You are on page 1of 23

Special Collection: III Workshop on Adsorption, Catalysis and Porous Materials

Adsorption Science & Technology


Viscosity reduction of extra 0(0) 1–23
! The Author(s) 2017
heavy crude oil by magnetite DOI: 10.1177/0263617417704309
journals.sagepub.com/home/adt

nanoparticle-based
ferrofluids
Juan E Aristizábal-Fontal, Farid B Cortés and
Camilo A Franco
Grupo de Investigación Fenómenos de superficie – Michael Polanyi, Facultad de Minas, Universidad
Nacional de Colombia—Sede Medellı́n, Colombia

Abstract
The main objective of this work is to synthesize and evaluate magnetite (Fe3O4) nanoparticle-
based ferrofluids for reducing the viscosity of an extra heavy crude oil. The carrier fluid of the
nanoparticles was synthesized using an engine lubricant recycled from the automotive industry
and hexadecyltrimethylammonium bromide as a surfactant. Fe3O4 nanoparticles were synthesized
by coprecipitation method. The effect of the concentration of nanoparticles in the viscosity
reduction degree was determined for dosages between 0 and 50,000 mg/L. Different dosages
of carrier fluid were evaluated between 0 and 10% v/v. The effects of the amount of brine
emulsified, temperature, time, and shear rate were assessed. Overall, the results showed that
viscosity and shear stress of extra heavy crude oil could be reduced up to 81 and 78% in the
presence of ferrofluid, respectively. The rheological behavior of extra heavy crude oil in the
presence and absence of ferrofluid was assessed by Cross, Ostwald-de Waele, and Herschel-
Bulkley models.

Keywords
Extra heavy oil, ferrofluid, viscosity, nanoparticles, Fe3O4, adsorption

Submission date: 6 October 2016; Acceptance date: 10 March 2017

Introduction
The increase in global demand for petroleum fuels and the decrease in supply from
conventional reservoirs has stimulated in the global oil industry the advancement in
technologies for the exploitation of reservoirs of heavy crude oil (HO) and extra heavy

Corresponding author:
Camilo A Franco, Grupo de Investigación Fenómenos de superficie – Michael Polanyi, Facultad de Minas, Universidad
Nacional de Colombia—Sede Medellı́n, Cr 80 # 65-223 M3-100(1), Antioquia 050041216, Colombia.
Email: caafrancoar@unal.edu.co

Creative Commons CC-BY: This article is distributed under the terms of the Creative Commons
Attribution 4.0 License (http://www.creativecommons.org/licenses/by/4.0/) which permits any use,
reproduction and distribution of the work without further permission provided the original work is attributed as specified
on the SAGE and Open Access pages (https://us.sagepub.com/en-us/nam/open-access-at-sage).
2 Adsorption Science & Technology 0(0)

crude oil (EHO) (Ali, 1974). HOs and EHOs are those with API gravities less than 20 and 10 ,
respectively, and represent about 70% of total petroleum resources in worldwide (Alboudwarej
et al., 2006). Viscosity is the property that most affects production and recovery operations with
values between 20 cP and more than 1  106 cP (Alboudwarej et al., 2006; Curtis et al., 2003),
which complicates its production and pipeline transportation, due to its poor fluidity and
high pressure differentials generated. Among the different compounds of crude oil,
asphaltenes are usually most responsible for the high viscosities in HO and EHO.
Asphaltenes can be defined as the most polar fraction of crude oil that are insoluble in
alkanes (such as n-C5, n-C6, and n-C7) and soluble in aromatics (such as toluene, benzene,
and xylene) (Adams, 2014; Luo and Gu, 2005; Mullins et al., 2012). The location of the
heteroatoms such as O, S, N inside asphaltene molecule grants a highly polar characteristic
that provides self-associative properties (Franco et al., 2015a), which lead to the formation of
large aggregates whose average size increases as asphaltene concentration increases to form a
viscoelastic network and therefore dramatically increase the viscosity (Akbarzadeh et al., 2007).
In searching for economically viable options to locate these crude oils in viscosity
conditions required for its production and subsequent transport, and thus meet the
production volumes projected by the market, different methods to reduce the viscosity of
this kind of crudes have been used. These methods include practices at subsurface
conditions such as thermal processes in enhanced oil recovery (EOR) and ‘‘cold’’ processes
including injection of solvents and light crudes (Islam et al., 1992; McGuire et al., 2016; Picha,
2007), in situ deasphalting (Jamaluddin et al., 1996; Mokrys and Butler, 1993), and
nanoparticle-based treatments (Taborda et al., 2016; Zabala et al., 2016). At surface
conditions other techniques have been proposed like heating (Davletbaev et al., 2014),
dilution (Alomair and Almusallam, 2013; Garcı́a Zapata and De Klerk, 2014), oil in
water emulsions (Alfonso and Drubey, 2008; Md. Saaid et al., 2014; Nguyen and Balsamo,
2013; Sharma et al., 1998), partial upgrading (Colyar, 2009; Luhning et al., 2002; Motaghi
et al., 2010), annular flow (Bobok et al., 1996; Ghosh et al., 2009; Oliemans et al., 1987; Poesio
and Strazza, 2007), and electromagnetic fields induction (Tao and Xu, 2006).
However, there is a need to develop new and cost-effective technologies because some of
the previously mentioned methods do not have good perdurability, they are expensive,
or environmentally harmful (Escojido et al., 1991; Jamaluddin et al., 1996). In this order,
and due to its exceptional properties, nanoparticles have taken particular interest in the oil
industry in stages such as exploration, drilling, EOR methods, refinery processes, water
treatment, and crude oil transportation, among others (Ali et al., 2015; Bennetzen and
Mogensen, 2014; Betancur et al., 2014; Franco et al., 2013a, 2014, 2016; Giraldo et al.,
2013a; Guo et al., 2015; Hashemi et al., 2014; Riaza et al., 2014). Nanoparticles have
different morphologies such as spherical, tubular, or irregular shape and are characterized
by having a small size between 1 and 100 nm with a high ratio of surface area/volume, high
dispersibility, and selectivity to capture high molecular weight compounds such as crude oil
asphaltenes (Franco et al., 2013b; Lövestam et al., 2010). Additionally, these materials can
be constituted by a given material enabling them to obtain specific physical qualities and
functionalized on its surface according to the needs (Franco et al., 2015b).
Nassar et al. (2015) employed commercial nanoparticles of SiO2, Fe3O4, and Al2O3 for the
asphaltenes adsorption and reduction of the mean aggregate size. In general, the authors
found that the asphaltene adsorption process is quick and that the nanoparticles decrease
the average size of asphaltenes aggregates in different degrees depending on their chemical
nature. Betancur et al. (2016) synthesized magnetite nanoparticles for adsorption of
Aristizábal-Fontal et al. 3

asphaltenes extracted from an EHO. The authors obtained that nanoparticles were capable
of adsorbing up to 61 mg/g of asphaltenes and reduce up to 39% average aggregate size
because the interaction of asphaltene–nanoparticle becomes greater than that of asphaltene–
asphaltene. Therefore, it can be expected that as the nanoparticles are included in the
asphaltenes aggregation system the average aggregate size of this decreases, inhibiting
the formation of the viscoelastic network for reducing the crude oil viscosity.
In this sense, only Zabala et al. (2016) reported the use of oil-based nanofluids for
viscosity reduction in ‘‘cold’’ processes by static and dynamic laboratory tests and
subsequent application to field conditions for HOs and EHOs. The authors report that
the nanoparticles with more adsorptive capacity toward asphaltenes are more efficient in
the viscosity reduction of crude oils evaluated with percentages of 99% for low nanofluid
fractions of 3% v/v. Furthermore, by analysis of relative permeability curves, the authors
find that the nanofluid used besides reducing viscosity can change the wettable condition of
the porous medium to a water-wettable state. In field applications, including nanofluids
resulted in an increase of oil production rate of 280 and 310 bbl in the case of heavy and
extra heavy oil fields, respectively.
However, until now there are no studies in the specialized literature that report the use of
ferrofluids for HOs and EHOs viscosity reduction. Ferrofluids are stable dispersions of
magnetic nanoparticles in the presence of surfactant in determined carrier fluid (Kothari
et al., 2010). In the oil and gas industry, ferrofluids have been used in various applications
such as improved flow surfactants in EOR processes (Cocuzza et al., 2011), imaging of
fractures (Sengupta, 2012), corrosion inhibitors (Rahmani et al., 2014), and mobilization
of residual crude oil by viscosifying displacement fluid and reducing interfacial tension
(Matteo et al., 2012; Soares et al., 2014). One of the greatest advantages of ferrofluids is
that in the presence of an external field, the magnetic moments of each nanoparticle in the
ferrofluid will align along a preferential direction according to the applied field, keeping the
mechanical behavior of fluids and allowing to control its flow by convenience through
the manipulation of the magnetic effort (Kothari et al., 2010; Saint-Martin de Abreu
Soares, 2015). Hence, it is expected that by adding ferrofluids to HOs and EHOs, the flow
conditions of this type of unconventional hydrocarbons could be controlled and optimized at
both surface or subsurface. Additionally, due to the magnetic character of the nanoparticles,
these could be recovered, regenerated, and reused after the process is performed (Betancur
et al., 2016). Therefore, the objective of this study is to evaluate for first time the effect of the
addition of ferrofluids in EHO viscosity reduction. The nanoparticles were synthesized based
on the coprecipitation method. The nanofluids were prepared using an oil residual and anionic
surfactant and blended with the nanoparticles. Rheological tests were performed under
different conditions of time, water content, temperature, and shear rate. Ferrofluids were
prepared using magnetite nanoparticles and an engine lubricant recycled from the
automotive industry. In addition, the interaction of asphaltene–Fe3O4 nanoparticles was
studied through batch mode experiments and the adsorption isotherm was obtained. This
study is expected to generate a better overview about the application of ferrofluids in the
oil industry and to improve the mobility of HOs and EHOs at surface or reservoir conditions.

Materials
The crude oil used was an EHO with 6.4 API at 289 K and saturated, aromatic, resins, and
asphaltenes contents of 15.02, 18.62, 49.86, and 16.50 wt%, respectively. For the preparation
4 Adsorption Science & Technology 0(0)

of brine, KCl (99.5%, PanReac, Spain) and deionized water were used. Nanoparticles of
Fe3O4 were synthesized by the method of coprecipitation (Lu et al., 2007) using HCl (37%,
Emsure, Germany), NH3 (25%, Emsure, Germany), and salts of FeCl2 (99.9%, PanReac,
Spain) and FeCl3 (99.9%, PanReac, Spain). Magnetite nanoparticles obtained have a
spherical morphology and average particle size, surface area, and Curie temperature of
40 nm, 66 m2/g, and 805 K, respectively. Details of the characterization and method of
synthesis of Fe3O4 nanoparticles can be found in a previous study (Betancur et al., 2016).
Engine oil recovered from a local lubrication center was used, which was recycled, and
together with hexadecyltrimethylammonium bromide (CTAB) (99.9%, PanReac, Spain)
integrate the carrier fluid of ferrofluid. n-heptane (99%, Sigma Aldrich, USA) and toluene
(99.5%, Merck GaG, Germany) were used for evaluating the interaction between
asphaltenes and the Fe3O4 nanoparticles.

Methodology
Evaluation of the Fe3O4 nanoparticles–asphaltene interaction
The interaction between the synthesized nanoparticles and the asphaltenes was evaluated
through adsorption isotherms and measurements of the mean particle size of asphaltenes
aggregates. For this, batch adsorption experiments and dynamic light scattering
measurements were made, respectively. n-C7 asphaltenes were isolated from crude oil using
n-heptane as precipitant and following a standard procedure described in previous works
(Franco et al., 2013b). Batch adsorption experiments were conducted at 298 K according to
the method described by Guzmán et al. (2016) using an Genesys 10S UVVis
spectrophotometer (Thermo Scientific, Waltham, MA) for a solution volume-to-adsorbent
mass ratio of 10 g/L and for different initial asphaltenes concentrations (Ci ) in toluene
between 0 and 4000 mg/L. The amount adsorbed (Nads ) in m2/g is plotted against the
asphaltene concentration at the equilibrium (CE ).
Also, DLS measurements are performed for determined changes in the mean aggregate
size of asphaltenes in the presence of the Fe3O4 nanoparticles. DLS experiments were
conducted for a fixed concentration of 1000 mg/L of asphaltenes in a solution of 60% v/v
of toluene and 40% v/v of precipitant (n-heptane) using a NanoPlus-3 particle analyzer from
Particulate Systems at 298 K (Nassar et al., 2015). A dosage of 10 g/L of nanoparticles was
employed and measurements were taken by triplicate.

Ferrofluid preparation
The engine lubricating oil recycled was centrifuged and filtered using a filter paper with
pore size 0.45 mm to remove any solid particle contaminants remaining in the fluid from
their use. CTAB was utilized in a proportion of 5% v/v as a surfactant to prevent
aggregation of nanoparticles (Garcı́a-Cerda et al., 2003; Yuvarajan and Ramanan, 2016).
CTAB was homogenized in engine oil recovered by magnetic stirring at room
temperature and 600 r/min. Following this, a desired amount of magnetite nanoparticles
was added to maintain a concentration of 1500 mg/L. The carrier fluid (CTAB þ recycled
engine oil) and ferrofluid were characterized in viscosity () and density (r). The viscosity
was determined using a Kinexus – Pro þ rheometer (Malvern, UK) and density by
gravimetry using a 50 mL pycnometer. Table 1 shows the properties of the carrier fluid
and the ferrofluid.
Aristizábal-Fontal et al. 5

Table 1. Viscosity and density of recycled engine oil, carrier fluid, and
ferrofluid at 298K.

Fluid m (cP) r (g/mL)

Recycled engine oil 200.1 0.8832


Carrier fluid 251.1 0.8996
Ferrofluid (1500 mg/L 260.7 0.9082
Fe3O4 nanoparticles)

Emulsions preparation
The emulsions of water in crude oil (W/O) were prepared with EHO and brine at a
concentration of 2 wt% of KCl in deionized water. Brine was added to the crude oil
under constant stirring for 60 min at room temperature and 32,000 r/min to reach
concentrations of 10 and 20% regarding the total volume. Drop size of the prepared W/O
emulsions was determined by optical microscopy using a microscope RPL3B (Microscopes
INDIA, India). The drop size distribution of dispersed brine in the continuous phase of
crude oil was obtained using the tpsDig software (Rohlf, 2006) and Microsoft Excel.

Viscosity and rheology measurements


Viscosity measurements () and rheology were performed in a Kinexus – Pro þ rheometer
using a GAP of 0.3 mm with serrated plate–plate geometry optimal for high viscous material
measures. Measurements were made at different shear rates in a range from 0 to 100 s1 for
EHO with various dosages of carrier fluid, nanoparticles between 0 and 50,000 mg/L, and
emulsified brine from 0 to 20% v/v. Additionally, measurements were obtained in a
temperature range between 298 and 318K. The viscosity of the samples was recorded for
six days at a constant shear of 1 s1. The viscosity reduction degree (VRD) was determined
according to equation (1) (Hasan et al., 2010)
ðref  treat Þ
VRD% ¼  100 ð1Þ
ref

where ref and treat are the reference viscosity and viscosity after the inclusion of
certain treatment.

Modeling
Arrhenius equation
Arrhenius equation (1887) is a mathematical expression that can be used to model the change
in viscosity as a function of temperature with high accuracy. A generalization supported by the
equation is that for the reagents transformed into products, it is necessary to acquire a
minimum amount of energy first called activation energy at a given temperature value. In
Equation (2) Arrhenius model proposed is presented
Ea
 ¼ AeRT ð2Þ
6 Adsorption Science & Technology 0(0)

Table 2. Ostwald-de Waele, Herschel-Bulkley, and Cross rheological models (Nik et al., 2005) and their
respective variables.

Model Parameters Equations

Ostwald-de Waele (kP): Viscosity  ¼ Kð ðn1Þ Þ (3)


K(kP/s): Consistency index
 (s1): Shear rate
n: Flow behavior index
Herschel-Bulkley (kP): Viscosity  ¼ KH ð ðnH 1Þ Þ þ 1, (4)
KH (kP/s): Consistency index
 (s1): Shear rate
nH : Flow behavior index
1, (kP): Viscosity at infinite shear rate
0, 1,
Cross (kP): Viscosity  ¼ 1, þ 1þðc Þm (5)
c (s): Characteristic relaxation time
 (s1): Shear rate
m: Constant
1, (kP): Viscosity at infinite shear rate
0, (kP): Viscosity at zero shear rate

where  in kiloPoise (kP) is the viscosity at a certain temperature T (K), A (kP) is the
Arrhenius constant or frequency factor, and Ea (kJ/mol) is activation energy that
indicates the sensitivity showed at the given temperature.

Rheological models
Due to the importance of non-Newtonian fluid in engineering, there are empirical models
that have allowed to describe their behavior (Sarpkaya, 1961; Shao and Lo, 2003), among
which are the Ostwald-de Waele, Cross, and Herschel-Bulkley models (Nik et al., 2005).
Models were used for modeling the rheological behavior of EHO in presence and absence of
treatment with ferrofluid or its individual components. Table 2 shows the expressions for
each model together with their respective variables.
To determine which model described better the experimental behavior obtained the root-mean-
square error definition (RMSE%) was used, determined by equation (6) (Giraldo et al., 2013b)
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
#ffi
u " k 
u1 X Cexp  Ccal 2
RSME% ¼ 100t ð6Þ
k 1 Ccal

where Cexp is the value obtained experimentally, Ccal is the value calculated by model, and k
is the number of experimental points.

Results
In the following sections, the results of the evaluation of the asphaltene–nanoparticle
interaction as well as viscosity and rheology tests for EHO evaluated in the presence and
absence of nanoparticles, ferrofluid, and water emulsified are presented.
Aristizábal-Fontal et al. 7

Figure 1. Adsorption isotherms of n-C7 asphaltene over the Fe3O4 nanoparticles surface at a fixed
temperature of 298 K.

Adsorption isotherm of n-C7 asphaltenes and evaluation of mean aggregate size


The interaction between nanoparticles and the EHO asphaltenes is a primary parameter for
understanding the driven mechanisms of viscosity reduction. Figure 1 shows the adsorption
isotherm of n-C7 asphaltene over the Fe3O4 nanoparticles surface at a fixed temperature of
298 K. From Figure 1 it can be observed that adsorption isotherm follows a Type Ib
behavior according to the most recent IUPAC classification (Thommes et al., 2015).
These systems are characterized for representing high affinity in the adsorbate–adsorbent
couple, mainly in Henry’s region at low values of CE . It can also be observed that
nanoparticles can adsorb up to 1.13 m2/g of n-C7 asphaltenes, indicating that the number
of individual asphaltene molecules in the bulk solution is reduced and hence, their
interaction with other asphaltenes is limited. This affirmation is corroborated by the DLS
measurements, where it was observed that after 24 h when the system can be considered at
equilibrium (Nassar et al., 2015), the mean aggregate size of asphaltenes change from
1017  25 nm for the system in the absence of nanoparticles to 601  25 nm after
nanoparticles inclusion. Results are in agreement with those reported by Betancur et al.
(2016) and Nassar et al. (2015), who studied the effect of nanoparticles in the kinetics of
asphaltene aggregation/fragmentation and found that the mean particle size of asphaltene
decreased drastically in the presence of the evaluated materials.

Effect of nanoparticles addition


In Figure 2 the effect of the addition of Fe3O4 nanoparticles in EHO viscosity at different
concentrations of nanoparticles between 0 and 50,000 mg/L for a fixed temperature of 298 K
and a shear rate of 1 s1 is shown. Due to the high content of asphaltenes present in this
EHO (16.5 wt%), they can self-associate, forming larger aggregates to constitute then a
viscoelastic network, which is responsible for the high viscosity up to 25 kP at evaluated
conditions (Akbarzadeh et al., 2007). Figure 2 shows that for concentrations of 500, 1000,
and 1500 mg/L of nanoparticles, the oil viscosity decreases as the dosage of the material
increases with reductions of 8, 12, and 20%, respectively. This decrease could be due to as
the number of particles increases, the adsorption of asphaltenes on the nanoparticles surface
8 Adsorption Science & Technology 0(0)

Figure 2. EHO viscosity in presence and absence of Fe3O4 nanoparticles at concentrations of 500, 1000,
1500, and 50,000 mg/L at 298 K and constant shear rate of 1 s1.
EHO: extra heavy crude oil.

is favored (Figure 1), inhibiting the formation of large aggregates and the subsequent
generation of the viscoelastic network. However, for a dosage of 50 g/L an adverse effect
on viscosity is observed, and may be because magnetite nanoparticles in high concentrations
tend to aggregate due to its magnetic characteristics, forming clusters and decreasing its
surface area per mass unit as well as its adsorptive capacity, being unable to inhibit the
tendency of the asphaltenes to aggregate each other (Nassar et al., 2015).

Effect of ferrofluid in EHO viscosity


After evaluating various concentrations of nanoparticles, a recommended dosage
of 1500 mg/L at carrier fluid concentrations between 0.15 and 10% v/v was selected.
In Figure 3 EHO viscosities are presented for a fixed dosage of 1500 mg/L of Fe3O4
nanoparticles and different dosages carrier fluid. It should be noted here that in this type
of experiments the ferrofluid (carrier fluid þ nanoparticles) is prepared before including to
the oil. Figure 3 shows that as the concentration of carrier fluid increases, the VRD increases
with values of 21, 35, 51, and up to 75% for dosages of 0.15, 1, 5, and 10%, respectively, and
even reductions of one order of magnitude in the latter case are observed. This may be due
mainly for two reasons: (i) dispersion of nanoparticles in the carrier fluid and therefore an
increase in the available area for interacting with asphaltenes and (ii) the dilutive effect of the
carrier fluid. The CTAB was included in the formulation of the ferrofluid as many of
the nanoparticles applications at reservoir conditions require the preparation of the fluid
on the surface before injection. However, in the possible scenario of in situ preparation of
ferrofluid, the use of CTAB as a surfactant could be avoided due to asphaltenes would act as
dispersing agents of the nanoparticles.
In Figure 4, the individual effect of each component of ferrofluid in reducing the
viscosity of EHO for dosages of nanoparticles and carrier fluid of 1500 mg/L and 10% v/
v, respectively, is shown. It is observed from Figure 4 that the VRD increases in order Fe3O4
nanoparticles < carrier fluid < ferrofluid, with values of 20, 50, and 75%, respectively. These
results indicate that the diluent of the carrier fluid significantly reduces the viscosity of the
selected EHO. However, it is observed that the addition of ferrofluid generates a synergistic
Aristizábal-Fontal et al. 9

Figure 3. EHO viscosity in presence and absence of ferrofluid with fixed concentration of 1500 mg/L of
Fe3O4 and carrier fluid concentrations of 0.15, 1, 5, and 10% v/v at 298 K and constant shear rate of 1 s1.
EHO: extra heavy crude oil.

Figure 4. Evaluation of ferrofluid effect and its individual components in EHO viscosity at 298 K and
constant shear rate of 1 s1. Nanoparticle dosage: 1500 mg/L; carrier fluid dosage: 10% v/v.
EHO: extra heavy crude oil.

effect on the VRD by obtaining a value 5% greater than that compared with the addition of
the individual effects of the ferrofluid components. Additionally, the results show that a
similar VRD can be obtained by adding 5% v/v of ferrofluid (Figure 3) compared to 10% v/v
carrier fluid, obtaining viscosities of 12.1 and 12.3 kP, respectively.

Effect of brine content in oil viscosity reduction by addition of ferrofluid


Effect of brine addition in EHO viscosity. Two EHO samples were evaluated with brine contents
of 10 (W10) and 20% v/v (W20). In Figure 5, the change in EHO viscosity versus time by
adding 10–20% v/v brine at 298 K and constant shear rate of 1 s1 is shown. As shown in
Figure 5, EHO viscosity increases as the water content in the system increases. These results
are consistent with those obtained by Arhuoma et al. (2009), Johnsen and Rønningsen
10 Adsorption Science & Technology 0(0)

Figure 5. Effect of brine in W/O emulsions for brine contents of 10 and 20% v/v at 298 K and constant
shear rate of 1 s1.

(2003), and Kokal (2005), who evaluated the effect of the amount of water in the behavior of
the viscosity of W/O emulsions for HOs from Saudi Arabia, North Sea, and Canada,
respectively. The increase in EHO viscosity may be due to the gradual strengthening of
interfacial film produced by asphaltene rearrangement, inhibiting coalescence of the drops
of the dispersed phase. Due to the formation of this rigid film at the interface, the drops of
the dispersed phase behave as hard spheres, and the emulsion takes similar behavior to a
liquid–solid dispersion (Schorling et al., 1999). Fluid viscosity with shear thinning behavior
as presented by emulsions (see Figure 12) is dependent on the drop size of the dispersed
phase, the latter being inversely proportional to the drop size and proportional to the
interfacial tension (Otsubo and Prud’homme, 1994). Additionally, emulsions stabilization
may be in part due to the content of KCl in the brine, generating an electroviscous effect
which arises from the overlap of the layers, becoming more significant the higher the
electrolyte concentration (Tadros, 1994).
In Figure 5 it is observed the increase in viscosity of W10 and W20 samples as a function
of time. When the viscosity of the samples is measured at the moment after preparation, it
increases from 25 kP to 36 and 47 kP for W10 and W20 in comparison to anhydrous EHO,
respectively. However, on day 6 an increase in viscosity of 12–16% compared to day 0 is
observed, reaching values of 40 and 54 kP for W10 and W20 samples, respectively. This
increase in viscosity could be due to as the average drop size of the dispersed phase decreases,
the viscosity of the emulsion increases (Otsubo and Prud’homme, 1994).
Figure 6 shows the drop size of brine in the W/O emulsion at 10 and 20% v/v measured
the same day that emulsions were prepared. The drop size distributions are typical of a tight
emulsion (Kokal, 2005), where the values of the average drop size (d50) were similar with
values of 4.5  0.5 mm in both W10 and W20 samples. However, the distribution width of
W20 is smaller compared with that of W10, which along with the amount of brine present in
the emulsion explains greater viscosity values for the W20 sample (Kokal, 2005).
Figure 7 shows the evolution of d50 value for the brine drops size as a function of time for
W10 and W20 samples. From Figure 7 it can be observed that for both W/O emulsions, the
mean drop size decreases as time increases. This behavior may be due to the redistribution of
asphaltenes in crude oil being located at the brine–oil interface, strengthening the interfacial
film, and hindering the aggregation and subsequent coalescence of the dispersed
Aristizábal-Fontal et al. 11

Figure 6. Drop size of the dispersed phase in W/O emulsion and their respective distribution at day 0 for
(a), (b) W10, and (c), (d) W20.

phase (Schorling et al., 1999), which along with the unadsorbed asphaltene reorganization
produces a significant increase in the viscosity of the samples (Figure 5).

Effect of nanoparticles and ferrofluid addition in the viscosity of emulsions. In Figure 8 the viscosity of
the samples W10 in the presence and absence of Fe3O4 nanoparticles and ferrofluid is
presented at a temperature of 298 K and shear rate fixed 1 s1. Similarly, results of
viscosity and for W20 samples in the presence and absence of nanoparticles and ferrofluid
are included in Figure S1 of the Supporting Information document. It can be seen from
Figure 8(a) that the viscosity of the emulsions decreases as the concentration of
nanoparticles increases. Similar results are observed for W20 sample in Figure S1(a). The
greatest VRD is observed after six days with values of 6, 10, and 18% and 7, 11, and 19% for
W10 and W20 samples with dosages of 500, 1000, and 1500 mg/L, respectively. Additionally,
to the effect of nanoparticles in inhibiting aggregation of asphaltenes, they could interact at
the brine–oil interface by weakening the interfacial film for enhancing the coalescence of the
drops of the dispersed phase, generating an increase in the drop size of the disperse phase. In
Figures 8(b) and S1(b), it can be observed that the viscosity of W10 and W20 decreases
considerably up to one order of magnitude in the presence of ferrofluid and exhibit a VRD of
81 and 80%, respectively. This reduction could be due to the diluent effect of the carrier fluid
along with the influence of addition of nanoparticles into asphaltenes aggregation system
and positioning at the oil–brine interface.
In Figure 9 the drop size of the dispersed phase at brine percentages of 10 and 20% v/v in
emulsion with EHO at day 6 is presented. Figure 9 shows that after addition of nanoparticles
in the emulsion the drop size increases regarding the samples in the absence of nanoparticles
12 Adsorption Science & Technology 0(0)

Figure 7. Mean brine drop size in emulsion W/O 10% v/v and 20% v/v for six days.

Figure 8. Effect of nanoparticles and ferrofluid in the emulsified oil with 10% brine at 298 K and constant
shear rate 1 s1.

(Figure 7) with values of 3.4 and 3.8 mm, representing increases of 31 and 27% for W10 and
W20 samples, respectively. Also, the CTAB included in the ferrofluid formulation will locate
at the oil–water interface due to its ionic nature and could contribute to the emulsion
destabilization through changes in the electrostatic interactions between dispersed phase
droplets (Xin et al., 2013).
Aristizábal-Fontal et al. 13

Figure 9. Brine drop size and their respective distribution at day 6 for (a), (b) W10, and (c), (d) W20
samples W/O emulsions after being treated with Fe3O4 nanoparticles at a fixed concentration of 1500 mg/L.

Figure 10. Effect of shear rate on EHO viscosity at 298 K. The symbols are the experimental data, and the
lines represent the Ostwald-de Waele, Herschel-Bulkley, and Cross models.
EHO: extra heavy crude oil.

Shear rate effect


Shear rate effect in extra heavy oil viscosity in presence and absence of nanoparticles and ferrofluid. In
Figure 10 the shear rate effect on viscosity for the selected EHO at 298 K is presented
together with the rheological models evaluated. In Figure 10 the typical non-Newtonian
14 Adsorption Science & Technology 0(0)

Table 3. Rheological parameters of Cross model for EHO in the absence and presence of 1500 mg/L of
magnetite nanoparticles, 10% v/v carrier fluid, and 10% v/v of ferrofluid at 298 K.

Sample

10% v/v 10% v/v


Model Parameters EHO 1500 mg/L Carrier fluid Ferrofluid

Cross c (s  102) 3.07 3.07 3.06 2.42


m 1.61 1.62 1.61 2.63
1, (kP) 0.28 0.24 0.13 0.07
0, (kP) 22.78 18.42 11.44 5.15
RSME% 3.48 3.50 3.45 5.70

EHO: extra heavy crude oil.

Figure 11. Effect of ferrofluid and each of its components on (a) viscosity and (b) shear stress of EHO at
298 K as a function of shear rate. The symbols are the experimental data, and the solid lines represent the
Cross model.
EHO: extra heavy crude oil.

behavior of this type of fluid is observed as viscosity changes by increasing the shear rate.
Additionally, in Table 3 the rheological parameters of Ostwald-de Waele, Herschel-Bulkley,
and Cross models at 298 K are shown. The selected models describe well the experimental
data, being the Cross model that presents the best fit according to the value of RSME%.
Aristizábal-Fontal et al. 15

In Figure 11 the experimental results of (a) viscosity and (b) shear stress as a function of
shear rate for EHO before and after treatment with the ferrofluid as well as each one of
its components are shown. Results of the obtained parameters for the Cross rheological
models evaluated are listed in Table 3. Also, parameters for the Ostwald-de Waele and
Herschel-Bulkley models are shown in Table S1 of the Supporting Information document.
In all cases, the Cross model described the experimental data in a right way with
RSME% < 6%. It is observed from Figure 11 that after treatment with the individual
components of the ferrofluid, the pseudoplastic behavior prevails. However, for the
sample treated with the ferrofluid, the viscosity is lower in the entire range of shear rate
evaluated in comparison with the individual components. For a fixed value of shear rate
of 100 s1 the trend of viscosity follows the order EHO > 1500 mg/L of nanoparticles >
10% v/v carrier fluid > 10% v/v ferrofluid, with values 3.21, 2.60, 1.61, 0.50 kP,
respectively. The results confirm the synergistic effect between the nanoparticles and the
carrier fluid in reducing the viscosity of the HO and is supported by the results of the
Cross model (Table 3).
In Figure 11(b) the behavior of the shear stress is shown, and it can be observed that it
increases as shear rate increases (Hasan et al., 2010). The shear stress increases significantly
at low shear rates, reflecting the viscoelastic behavior of EHO. Hence, it can be said that
there are no significant phenomena of disaggregation of asphaltenes. At higher shear rates,
the shear stress overcomes the value necessary for propitiating the asphaltenes desegregation,
which may explain the constant values for shear rates > 30 s1 (Pal, 1996).
In Tables 3 and S1 the obtained parameters of rheological models for EHO in absence
and presence of ferrofluid and treatment with each one of its components are presented.
The flow behavior indexes (n and nH ) show if the fluid has a Newtonian behavior when
values close to 1.0 are obtained. Values < 1 are typical of pseudoplastic fluids (Nik et al.,
2005). Consistency indexes (K and KH ) are indicative of the fluid apparent viscosity. Extreme
viscosity parameters (0, and 1, ) account for the fluid behavior when subjected to stress
at shear rates of zero and infinity, respectively. The characteristic relaxation time (c )
refers to the time required for a fluid to respond to a perturbation generated by stirring,
being equal to zero for a Newtonian. Table S1 shows that after addition of treatment
with ferrofluid, as well as each one of its components, the consistency index for Ostwald-
de Waele and Herschel-Bulkley models follows the order EHO > 1500 mg/L of
nanoparticle > 10% v/v of carrier fluid > 10% v/v of ferrofluid, while flow behavior index
increases. Likewise, Table 3 shows how 0, and 1, decrease in the order magnetite
nanoparticles > carrier fluid > ferrofluid. Also, it is observed that c takes values close to 0
for the EHO treated with the synthesized ferrofluid, which explains the behavior observed in
Figure 11(a).

Shear rate effect on emulsified EHO in presence and absence of treatment. Similar results are
obtained for the emulsified samples in comparison with the crude oil in the absence of
water. In Figure S2 of the Supporting Information document, the rheological behavior of
W10 and W20 samples at 298 K both in absence and presence of treatment with 1500 mg/L
of magnetite nanoparticles and 10% v/v of ferrofluid is shown as a function of shear rate
together with the Cross model fitting. It is observed from Figure S2 that treatment with
nanoparticles effectively decreases the viscosity of the samples in the range of shear rate
evaluated. These results could lead to lower demands on energy needs at surface or well
conditions, with reductions of shear stress up to 17 and 78%, and 18 and 76% for W10 and
16 Adsorption Science & Technology 0(0)

Table 4. Parameters of Cross rheological model for W10 and W20 samples in absence and presence of
1500 mg/L magnetite nanoparticles and 10% v/v of ferrofluid at 298 K.

Sample

W10 W20

Without Without
Model Parameters treatment 1500 mg/L Ferrofluid treatment 1500 mg/L Ferrofluid

Cross c (s  102) 6.60 6.59 6.56 6.60 6.60 6.59


m 1.37 1.38 1.41 1.37 1.37 1.42
1, (kP) 0.53 0.45 0.23 0.60 0.55 0.31
0, (kP) 35.19 29.15 7.49 46.80 38.37 10.43
RSME% 3.14 3.24 0.58 2.80 3.03 1.74

Figure 12. Comparative average viscosity reduction for the best dosage of nanoparticles and ferrofluid at
298 K and a fixed shear rate value of 1 s1.

W20 in the presence of treatment of 1500 mg/L of Fe3O4 nanoparticles and 10% v/v of
ferrofluid, respectively.
In Table 4, the corresponding values to each of the parameters evaluated by Cross
rheological model for W10 and W20 in absence and presence of treatment with magnetite
nanoparticles and ferrofluid at recommended concentrations at 298 K are presented.
Information of parameters of Ostwald-de Waele and Herschel-Bulkley can be found in
Table S2 of the Supporting Information document. The Cross model has a better fit
toward the experimental data with RSME% values < 4%. K, KH , and 0, decrease in the
order emulsion untreated > emulsion with 1500 mg/L of nanoparticles > emulsion with 10%
v/v of ferrofluid. Also, Table S2 shows that n and nH are < 1.0 in all cases, indicating a
pseudoplastic behavior of the selected samples.
Figure 12 summarizes the VRD obtained for EHO, W10, and W20 samples after
treatment with Fe3O4 nanoparticles and ferrofluid, achieving considerable reductions for
the latter. Higher degrees of viscosity reduction up to 81% are obtained for the emulsified
samples by using a ferrofluid based on low-cost materials.
Aristizábal-Fontal et al. 17

Figure 13. Effect of ferrofluid and temperature in (a) viscosity and (b) shear stress of EHO at 308 and
318 K. The symbols are the experimental data, and the solid lines represent the Cross model.
EHO: extra heavy crude oil.

Temperature effect on the EHO viscosity in the presence and absence of treatment. Figure 13 shows
the rheological behavior of EHO before and after treatment with ferrofluid at a dosage of
10% v/v and for temperatures of 308–318 K. It can be seen in Figure 13(a) that for a fixed
shear rate of 100 s1, the viscosity decreases with increasing the system temperature up to
84 and 81% for EHO in absence and presence of treatment with ferrofluid at 308 K with
values of 0.94 and 0.35 kP, respectively. Same observations can be made for samples at 318 K
with viscosity reductions of 97 and 96%, respectively. It is noteworthy that the oil viscosity
in the presence of ferrofluid at a temperature of 308 K is close to that without ferrofluid at
318 K, indicating the effectiveness of treatment regarding energy saving for mobilization
of this kind of hydrocarbons. From Figure 13(b) shear stress reductions due to the effect
of temperature on EHO and ferrofluid inclusion are observed with values of 63 and 88% at
35 and 318 K, respectively.
18 Adsorption Science & Technology 0(0)

Table 5. Parameters of Cross rheological model for EHO in absence and presence of treatment with
ferrofluid at 308 and 318 K.

308 K 318 K

Model Parameters EHO 10% v/v Ferrofluid EHO 10% v/v Ferrofluid
2
Cross c (s  10 ) 2.12 1.45 1.16 0.33
m 1.63 1.98 1.80 0.85
1, (kP) 0.10 0.04 0.05 0.03
0, (kP) 3.76 1.06 0.75 0.27
RSME% 1.93 2.75 0.94 0.87

EHO: extra heavy crude oil.

Figure 14. Temperature effect on the rheology of EHO with and without ferrofluid for shear rate of 1 s1.
The symbols are the experimental data, and the solid lines represent the Arrhenius model.
EHO: extra heavy crude oil.

Parameters obtained for modeling EHO viscosity in absence and presence of


treatment with ferrofluid at 308 and 310 K for Cross models are presented in Table 5.
Also, in Table S3 of the Supporting Information document, the parameters of the
Ostwald-de Waele and Herschel-Bulkley models are listed. It is observed that parameters
K and KH for Ostwald-de Waele and Herschel-Bulkley models, respectively, decrease by
increasing system temperature. It can also be seen from Table 5 that parameter c for
Cross model decreases after ferrofluid addition and with temperature increases. c reaches
values of 1.45  102 and 0.33  102 s at 308 and 318 K, respectively, which indicates a
Newtonian-like behavior of the fluid.
Figure 14 shows the temperature effect on viscosity of crude oil for shear rate of 1 s1,
together with the Arrhenius model. Similarly, in Figure S3 of the Supporting Information
document change in viscosity for shear rates of 10 and 50 s1 is shown. Also, in Table 6 the
parameters evaluated by the Arrhenius model are presented. From Figures 14 and S3 it can
Aristizábal-Fontal et al. 19

Table 6. Parameters of Arrhenius equation for EHO viscosity in the presence and absence of ferrofluid for
temperatures of 298, 308, and 318 K.

EHO Ferrofluid

Parameters 1 s1 10 s1 50 s1 1 s1 10 s1 50 s1

A (kP) 9.27  1024 1.93  1022 2.34  1019 3.73  1022 4.60  1021 6.00  1015
Ea (kJ/mol) 139.39 131.25 111.53 126.80 120.06 82.95
RSME% 4.39 4.56 9.44 5.84 4.80 5.01

EHO: extra heavy crude oil.

be seen that at the three different values of shear rate selected, the viscosity of the EHO in
presence and absence of the ferrofluid decreases exponentially as temperature increases.
Also, Arrhenius model described the experimental data accurately, with RSME% values
lower than 10%. Increase in shear rate values leads to decreases in the activation energy
(Table 6). The same situation is observed when comparing EHO before and after treatment
at a fixed value of shear rate.
Results indicate that with the inclusion of the ferrofluid, a substantial decrease in the
thermal requirement for enhancing the transport/production of HO and EHO with a cost-
effective treatment could be reached.

Conclusions
Successfully, the synthesis and evaluation of a magnetite nanoparticle-based ferrofluid for
viscosity reduction on EHOs were made. It was observed that nanoparticles are able to
adsorb asphaltenes and reduce their mean aggregate size. Obtained adsorption isotherms
showed a Type Ib behavior according to the IUPAC, indicating the high affinity between the
couple asphaltene–nanoparticle. Ferrofluid was prepared using CTAB and engine lubricant
oil recycled from the automotive industry, giving an added value to a typical residue of the
sector. The ferrofluid was efficient in reducing viscosity, with reductions of 75, 81, and 80%
for the extra heavy oil and emulsified oil with 10 and 20% v/v of brine. Reductions in the
shear stress of 65, 78, and 76% were also observed, respectively. Cross model successfully
reproduced the experimental data with RSME% lower than 6%. The extra heavy oil and the
selected W/O emulsions showed a pseudoplastic behavior. However, when ferrofluid is
included and the temperature is increased, a Newtonian-like behavior was observed.
Additionally, it was noted that the inclusion of the nanoparticles in the carrier fluid has a
synergistic effect in the extra heavy oil viscosity reduction, leading to higher VRD than those
obtained by the addition of the individual components of the ferrofluid. Results open a wider
landscape about the use of nanotechnology in the oil and gas industry with a low-cost and
cost-effective implementation.

Acknowledgements
This work was recognized with the ‘‘Francisco Rodrı́guez-Reinoso’’ award for the best contribution by
a junior researcher during the ‘‘III Workshop on Adsorption, Catalysis and Porous Materials’’ held in
Bogotá, Colombia on 29–31 August 2016.
20 Adsorption Science & Technology 0(0)

Declaration of Conflicting Interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or
publication of this article.

Funding
The author(s) disclosed receipt of the following financial support for the research, authorship, and/or
publication of this article: The authors thank Colciencias and Universidad Nacional de Colombia—
Sede Medellin for the financial and logistical support in the realization of this investigation.

Supplemental Material
The online supplementary figures are available at http://journals.sagepub.com/doi/suppl/10.1177/
0263617417704309.

References
Adams JJ (2014) Asphaltene adsorption, a literature review. Energy and Fuels 28: 2831–2856.
Akbarzadeh K, Hammami A, Kharrat A, et al. (2007) Los asfaltenos: Problemáticos pero ricos en
potencial. Oilfield Review 19: 25.
Alboudwarej H, Felix JJ, Taylor S, et al. (2006) La importancia del petróleo pesado. Oilfield Review 18:
38–59.
Alfonso HL and Drubey YD (2008) Propiedades reológicas de emulsiones de petróleo pesado en agua.
Ingeniare. Revista chilena de ingenierı´a 16: 244–249.
Ali N, Zhang B, Zhang H, et al. (2015) Interfacially active and magnetically responsive
composite nanoparticles with raspberry like structure; synthesis and its applications for heavy crude
oil/water separation. Colloids and Surfaces A: Physicochemical and Engineering Aspects 472: 38–49.
Ali S (1974) Heavy oil recovery-principles, practicality, potential, and problems. SPE Rocky Mountain
Regional Meeting. Montana, USA: Society of Petroleum Engineers, pp.1–10.
Alomair OA and Almusallam AS (2013) Heavy crude oil viscosity reduction and the impact of
asphaltene precipitation. Energy and Fuels 27: 7267–7276.
Arhuoma M, Dong M, Yang D, et al. (2009) Determination of water-in-oil emulsion viscosity in
porous media. Industrial and Engineering Chemistry Research 48: 7092–7102.
Arrhenius S (1887) Über die Dissociation der in Wasser gelösten Stoffe. Zeitschrift für physikalische
Chemie 1: 631–648.
Bennetzen MV and Mogensen K (2014) Novel Applications of Nanoparticles for Future Enhanced Oil
Recovery. In: International Petroleum Technology Conference, Kuala Lumpur, Malaysia:
International Petroleum Technology Conference, pp.1–14.
Betancur S, Alzate GA and Cortés FB (2014) Optimization of drilling fluids using functionalized
nanoparticles: Loss filtration reduction and thickness mudcake. Boletı´n de Ciencias de la Tierra
35: 5–14.
Betancur S, Franco CA and Cortés FB (2016) Magnetite-silica nanoparticles with a core-shell structure
for inhibiting the formation damage caused by the precipitation/deposition of asphaltene. In:
Pittman SR (ed.) Asphaltenes: Fundamentals, Applications and Future Developments. New York:
Nova Science Publishers, Inc., pp.85–138.
Bobok E, Magyari D and Udvardi G (1996) Heavy oil transport through lubricated pipeline. European
Petroleum Conference. Milan, Italy: Society of Petroleum Engineers, pp.239–244.
Cocuzza M, Pirri F, Rocca V, et al. (2011) Is the oil industry ready for nanotechnologies? Offshore
Mediterranean Conference and Exhibition. Ravenna, Italy: Offshore Mediterranean Conference,
pp.1–17.
Aristizábal-Fontal et al. 21

Colyar J (2009) Has the time for partial upgrading of heavy oil and bitumen arrived? Petroleum
Technology Quarterly 14: 43–56.
Curtis C, Kopper R, Decoster E, et al. (2003) Yacimientos de petróleo pesado. Oilfield Review, II 23:
32–55.
Davletbaev A, Kovaleva L and Babadagli T (2014) Heavy oil production by electromagnetic heating in
hydraulically fractured wells. Energy and Fuels 28: 5737–5744.
Escojido D, Urribarri O and González J (1991) Part 1: Transportation of heavy crude oil and natural
bitumen. In: Thirteenth World Petroleum Congress, pp.15–22.
Franco C, Cardona L, Lopera S, et al. (2016) Heavy oil upgrading and enhanced recovery in a
continuous steam injection process assisted by nanoparticulated catalysts. In: SPE Improved Oil
Recovery Conference. Tulsa, Oklahoma, USA: Society of Petroleum Engineers, pp.1–16.
Franco CA, Lozano MM, Acevedo S, et al. (2015a) Effects of resin I on asphaltene adsorption onto
nanoparticles: A novel method for obtaining asphaltenes/resin isotherms. Energy and Fuels 30:
264–272.
Franco CA, Montoya T, Nassar NN, et al. (2013a) Adsorption and subsequent oxidation of
Colombian asphaltenes onto nickel and/or palladium oxide supported on fumed silica
nanoparticles. Energy and Fuels 27: 7336–7347.
Franco CA, Nassar NN and Cortés FB (2014) Removal of oil from oil-in-saltwater emulsions by
adsorption onto nano-alumina functionalized with petroleum vacuum residue. Journal of Colloid
and Interface Science 433: 58–67.
Franco CA, Nassar NN, Montoya T, et al. (2015b) Influence of asphaltene aggregation on the
adsorption and catalytic behavior of nanoparticles. Energy and Fuels 29: 1610–1621.
Franco CA, Nassar NN, Ruiz MA, et al. (2013b) Nanoparticles for inhibition of asphaltenes damage:
Adsorption study and displacement test on porous media. Energy and Fuels 27: 2899–2907.
Garcı́a-Cerda L, Rodrı́guez-Fernández O, Betancourt-Galindo R, et al. (2003) Sı́ntesis y propiedades
de ferrofluidos de magnetita. Superficies y Vacı´o 16: 28.
Garcı́a Zapata JL and de Klerk A (2014) Viscosity changes during mild oxidation of oilsands-derived
bitumen: Solvent effects and selectivity. Energy and Fuels 28: 6242–6248.
Ghosh S, Mandal T, Das G, et al. (2009) Review of oil water core annular flow. Renewable and
Sustainable Energy Reviews 13: 1957–1965.
Giraldo J, Benjumea P, Lopera S, et al. (2013a) Wettability alteration of sandstone cores by alumina-
based nanofluids. Energy and Fuels 27: 3659–3665.
Giraldo J, Nassar NN, Benjumea P, et al. (2013b) Modeling and prediction of asphaltene adsorption
isotherms using Polanyi’s modified theory. Energy and Fuels 27: 2908–2914.
Guo K, Li H and Yu Z (2015) Metallic nanoparticles for enhanced heavy oil recovery: Promises and
challenges. Energy Procedia 75: 2068–2073.
Guzmán JD, Betancur S, Carrasco-Marı́n F, et al. (2016) Importance of the adsorption method used
for obtaining the nanoparticle dosage for asphaltene-related treatments. Energy and Fuels 30:
2052–2059.
Hasan SW, Ghannam MT and Esmail N (2010) Heavy crude oil viscosity reduction and rheology for
pipeline transportation. Fuel 89: 1095–1100.
Hashemi R, Nassar NN and Almao PP (2014) Nanoparticle technology for heavy oil in-situ upgrading
and recovery enhancement: Opportunities and challenges. Applied Energy 133: 374–387.
Islam M, Erno B and Davis D (1992) Hot gas and waterflood equivalence of in situ combustion.
Journal of Canadian Petroleum Technology 31: 44–52.
Jamaluddin A, Nazarko T, Sills S, et al. (1996) Deasphalted oil—A natural asphaltene solvent. SPE
Production and Facilities 11: 161–165.
Johnsen EE and Rønningsen HP (2003) Viscosity of ‘live’ water-in-crude-oil emulsions:
Experimental work and validation of correlations. Journal of Petroleum Science and Engineering
38: 23–36.
Kokal SL (2005) Crude oil emulsions: A state-of-the-art review. SPE Production and facilities 20: 5–13.
22 Adsorption Science & Technology 0(0)

Kothari N, Raina B, Chandak KB, et al. (2010) Application of ferrofluids for enhanced surfactant
flooding in IOR. In: SPE EUROPEC/EAGE Annual Conference and Exhibition. Barcelona, Spain:
Society of Petroleum Engineers, pp.1–7.
Lüvestam G, Rauscher H, Roebben G, et al. (2010b) Considerations on a definition of nanomaterial for
regulatory purposes. Luxembourg: Publications Office of the European Union. ISBN 978-92-79-
16014-1. doi: 10.2788/98686.
Lu AH, Salabas EeL and Schüth F (2007) Magnetic nanoparticles: Synthesis, protection,
functionalization, and application. Angewandte Chemie International Edition 46: 1222–1244.
Luhning R, Anand A, Blackmore T, et al. (2002) Pipeline transportation of emerging partially
upgraded bitumen. In: Canadian International Petroleum Conference. Calgary, Alberta: Petroleum
Society of Canada, pp.1–18.
Luo P and Gu Y (2005) Effects of asphaltene content and solvent concentration on heavy oil viscosity.
In: SPE International Thermal Operations and Heavy Oil Symposium. Calgary, Alberta, Canada:
Society of Petroleum Engineers, pp.1–7.
McGuire PL, Okuno R, Gould TL, et al. (2016) Ethane-Based EOR: An Innovative and Profitable
EOR Opportunity for a Low Price Environment. In: SPE Improved Oil Recovery Conference. Tulsa,
Oklahoma, USA: Society of Petroleum Engineers, pp.1–25.
Matteo C, Candido P, Vera R, et al. (2012) Current and future nanotech applications in the oil
industry. American Journal of Applied Sciences 9: 784.
Md. Saaid I, Mahat SQA, Lal B, et al. (2014) Experimental investigation on the effectiveness of
1-Butyl-3-methylimidazolium perchlorate ionic liquid as a reducing agent for heavy oil
upgrading. Industrial and Engineering Chemistry Research 53: 8279–8284.
Mokrys I and Butler R (1993) In-situ upgrading of heavy oils and bitumen by propane deasphalting:
the VAPEX process. In: SPE Production Operations Symposium. Oklahoma City, Oklahoma:
Society of Petroleum Engineers, pp.409–424.
Motaghi M, Saxena P and Ravi R (2010) Partial upgrading of heavy oil reserves. Petroleum Technology
Quarterly 15: 55–64.
Mullins OC, Sabbah H, Eyssautier Jl, et al. (2012) Advances in asphaltene science and the
Yen–Mullins model. Energy and Fuels 26: 3986–4003.
Nassar NN, Betancur S, Acevedo S, et al. (2015) Development of a population balance model to
describe the influence of shear and nanoparticles on the aggregation and fragmentation of
asphaltene aggregates. Industrial and Engineering Chemistry Research 54: 8201–8211.
Nguyen D and Balsamo V (2013) Emulsification of heavy oil in aqueous solutions of poly
(vinyl alcohol): A method for reducing apparent viscosity of production fluids. Energy and Fuels
27: 1736–1747.
Nik WW, Ani F, Masjuki H, et al. (2005) Rheology of bio-edible oils according to several rheological
models and its potential as hydraulic fluid. Industrial Crops and Products 22: 249–255.
Oliemans R, Ooms G, Wu H, et al. (1987) Core-annular oil/water flow: The turbulent-lubricating-film
model and measurements in a 5 cm pipe loop. International Journal of Multiphase Flow 13: 23–31.
Otsubo Y and Prud’homme R (1994) Rheology of oil-in-water emulsions. Rheologica Acta 33: 29–37.
Pal R (1996) Effect of droplet size on the rheology of emulsions. AIChE Journal 42: 3181–3190.
Picha MS (2007) Enhanced oil recovery by hot CO2 flooding. In: SPE Middle East oil and gas show and
conference. Society of Petroleum Engineers, pp.1–5.
Poesio P and Strazza D (2007) Experiments on start-up of an oil-water core annular flow through a
horizontal or nearly horizontal pipe. In: 13th International Conference on Multiphase Production
Technology. Edinburgh, UK: BHR Group, pp.79–87.
Rahmani AR, Bryant S, Huh C, et al. (2014) Crosswell magnetic sensing of superparamagnetic
nanoparticles for subsurface applications. SPE Journal 1067–1082. DOI: https://doi.org/10.2118/
166140-PA.
Aristizábal-Fontal et al. 23

Riaza S, Cortés FB and Otalvaro J (2014) Emulsiones con crudo pesado en presencia de
nanopartı́culas. Boletı´n de Ciencias de la Tierra 36: 55–68.
Rohlf F (2006) tpsDig, version 2.10. Stony Brook: Department of Ecology and Evolution, State
University of New York.
Saint-Martin de Abreu Soares F (2015) A Pore Scale Study of Ferrofluid-Driven Mobilization of Oil.
Austin, Texas: The University of Texas at Austin.
Sarpkaya T (1961) Flow of non-Newtonian fluids in a magnetic field. AIChE Journal 7: 324–328.
Schorling P-C, Kessel D and Rahimian I (1999) Influence of the crude oil resin/asphaltene ratio on the
stability of oil/water emulsions. Colloids and Surfaces A: Physicochemical and Engineering Aspects
152: 95–102.
Sengupta S (2012) An Innovative approach to image fracture dimensions by injecting ferrofluids. In:
Abu Dhabi International Petroleum Conference and Exhibition. Abu Dhabi, UAE: Society of
Petroleum Engineers, pp.1–6.
Shao S and Lo EY (2003) Incompressible SPH method for simulating Newtonian and non-Newtonian
flows with a free surface. Advances in Water Resources 26: 787–800.
Sharma K, Saxena V, Kumar A, et al. (1998) Pipeline transportation of heavy/viscous crude oil as
water continuous emulsion in North Cambay Basin (India). In: SPE India Oil and Gas Conference
and Exhibition. New Delhi, India: Society of Petroleum Engineers, pp.301–305.
Soares FS-MdA, Prodanovic M and Huh C (2014) Excitable nanoparticles for trapped oil
mobilization. In: SPE Improved Oil Recovery Symposium. Tulsa, Oklahoma, USA: Society of
Petroleum Engineers, pp.1–16.
Taborda EA, Franco CA, Lopera SH, et al. (2016) Effect of nanoparticles/nanofluids on the rheology
of heavy crude oil and its mobility on porous media at reservoir conditions. Fuel 184: 222–232.
Tadros TF (1994) Fundamental principles of emulsion rheology and their applications. Colloids and
Surfaces A: Physicochemical and Engineering Aspects 91: 39–55.
Tao R and Xu X (2006) Reducing the viscosity of crude oil by pulsed electric or magnetic field. Energy
and Fuels 20: 2046–2051.
Thommes M, Kaneko K, Neimark AV, et al. (2015) Physisorption of gases, with special reference to
the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure and
Applied Chemistry 87: 1051–1069.
Xin X, Zhang H, Xu G, et al. (2013) Influence of CTAB and SDS on the properties of oil-in-water
nano-emulsion with paraffin and span 20/Tween 20. Colloids and Surfaces A: Physicochemical and
Engineering Aspects 418: 60–67.
Yuvarajan D and Ramanan MV (2016) Effect of magnetite ferrofluid on the performance and
emissions characteristics of diesel engine using methyl esters of mustard oil. Arabian Journal for
Science and Engineering 41: 2023–2030.
Zabala R, Franco C and Cortés F (2016) Application of nanofluids for improving oil mobility in heavy
oil and extra-heavy oil: A field test. In: SPE Improved Oil Recovery Conference. Tulsa, Oklahoma,
USA: Society of Petroleum Engineers, pp.1–14.

You might also like