Development of A High-Temperature-Resistant Polymer-Gel System For Conformance Control in Jidong Oil Field

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

REE186235 DOI: 10.

2118/186235-PA Date: 28-June-18 Stage: Page: 1 Total Pages: 10

Development of a High-Temperature-
Resistant Polymer-Gel System for
Conformance Control in Jidong Oil Field
Daoyi Zhu, Jirui Hou, Qi Wei, and Yuguang Chen, China University of Petroleum, Beijing

Summary
The PG Reservoir in Jidong Oil Field is at a depth of approximately 4500 m with an extremely high temperature of approximately
150 C. The average water cut has reached nearly 80%, but the oil recovery is less than 10% after only 2 years of waterflooding process.
It is of great importance to develop a high-temperature-resistant plugging system to improve the reservoir conformance and control
water production. An in-situ polymer-gel system formed by the terpolymer and a new crosslinker system was developed, and its proper-
ties were systematically studied under the condition of extremely high temperature (150 C). Suitable gelation time and favorable gel
strength were obtained by adjusting the concentration of the terpolymer (0.4 to 1.0%) and the crosslinker system (0.4 to 0.7%). An
increase of polymer and crosslinker concentration would decrease the gelation time and increase the gel strength. The gelant could
form continuous 3D network structures and thus have an excellent long-term thermal stability. The syneresis of this gel system was
minor, even after being heated for 5 months at the temperature of 150 C. The gel system could maintain most of the initial viscosity
and viscoelasticity, even after experiencing the mechanical shear or the porous-media shear. Core-flow experiments showed that the gel
system could have great potential to improve the conformance in Jidong Oil Field.

Introduction
Excessive water production is one of the most-serious problems in oil-producing wells all over the world. It not only causes corrosion,
underground and surface water pollution, expensive separation cost, and nonproductive pumping, but it also leads to the decrease of
sweep efficiency and oil recovery (Seright et al. 2001; Bai and Zhang 2011; Goudarzi et al. 2015). Therefore, it is imperative to solve
the problem of excessive water production. Polymer-gel technology has proved to be one of the most-effective methods in the field of
conformance control and enhanced oil recovery (EOR) (Sydansk and Argabright 1987; Portwood 2005). In most cases, a polymer-
gelation system—mainly composed of a proprietary, aqueous polymer solution and a crosslinker system—is injected into a reservoir. A
gel with continuous 3D network structures, called bulk gel, will be formed in the reservoir. Bulk gels can have a broad range of
controllable gelation time and strength by adjusting the concentration of polymer and crosslinker system so they can be applied to
control water production (Bryant et al. 1997; Al-Muntasheri et al. 2008; El-Karsani et al. 2014).
Most of the successful applications of in-situ polymer-gel systems are composed of partially hydrolyzed polyacrylamide (HPAM)
and inorganic crosslinkers, such as aluminum (Al3þ), chromium (Cr3þ), and zirconium (Zr4þ) (Zhu et al. 2017). However, they are
usually applied at low-temperature (<80 C) reservoirs because the intermolecular force between polymer and crosslinkers is a weak
coordinate bond between the carboxylate ions (–COO–) of the polymer and multivalent-ion complexes (Brattekås et al. 2017). In recent
years, more and more water-control problems have required attention for either high-temperature or extremely high-temperature reser-
voirs. Many polyacrylamides and biopolymers have been reported for use in harsh EOR conditions. New functional groups have been
tried to improve the long-term thermal and salt stability. These functional groups on the polymer chain mainly include acrylamide,
acrylic acid, sodium acrylamide-tertiary-butyl sulfonate, and N-vinyl pyrrolidone (Gaillard et al. 2017). These polymers are named
copolymers or terpolymers, which refer to the polymer production using either two or three monomers. Biopolymers such as xanthan,
scleroglucan, and schizophyllan have also been applied in high temperatures and salinity. They can exhibit excellent temperature resist-
ance up to 90 to 135 C and exhibit high tolerance to salinity (Rivenq et al. 1992; Leonhardt et al. 2014). Organic crosslinker systems,
such as phenol/aldehyde and polyethyleneimine (PEI) (El-Karsani et al. 2015; Zhao et al. 2015), are used at high-temperature (>80 C)
reservoirs. The organic crosslinker system composed of aldehydes can crosslink with polymers through dehydration/condensation reac-
tion between the amide groups (–CONH2) of the polymer and the hydroxyl groups (–CH2OH) of the crosslinker system, which is a
stronger intermolecular reaction. Also, the crosslinking reactions between HPAM and PEI are believed to occur between the imine
nitrogen on PEI and the amide groups on the HPAM (transamidation) (Al-Muntasheri et al. 2009). Although organic gel systems have
been widely used for conformance control, few studies have been performed for extremely high-temperature (>138 C) reservoirs. In
addition, most of the studies focused only on the properties of the bulk gel, whereas the properties of either the gelant or the bulk gel in
the porous media, such as the mechanical shear effect and the porous-media shear effect, were not well-studied. Seright (2003) and
Brattekås et al. (2017) have conducted extensive research on gel-extrusion performance through fractures. They found that with the
decrease of fracture width, pressure gradients required to extrude gels increased.
In this study, an in-situ polymer-gel system formed by terpolymer and the crosslinker system OC-3 was developed to solve conform-
ance problems in the PG Reservoir in Jidong Oil Field, east China. The temperature of the PG Reservoir ranges from 140 to 160 C and
the salinity is 5050 mg/L. A terpolymer with the low molecular weight of 3 to 5 million daltons was used. A high-temperature-
resistance crosslinker system was selected because it is less toxic and has longer gelation time and stronger gel strength than the
phenol/formaldehyde crosslinker system. Factors that affect the gelation time and gel strength were also studied, including the
concentration of the polymer and the crosslinker and the shear effect. An environmental scanning electron microscope (ESEM) was
used to observe the microstructures of the gel system, and both fractured- and homogeneous-core-flow experiments were conducted to
study the plugging capacity and conformance-control performance of the gel system under extremely high temperature (150 C).

Copyright V
C 2018 Society of Petroleum Engineers

This paper (SPE 186235) was accepted for presentation at the SPE/IATMI Asia Pacific Oil and Gas Conference and Exhibition, Jakarta, 17–19 October 2017, and revised for publication.
Original manuscript received for review 10 September 2017. Revised manuscript received for review 13 February 2018. Paper peer approved 19 February 2018.

2018 SPE Reservoir Evaluation & Engineering 1

ID: jaganm Time: 14:41 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028


REE186235 DOI: 10.2118/186235-PA Date: 28-June-18 Stage: Page: 2 Total Pages: 10

Experimental Studies
Materials. Core. Natural sandstone core samples were provided by Jidong Oil Field, China. They were toluene-extracted to remove
oil before the fabrication of the fractured model. The lithology is sandstone, the rock is water-wet, and the permeability to gas was
approximately 300 md.
Brine. The produced water was provided by Jidong Oil Field, China; its salinity was 5050 mg/L, and pH value was 6.4. The ion
composition of the brine is summarized in Table 1. This brine was used in all experiments.

Cations Anions
+ + 2+ 2+ – 2– – 2–
Ion Composition Na and K Ca Mg Cl SO4 HCO3 CO3 Salinity
Concentration (mg/L) 1498 11 5 408 27 2914 189 5050

Table 1—Ion composition of the produced water.

Polymer. The terpolymer ZP-4 used in our research was synthesized by acrylamide, sodium acrylamide-tertiary-butyl sulfonate,
and N-vinyl pyrrolidone. The molecular weight was 3 to 5 million daltons and the hydrolysis degree was 15 to 20%.
Crosslinker. The organic crosslinker system (OC-3) consisting of hexamethylenetetramine (HMTA) and resorcinol was made in
our laboratory. It can be applied in curing temperatures higher than 150 C. In addition, it is less toxic and has longer gelation time than
the phenol/formaldehyde crosslinker system (Bryant et al. 1997; Moradi-Araghi 2000). The crosslinking mechanism between the poly-
mer and crosslinker is as follows.
First, resorcinol reacts with the formaldehyde (generated by HMTA), forming 2,4,6-tris(hydroxymethyl)benzene-1,3-diol, before
being further carried out to the polycondensation reaction to form crosslinked clusters, which are shown in Step 1 of Scheme 1. Second,
the amide groups (–CONH2) in the polymer undergo a methylation reaction with highly active methylene glycol (hydrolysis of formalde-
hyde); however, only part of the amide groups undergoes this process, as shown in Step 2. Finally, the products obtained by the preceding
two steps will undergo dehydration/condensation reactions to generate a hydrogel with a 3D network structure, as shown in Step 3.
Additives. Thiourea (used as an antioxidant) and sodium chloride were analytically pure.

Step 1

OH OH
HOH2C CH2OH HOH2C OH HOH2C OH HOH2C OH
O
H2 H2
C C CH2OH
C n
OH OH
H H HOH2C OH HOH2C OH HOH2C OH
CH2OH

Step 2
O
H H
CH2 C CH2 C
n C n
C H H C
H2N O HOH2CHN O

Step 3
H H
CH2 C CH2 C
n

O n n
C C C
HC O O
NHCH2 H2CHN H2CHN
CH2

CH2O OH OH2C OH OH2C OH


H2 H2
C C CH2OH

OH2C OH OH2C OH OH2C OH


n

NHCH2 H2CHN H2CHN


O O
C C C
HC
O
CH2

CH2 C CH2 C
H n H n

Scheme 1—Crosslinking mechanism of the studied polymer-gel system.

Gelant Preparation. Polymer gelants formed by different concentrations of the terpolymer ZP-4 and the crosslinker system OC-3,
0.05% antioxidant, and 0.05% heat stabilizer were prepared at room temperature as follows: 10 g of terpolymer solid particles was
added to the 1000 g of brine and stirred at the rate of 400 6 20 rev/min for 4 hours, and then aged for 24 hours. The terpolymer solution

2 2018 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 14:41 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028


REE186235 DOI: 10.2118/186235-PA Date: 28-June-18 Stage: Page: 3 Total Pages: 10

(1.0%) was then diluted with brine to the required concentrations. Then, the preweighed crosslinker system OC-3, antioxidant, and heat
stabilizer were added to the solutions and then stirred for 20 minutes to produce a homogeneous gelant. Finally, approximately 20 cm3
of gelant was slowly injected into a thick-walled ampule with a thickness of 2.5 mm and a pressure limit of 1500 kPa. The ampule neck
was flame-sealed after the ampule was vacuumed for 2 to 3 hours for successful seals. Then, the crosslinking reaction was started when
the ampule was heated in the oven at 150 C.

Measurement and Analysis. Determination of Gelation Time and Gel Strength. The gelation time and gel strength were determined
by the gel-strength-code method (Sydansk and Argabright 1987). By this method, the gel strengths of different flowing and suspension
states were divided into nine categories (Codes A through I), depicted in Fig. 1. In this study, the ampules were inverted and observed
every half-hour until the gel strength remained constant, and that time was marked as the gelation time. Note that the gelant in Fig. 1a
nearly has the same viscosity as the original gelant, whereas the gel in Fig. 1b is only slightly more viscous than the initial gelant, so
the lighter and darker colors were adopted to indicate these differences.

A B C D E F G H I

Fig. 1—Schematic of different gel-strength codes, adapted from Sydansk and Argabright (1987). Codes (A) through (I) are as fol-
lows: Code A: no detectable continuous gel formed; Code B: highly flowing gel; Code C: flowing gel; Code D: moderately flowing
gel; Code E: barely flowing gel; Code F: highly deformable nonflowing gel; Code G: moderately deformable nonflowing gel; Code H:
slightly deformable nonflowing gel; and Code I: rigid gel.

Viscosity of the Gelant Sheared at 150º C. The viscosity of the gelant was measured with a commercial modular rheometer, which
can be used to study the shear effect of the gelant under the high shear rate (100 1/seconds) and to test whether it can still form bulk
gel at the shear rate of 7.1 1/seconds at 150 C. The mechanical shear time was 1 hour. The viscosity was measured in a controlled-
rate mode.
Rheological Property of the Bulk Gel. The gel performance (i.e., storage modulus G0 and loss modulus G00 ) of the preformed bulk
gel before and after being sheared in an artificial core (300 md, 5.0 cm long, and 2.5 cm in diameter) was measured with a commercial
modular rheometer with a 20-mm plate/plate system at 60 C. Because the system was open and the water in the bulk gel system would
be lost by evaporation at high temperatures, 60 C was chosen as the experimental temperature to study the gel performance in our
study. The rheological properties were obtained in a controlled-rate mode. The frequency was 1 Hz, the gap was 1.0 mm, and the shear
stress was 10 Pa; parameters were selected to remain within the linear range.
Thermal Stability of Bulk Gel. The thermal stability of the bulk gel system was studied by the gel strength (Codes A through I) and
the syneresis of the bulk gel when the ampules were heated at 150 C for 5 months. Syneresis (S) can be defined as
0

S¼ WW  100%; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ
W
where W0 is the weight of aged gel without free water and W is the weight of the original gelant. If the syneresis ranges from 1 to 5%,
the syneresis of the bulk gel can be defined as very low syneresis, and if the syneresis ranges from 6 to 30%, the syneresis of the bulk
gel can be defined as low syneresis (Eriksen et al. 1997).
Microstructure of Bulk Gel. An ESEM was used to observe the microstructure of the bulk gel system. In this experiment, the bulk
gel samples were freeze-dried by liquid nitrogen, which can preserve the structure of the bulk gel system. The gel samples were then
placed on a copper-cylinder surface attached with conductive adhesive and sprayed with gold to make it more conductive.
Plugging Capacity of the Gel System in Fractured Cores. The method of making fractures in a core is shown in Fig. 2. The origi-
nal core (300 md) was split in half; three copper wires are placed evenly on one of the fractured cores; and the cylinder of the core was
sealed by epoxy resin. The opening of the fracture is controlled by the diameter (0.20 and 0.40 mm) of the copper wires. These fractured
cores were then placed into the core holders for the following experiment.
Core-flow experiments were used to measure the plugging capacity of gel in fractured cores; a schematic of the core-flow setup is
shown in Fig. 3. In this experiment, the confinement pressure was maintained at 5 MPa. The water-injection process was first performed
until the pressure drop across the core reached a stable value, and it was marked as Dp0 . Then 0.6 fracture volume (FV) of the gelant,
suggested by the oilfield company, was injected into the core after 0.15 pore volumes (PV) of the terpolymer solution with the same
polymer concentration of the gelant was preflooded. Note that the injection volume can be adjusted according to the plugging require-
ment when being applied in specific oilfield applications. The core holder was then heated in a 150 C oven for 1 week. Water injection
was performed again until the pressure drop reached another stable value, marked as Dp1 . Plugging efficiency (E) was usually used to
measure the plugging capacity of gel systems, which is defined at constant-injection rate as
Kw0  Kw1 Dp1  Dp0
E¼ ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ
Kw0 Dp1
where Kw0 , Kw1 , Dp0 , and Dp1 are the absolute permeability to water and the pressure drop before and after the gel
treatment, respectively.

2018 SPE Reservoir Evaluation & Engineering 3

ID: jaganm Time: 14:41 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028


REE186235 DOI: 10.2118/186235-PA Date: 28-June-18 Stage: Page: 4 Total Pages: 10

(a) (d)

(b) (c)

Fig. 2—Schematic of making fractures in cores. The red lines represent copper wire of diameter 0.2 or 0.4 mm.

150°C

Polymer solution

Gelant
Brine

Pressure meter Core holder

Collector
Syringe pump Valve

Fig. 3—Schematic of the core-flow setup.

Plugging Performance in the Homogenous Core. A homogeneous-core-flow experiment was conducted to measure the plugging
efficiency of the terpolymer-gel system in the porous media without fractures. Fig. 3 shows the experimental setup. A homogeneous
300-md sandstone core was used. Its porosity was 23.72%. The backpressure was atmospheric pressure, and confinement pressure was
maintained at 5 MPa. Water injection was first performed until injection pressure reached a stable value, and 1.5 PV of the terpolymer
gelant was then injected. The core holders were heated in a 150 C oven for 3 days. Then, chase water was injected when the injection
pressure reached another stable value. Plugging efficiency was calculated using Eq. 2.

Results and Discussion


Factors of Gelation Time and Gel Strength. Polymer Concentration. The terpolymer concentration is very significant to the gela-
tion performance of the gel system. It can affect the gelation time, the gel strength, and the long-term thermal stability of the gel system.
If the gelation time is too short, the gel system may not be transported to the target position, and if the gel strength and the long-term
thermal stability are not good enough, it will affect the plugging capacity of the gel system (Zhao et al. 2015). In this experiment, the
terpolymer ZP-4 concentration ranged from 0.2 to 1.0%, whereas the concentrations of the crosslinker system OC-3, the antioxidant,
and the heat stabilizer were kept constant at 0.6, 0.05, and 0.05%, respectively. The gelation reaction was started at 150 C.
Table 2 and Fig. 4 show the effect of the terpolymer concentration on the gelation time and the gel strength. The gelation time can
be controlled (ranging from 2 to 12 hours) and the gel strength was varied from Code D to Code H. With the increase of the terpolymer
concentration, the gelation time became shorter and the gel strength became stronger than those composed of lower terpolymer concen-
trations. As the terpolymer concentration increases, more crosslinking sites will be created in the polymer side chains, so that the cross-
linker systems can crosslink more with the amide groups (–CONH2) of the terpolymer. In other words, a suitable gelation time and a
favorable gelation performance can be obtained by adjusting the terpolymer concentrations.

Gelation Gelation
Number ZP-4 (%) Time (hours) Strength Code
G1 0.2 12 D–E
G2 0.4 12 E–F
G3 0.6 7 F–G
G4 0.8 4.5 G
G5 1.0 2 G–H

Table 2—Effect of the terpolymer concentration on gelation perfor-


mance (0.6% OC-3).

4 2018 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 14:41 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028


REE186235 DOI: 10.2118/186235-PA Date: 28-June-18 Stage: Page: 5 Total Pages: 10

G1 G2 G3 G4 G5

Fig. 4—Gel systems composed by different terpolymer concentrations at the gelation time. The gelant compositions of G1 through
G5 are shown in Table 2.

Note that when the terpolymer concentration was 0.2%, the gelation time was 12 hours; however, the gelation strength was Code
D–E, which is not strong enough to be applied in the conformance-control operations. When the terpolymer concentration was 1.0%,
the gelation time was 2 hours, which is too short to transport the gelant to the target depth.
Crosslinker Concentration. Organic crosslinker systems can link one polymer chain to another by covalent bonds. In other words,
the amide groups (–CONH2) of the terpolymer could form a 3D network structure by crosslinking with the hydroxyl groups (–CH2OH)
at the phenolic ring of the organic crosslinker system. Hence, the crosslinker system is also very critical to the gelation performance. In
this experiment, the concentration of organic crosslinker system OC-3 ranged from 0.4 to 0.7%, whereas the concentrations of the ter-
polymer, the antioxidant, and the heat stabilizer were kept constant at 0.6, 0.05, and 0.05%, respectively. The gelation reaction was
started at 150 C.
Table 3 shows the effect of crosslinker-system concentration on the gelation time and the gel strength. The gelation time can be con-
trolled (ranging from 4 to 12 hours) and the gel strength ranged from Code D through Code H. The gelation time decreased with the
increase of the crosslinker-system concentration; however, the gelation strength was increased from Code D through Code H. With the
increase of the crosslinker-system concentration, the amount of the hydroxyl groups (–CH2OH) in the gelant are increasing, which can
accelerate the reaction rate of forming 3D network structures. Therefore, the gelation time shortened and the gel strength increased with
the increase of the crosslinker-system concentration. However, suitable gelation time and favorable gelation performance can be
obtained by adjusting the concentrations of the crosslinker system, which can broaden its oilfield application in conformance control.

Gelation Gelation
Number OC-3 (%) Time (hours) Strength Code
G6 0.4 12 D
G7 0.5 10 E–F
G3 0.6 7 F–G
G8 0.7 4 H

Table 3—Effect of crosslinker-system concentration on gelation


performance (0.6% ZP-4).

Long-Term Thermal Stability. Long-term thermal stability of the bulk gel system was studied by the strength and the syneresis of the
bulk gel after the ampules were heated in a 150 C oven for 5 months, as shown in Fig. 5. Gel G1 has the highest syneresis, and the syn-
eresis was approximately 50%, which indicates that the gel system was not very stable at an extremely high temperature for a long
time. This is mainly because the terpolymer concentration of the gelant (G1) is only 0.2%, so the amount of polymer chain is insuffi-
cient to crosslink with the hydroxyl groups (–CH2OH) of the crosslinker system, and thus the gel strength (as shown in Fig. 4) was too
weak to keep stable for 5 months. Gel G2 had the second-highest syneresis, and the syneresis was approximately 5%, which shows that
the gel system was stable at an extremely high temperature, but the gel strength decreased too much. Gels G3 and G4 were stable and
had very little syneresis even after being heated for 5 months at 150 C. Gel G5 was also kept stable at the high temperature. However,
the gelation time was very short, and part of the free water did not have enough time to enter the gel grids, so there was a little amount
of free water at the bottom of the ampule after a long time of heating. The formed gel systems moved down as a whole in the middle of
the ampules when being inverted, mainly because of some gas released during the decomposition of the crosslinker system. From
above, the bulk gel system shows a long-term thermal stability.

Shear Effect. When a gelant is transported through the injection facilities (mechanical shear) or flows in the formation (porous-media
shear) (Seright 1983), the polymer chains will be subjected to stress by friction forces, and thus polymer molecules break and shear degra-
dation occurs. This might result in significant viscosity loss of the gelant and the uncertainty of the gelation process if the gelant is sensi-
tive to shear effect (Caulfield et al. 2002), so it is imperative to study the effect of shear on the gelant. Two separate experiments were
conducted to investigate the effect of shear. One was to measure the gelation performance of the gelant under different shear rates (100
1/seconds for the first hour) using the commercial modular rheometer. The curing temperature ranged from 20 to 150 C with an increasing

2018 SPE Reservoir Evaluation & Engineering 5

ID: jaganm Time: 14:42 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028


REE186235 DOI: 10.2118/186235-PA Date: 28-June-18 Stage: Page: 6 Total Pages: 10

rate of 0.036 C/min to simulate the heating process in the wellbore. Another was to measure the rheological property of the bulk gel sys-
tems before and after the gelants (shear rate of 7.1 1/seconds) were sheared by a core (5-mm length) at the injection rate of 0.3 cm3/min.

G1 G2 G3 G4 G5

Fig. 5—Gel systems composed of different terpolymer concentrations when heated at 1508C for 5 months. The gelant contents of
G1 through G5 are shown in Table 2.

Fig. 6 shows the mechanical shear effect in a high-temperature closed measurement system at the shear rate of 100 1/seconds for the
first hour with the curing temperature ranging from 20 to 150 C, which is used to simulate the shear effect in wellbores. Then, the shear
rate decreased to 7.1 1/seconds to simulate the dynamic gelation process in porous media. The gelant samples were all composed of
0.8% ZP-4, 0.6% OC-3, 0.05% antioxidant, and 0.05% heat stabilizer. Results showed that the viscosity of the gelant decreased from
130 to 21 mPas with the curing temperature increasing from 20 to 150 C, which shows that after mechanical shear the gelant could
also maintain a certain amount of viscosity. Then, the gelant was aged at the temperature of 150 C and sheared at the shear rate of
7.1 1/seconds. After being cured for another 2.3 hours, the gelant viscosity gradually increased because of the beginning of the cross-
linking reaction between the terpolymer and the crosslinker OC-3. At the curing time of 4 hours, the viscosity increased sharply because
of the preliminary crosslinking reaction. In this period, small, weak crosslinking particles formed in the solution (Chauveteau et al.
2000), and therefore the viscosity decreased soon afterward, as shown in Fig. 6. With the curing process continued, the crosslinking
degree increased and gel particles started to form bulk gels, which need a certain amount of time (Chauveteau et al. 2000). Because of
this, there was a large fluctuation at the time ranging from 5.2 to 5.8 hours. After that, bulk gels could be formed and the viscosity
nearly reached stability at approximately 3000 to 3500 mPas. According to this experiment, the gel system showed an excellent
mechanical shear capacity. Note that the viscosity of the gelant was approximately 21 mPas at the curing temperature of 150 C, and
therefore the gel system has a good injection ability to make it transport to the target zones.

Mechanical Porous media


shear shear
160 120
6000
140
100
5000
120 Curing temperature
Curing Temperature (ºC)

Shear Rate (1/seconds)


Viscosity (mPa·s)

Viscosity 80
100 Shear rate 4000

80 60
3000

60
2000 40
40

1000 20
20

0 0 0
0 1 2 3 4 5 6 7
Curing Time (hours)

Fig. 6—Effect of shearing on the gelation performance. Gelant was sheared at a shear rate of 100 1/seconds for the first hour to
simulate the mechanical shear at or near the wellbore. Then, the gelant continued to be cured at the shear condition of 7.1 1/sec-
onds and 1508C.

Porous-media shear of the bulk gel was conducted in an artificial homogeneous sandstone core by injecting the bulk gel into the
matrix at the injection rate of 0.3 cm3/min. The gelant samples were all composed of 0.6% ZP-4, 0.6% OC-3, 0.05% antioxidant, and

6 2018 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 14:42 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028


REE186235 DOI: 10.2118/186235-PA Date: 28-June-18 Stage: Page: 7 Total Pages: 10

0.05% heat stabilizer. Gelant with a volume of 30 cm3 sheared by the porous media was collected in ampule bottles. The sheared bulk
gel was then heated at 150 C until it formed a stable gel. Fig. 7 shows the gel performance of the bulk gel before and after being
sheared in the porous media. After being sheared in the core, the storage modulus G0 and loss modulus G00 of the bulk gel were both
decreased wherein the decrease in G0 was greater than G00 , which shows that the porous-media shear has a greater effect on G0 . How-
ever, the storage modulus G0 >10 Pa and the loss modulus G00 > 2 Pa after the bulk gel sample were sheared in the porous media, which
indicates that the gelant sample also has a strong porous-media-shear stability. Note that in this test the injection pressure could not
maintain stability and it was higher than 1 MPa, which indicates that the preformed bulk gels cannot easily be squeezed into such low-
permeability porous media.

16

14

12
G′ or G″ (Pa)

10

G′ (before)
8
G″ (before)
G′ (after)
6 G″ (after)

0 2 4 6 8 10
Shear Rate (Pa)

Fig. 7—Gel performance of the bulk gel before and after the gelant was sheared in the core. The stress-scanning tests were con-
ducted at 608C. The frequency was 1 Hz and the gap was 1.0 mm.

Microstructure of Bulk Gel. Fig. 8a shows that the bulk gel (G3) in the ampule was a moderately deformable nonflowing gel and
Fig. 8b demonstrates that the bulk gel (G3) was also a viscoelastic fluid. The microstructure of the frozen, dried gel G3 was observed
by an ESEM. Figs. 8c and 8d show a continuous 3D network structure. The amide groups (–CONH2) of the terpolymer provide cross-
linking sites to form a 3D network structure by crosslinking with the hydroxyl groups (–CH2OH) of the phenolic ring of the crosslinker
system. The mesh size was approximately 10 mm, and the thickness of the grid was substantially large, so the network structures of the
bulk gel were not easily broken during the long-term thermal vibrations under the high temperature, and free water did not easily escape
from the meshes. This stable structure shows an excellent water-holding capacity and a very low syneresis. Therefore, this gel system
has a long-term thermal stability in extremely high-temperature conditions.

Plugging Capacity of the Gel System in Fractured Core. Two fractured-core-flow experiments were performed to study the plug-
ging capacity of the gel systems when fractures cause conformance problems. The petrophysical properties of the natural cores from
Jidong Oil Field used in the experiment are shown in Table 4. The schematic of fracturing the natural core is shown in Fig. 2, and the
fracture opening was controlled by the diameters of the copper wire, which were 0.20 and 0.40 mm, respectively. The compositions of
the gel system used in the fractured-core-plugging experiment are also shown in Table 4.
The results indicate that when the fracture openings increased from 0.20 to 0.40 mm, the pressure drop across the core decreased
because the water-channel width increased, and thus the flow resistance decreased. After being injected with 0.6-FV gelant and 0.15-
PV polymer solution and kept at 150 C for 1 week, the fractured cores were flooded by brine again at the rate of 0.3 cm3/min. The
pressure drops were 24.32 and 34.76 kPa, respectively, which shows that stable gel systems were formed in the fractures. The plugging
efficiency reached 98.48 and 99.40%, respectively. The gel system composed of the terpolymer and the crosslinker system could be
used for conformance control in extremely high-temperature reservoirs with small fractures by adjusting the compositions of the gel
system. It is worth noting that gel treatments in many oilfield applications were less effective than expected or tested in reducing water
production from fractured wells because gel might be extruded through fractures and washed out during chase-brine injection (Seright
2002). More experiments related to our gel-system behavior in fractures should be conducted before its larger application in Jidong Oil
Field, such as for gel and gelant extrusion, water leakoff, and gel dehydration.
Plugging Capacity of the Gel System in Homogeneous Core. One homogeneous-core-flow experiment was performed to study the
plugging capacity of the gel systems when the conformance is caused by permeability contrast without fractures. Core-300 with the gas
permeability of 300 md was from Jidong Oil Field and cleaned with toluene before the coreflooding test. The compositions of the gel
system used in this plugging experiment are 0.6% ZP-4 and 0.6% OC-3. Coreflooding-test results are shown in Fig. 9.
The differential pressure during waterflooding is 15.90 kPa (i.e., the water permeability is 106.81 md). When the gelant was injected
into the homogeneous core, the differential pressure increased sharply and reached 255.9 kPa when the injected PV was 1.5 PV, which
shows that high-pressure gradient pressures are needed when the gelant is injected to the porous media of Jidong Oil Field. However,
after the core was aged at 150 C for 3 days, the differential pressure during chase waterflooding reached more than 198.3 kPa, which
shows strong plugging efficiency (i.e., 11.47) in the 300-md core. After the gel system is formed in the porous media, the gel will form
a continuous 3D network structure to block the pore throats in the core. As depicted by the yellow circles in Fig. 10, most of the pore
throats in the porous media were plugged by this terpolymer gel systems. Therefore, this gel system shows great potential to plug the
porous media; however, in a real oilfield application, we should design injection parameters (e.g., gel composition, gel volume, injec-
tion rate) according to the requirement of maximum injection pressure.

2018 SPE Reservoir Evaluation & Engineering 7

ID: jaganm Time: 14:43 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028


REE186235 DOI: 10.2118/186235-PA Date: 28-June-18 Stage: Page: 8 Total Pages: 10

(a) (b)

(c) (d)

5/23/2016 HV HFW mag WD del 100 μm 5/23/2016 HV HFW mag WD det 20 μm


7:25:25 PM 20:00 kV 298 μM 1 000× 11.7 mm ETD Quanta FEG 7:32:30 PM 20:00 kV 59.7 μM 5 000× 11.8 mm ETD Quanta FEG

Fig. 8—Microstructure of the gel system composed of 0.6% terpolymer and 0.6% OC-3. (a, b) Pictures of the bulk gel system; (c, d)
ESEM images of the bulk gel at 1,000X and 5,000X, respectively.

Matrix Gas Fracture


Permeability Porosity Opening Δp0 Δp1 Plugging
Number (md) (%) (mm) Gel System (kPa) (kPa) Efficiency (%)
P2 314 22.37 0.20 0.4% ZP-4 and 0.6% OC-3 0.37 24.32 98.48
P3 322 22.54 0.40 0.6% ZP-4 and 0.6% OC-3 0.21 34.76 99.40

Table 4—Plugging capacity of gel system in fractured cores.

300
Water Gelant Chase water

250

200
Pressure (kPa)

150

100

50 Core-300: Kw = 106.81 md
Displacing rate: 0.5 cm3/min

0
0 1 2 3 4 5 6 7 8
PV

Fig. 9—Coreflooding experiment in 300-md homogeneous core.

8 2018 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 14:43 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028


REE186235 DOI: 10.2118/186235-PA Date: 28-June-18 Stage: Page: 9 Total Pages: 10

12/3/2016 HV HFW mag WD det 10 μm


4:22:19 PM 20:00 kV 29.8 μM 10 000 × 15.4 mm ETD Quanta FEG

Fig. 10—ESEM pictures of the high-permeability core after being treated by the terpolymer-gel system. The yellow circles show the
gels in the pore throats.

Conclusion
In this work, in-situ polymer gels formed by the terpolymer and the organic crosslinker system OC-3 were systematically studied as a
plugging agent for conformance control in extremely high-temperature reservoirs. The gelation time and the gel strength of the in-situ
polymer-gel systems could be controlled by changing the concentrations of the polymer and crosslinker system. Different gel strengths
could meet the various requirements of reducing the permeability in the near-wellbore or far-from-wellbore region, and different gela-
tion times could be obtained to make the gelant easily enter the target areas. The in-situ polymer-gel system had a good mechanical-
shear and porous-media-shear stability and excellent long-term thermal stability; the syneresis of the gel system was very low even after
being heated for 5 months at 150 C. A continuous 3D network structure can be formed by the reaction between the amide groups
(–CONH2) of the polymer and the hydroxyl groups (–CH2OH) of the phenolic ring of the crosslinker system, which contributes to the
long-term stability and the low syneresis of the gel system, and thus the gel system is ready for field testing to solve the excessive
water-production problems in extremely high-temperature reservoirs.

Nomenclature
E ¼ plugging efficiency, %
G0 ¼ storage modulus, Pa
G00 ¼ loss modulus, Pa
Kw0 ¼ absolute permeability to water before the gel treatment, md
Kw1 ¼ absolute permeability to water after the gel treatment, md
S ¼ syneresis of the bulk gel system, %
W ¼ weight of the gelant solution, g
W0 ¼ weight of the aged gel without free water, g
Dp0 ¼ pressure drop before the gel treatment, kPa
Dp1 ¼ pressure drop after the gel treatment, kPa

Acknowledgment
This work was supported by the National Science and Technology Major Project (No. 2016ZX05014-004-004) and the China Scholar-
ship Council (No. 201606440051). The paper is a revised and improved version of the conference paper (SPE-186235-MS) presented at
the SPE/IATMI Asia Pacific Oil and Gas Conference and Exhibition, Jakarta, 17–19 October 2017.

References
Al-Muntasheri, G. A., Nasr-El-Din, H. A., Al-Noaimi, K. et al. 2009. A Study of Polyacrylamide-Based Gels Crosslinked With Polyethyleneimine. SPE
J. 14 (2): 245–251. SPE-105925-PA. https://doi.org/10.2118/105925-PA.
Al-Muntasheri, G. A., Nasr-El-Din, H. A., and Zitha, P. L. 2008. Gelation Kinetics and Performance Evaluation of an Organically Crosslinked Gel at
High Temperature and Pressure. SPE J. 13 (3): 337–345. SPE-104071-PA. https://doi.org/10.2118/104071-PA.
Bai, B. and Zhang, H. 2011. Preformed-Particle-Gel Transport Through Open Fractures and Its Effect on Water Flow. SPE J. 16 (2): 388–400. SPE-
129908-PA. https://doi.org/10.2118/129908-PA.
Brattekås, B., Steinsbø, M., Graue, A. et al. 2017. New Insight Into Wormhole Formation in Polymer Gel During Water Chase Floods With Positron
Emission Tomography. SPE J. 22 (1): 32–40. SPE-180051-PA. https://doi.org/10.2118/180051-PA.
Bryant, S. L., Bartosek, M., Lockhart, T. P. et al. 1997. Polymer Gelants for High Temperature Water Shutoff Applications. SPE J. 2 (4): 447–454. SPE-
36911-PA. https://doi.org/10.2118/36911-PA.
Bryant, S. L., Borghi, G. P., Bartosek, M. et al. 1997. Experimental Investigation on the Injectivity of Phenol-Formaldehyde/Polymer Gelants. Presented
at International Symposium on Oilfield Chemistry, Houston, 18–21 February. SPE-37244-MS. https://doi.org/10.2118/37244-MS.

2018 SPE Reservoir Evaluation & Engineering 9

ID: jaganm Time: 14:44 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028


REE186235 DOI: 10.2118/186235-PA Date: 28-June-18 Stage: Page: 10 Total Pages: 10

Caulfield, M. J., Qiao, G. G., and Solomon, D. H. 2002. Some Aspects of the Properties and Degradation of Polyacrylamides. Chem. Rev. 102 (9):
3067–3084. https://doi.org/10.1021/cr010439p.
Chauveteau, G., Omari, A., Tabary, R. et al. 2000. Controlling Gelation Time and Microgel Size for Water Shutoff. Presented at SPE/DOE Improved Oil
Recovery Symposium, Tulsa, 3–5 April. SPE-59317-MS. https://doi.org/10.2118/59317-MS.
El-Karsani, K. S. M., Al-Muntasheri, G. A., and Hussein, I. A. 2014. Polymer Systems for Water Shutoff and Profile Modification: A Review Over the
Last Decade. SPE J. 19 (1): 135–149. SPE-163100-PA. https://doi.org/10.2118/163100-PA.
El-Karsani, K. S. M., Al-Muntasheri, G. A., Sultan, A. S. et al. 2015. Gelation of a Water-Shutoff Gel at High Pressure and High Temperature: Rheologi-
cal Investigation. SPE J. 20 (5): 1103–1112. SPE-173185-PA. https://doi.org/10.2118/173185-PA.
Eriksen, O. I., Daasvatn, K., Vigerust, B. et al. 1997. Gel Formation and Thermal Stability of Gels Made From Novel Water-Soluble Polymers for
Enhanced Oil Recovery Applications. Presented at International Symposium on Oilfield Chemistry, Houston, 18–21 February. SPE-37247-MS.
https://doi.org/10.2118/37247-MS.
Gaillard, N., Thomas, A., Bataille, S. et al. 2017. Advanced Selection of Polymers for EOR Considering Shear and Hardness Tolerance Properties. Pre-
sented at IOR 2017–19th European Symposium on Improved Oil Recovery, Stavanger, Norway, 24–27 April. https://doi.org/10.3997/2214-
4609.201700333.
Goudarzi, A., Zhang, H., Varavei, A. et al. 2015. A Laboratory and Simulation Study of Preformed Particle Gels for Water Conformance Control. Fuel
140 (15 January): 502–513. https://doi.org/10.1016/j.fuel.2014.09.081.
Leonhardt, B., Ernst, B., Reimann, S. et al. 2014. Field Testing the Polysaccharide Schizophyllan: Results of the First Year. Presented at SPE Improved
Oil Recovery Symposium, Tulsa, 12–16 April. SPE-169032-MS. https://doi.org/10.2118/169032-MS.
Moradi-Araghi, A. 2000. A Review of Thermally Stable Gels for Fluid Diversion in Petroleum Production. J. Pet. Sci. Eng. 26 (1–4): 1–10. https://
doi.org/10.1016/S0920-4105(00)00015-2.
Portwood, J. T. 2005. The Kansas Arbuckle Formation: Performance Evaluation and Lessons Learned From More Than 200 Polymer-Gel Water-Shutoff
Treatments. Presented at SPE Production Operations Symposium, Tulsa, 16–19 April. SPE-94096-MS. https://doi.org/10.2118/94096-MS.
Rivenq, R. C., Donche, A., and Nolk, C. 1992. Improved Scleroglucan for Polymer Flooding Under Harsh Reservoir Conditions. SPE Res Eeng 7 (1):
15–20. SPE-19635-PA. https://doi.org/10.2118/19635-PA.
Seright, R. S. 1983. The Effects of Mechanical Degradation and Viscoelastic Behavior on Injectivity of Polyacrylamide Solutions. SPE J. 23 (3):
475–485. SPE-9297-PA. https://doi.org/10.2118/9297-PA.
Seright, R. S. 2002. An Alternative View of Filter Cake Formation in Fractures. Presented at SPE/DOE Improved Oil Recovery Symposium, Tulsa,
13–17 April. https://doi.org/10.2118/75158-MS.
Seright, R. S. 2003. An Alternative View of Filter-Cake Formation in Fractures Inspired by Cr(III)-Acetate-HPAM Gel Extrusion. SPE Prod & Fac 18
(1): 65–72. SPE-81829-PA. https://doi.org/10.2118/81829-PA.
Seright, R. S., Lane, R. H., and Sydansk, R. D. 2001. A Strategy for Attacking Excess Water Production. Presented at the SPE Permian Basin Oil and
Gas Recovery Conference, Midland, Texas, 15–17 May. SPE-70067-MS. https://doi.org/10.2118/70067-MS.
Sydansk, R. D. and Argabright, P. A. 1987. Conformance Improvement in a Subterranean Hydrocarbon-Bearing Formation Using a Polymer Gel. US
Patent No. 4,683,949.4.
Zhao, G., Dai, C., Zhang, Y. et al. 2015. Enhanced Foam Stability by Adding Comb Polymer Gel for In-Depth Profile Control in High Temperature Res-
ervoirs. Colloid. Surface. A 482 (5 October): 115–124. https://doi.org/10.1016/j.colsurfa.2015.04.041.
Zhu, D., Bai, B., and Hou, J. 2017. Polymer Gel Systems for Water Management in High-Temperature Petroleum Reservoirs: A Chemical Review.
Energy Fuels 31 (12): 13063–13087. https://doi.org/10.1021/acs.energyfuels.7b02897.

Daoyi Zhu is a PhD degree student at China University of Petroleum, Beijing, a joint PhD degree student at Missouri University of
Science and Technology, and a member of the editorial board of Petroleum Industry Press. Previously, he worked for 1 year
at China University of Petroleum as a research assistant. Zhu’s research interests include colloidal and interfacial chemistry,
oilfield chemistry, chemical enhanced oil recovery (EOR), conformance control, and reservoir engineering. He has authored or
coauthored more than 10 technical papers and holds one patent. Zhu holds a bachelor’s degree in applied chemistry from
Yangtze University, China, and a master’s degree in chemical engineering from China University of Petroleum, Beijing. He is a
member of SPE.
Jirui Hou is a professor and a PhD supervisor at China University of Petroleum, Beijing, the director of Research Institute of
Enhanced Oil Recovery, the associate director of the China National Petroleum Corporation Tertiary Oil Recovery Key
Laboratory–Applied Fundamental Theory Research Laboratory of EOR for Low Permeability Oilfield, and a member of the China
Standard Committee of Oilfield Chemicals Industry. His research interests include chemical flooding, CO2 flooding, gas-
channeling treatment, and conformance-control theory and technology. Hou has authored or coauthored more than 130
technical papers and holds more than 10 patents. He holds a PhD degree in applied chemistry from Dalian University of
Technology, China.
Qi Wei is a master’s degree student at China University of Petroleum, Beijing. His research interests include CO2 EOR, foam EOR,
and conformance control. Wei has authored or coauthored more than five technical papers. He holds a bachelor’s degree in
petroleum engineering from Yangtze University. Wei is a member of SPE.
Yuguang Chen is a master’s degree student at China University of Petroleum, Beijing. His research interests include EOR,
oilfield chemistry, and conformance control. Chen has authored or coauthored more than five technical papers. He holds a
bachelor’s degree in petroleum engineering from China University of Petroleum (East China). Chen is a member of SPE.

10 2018 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 14:44 I Path: S:/REE#/Vol00000/180028/Comp/APPFile/SA-REE#180028

You might also like