Sei sulla pagina 1di 17

Fuel 185 (2016) 886902

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Review article

In-situ heavy and extra-heavy oil recovery: A review


Kun Guo a,b, Hailong Li c, Zhixin Yu a,b,
a

Department of Petroleum Engineering, University of Stavanger, 4036 Stavanger, Norway


The National IOR Centre of Norway, University of Stavanger, 4036 Stavanger, Norway
c
Department of Energy, Building and Environment, Mlardalen University, 72123 Vsters, Sweden
b

h i g h l i g h t s
 A review of the existing in-situ heavy oil recovery techniques is presented.
 Various aspects of traditional recovery methods are systematically discussed.
 The in-situ catalytic upgrading and recovery of heavy crude oil is elaborated.
 Standards and methodologies are summarized to establish the technological criteria.

a r t i c l e

i n f o

Article history:
Received 23 June 2016
Received in revised form 8 August 2016
Accepted 11 August 2016
Available online 20 August 2016
Keywords:
Heavy oil recovery
Thermal injection
Chemical injection
Gas injection
Catalysis
In-situ upgrading

a b s t r a c t
Due to the growing global energy demand and increasingly limited availability of conventional or easyto-produce crude oils, extensive attention is being paid to the exploitation of unconventional heavy and
extra-heavy oils. However, their inherent properties, characterized by high viscosity and poor mobility,
coupled with the complex reservoir configuration, make the desired recovery processes very challenging.
Although several in-situ recovery techniques have been employed in oil reservoirs worldwide, most of
them are still suffering from low sweep and displacement efficiencies, high capital investment, potential
formation damage and negative environmental footprints.
This paper aims to provide a comprehensive review of the existing in-situ heavy oil recovery techniques, which fall into three categories of thermal injection, chemical injection and gas injection.
Different aspects including the fundamental principles, main features, applicability, and limitations of
these recovery processes are elaborated sequentially to illustrate the current technology status.
Underlying mechanisms causing the relatively low recovery factors will also be pinpointed.
Furthermore, this paper focuses on the technology using novel and active catalysts for simultaneous
heavy oil upgrading and recovery, especially in the case of metallic nanocatalysts. Rationales, advantages
and challenges regarding this in-situ catalytic upgrading technology will be extensively described for
their potential implementation in fields. It is noteworthy that many recovery techniques are still limited
to the laboratory scale with needs for further investigations. Therefore, this paper also covers the evaluation standards and analytical methodologies of heavy and extra-heavy oil recovery to establish experimental screening criteria. In the end, economic and environmental aspects of the in-situ catalytic
upgrading technology have been briefly discussed. The objective of this review is to present a wide range
of expertise related to the in-situ heavy oil recovery processes, and to introduce the in-situ catalytic
upgrading technology as an effective and environmental friendly heavy oil recovery process.
2016 Elsevier Ltd. All rights reserved.

Contents
1.
2.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Heavy and extra-heavy oils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Physical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author at: Department of Petroleum Engineering, University of


Stavanger, 4036 Stavanger, Norway.
E-mail address: zhixin.yu@uis.no (Z. Yu).
http://dx.doi.org/10.1016/j.fuel.2016.08.047
0016-2361/ 2016 Elsevier Ltd. All rights reserved.

887
888
888
888

K. Guo et al. / Fuel 185 (2016) 886902

3.

4.

5.

6.

Recovery technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Thermal injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.1.
Cyclic steam stimulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.2.
Steam flooding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.3.
Steam-assisted gravity drainage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.4.
In-situ combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Chemical injection. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1.
Surfactant flooding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.2.
Polymer flooding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.3.
Alkaline surfactant polymer flooding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.4.
Solvent flooding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
Gas injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.
In-situ catalytic upgrading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.1.
Catalysts for in-situ upgrading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.2.
Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.3.
Catalyst preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.4.
Structural parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.5.
Transport behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Quality enhancement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.
Viscosity and C/H ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.
Coke formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.
Heteroatoms content. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.4.
Composition analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Promises and challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.
Formation damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.
Economical cost. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.
Environmental impact. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Projection of the oil demand by the Organization of the Petroleum Exporting Countries (OPEC) will reach 111.1 million barrels
per day by 2040, with a 23.1% increment compared to current data
[1]. Meanwhile, the consumption of conventional light oils has
resulted in declining reserves of these resources. As fossil fuel will
remain to be the main energy source for the coming decades, there
is an urgent need to exploit alternative fossil resources. Therefore,
substantial efforts have been devoted to the effective production
of heavy and extra-heavy (i.e., natural bitumen or oil sands) oils
from reservoirs, which account for ca. 70% of total world oil reserves
[2,3].
Compared to the production of conventional oils, heavy oil
recovery is more problematic due to the inherent properties, such
as high viscosity or even immobility, high carbon/hydrogen (C/H)
ratios and high heteroatom contents [4,5]. Complex formation configuration could also add additional difficulty in the oil production.
Notwithstanding, the key mechanism for effective recovery has
been identified to be the oil viscosity reduction and the resulting
improved oil mobility. With this in mind, according to the
temperature-viscosity correlation, various external energy sources
are supplied to heat up the heavy oils, which eases their flow and
extraction from the underground. Traditionally, oil production is
divided into three steps: primary, secondary and tertiary recoveries. In the production of heavy oils, primary and secondary recoveries are dominated by cold production and water flooding, as in
the cases of Canadian oil sands, Venezuelan heavy oils and UK Continental Shelf. These recovery methods are however either limited
to relatively shallow reservoirs or only effective to lighter heavy
oils [6]. To realize a high recovery factor, tertiary recovery, or more
widely known as enhanced oil recovery (EOR), is rather critical and
necessary to extract more oils left behind by the primary and secondary recoveries.

887

888
888
888
888
889
890
891
891
892
892
892
893
893
893
896
896
897
897
897
897
898
898
898
898
898
898
899
899
899
899

Until recently, a variety of EOR techniques have been proposed


and developed, which can be conveniently classified into thermal,
chemical and gas injection [6,7]. Among these methods, thermal
injection is recognized as an effective one with high recovery factors up to 70% of the original oil in place (OOIP). Typical thermal
recovery includes steam-assisted gravity drainage, cyclic steam
stimulation and in-situ combustion. However, these technically
successful methods are still challenged both economically and
environmentally because of high cost of heat supply along with
excessive carbon dioxide (CO2) emission and costly posttreatment and maintenance [8]. Chemical and gas injection are
also, to some extent, commercially viable, especially the CO2 flooding, which is gaining considerable interest recently due to its
potential for CO2 sequestration. Unfortunately, these methods are
often suffering from poor mobility control and severe viscous fingering, resulting in insufficient sweep and displacement efficiencies. Due to the high interfacial tension (IFT) and distinct
viscosity contrast between displacing fluid and oil, particularly in
terms of bitumen or oil sands, these techniques are usually ineffective and far from commercial expectations [4]. The abovementioned obstacles thus have become the driving forces behind
the desire for a better recovery solution.
Simultaneous recovery and upgrading of heavy oils in reservoirs
using thermal recovery along with injection of active catalysts are
emerging as a promising technology that combines the advantages
of both thermal recovery and in-situ catalysis [911]. Growing
interests are being dedicated to investigate the feasibility of applying active catalysts for in-situ heavy oil recovery, especially metallic nanoparticle assisted recovery method [12]. The integration of
nanotechnology with enhanced heavy oil recovery could provide
a novel pathway and have the potential to achieve higher recovery
efficiency and better recovery factor. In this review, a chapter will
specially focus on different aspects of the nanoparticle-assisted
recovery in detail compared to other recovery methods.

888

K. Guo et al. / Fuel 185 (2016) 886902

Thereafter, theoretical knowledge and practical characterization methods on the evaluation of oil quality are presented systematically. Promises and challenges faced by current recovery
technologies are illustrated with regard to formation damage and
economical and environmental aspects. This report intends to
review the recent progresses in in-situ heavy and extra-heavy oils
recovery and aims to provide systematic knowledge for petroleum
engineers in developing new technologies to achieve higher recovery factor from reservoirs.

EOR is essential to be followed. It is worth mentioning that oil viscosity is closely associated with temperature. Oil viscosity
decreases with the increase of temperature. Accordingly, external
thermal sources can be introduced to heat the oil and reduce the
viscosity, which are generally known as steam-based thermal
injection. In order to improve the displacement efficiency, recovery
techniques using chemicals and gases have also been successfully
implemented. All these technologies including the hybrid ones will
be illustrated in the next sections.

2. Heavy and extra-heavy oils

3.1. Thermal injection

2.1. Definition

3.1.1. Cyclic steam stimulation


Cyclic steam stimulation (CSS), also known as Huff and Puff,
was accidently discovered by Shell Oil in the Mene Grande Tar
Sands of Venezuela five decades ago [15]. Back then, steam was
first injected through a vertical well until cracks were generated
on the ground releasing steam, water and heated oil. Surprisingly,
oil then kept flowing out from the injection well. This phenomenon
is later referred to as steam soak.
CSS is a three-step process that begins with the injection of
high-pressure steam for several weeks, which is followed by the
soak step allowing the injected steam to sufficiently diffuse
through the reservoir. During this period, steam and hot water heat
the viscous oils and increase their mobility. Subsequently, the third
step starts by using the same well as production well to extract the
mobile oils. Fig. 1 illustrates the schematic diagram of a CSS cycle.
The production rate stays at a high level for a short period before
declining gradually over several months [6]. This cycle is then
restarted successively to reach a desirable recovery factor. It is
worth noting that water production or steam-oil ratio gradually
increases with repeated cycles, resulting in a decreased economic
profit. Therefore, CSS is usually applied as the primary production
process before other follow-up recovery techniques such as in-situ
combustion, CO2 injection and steam flooding. Furthermore, the
resulting high pressure during the CSS process also increases the
requirements to maintain the reservoir formation. Accordingly,
the amount of required steam and duration for each cycle should
be thoroughly considered, which largely depend on the reservoir
thickness and formation permeability as well as oil viscosity
[16,17]. To increase the effectiveness of CSS, injected steam is modified by incorporating chemical additives such as solvents, surfactants and gases [18].
Although CSS has been commercially operated in heavy oil
reservoirs in Canada, Brazil, Venezuela and USA, typically a relatively low recovery factor of around 30% of OOIP is achieved
[4,19,20]. Besides, the use of steam, like all steam-based injection
methods, means an intensive energy consumption as well as large
CO2 emission ascribed to the steam generation from methane
(CH4), increasing both the economic and environmental costs.

Heavy or extra-heavy oils are highly viscous oils that cannot


easily flow to production wells at normal reservoir conditions.
Heavy is defined because the density or specific gravity is higher
than that of lighter oils (i.e., conventional oils). It is widely
accepted to adopt the specific gravity and viscosity as two standards to classify light, heavy and extra-heavy oils, which are exemplified in Table 1 [2,4,5]. Specific gravity is measured based on the
American Petroleum Institute in units called API degrees (API); the
lower the number of API degrees, the higher the specific gravity of
the oil. Viscosity is measured in centipoises (cP) that represents the
oils resistance to flow; the higher the value, the higher the
viscosity.
2.2. Physical properties
Heavy oil and bitumen are mainly hydrocarbons with high
molecular weights resulting in high boiling points. Typical carbon
atoms of most of these hydrocarbons are above 60. Their components can be categorized as saturate, aromatic, resin and asphaltene (SARA) according to the polarizability and polarity. Saturates
comprise nonpolar fractions including linear, branched, and cyclic
saturated hydrocarbons. Aromatics that contain one or more aromatic rings are relatively polarizable. The other two components,
resin and asphaltene, are both polar substances. The major difference between them lies in that resins are miscible with nheptane whereas asphaltenes are n-heptane insoluble and toluene
soluble. Additionally, SARA also contain a certain amount of nitrogen (N), oxygen (O), sulfur (S) and metal elements, especially in the
case of asphaltene, which is also the main ingredient of heavy oil
and bitumen. It has been demonstrated that reservoirs show big
geological differences in producing oils with different SARA proportion and sulfur content. For example, high sulfur heavy oils
mostly exist in North America, South America and Middle East
countries [10].
3. Recovery technologies
Cold production and surface mining are low-risk primary recovery approaches for heavy oils, but they are limited to the depth of
reservoirs and relatively light grades inside [13,14]. Water flooding, as a popular secondary recovery method, is also not viable
due to the huge viscosity difference between water and oils, leading to low sweep efficiency [8]. To achieve a high recovery factor,

Table 1
API gravity and viscosity of light, heavy and extra-heavy oils.

Light oil
Heavy oil
Extra-heavy oil (bitumen)

API gravity ()

Viscosity (cP)

Density (kg/m3)

>22
1022
<10

<100
>100
>10,000

<934
9341000
>1000

3.1.2. Steam flooding


Steam flooding, also named steam drive or steam stimulation, is
used as a recovery technology to extract denser and thicker oils
than conventional ones at the field scale. It requires a continuous
supply of steam through vertical injection wells to heat the heavy
oils, as illustrated in Fig. 2. The increasing temperature reduces
their viscosity and thereby enables them to flow more easily to
the production wells. Meanwhile, another mechanism of physical
displacement, similar to that of water flooding, is also involved
because a zone of condensed water is formed between steam and
oil, which pushes the heated oil towards the production well.
Commercial application of steam flooding has started since the
early 1960s in California, where steam flooding was intensively
used to supplement CSS [21]. From then on, this technology has
been field tested under a wide variety of reservoir conditions,

K. Guo et al. / Fuel 185 (2016) 886902

889

Fig. 1. Schematic illustration of cyclic steam stimulation: CSS begins with steam injection, followed by the soak step allowing the injected steam to heat up the oil; the heated
oil is produced from the same well.

Fig. 2. Schematic illustration of steam flooding: a continuous supply of steam is injected through vertical wells to heat the heavy oils and push the heated oil towards the
vertical production well.

and typical recovery factors are around 50% of OOIP [8,13]. Nevertheless, with the aid of simulation models, researchers have realized a more plausible mechanism behind this technique.
Essentially, the injected steam moves rapidly to the top of internal
space and then penetrates through a thin layer towards the production well. Once steam breakthrough reaches the production
well, differential pressure between injection and production wells
is dramatically reduced, which then fails to drive the oils effectively. It is therefore critical to control the bypass of steam through
the oil. Furthermore, similar to CSS, steam flooding also faces big
economic and environmental challenges in terms of steam supply
and CO2 emission. The water used brings issues such as the handling of large volumes of liquids, and extra dehydration facility
for oil and water separation.
3.1.3. Steam-assisted gravity drainage
As a superior form of steam injection, steam-assisted gravity
drainage (SAGD) was invented by Dr. Roger Butler for the recovery

of Alberta bitumen in the 1970s [22,23]. Since then, SAGD has been
well developed and commercially applied as the most important
thermal injection method for in-situ recovery of heavy and extraheavy oil resources. High recovery factors of up to 70% of OOIP
have been achieved together with high production rates. Despite
pilot tests are reported in USA, China and Venezuela, successful
cases only exist in Canada and more specifically in Fort McMurray
and Athabasca [2427].
Typically, a pair of horizontal wells, with one well 46 m higher
than the other, are drilled to reach the pay zone. Steam is injected
through the upper injection well to heat the oils and thus reduce
their viscosity, which assists them to flow towards the lower production well and be pumped to the surface, as represented in Fig. 3.
More specifically, the injected steam rises to the steam-oil interfaces and gradually forms a steam chamber. This chamber expands
continuously to extrude condensed water and heated oils to the
production well. Due to the gravity effect, the tendency of the
steam to flow directly to the production well and thus bypass

890

K. Guo et al. / Fuel 185 (2016) 886902

Fig. 3. Schematic illustration of steam-assisted gravity drainage: a pair of horizontal wells are drilled; steam is injected through the upper well and a steam chamber is
formed; this chamber expands and drives the heated oil towards the lower production well.

the reservoir could be largely reduced or even eliminated, suggesting a higher sweep efficiency.
The applicability of SAGD is restricted by a series of geological
and geometric factors, which include sufficient reservoir thickness
to allow spatial placement of the two wells, high horizontal and
vertical homogeneity to ensure the formation of a uniform steam
chamber, and good permeability to enable the efficient flow of
the fluids [2832]. In addition, the presence of a gas cap or a bottom aquifer could reduce the efficiency of the SAGD process. Since
steam is also involved in the SAGD, large volumes of water is
required. It is estimated that one barrel of produced oil corresponds to over four barrels of consumed water [33]. This huge
amount of water then has to be recycled from the production well
and the cost of this post treatment can be significant in the total
project budget. Another issue is the use of natural gas as fuels,
which results in considerable amount of greenhouse gas (GHG)
emission.
To control the cost of external fuels, the idea of upgrading part
of the extracted heavy oils to proper gas fuels is proposed and practiced. This practical solution can reduce the demand for natural

gas. Furthermore, to reduce the energy consumption, a chemical


solvent could be supplemented, which is known as the expanding
solvent SAGD method [34,35]. A further solution to enhance the
recovery efficiency is to add a small amount of non-condensable
gases such as natural gas or nitrogen to the process, which is
referred to as steam and gas push [36]. Several variations of SAGD
such as fast SAGD and single well SAGD have been investigated,
but these methods are still at an early stage of development and
not expected to achieve technological breakthrough in the near
future [6].
3.1.4. In-situ combustion
In-situ combustion (ISC), or fire flooding, defines recovery processes by burning of oil originally existed in the reservoir [37,38].
Air, or more precisely a combustion promoter, is injected to react
with hydrocarbons that serve as fuels. These exothermic oxidation
reactions provide enormous heat to help improve the mobility of
the unburned heavy oils and then drain them to the production
wells, as shown in Fig. 4. Instead of steam, compressed air is
injected in ISC, which greatly reduces the energy demand relative

Fig. 4. Schematic illustration of in-situ combustion: air is injected to burn the oil and large amount of heat is released to ease the flow of heavy oil towards the production
well.

K. Guo et al. / Fuel 185 (2016) 886902

to steam generation. Moreover, this process does not involve water


recycling and has much lower GHG emission. Therefore, as the oldest thermal recovery technique, ISC has become the second most
important one for heavy and extra-heavy oils in the past decades
[39].
ISC starts with the ignition. When the combustion conditions
are satisfied, i.e. with formation temperature above 60 C, spontaneous ignition could occur after a short period of oxidizing gas
injection [3]. Otherwise, oil needs to be ignited by manually heating the formation near the injection wellbore. If ignition occurs at
the production well, the process is named reverse combustion
rather than conventional forward combustion [3]. After ignition,
a gas chamber composed of air and combustion gases and a combustion front are formed. The combustion front is then propagated
by continuous gas chamber expansion towards the production
wells, displacing the unburned heavy oils. The mechanism of ISC
is generally described by the formation of several well-defined
zones, which include the burned zone, combustion zone, cracking
zone, evaporation and visbreaking zone, steam zone, water bank
and oil bank with different temperatures and saturation of oil
and water [40,41].
ISC is regarded as a potentially hazardous recovery technique
because of many failures during early field tests, which are mainly
due to applying this technique to the wrong type of reservoirs or
inappropriate engineering parameters [42]. ISC is normally suitable for high permeability, homogeneous sandstones and shallow
onshore reservoirs. Although geological reservoir characterization
is well performed in some cases, there is still no guarantee of success due to difficulties in controlling the operation. That is why
production engineers would take ISC as the last option among all
the suitable techniques. Moreover, although ISC avoids the
incurred steam costs, it requires powerful and expensive compressors to inject air. The high pressure could also damage the reservoir
formation. With the objective of improving the control of conventional ISC, toe-to-heel air injection (THAI) is proposed by combining ISC with a horizontal production well, as shown in Fig. 5
[43,44]. In THAI, the combustion front moves from the toe position to the heel along the production well. This theory has been
experimentally demonstrated to get a stable front propagation and
the moving distance for mobile oils is much shorter, suggesting up
to 85% of IIOP can be extracted along with high production rates
[45]. In recently years, the catalytic add-on of THAI (THAI-CAPRI,

891

CAtalytic upgrading PRocess in-situ) is developed to integrate the


thermal recovery with in-situ catalytic upgrading of heavy oil.
The difference between them is that THAI-CAPRI has a catalyst
layer fixed onto the horizontal production well. As the heated oil
moves towards this well, thermal catalytic cracking will occur to
upgrade the heavy oil, which acts like an underground refinery.
3.2. Chemical injection
As mentioned above, water flooding is basically not applicable
because the viscosity difference between heavy oil and water is
too big, which results in a serious issue known as viscous fingering
[46]. It depicts the situation that when a fluid displaces a more viscous one, the interface between them does not keep flat and perpendicular to the flow axis [47]. In case of water flooding, the
less viscous water tends to bypass once it breaks through the viscous heavy oils because of the high IFT and immiscibility between
these two phases. This fingering instability thus results in poor
sweep and displacement efficiencies. In order to prevent this phenomenon, different chemicals are incorporated to modify the properties of injected water. According to the chemical types, these
methods are generally categorized as surfactant flooding, polymer
flooding, alkaline flooding and solvent flooding, which will be discussed sequentially in the next sections [6,7,48]. It should be
emphasized that chemical injection is mostly used for heavy oil
EOR but not very suitable for immobile bitumen or oil sands. In
addition, instead of reducing the oil viscosity, which is the main
mechanism of the thermal methods, lowering the IFT, mobility
control and wettability alteration are the alternative solutions to
improve the mobility of viscous oils in all chemical-based recovery
technologies [4952]. In particular, when immiscible water and oil
are in contact, an interface will form between the two phases
together with the inevitable IFT. This high IFT is responsible for
the resistance of flow of residual oils. IFT reduction is thus a key
factor in the mobility improvement of heavy oils.
3.2.1. Surfactant flooding
Surfactant flooding refers to the recovery method in which a
small amount of surfactant is added to the aqueous injection fluid
to displace the oil. The rationale behind the use of surfactant is to
recover the residual oil trapped by capillary pressure, which is the
major factor controlling the oil-water distribution in porous

Fig. 5. Schematic illustration of toe-to-heel air injection: air is injected to burn the oil and large amount of heat is released; the heated oil is driven along a horizontal
production well.

892

K. Guo et al. / Fuel 185 (2016) 886902

reservoir rocks [53]. If surfactants are properly chosen, they can


serve as wetting agents to lower the IFT, enabling easy spreading.
More specifically, a surfactant is usually a compound that is
amphiphilic, or it contains both hydrophobic and hydrophilic
groups, which are oil-soluble and water-soluble respectively.
Therefore, the surfactant can present at the interface between
water and oil and reduce the IFT significantly. The reduced IFT then
allows the formation of dispersed oil droplets that are stabilized or
emulsified by surfactant molecules inside the water phase, i.e. oilin-water (O/W) emulsion, or the opposite water-in-oil (W/O)
emulsion. In either case, the water phase viscosity and the oil
mobility are substantially improved.
The incremental recovery factor of 1020% of OOIP for tertiary
recovery, if not achievable for primary and secondary recovery, is
possible by surfactant flooding [3]. The technical success is closely
associated with the realization of distributing viscous oils as stable
and durable droplets in water phase. Consequently, the surfactant
properties, reservoir rock wettability, brine salinity, and reservoir
temperature and pressure are critical factors that should be considered [54]. For example, a surfactant with high HLB (hydrophilic
lipophilic balance) is preferred for generating an O/W emulsion.
The resulting high capillary number (Nc) also implies a decrease
in residual oil saturation and an increase in oil recovery. It should
also be noted that certain surfactants can be very expensive for
industrial practice or might bring some undesirable components
that affect the produced oil quality.
Surfactant flooding was a popular technology in the 1970s and
80s [7]. Since then, the advancement in surfactant manufacturing
has reduced the concentration of added surfactant from up to
10% to below 0.5% nowadays [46]. However, prior to the field
implementation of this technique, key process parameters need
to be optimized through a series of test procedures, resulting in relatively long operation cycles. In addition, this technique also
requires large capital expenditure for the blending equipment
and special injection pumps. The influence of the injected surfactants on the environment and human health also remains a controversial issue.

3.2.2. Polymer flooding


When operating polymer flooding, a polymer with high molecule weight and high viscosity is added into the water phase to
decrease the mobility of water and thus improve the displacement
efficiency [55]. The concept of mobility control lays the foundation
of polymer flooding method [56]. Generally, mobility ratio is
defined as the mobility of displacing phase (i.e., water in water
flooding) divided by the mobility of the displaced phase (i.e., oil).
A low value of mobility ratio (61) is favored because that mean
the displaced oil has a higher tendency to move than the displacing
water. Therefore, the phenomenon that water flows through more
permeable channels and bypasses the viscous oils is effectively
reduced, resulting in increased volumetric sweep efficiency.
Compared with surfactant flooding, polymer flooding is more
sensitive to rock permeability. The fact that the rock tends to retain
the polymer during the transport process means this technique is
only applicable to highly porous rock. Furthermore, despite the
increase of polymer molecular weight and size can improve the
aqueous viscosity, it will also increase the difficulty for polymers
to maintain mobility. Similar trade-off is also to be made with
respect to the polymer concentration. The interaction between
polymer molecules and the pores of reservoir rock should be considered to avoid polymer retention, which could originate from
surface adsorption, mechanical entrapment and polymer degradation. It is also reported that pH value, brine salinity, clay content
and reservoir temperature could exhibit considerable impacts on
the technical and economic success of polymer flooding [5759].

3.2.3. Alkaline surfactant polymer flooding


Alkaline surfactant polymer (ASP) flooding uses two sources of
surfactants and a polymer. One of the surfactants is in-situ produced by the injected alkaline chemicals that react with the
surface-active components originated from oil. Typically, the ASP
formulation consists of about 0.51% alkali, 0.1% surfactant and
0.1% polymer, which are injected together into the reservoir as a
slug [60]. A slug is added normally at 0.30.4 PV (pore volume)
for the effective performance. The alkali, such as sodium carbonate
or sodium hydroxide, can react with organic acids that commonly
exist in heavy oils, resulting in the formation of natural surfactants.
The injected surfactant helps to reduce the IFT between oil and
water and thus improve the accessibility between the alkalis and
organic acids. The polymer, mainly partially hydrolyzed polyacrylamide, can increase the viscosity of the aqueous solution to minimize channeling and enhance mobility control [61].
ASP combines the benefits of surfactant flooding, polymer
flooding and the use of alkali, while it also means that the process
requirements are even stricter for successful implementation. As
discussed before, key parameters to be considered are formation
brine salinity, rock permeability, clay content and reservoir temperature [6264]. It should be pointed out that most ASP applications are limited to sandstone reservoirs instead of carbonate [7].
This is due to the fact that anionic surfactants tend to adsorb to
carbonates while cationic surfactants are expensive; and that high
content of divalent cations often exist in carbonates, which results
in disadvantageous precipitation and alkaline waste.
3.2.4. Solvent flooding
Solvent flooding is quite different from the other chemical
flooding fundamentally. Strictly speaking, solvent flooding works
similar as the thermal injection despite several chemical solvents
are applied. In these methods, solvents are either partially introduced into the steam, such as LASER (liquid addition to steam to
enhance recovery) and other hybrid steamsolvent processes, or
completely replacing the steam, such as VAPEX (vapor extraction)
and CSI (cyclic solvent injection) [6569].
In terms of energy consumption, environmental impact, capital
investment and safety issue, solvent-based recovery methods exhibit distinct advantages over the thermal injection technology.
Admittedly, compatibility-induced viscosity reduction by miscible
solvents is more cost-effective than heat-enhanced mode by the
steam. Much less surface facilities are required for the solvent
flooding than that for the steam injection, which will reduce the
capital investment and operating cost. Moreover, when sufficient
solvent is dissolved into the heavy oil, asphaltene precipitation
could occur so that the produced heavy oil is in-situ deasphalted
and upgraded. Compared to LASER and hybrid steamsolvent
methods, CSI and VAPEX do not utilize steam. CSI is analogous to
the CSS process, in which a solvent mixture instead of steam is
injected into the reservoir and followed by a soaking period and
a production period. Normally, the solvents should be mainly gaseous, exhibit good solubility in oil, and be relatively affordable.
With this in mind, several blends consisting of readily available
CH4 or CO2 as carrier gas and additional propane or butane have
been experimentally tested [70]. Before testing CSI at field scale,
systematic investigations should be conducted to evaluate operational parameters such as solvent amount, soaking time and slug
injection strategies. In addition, although solvents provide better
performance than steam, the cost of solvents is much higher than
that of steam. Thus, there will be a compromise between the production efficiency and the solvent cost.
VAPEX, using vaporized solvents (normally propane or propaneCH4 or propanebutane), is also a solvent analogue of SAGD
with the same horizontal well configuration for heavy oil and bitumen recovery [69]. The oil-soluble solvents diffuse and adsorb to

K. Guo et al. / Fuel 185 (2016) 886902

the viscous hydrocarbons and thus reduce their viscosity so that


they can easily flow towards the production well and be pumped
out. In addition to being economical and eco-friendly, VAPEX is
also very suitable for heavy oil and bitumen recovery, which are
often correlated with thin pay zones, high water content, and poor
thermal conductivities of rock formations.
VAPEX was invented in 1989 by Bulter, who also proposed the
SAGD technology. Since then, VAPEX has emerged as a promising
alternative to the thermal-based heavy oil recovery methods over
the past two decades. However, VAPEX faces similar challenges
as CSI in terms of the solvent cost. The diffusivity and injectivity
of the solvents also need to be addressed to ensure the injection
efficiency. Interestingly, recovery with VAPEX is quite active in oilfields of Canada and further development of this technology can be
anticipated.
3.3. Gas injection
Gas injection is one of the oldest EOR techniques used for light
oils, while considerable interest has also been shown to the recovery of heavy oils. Basically, a readily available gas, like natural gas,
CO2, nitrogen or flue gas that exhibits certain miscibility with oil in
a compressed form, is injected into the reservoir to maintain the
pressure, under which gas saturation at the interface zone
increases. This high gas saturation can significantly generate oil
swelling, reduce the IFT between two interacting fluids and thus
improve the oil displacement efficiency. However, issues related
to gas segregation, viscous fingering or channeling, mobility control remain problematic for the gas flooding techniques. Here we
mainly discuss CO2 flooding as a representative gas injection
method with underlying mechanism, major advantages and potential challenges.
One of the most widely used fluids for miscible displacement is
CO2 due to several benefits, such as the superiority in viscosity
reduction, multiple interactions with oil, low operational cost,
and in particular the advantage for CO2 sequestration in oil reservoirs to mitigate GHG emission and climate change [71]. CO2 flooding works on principles that at the condition of certain
temperature and pressure the injecting CO2 could partially or completely dissolve in the viscous oils, which leads to reduction of the
oil viscosity and hence ease their extraction to the production well.
The high miscibility between CO2 and oil is realized by lowering
their IFT in the CO2 swept zone [72].
Unfortunately, owing to the fact that CO2 viscosity (0.01 cP) is
much smaller than that of heavy oils, CO2 flooding suffers severe
viscous fingering. The conformance and displacement front can
be hardly maintained. Easy breakthrough occurs at relatively permeable regions and consequently large amounts of recoverable oils
remain untouched. Besides, the probability of the lower density
CO2 to move upwards the reservoir, a situation known as gravity
override, further deteriorates the recovery efficiency [3].
To confront the above limitations, CO2 flooding is selectively
applied to reservoirs with depth over 800 m, where CO2 will be
in a supercritical state and the oil gravity will be larger [4]. In addition, the idea of synergistically combining chemical and gas recovery is realized by adding surfactants to generate CO2 foams so that
the viscosity of injecting fluid is significantly increased. Foam
assisted CO2 flooding has been a focused research area in recent
years, with both laboratory and field scale tests being extensively
conducted for commercial applications [7274].
As an alternative solution, water alternating gas (WAG) recovery has been initially employed to complement the gas flooding
since the late 1950s, especially in offshore Norway. WAG utilizes
intermittent slugs of water and gas to reduce gas channeling from
injector to producer thus improve sweep efficiency during gas
injection processes. Oil recovery by this process is attributed to

893

improved contact from the positive impact of both gas segregation


to the top and water accumulation at the bottom [75].
3.4. In-situ catalytic upgrading
Distinct advantages of nanotechnology have attracted continuous attempts for applications in various fields of electronics, materials, pharmaceutical, etc. The exceptional properties exhibited by
a material when its size in any dimension falls in the range of 1
100 nm have also enabled the nanomaterial to supplement the current EOR techniques. Previous results [7679] have shown that by
adding nanosized particles (silica, alumina, titania, etc.) into the
injection fluids, IFT between the fluid and oil is greatly reduced
because of the preferential distribution of particles at the fluid/oil
interface. Therefore, emulsions stabilized by the nanoparticles
(NPs), known as Pickering emulsions, are used to improve the displacement efficiency. Furthermore, NPs with different surface
functionality ranging from strongly hydrophilic (or lipophobic),
neutrally wet and strongly hydrophobic (or lipophilic) can serve
as agents to modify the wettability of reservoir rocks because they
tend to adsorb to the rock surface. Proper wettability modification
helps detach the capillary-trapped oils and thus improves the
sweep efficiency. Nevertheless, this technique is more feasible for
recovery of light oils rather than heavy counterparts. Therefore,
the underlying recovery mechanism is out of the scope in this
review.
On the other hand, the combination of nanotechnology and
catalysis has driven nanocatalysis in miscellaneous reactionrelated fields, in particular enhanced heavy oil recovery. Undoubtedly, viscosity reduction is the key to the success of in-situ recovery
of heavy and extra-heavy oils. Conventional thermal injection
methods have proved that heat-induced mechanism is effective
to improve mobility. However, inherent economic and environment concerns related to such processes are driving people in both
academia and industry to search alternative solutions. In-situ
upgrading by nanocatalysis is one potential technology that could
bring a cost-efficient process with high recovery factor. More
specifically, in-situ catalysis allows the upgrading and recovery to
occur simultaneously. With the catalytic decomposition of long
chain hydrocarbons, lighter fractions with smaller molecular
weight are formed, which implies viscosity reduction and mobility
improvement of the producing heavy oils. Although emerging as a
relatively new technology, this concept is being enormously investigated nowadays.
3.4.1. Catalysts for in-situ upgrading
In the 1980s, Hyne et al. [80] discovered that steam injection
brought not only the physical reduction of heavy oil viscosity by
temperature increase, but also the chemical reactions with some
oil components, leading to beneficial changes of the oil property
and composition. This reaction route was named aquathermolysis by these researchers. It should be emphasized that their further studies have revealed the chemical changes are reversible
because heteroatoms (S, N, and O) interact with other molecules
by van der Waals forces, which leads to polymerization and thus
the recurrence of high viscosity.
Inspired by this aquathermolysis, researchers moved forward
by introducing the idea of catalytic reaction into the reservoir.
Proper catalysts can help facilitate the chemical reactions of heavy
feeds to a large extent, leading to the effective and irreversible viscosity reduction. This in-situ catalysis thereby sheds light on the
realization of underground upgrading and enhanced recovery.
With this in mind, ongoing studies are being focused on the exploration of active, stable and reliable catalysts at laboratory and field
scales. Generally, reported catalysts for enhance recovery of heavy
oil and bitumen can be roughly classified as (1) water-soluble cat-

894

K. Guo et al. / Fuel 185 (2016) 886902

alysts, (2) oil-soluble catalysts, (3) amphiphilic catalysts, (4) minerals and zeolites, (5) solid superacids and (6) dispersed NPs.
Table 2 lists the reported catalysts in literature under different
operation conditions with corresponding test results. In most
cases, the experiments are conducted by simulating the reservoir
conditions with a reactor containing oil sample, water and catalyst
at various reaction temperature, pressure and duration. Viscosity
reduction are readily achieved in the reactor.
3.4.1.1. Water-soluble catalysts. For water-soluble, oil-soluble and
amphiphilic catalysts, they have the advantage of ensuring sufficient contact between the active species and reactants. Hyne
et al. [80] and Rivas et al. [81] observed the Ni(II) and Co(II)
salts had an impact on the aquathermolysis. Later on, Clark et al.

[8284] further confirmed this effect by systematically studying


a series of Al(III), first row transition metal Sr(III), VO(II), Cr(III),
Ni(II) and Cu(II), and Group VIIIB metal Fe(II), Co(II), Ni(II), Rh
(III), Os(III), Ir(III), Ru(III), Pt(II) and Pt(IV) salts. They concluded
that Al(III) and first row metal species were very reactive in the
aquathermolysis of thiophene and tetrahydrothiophene, two typical model compounds for organosulfur molecular types in heavy
oils. Besides, Pt(IV) is the most active metal in Group VIIIB to
reduce sulfur content, with 40% and 56% sulfur reduction for the
thiophene and tetrahydrothiophene, respectively. It is speculated
that the electron configuration, especially d orbital, should be
responsible for the good affinity of transition metal species to coordinate with the organic sulfur bonds, leading to the CAS bond
cleavage and thus reduced molecular weight. However, reactive

Table 2
Catalytic performance of various catalysts used for aquathermolysis.
Oil samplea

Catalysts

2.4 kPa

30 min
28 days
28 days

THT and TP

Ru(III) chloride

Peace River bitumen


Cold Lake bitumen
Liaohe extra-heavy oil

200
240
415

2+

SO2
4 /ZrO2
SO2
4 /ZrO2

Ni doped
Sn2+ doped
WO3/ZrO2
Ultradispersed Mo
Dispersed iron catalyst
Nano Ni catalyst
Nano Ni catalyst
Ultradispersed Ni
nanocatalyst
Ni NPs
Fe2O3/SiO2
NiWMo submicronic catalysts
Ultradispersed NiWMo NPs
Ultradispersed NiWMo NPs
FeNi/SiO2

Duration

240
240
240

Molybdenum oleate
Ni(II) and Co(II) oleate
Fe(III) naphthenate
Ni(II) organic salt
Co(II) organic salt
Aromatic sulfonic iron
Gemini catalyst
Alkyl ester sulfonate copper
Aromatic sulfonic Fe
Aromatic sulfonic Mo
Aromatic sulfonic Mo
Aromatic sulfonic Ni
Aromatic sulfonic Cu
Fe dodecylbenzene sulfonate
Ni dodecylbenzene sulfonate
Fe dodecylbenzene sulfonate
Ni dodecylbenzene sulfonate
Mineral
Natural zeolite Bogor
Nano-keggin-K3PMo12O40

Pressure

Bitumen, Eastern Venezuela


THT and TP
THT and TP

VO(II), Ni(II) and Fe(III) salt


Ni(II) sulfate

Experimental results

Temp.
(C)
Ni(II) sulfate
VO(II) salts
Al(III) sulfate
Cr(III) sulfate
Pt(IV) chloride

Fe(II) salt

Operation conditions

240

3.4 MPa

VRc (%)

3h
10
25 MPa

72 h

THN
THN
THN
DHA

240
250

24 h
24 h

240
180
340
280
280
200
350
240
200

24 h
24 h
1 week
24 h
24 h
24 h
1h
24 h
24 h

Shengli extra-heavy oil


Shengli San56-13-19 heavy oil
Shengli extra-heavy oil
Egyptian heavy crude oil

200
200
280
250
300
250

15 MPa
810 MPa
810 MPa
67 MPa
67 MPa

Toluene
THN
FA
FMM

FEG

24 h
24 h
24 h
24 h
78 MPa

240
200
280

24 h

Heavy oil
Heavy oil
Zhen411 extra-heavy oil
G540 extra-heavy oil
Shengli heavy oil

240

34 MPa

24 h

Liaohe heavy oil


Liaohe Shuguang heavy oil
Hamaca extra-heavy crude oil

220
280
410420

2 MPa
810 MPa
11 MPa

6h
24 h
1h

Liaohe extra-heavy oil


Shengli San56-13-19 heavy oil
Athabasca VR and VGO

280
200
370

6.4 MPa

24 h

Athabasca bitumen
Orinoco Basin extra-heavy oil
Athabasca bitumen
Athabasca bitumen
Athabasca bitumen
Shengli extra-heavy oil

240
280315
320380
340
320340
150

SRc (%)

Refs.
ARb (%)

90.0

7 days

Liaohe heavy oil


Shengli Shanjiasi extra-heavy
oil
Liaohe heavy oil
Liaohe heavy oil
Shengli heavy oil
Heavy oil
Liaohe heavy oil
EX35 extra-heavy oil
Fengcheng DF2005 heavy oil
Shengli extra-heavy oil
Extra-heavy oil DF32005

Shengli heavy oil

H
donorb

48 h
6h
24 h

FMM
CH4
H2

110 bar

24 h

H2

1600 psi
3.45 MPa
3.5 MPa
3.5 MPa

12 h
24 h
370 h
48 h
140 h
48 h

Decalin
THN

H2

THT = tetrahydrothiophene, TP = thiophene, VR = vacuum residue, VGO = vacuum gas oil.


THN = tetralin, DHA = dihydroanthracene, FA = formic acid, FMM = formamide, FEG = formate ester group.
VR = viscosity reduction ratio, SR = sulfur reduction, AR = asphaltene reduction.

98.6
99.2
87.0
82.6
72.5
62.6
93.2
93.6
71.6
87.0
89.5
90.7
99.3
90.7
95.6
99.3
97.2
96.3
95.5
79.0
73.0
45.3
48.5
36.9
68.5
92.3
90.0
57.7
48.9
82.2
94.1
99.7
99.8
98.3
90.4

46.0
64.0 (THT), 55.0 (TP)
62.0 (THT), 59.0 (TP)
40.0 (TP)
56.0 (THT)
36.1
31.8
66.7

62.0
55.4
66.7
87.5
42.7
86.9
87.7

48.6
64.7
15.5
18.5

7.1
1.1
22.2
66.7
14.2
22.0
73.4
20.6

[81]
[82]
[83]
[84]
45.2
29.2
46.2
23.5
44.0
39.0
30.4
55.1
44.6
58.9
63.6
23.4
55.5
26.2
6.3
8.3
7.7
43.5
37.3
27.3
25.4
1.8
11.4
8.8
24.3
6.6
95.5
4.0
22.2
55.2

15.3
34.9
43.8

81
99.2
42.8
23.2
77.2

28.0
37.5
35.3

31.0

[85]
[86]
[88]
[90]
[92]
[93]
[95]
[97]
[99]
[101]
[102]
[103]
[104]
[105]
[106]
[107]
[109]
[110]
[113]
[116]
[119]
[121]
[122]
[124]
[125]
[128]
[131]
[132]
[135]
[139]
[141]
[143]
[144]
[146]

K. Guo et al. / Fuel 185 (2016) 886902

intermediates are also formed with the tendency for polymerization reactions unless there is external hydrogen to stabilize them
[85]. Accordingly, Zhong et al. [86] investigated the effect of hydrogen donor on the catalytic activity. With the addition of tetralin as
a hydrogen donor, viscosity reduction is improved to 90% compared to 60% without tetralin at 240 C for 3 days. Ovalles et al.
[87] reported that the direct use of methane as a hydrogen source
improved the upgrading of extra-heavy oils. These remarkable
improvements highlight the importance of side reaction inhibition
by utilizing external hydrogen additives, which is also confirmed
by other studies [8890]. Moreover, Chens group [91] concluded
that different metal ions had their distinctive roles in catalyzing
the aquathermolysis reactions by investigating Fe3+ and Cu2+ ions.
3.4.1.2. Oil-soluble catalysts. The limitation that water-soluble catalysts cannot effectively mix with the oil phases inspires the development of oil-soluble catalysts. Wen et al. [92] synthesized
molybdenum oleate as an oil soluble catalyst and experimental
results showed that with 0.5 wt.% catalyst the viscosity reduction
of heavy oils could be up to 90%. Furthermore, this catalyst was
proved effective to reduce the viscosity at field tests at Liaohe oilfield in China. Zhao et al. [93] studied Ni(II) and Co(II) oleate in the
presence of petroleum sulfonate as an emulsifier and toluene as a
hydrogen donor. They demonstrated that the thermal stability and
durability of this catalyst as well as its effectiveness in viscosity
reduction were better than that of water-soluble catalysts. Impressively, with a low content of 0.4 wt.%, the catalyst reduced the viscosity by 90% at 180 C and sulfur content decreased by 87.5%.
Besides, Ni(II) and Fe(II) naphthenate [94,95], organic Ni(II) and
Co(II) salts [9699] and Fe(III) tris(acetylacetonate) [100] catalysts
with various hydrogen donors of tetralin, formamide, and formic
acid have also achieved remarkable performance in the reduction
of heavy oil viscosity. Therefore, it is commonly accepted that
oil-soluble catalysts show better catalytic activity than that of
water-soluble catalysts [94].
3.4.1.3. Amphiphilic catalysts. Even though oil-soluble catalysts
have shown prominent activity through better contact, they still
suffer from the problem of catalyst separation from the oil phase,
which leads to insufficient utilization of the catalytically active
metal species. Therefore, an amphiphilic catalyst that combines
the benefits of water-soluble and oil-soluble ones is proposed.
Chens group [101] designed aromatic sulfonic iron composed of
an active metal cation and an amphiphilic anion. This amphiphilic
anion improved the dispersion of cations into the oil and helped
the catalyst remain stable at the water-oil interface. Accordingly,
an apparent viscosity reduction of 90.7% at 200 C was realized;
field tests also achieved eminent enhancement with stable viscosity reduction of up to 82.3%. The same group [102] also synthesized
a new type of Gemini catalyst with transition metal as the active
center and Gemini surfactant as the ligand. Over 90% viscosity
reduction at both laboratory and field scale tests was achieved at
a relatively low temperature of 170 C. An interesting study was
performed by Chao et al. [103] by combination of hydrogen precursor and active species. The authors utilized alkyl ester sulfonate
copper as a bifunctional catalyst and observed 90.7% viscosity
reduction after 24 h using 0.3 wt.% catalyst at 240 C. They
declared that this eco-friendly catalyst could be prospective for
field applications. Additionally, several chelates [104108] and
metal dodecylbenzenesulfonates [109,110] are also reported with
superior performance. However, the temperature-dependent activity needs further validation taking into consideration that the temperature in the reservoirs will normally be lower. It is thus
desirable to design and develop catalytic reactions that can proceed at moderate conditions.

895

3.4.1.4. Minerals, zeolites, and solid superacids. Another important


type of catalysts are the heterogeneous ones including minerals
and zeolites, solid superacids and dispersed NPs. Mineral is an ideal
catalyst due to its natural abundance in the reservoir. This concept
was first testified by Monin and Audibert [111] who found that
seven different types of minerals all behaved similarly by promoting the cracking reactions of heavy crude oils. Fan et al. [112,113]
and Ovalles et al. [89] further observed a synergetic effect of minerals and steam on the aquathermolysis of heavy oils. Natural zeolites are economic and abundant minerals. When activated
properly to possess certain acidity, they also showed considerable
activity in heavy oil viscosity reduction. Junaid et al. [114] concluded that natural zeolite was effective in heteroatom removal
by examining chabazites and clinoptilolites as catalysts. Researchers [114117] also found that the porous structure enabled the
zeolites to adsorb unwanted components and the acidic properties
offered large amounts of hydrogen ions to stabilize the cracked
intermediate molecules, which underwent an ionic mechanism different from the free-radical mechanism of thermal hydrocracking.
Similarly, solid superacids with strong acidity have also been
employed as catalysts. Strausz et al. [118] reported that tetrafluoroboric acid (HFBF3) as an active catalyst led the reactions to proceed by depolymerization and hydrogenation with the main
products of volatile organic compounds and alkylbenzenes. Results
from this study triggered the investigation of other heteropoly
acids and modified zirconia as alternatives since the water solubility of HFBF3 poses recovery problems. Chen et al. [119,120] discovered that both nano-keggin K3PMo12O40 and H3PMo12O40 showed
apparent large viscosity reduction due to their unique properties
in acidity, redox and pseudo-liquid phase reaction field. Significant
changes of oxygen-containing components were observed with
effective cracking or dissolving. In addition, zirconia activated by
sulfate groups or metal oxides are also widely used superacids.
Lis group [121] demonstrated that SO2
4 /ZrO2 superacid doped
with Ni2+ or Sn2+ catalyzed the visbreaking of Shengli heavy oils.
The content of resin, asphaltene, sulfur and nitrogen decreased significantly. Wang et al. [122] compared WO3/ZrO2 as a catalyst
obtained by hydrothermal and conventional impregnation methods in the absence of water. It was proved that WO3/ZrO2 prepared
by hydrothermal method had better-dispersed WO3 and more acid
sites than that by the other methods. This enhanced acidity was
correlated with the increased catalytic activity and the absence
of water did not affect the pyrolytic stability of heavy crude oil.
Although the studied catalysts have displayed satisfactory
activities in the upgrading of heavy oil and bitumen, issues regarding catalyst recycling, deactivation and applicability are still troublesome for commercial implementation. Specifically, watersoluble catalysts face shortcomings in the inadequate contact
between catalysts and the oil phase, whereas oil-soluble catalysts
pose the problem of catalyst separation. The catalytic activity of
most minerals is insufficient, or even negligible, and the injection
of mineral particles is practically difficult. In addition, high activities of solid acids at low reservoir temperatures are problematic. To
operate at relatively mild conditions, it is imperative to continue
looking for highly active and durable catalysts. Among the reported
catalysts, nanomaterials have drew increasing attention recently.
Owing to the unique properties of the nanoscale materials, such
as high inherent catalytic activity, high surface area to volume
ratio, efficient transport inside porous rocks and controllable synthesis for specific functionalities, nanocatalysts are being developed to catalyze the in-situ chemical reactions in reservoirs, in
which heavy oil and bitumen are converted to lighter products that
meet pipeline and refinery requirements.
3.4.1.5. Metallic NPs. Metallic NPs have long been used in the
upgrading of heavy oil and bitumen in petroleum refineries, but

896

K. Guo et al. / Fuel 185 (2016) 886902

not in the subsurface reservoirs. Until now, various transition metals and oxides such as Mo [123,124], Fe [125127], Ni [128133],
Cu [134,135], Fe2O3 [136139], CuO [126,140] and alloys [141
146] based catalysts have been reported. For example, Li et al.
[128] prepared nano-Ni microemulsion and applied them in the
aquathermolysis of extra-heavy oil. The study proved that Ni promoted visbreaking, sulfur removal and asphaltene conversion and
the synergetic effects of upgrading, emulsification and diluting
resulted a high viscosity reduction of 98.9%. Hashemi et al. [143
145] reported microemulsions containing trimetallic (W, Ni, and
Mo) colloidal nanoparticles as catalysts, and their injectivity in a
packed bed reactor along with enhanced bitumen recovery was
demonstrated. To clarify the underlying mechanism, Babadaglis
[126] and Ovalless [139] groups explored the influence of metal
type, size and concentration of NP on the catalytic activity. They
both recognized that these parameters were very important and
should be optimized to achieve better overall performance. The
high thermal conductivity of metallic nanoparticles can also
improve the energy efficiency. Accordingly, it can be proposed that
metallic NP assisted in-situ heavy oil recovery is a new and promising alternative to supplement conventional thermal methods
although substantial research and development need to be conducted for the commercial field application.
3.4.2. Mechanism
The viscosity reduction of heavy oils relies on two pathways,
which are by virtue of either redistributing internal hydrogen
atoms or introducing external hydrogen donors. For the former
one, hydrogen is migrated from one part of heavy hydrocarbons
to the others, which means upgrading would occur together with
potential formation of detrimental coke. It is therefore reasonable
to introduce external hydrogen sources, such as water, tetralin
and CH4, to avoid coke formation. However, the hydrotreating processes are kinetically sluggish such that active catalysts should be
applied. The incorporation of catalysts can facilitate the
hydrotreating reactions to proceed at mild conditions and eventually turn the reservoirs into subsurface reactors.
As mentioned previously, heavy oil and bitumen are heavy
hydrocarbons featured with high C/H ratios and considerable
amounts of N, S, O and metal elements. When steam is injected,
heat energy is supplied to break large molecules into smaller and
lighter ones. Hyne et al. [80] proposed the following reaction
mechanism in aquathermolysis:

RCH2 CH2 SCH3 2H2 O RCH3 CO2 H2 H2 S CH4


The cleavage of CAS bond results in not only reduced molecular
size and sulfur removal but also gaseous products. Early studies
have confirmed the production of these gases. More interestingly,
the comparative analysis of Canadian and Venezuelan heavy crude
oils reveals that sulfur content is the key factor in aquathermolysis
and the viscosity reduction largely depends on the CAS bond
breakage. However, the aquathermolysis of the Liaohe heavy crude
oil with low sulfur content (<0.5%) in Northeast China is also
demonstrated. This controversial mechanism indicates the complexity of the chemical reactions involved in the aquathermolysis
process. Typically, several favorable hydroprocessing or
hydrotreating reactions, i.e., hydrocracking (HCK), hydrodesulfurization (HDS) hydrodenitrogenation (HDN), hydrodeoxygenation
(HDO) and hydrodemetallization (HDM), are involved with a series
of CAS, CAN, CAO, CAC, C@S, C@O, C@N and C@S bonds in SARA
fractions [103,106108]. Due to the fact that resins and asphaltenes are the main components accounting for the high viscosity,
increasing saturate and aromatic contents while decreasing resin
and asphaltene contents are always anticipated by the effective
upgrading and recovery.

HCK entails the theoretical feasibility for this underground


refinery. Nonetheless, it is only thermodynamic favorable at high
temperatures. An active catalyst needs to be introduced to lower
the activation energy barrier. Catalytic HCK will be a two-stage
process combining cracking and hydrogenation processes promoted by catalysts. Cleavage of CAC and CAR (R = S, N, and O)
bonds induced by the coordination effect of transition metal atoms
is the main route for the exposed intermediate carbon radicals to
adsorb free hydrogen atoms. Among all the chemical bonds existing in hydrocarbon molecules, CAS bond is found to possess the
lowest bond dissociation energy, which means that CAS bond
would be the most catalytically active [147]. Considering that the
asphaltene fraction contains percentage levels of sulfur, CAS bond
cleavage is of paramount importance in asphaltene removal. Furthermore, several studies have shown that heteroatom (N, O and
S) containing groups in the heavy oil could cause polymerization
with the formation of hydrogen bonds, adversely leading to the viscosity regression. In-situ sulfur removal could also alleviate the
burden of the desulfurization process in refineries. Accordingly,
the HDS, HDN, HDO and HDM reactions, in particular the HDS, play
an important role in upgrading heavy oils to high quality crude
oils.
The introduction of catalysts can also provide additional reaction pathways for the viscosity reduction. Wang et al. [104]
observed isomerization, ring opening, oxygenation, alcoholization,
and esterification, and reconstruction reactions when using Fe3+
and Mo6+ as catalytic ions. The same group also conducted a study
[91] to show that Cu2+ ions mainly caused the depolymerization
and cleavage of some bridge bonds of the macromolecular ring system, whereas Fe3+ ions primarily led to the isomerization of side
chains and ring-opening of heterocyclic rings. Wangs group
[122] suggested that with the aid of superacid WO3/ZrO2, the presence of water was not required. Galarraga and Almao [141]
reported that NiWMo submicronic catalysts reduced both the viscosity and microcarbon residue (MCR). Impressively, Shokrlu et al.
[126] observed the viscosity reduction once the metal NPs were
mixed with oil samples at low temperatures owing to the exothermic chemical reactions. This result underlines the potential of utilizing highly active NPs as in-situ catalysts in heavy oil upgrading
and recovery. Furthermore, they have also proposed another beneficial factor of superior thermal conductivity of metals since efficient heat transfer is crucial in the thermal recovery methods.
Most of the current investigations are limited to the simulated
reservoir conditions in laboratory, the multiple beneficial functions
of metallic NPs entitle this promising technology to be further
investigated and developed. As a result, the following section will
focus on the application of metallic NPs as in-situ catalysts.
3.4.3. Catalyst preparation
Synthesis of metallic nanomaterials is enriched along with the
fast advancement of nanotechnology in multidisciplinary applications and is already a well-established field. Precise manipulation
of the size, shape, composition, type, and surface functionality
has all been possible by the selective control of synthesis parameters. Generally, synthetic approaches can be divided into the topdown and bottom-up methods. Top-down methods are normally
done with physical decomposition of bulk materials into nanoscale
materials, such as milling and grinding. Conversely, bottom-up
methods involve the initial chemical reactions of molecular or
atomic precursors, and subsequent nucleation and growth to the
formation of nanomaterials. It is widely accepted that the chemical
approaches are more suitable for the controlled synthesis of multifunctional nanomaterials.
Among different chemical synthetic approaches, the
microemulsion method is widely used for the preparation of
metallic NPs in heavy oil phase. Typically, a W/O microemulsion

K. Guo et al. / Fuel 185 (2016) 886902

897

Fig. 6. Nanoparticle-assisted in-situ recovery: nanoparticle suspension and heat source are injected into the reservoir; these nanoparticles serve as catalysts to decompose
heavy hydrocarbons into lighter products, which eases the flow towards the production well.

is prepared by mixing water, oil (e.g., vacuum gas oil) and an emulsifying agent (e.g., surfactants). A solution containing corresponding salt precursors is then added and blended uniformly with the
W/O emulsion. Subsequently, another solution, precipitation agent
or reduction agent, is added to activate the nucleation and growth
of NPs, which exist inside individual water droplets that are covered by emulsifier and suspended steadily in the oil phase. Besides,
surface-modified silica can also be introduced to support the dispersed NPs with the formation of stable suspension [127,139,146].
3.4.4. Structural parameters
Structureactivity relationship is always a key point in the functional design and synthesis of NPs. The effect of metal type, size,
shape, composition, dispersity, concentration and support on the
activity should be elucidated to maximize the catalytic activity,
which will further help the NP-assisted recovery method be
adapted to the harsh reservoir conditions.
Among the transition metals, Ni is regarded as the most promising one because of its high activity and low cost. Shokrlu and Babadagli [126] studied the Fe, Cu, Ni, Fe2O3 and CuO NPs with different
nano and micron sizes as viscosity reducers and identified the optimal variables such as metal type, size and concentration. They concluded that optimal concentration of NPs strongly relied on the
heavy oil composition, especially the asphaltene content. Proper
selection of metal type and size should also be based on the composition analysis. Wang et al. [122] declared that diffusion of WO3
onto the ZrO2 was improved by hydrothermal method. The welldispersed WO3/ZrO2 offered additional activity enhancement. In
addition, the importance of catalyst composition and synergetic
effect of metal alloying is discovered by Nassars group [144,145].
They assembled trimetallic (WAMoANi) NPs with a specific metal
ratio and stated that Mo was used as a hydrogenolysis metal, while
W was inserted to improve the hydrogenating activity of the
trimetallic catalyst and Ni acted as a promoter. It is noteworthy
that exposed facet of NPs is also a factor that deserves further
study because it is well know that different crystal planes exhibit
different catalytic activities.
3.4.5. Transport behavior
The majority of studies of NPs as catalysts are performed to
evaluate the catalytic activity in absence of porous matrix. However, to unlock the heavy underground feeds, assessment of NP

mobility inside the porous reservoir rocks is also a critical factor.


Successful application of nanocatalysts in reservoirs would largely
depend on the availability of NPs to get contact with heavy hydrocarbon molecules effectively. Catalysts should theoretically be able
to flow through the submicron and micron-sized channels that are
filled with heavy oil. A typical scheme is represented in Fig. 6.
However, the high surface energy of NPs boosts the spontaneous
agglomeration, precipitation and adsorption, causing the undesired
deactivation of certain active sites. Therefore, the transport behavior of NPs should be monitored and controlled to maximize the catalytic performance.
Pioneering work was done by Zamani et al. [148] who claimed
that it was possible for the propagation of ultradispersed submicron particles suspended in oil phase through the sand beds.
Unfortunately, part of the particles were retained by the sand,
especially at the injection entrance. Irreversible retention of particles was attributed to the surface attachment deep inside the sand
and the mechanical entrapment at the inlet. Later, Nassars group
[142145] conducted a series of experiments to discuss the transport feasibility of multimetallic NPs in detail. At typical SAGD conditions of 3.5 MPa and 300320 C, breakthrough was achieved for
both low- and high-permeability oil sands packed columns. Moreover, metal type, temperature and oil sands porosity could affect
the transport of NPs, but neither major permeability damage nor
pore plugging were observed. Although the transport of NPs in
reservoir is feasible, it should be noted that there still lacks basic
understanding of NPs behavior in reservoir rocks. In addition, modelling of NPs behavior in porous media is too complex to be established with adequate accuracy.
4. Quality enhancement
Based on the properties of heavy and extra-heavy oils, extensive
approaches and apparatus have been developed to characterize the
important aspects of the upgraded oil quality.
4.1. Viscosity and C/H ratio
Viscosity is considered as one of the major factors for heavy oil
to meet strict refinery and pipeline specifications. This physical
property depicts its resistance to shearing flows, which can be read
directly on a viscometer or rheometer at a certain temperature.

898

K. Guo et al. / Fuel 185 (2016) 886902

Viscosity data of oil sample before and after the catalytic tests will
be compared. Similarly, the change of C/H ratio can also be measured as an indication of the extent of upgrading. Lighter products
produced from the hydrogenation of heavy molecules have lower
C/H ratios.
4.2. Coke formation
In a thermal cracking process without external hydrogen addition, the original hydrogen will be redistributed in the feedstock,
which means the generation of light components with decreased
C/H ratio as well as heavier coke. Especially when temperature
reaches above 300 C, pyrolysis of heavy feeds will become the predominant reaction mechanism instead of aquathermolysis. Accordingly, MCR could be measured as another indicator for the extent of
heavy oil upgrading, i.e., products with improved quality will possess a lower content of coke formation. Furthermore, it is suggested that the inhibition of coke formation is closely associated
with quality enhancement. Because of the improved conversion
of hydrogen atom to free radicals by injected catalysts, the transfer
of hydrogen to combine intermediate products is enhanced and
thus coke formation is largely depressed. Researchers also proposed that the presence of NPs could bring more reaction pathways for intermediate radicals.
4.3. Heteroatoms content
S, N, O and small amount of heavy metals are common heteroatoms in heavy oil and bitumen. The heteroatom removal is necessary in any oil refining processes due to the strict environmental
regulations on transportation fuels. A series of HDS, HDN, HDO
and HDM reactions are accelerated due to the favorable activities
of the catalysts. In particular, the effectiveness of sulfur removal
has been confirmed in many previous studies by examining the
sulfur content before and after catalyst addition.
4.4. Composition analysis
To understand thoroughly the quality enhancement, it is natural to characterize the composition of the oil samples. Given that
the heavy oil and bitumen are complicated hydrocarbon mixtures,
chromatography technique is widely used to analyze the gas and
liquid products. For example, thin layer chromatography is capable
to calculate the respective content of SARA components and column chromatography can separate them as relatively pure components for further detailed analysis. Moreover, gas chromatography
(GC) can give information of the gas products including H2S, CO2
and light hydrocarbons (C1C4). Improved GC columns can be used
to perform simulated distillation analysis covering the hydrocarbon range of C5C100. Their fractions in the oil sample can be determined accurately, which gives a direct observation of the
composition distribution after the catalyst addition. Other physical
methods such as nuclear magnetic resonance, spectroscopy analysis are useful tools to elucidate the corresponding reaction
mechanism.
5. Promises and challenges
Compared to the conventional thermal, chemical and gas injection methods, in-situ catalytic recovery of heavy oil and bitumen
has the advantages of (1) proceeding at relatively mild operation
conditions; (2) upgrading the oils to great extent with reduced
waste product production; (3) combination of upgrading and
recovery inside the reservoir to minimize the requirements for surface upgrading facilities; (4) reduce economic and environmental

costs for heavy oil production; and (5) applicability for both
medium- and extra-heavy oils in different reservoir configuration
by tailoring the catalysts. Here we discuss practical issues to be
considered from the industrial application perspective, through
comparison with conventional recovery methods. Future research
directions for improving the in-situ catalytic upgrading with potential challenges are also addressed.
5.1. Formation damage
Bitumen and oil sands normally exist in relatively shallow
reservoirs. Therefore, in-situ upgrading of these oil should be operated at a relatively low pressure, well below the reservoir fracture
pressure, to ensure containment [11]. For safety issues at normal
depths at which oil sands deposits in Canada are found, a pressure
limitation of maximum 5 MPa is chosen [11]. In this case, if steam
injection methods are utilized, workable temperature is limited to
280 C. However, the incorporation of catalysts has the potential to
decrease the temperature, alleviating the risk of formation rupture.
A big challenge during the production of heavy oil and bitumen
is the precipitation of active asphaltenes within the porous matrix,
leading to detrimental pore plugging. Asphaltene precipitation via
mechanical or chemical processes can reduce the effective permeability of liquid phase and hence lower the recovery factor. Therefore, finding an inhibitor to prevent or delay their precipitation or
converting the unwanted hydrocarbons is of great importance. Due
to the high surface area of nanocatalysts, they can adsorb asphaltene molecules and stabilize these molecules in the oil phase with
improved mobility. Moreover, catalytic HDS reaction can contribute to the conversion of asphaltenes to light products. By integrating of nanofluids containing SiO2, Al2O3 or TiO2 supports with
tuned HLB, it is possible to modify the wettability of reservoir rocks
to water wet, which helps the detachment of strongly adsorbed oil
molecules.
Despite these merits, in-situ catalysis is still challenged by the
possibility of catalyst retention inside the formation. Actually, it
is likely that injected NPs might either agglomerate into larger particles or adsorb to the rock surface especially at the harsh temperature and pressure. To address these issues, ultradispersed
suspension with high stability and selectivity should be prepared.
5.2. Economical cost
The conventional thermal methods involve large capital investment in the steam generation, injection and recycling facilities. In
order to control costs, petroleum engineers are working continuously with attempts of replacing the heat supply by costeffective solar-generated steam [149], microwave [150] and electrical current [151]. For the same purpose, catalysts have also been
integrated in conventional CSS and ISC methods. Farooqui et al.
[152] investigated the effect of Ni NPs on the recovery factor of a
simulated CSS process. In one test, they found that the Ni suspension led to a further increment of oil recovery after 5 cycles of CSS,
which otherwise could not be achieved by only injecting steam.
The produced oil was also in better quality than that without Ni
NPs. Rezaeis group [153] performed a study on the ISC to discuss
the thermocatalytic conversion of crude oil in the presence of
NPs. They concluded that NPs could increase both the amount
and reactivity of deposited fuel at the combustion front and could
catalyze the high temperature combustion reactions to produce
high-quality products. These results indicate the great potential
in combining the in-situ catalysis with thermal injection to reduce
the capital investment.
Hydrogen donor is necessary for an ideal catalytic upgrading
with HCK, HDS, HDO, HDN and HDM reactions, which entails that
sufficient hydrogen sources should be supplied. For an efficient

K. Guo et al. / Fuel 185 (2016) 886902

899

operation, natural gas can be used as one option since thermal


techniques rely mostly on natural-gas-fired steam generation.
However, it should be pointed out that handling these flammable
gases at high pressure and temperature conditions also arouses
safety issues. It is therefore necessary to look for hydrogen donors
with high abundance, low price and good accessibility. Another
part of investment comes from the metal-based catalysts. Interestingly, recent work by Cuis group [147] raised the idea of utilizing
carbon nanomaterials as alternative catalysts. With this carbon
nanocatalyst, a high viscosity reduction of over 96% in short reaction time is obtained at a low temperature of about 150 C. The
replacement of transition metals with carbon nanomaterials will
be helpful to reduce the operating cost.

improved quality. Remarkable performance have been reported


with several types of catalysts, among which the metallic NPs exhibit the best activity at relatively mild temperatures and pressures.
Moreover, the feasibility of successful transport of NPs through
porous matrix is demonstrated although particle retention is also
observed.
Through the comparison of in-situ catalytic upgrading with conventional recovery methods, we have highlighted the benefits as
well as the challenges faced with in-situ catalysis. So far, a lot of
research have been devoted to the development of catalysts with
high activity, stability and reliability, but most of the studies are
limited to laboratory study and pilot tests. Considerable efforts
are necessary for successful commercial applications.

5.3. Environmental impact

Acknowledgement

Undoubtedly, if in-situ catalysis is successfully combined with


the thermal injection methods, the environmental footprint for
heavy oil production will be reduced. For instance, in Hashemis
study [145] it is evidenced that when metallic NPs are added,
GHG emission, especially CO2, is reduced by 50% compared to the
case without NPs. However, in the same experiment, the overall
gas production, including CO2, CO, H2S and hydrocarbon gases, is
almost doubled in the presence of NPs, suggesting that more H2S
might have been produced. Therefore, this toxic gas should be
avoided or be converted to unharmful products. Interestingly,
Montgomery et al. [154] revealed that H2S generation only occurred
within a specific temperature and pressure window. For high sulfur
containing heavy oils, the generation of H2S was minimal if the process was controlled within specific operation window.

The authors would like to acknowledge the support from the


National IOR Centre of Norway at University of Stavanger.

6. Conclusions
With the increasing attention in the exploitation and production of unconventional resources, e.g., heavy oil and bitumen, this
paper provides a comprehensive review on the current and
prospective heavy oil recovery methods. The properties of heavy
and extra-heavy oils are characterized by high viscosity or poor
mobility, long and large molecules with high C/H ratios, and considerable amounts of S, N, O and heavy metals. These heavy hydrocarbons are mainly composed of saturates, aromatics, resins and
asphaltenes.
Thermal injection is the most widely used method for the in-situ
recovery of heavy oil and bitumen. The heat-induced viscosity
reduction effectively improves the mobility of heavy feeds and
enhances their flow. However, challenging issues in steam generation, GHG emission and water management are limiting the practical application of this technology. Chemical injection has the
capability to enhance the oil mobility by reducing the IFT with
the formation of W/O or O/W emulsions and by increasing the
water viscosity with improved mobility control, but it is only suitable for reservoirs with high permeability. Gas injection is
regarded as an eco-friendly technology because it utilizes abundant CO2 or CH4 to extract the heavy oils. Nonetheless, the big viscosity difference between the gas and heavy oil results in viscous
fingering thus poor sweep and displacement efficiencies. For this
reason, gas injection is hardly applicable to the recovery of bitumen or oil sands.
By combination of nanotechnology and catalysis, in-situ catalytic upgrading has emerged as a promising technology that turns
the reservoir into a subsurface refinery to implement the upgrading and recovery simultaneously. Aquathermolysis is proposed to
account for a successful upgrading. In a typical catalytic aquathermolysis process, a series of HCK, HDS, HDO, HDN and HDM reactions are facilitated to a greater extent, generating products with

References
[1] Countries OotPE. World oil outlook 2014. Vienna (Austria): OPEC; 2014.
[2] Dusseault MB, Shafiei A. Oil sands. Ullmanns encyclopedia of industrial
chemistry. Wiley-VCH Verlag GmbH & Co. KGaA; 2000.
[3] Huc AY. Heavy crude oils: from geology to upgrading: an overview. Paris
(France): Editions Technip; 2011.
[4] Speight JG. Enhanced recovery methods for heavy oil and tar sands. Houston
(Texas): Gulf Publishing Company; 2009.
[5] Speight JG. Heavy and extra-heavy oil upgrading technologies. USA: Gulf
Professional Publishing; 2013.
[6] Shah A, Fishwick R, Wood J, Leeke G, Rigby S, Greaves M. A review of novel
techniques for heavy oil and bitumen extraction and upgrading. Energy
Environ. Sci. 2010;3(6):70014.
[7] Alvarado V, Manrique E. Enhanced oil recovery: an update review. Energies
2010;3(9):152975.
[8] Zhao DW, Wang J, Gates ID. Thermal recovery strategies for thin heavy oil
reservoirs. Fuel 2014;117:43141.
[9] Maity SK, Ancheyta J, Marroqun G. Catalytic aquathermolysis used for
viscosity reduction of heavy crude oils: a review. Energy Fuels 2010;24
(5):280916.
[10] Muraza O, Galadima A. Aquathermolysis of heavy oil: a review and
perspective on catalyst development. Fuel 2015;157:21931.
[11] Almao PP. In situ upgrading of bitumen and heavy oils via nanocatalysis. Can J
Chem Eng 2012;90(2):3209.
[12] Hashemi R, Nassar NN, Almao PP. Nanoparticle technology for heavy oil insitu upgrading and recovery enhancement: opportunities and challenges.
Appl Energy 2014;133:37487.
[13] Zhao DW, Wang JJ, Gates ID. An evaluation of enhanced oil recovery strategies
for a heavy oil reservoir after cold production with sand. Int J Energy Res
2015;39(10):135565.
[14] Thomas S. Enhanced oil recovery an overview. Oil Gas Sci Technol 2007;63
(1):919.
[15] de Haan HJ, van Lookeren J. Early results of the first large-scale steam soak
project in the Tia Juana field, Western Venezuela. J Petril Technol 1969;21
(1):10110.
[16] Cokar M, Gates ID, Kallos MS. Reservoir simulation of steam fracturing in
early-cycle cyclic steam stimulation. SPE Reserv Eval Eng 2013;15(6):67687.
[17] Jiang Q, Thornton B, Russel-Houston J, Spence S. Review of thermal recovery
technologies for the clearwater and lower grand rapids formations in the Cold
Lake area in Alberta. J Can Petrol Technol 2013;49(9):213.
[18] Alvarez J, Han S. Current overview of cyclic steam injection process. J Petrol
Sci Res 2013;2(3):11627.
[19] Butler RM, Yee CT. Progress in the in situ recovery of heavy oils and bitumen. J
Can Petrol Technol 2002;41(1):3140.
[20] de Souza JC, Cursino DFS, Padua KG. Twenty years of vapor injection in heavyoil fields. SPE Latin American and caribbean petroleum engineering
conference. Rio de Janeiro (Brazil): Society of Petroleum Engineers; 2005.
[21] Hanzlik EJ, Mims DS, Padua KG. Forty years of steam injection in california
the evolution of heat management. SPE international improved oil recovery
conference in Asia Pacific. Kuala Lumpur (Malaysia): Society of Petroleum
Engineers; 2003.
[22] Butler RM. A new approach to the modelling of steam-assisted gravity
drainage. J Can Petrol Technol 1985;24(3):4251.
[23] Butler RM, Stephens DJ. The gravity drainage of steam-heated heavy oil to
parallel horizontal wells. J Can Petrol Technol 1981;20(2):906.
[24] Yang L, Zhou DS, Sun YH. SAGD as follow-up to cyclic steam stimulation in a
medium deep and extra heavy oil reservoir. International oil & gas conference
and exhibition in China. Beijing (China): Society of Petroleum Engineers;
2006.

900

K. Guo et al. / Fuel 185 (2016) 886902

[25] Mendoza HA, Finol JJ, Butler RM. SAGD, pilot test in Venezuela. Latin
American and Caribbean petroleum engineering conference. Caracas
(Venezuela): Society of Petroleum Engineers; 1999.
[26] Jimenez J. The field performance of SAGD projects in Canada. In: International
petroleum technology conference, international petroleum technology
conference, Kuala Lumpur, Malaysia; 2008.
[27] Grills TL, Vandal B, Hallum F, Trost P. Case history: horizontal well SAGD
technology is successfully applied to produce oil at LAK Ranch in Newcastle
Wyoming. SPE international thermal operations and heavy oil symposium
and international horizontal well technology conference. Calgary (Alberta,
Canada): Society of Petroleum Engineers; 2002.
[28] Kamath VA, Sinha S, Hatzignatiou DG. Simulation study of steam-assisted
gravity drainage process in Ugnu tar sand reservoir. SPE western regional
meeting. Anchorage (Alaska): Society of Petroleum Engineers; 1993.
[29] Ashrafi M, Souraki Y, Torsaeter O. Numerical simulation study of field scale
SAGD and ES-SAGD processes investigating the effect of relative
permeabilities. Energy Environ Res 2013;3(1):93105.
[30] Elliott KT, Kovscek AR. Simulation of early-time response of single-well steam
assisted
gravity
drainage
(SW-SAGD).
SPE
western
regional
meeting. Anchorage (Alaska): Society of Petroleum Engineers; 1999.
[31] Akin S, Bagci S. A laboratory study of single-well steam-assisted gravity
drainage process. J Petrol Sci Eng 2001;32(1):2333.
[32] Kisman KE, Yeung KC. Numerical study of the SAGD process in the Burnt Lake
oil sands lease. SPE international heavy oil symposium. Calgary (Alberta,
Canada): Society of Petroleum Engineers; 1995.
[33] Gates ID, Chakrabarty N. Optimization of steam assisted gravity drainage in
McMurray reservoir. J Can Petrol Technol 2013;45(9):5462.
[34] Nasr TN, Beaulieu G, Golbeck H, Heck G. Novel expanding solvent-SAGD
process ES-SAGD. J Can Petrol Technol 2013;42(1):136.
[35] Govind PA, Das SK, Srinivasan S, Wheeler TJ. Expanding solvent SAGD in
heavy oil reservoirs. International thermal operations and heavy oil
symposium. Calgary (Alberta, Canada): Society of Petroleum Engineers; 2008.
[36] Butler R. The steam and gas push (SAGP). J Can Petrol Technol 2013;38
(3):5461.
[37] Chu C. A study of fireflood field projects. J Petrol Technol 1977;29(2):11120.
[38] Cheih C. State-of-the-art review of fireflood field projects. J Petrol Technol
1982;34(1):1936.
[39] Turta AT, Chattopadhyay SK, Bhattacharya RN, Condrachi A, Hanson W.
Current status of commercial in situ combustion projects worldwide. J Can
Petrol Technol 2007;46(11):814.
[40] Wu CH, Fulton PF. Experimental simulation of the zones preceding the
combustion front of an in-situ combustion process. Soc Petrol Eng J 1971;11
(1):3846.
[41] Castanier LM, Brigham WE. Upgrading of crude oil via in situ combustion. J
Petrol Sci Eng 2003;39(12):12536.
[42] Mahinpey N, Ambalae A, Asghari K. In situ combustion in enhanced oil
recovery (EOR): a review. Chem Eng Commun 2007;194(8):9951021.
[43] Xia TX, Greaves M. Upgrading Athabasca tar sand using toe-to-heel air
injection. J Can Petrol Technol 2002;41(8):517.
[44] Xia TX, Greaves M, Turta AT, Ayasse C. Thaia short-distance displacement
in situ combustion process for the recovery and upgrading of heavy oil. Chem
Eng Res Des 2003;81(3):295304.
[45] Greaves M, Xia TX, Turta AT, Ayasse C. Recent laboratory results of THAI and
its comparison with other IOR processes. SPE/DOE improved oil recovery
symposium. Tulsa (Oklahoma): Society of Petroleum Engineers; 2000.
[46] Thomas S, Farouq Ali SM. Status and assessment of chemical oil recovery
methods. Energy Sources 1999;21(12):17789.
[47] Homsy G. Viscous fingering in porous media. Annu Rev Fluid Mech 1987;19
(1):271311.
[48] Zolotukhin AB, Bokserman A, Kokorev V, Nevedeev A, Ushakova A,
Shchekoldin K. New upstream and downstream technologies for extra
heavy oils. SPE heavy oil conference Canada. Calgary (Alberta,
Canada): Society of Petroleum Engineers; 2012.
[49] Dehghan AA, Masihi M, Ayatollahi S. Interfacial tension and wettability
change phenomena during alkalisurfactant interactions with acidic heavy
crude oil. Energy Fuels 2015;29(2):64958.
[50] Dong M, Ma S, Liu Q. Enhanced heavy oil recovery through interfacial
instability: a study of chemical flooding for Brintnell heavy oil. Fuel 2009;88
(6):104956.
[51] Zhang H, Dong M, Zhao S. Which one is more important in chemical flooding
for enhanced court heavy oil recovery, lowering interfacial tension or
reducing water mobility? Energy Fuels 2010;24(3):182936.
[52] Mohammed M, Babadagli T. Wettability alteration: a comprehensive review
of materials/methods and testing the selected ones on heavy-oil containing
oil-wet systems. Adv Colloid Interface Sci 2015;220:5477.
[53] Fletcher PD, Savory LD, Woods F, Clarke A, Howe AM. Model study of
enhanced oil recovery by flooding with aqueous surfactant solution and
comparison with theory. Langmuir 2015;31(10):307685.
[54] Al-Wahaibi T, Al-Wahaibi Y, Al-Hashmi A-AR, Mjalli FS, Al-Hatmi S.
Experimental investigation of the effects of various parameters on viscosity
reduction of heavy crude by oilwater emulsion. Petrol Sci 2015;12
(1):1706.
[55] Fletcher P, Cobos S, Jaska C, Forsyth JPJ, Crabtree M, Gaillard N, Favero C.
Improving heavy oil recovery using an enhanced polymer system. SPE
improved oil recovery symposium. Tulsa (Oklahoma, USA): Society of
Petroleum Engineers; 2012.

[56] Abdulbaki M, Huh C, Sepehrnoori K, Delshad M, Varavei A. A critical review on


use of polymer microgels for conformance control purposes. J Petrol Sci Eng
2014;122:74153.
[57] Pu WF, Liu R, Wang KY, Li KX, Yan ZP, Li B, et al. Water-soluble coreshell
hyperbranched polymers for enhanced oil recovery. Ind Eng Chem Res
2015;54(3):798807.
[58] Choi SK, Sharma MM, Bryant SL, Huh C. PH-sensitive polymers for novel
conformance-control and polymer-flood applications. SPE Reservoir Eval Eng
2010;13(6):92639.
[59] Chen Z, Zhao X, Wang Z, Fu M. A comparative study of inorganic alkaline/
polymer flooding and organic alkaline/polymer flooding for enhanced heavy
oil recovery. Colloids Surf A 2015;469:1507.
[60] Sheng JJ. A comprehensive review of alkaline-surfactant-polymer (ASP)
flooding. Asia-Pac J Chem Eng 2014;9(4):47189.
[61] Doorwar S, Mohanty KK. Viscous fingering during non-thermal heavy oil
recovery. SPE annual technical conference and exhibition. Denver (Colorado,
USA): Society of Petroleum Engineers; 2011.
[62] Ge J, Feng A, Zhang G, Jiang P, Pei H, Li R, et al. Study of the factors
influencing alkaline flooding in heavy-oil reservoirs. Energy Fuels 2012;26
(5):287582.
[63] Ding B, Zhang G, Ge J, Liu X. Research on mechanisms of alkaline flooding for
heavy oil. Energy Fuels 2010;24(12):634652.
[64] Pei H, Zhang G, Ge J, Jin L, Liu X. Analysis of microscopic displacement
mechanisms of alkaline flooding for enhanced heavy-oil recovery. Energy
Fuels 2011;25(10):44239.
[65] Leaute RP, Carey BS. Liquid addition to steam for enhancing recovery (LASER)
of bitumen with CSS: results from the first pilot cycle. In: Canadian
international petroleum conference. Calgary (Alberta): Petroleum Society of
Canada; 2005.
[66] Leaute RP, Mohanty KK. Liquid addition to steam for enhancing recovery
(LASER) of bitumen with CSS: evolution of technology from research concept
to a field pilot at Cold Lake. SPE international thermal operations and heavy
oil
symposium
and
international
horizontal
well
technology
conference. Calgary (Alberta, Canada): Society of Petroleum Engineers;
2002.
[67] Lin L, Ma H, Zeng F, Gu Y. Liquid addition to steam for enhancing recovery
(LASER) of bitumen with CSS: evolution of technology from research concept
to a field pilot at Cold Lake. SPE heavy oil conference-Canada. Calgary
(Alberta, Canada): Society of Petroleum Engineers; 2014.
[68] Torabi F, Yadali Jamaloei B, Stengler BM, Jackson DE. The evaluation of
CO2-based vapour extraction (VAPEX) process for heavy-oil recovery. J Petrol
Explor Prod Technol 2012;2(2):93105.
[69] Upreti SR, Lohi A, Kapadia RA, El-Haj R. Vapor extraction of heavy oil and
bitumen: a review. Energy Fuels 2007;21(3):156274.
[70] Ivory J, Chang J, Coates R, Forshner K. Investigation of cyclic solvent injection
process for heavy oil recovery. Canadian international petroleum
conference. Calgary (Alberta): Petroleum Society of Canada; 2009.
[71] Talebian SH, Masoudi R, Tan IM, Zitha PLJ. Foam assisted CO2-EOR: a review of
concept, challenges, and future prospects. J Petrol Sci Eng 2014;120:
20215.
[72] Andrianov A, Farajzadeh R, Mahmoodi NM, Talanana M, Zitha PLJ. Immiscible
foam for enhancing oil recovery: bulk and porous media experiments. Ind Eng
Chem Res 2012;51(5):221426.
[73] Worthen A, Bagaria H, Chen Y, Bryant SL, Huh C, Johnston KP. Nanoparticle
stabilized carbon dioxide in water foams for enhanced oil recovery. SPE
improved oil recovery symposium. Tulsa (Oklahoma, USA): Society of
Petroleum Engineers; 2012.
[74] Nguyen P, Fadaei H, Sinton D. Nanoparticle stabilized CO2 in water foam for
mobility control in enhanced oil recovery via microfluidic method. In: SPE
heavy oil conference-Canada. Calgary (Alberta, Canada): Society of Petroleum
Engineers; 2014.
[75] Zhang YP, Sayegh S, Huang S. Enhanced heavy oil recovery by immiscible
WAG injection. In: Canadian international petroleum conference. Calgary
(Alberta): Petroleum Society of Canada; 2006.
[76] Bayat EA, Junin R, Samsuri A, Piroozian A, Hokmabadi M. Impact of metal
oxide nanoparticles on enhanced oil recovery from limestone media at
several temperatures. Energy Fuels 2014;28(10):625566.
[77] Hendraningrat L, Torster O. Effects of the initial rock wettability on silicabased nanofluid-enhanced oil recovery processes at reservoir temperatures.
Energy Fuels 2014;28(10):622841.
[78] Zhang H, Nikolov A, Wasan D. Enhanced oil recovery (EOR) using nanoparticle
dispersions: Underlying mechanism and imbibition experiments. Energy
Fuels 2014;28(5):30029.
[79] Moghaddam RN, Bahramian A, Fakhroueian Z, Karimi A, Arya S. Comparative
study of using nanoparticles for enhanced oil recovery: wettability alteration
of carbonate rocks. Energy Fuels 2015;29(4):21119.
[80] Hyne JB, Greidanus JW, Tyrer JD, Verona D, Rizek C, Clark PD, et al.
Aquathermolysis of heavy oils. In: The second international conference on
heavy crude and tar sands, Caracas, Venezuela; 1982. p. 2530.
[81] Rivas OR, Campos RE, Borges LG. Experimental evaluation of transition metals
salt solutions as additives in steam recovery processes. In: SPE Annual
technical conference and exhibition. Houston (Texas): Society of Petroleum
Engineers; 1988.
[82] Clark PD, Hyne JB. Chemistry of organosulphur compound types occurring in
heavy oil sands: 3. Reaction of thiophene and tetrahydrothiophene with
vanadyl and nickel salts. Fuel 1984;63(12):164954.

K. Guo et al. / Fuel 185 (2016) 886902


[83] Clark PD, Dowling NI, Hyne JB, Lesage KL. The chemistry of organosulphur
compound types occurring in heavy oils: 4. The high-temperature reaction of
thiophene and tetrahydrothiophene with aqueous solutions of aluminium
and first-row transition-metal cations. Fuel 1987;66(10):13537.
[84] Clark PD, Dowling NI, Lesage KL, Hyne JB. Chemistry of organosulphur
compound types occurring in heavy oil sands: 5. Reaction of thiophene and
tetrahydrothiophene with aqueous group VIIIB metal species at high
temperature. Fuel 1987;66(12):1699702.
[85] Clark PD, Kirk MJ. Studies on the upgrading of bituminous oils with water and
transition metal catalysts. Energy Fuels 1994;8(2):3807.
[86] Zhong LG, Liu YJ, Fan HF, Jiang SJ. Liaohe extra-heavy crude oil underground
aquathermolytic treatments using catalyst and hydrogen donors under steam
injection conditions. In: SPE international improved oil recovery conference
in Asia Pacific. Kuala Lumpur (Malaysia): Society of Petroleum Engineers;
2003.
[87] Ovalles C, Hamana A, Rojas I, Bolvar RA. Upgrading of extra-heavy crude oil
by direct use of methane in the presence of water. Fuel 1995;74(8):11628.
[88] Liu Y, Fan H. The effect of hydrogen donor additive on the viscosity of heavy
oil during steam stimulation. Energy Fuels 2002;16(4):8426.
[89] Ovalles C, Vallejos C, Vasquez T, Rojas I, Ehrman U, Benitez JL, et al. Downhole
upgrading of extra-heavy crude oil using hydrogen donors and methane
under steam injection conditions. Petrol Sci Technol 2003;21(12):25574.
[90] Fan Z, Zhao F, Wang J, Gong Y. Upgrading and viscosity reduction of super
heavy oil by aqua-thermolysis with hydrogen donor. J Fuel Chem Technol
2006;34(3):3158.
[91] Li J, Chen Y, Liu H, Wang P, Liu F. Influences on the aquathermolysis of heavy
oil catalyzed by two different catalytic ions: Cu2+and Fe3+. Energy Fuels
2013;27(5):255562.
[92] Wen S, Zhao Y, Liu Y, Hu S. A study on catalytic aquathermolysis of heavy
crude oil during steam stimulation. In: International symposium on oilfield
chemistry. Houston (Texas, U.S.A.): Society of Petroleum Engineers; 2007.
[93] Zhao XF, Tan XH, Liu Y. Behaviors of oil-soluble catalyst for aquathermolysis
of heavy oil. Ind Catal 2008;16(11):314.
[94] Yi Y, Li S, Ding F, Yu H. Change of asphaltene and resin properties after
catalytic aquathermolysis. Petrol Sci 2009;6(2):194200.
[95] Xu HX, Pu CS, Wu FP. Mechanism of underground heavy oil catalytic
aquathermolysis. J Fuel Chem Technol 2012;40(10):120611.
[96] Zhao F, Liu Y, Wu Y, Zhao X, Tan L. Study of catalytic aquathermolysis of heavy
oil in the presence of a hydrogen donor. Chem Technol Fuels Oils 2012;48
(4):27382.
[97] Zhao FJ, Wang X, Wang YL, Shi Y. The catalytic aquathermolysis of heavy oil in
the presence of a hydrogen donor under reservoirs conditions. J Chem Pharm
Res 2014;6(5):203741.
[98] Petrukhina NN, Kayukova GP, Romanov GV, Tumanyan BP, Foss LE, Kosachev
IP, et al. Conversion processes for high-viscosity heavy crude oil in catalytic
and noncatalytic aquathermolysis. Chem Tech Fuels Oils 2014;50(4):31526.
[99] Zhao F, Liu Y, Fu Z, Zhao X. Using hydrogen donor with oil-soluble catalysts
for upgrading heavy oil. Russ J Appl Chem 2015;87(10):1498506.
[100] Galukhin AV, Erokhin AA, Nurgaliev DK. Effect of catalytic aquathermolysis
on high-molecular-weight components of heavy oil in the Ashalcha field.
Chem Technol Fuels Oils 2015;50(6):55560.
[101] Chen Y, Wang Y, Wu C, Xia F. Laboratory experiments and field tests of an
amphiphilic metallic chelate for catalytic aquathermolysis of heavy oil.
Energy Fuels 2008;22(3):15028.
[102] Chen Y, Yang C, Wang Y. Gemini catalyst for catalytic aquathermolysis of
heavy oil. J Anal Appl Pyrol 2010;89(2):15965.
[103] Chao K, Chen Y, Liu H, Zhang X, Li J. Laboratory experiments and field test of a
difunctional catalyst for catalytic aquathermolysis of heavy oil. Energy Fuels
2012;26(2):11529.
[104] Wang Y, Chen Y, He J, Li P, Yang C. Mechanism of catalytic aquathermolysis:
influences on heavy oil by two types of efficient catalytic ions: Fe3+and Mo6+.
Energy Fuels 2010;24(3):150210.
[105] Wu C, Lei G, Yao C, Jia X. In situ upgrading extra-heavy oil by catalytic
aquathermolysis treatment using a new catalyst based an amphiphilic
molybdenum chelate. In: International oil and gas conference and
exhibition in China. Beijing (China): Society of Petroleum Engineers; 2010.
[106] Wu C, Lei GL, Yao CJ, Sun KJ, Gai PY, Cao YB. Mechanism for reducing the
viscosity of extra-heavy oil by aquathermolysis with an amphiphilic catalyst.
J Fuel Chem Technol 2010;38(6):68490.
[107] Chao K, Chen Y, Li J, Zhang X, Dong B. Upgrading and visbreaking of superheavy oil by catalytic aquathermolysis with aromatic sulfonic copper. Fuel
Process Technol 2012;104:17480.
[108] Wu C, Su J, Zhang R, Lei G, Cao Y. The use of amphiphilic nickel chelate for
catalytic aquathermolysis of extra-heavy oil under steam injection
conditions. Energy Sources, Part A 2014;36(13):143744.
[109] Desouky S, Sabagh AA, Betiha M, Badawi A, Ghanem A, Khalil S. Catalytic
aquathermolysis of Egyptian heavy crude oil. Inter J Chem Mater Sci Eng
2013;7(8):28691.
[110] Wang J, Liu L, Zhang L, Li Z. Aquathermolysis of heavy crude oil with
amphiphilic nickel and iron catalysts. Energy Fuels 2014;28(12):74407.
[111] Monin JC, Audibert A. Thermal cracking of heavy-oil/mineral matrix systems.
SPE Reserv Eng 1988;3(4):124350.
[112] Fan H, Liu Y, Zhong L. Studies on the synergetic effects of mineral and steam
on the composition changes of heavy oils. Energy Fuels 2001;15(6):14759.
[113] Fan H, Zhang Y, Lin Y. The catalytic effects of minerals on aquathermolysis of
heavy oils. Fuel 2004;83(1415):20359.

901

[114] Junaid ASM, Street C, Wang W, Rahman MM, An W, McCaffrey WC, et al.
Integrated extraction and low severity upgrading of oilsands bitumen by
activated natural zeolite catalysts. Fuel 2012;94(1):45764.
[115] Junaid ASM, Wang W, Street C, Rahman M, Gersbach M, Zhou S, et al.
Viscosity reduction and upgrading of Athabasca oilsands bitumen by natural
zeolite cracking. Inter J Chem Mater Sci Eng 2010;4(9):338.
[116] Merissa S, Fitriani P, Iskandar F, Abdullah M, Khairurrijal. Preliminary study
of natural zeolite as catalyst for decreasing the viscosity of heavy oil. In:
Padjadjaran international physics symposium 2013. Universitas Padjadjaran,
West Java-Indonesia: AIP Publishing LLC; 2013. p. 1314.
[117] Junaid ASM, Rahman MM, Rocha G, Wang W, Kuznicki T, McCaffrey WC, et al.
On the role of water in natural-zeolite-catalyzed cracking of Athabasca
oilsands bitumen. Energy Fuels 2014;28(5):336776.
[118] Strausz OP, Mojelsky TW, Payzant JD, Olah GA, Prakash GKS. Upgrading of
Albertas heavy oils by superacid-catalyzed hydrocracking. Energy Fuels
1999;13(3):55869.
[119] Chen Y, Wang Y, Lu J, Wu C. The viscosity reduction of nano-kegginK3PMo12O40 in catalytic aquathermolysis of heavy oil. Fuel 2009;88
(8):142634.
[120] Chen Y, He J, Wang Y, Li P. GC-MS used in study on the mechanism of the
viscosity reduction of heavy oil through aquathermolysis catalyzed by
aromatic sulfonic H3PMo12O40. Energy 2010;35(8):345460.
[121] Jing P, Li Q, Han M, Sun D, Jia L, Fang W. Visbreaking of heavy petroleum oil
2+
catalyzed by SO2
or Sn2+. Front Chem
4 /ZrO2 solid super-acid doped with Ni
Eng China 2008;2(2):18690.
[122] Wang H, Wu Y, He L, Liu Z. Supporting tungsten oxide on zirconia by
hydrothermal and impregnation methods and its use as a catalyst to reduce
the viscosity of heavy crude oil. Energy Fuels 2012;26(11):651827.
[123] Ovalles C, Filgueiras E, Morales A, Rojas I, de Jesus JC, Berrios I. Use of a
dispersed molybdenum catalyst and mechanistic studies for upgrading extraheavy crude oil using methane as source of hydrogen. Energy Fuels 1998;12
(2):37985.
[124] Zhao F, Huang J, Li M, Liu S, Guo Y, Zhang P. Study on hydrogen donors
catalytic upgrading of heavy oil using ultradispersed catalyst. J Chem Pharm
Res 2015;7(4):13707.
[125] Ovalles C, Filgueiras E, Morales A, Scott CE, Gimenez FG, Embaid BP. Use of a
dispersed iron catalyst for upgrading extra-heavy crude oil using methane as
source of hydrogen. Fuel 2003;82(8):88792.
[126] Shokrlu YH, Babadagli T. Viscosity reduction of heavy oil/bitumen using
micro- and nano-metal particles during aqueous and non-aqueous thermal
applications. J Petrol Sci Eng 2014;119:21020.
[127] Yang Z, Liu X, Su C, Li X, Zhang Z, Zhao M. Preparation of silica supported
nanoscale zero valence iron and its feasibility in viscosity reduction of heavy
oil. Micro Nano Lett 2014;9(5):3558.
[128] Li W, Zhu JH, Qi JH. Application of nano-nickel catalyst in the viscosity
reduction of Liaohe extra-heavy oil by aqua-thermolysis. J Fuel Chem Technol
2007;35(2):17680.
[129] Greff J, Babadagli T. Use of nano-metal particles as catalyst under
electromagnetic heating for in-situ heavy oil recovery. J Petrol Sci Eng
2013;112:25865.
[130] Shokrlu YH, Babadagli T. In-situ upgrading of heavy oil/bitumen during steam
injection by use of metal nanoparticles: a study on in-situ catalysis and
catalyst transportation. SPE Reservoir Eval Eng 2013;16(3):33344.
[131] Wu C, Su J, Zhang R, Lei G, Cao Y. The use of a nano-nickel catalyst for
upgrading extra-heavy oil by an aquathermolysis treatment under steam
injection conditions. Petrol Sci Technol 2013;31(21):22118.
[132] Alkhaldi S, Husein MM. Hydrocracking of heavy oil by means of in situ
prepared ultradispersed nickel nanocatalyst. Energy Fuels 2014;28(1):
6439.
[133] Shokrlu YH, Babadagli T. Kinetics of the in-situ upgrading of heavy oil by
nickel nanoparticle catalysts and its effect on cyclic-steam-stimulation
recovery factor. SPE Reservoir Eval Eng 2014;17(3):35564.
[134] Greff J, Babadagli T. Catalytic effects of nano-size metal ions in breaking
asphaltene molecules during thermal recovery of heavy-oil. SPE annual
technical conference and exhibition. Denver (Colorado, USA): Society of
Petroleum Engineers; 2011.
[135] Hendraningrat L, Souraki Y, Torsater O. Experimental investigation of decalin
and metal nanoparticles-assisted bitumen upgrading during catalytic
aquathermolysis. In: SPE/EAGE European unconventional conference and
exhibition. Vienna (Austria): Society of Petroleum Engineers; 2014.
[136] Nassar NN, Husein MM. Ultradispersed particles in heavy oil: Part I,
preparation and stabilization of iron oxide/hydroxide. Fuel Process Technol
2010;91(2):1648.
[137] Nassar NN, Husein MM, Almao PP. Ultradispersed particles in heavy oil: Part
II, sorption of H2S(g). Fuel Process Technol 2010;91(2):16974.
[138] Abdrafikova IM, Kayukova GP, Petrov SM, Ramazanova AI, Musin RZ, Morozov
VI. Conversion of extra-heavy Ashalchinskoe oil in hydrothermal catalytic
system. Petrol Chem 2015;55(2):10411.
[139] Ovalles C, Rivero V, Salazar A. Downhole upgrading of orinoco basin
extra-heavy crude oil using hydrogen donors under steam injection
conditions. Effect of the presence of iron nanocatalysts. Catalysts 2015;5
(1):28697.
[140] Shokrlu YH, Babadagli T. Effects of nano-sized metals on viscosity reduction
of heavy oil/bitumen during thermal applications. Canadian unconventional
resources & international petroleum conference. Calgary (Alberta,
Canada): Society of Petroleum Engineers; 2010.

902

K. Guo et al. / Fuel 185 (2016) 886902

[141] Galarraga CE, Almao PP. Hydrocracking of Athabasca bitumen using


submicronic multimetallic catalysts at near in-reservoir conditions. Energy
Fuels 2010;24(4):23839.
[142] Hashemi R, Nassar NN, Almao PP. Transport behavior of multimetallic
ultradispersed nanoparticles in an oil-sands-packed bed column at a high
temperature and pressure. Energy Fuels 2012;26(3):164555.
[143] Hashemi R, Nassar NN, Almao PP. Enhanced heavy oil recovery by in situ
prepared ultradispersed multimetallic nanoparticles: a study of hot fluid
flooding for Athabasca bitumen recovery. Energy Fuels 2013;27
(4):2194201.
[144] Hashemi R, Nassar NN, Almao PP. In situ upgrading of Athabasca bitumen
using multimetallic ultradispersed nanocatalysts in an oil sands packed-bed
column: Part 1. Produced liquid quality enhancement. Energy Fuels 2014;28
(2):133850.
[145] Hashemi R, Nassar NN, Almao PP. In situ upgrading of athabasca bitumen
using multimetallic ultradispersed nanocatalysts in an oil sands packed-bed
column: Part 2. Solid analysis and gaseous product distribution. Energy Fuels
2014;28(2):135161.
[146] Liu X, Yang Z, Zhao M, Li X, Su C, Zhang Z. Preparation of silica-supported
nanoFe/Ni alloy and its application in viscosity reduction of heavy oil. Micro
Nano Lett 2015;10(3):16771.
[147] Li K, Hou B, Wang L, Cui Y. Application of carbon nanocatalysts in upgrading
heavy crude oil assisted with microwave heating. Nano Lett 2014;14
(6):30028.

[148] Zamani A, Maini B, Almao PP. Experimental study on transport of ultradispersed catalyst particles in porous media. Energy Fuels 2010;24
(9):49808.
[149] Sandler J, Fowler G, Cheng K, Kovscek AR. Solar-generated steam for oil
recovery: reservoir simulation, economic analysis, and life cycle assessment.
Energy Convers Manage 2014;77:72132.
[150] Hascakir B, Acar C, Akin S. Microwave-assisted heavy oil production: an
experimental approach. Energy Fuel 2009;23(12):60339.
[151] Hill DG, Chilingar GV, Wittle JK. Direct current electrical enhanced oil
recovery in heavy-oil reservoirs to improve recovery, reduce water cut, and
reduce H2S production while increasing API gravity. In: SPE Western
Regional and Pacific Section AAPG Joint Meeting. Bakersfield (California,
USA): Society of Petroleum Engineers; 2008.
[152] Farooqui J, Babadagli T, Li HA. Improvement of the recovery factor using
nano-metal particles at the late stages of cyclic steam stimulation. In: SPE
Canada heavy oil technical conference. Calgary (Alberta, Canada): Society of
Petroleum Engineers; 2015.
[153] Rezaei M, Schaffie M, Ranjbar M. Thermocatalytic in situ combustion:
influence of nanoparticles on crude oil pyrolysis and oxidation. Fuel
2013;113:51621.
[154] Montgomery W, Sephton MA, Watson JS, Zeng H, Rees AC. Minimising
hydrogen sulphide generation during steam assisted production of heavy oil.
Sci Rep 2015;5:815964.

Potrebbero piacerti anche