Vous êtes sur la page 1sur 14

SPE-170800-MS

Surfactant-Based Fluids Containing Copper-Oxide Nanoparticles for Heavy


Oil Viscosity Reduction
Aditya Srinivasan and Subhash N. Shah, Well Construction Technology Center, The University of Oklahoma

Copyright 2014, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in Amsterdam, The Netherlands, 2729 October 2014.
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
Limited research is reported in the area of nano-fluids applicability in heavy oil recovery. Some of the
largest reserves in the world are heavy oil reservoirs. Thermal and chemical methods are widely used for
heavy oil recovery. This research work investigates the viscosity reduction of heavy oil by adding a
surfactant-based fluid containing nanoparticles. The viscosity reducing effect of only the nano-sized
copper-oxide particles without surfactant-based fluid on heavy oil is also examined. Copper (II) oxide
particles were selected due to their higher thermal conductivity, easy availability, and lower cost
compared to metal nanoparticles. Two heavy oil samples (20-21 API) having different viscosities and
acquired from two different wells located in Oklahoma, USA were used for testing. Heavy oil samples
mixed with nano-emulsion containing copper-oxide nanoparticles were investigated at different nanoparticle concentration and oil to emulsion volume ratio. Oil to emulsion volume ratio of (1:0.25) was used
to determine optimum nanoparticle concentration. The optimum nanoparticle concentration was found to
be 0.02% wt., exhibiting the maximum viscosity reduction.The viscosity of oil samples was measured at
various temperatures. The performance of this nano-emulsion was also compared to an industrial viscosity
reducer. In addition interfacial tension tests were performed with the oil samples. A significant reduction
in the interfacial tension of heavy oil was observed after the formation of oil-in-water emulsion. Inherent
properties of copper-oxide nanoparticles in tandem with the oil-in-water emulsion might utilize the
advantages of both thermal and chemical recovery methods to obtain an enhanced heavy oil viscosity
reduction. In conclusion, this nano-emulsion can be an effective oil recovery enhancement option in a
heavy oil reservoir.
Keywords: Heavy oil; Nanoparticles; Viscosity reduction; Copper oxide; Surfactant-based fluids; Oil-in-water
emulsion

Introduction
The demand for oil and gas is likely to increase by 50% globally in the next 20 years [Kong and Ohadi,
2010]. Production of oil and gas has to be increased from both conventional and unconventional resources
in order to meet future energy requirements. Currently, most of the easily accessible oil and gas resources
are produced extensively leading to a decline in conventional petroleum. A few solutions to meet this
growing demand are new oil and gas reservoir discoveries, increasing the recovery of existing wells and

SPE-170800-MS

Table 1Chemical Composition of Heavy Oil Samples


Properties

Burruss Well-4 (S1)

Burruss Well-5 (S2)

Oil Gravity
Wt% Sulphur
Wt% Water
Wt% Asphaltenes
Wt% Paraffins

21.3
0.817
0.67
3.3
1.6

21
0.815
0.96
4.2
2.1-2.2

producing from unconventional resources such as


shale gas, shale oil and heavy oil. Heavy oil resources have a huge potential to meet the future
demand for petroleum products. However, their recovery rates are less than 5-30% of oil in place. The
in-situ viscosity reduction of heavy oil is necessary
for economic production. Conventional enhanced
oil recovery (EOR) methods like chemical, thermal
and gas processes have been successful in improving the recovery rates of heavy oil.
Research in chemical EOR has been carried out
since the past 30 years and the industry now has a
better understanding and experience. Hence, the Figure 1Effect of Nanoparticle Concentration on Apparent Viscosity
feasibility of chemical EOR application in heavy oil of Heavy Oil at 70F
reservoirs is currently being explored [Mai et al.,
2009]. Surfactant-based fluids are used in chemical
EOR due to their ability to form in-situ emulsions
[Bryan and Kantzas, 2007; Bryan et al., 2008].
These fluids exhibit good rheological properties
over a wide temperature range. The use of nanotechnology has provided a means of enhancing the
rheological properties of fluids. A variety of nanofluids can be designed for a specific application by
adding nanoparticles to different base fluids. It has
been found that surfactant-based nano-fluids can be
used to improve recovery of both conventional and
unconventional oil [Qiu, 2010; Birkeland, 2013].
Nanotechnology applications in the petroleum in- Figure 2Effect of Nanoparticle Concentration on Apparent Viscosity
dustry are recent and can help develop economical, of Heavy Oil at 175F
efficient and environment friendly recovery methods [Evdokimov et al., 2006; McElfresh et al., 2012; Fakoya and Shah, 2013].
This study is aimed at investigating the impact of surfactant-based nano-fluids on the rheological and
interfacial tension (IFT) properties of heavy oil and heavy oil-in-water emulsions.

Experimental Setup
Materials and Equipment
Heavy oil, Triton X-100 surfactant, xylene solvent, 2% NaCl brine and 50 nm copper-oxide nanoparticle powder were used in preparing emulsion samples used for rheological and IFT tests. Additionally,
a chemical viscosity reducer supplied by ChemEOR [Shuler et al., 2010] used commercially in the oil and

SPE-170800-MS

gas industry was mixed with heavy oil for the purpose of comparing its performance with the nanoemulsions. The Model 900 OFITE viscometer was
used for rheological measurements. The M6500
Spinning Drop Apparatus and the ATTENSION
Drop Shape Analyzer were employed for IFT measurements.
Fluids Preparation
Two different heavy oil samples henceforth referred
to as S1 and S2 were obtained from the Burruss
fields near Oklahoma City, Oklahoma, USA. Table
1 illustrates the properties of the two heavy oil
Figure 3Rheology Curves of Heavy Oil S1 (0.1% wt. Copper-Oxide
samples.
Nanoparticles)
Firstly, the rheology of heavy oil samples and
heavy oil samples mixed with two different concentrations of copper-oxide nanoparticles (0.1 and
0.5% wt.) were tested using the OFITE viscometer.
The purpose of this test was to evaluate the effect of
copper-oxide nanoparticles on heavy oil viscosity
reduction.
Based on experiments conducted by Qiu (2010),
Triton X-100 was used as the surfactant and xylene as the solvent in 2% NaCl brine (base fluid) in
order to formulate the base emulsion. The fluids
were mixed in a proportion of 95, 1.8 and 3.2% wt.
brine, solvent and surfactant respectively to obtain a
stable base emulsion. Furthermore, viscosity tests
were conducted after mixing the base emulsion with
heavy oil samples at different oil to emulsion volFigure 4 Rheology Curves of Heavy Oil S2 (0.1% wt. Copper-Oxide
ume ratios.
Thereafter, copper-oxide nanoparticles were Nanoparticles)
added to base emulsion (based on base emulsion
weight) to prepare several concentrations of each fluid type. The 1:0.25 oil to emulsion volume ratio was
chosen for nano-emulsion tests.
Base Emulsion [95, 1.8 and 3.2% wt. brine, solvent and surfactant (No nanoparticles)] The
required weight of Triton X-100 and xylene to obtain 3.2 and 1.8% wt. respectively was added to 2% NaCl
brine (95% wt.) and mixed in a blender for 15 20 min while avoiding air entrapment. A foamy solution
was obtained, agitation was stopped, and the fluid was allowed to hydrate for 24 hrs.
Nano-Emulsion [95, 1.8 and 3.2% wt. brine, solvent and surfactant containing nanoparticles] I n
preparing the nano-emulsions, the steps required for mixing the base emulsion were repeated; followed
by the addition of 0.01, 0.02, 0.05, 0.1, and 0.5 % wt. 50 nm copper-oxide nanoparticles while agitation
continued.
Heavy Oil Emulsion [Different volume ratios (No nanoparticles)] The fluid was prepared by adding
the required volume of base emulsion (according to volume fraction) to 250 ml heavy oil in the blender
and mixing it until a homogeneous oil-in-water emulsion was obtained.

SPE-170800-MS

Heavy Oil Containing Nanoparticles Heavy oil


was mixed with 0.1 or 0.5% wt. 50 nm copper-oxide
nanoparticles and blended at high speed (3000 rpm)
for 5 mins to ensure nanoparticle dispersion
throughout the sample.
Heavy Oil containing Industrial Viscosity Reducer
Prepared by blending 1.5% wt. industrial viscosity
reducer with 2% NaCl brine and adding the mixture
to 250 ml heavy oil. Details of the fluids tested are
presented in Table A-1.

Experimental Procedures
Model 900 OFITE Viscometer
Figure 5Effect of Emulsion Volume Fraction on Apparent Viscosity of
Fluid sample was subjected to different shear rates Heavy Oil S1 at 70F
consisting of 7 data points in the range from 5.1 to
511 s1. However, the data points in the range of 102.1 to 511 s1 were considered for analysis based on
the suppliers recommendation for F-2 spring since it was found to be the most appropriate to carry out
rheology tests for the heavy oil samples. Viscosity data were acquired at different temperatures from 70
to 175F.
Spinning Drop Apparatus
The M6500 spinning drop tensiometer cannot measure IFT values above 7 mN/m and is generally used
for low and ultra-low IFT measurements. Hence, it was used to measure IFT of the heavy oil emulsions
and nano-emulsions. A small amount of sample (heavy oil emulsion) was injected into a capillary tube
containing the aqueous phase (deionized water). The sample was then rotated at 4100 rpm in the spinning
drop apparatus. The IFT was calculated from the measured length of the oil drop using the modified
Vonnegut equation provided by the manufacturer as shown below [Vonnegut, 1942].
(1)
ATTENSION Drop Shape Analyzer
The ATTENSION drop shape analyzer can measure IFT in the range of 0.1-1000 mN/m and is not suitable
for ultra-low IFT. Hence, it was used only for heavy oil IFT measurements. The heavy oil sample was
injected into a hooked needle immersed in a cuvette containing the aqueous phase (deionized water) for
reverse pendant drop measurements. The precision of optical tensiometry depends on the quality of the
pictures and the analysis software. The ATTENSION software analyzes the images and calculates IFT of
the sample using necessary data.

Ostwald-de Waele or Power Law Model


This model is suitable for characterizing fluids that exhibit a non-Newtonian pseudoplastic behavior. A
description of the model is presented below.
(2)
where w is wall shear stress, n is flow behavior index, and Kv is viscometer consistency index.
Dividing both sides of Eq. 2 by,
(3)
But,

SPE-170800-MS

Figure 6 Effect of Emulsion Volume Fraction on Apparent Viscosity of Heavy Oil S1 at 175F

(4)
where a is apparent viscosity.
Therefore, Eq. 3 becomes:
(5)
Taking the logarithm of both sides of Eq. 5 gives:
(6)
A logarithmic plot of a versus
will result in
a straight line with (n-1) as slope, and Kv as inter 1 for the shear thinning region.
cept at

Figure 7Oil-in-Water Emulsion Droplet in a Test Tube containing


Decalin

Results and Discussion


Rheology of Heavy Oil Containing Nanoparticles
Effect of Nanoparticle Concentration on Viscosity of Heavy Oil Figures 1 and 2 show the apparent
viscosity versus wall shear rate plots for heavy oil samples S1 and S2 containing 0 (no nanoparticles), 0.1
and 0.5% wt. of 50 nm copper-oxide nanoparticles at 70 and 175F.
From Figs. 1 and 2, it can be seen that for all practical purposes the heavy oil samples exhibit a
Newtonian behavior due to constant viscosity profile at various shear rates. There is a noticeable viscosity
reduction of both samples by addition of copper-oxide nanoparticles at room temperature. This result is
in agreement with the studies reported by Hascakir et al. (2008) who observed a viscosity reduction by
addition of micron-sized iron particles to heavy oil at room temperature. The reason for viscosity
reduction is attributed to a decrease in percentage of polar compounds in the oil, weakening the hydrogen
bonding [Kershaw et al., 1980].
The viscosity reduction effect of copper-oxide nanoparticles is confirmed at higher temperatures with
0.1% wt. concentration giving the best result for both samples. This result is contrary to the findings of
Eastman et al. (1996) who studied the thermal conductivity enhancement of water dispersions of 36nm
copper-oxide nanoparticles, and found that it increases linearly with increasing particle concentration. It
can be argued that the optimum nanoparticle concentration for the heavy oil experiments is less than 0.5%
wt. concentration and further addition of nanoparticles does not necessarily improve viscosity reduction.

SPE-170800-MS

Effect of Temperature on Viscosity of Heavy Oil


containing Nanoparticles Figures 3 and 4 show
the plots of apparent viscosity versus wall shear rate
for the two heavy oil samples containing copperoxide nanoparticles (0.1% wt.) at 70 to 175F. It is
evident that both heavy oil samples follow Newtonian behavior at all temperatures and a drastic viscosity reduction is observed at higher temperatures.
This is due to the de-aggregation of long chain
asphaltenes at higher temperatures that contribute to
the viscosity of heavy oil.
Rheology of Heavy Oil Emulsions
Effect of Emulsion Volume Fraction on Viscosity
of Heavy Oil Figures 5 and 6 present apparent Figure 8 Power Law Parameters vs. Temperature for Heavy Oil S1
viscosity versus wall shear rate plots for heavy oil and S2 Emulsions
sample S1 containing emulsion in 1:1, 1:0.75, 1:0.5
and 1:0.25 oil to emulsion volume ratios at 70 and
175F respectively. The water molecules from brine
surround the heavy oil molecules after mixing and
xylene dissolves in viscous oil, reducing the viscosity of the original crude oil. Hence, the crude oil
immediately emulsifies forming low viscosity, oilin-water emulsions with the assistance of the surfactant and solvent. Xylene plays an important part
in mobilizing and dissolving the viscous oil whereas
the Triton X-100 surfactant helps in forming an
in-situ emulsification. It is assumed that the oil-inwater emulsion formed in the heavy oil reservoir
does not invert to a water-in-oil emulsion.
The fluids with higher emulsion volume fractions Figure 9 Effect of Nanoparticles Concentration on Apparent Viscosity
(1:1 and 1:0.75 oil: emulsion ratio) exhibit a New- of Heavy Oil S1 Nano-Fluids at 70F
tonian behavior and a drastic viscosity reduction is
observed even at room temperature (Fig. 5). However, at lower shear rates (102 and 170 s1) the
Newtonian values are lower than the sensitivity of the instrument at all temperatures.
From Figs. 5 and 6, it can be concluded that increasing emulsion volume fraction decreases the heavy
oil viscosity. The reason behind this phenomenon is increasing emulsion volume fraction increases the
number of water molecules surrounding crude oil and maximizes the dissolving effect of xylene leading
to a further viscosity reduction. Hence, with increasing temperature, the oil-in-water emulsions with
higher emulsion volume fractions tend to have an even lower viscosity at low and moderate shear rates
due to thermal effect. From Fig. 6, it can be seen that the viscosity of 1:1 oil to emulsion ratio fluid is low
even at high shear rates, at a temperature of 175F. Hence, addition of emulsion to heavy oil at volume
ratio greater than 1:1 can be a dilution effect.
In contrast, the 1:0.25 and 1:0.5 oil to emulsion ratio fluids exhibit non-Newtonian pseudoplastic
behavior at all temperatures. According to the studies conducted by Ilia and Abdurahman (2010), it was
found that this behavior is due to the frequency of droplet interaction and increased friction between oil
and water causing the droplets to aggregate and coalesce. Another explanation was given by Abdurahman
(2012) who stated that an increase in heavy oil volume fraction in an oil-in-water emulsion increases the

SPE-170800-MS

surfactant concentration in the aqueous continuous


phase (water). Hence, the oil-in-water emulsion behaves like a shear thinning fluid.
The droplet size distribution and oil fraction are
the main parameters that affect the viscosity and
rheological behavior of the oil-in-water-emulsions
[Ahmad et al., 1999; Nadirah et al., 2014]. Studies
conducted by Nadirah et al. (2014) on heavy oil-inwater emulsions indicated a non-Newtonian pseudoplastic profile at oil content greater than 60%
(volume). The 1:0.25 and 1:0.5 oil to emulsion ratio
fluids contain 80 and 67% oil respectively and follow the similar profile. Similar results were obtained for heavy oil S2 emulsions (Srinivasan,
2014).

Figure 10 Effect of Nanoparticles Concentration on Apparent Viscosity of Heavy Oil S1 Nano-Fluids at 175F

Confirmation Test for Oil-in-Water Emulsion In


order to confirm the nature of emulsion, a test was
carried using decalin (decahydronaphthalene) as a
solvent. A drop of the oil-in-water emulsion is injected into a test tube containing decalin. The emulsion droplet sinks to the bottom of the test tube as
seen in Fig. 7. This test confirms the emulsion to be
an oil-in-water emulsion since decalin acts as an oil
phase and does not interact with water (continuous
phase). If the droplet was a water-in-oil emulsion it
would have dispersed in the presence of decalin
since oil is the continuous phase.
Rheological Characterization of Heavy Oil EmulFigure 11Heavy Oil S1 Viscosity Reduction Comparison at 70F
sions The power law model is appropriate for the
oil-in-water emulsions investigated since a nonNewtonian pseudoplastic behavior was observed.
The points in the shear thinning region are used. The
variation of n and Kv with temperature for both the
oil-in-water emulsions is shown in Fig. 8. An increasing trend of n signifies a decrease in nonNewtonian behavior. Likewise, a decreasing trend
of Kv represents a decrease in apparent viscosity. It
can be concluded from Fig. 8 that the heavy oil S2
emulsion shows a greater non-Newtonian behavior
and apparent viscosity compared to heavy oil S1
emulsion, except at temperatures above 130 F. This
can be explained by the presence of a higher frac- Figure 12Power Law Parameters vs. Temperature for Heavy Oil S1
tion of asphaltenes and paraffins in heavy oil S2 Nano-Fluids
compared to S1, giving it a higher viscosity and Kv
value (Table 1).
At elevated temperatures, heavy oil S1 emulsion exhibits greater non-Newtonian behavior due to lower
values of n. Due to the presence of higher water content in heavy oil S2 (Table 1), the surfactant

SPE-170800-MS

Figure 13Relationship between Residual Oil Saturation and Capillary Number (Thomas, 2007)

concentration is diluted in heavy oil S2 emulsion at elevated temperatures leading to a reduction in


non-Newtonian behavior. In general, it can be seen that the non-Newtonian behavior decreases with
increasing temperature for both the emulsions, except at 150F. Intermixing or re-aggregation of long
chain hydrocarbons (like asphaltene) in heavy oil samples at 150F might be responsible for this anomaly.
Rheology of Heavy Oil Nano-Emulsions
Effect of Nanoparticle Concentration on Viscosity of Heavy Oil Nano-Fluids Figures 9 and 10 display the apparent viscosity versus wall shear rate plots for heavy oil S1 mixed with base emulsion
containing 0, 0.01, 0.02, 0.05, 0.1 and 0.5% wt. of CuO nanoparticles at 70 and 175F respectively. The
plots also show the rheology of a commercial viscosity reducer added to the heavy oil samples for
comparison with nano-emulsions. The non-Newtonian pseudoplastic behavior of the nano-fluids is
noticeable at all temperatures and nanoparticle concentrations.
The addition of nanoparticles to surfactant solutions changes their macroscopic properties and phase
behavior. The presence of nanoparticles in surfactant solutions causes interactions in which the endcaps
of the entangled wormlike micelles attach themselves to an equilibrium adsorbed surfactant layer that
covers the nanoparticles surfaces. This attachment creates micelle-particle junctions [Fakoya and Shah,
2013].
The cumulative effect of the interactions between the micelles and nanoparticles can increase base
emulsion viscosity and stability. Hence, addition of nanoparticles to base emulsion results in a favorable
mobility ratio during injection. In addition, nanoparticles help in in-situ crude oil emulsification that could
mobilize heavy oil from the reservoir and reduce fluid loss into the formation [Qiu, 2010]. As seen in Fig.
9, the addition of nanoparticles reduces the viscosity of heavy oil emulsion even at room temperature
[Hamedi et al., 2010].
According to studies conducted by Xu et al. (2013), the presence of a secondary fluid in a suspension
containing nanoparticles results in the fluid forming a hydrophobic membrane around the particle surface
and decreases the viscosity and yield stress of the suspension. Due to the presence of a hydrophobic layer,
water is repelled from the surface of these particles increasing the amount of free water.
Thus, this phenomenon might be one of the reasons for the viscosity reduction in the heavy oil-in-water
nano-emulsion at room temperature. The hydrophobic layer also prevents the formation of micellenanoparticle aggregates leading to a further viscosity reduction. A viscosity reduction can be seen at all
temperatures due to the addition of CuO nanoparticles irrespective of concentration compared to the
emulsion with no nanoparticles. Similar results were obtained for heavy oil S2 nano-fluids (Srinivasan,
2014).

SPE-170800-MS

Upon addition of the copper-oxide nanoparticles,


properties of the base emulsion such as thermal
conductivity, density, viscosity and specific heat
increase [Zhang et al., 2010]. A noticeable viscosity
reduction is observed in Figs. 9 and 10 at elevated
temperatures due to the excellent thermal conductivity properties of CuO nanoparticles. A high thermal conductivity greatly improves heat transfer
throughout the fluid enhancing viscosity reduction.
The enhancement of thermal properties of fluids
using metal particles have been investigated by various researchers [Hamilton, 1962; Masuda, 1993;
Choi, 1995; Hamedi, 2010]. It is found that the heat
Figure 14 Measured Values of Interfacial Tension with Increasing
transfer improvement is due to the high surface area Nanoparticle Concentration
of metal particles per unit volume of heavy oil. In
addition, increase in temperature corresponds to a high thermal energy. This thermal energy increases the
Brownian motion of the nanoparticles, thereby limiting the formation of micelle-nanoparticle junctions.
However, increasing nanoparticle concentration does not increase viscosity reduction as seen in Figs.
9 and 10. The viscosity reduction is maximum at 0.02% wt. concentration of CuO at all temperatures.
Hence, it can be concluded that 0.02% wt. is the optimum concentration of CuO nanoparticles for the
nano-emulsions used in this study.
From Figs. 9 and 10 it can be observed that the commercial chemical viscosity reducer added to heavy
oil follows Newtonian behavior and reduces the viscosity of both the heavy oil samples.
Comparison of Viscosity Reduction due to different Fluids Although the commercial viscosity reducer lowers heavy oil viscosity at all temperatures, the base emulsion and nano-emulsions give a much
better viscosity reduction in comparison especially at higher shear rates (Fig. 11). The heavy oil
nano-emulsion containing 0.02% wt. CuO nanoparticles gives the best viscosity reduction. Similar results
were obtained for heavy oil S2 (Srinivasan, 2014). Hence, injection of base emulsion containing 0.02%
wt. CuO nanoparticles is recommended to maximize heavy oil recovery for both the Burruss wells.
Rheological Characterization of Heavy Oil Nano-Emulsions The power law model is appropriate for
the oil-in-water nano-emulsions investigated since a non-Newtonian pseudoplastic behavior was observed. The variation of n and Kv with temperature for heavy oil S1 nano-fluids is shown in Fig. 12. The
nano-fluid sample shows a decreasing trend of Kv at all nanoparticle concentrations indicating viscosity
reduction. A decrease in non-Newtonian behavior can also be seen at all temperatures, except at 150F.
Similar results were obtained for heavy oil S2 nano-fluids (Srinivasan, 2014). The re-aggregation of the
long chain hydrocarbons present in heavy oil with nanoparticles at this temperature might be responsible
for this irregularity. An increase in temperature affects the aggregation of asphaltenes by decreasing the
aggregate size, but there is some controversy regarding the temperature effect in literature [Aske et al.,
2002]. Speight (1999) states that the precipitation of asphaltenes increases with temperature. A precipitation maximum at 80F has been reported in paraffinic solvents [Anderson and Birdi, 1990]. The
nano-fluid at 0.02% wt. concentration shows the least non-Newtonian behavior and apparent viscosity
respectively for both heavy oil samples, except at 150F for S2 nano-emulsion.
Interfacial Tension of Heavy Oil Nano-Fluids
It can be seen from Table A-2 that there is a thousand-fold decrease in the IFT of both the heavy oil
samples due to the addition of surfactant based nano-fluid. For example, the IFT of heavy oil S1 is reduced
from 19.38 to 0.018 mN/m due to addition of nano-emulsion containing 0.02% wt. CuO nanoparticles.

10

SPE-170800-MS

The capillary number is given by:


(7)
where is the Darcy velocity, is the interfacial tension and is viscosity of the displacing fluid.
Now, the viscosity of heavy oil S1 reduces to 152 cP from 602 cP (Fig. 11) which is about a four-fold
reduction. Assuming Darcy velocity to be constant, the new capillary number (Nc) can be written as:
(8)
Now with,

and

; Nc becomes

or,
Hence,
(9)
According to Thomas (2007), an increase in capillary number by three-orders of magnitude (a factor
of 1000) can reduce residual oil saturation by 50% (Fig. 13).
Assuming that the oil saturation in the reservoir is 35% before nano-fluid treatment, it decreases to
about 20% after nano-fluid injection due to increase in capillary number by a factor of 250 (Fig. 13). So
around 15% of residual oil can be recovered from the reservoir as a result of nano-fluid injection.
Therefore, the treatment of nano-emulsion to the heavy oil samples tested has the potential to significantly
reduce the residual oil saturation.
Effect of Nanoparticle Concentration on IFT Table A-2 shows that 0.02% wt. nanoparticle concentration provides the least IFT for both the heavy oil emulsion samples. The results of interfacial tension
for the heavy oil emulsions are plotted in Fig. 14.
The 0.02% wt. nanofluid exhibits lower IFT value than 0.1% wt. which is consistent with the
hypothesis of Vafaei et al. (2009) that the maximum reduction in IFT occurs at an optimum nanoparticle
concentration.
As the concentration of nanoparticles in the fluid increases, more nanoparticles are initially driven to
the liquidliquid interface and the surface tension continues to decrease due to electrostatic repulsion and
the lower surface energy of the effective interface containing nanoparticlewater, nanoparticle oil and
oilwater surfaces versus the original oilwater interface. A saturation is reached when surface tension
cannot further decrease with an increase in nanoparticle concentration and the opposite trend starts to
develop.
Hendraningrat et al. (2013) investigated the effect of nanoparticle concentration up to 0.05% wt. in
nanofluids on interfacial tension with brine as the base fluid. The increasing nanoparticle concentration
showed a decrease in the interfacial tension. Thus, nanoparticle concentration can either increase or
decrease IFT depending on when the optimum concentration is reached.

Development of Correlations
Relative viscosity (r) is the ratio of suspension viscosity (s) to solution viscosity (o) at the same shear
rate and temperature. Several authors have used the relative viscosity concept to analyze the rheological
behavior of Newtonian and non-Newtonian fluids containing particles.
The nano-fluids studied display a decrease in s; and in addition, they show different values of s at
different temperature and shear rate. For ease of prediction of s, two empirical r correlations are
developed.
(10)

SPE-170800-MS

11

(11)
where T temperature (F), solid volume fraction, and shear rate (sec1).
The correlations are valid for 0.000016 0.0008, and 102 511 sec1. Also, r 1 at
0.
Equation 10 is applicable for heavy oil nano-fluids in the temperature range of 70 T 130F, and
where A 1.7289, B 0.7398, C 55.64, D 0.3432, E 0.3086, F 1.0225 and G 0.09551.
Equation 11 is also applicable for heavy oil nano-fluids but in the temperature range of 130 T
175F, and where A 0.01317, B 70.31, C 0.00895, D 0.00142 and E 4.9094.
Example Calculations using Relative Viscosity Correlations The prediction of s from either Eq. 10
or 11 involves two simple calculation steps as exemplified below.
Estimate the suspension viscosity of a heavy oil nano-fluid at 150 F and 340 s1 when a base emulsion
containing 50 nm copper-oxide nanoparticles at a solid volume fraction of 0.00008 is added to heavy oil.
Base fluid viscosity at this temperature and shear rate conditions is 64 cP.
1. Calculate r from Eq. 11 and with o 64 cP. r exp[(0.01317 64) (70.31 0.00008)
(0.00895 150) (0.00142 340) 4.9094]/64
r 0.795
2. Compute s from r.
s r o 0.795..64 51 cP
Thus, the viscosity of heavy oil reduced down to 51 cP from 64 cP. The results at this , and shear rate
conditions but different temperatures are summarized in Table A-3. Experimental results are included for
comparison purpose. The predicted values of agree with experimental results with an absolute percentage
deviation of 2.85%.

Conclusions
y The addition of 50 nm CuO nanoparticles to the heavy oil samples reduces their viscosity at both
ambient and elevated temperatures.
y The addition of base emulsion to the heavy oil samples led to a drastic viscosity reduction at room
temperature due to the formation of oil-in-water emulsion. This viscosity reduction increased with
increasing base emulsion volume fraction.
y The addition of nanoparticles reduces the viscosity of heavy oil-in-water emulsion at all temperatures and concentrations. The optimum CuO nanoparticle concentration found was 0.02% wt.
y The base emulsion and nano-emulsion exhibited much better viscosity reduction compared to the
commercial viscosity reducer, especially at higher shear rates.
y IFT tests showed excellent results with the nano-emulsion containing 0.02% wt. CuO nanoparticles
giving the best IFT reduction.

Acknowledgment
The principal author would like to thank the Well Construction Technology Center, The University of
Oklahoma for providing the financial assistance. Authors are grateful to Dr. Patrick Shuler of ChemEOR
for providing the commercial viscosity reducer sample.
Nomenclature
D Diameter of stabilized oil drop, mm
Kv Viscometer consistency index, lbf secn/ft2
n Flow behavior index, dimensionless

12

SPE-170800-MS

Nc Capillary Number, dimensionless


T Temperature, F
Darcy velocity, m/s
Greek Symbols
o Solution viscosity, cP
r Relative viscosity, dimensionless
s Suspension viscosity, cP
Solid volume fraction, dimensionless
w Wall shear stress, lbf/ft2
Fluid viscosity, cP
a Apparent fluid viscosity, cP
IFT in Vonnegut equation, mN/m
Shear rate, sec1
Wall shear rate, sec1
w
IFT, mN/m
rotational speed of spinning drop apparatus, rpm
Density difference between oil and water, g/ml

References
Abdurahman, N.H., Rosli, Y.M., Azhari, N.H., and Hayder, B.A. (2012), Pipeline transportation of
viscous crudes as concentrated oil-in-water emulsions. Journal of Petroleum Science and Engineering,
90 91: 139 144.
Anderson, S.I., and Birdi, K.S. (1990), Influence of Temperature and Solvent on the Precipitation of
Asphaltenes, Fuel Science Technology International, 8: pp. 593615.
Aske, N., Kallevik, H., Johnsen, E., and Sjoblom, J. (2002), Asphaltene Aggregation from Crude Oils
and Model Systems Studied by High-Pressure NIR Spectroscopy, Energy & Fuels, 16: pp. 12871295.
Birkeland, M., (2013), Investigation of Nanoparticle Effect on Wettability and Interfacial tension: An
Experimental Study of a Two-Phase System of Heavy Oil and Di Water, Norwegian University of Science
and Technology, Trondheim, Norway.
Bryan, J. & Kantzas, A. (2007), Enhanced Heavy-Oil Recovery by Alkali-Surfactant Flooding. Paper
SPE 110738-MS presented at the SPE Annual Technical Conference and Exhibition, Ananheim, California, USA, 11-14 November.
Bryan, J., Mai, A. & Kantzas, A. (2008), Investigation into the Processes Responsible for Heavy Oil
Recovery by Alkali-Surfactant Flooding, paper presented at the SPE/DOE Symposium on Improved Oil
Recovery, Tulsa, Oklahoma, USA, 20-23 April.
Eastman, J. A., Choi, S. U. S., Li, S., Thompson, L. J., and Lee, S. (1996), Enhanced thermal
conductivity through the development of nanofluids, Fall Meeting of the Materials Research Society
(MRS), Boston, USA.
Evdokimov, I.N., Eliseev, N.Y., Losev, A.P., and Novikov, M.A. (2006), Emerging PetroleumOriented Nanotechnologies for Reservoir Engineering, paper presented at the SPE Russian Oil and Gas
Technical Conference and Exhibition, Moscow, Russia, 3-6 October.
Fakoya, M.F., and Shah, S.N. (2013), Rheological Properties of Surfactant Based and Polymeric
Nano-Fluids, paper presented at the SPE Coiled Tubing and Well Intervention Conference and Exhibition,
Woodlands, Texas, USA, 26-27 March.

SPE-170800-MS

13

Hascakir, B., Babadagli, T. and Akin, S. (2008), Experimental and numerical modeling of heavy-oil
recovery by electrical heating, paper presented at the International Thermal Operations and Heavy Oil
Symposium, Calgary Canada, October.
Hendraningrat, L., Li, S., and Torsaeter, O. (2013), A coreflood investigation of nanofluid enhanced
oil recovery in low-medium permeability berea sandstone, paper prepared for the SPE International
Symposium on Oilfield Chemistry, Woodlands, Texas, USA, 8-10 April.
Ilia, A.A.N. and Abdurahman, H.N. (2010), Effect of Viscosity and Droplet Diameter on water-in-oil
(w/o) Emulsions: An Experimental Study, World Academy of Science, Engineering and Technology,
38:691694
Kershaw, J.R., Barass, G., and Gray, D. (1980), Chemical Nature of Coal Hydrogenation Oils part I,
The Effect of Catalysis. Fuel Processing Technology, 32: pp. 115129.
Kong, X. & Ohadi, M., (2010), Application of micro and nano technologies in the oil and gas industry
an overview of the recent progress, paper presented at the Abu Dhabi International Petroleum Exhibition
and Conference, Abu Dhabi, UAE.
ai, A., Bryan, J., Goodarzi, N. & Kantzas, A. (2009), Insights into Non-Thermal Recovery of Heavy
Oil, Journal of Canadian Petroleum Technology, 48(3), pp. 2735.
McElfresh, P., Olguin, C., and Ector, D. (2012), The Application of Nanoparticle Dispersions to
Remove Paraffin and Polymer Filter Cake Damage, paper presented at the SPE International Symposium
and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA, 1517 February.
McElfresh, P., Holcomb, D., and Ector, D (2012), The Application of Nanofluid Technology to
Improve Recovery in Oil and Gas Wells, paper presented at the SPE International Oilfield Nanotechnology Conference, Noordwijk, The Netherlands, 1214 June.
Nadirah, L.M.S., Abdurahman, H.N., and Rizauddin, D. (2014), Rheological Study of Petroleum Fluid
and Oil in Water Emulsion, International Journal of Engineering Sciences and Research Technology,
3(1), 1 January.
Qiu, F. (2010), The Potential Applications in Heavy Oil EOR With the Nanoparticle and Surfactant
Stabilized Solvent-Based Emulsion, paper prepared for presentation at the Canadian Unconventional
Resources & International Petroleum Conference held in Calgary, Alberta, Canada, 19 21 October.
Shokrlu, Y.H., and Babadagli, T. (2010), Effects of Nano Sized Metals on Viscosity Reduction of
Heavy Oil/Bitumen during Thermal Applications, paper presented at the Canadian Unconventional
Resources & International Petroleum Conference, Calgary, Alberta, Canada, 19-21 October.
Shuler P., Tang Y., and Tang H. (2010), Heavy Oil Production Enhancement by Viscosity Reduction,
paper presented at the Western North America Regional Meeting held in Anaheim, California, USA.
Speight, J.G. (1999), The Chemistry and Technology of Petroleum. 3rd ed. New York: Marcel Dekkar.
Srinivasan, A. (2014), Surfactant-Based Fluids Containing Copper-Oxide Nanoparticles for Heavy Oil
Viscosity Reduction. Unpublished masters thesis, University of Oklahoma, Norman, USA.
Vafaei, S., Purkayastha, A., Jain, A., Ramanath, G., and Borca-Tasciuc, T. (2009), The Effect of
Nanoparticles on Liquid-Gas Surface Tension of Bi2Te3 Nanofluids, Nanotechnology, vol. 20, pp. 6 12.
Vonnegut, B. (1942), Rotating Bubble Method for the Determination of Surface and Interfacial
Tension, Review of Scientific Instruments, vol. 13, pp. 6 9.
Xu, M., Liu, H., Zhao, H., and Li, W. (2013), How to Decrease the Viscosity of Suspension with
Second Fluid and Nanoparticles, Nature, Scientific Reports, Article number: 3137.
Zhang, T., Davidson, A., Bryant, S.L., and Huh, C. (2010), Nanoparticle-Stabilized Emulsions for
Applications in Enhanced Oil Recovery, paper presented at the SPE Improved Oil Recovery Symposium,
Tulsa, Oklahoma, USA, 24-28 April.

14

SPE-170800-MS

Appendix

Table A-1Test Matrix


Copper-Oxide Nanoparticles Concentration, %wt.
Fluid System

0.01

0.02

0.05

0.1

0.5

Heavy Oil
1:1 Heavy Oil to Emulsion Ratio
1:0.75 Heavy Oil to Emulsion Ratio
1:0.5 Heavy Oil to Emulsion Ratio
1:0.25 Heavy Oil to Emulsion Ratio
Heavy Oil with Industrial Viscosity Reducer

Table A-2Interfacial Tension Results for Heavy oil and Heavy oil Nano-Fluids
Properties
Sample Type
Heavy
Heavy
Heavy
Heavy
Heavy
Heavy
Heavy
Heavy

Oil
Oil
Oil
Oil
Oil
Oil
Oil
Oil

S1
S1
S1
S1
S2
S2
S2
S2

Specific Gravity

Instrument Reading (mm)

Time (min)

IFT (mN/m)

0.92
0.86
0.86
0.861
0.96
0.88
0.88
0.881

0.416
0.381
0.390

0.452
0.411
0.430

40
40
35

40
40
30

19.38
0.024
0.018
0.02
21.76
0.026
0.02
0.023

Nano-Emulsion (0% wt.)


Nano-Emulsion (0.02% wt.)
Nano-Emulsion (0.1% wt.)
Nano-Emulsion (0% wt.)
Nano-Emulsion (0.02% wt.)
Nano-Emulsion (0.1% wt.)

Table A-3Summary of Results at 0.00008, and 340 s1


Temperature, F

r calculated

70
100
150
175
Average % Deviation

0.403
0.659
0.795
0.792

o (cP)

s (cP) calculated

s (cP) experimental

Deviation (%)

604
244
64
35

243.66
161.88
51.00
27.84

246.62
172.53
51.61
27.07

1.2
6.17
1.18
2.84
2.85

Vous aimerez peut-être aussi