You are on page 1of 5

Journal of Industrial and Engineering Chemistry 21 (2015) 741–745

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


journal homepage: www.elsevier.com/locate/jiec

Enhanced oil recovery performance and viscosity characteristics


of polysaccharide xanthan gum solution
Hee Yeon Jang a, Ke Zhang b,c, Bo Hyun Chon a,**, Hyoung Jin Choi c,*
a
Department of Energy Resources Engineering, Inha University, Incheon 402-751, Republic of Korea
b
School of Chemical Engineering and Technology, Harbin Institute of Technology, Harbin 150001, People’s Republic of China
c
Department of Polymer Science and Engineering, Inha University, Incheon 402-751, Republic of Korea

A R T I C L E I N F O A B S T R A C T

Article history: This study compared the potential applications of xanthan gum and hydrolyzed polyacrylamide (HPAM)
Received 2 January 2014 as polymer-flooding agents for heavy oil recovery applications under a range of salinity conditions.
Accepted 4 April 2014 Rheological measurements were carried out to examine the change in shear viscosity when the polymer
Available online 13 April 2014
was applied under a range of reservoir conditions. The results showed that the shear viscosity of the
xanthan gum solution was less sensitive to increasing temperatures and salinity than that of the HPAM
Keywords: solution. Accordingly, a xanthan gum injection is more effective than HPAM under higher salinity
Xanthan gum
reservoir conditions.
Rheology
Polymer flooding
ß 2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
Enhanced oil recovery reserved.

1. Introduction the poor water/oil mobility ratio responsible for the conformance
control problems that lead to poor water flood performance on
Since the introduction of chemical enhanced-oil-recovery (EOR) intermediate heavy oil [6,7]. Therefore, the polymer-induced EOR
methods several decades ago, it was often considered to be technique is widely accepted for various mature oil fields [2].
unprofitable to apply those technologies to heavy oils because of Currently, both water-soluble polyacrylamide and polysacchar-
the low oil prices. On the other hand, with the current increases in ides are used widely in oil fields to enhance oil recovery. Among
the demand and prices of oil, which are led by increasing these, xanthan gum, a natural polysaccharide and an important
oil consumption in countries, such as China and India, heavy oil industrial biopolymer, has attracted considerable attention as an
reservoirs have attracted considerable attention from the oil EOR agent in oil drilling, fracturing, pipeline cleaning [8]. Xanthan
industry. Therefore, heavy oil production needs to be increased gum is an extracellular polysaccharide produced by the fermenta-
using the EOR techniques [1,2]. Common EOR methods include tion of a cellulosic backbone consisting of five monosaccharides
thermal, miscible, and chemical processes, of which chemical by the bacterium, Xanthomonas campestris, to give a pentasacchar-
treatment is one of the main methods to limit water production [3]. ide repeating unit. The cellulosic backbone is substituted at C-3 on
This has been operated by building up flow resistance, which is the alternate beta-1,4-D-glucopyranosyl residues with the trisac-
achieved by increasing the shear viscosity of the injected water charide side chains of beta-D-rhamnopyranosyl, beta-1,4-D-glu-
and/or reducing its permeability through polymer adsorption [4]. curonopyranosyl and alpha-1,2-D-mannopyranosyl with various
In particular, a polymer flooding method is considered one of amounts of acetyl and pyruvate substituents [9]. The backbone of
the most promising chemical EOR processes in many reservoirs the polymer is similar to that of cellulose. Therefore, in addition to
due to its lower cost. Because high molecular weight water-soluble the EOR area, the xanthan gum is also used widely in engineering,
polymers can increase the viscosity of the aqueous phase easily, it such as a viscosity-enhancing agent in foods, in cosmetics and
also results in an increase in sweep efficiency during EOR processes pharmaceuticals, and turbulent drag reduction with interesting
[5]. In addition, the shear viscosity of a polymer solution corrects characteristics including good temperature stability, fine emulsion
stabilization, and compatibility with food ingredients [8,10].
The strong ability to increase the shear viscosity, coupled with
* Corresponding author. Tel.: +82 328607486.
its excellent stability under high salinity, high temperature and
** Corresponding author. mechanical shear conditions, makes it suitable for EOR. Because
E-mail addresses: bochon@inha.ac.kr (B.H. Chon), hjchoi@inha.ac.kr (H.J. Choi). the viscosity of the displacing fluid is vital in EOR projects, it is

http://dx.doi.org/10.1016/j.jiec.2014.04.005
1226-086X/ß 2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
742 [(Fig._1)TD$IG]
H.Y. Jang et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 741–745

important to control the viscosity of xanthan gum under given


salinity and temperature conditions [11]. In both EOR and
polymer-induced turbulent drag reduction, very dilute solutions
of flexible polymers can have extremely high extensional
viscosities, which can be orders of magnitude higher than those
expected based on a Newtonian fluid [12].
In this study, the polymer flooding characteristics of xanthan
gum were examined to explore the possibility of improving heavy
oil recovery in a range of salinities and compared to those of
hydrolyzed polyacrylamide (HPAM) solutions. In addition, rheo-
logical analysis of xanthan gum was also carried out to assess its
relation to the polymer flooding.

2. Experimental

The polymer solutions were prepared by dissolving the


respective polymer powder in distilled water using a magnetic
stirrer for five days at room temperature. A salt concentration of up
to 20 wt% and two types of commercial polymers, xanthan gum Fig. 1. Shear stress versus shear rate of xanthan gum solution at different
with a mean average molecular weight of 5  106 Da and HPAM concentrations at 25 8C. The solid lines are obtained from Herschel–Bulkley model.
with a molecular weight of 13  106 Da, were used. The xanthan
gum was 99% pure (Sigma–Aldrich Co., Germany). HPAM 3. Results and discussion
(AN910VHM, OCI-SNF) was obtained from OCI, Korea with a
12 mol% degree of hydrolysis [6]. These polymers are soluble in Fig. 1 shows the shear stress versus shear rate of xanthan gum
aqueous sodium chloride solutions, and have been used success- solutions at different concentrations measured using a rotational
fully to reduce the mobility of injected fluids and EOR from oil rheometer at 25 8C at 3 wt% salinity. The xanthan gum solutions
fields as well as in water-control operations. exhibited non-Newtonian behavior, and an increase in the xanthan
Rheological measurements were conducted using a rotational gum concentration led to an increase in yield stress. The Herschel–
rheometer (Physica, MCR 300, Austria) with a Couette cell Bulkley model (Eq. (1)) [12] was used to fit the experimental data.
geometry (DG 26.7) to examine the characteristics of the xanthan
gum solutions at concentrations ranging from 1500 to 8000 ppm in t ¼ t 0 mġ n (1)
a 3 wt% saline solution. Some polymer solutions were prepared by
mixing with synthetic brine (1–20 wt% salinity), which resembled where t is the shear stress, ġ is the shear rate, t0 is the yield stress,
the concentrations of local reservoirs that contain 99% sodium m is the consistency coefficient, and n is the flow behavior index.
chloride. The polymer solutions of the 3 wt% saline solutions were Table 1 lists the parameters obtained from this Herschel–Bulkley
tested at 25, 50, 70 and 85 8C. Rheological tests were carried out to model.
examine the effects of salinity and temperature on the polymer Fig. 2 shows the steady shear viscosity of the xanthan gum
solutions at a concentration of 1500 ppm. solutions at different concentrations as a function of the shear rate.
Furthermore, to perform the polymer flooding tests, the core An increase in shear viscosity was observed for the increased
holder was filled with glass beads, and compacted tightly for more xanthan gum concentration. For polymer flooding in the EOR, the
than 1 h. Through flow experiments, the porosity and permeability, increased shear viscosity of the xanthan gum can increase the
measured using glass beads ranging in diameter from 63 mm to residual resistance factor and improve the sweep efficiency [9].
106 mm, were 36.9% and 3.79 Darcy, respectively. These experi- The data clearly shows that the Newtonian viscosity region noted
ments were conducted in a glass-bead pack saturated with 3.0 wt% at low shear rates was not observed at high concentrations.
saline water and a viscous crude oil with 450 cP (obtained from SK Instead, significant shear-thinning behavior was observed, which
Energy, Korea) at 25 8C. After packing the core holder with glass was attributed to the disentanglement and alignment of the
beads, it was placed vertically, and the brine was injected from its xanthan gum polymer chains along the flow direction. In a drilling
inlet. Heavy oil was also injected in the same direction using fluid of the EOR, shear thinning polymer solutions were designed
an ISCO pump and a transfer cylinder until the oil reached the that can suspend drilling cutting at low shear rates, but offer little
outlet. A detailed schematic diagram of the experimental set-up is resistance to flow at high shear rates [5]. On the other hand,
described elsewhere [7]. To restore the wettability, the core polysaccharides that display interesting solution characteristics
holder was filled with the formation brine and heavy oil, and left tend to form ordered structures in aqueous environments [13].
to saturate for 24 h. The fluid injection rate was maintained at Large molecules in an aqueous solutions form aggregates and
4 ml/min for the flooding experiments, and all of which were entangle, leading to high viscosity at low shear rates. Increasing the
carried out at the same injection rate. In the water flood tests, shear rate decreases the steady shear viscosity due to the
water was injected at a 0.6 pore volume. Because water uncoiling and partial alignment of the polymer chains at the high
breakthrough occurred after a 0.6 pore volume, the polymer
injection tests were started from this point until the water flood
reached 3 pore volumes [7]. Table 1
Fitting parameters for the xanthan gum solutions at various concentrations
The xanthan gum polymer solution injection tests began at
calculated from Herschel–Bulkley model.
oil saturation after the water flood, and the effects of oil
recovery were examined. The increased shear viscosity of the t0 m n
polymer solution improved the tertiary oil recovery. Within 1500 ppm 15 100 0.46
the optimal viscosity range of the polymer solution, an increase 3000 ppm 300 300 0.47
5000 ppm 1500 401 0.53
in the effective viscosity of the polymer solution resulted in
8000 ppm 4300 469 0.57
an increase in oil production [7].
[(Fig._2)TD$IG] [(Fig._3)TD$IG]
H.Y. Jang et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 741–745 743

Fig. 3. Xanthan gum (1500 ppm) viscosity at various brine concentrations at 25 8C


versus shear rate.
[(Fig._4)TD$IG]
Fig. 2. Effect of xanthan gum concentration (3 wt% NaCl) on shear viscosity at 25 8C.
The solid lines are obtained from Eq. (2).

shear rate region. Xanthan gum solutions are highly pseudoplastic.


Although they show a shear-thinning behavior under an applied
shear rate, the initial viscosity is recovered almost instantaneously
when the shear rate is removed [14]. The shear viscosity and shear
rate of the xanthan gum solutions were fitted with Eq. (2).
The Carreau model described the shear-thinning behavior of the
polymers well. This is more realistic than the power law model
because it represents the data quite well at both high and low shear
rates. The curve fitting parameters were h0, l and n, where h0 is the
zero-shear viscosity, h1 is the infinite-shear viscosity, ġ is the
shear rate, and the parameter l has a time dimension, as shown in
Table 2 [9,15,16].

h  h1 n1=2
¼ ½1 þ ðlġ Þ2  (2)
ho  h1 Fig. 4. Xanthan gum (1500 ppm) viscosity at high brine concentrations at 25 8C
versus shear rate.
Using aqueous xanthan gum solutions at a relatively low
concentrations, the Cross equation was adopted successfully to fit
the shear viscosity as a function of the shear rate [9]. molecule is in an ordered state, it is relatively insensitive to further
To examine the effects of salt on the viscosity of xanthan gum, increases in salt concentration [17].
the shear viscosity was measured as a function of the shear rate for The addition of NaCl to the HPAM solution (1500 ppm) reduced
different salt concentrations at 25 8C, as shown in Figs. 3 and 4. the polymer solution viscosity significantly (Fig. 5), as reported
The addition of 1 wt% salt made the shear viscosity lower than that previously [7]. This was attributed to the addition of Na+ cations,
of a salt-free solution but further salt addition had little effect. In which effectively screen the negatively charged carboxyl groups to
the salt-induced transition from a disordered to ordered state, the reduce electrostatic repulsion within the polymer chains [18].
polymer backbone acquires a helical conformation, whereas The shear viscosity of HPAM in 0, 10 and 20 wt% saline solutions
the charged trisaccharide side-chains collapse onto the backbone measured at a 11.7 s1 shear rate was 108, 13.1 and 14.9 mPa s,
(due to charge screening effects) to stabilize the ordered respectively. According to the viscosities of xanthan gum and
conformation [16]. The xanthan gum viscosity was relatively HPAM for different salt concentrations measured at an 11.7 s1
insensitive to salt concentrations of up to 20 wt%, i.e. 55.3, 31, 27.8, shear rate, the HPAM solution was more sensitive to salinity than
16.5, 20.5 and 25. 1 mPa s with an experimental error of 5% for 0, 1, to the xanthan gum solution. Nevertheless, HPAM with zero
3, 5, 10 and 20 wt% salts, respectively, measured at a 11.7 1 s1 salinity had higher viscosity than the xanthan gum solution with
shear rate. On the other hand, this generalization ignores the initial zero salinity.
effects that occur at very low salt levels, which drives the transition Xanthan gum also has both helix and random coil conforma-
from a disordered state to an ordered conformation. Once the tions that depend strongly on the dissolution temperature [8].
Fig. 6 shows the shear rate dependence of the steady shear
Table 2
Fitting parameters for the xanthan gum solutions at various concentrations
calculated from Carreau model. Table 3
Viscosity (cP) of xanthan gum and HPAM aqueous solutions (1500 ppm, 3 wt% NaCl)
h0 h1 l n at 11.7 s1 shear rate.
1500 ppm 126 3.04 0.45 0.18
11.7 s1 25 8C 50 8C 70 8C 85 8C
3000 ppm 1310 8.14 1.2 0.1
5000 ppm 6960 15.2 3.0 0.08 Xanthan gum 27.8 16.1 10.4 11.8
8000 ppm 22,300 25.8 3.5 0.03 HPAM 10.2 5.49 3.3 3.79
[(Fig._5)TD$IG]
744 H.Y. Jang et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 741–745

Table 5
Comparison of water and polymer flood oil recovery.

Flood types Water flood Tertiary Final recovery %,


recovery recovery OOIP

%, OOIP Injected, %, OOIP %, ROIP


(PV)

Water flood 39.0 3.0 39.0

Polymer flood
HPAM 3 wt% 24.0 0.6 37.9 81.3 61.9
Xanthan gum 3 wt% 24.2 0.6 44.9 90.2 69.1
Xanthan gum 10 wt% 24.5 0.6 42.9 88.9 67.4

cause chain expansion due to the repulsion of ionic groups, which


leads to higher shear viscosity [5]. As shown in Table 3, the xanthan
gum solutions were less sensitive to high temperatures than HPAM
Fig. 5. HPAM (1500 ppm) viscosity at high brine concentrations at 25 8C versus solutions. The effect of temperature on the viscosity of the HPAM
shear rate. solutions was complicated because their viscosity can also increase
[(Fig._6)TD$IG] due to the enhanced hydrolysis of HPAM at high temperatures [16].
The aims of this study were (i) to consider the feasibility of
varying the salinity of a polymer flood and (ii) to compare how
much heavy oil can be recovered in a water flood versus a polymer
flood. To examine the applicability of a polymer flood in the field,
experiments were conducted to obtain data on the properties, such
as porosity, absolute permeability and oil saturation. The results
are summarized in Table 4, where Soi* is the initial oil saturation
and Soi** is the oil saturation after the water flood experiments.
The measured initial oil saturation ranged from 91% to 92.9% of the
distribution, whereas porosity measurements ranged from 36.9%
to 38.1%. Table 5 lists the water and polymer flood test results. In
the water flood, an oil recovery of 39% and a water cut of
approximately 98% were obtained after injecting 3.0 PV. In
contrast, the injection of a HPAM aqueous solution under the
same salinity recovered 61.9% of the original oil in place (OOIP) [7].
In the case of he second test, which used a xanthan gum flood at
3 wt% NaCl, a total oil recovery of 69.1% was obtained by an injection
Fig. 6. Viscosity of xanthan gum aqueous solutions (1500 ppm, 3 wt% NaCl). of 0.6–3.0 PV at an oil saturation of 74.8% after the water flood. When
0.6 PV was reached, the water broke through and the oil recovery
was 24.2% using the water flood. The water cut reached
viscosity for aqueous xanthan gum solutions at 25, 50, 70, and approximately 95%, as shown in Fig. 7. Similarly, the total oil
85 8C. The viscosity decreased marginally with increasing temper- recovery was 67.4% for the xanthan gum solution based flooding test
ature, showing a lack of temperature sensitivity compared to the at 10 wt% salinity. All the experiments maintained an injection rate
partial HPAM [7]. The change in viscosity was attributed to a of 4 ml/min, which was controlled by the ISCO pump. Powdered
conformation transition from a rod-like structure (ordered) to a xanthan gum and the 3 wt% and 10 wt% salinity solutions were
more flexible structure (disordered) when the temperature was blended to prepare the xanthan gum solution for the second polymer
increased because the rigidity of the (1-4)-b-D-glucan chain
decreases progressively with increasing temperature between 40
[(Fig._7)TD$IG]
flood test. The main purpose of using polymers in the EOR process

and 60 8C [19].
The shear viscosity of a pure HPAM solution was also reported
to be strongly dependent on temperature, i.e. it decreases with
increasing temperature [7]. With increasing temperature, the
intermolecular interactions decrease due to an increase in the
thermal motion of the molecules [20]. Nevertheless, the HPAM has
been adopted widely in the EOR because its carboxylate groups

Table 4
Properties measured in injection experiments.

Flood types Porosity Absolute permeability Soi* Soi**


(%) (Darcy) (%) (%)

Water flood 36.9 3.79 91.0 73.1

Polymer flood
HPAM (3 wt%) 37.8 3.15 91.9 76.0
Xanthan gum (3 wt%) 37.2 3.41 91.7 74.8
Fig. 7. Comparison of oil recoveries in xanthan gum polymer flood tests at 3 wt% and
Xanthan gum (10 wt%) 38.1 3.38 92.9 75.2
10 wt% NaCl.
H.Y. Jang et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 741–745 745

Table 6 was to increase the solution viscosity and reduce the rock
Comparison of water and polymer flood pressure data.
permeability, both of which help increase the sweep efficiency by
Flood Pressure at water Pressure at reducing the mobility of the displacing fluid [20].
types flood recovery tertiary recovery Table 6 lists the water and polymer flood pressure data. Figs. 8
Initial Initial Injected, Maximum Initial Stable and 9 revealed an unstable pressure drop after breakthrough in
in-let pressure (PV) in-let pressure pressure, various flood tests. This effect can be explained by the occurrence
pressure, drop, pressure, drop, psi of the viscous fingering at the outlet face. In addition, the
psi psi psi psi
maximum inlet pressure was high in the xanthan gum flood
Water flood 190.7 74.3 3.0 – – 1.8 (Table 6), which is closely related to the polymer viscosity.
Polymer flood
HPAM 3 wt% 214.1 77.8 0.6 190.4 125.8 7.3 4. Conclusions
Xanthan gum 219.7 75.2 0.6 431.6 260.2 47.8
3 wt%
This study examined the applicability of xanthan gum at
Xanthan gum 217.8 74.5 0.6 422.9 253.1 45.4
10 wt% different salinities to heavy oil recovery in a high salinity
[(Fig._8)TD$IG] reservoir. The rheological characteristics of the two types of
polymer solutions at different salinities, concentrations and
temperatures were examined and compared in terms of the
viscosity. The steady shear viscosity of the xanthan gum and
HPAM solutions exhibited non-Newtonian fluid characteristics at
each concentration at 25 8C. The rheological test results showed
that the HPAM solution was more sensitive to high temperatures
and salinity than the xanthan gum solution. In the flood test
performed with the xanthan gum solution at 3 wt% salinity, the
oil recovery was 7.2% higher than that using the HPAM solution.
The xanthan gum solution floods at 3 wt% and 10 wt% salinity did
not show a significant difference in oil recovery. These results
will be useful in studies testing the feasibility of applying
xanthan gum flooding under high-salinity reservoir conditions.

Acknowledgment

One of the authors (HJC) wishes to acknowledge the partial


financial support from National Research Foundation of Korea
(NRF-2009-0080253).
Fig. 8. Pressure drops in water and polymer floods.
References
[(Fig._9)TD$IG]
[1] W. Wang, Y.Z. Liu, Y.G. Gu, Colloid Polym. Sci. 281 (2003) 1046.
[2] Q. Chen, Y. Wang, Z. Lu, Y. Feng, Polym. Bull. 70 (2013) 391.
[3] F. Chen, H. Jiang, X. Bai, W. Zheng, J. Ind. Eng. Chem. 19 (2013) 450.
[4] I.R. Rangel, R.L. Thompson, R.G. Pereira, F.L.B. de Abreu, J. Braz. Soc. Mech. Sci. Eng.
34 (2012) 285.
[5] K.C. Taylor, H.A. Nasr-El-Din, J. Petrol. Sci. Eng. 19 (1998) 265.
[6] R. Qiao, W. Zhu, J. Ind. Eng. Chem. 16 (2010) 278.
[7] (a) J.C. Jung, K. Zhang, B.H. Chon, H.J. Choi, J. Appl. Polym. Sci. 127 (2013) 4833;
(b) K.M. Ko, B.H. Chon, S.B. Jang, H.Y. Jang, J. Ind. Eng. Chem. 20 (2014) 228.
[8] A. Palaniraj, V. Jayaraman, J. Food Eng. 106 (2011) 1.
[9] L. Xu, G. Xu, T. Liu, Y. Chen, H. Gong, Carbohydr. Polym. 92 (2013) 516.
[10] C.A. Kim, H.J. Choi, C.B. Kim, M.S. Jhon, Macromol. Rapid Commun. 19 (1998) 419.
[11] A.A. Alquraishi, D. Fares, F.D. Alsewailem, Carbohydr. Polym. 88 (2012) 859.
[12] H.A. Barnes, J.F. Hutton, K. Walters, An Introduction to Rheology, Elsevier,
Amsterdam, 1989p. 78.
[13] J.F. Kennedy, I.J. Bradshaw, Prog. Ind. Microbiol. 19 (1984) 319.
[14] H.J. Choi, S.H. Park, J.S. Yoon, H.S. Lee, S. Choi, Polym. Eng. Sci. 35 (1995) 1636.
[15] Y.H. Hyun, S.T. Lim, H.J. Choi, M.S. Jhon, Macromolecules 34 (2001) 8084.
[16] G. Muller, M. Anhourrache, J. Lecourtier, G. Chauveteau, Int. J. Biol. Macromol. 8
(1986) 167.
[17] A. Jeanes, J.E. Pittsley, F.R. Senti, J. Appl. Polym. Sci. 5 (1961) 519.
[18] A. Samanta, A. Bera, K. Ojha, A. Mandal, J. Chem. Eng. Data 55 (2010) 4315.
[19] M. Milas, M. Rinaudo, Carbohydr. Res. 76 (1979) 189.
[20] N.M. Quy, P.G. Ranjith, S.K. Choi, P.H. Giao, D. Jasinge, J. Petrol. Sci. Eng. 66 (2009)
Fig. 9. Pressure drops in xanthan gum at 3 wt% and 10 wt% salinity. 75.

You might also like