Sie sind auf Seite 1von 14

SPE-179677-MS

Application of Nanofluids for Improving Oil Mobility in Heavy Oil and


Extra-Heavy Oil: A Field Test
R. Zabala, Ecopetrol; C. A. Franco, and F. B. Cortés, Universidad Nacional de Colombia - Sede Medellín

Copyright 2016, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Improved Oil Recovery Conference held in Tulsa, Oklahoma, USA, 11–13 April 2016.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
An important factor during the life of a heavy crude reservoir is the oil mobility. It depends on two factors,
oil viscosity and oil relative permeability. Two characteristics of nanoparticles that make them attractive
for assisting IOR and EOR processes are their size (1 to 100 nm) and ability to manipulate their behavior.
Due to their nano-sized structure, nanomaterials have large tunable specific surface areas that lead to an
increase in the proportion of atoms on the surface of the particle, indicating an increasing in surface
energy. Nanoparticles are also able to flow through typical reservoir pore spaces with sizes at or below
1 micron without the risk to block the pore space. Nanofluids or ⬙smart fluids⬙ can be designed by tuning
nanoparticle properties, and are prepared by adding small concentrations of nanoparticles to a liquid phase
in order to enhance or improve some of the fluid properties. However the use of nanoparticles and
nanofluids for oil mobility has been poorly studied. Hence, the scope of this work is to present the field
evaluation of nanofluids for improving oil mobility and mitigate alteration of wettability in two Colom-
bian heavy oil fields; Castilla and Chichimene. Asphaltenes sorption tests with two different types of
nanomaterials were performed for selecting the best nanoparticle for each type of oil. An oil based
nanofluid (OBN) containing these nanoparticles was evaluated as viscosity reducer under static condi-
tions. Displacement tests through a porous media in core plugs from Castilla and Chichimene at reservoir
conditions were also performed. OBN was evaluated to reduce oil viscosity varying oil temperature and
water content. Maximum change in oil viscosity is achieved at 122°F and 2% of nanofluid dosage. The
use of the nanofluid increased oil recovery in the core flooding tests, caused by the removal of asphaltenes
from the aggregation system, reduction of oil viscosity, and the effective restoration of original core
wettability. Two field trials were performed in Castilla (CNA and CNB wells), by forcing 200 bbl and 150
bbl of nanofluid respectively as main treatment within a radius of penetration of ~3 ft. Instantaneous oil
rate increases of 270 bopd in CNA and 280 bopd in CNB and BSW reductions of ~11% were observed.
In Chichimene also two trials were performed (CHA and CHB), by forcing 86 bbl of and 107 bbl of
nanofluid as main treatment within a radius of penetration of ~3 ft. Instantaneous oil rate increases of 310
bopd in CHA and 87 bopd in CHB were achieved not BSW reduction has been observed yet. Interventions
were performed few months ago and long term effects are still under evaluation. Results look promising
making possible to think extending application of nanofluid in other wells in these fields.
2 SPE-179677-MS

Introduction
An important factor during the life of a heavy crude reservoir is the oil mobility. It depends on two factors,
oil viscosity and oil relative permeability (Giraldo et al. 2013, Franco, Nassar, et al. 2013). Two
characteristics of nanoparticles that make them attractive for assisting IOR and EOR processes are their
size (diameter ranging from 1 to 100 nm) and the ability to manipulate their behavior (Zabala et al. 2014,
Franco et al. 2014, Franco et al. 2015). Due to their nano-sized structure, nanomaterials have large tunable
specific surface areas that lead to an increase in the proportion of atoms on the surface of the particle,
indicating an increasing in surface energy. Nanoparticles are also able to flow through typical reservoir
pore spaces with sizes at or below 1 micron without the risk to block the pore space (Franco et al. 2013).
Nanofluids or ⬙smart fluids⬙ can be designed by tuning nanoparticle properties, and are prepared by adding
small concentrations of nanoparticles to a liquid phase in order to enhance or improve some of the fluid
properties (Zitha 2005, Sherik and Nabulsi 2009). However the use of nanoparticles and nanofluids for oil
mobility has been poorly studied.
Hamedi et al. (2010), studied the effect of micro (Iron and copper) and nano-sized (Iron and Nickel)
metal particles for heavy oil viscosity reduction. The tests of viscosity reduction were performed by
adding 0.1, 0.5 and 1 wt% of the material to heavy oil (14.7 API) at 77, 122 and 176°F using an AR-G2
rotational rheometer. The authors found that in general the highest reduction was achieved by adding 0.1
wt% of the particles. However the percentage of viscosity reduction was not higher than 9.5% and was
achieved by adding 0.1 wt% of micro-sized copper particles. Recently some researchers have suggested
that the mobility and effectiveness of wettability modifiers such as surfactants can be increased by the
addition of nanoparticles (Giraldo et al. 2013, Franco, Nassar, et al. 2013, Karimi et al. 2012, Ju, Fan, and
Ma 2006, Ju and Fan 2009, Maghzi et al. 2011). Karimi et al. (2012) studied the effect of zirconium oxide
(ZrO2)-based nanofluids on wettability alteration of a carbonate reservoir rock. Several nanofluids were
made composed of ZrO2 nanoparticles and mixtures of nonionic surfactants with hydrophilic-lipophilic
balance (HLB) ranging from 15 to 1.8. Two different nanoparticles concentrations were tested (50000 and
100000 mg/L). The impact of wettability alteration on oil recovery was evaluated by free imbibitions tests.
The authors concluded that designed ZrO2-based nanofluids were effective wettability modifiers for
carbonate systems since they could change the rock wettability from a strongly oil-wet to a strongly
water-wet condition.
In other study, Giraldo et al. (2013) studied the effectiveness of alumina-based nanofluids in altering
the wettability of sandstone cores with an induced oil-wet wettability. They used five nanofluids with
different particles concentration ranging from 100 mg/L to 10000 mg/L, and were prepared by dispersing
the alumina nanoparticles in a commercial surfactant. By contact angle and imbibition test they showed
that designed nanofluids could alter significantly the rock wettability from a strongly oil-wet to a strongly
water-wet condition. Imbibition tests also allowed identifying the effect of nanoparticles concentration on
the suitability of the treatment for enhancing the imbibition process and restoring the original core
wettability. The best performance was achieved when they used a concentration of 100 mg/L. They also
performed a core displacement test by injecting an alumina-based nanofluid in a sand pack. The treatment
was effective in altering the sand pack wettability from an oil-wet to a strongly water-wet condition.
Franco et al.1 carried on an experimental study to analyze the effect of the chemical nature of twelve types
of nanoparticles on asphaltenes adsorption, and the delay or inhibition of the deposition an precipitation
of these asphaltenic compounds at reservoir conditions. They found that nanoparticles that favor Langmuir
type isotherms are good candidates for inhibiting the asphaltene precipitation on the rock surface. Only
the nanoparticles that adsorb strongly the more polar compounds are capable of neutralizing the polar
forces that remain active during weak adsorption to cause multi-layer adsorption. That inhibition prevents
both flocculation-precipitation, which these polar compounds seem to be the building blocks of, and
eliminate their tendency to adsorb in multi-layers which could be due to remaining polarity of the initially
SPE-179677-MS 3

adsorbed asphaltenes. The injection of nanofluids into porous media showed an inhibition in the
agglomeration, precipitation and deposition of asphaltenes on the rock surfaces, which was based on a
three-step displacement test. Additionally, the nanoparticle treatment demonstrated an enhanced perdu-
rability of the system. The nanoparticles were able to restore production and led to improvements in
recovery due to their ability to adsorb and stabilize the asphaltenes content of the system. Zabala et al.
(2014) analyzed the effect of two nanoparticles on the adsorption of asphaltenes extracted from two
Colombian heavy oils fields, Castilla and Chichimene. This was in order to select the best nanomaterial
to enhance oil based nanofluid to restore wettability of cores from oil-wet to water-wet and increase the
oil mobility trough the porous media at reservoir conditions. The use of the nanofluids increased oil
recovery in the core flooding tests, caused by asphaltenes inhibition, reduction of oil viscosity, and the
effective restoration of original core wettability.
This paper describe the recently trails in two fields in Colombia with a novel technique based in oil
based nanofluidos carrying the tailor made nanomaterials according to previous scenario, it was performed
two stimulation jobs in Castilla and two jobs in Chichimene Fields. Nanotechnology was successfully
implemented. Castilla heavy oil field is located in Colombia’s Llanos Basin 50 Km to the SE of
Villavicencio City (Fig. 1). This field was discovered by Chevron in 1969 but it has been operated by
ECOPETROL S.A since 2002. There are three operating areas called Castilla, Castilla Norte and Castilla
Este. However, geology, stratigraphy and reservoir units are the same. Structural interpretations confirm
that Castilla Field is formed by a NE-SW trending asymmetric anticline that forms a three-way dip closure
against the Castilla Fault complex to the southeast.

Figure 1—Castilla and Chichimene Field Locations.

Fig. 2 shows a) Castilla and b) Chichimene reservoir units. The main producing stratigraphic units are
the Massive Guadalupe (Fig. 2a), or K2, and Upper Guadalupe, or K1 which has been divided into two
packages: the upper (K1 Superior) and lower (K1 Inferior) members. Additionally, there is a Tertiary
stratigraphic unit, called San Fernando Formation or operationally T2, but it is currently a secondary
target. Petrophysical properties are fairly good: porosity varies from 16% to 22%, permeability ranges
4 SPE-179677-MS

from 500 to 10000 mD, initial water saturation is around 15%. API gravity is around 13° and its viscosity
is close to 150 cP. On the other hand, The Free Water Level (FWL) is tilted, saturated with fresh water
(active stronger aquifer). Chichimene heavy oil field is located near to Castilla Field. This field was
discovered by Chevron in 1969 but it has been operated by ECOPETROL S.A since 2000. Structural
interpretations confirm that Chichimene Field is formed by a NE-SW trending asymmetric anticline,
bounded by a reverse fault complex to the east and southeast. The main producing stratigraphic units are
the San Fernando Formation (Fig. 2b), or T2, and Upper Guadalupe, or K1 which has been divided into
two packages: the upper (K1 Superior) and lower (K1 Inferior) members, but are currently a secondary
target. Petrophysical properties are fairly good: porosity varies from 16% to 20%, permeability ranges
from 800 to 4000 mD, initial water saturation is around 22%. API gravity is around 8° and its viscosity
is close to 350 cP.

Figure 2—a) Castilla and b) Chichimene Reservoir Units.

Experimental Description
Materials
Crude Oils The ⬙Chichimene⬙ extra heavy oil (8 °API at 77°F, 1% water content and 16% by weight of
asphaltenes) and ¨Castilla¨ heavy oil (12 °API at 77°F, 25% water content and 8% by weight of
asphaltenes) were used to evaluate the development of the nanofluid in porous media. Also the crude oil
was used as a source of asphaltenes. In this document ⬙Chichimene⬙ and ⬙Castilla⬙ crude oils are named
⬙CHO⬙ (Chichimene oil) and ⬙CAO⬙ (Castilla Oil) respectively.
Solvents and reagents. n-heptane (99%) was used for asphaltenes extraction from crude oil and toluene
(99%) was used to dissolve the asphaltenes at the desired concentration. Nanoparticles A (NPA) and
nanoparticles B (NPB) supplied by Petroraza (Colombia) were used to perform the asphaltenes adsorption
experiments. Commercial nanoparticles C (NPC), nanoparticles D (NPD) and nanoparticles E (NPE) were
also used to perform the adsorption tests. Commercial oil based nanofluid (OBN, 15 °API; viscosity, 8.14
cp and pH 10.23 at 76.3°F) and commercial pre-flux were used to perform de displacement tests.
Methods
The asphaltene extraction protocol Asphaltenes were precipitated from CHO and CAO following a
standard procedure (Franco et al. 2013, Franco et al. 2014). The model solutions for the batch adsorption
SPE-179677-MS 5

experiments were prepared by dissolving a desired amount of the obtained asphaltenes in toluene. The
initial concentration of asphaltene solutions used in the adsorption experiments ranged from 250 to 1500
mg/L.
Surface area and particle size measurements. The surface areas (SBET) of the prepared nanoparticles
were estimated following the Brunauer-Emmett-Teller (BET) method (Rouquerol et al. 2013). The mean
crystallite size of the particles (dp: nanoparticle diameter) was obtained by applying the Scherrer equation
to the main diffraction peak.
Equilibrium adsorption isotherms. According to the procedure described in previous studiesthe ad-
sorbed amount of asphaltenes (mg of asphaltenes/g of nanoparticles) was determined by the mass balance
in Eq. 1 (Franco et al. 2013, Franco et al. 2014):
(1)

where, C0 is the initial concentration of asphaltenes in the solution (mg/L), CE is the equilibrium
concentration of asphaltenes in the supernatant (mg/L), V is the solution volume (L), and W is the amount
of nanoparticles added to the solution (g).
The Viscosity Measurements A FUNGILAG rotational viscometer SMART R was used to perform the
viscosity measurements at 86, 122 and 175°F for CHO and 86, 122 and 156°F for CAO. First, the base
curves without treatment were constructed. Then 0.5, 1, 2 and 3 wt% of nanofluid were added and the
viscosity measurements were performed.

The fluid injection test


Porous media Two cores were used to study the transport behavior of the nanofluid through the porous
media. For CHO the selected core was ⬙CHI 30 – 7920.5⬙. The porous media has an absolute permeability
of 622 mD, porosity of 21 %, diameter of 3.75 cm and length of 7.00 cm. For CAO the selected core was
⬙Castilla N0046 – 7432.5⬙. In this case, the plug has an absolute permeability of 500 mD, porosity of 11
%, diameter of 3.80 cm and length of 6.75 cm.
Preparation of injection fluids For the fluid injection test, two synthetic brines were prepared for each
field to get an accurate approximation of the reservoir conditions. The nanofluid was magnetically stirred
at 77°F for 6h and then sonicated at the same temperature for 24 h, whereby nanoparticles remained stable
in suspension. According to the results of the adsorption isotherms NPA was used in the displacement test.
Experimental set-up and procedure Fig. 3 shows a schematic representation of the experimental setup.
For CHO, all tests were carried out at a temperature of 210°F, pore pressure of 3002 psi. For CAO, the
test temperature was 188°F, and the pore pressure and the overburden pressure were 2495 and 3002 psi
respectively. For the displacement tests on the porous media, the chemical nature of the nanoparticles and
their concentration in the aqueous solution were selected based on the isotherm results. The main objective
of this displacement test was to evaluate the effectiveness of the nanoparticles by improving the oil
mobility and changing the core wettability to a water-wet medium from an oil-wet system. The
displacement tests were carried in four steps: 1) constructing the initial curves. 10 pore volumes (PV) of
water were injected to measure the absolute permeability at the temperature, pore pressure and overburden
pressure desired. Then, at the same temperature, and under the saturation condition of residual water
(Swr), the crude oil was injected until the pressure no longer changed. Finally, 20 PV of water was
injected at the test temperature to construct the relative permeability curves as functions of the water
saturation. Accordingly, the recovery curves (Np) and the effective permeability to water at the saturation
of oil residual (Sor) conditions were measured. 2) Becoming the core in an oil-wet medium and
constructing the base curves. First the core is saturated by injecting 20 PV of crude oil and aged for 10
6 SPE-179677-MS

days injecting each day 1 PV of crude oil. Then for verifying if the system is oil wettable, the base curves
are constructed following the same procedure as for the initial curves. 3) Evaluating the oil based
nanofluid on the crude oil mobility and the restoration of the wettability. 10 PV of crude oil were injected
to saturate the porous media. At that point, 0.3 PV of pre-flux were injected to put in conditions the porous
media for the nanofluid injection. Then, the core was aged with the nanofluid for 4 hours by injecting 0.3
PV of the product. After that, changes in pressure were verified to check the improvement of the oil
mobility and changes in the core wettability by injecting 20 PV of crude oil in the production sense (i.e.
closing both inlet and outlet valves, unplug the core holder and twist it horizontally to plug the input valve
with the core holder outlet and vice versa). Then, 20 PV of water was injected to measure the relative
permeability and Np curve. 4) Finally, observing the perdurability of the nanofluid treatment on the
porous media. Crude oil is injected to the porous media until the Ko is approximately equal to the base
Ko.

Figure 3—Schematic representation of the experimental setup. Legend: (1) the core holder, (2) the sand packed bed, (3) the
displacement cylinder, (4) the displacement positive pump, (5) the mineral oil pump, (6) the pressure transducers, (7) the valves, and
(8) sample output.

Laboratory Results and Discussion


Nanoparticles Characterization
Table 1 resumes the evaluated properties of the nanoparticles. Results shown that NPA nanoparticles have
bigger surface area and particle diameter than NPB. Based on the measurement of N2 adsorption, NPA
have a surface area of 223.2 m2/g and NPB of 119.1 m2/g. The dp of NPA was 35 nm and for NPB was
7 nm.
SPE-179677-MS 7

Table 1—Estimated values of nanoparticles diameter (dp) and


surface area of the selected nanoparticles.
Material dp (nm) Surface area (m2/g)

NPA 25 223
NPB 7 120
NPC 28 102
NPD 15 99
NPE 19 103

The batch adsorption test: The equilibrium isotherm of asphaltenes adsorption onto the
nanoparticles
Fig. 4 presents the maximum adsorption capacities of asphaltenes extracted from CHO and CAO onto
NPA, NPB, NPC, NPD and NPE at 77°F. For all materials evaluated the adsorptive capacity is higher for
CHO and NPA and NPB showed higher adsorption capacity than the other samples.

Figure 4 —Experimental data for maximum adsorptive capacities of CHO and CAO asphaltenes onto NPA, NPB, NPC, NPD and NPD at
77°F. Adsorbent dose, 10 mg/ml; shaking rate, 300 rpm.

Fig. 5 shows the experimental adsorption isotherms at 77°F versus the calculated adsorption isotherms
at a) 210°F for CHO and b) 188°F for CAO. As seen, the asphaltenes extracted from CHO showed more
adsorption affinity towards NPA and NPB. That is, the NPA and NPB samples adsorbed more asphaltenes
from CHO than from CAO. This observation can be attributed to the intermolecular forces (i.e., the polar
interactions and electrical forces between the localized charges, which resulted from either permanent or
induced dipoles) between the most polar components of the asphaltenes molecules from CHO (Lower API
gravity than CAO) and the surfaces of the nanoparticles. As a comparison, for a concentration of 281
mg/L, NPA adsorbed 121 mg of asphaltenes extracted from CHO per gram of nanoparticles while at the
same concentration, the asphaltenes adsorbed from CAO where 107 mg/g. For NPB there was less
difference between the amounts of asphaltenes adsorbed. For the adsorption of the asphaltenes extracted
from CHO, the amount adsorbed was 64 mg/g at a concentration of 780 mg/L and for the asphaltenes
extracted from CAO, 61 mg/g at the same concentration. In Fig. 5 it can be also observed that NPA present
a higher uptake than NPB for both CHO and CAO, this can be attributed to SBET value of NPA that is
46% higher than the SBET of NPB. While NPA captured 121 mg/g of asphaltenes extracted from CHO
at a concentration of 281 mg/L, the NPB captured approximately 62 mg/g at the same concentration. For
8 SPE-179677-MS

CAO, a similar trend was observed, while the amount adsorbed at 280 mg/L for NPA was 106 mg/g, for
NPB was 58 mg/g. As expected, the amount of asphaltenes adsorbed decreased as the temperature
increased for both CHO and CAO due the exothermic behavior of the process. The temperature is strongly
related with the size of the asphaltene aggregates. As temperature increases, the aggregate size decreases,
and then the adsorption capacity decreases.39 It can be seen in Fig. 6-a that the amount adsorbed of the
asphaltenes extracted from CHO onto NPA at the reference concentration of 281 mg/L decreased from
121 to 80 mg/g, and for NPB at a concentration of 780 mg/L decreased from 64 to 49 mg/g. In the case
of the asphaltenes extracted from CAO (Fig. 6-b), the amount adsorbed onto NPA went from 113 to 100
mg/g at 364 mg/L, and for NPB from 61 to 44 mg/g at 800 mg/L.

Figure 5—Experimental data for adsorption isotherms of asphaltenes onto NPA and NPB at 77°F, and calculated isotherms for a) CHO
and b) CAO reservoir temperature 210°F and 188°F respectively.

Figure 5—a) Viscosity of CHO in function of the treatment dosage at 86 ({), 122 (□) and 175°F (⌬) and b) Viscosity of CAO in function
of the treatment dosage at 86 (□), 122 ({) and 156°F (o).
SPE-179677-MS 9

Figure 6 —The relative permeability curves for the base, the initial and the treated system after nanoparticle injection in their respective
formation plug saturated with a) CHO y b) CAO.

The Viscosity Measurements


Fig. 5a shows the effect of the oil based nanofluid on the CHO viscosity at 86, 122 and 175°F. It can be
observed that viscosity decreases between 75% and 99% for all temperatures studied. Fig. 5b shows the
curves of viscosity reduction for CAO at 86, 122 and 156°F, the viscosity decreases between 97% and
99.9%.
The Core Displacement Tests
According to the batch adsorption test results, NPA nanoparticles were used in the displacement test. For
each crude oil type, a core displacement was carried out by injecting the OBN at the respective reservoir
conditions. Three scenarios were evaluated: 1) Initial condition (Initial, washed core), 2) base case
scenario (core aged 10 days with crude oil) and 3) scenario after nanofluid injection. For Chichimene
crude oil (CHO), the Sor decreased up to 52% from scenario 2 to scenario 3, while Kro increased up to
358% from scenario 2 to scenario 3 (Fig. 6a). In the other hand, for Castilla crude oil (CAO), the Sor
decreased up to 57% from scenario 2 to scenario 3, while Kro increased up to 199% from scenario 2 to
scenario 3 (Fig. 6b). The use of the OBN increased hydrocarbon flow successfully, as a result of the
inhibition of asphaltene adsorption onto the rock surface, the critical reduction of oil viscosity, and the
effective restoration of original core wettability.

Field Application
Formation damage models were developed for Castilla and Chichimene fields. Fig. 7 shows the most
important sources of damage present in these areas. In Chichimene induced damage, emulsion damage
and problems in the relative permeability curves (Kr’s) are the most important sources of damage with
relative weights in the total skin of 29, 31.9 and 37% respectively (Fig. 7a). Other sources of damage are
present but in lower proportions like mineral and organic scales. In Castilla the damage is produced by
organic scales due to the precipitation of asphaltenes (30%), and mineral scales (14%). However, main
damage is induced during drilling, cementing and completion operations (56%) (Fig. 7b). The objective
of the stimulation process with the OBN in these fields is to mitigate the sources of damage associated
with organic deposits, emulsions and problems in the relative permeability curves (Kr’s), the actual main
sources of damage. For the trials a candidate selection process was implemented in the fields, after that
in Chichimene, CHA and CHB wells were selected as the pilots as well as CNA and CNB wells in
Castilla.
10 SPE-179677-MS

Figure 7—Distribution of the sources of formation damage in a) Chichimene and b) Castilla.

Chichimene Field Results


In Chichimene the intervention jobs were designed within a treatment radius of ~3 ft, squeezing 86 bbl
and 107 bbl of nanofluid as main treatment in CHA and CHB respectively. Other stages in the well
program were included. In order to remove organic deposits with an organic treatment was used, followed
by a step to remove calcite with a scale dissolver. The nanofluid and other fluids were pumped at matricial
low rate, and then wells were shut-in down for soaking for 12 hours. After the jobs, the production of the
wells showed instantaneous oil rate increases of around 310 bopd in CHA and 87 bopd in CHB, not BSW
reductions were observed. As an example complete results from CHA are going to be described in this
paper. CHA represents a well with the typical formation damage of the field (Fig. 7a). Determination of
skin damage with nodal analysis before and after the stimulation with the nanofluid shows a change of the
skin from 23 to 6.2 as it is shown in Fig. 8a, this reduction of skin damage 73% represent 310 bopd of
instantaneous oil rate increase.

Figure 8 —a) Skin before and after the job: i-Pre stimulation 263 BFPD, 5.65% BSW, 248 BOPD and Pwf ⴝ 1303 psi; ii-Post stimulation
641 BFPD, 12.95% BSW, 558 BOPD, Pwf ⴝ 1278 psi and ⌬Qo ⴝ 310. b) Production follows up.
SPE-179677-MS 11

Until now production gain have been extended until day 269th after the well was treated, oil rate still
above base line (Fig. 8b). According to results looks an efficient treatment. It is worth to mention that the
capital inverted in the stimulation operation was recovered in the first three months and the following
months represent economic benefits. As soon as the well was opened in production the first observation
was the reduction in the crude oil viscosity and improvement in oil mobility. Well performance coincides
with the reduction in viscosity of the oil produced in the first 120 days (Fig. 9a), after that the
measurements of this property reached base line conditions. In the first 9 days were observed viscosity
reductions of around 98%, it’s important to remember that the nanofluid was pumped into formation just
one time and it’s not a continuous injection. Improvement in oil mobility is still noted by almost the same
269 days that maintenance for residual nano-particles was recorded in the well. Results are shown in the
Fig. 9b. Due to the high polarity of the employed nanoparticles, tracking of existing residual nano-
particles has been monitored by measuring content of nano-particles in produced water, and for some
samples without free water, lixiviation with distilled water was done to obtain the aqueous sample for
residual measuring. The residual concentration of nanoparticles after 269 days is 56 ppm (Fig. 9b). This
value is indicting that treatment life is going to be longer.

Figure 9 —a) Viscosity of produced fluids from well CHA and b) residual of nanoparticles.

Castilla Field Results


Stimulation jobs in Castilla were designed within a penetration radius of ~3 ft, 200 bbl and 150 bbl of
nanofluid were pumped as main treatment in CNA and CNB respectively. In CNA other stages in the well
program were included. In order to remove organic deposits with an organic treatment was used, followed
by a step to remove calcite with a scale dissolver. In CNB only nanofluid was squeezed. The nanofluid
and other fluids were pumped at matricial low rate, and then wells were shut-in down for soaking between
12-18 hours. Instantaneous oil rate increases of 270 bopd in CNA and 280 bopd in CNB and BSW
reductions of ~11% were observed. Like in the last section results from well CNA are going to be
described in this paper. Formation damage distribution described for Castilla field (Fig. 7b) is fulfilled in
CNA well. Determination of skin damage with nodal analysis before and after the stimulation with the
nanofluid shows a change of the skin from 47 to 19 as it is shown in Fig.e 10a, this reduction of skin
damage 60% represent 270 bopd of instantaneous oil rate increase.
12 SPE-179677-MS

Figure 10 —a) Skin before and after the job: i-Pre stimulation 201 BFPD, BSW 20%, 160 BOPD and Pfw ⴝ 1280 psi; ii-Post stimulation
485 BFPD, BSW 9.67%, 438 BOPD, Pfw ⴝ 938 and ⌬Qo ⴝ 278. b) Production follows up.

Production gains have been extended until day 174th after the well reached base line (Fig. 10b).
According to results looks an efficient treatment however results from Chichimene were better. It is worth
to mention that the capital inverted in the stimulation operation was recovered in the first four months. As
soon as the well was opened in production the first observation was the reduction in the crude oil viscosity
and improvement in oil mobility. Also effects in BSW reduction were noted about 11% points less than
before job this is because the system becomes strong water wet. Effects in BSW and viscosity reduction
and improvement in oil mobility were noted by almost the same 174 days that maintenance for residual
nano-particles was recorded in the well. Results are shown in Fig. 11. The residual concentration of
nanoparticles after 200 days is negligible (fig. 11b). In the first 30 days were observed viscosity reductions
of around 47%.

Figure 11—a) Viscosity of produced fluids well CNA and b) residual of nanoparticles.
SPE-179677-MS 13

Conclusions
● Lab results showed that nanoparticles A (NPA) lead to a higher asphaltenes uptake than nano-
particles B (NPB) for both CHO and CAO.
● The treatment based on Nanoparticles A was effective to reduce the CHO and CAO viscosity. A
new protocol at laboratory scale was implemented to evaluate the behavior of the nanofluid in the
porous media at reservoir temperature and pressures. The injection of nanofluidos into porous
media showed that nanoparticles are able to restore rock wettability from oil-wet to water-wet and
increase the oil mobility. The nanoparticles also were able to restore production and led to
improvements in recovery percentage due to their ability to change the wettability of the core.
● Formation damage models were developed, identifying that Castilla and Chichimene fields are
good candidates for the use of a treatment based on nanoparticles. A candidate selection process
was implemented in the fields, in Chichimene, CHA and CHB wells were selected as the pilots as
well as CNA and CNB wells in Castilla.
● Main sources of formation damage in these fields are damage associated with organic deposits,
emulsions and problems in the relative permeability curves (Kr’s). The objective of the stimulation
process with the OBN is to mitigate these mechanisms.
● In Chichimene field interventions of the wells showed instantaneous oil rate increases of around
310 bopd in CHA and 87 bopd in CHB, not BSW reductions were observed.
● A nodal analysis in CHA before and after the stimulation with the nanofluid shows a change of
the skin from 23 to 6.2, this represent a reduction in the skin damage about 73%. In the first 9 days
were observed viscosity reductions of around 98%, it’s important to remember that the nanofluid
was pumped into formation just one time and it’s not a continuous injection.
● Production benefits from stimulation job have been extended until day 269th after the well was
treated, nowdays oil rate still above base line. The residual concentration of nanoparticles after 269
days is 56 ppm. This value is indicating that treatment life is going to be longer.
● Stimulations jobs in Castilla field showed instantaneous oil rate increases of 270 bopd in CNA and
280 bopd in CNB and BSW reductions of ~11% were observed.
● A nodal analysis in CNA well before and after the treatment showed reduction in the skin damage
about 73% (skin from 47 to 19). In the first 30 days was observed a viscosity reduction of around
47%.
● Results from CNA finished 174th days after the well was treated. The residual concentration of
nanoparticles after 200 days is negligible.
● Jobs results in Chichimene field were extended longer than Castilla. This could be because the
CHO is heavier than CAO.
● Final results are still on evaluation but results are promising (BSW reduction in Castilla is a fact)
and final incremental oil should follow; now is possible to extend application of nanofluid in other
wells in these fields.

6. Acknowledgements
The authors would like to thank Ecopetrol for granting permission to present and publish this paper. The
authors acknowledge Universidad Nacional de Colombia for logistical and financial support.

References
Franco, Camilo A, Monica M Lozano, Socrates Alejandro Acevedo et al.2015. Effects of Resin I on Asphaltene
Adsorption onto Nanoparticles: A novel method for obtaining Asphaltenes/Resins Isotherms (in Energy & Fuels.
Franco, Camilo A, Tatiana Montoya, Nashaat N Nassar et al.2013. Adsorption and Subsequent Oxidation of Colombian
Asphaltenes onto Nickel and/or Palladium Oxide Supported on Fumed Silica Nanoparticles (in Energy & Fuels 27
(12): 7336 –7347.
14 SPE-179677-MS

Franco, Camilo A, Nashaat N Nassar, Tatiana Montoya et al. 2014. NiO and PdO Supported on Fumed Silica
Nanoparticles for Adsorption and Catalytic Steam Gasification of Colombian C7 Asphaltenes. In Handbook on Oil
Production Research, ed. Jacquelyn Ambrosio, Chap. 3. Nova Science Publishers.
Franco, Camilo A, Nashaat N Nassar, Marco A Ruiz et al.2013. Nanoparticles for inhibition of asphaltenes damage:
adsorption study and displacement test on porous media (in Energy & Fuels 27 (6): 2899 –2907.
Franco, Camilo, Edgar Patiño, Pedro Benjumea et al.2013. Kinetic and thermodynamic equilibrium of asphaltenes sorption
onto nanoparticles of nickel oxide supported on nanoparticulated alumina (in Fuel 105: 408 –414.
Giraldo, Juliana, Pedro Benjumea, Sergio Lopera et al.2013. Wettability alteration of sandstone cores by alumina-based
nanofluids (in Energy & Fuels 27 (7): 3659 –3665.
Hamedi Shokrlu, Yousef, Tayfun Babadagli. Effects of nano-sized metals on viscosity reduction of heavy oil/bitumen
during thermal applications. Society of Petroleum Engineers.
Ju, Binshan, Tailiang Fan. 2009. Experimental study and mathematical model of nanoparticle transport in porous media
(in Powder Technology 192 (2): 195–202.
Ju, Binshan, Tailiang Fan, Mingxue Ma. 2006. Enhanced oil recovery by flooding with hydrophilic nanoparticles (in China
Particuology 4 (01): 41–46.
Karimi, Ali, Zahra Fakhroueian, Alireza Bahramian et al.2012. Wettability alteration in carbonates using zirconium oxide
nanofluids: EOR implications (in Energy & Fuels 26 (2): 1028 –1036.
Maghzi, Ali, Ali Mohebbi, Riyaz Kharrat et al.2011. Pore-scale monitoring of wettability alteration by silica nanoparticles
during polymer flooding to heavy oil in a five-spot glass micromodel (in Transport in porous media 87 (3): 653–664.
Rouquerol, Jean, Françoise Rouquerol, Philip Llewellyn et al.2013. Adsorption by powders and porous solids: principles,
methodology and applications, Academic press (Reprint).
Sherik, Abdelmounam M, Khalid M Nabulsi. 2009. Applications of nanotechnology in oil and gas (in International
Journal of Nano and Biomaterials 2 (1-5): 409 –415.
Zabala, R, E Mora, OF Botero et al.Nano-Technology for Asphaltenes Inhibition in Cupiagua South Wells.
Zitha, PLJ. 2005. Smart fluids in the oilfield (in Exploration & Production: The Oil & Gas Review: 66 –68.

Das könnte Ihnen auch gefallen