, The University of Texas at Austin.
Lund, O. and Ikoku, C. U. 1981. “Pressure Transient Behavior of Non-Newtonian/Newtonian Fluid
Composite Reservoirs,” SPEJ (April 1981): 21-02.
Mahani, H., Sorop, T. G., van den Hoek, P. J., Brooks, A. D., and Zwaan, M. 2011. “Injection Fall-off
Analysis of Polymer Flooding EOR,” paper SPE 145125 presented at SPE Reservoir Characterization
and Simulation Conference and Exhibition, Abu Dhabi, UAE, 9-11 October.
Meter, D. and Bird, R. 1964, “Tube Flow of Non-Newtonian Polymer Solutions”, AIChE Journal
(November 1964): 878-81.
Odeh, A. and Yang, H. 1979. “Flow of Non-Newtonian Power-Law Fluids Through Porous Media,” SPEJ
(June 1979): 155-63.
Ostwald. 1929. “de Waele-Ostwald Equation,” Kolloid Zeitschrift (1929) 47(2) 176-87.
Seright, R.S. 1983. “The Effects of Mechanical Degradation and Viscoelastic Behavior on Injectivity of
Polyacrylamide Solutions”, SPEJ (June 1983): 475-485.
Sorbie, K. S. and Roberts, L. J. 1984. “A Model for Calculating Polymer Injectivity Including the Effects
of Shear Degradation,” paper SPE 12654 presented at the SPE/DOE Fourth Symposium on Enhanced
Oil Recovery, Tulsa, OK, USA, 15-18 April.
Sorbie, K. S. 1991. Polymer-Improved Oil Recovery, CRC Press, Inc., Boca Raton, Florida.
van den Hoek, P. J., Mahani, H., Sorop, T. G., Brooks, A. D., Zwaan, M., Sen, S., Shuaili, K., and Saadi,
F. 2012. “Application of Injection Fall-Off Analysis in Polymer Flooding,” paper SPE 154376
presented at the 74th EAGE Conference & Exhibition incorporating SPE EUROPEC 2012,
Copenhagen, Denmark, 4-7 June.
Van Poollen, H. K. and Jargon, J.R. 1969. “Steady-state and Unsteady-state Flow of Non-Newtonian
Fluids Through Porous Media,” SPEJ (March 1969) pp 80-8.
Vongvuthipornchai, S. and Raghavan, R. 1987. “Well Test Analysis of Data Dominated by Storage and
Skin: Non-Newtonian Power-Law Fluids,” SPEFE (December 1987) pp 618-28.
Yeh, N. S. and Agarwal, R. G. 1990. “Pressure Transient Analysis of Injection Wells in Reservoirs with
Multiple Fluid Banks,” paper SPE 19775 presented at SPE ATCE, San Antonio, Texas, USA.
16 SPE-181473-MS

Appendix A: Mathematical Details of Method A

Discretize the Power-Law Model


The original equation for Power-law non-Newtonian fluid in single-bank system is given by Ikoku and
Ramey 1978, Vongvuthipornchai and Raghavan 1987:
n 1
 2 p n p p  qB   c
  r1 n Ff  
r r r t  2 h  k
2

n j 1
 2 p j n j p j 1 n j p j  qB    c 
For each bank j :  r Ffj    
r 2 r r t  2 h   k  j
p jD  c / k  j Ffj  qB /  2 hrw  
n 1
 2 p jD n j p jD
j
1 n j
Non-dimensionalization:  r
rD2 rD rD t D  c / k 1 Ff 1  qB /  2 hrw   n1 1
D
 
 c / k  j Ffj  qB /  2 hrw  j
n 1
1     p jD  p jD j 1 n j
Same as:  rD   F j , where *  rD , F j 
rD rD   j rD  t D   c / k 1 Ff 1  qB /  2 hrw  n1 1

The dimensionless variables defined in the above equations are given by:
n1 1 n1 n1 1
2 kh  qB  K 3  qB  kt
150k 
(1 n1 )/2
pD  p ,  *  Ff 1    9     , tD 
qB *
 2 hrw  12  n1   2 hrw   c *rw2

Solve for Laplace Pressure

Propose the formula of Laplace pressure by analogy (Abbaszadeh and Kamal 1989; Ikoku and Ramey
1978; Vongvuthipornchai and Raghavan 1987):
 1
  
 

p jD  rD j  Aj K vj rD j zF j /  j  B j I vj rD j zF j /  j 
 
Inner boundary and initial conditions after Laplace transform:
 p  1
zCD pwD   rD 1D  
 rD r 1 z D

 p 
pwD  p1D  s  rD 1D 
 rD rD 1

Follow the formula of p1D and define B j   Aj j , j  1, 2,..., N :


  p1D   z  z  z  z
    A1 zK1v1    B1 zI1v1     A1 zK1v1    A11 zI1v1  
 rD rD 1  1   1   1   1 
 z  z   z  z 
pwD  A1K v1    A11I v1    s  A1 zK1v1    A11 zI1v1   
 1   1    1   1  
SPE-181473-MS 17

p   p  1
Plug in pwD and  1D  into equation zCD pwD   rD 1D   to obtain A1 :
 rD rD 1  rD rD 1 z
1 1
A1  
z  z   z  z    z   z  z   
zK1v1    zCD  K v1    s zK1v1     1  zI1v1    zCD  I v1    s zI1v1   
 1    1   1     1    1   1   

Finally, the wellbore pressure is:


 z  z   z  z 
K v1    1 I v1    s z  K1v1    1 I1v1  
1  1   1    1   1  
pwD  
z  z   z  z    z   z  z   
zK1v1    zCD  K v1    s zK1v1     1  zI1v1    zCD  I v1    s zI1v1   
 1    1   1     1    1   1   

Where 1 is given by:


 j 1 j 1Kv  Z j , j    j 1K1v  Z j , j 
j 
 j 1 j 1 I v  Z j , j    j 1 I1v  Z j , j 

Z j , j  rjDj zF j /  j

Z j , j 1  rjDj1 zF j 1 /  j 1
 j 1  K1vj 1  Z j , j 1    j 1 I1vj 1  Z j , j 1 
 j 1   M j / M j 1  F j 1 / F j
 j 1  K vj 1  Z j , j 1    j 1 I vj 1  Z j , j 1 

The reservoir boundary effect is contained in  N :


 N  0 for infinite acting radial flow

N 

K vN reD N zF N /  N  for a constant pressure boundary
I vN r N
eD zF /  
N N

N  
K1vN  r zF /   for a no-flow boundary
N
eD N N

I1vN  r zF /  
N
eD N N

Proof: how the definitions above satisfy the pressure and rate conditions
between banks

From pressure conditions: p jD  p j 1D , rD  rfjD  t D 



Aj Kv rjDzF j /   B j I v rjD

 
zF j /   Aj 1Kv rjD

  
zF j 1 /   B j 1I v rjD


zF j 1 /  
Define:

Z j , j  rjDj zF j /  j
18 SPE-181473-MS


Z j , j 1  rjDj1 zF j 1 /  j 1
B j   Aj j , j  1, 2,..., N
 j 1  K vj 1  Z j , j 1    j 1 I vj 1  Z j , j 1 
Yield:
Aj K vj  Z j , j   Aj j I vj  Z j , j   Aj 1K vj 1  Z j , j 1   Aj 1 j 1I vj 1  Z j , j 1   Aj 1 j 1
K vj  Z j , j    j I vj  Z j , j 
Aj 1  Aj
 j 1

 p jD  p j 1D
From flow rate conditions:  j   j 1 , rD  rfjD  t D 
rD rD
       
 Aj zF j K1vj Z j , j  B j zF j I1vj Z j , j  M j / M j 1  Aj 1 zF j 1 K1vj 1 Z j , j 1  B j 1 zF j 1 I1vj 1 Z j , j 1 
   
Define:
 j 1   M j / M j 1  F j 1 / F j
 j 1  K1vj 1  Z j , j 1    j 1 I1vj 1  Z j , j 1 
Yield:
K vj  Z j , j    j I vj  Z j , j 
Aj  K1vj  Z j , j    j I1vj  Z j , j     j 1 Aj 1  K1vj 1  Z j , j 1    j 1 I1vj 1  Z j , j 1     j 1 Aj 1 j 1   j 1 j 1 Aj
 j 1
K vj  Z j , j    j I vj  Z j , j 
K1vj  Z j , j    j I1vj  Z j , j    j 1 j 1
 j 1
 j 1 j 1 Kvj  Z j , j    j 1 K1vj  Z j , j 
Solve for:  j 
 j 1 j 1 I vj  Z j , j    j 1 I1vj  Z j , j 


Outer boundary conditions on  N : pND  reD N 1  AN K vN reD N zF N /  N  BN I vN reD N zF N /  N 
    
B
For infinite acting radial flow, p ND must remain bounded as reD   . Thus BN  0 ,  N   N  0
AN


For a constant pressure boundary, pND  reD N 1  AN K vN reD N zF N /  N  BN I vN reD N zF N /  N   0 .
    
Thus  N  
N
BN K vN reD zF N /  N

 
AN I vN reD N zF N /  N  
For a no-flow boundary,
  pND 
   
  AN zF N K1vN reD N zF N /  N  BN zF N I1vN reD N zF N /  N  0 .  
 rD rD reD

Thus  N  
BN

K1vN reD N zF N /  N  
AN I1vN reD N zF N /  N  
SPE-181473-MS 19

Appendix B: Shear Rate Dependent Viscosity Models

Power-Law Viscosity Model  p  n, K , r 


n 1 n
 qB  K 3
, where Ff   9   150kkrp S p 
1 n  /2
 p  Ff  
 2 hr  12  n
n 1
 qB 
  Ff 
*
 at wellbore
 2 hrw 

Meter’s Viscosity Model  p  , C p , r 


 0p   w
 p  w   1
, where
  
1  
  1/2 
 0p  w 1  a1C p  a2C p2  a3C 3p  ,  
3.97 E q
kkrp S p 2 hr

View publication stats

You might also like