You are on page 1of 36

SPE-193646-MS

A Comprehensive Review Heavy Oil Reservoirs, Latest Techniques,


Discoveries, Technologies and Applications in the Oil and Gas Industry

Cenk Temizel, Aera Energy; Celal Hakan Canbaz, Schlumberger; Minh Tran, USC; Elsayed Abdelfatah, University
of Calgary; Bao Jia, University of Kansas; Dike Putra, Rafflesia Energy; Mazda Irani, Ashaw Energy; Ahmad
Alkouh, College of Technological Studies

Copyright 2018, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE International Heavy Oil Conference and Exhibition held in Kuwait City, Kuwait, 10-12 December 2018.

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

Petroleum in general is found in sub-surface reservoir formation amongst pores existent in the formation. For
several years due to lack of information regarding production and technology, free-flowing, low viscosity
oil has been produced known as conventional crude oil. Fortunately, in recent times, due to advancement
of technology, high viscosity with higher Sulphur content-based crude has been produced known as heavy
oil. There are also exists significant difference in volatile materials as well as processing techniques used
for the two types of crude. (IEA, 2005; Ancheyta et al., 2007). The oil viscosity is a huge problem in regard
to heavy oil as both recovery and processing charges increase proportional to Sulphur content and viscosity
of the crude.
Heavy Oil can be used by definition internationally to describe oil with high viscosity (Although the
Oxford dictionary might have several variations of the same, within the contents of this paper, we refer to
heavy oil as high viscosity crude). Heavy oil generally contains a lower proportion of volatile constituents
and larger proportion of high molecular weight constituents as compared to conventional crude oil (often
referred to as light oil, we shall describe the characteristics of the types of oil further in the introduction).
The heavy oil just doesn't contain a composition of paraffins and asphaltenes but also contains higher traces
of wax and resins in its composition. These components have larger molecular structures leading to high
melting and pour points. This makes the oil a bad candidate for flow profiles and adversely affects the
mobility of the crude. (Speight, 2016). It is crucial to know the heavy oil constitution as it affects:

• Recovery: Low viscosity and high melting points

• Processing: Higher Resin, Sulphur and aromatic content

• Transportation: Low Viscosity

These all together impact the economics related to E&P (Exploration and Production) of heavy oil
resources. These resources generally have a higher of production associated with them and are one of the
first candidates to be affected by reduction of crude prices as seen in 2014 and early 2015. Crude oil can
2 SPE-193646-MS

generally be classified into its types by using its API values that are generally obtained through lab testing.
Table B1 provides a few popular crude types and their associated API Values.

Heavy Oil and Heavy Oil Resources and Recovery


Petroleum in general is found in sub-surface reservoir formation amongst pores existent in the formation. For
several years due to lack of information regarding production and technology, free-flowing, low viscosity
oil has been produced known as conventional crude oil. Fortunately, in recent times, due to advancement
of technology, high viscosity with higher Sulphur content-based crude has been produced known as heavy
oil. There are also exists significant difference in volatile materials as well as processing techniques used
for the two types of crude. (IEA, 2005; Ancheyta et al., 2007). The oil viscosity is a huge problem in regard
to heavy oil as both recovery and processing charges increase proportional to Sulphur content and viscosity
of the crude.
Heavy Oil can be used by definition internationally to describe oil with high viscosity (Although the
Oxford dictionary might have several variations of the same, within the contents of this paper, we refer to
heavy oil as high viscosity crude). Heavy oil generally contains a lower proportion of volatile constituents
and larger proportion of high molecular weight constituents as compared to conventional crude oil (often
referred to as light oil, we shall describe the characteristics of the types of oil further in the introduction).
The heavy oil just doesn't contain a composition of paraffins and asphaltenes but also contains higher traces
of wax and resins in its composition. These components have larger molecular structures leading to high
melting and pour points. This makes the oil a bad candidate for flow profiles and adversely affects the
mobility of the crude. (Speight, 2016). It is crucial to know the heavy oil constitution as it affects:

• Recovery: Low viscosity and high melting points

• Processing: Higher Resin, Sulphur and aromatic content

• Transportation: Low Viscosity

These all together impact the economics related to E&P (Exploration and Production) of heavy oil
resources. These resources generally have a higher of production associated with them and are one of the
first candidates to be affected by reduction of crude prices as seen in 2014 and early 2015. Crude oil can
generally be classified into its types by using its API values that are generally obtained through lab testing.
Table B1 provides a few popular crude types and their associated API Values.
Table B2 illustrates the differences in types of oil. The difference is based on API gravity and type of
recovery methods generally used. Chemistry however cannot be a distinguishing factor as the properties
are spread out. For further detailed distinction, readers are referred to USGS, 2007. Today there exists over
700 billion barrels of proven heavy Oil resources. Most of these resources are present in North American
region. Canada has almost one-third of the world's heavy oil resources. Followed by Venezuela in a close
second. A lot of the extra-heavy oil resources are yet of be developed and has more potential with the
advancement of technology. In this paper we will be reviewing the properties of crude oil, recovery and
production procedures and the major problem today in the heavy oil industry.
Despite a recent slump in oil price, oil and gas companies are always actively seeking to augment their
portfolios with additional petroleum reserves. These efforts help sustain the economic positions of petroleum
companies in the long run, providing financial flexibility to rationalize projects in case oil price rises
again. Heavy oil's abundant remaining reserves coupled with leaping technological advances in production
processes justify heavy oil project a promising prospect when price rebounds.
Petroleum is found in existing pores of subsurface reservoir formations. For several years due to
technology deficiency and production limitation, free-flowing, low viscosity oil has been produced known
as conventional crude oil or light oil. In recent times, due to advancement of exploitation technology,
SPE-193646-MS 3

high viscosity with higher sulphur content crude has been produced known as heavy oil. Viscosity and
API gravity are generally used to classify and differentiate light oil, heavy oil and bitumen (see Figure
A1). Heavy oil is defined as liquid petroleum in porous formations that has API gravity from 10° to 20°
(equivalent density from 943 to 1000 kg/m3 at 15.6°C), viscosity from 100 cP to 10000 cP at reservoir
temperature (Farouq et al., 2006). Heavy and extra heavy crude oil are biodegraded versions of oil that
remain when lighter components are lost or spent by biological, chemical and physical processes (Cornelius,
1987). Thus, heavy oil contains a lower proportion of volatile constituents and a larger fraction of heavy
constituents as compared to conventional crude oil, often referred to as light oil. Besides paraffin and
asphaltene, heavy oil's composition also contains higher traces of wax and resins. These components have
larger molecular structures and thus higher melting and pour points.
Heavy oil deposit usually occurs at shallow depths (<200 m), where the reservoir is hydraulically
connected to ground water aquifers (Hein, 2006). In the US, heavy oil is produced in situ mostly at shallow-
depth reservoirs, primarily taking advantage of steam injection technologies. Most heavy oil deposits occur
in shallow (3000 feet or less), high permeability (order of darcies), high porosity and oil saturation, relatively
thick (more than 50 feet) poorly consolidated sand formations (Farouq et al. 2006). Variations of formation
pressure and temperature in different heavy oil formations affect in situ oil viscosity and consequently
dictate different optimal recovery methods (See Table B3).
Most reporting agencies do not differentiate oil API gravity in their databases, thus creating difficulties
to keep track of heavy oil's reserves, production and consumption. In the US, most heavy oil resources
are concentrated in California (78%), Texas (15%) and Alaska, of which the bulk of heavy oil production
comes from California (Hein, 2006). The total original oil in place (OOIP) of the lower 48 states (Alaska
excluded) is 84 billion barrels, of which the cumulative production reached 11 billion barrels in 2007
(Utah Heavy Oil Program UHOP, 2007). According to the US Energy Information Administration, the
total US liquid petroleum production is 9480 MBOPD (thousand barrels of oil per day), of which around
490 MBOPD (5.2%) is from heavy oil production. In California, the most prolific state in terms of
heavy oil production, total oil production trends downwards from 660 MBOPD (1996) to 472 MBOPD
(2017) (California Department of Conservation CDC, 2017). Despite its decent reserves, Alaska heavy oil
production never really takes off, reaching 27 MBOPD (UHOP 2007). Resources that include bitumen and
heavy oil resources mostly grouped together in reserves estimation and reporting analysis. Proven bitumen
and heavy oil resources are estimated at 5.6 trillion barrels, significantly larger than the conventional crude
oil reserves (1.02 trillion barrels) and possible/undiscovered reserves (1.7 trillion barrels) (Hein, 2006).
More than 80% of the world heavy oil resources are in Canada, U.S.A and Venezuela. The Orinoco Oil Belt
(Venezuela) is the largest heavy oil field reserves-wise at 1.2 trillion barrels. Three largest North America
deposits are the Hoydminster deposits in western Canada (102 billion barrels OOIP), California deposits
(76 billion barrels OOIP) and North Slope Alaska deposits (30 billion barrels OOIP) (see Figure A2 and
UHOP, 2007). In Canada, conventional heavy production is predicted to stabilize at 530 MBOPD (14% of
total Canada oil production) while most of Canada production growth happens in the oil sands sector (see
Figure A3 and Canadian Association of Petroleum Producers CAPP, 2017). In Mexico, annual heavy oil
production reduced by 6% per year in 2006 (PEMEX, 2006).
Development of heavy oil resources encounters numerous challenges, ranging from resource delineation,
subsurface heat management to refining complexity and environmental impact. It is important to have the
properties of heavy oil constituents as they affect ultimate recovery, processing strategy and transportation
scheme. These aspects impact the economics of heavy oil resource's Exploration and Production (E&P).
Heavy oil requires a well-managed upgrading process to be compatible with current refining technology.
Upgrading is the process of converting these lower API oils to higher API oil so that they are more suitable
for conventional refinery feed stream (UHOP, 2007). Heavy oil field development generally has a higher
cost of production, therefore tend to be one of the first candidates to be negatively affected by crude oil
price slump starting at the end of 2014. Several inputs are important in determining the feasibility of a
4 SPE-193646-MS

heavy oil project such as: oil quality, refining market, existing infrastructure, energy sources for steam
generation, global oil price. Heavy oil economics are mainly affected by three aspects: oil price, price of
energy for steam generation and recoverable oil in place (UHOP, 2007). California 13°API heavy crude oil
is considered inferior in quality compared to lighter crude, thus it is usually traded at a discount (around
$5-$10 per barrel) to WTI crude price (see Figure A4).

Composition of Crude Oil


Crude oil contains all the hydrocarbons undersurface gaseous in liquid and solid form and other organic
compounds of hetero-elements and heavy metals. Chemical structure of hydrocarbons in crude oil
categorized according to polarity and polarizability status: Saturate, Aromatic, Resin and Asphaltane
(SARA). The presence of heavy metals and hetero-elements makes oil recovery difficult and expensive
compared to light oils. Therefore, quantification of hetero-elements and heavy metals are become important
as well as SARA analysis.

Hydrocarbon content
Saturated hydrocarbons which are the most precious portion of crude oil have sigma bonds between each
carbon and maximum amount of hydrogen in the carbons. Saturates comprise nonpolar organic compounds
such as paraffin (CnH2n+2), branched paraffin (CnH2n+2), naphthenes or cycloparaffins (CnH2n), saturated
carboxylic acids (R-COOH, R: CnH2n+1), etc. The percentage of saturates varies 2-50% based on content
of crude oil. More specifically, temperature, oil components and molecular configuration influence on
saturated hydrocarbon degradation of conventional crude oil. Thus, it contains higher amounts of saturated
hydrocarbons compared to heavy oil with the least saturated hydrocarbon content.
Aromatics contain one or more circles in their molecular structures with sigma and pi bonds between
carbons. Aromatics are polarizable hydrocarbons i.e. single ring (CnHn), two or multiple aromatic ring
compounds and their derivatives. 15-50% aromatic organic compound content is found in crude oil. Benzene
(C6H6) and benzene derivatives such as toluene (C6H6-CH3), xylene (C6H4-(CH3)2), ethylbenzene (C6H5-
C2H5) are the most existing compounds in crude oil. These chemicals present higher amount in conventional
crude oil than heavy oil since they are inflammable gases. Naphthalene (C10H8), anthracene (C14H10), pyrene
(C16H10) which are formed by attached one or more cyclic compounds commonly exist in heavy oil.
Resins have more complicated molecular structure than saturates and aromatics; thus, presence of
resins leads to increase of density and viscosity of crude oil. Resins are polar organic compounds that
have molecular weight of 500-2000 Da with saturated, unsaturated and cyclic hydrocarbon in their chain.
Resins which are found mostly in heavy oils contain higher amounts of hetero-elements than saturates and
aromatics. Although resins have various molecular shaper, all of them are soluble in five or more carbon
linear paraffins such as n-pentane and n-heptane. Resins are the major component of heavy oils.
Asphaltanes present mostly in heavy oils and consist of carbon, hydrogen, hetero-elements and trace
amounts of heavy metals. Asphaltene has high amounts of aromatic hydrocarbons (i.e. naphthanes) (~50%),
sulfur compounds (5-8 wt%), oxygen (~1 wt%), nitrogen (~1.5 wt%) and heavy metals in its molecular
structure. Asphaltene presents in tiny amounts in petroleum gases and conventional liquid crude oil
whereas it exists in heavy oils more than 10%. Asphaltanes are the most polar component with aromatic
rings, branched and unsaturated hydrocarbons in the side chains of their molecular structure; hence have
the highest density (i.e. ~1200 kg/m3) of crude oil. Existence of asphalthanes is also correlated with
viscosity increase; thus, heavy oils have higher asphaltane content which makes oil recovery complicated.
Asphaltanes are dissolved in benzene and derivatives and insoluble in nonpolar hydrocarbons despite their
unique molecular structure of asphaltane. Since asphalthanes are not soluble in alkanes such as n-heptane,
n-heptane, resins and asphaltane can be separated easily.
Asphaltane is well known for its large molecular weight of 0.5-10 kDa in the crude oil. However, recently,
molecular weight of asphaltene has found around 750 g/mol and the previous high molecular weights were
SPE-193646-MS 5

related to self-association of asphaltene molecules. Asphaltene can aggregate at very low concentration
(~0.1 g/L) as a result of van der Waals interaction between its molecules. Micellization of asphaltane
starts with asphaltane-asphaltane self-association. It is assumed that asphaltane molecules form asphaltane
micelles with an aromatic residue core and lighter hydrocarbons are in the shell. The micelles continue with
asphaltane aggregation and grown aggregates according to solubility of asphaltane. Solubility of asphaltanes
is affected by temperature, pressure, viscosity and injected chemicals such as permeability modifiers as
well as content of saturates, aromatics and resin in the reservoir. Specifically, resin has an important role on
steadiness, movability and solubility of asphaltane by their noncovalent interactions.
Asphaltane micellization improves oil flowability by viscosity decrease effect of asphaltane self-
association. Additionally, asphaltane micelles enhance porous plugging that lessens permeability. Linear
paraffin or saturated alkanes can be injected to lower asphaltane solubility to upgrade heavy oil flow.
Carbon number of injected paraffin has an impact on solubility of asphaltane and resin. In some cases, both
asphaltane and resin precipitation upgrade heavy oil flow, in some cases only asphaltane precipitation is
preferred. For instance, propane is applied to decrease both resin and asphaltene solubility. Higher paraffins,
decane or hexane is used to solubilize resin and precipitate asphaltene. On the other hand, asphaltane
aggregation is stable that causes viscosity increase and hence flow ability decrease. Micelles or early
stages of precipitation of asphaltane improves oil mobility; however, further asphaltane aggregates can lead
unfavorable results.

Hetero-element and heavy metal content


The content of sulfur, oxygen and nitrogen (S-O-N) hetero-elements, heavy metals and nonvolatile
components in heavy oils are higher than conventional oils. Sulfur based components (1.1-6 wt% in heavy
oils) are methanethiols (CH3-SH), thiols (-C-SH or R-SH), thiophenols (C6H5-SH), disulfides (R1-S-S-
R2), benzothiophene (C8H6S) and their derivatives. Oxygen containing compounds (0.1-3 wt%) are found
in carboxylic acids (R-COOH), phenols (C6H5-OH), furans (C4H4O) and some aromatic hydrocarbons.
Nitrogen content (0.2-1.5 wt%) comes from pyridine (C6H5N), indoles (C8H7N), carbazoles (C12H9N),
porphyrin (C20H12N4) and their derivatives.
Sulfur is the third common (after carbon and hydrogen) chemical element in crude oil with the average
weight. Sulfur substances are not thermally stable unlike nitrogen compounds. Sulfur based compounds
present either in organic or inorganic substances. Sulfur presents mostly in organic compounds such
as thiols, disulfides, etc. Inorganic sulfur present in hydrogen sulfide (H2S), pyrite (FeS2). Thioethers
and thiolanes can be eliminated by hydrodesulfurization or thermal treatment. However, the other sulfur
compounds are not excluded easily with these treatments. Oxidation pretreatment improves desulfurization
of crude oil for those compounds. Hydrodesulfurization is a method that produces hydrogen sulfide and
discard sulfur compounds. Hydrogen sulfide is converted to elemental sulfur or sulfuric acid to be eliminated
from crude oil. Hydrogen sulfide is a corrosive/harmful compound; therefore, nanoparticles might be used
in the future to obtain a safer method to remove sulfur due to adverse effects of hydrogen sulfide.
Heavy metals naturally found in crude oils are vanadium (V), iron (Fe), nickel (Ni), sodium (Na),
potassium (K), lithium (Li), copper (Cu), calcium (Ca), lead (Pb), titanium (Ti), manganese (Mn), mercury
(Hg), silver (Ag), aluminum (Al), gold (Au), zinc (Zn) and barium (Ba), magnesium (Mg). The most
common and unwanted heavy metals are nickel (up to 150 ppm) and vanadium (0.1-1200 ppm). Nickel and
vanadium heavy metals dissolve in oil; thus, it is laborious to remove from oil. Presence of any heavy metals
causes negative circumstances such as catalyst deactivation, corrosion during oil purification. Heavy metals
are found in either inorganic salts or organometallic substances. Calcium, magnesium, zinc and titanium
exist in naphthenic acids. Vanadium, nickel, copper metals containing compounds form metalloporphyrin's
which are dissolved in asphaltane aggregation. Current heavy metal elimination methods are using a solvent
for deasphalting and sulfuric acid (H2SO4). The released heavy metal gases would be detrimental to the
6 SPE-193646-MS

environment (e.g. air, water pollution) if the fuels were not refined from heavy metals. Thus, heavy metal
removal is needed to upgrade product quality, maintenance and operation adverse events and decrease
harmful environmental effects.

Heavy Oil Chemistry


Conventional crude oil has been utilized for decades compared to other types of oil due to their cost
effectiveness and high-level recovery. Considering portion of light oil in the reservoirs has started to lessen,
heavy oil has still the largest share of the petroleum reserves. Thus, heavy oil reservoirs gain in importance
to as future energy sources compensate global energy demand; however, upgrading processes are necessary.
Heavy oils are originally located in deep similar to conventional crude oils and they drift to ground level
where light hydrocarbons and volatile compounds are lost, biodegradation and physical degradation occur.
It is estimated that optimum temperature for microbial biodegradation is around 80 °C; consequently, most
heavy oils are located in shallow places. Heavy oil has higher carbon, sulfur and heavy metal content
compared to light oil. As saturated compounds degrade, Carbon/Hydrogen (C/H) ratio, viscosity, density,
acidity of heavy oils increase and Hydrogen/Carbon (H/C<1.5) reduces.
Viscosity. Heavy oils are highly viscous and viscosity reduction is necessary to enhance heavy oil
properties. Applied techniques to upgrade oil flow are (1) temperature increase, (2) dilution, (3) surfactant
inclusion, (4) addition of pour point depressants, (5) catalytic cracking. Increasing temperature is most
commonly used technique to decrease viscosity to overcome transportation problems. Dilution with lighter
oils or solvents (e.g. alcohols) is utilized for viscosity modification. Surfactants are added to form oil-in-
water self-assemblies to suspend heavy oil in emulsions that improves flow ability of heavy oils. Pour point
described as the lowest degree of temperature that oil flows without restriction. Pour point depressants are
utilized to control wax crystal formation under reservoir conditions. Metal ions (such as Cu2+, Fe3+, Mo6+)
and clays (e.g. zeolite) have been added for catalytic cracking.
Acidity. Heavy oils are more acidic compared with conventional crude. The reason is carboxylic
acids, phenols, polycyclic aromatic compounds and derivatives formation in heavy oils. TAN value is
important quantification for petroleum refinery. Total acid number (TAN) is used to estimate oil acidity
by determination of required potassium hydroxide (KOH) in milligrams to neutralize pH of a gram of oil
(ASTM D664 standard method). TAN is calculated with the eq. 1 and higher than 0.5 mg KOH/g oil is
considered acidic oil.

[1]

CP: concentration of potassium hydroxide (KOH)


VP: required volume of KOH
Wo: oil sample amount in grams
Naphthenic acids are formed by biodegradation of fossil fuels and exist mostly in heavy oil reservoirs.
TAN value usually correlated with naphthenic acid concentration; however, TAN shows acidity of acidic
compounds such as carboxylic acids, phenol derivatives and inorganic acids in an oil sample. Naphthenic
acids which contain saturated aliphatic, aromatic and heterocyclic carboxylic acids cause corrosion
problems such as equipment malfunction, operational difficulty (Chemical formula is CnH2n-zO2 where n
defines the carbon number and z defines deficiency of hydrogen (z ≥ 0 with a positive even number). The
positive even number indicates the presence of aromatic ring in the hydrocarbon chain of carboxylic acid.
Half of z number,z/2, presents the number of aromatic rings in the chain. Long chain naphthenic acids
could be thermally cracked into shorter chain carboxylic acids. Naphthenic acids also have two possible
reactions with other chemicals. (1) Naphthenic acids cause corrosion in the presence of metals as a result
SPE-193646-MS 7

of oxidation reaction. (2) Esterification reaction takes place between alcohols and naphthenic acids. Esters
are less corrosive compared to naphthenic acids.

American Petroleum Institute (API) Gravity


It is calculated by the eq. 2 to categorize oils by using their density. Conventional crude oil has 30-40° API
gravity which displays lighter oil density and specific gravity. As oil density and specific gravity increases,
API gravity number lessens and API gravity of heavy oils could be as low as 2°. Heavy oils are classified
into three groups according to their API: heavy, extra-heavy oil and tar sand. The lower API gravity number
shows the higher density of oil.
[2]
ρ= density of oil sample at 60 °F/15.6 °C (kg/m3)
Heavy Oil. API gravity number of heavy oils is 10-22.3. Heavy oil has a viscosity of 100-5,000 centipoises.
Heavy oils are mobile under reservoir environment and hence can be recovered by enhanced oil recovery
techniques. Heavy oils might have more than 60 carbons in their compounds and > 2wt% sulfur content
that are the reasons of high boiling point.
Extra-Heavy Oil. Extra heavy oil has mobility under reservoir conditions with API gravity number <10°.
Viscosity of extra-heavy oil changes between 5,000 and 10,000 centipoises.
Tar Sand / Ultra-heavy oil. Tar sand or ultra-heavy oil which contains sandstones and bitumen is motionless
in reservoir environment and cannot recover easily by EOR techniques. Tar sand is semi-solid with viscosity
of 10,000-10,000,000 cP and 2-4° API gravity number. The bitumen content of tar sand which is soluble
in carbon disulfide is a highly viscous sticky hydrocarbonaceous substances with hetero-elements and
heavy metals. Bitumen consists of polyaromatic chains in its chemical structure and varies from a reservoir
to another. Bitumen has high amount of sulfur compounds (mostly >4 wt%) than nitrogen and oxygen
constitutes. Bitumen has less hydrogen compared to heavy oils. It has high boiling point (i.e. >500 °C) due
to its complex structure. The half of bitumen comprises vacuum residues which can be purified by coking
(carbon rejection) or hydro-treating (hydrogen addition). Main principle of coking processes is using heat
to crack larger molecules to smaller products. The presence of hetero-elements and heavy metals might
affect cracking process. The purposes of hydro-treating are to remove hetero-elements especially sulfur
compounds and increase Hydrogen/Carbon ratio. Water, tetralin (C10H12), methyl formate (C2H4O2) and
syngas (CO+ H2), have been applied as hydrogen source for hydro-treating applications. Although some
solvents might cause undesired asphaltane precipitation, hydro-treating processes are promising.

Heavy Oil Recovery Techniques


In this section, we provide a comprehensive understanding and critical review of heavy oil recovery
techniques, EOR method selection for reservoirs depending on characteristics. We also present the
challenges and new technologies emerged recently to enhance the productivity of heavy oil reservoirs.
Completion design in heavy oil reservoirs, advanced wells and other technologies, production optimization
and artificial lift methods that help heavy oil recovery will be discussed. Heavy oil extraction is harder
compared to conventional petroleum due to the higher viscosity and specific gravity as well as low hydrogen/
carbon ratios with higher nitrogen, heavy metal, sulfur and Asphaltene content compared to light oils
(Speight 2013). The extreme viscosity of heavy oil makes it recovery very challenging.
The massive heavy oil reserves can play a significant role to meet energy needs for future generations
(Speight 2013, Meyer 1997). Despite the heavy oil resources of trillions of barrels, the recovery factors
from these resources are significantly low. This requires the development of evolutionary recovery
technologies that are economically and environmentally remarkable. There are many different methods for
8 SPE-193646-MS

heavy oil recovery. Here, we will present a synopsis of the different methods used to recover heavy oil
reservoirs. These recovery methods are subdivided into two major categories: thermal recovery methods
and nonthermal recovery methods (Speight 2013, Brook and Kantzas 1998, Mai et al. 2009, Selby et al.
1989, Nasehi Araghi and Asghari 2010, Huang et al. 2017, Ali 1976, Al Bahlani and Babadagli 2008). Table
3 summarizes the most important recovery methods for heavy oil reservoirs.

Nonthermal Recovery Methods


In thin formations where heat losses to surrounding layers become significant, nonthermal recovery methods
are required to enhance the productivity and increase recovery factor. Chemical flooding techniques are
promising for improving recovery of heavy oil (Brook and Kantzas 1998). Nonthermal methods are a viable
option for improving recovery of moderately viscous from reservoirs which are unsuitable for thermal
treatment (Selby et al. 1989). For example, too thin and deep formations cannot contain the heat within the
reservoir. This makes thermal methods inefficient. However, nonthermal methods recovery factor is still
low compared to thermal methods.
After primary recovery of heavy oil reservoir reaches an economic limit, mostly huge oil volumes are still
remains inside the reservoir with a depleted reservoir pressure. The basic EOR process which considered
for additional oil recovery is waterflooding. As a result of high viscosity, the mobility ratio of heavy oil is
very high and hence, the recovery factor is extremely low and associated with high water cut (Selby et al.
1989). However, Waterflooding still applied in many heavy oil reservoirs due to low cost. As mentioned
earlier that shallow heavy oil reservoirs, thermal recovery is not feasible. Due to low cost, waterflooding
has been applied for such reservoirs. However, the sweep efficiency and recovery factors are very low due
to high mobility ratio. High mobility ratio is basically sourced by the large viscosity differences between the
viscous oil phase in the reservoir and the injected water phase that creates viscous fingering leaving large
oil saturation behind (Ali 1976). Regardless of poor performance, currently more than 200 wells in western
Canada are operated using waterflooding (Nasehi Araghi and Asghari 2010).
Due to limited potential of waterflooding in heavy oil reservoirs by viscous fingering, polymers can be
used to increase sweep efficiency by increasing the displacing water viscosity. Polymer injection can be
used in such cases that limit the performance of waterflooding such as fractured reservoirs (Speight 2013).
Viscosity increment of water phase decreases the water/oil mobility ratio. Also, the shear thinning behavior
of polymer solutions can adjust the mobility ratio accordingly to the local permeability of the reservoir. In
some cases, polymer crosslinked gels are used for conformance control by shutting off high permeability
streaks (Brook and Kantzas 1998).
Alkali-surfactant flooding is not a new concept. Chemical flooding techniques have been proposed as
early as the 1950 and they are a well-recognized technique in EOR for conventional oil. Heavy oils contain
naturally occurring acids that would react with the injected alkali and generate in situ surface active agents at
the oil-water interface. The presence of these active agents at the interface reduces surface tension between
oil and water which can greatly minimize the capillary pressure trapping an oil ganglion in a rock pore (Mai
et al. 2009). Thus, a reduction in interfacial tension can lead to the enhanced recovery of oil beyond the
amount obtained from waterflooding. A secondary result of reducing the oil-water interfacial tension is the
potential for emulsifying the fluids in one another by applying shear to mix the fluids together. Simple IFT
reduction alone will not be sufficient to produce heavy oil; improved recovery must be due to the formation
of emulsions, either changing the effective viscosity of the in-situ fluids or altering their flow pathways.
CO2 injection into heavy oil reservoirs is proved to be an effective method of recovery. CO2 can effectively
dissolves into the heavy oil and reduces the viscosity of oil phase (Nasehi Araghi and Asghari 2010).
This would cause heavy oil to swell and improve the recovery of heavy oil compared to waterflooding.
By increasing the operating pressure above the minimum miscibility pressure, the recovery increases
significantly. Also, increasing the bulk volume of CO2 injected can tremendously increase oil recovery.
Post-CO2 waterflooding shows an improvement in recovery form lower permeability reservoirs (Figure A5)
SPE-193646-MS 9

Solvent-based methods are also useful to reduce heavy oil viscosity in wellbore using a diluting agent
(naphtha, light oil, etc…). This would ease the pumpability of heavy oil to the surface. Also, the diluent
can be mixing with heavy oil in the surface facilities to facilitate transport through a pipeline. Meanwhile,
solvent-based methods improve the reservoir heavy oil mobility in a nonthermal way. Vapor-assisted
petroleum extraction (VAPEX) (Butler and Mokrys 1991) is a nonthermal process where relatively cold
solvents are injected at low pressure through two parallel horizontal wells that are 15 feet apart in the same
vertical plan (Figure A6). Vaporized solvents are injected through the upper well to create a vapor chamber.
The vapor chamber grows and flows through to the oil face. Henceforth, the condensed vapor mixed with oil
reducing oil viscosity and then the mixture flows to the well in a lower depth via gravity drainage where it
is pumped to the surface (Speight 2013). However, this process can be also applied in a horizontal well or in
a pair of a vertical and horizontal well. Ahmadi et al. (2014) present experimental, statistical, and modeling
study to evaluate the recovery of VAPEX in heavy oil reservoirs. This study highlights the importance of
Peclet number as well as the effect of fractures inside the reservoir on the rate of oil production performance
by using VAPEX method.
VAPEX has been proposed as a viable recovery process for EOR in cases that the thermal methods are
technically and economically unviable (Butler and Mokrys 1991). VAPEX relies mainly on gravity drainage.
However, in extremely thin reservoirs the gravity drainage is insufficient to provide efficient recovery.
Another recovery process was proposed for such thin reservoirs by injecting hydrocarbon gas in huff-n-
puff mode.
As shown by Irani and Gates, (2017) for propane the hydrate will be formed at reservoir pressure and
temperature that VAPEX is running. This phenomenon was the main reason to fail the VAPEX process in
the field pilot although it works in lab that was in room temperature above the hydrate formation envelope
for propane.
Dong et al. (2006) discussed the application of Cyclic Solvent Process (CSP) using Methane in horizontal
wells for heavy-oil reservoirs with low reservoir thickness. The process relies on restoring solution-gas-
drive mechanism using the injected Methane. An appropriate amount of Methane should be injected, and
pressurized by water injection to reach the approximate original reservoir pressure. Dong et al. (2006)
conducted sandpack tests using dead-oil of different viscosity. They found that the process of pressure
cycling can create favorable conditions for an efficient contact of injected gas with the remaining oil phase
inside the reservoir. This leads restoring of the original mechanism of solution-gas drive and displaces a
significant amount of the remaining oil. Jia et al. (2015) introduced a mathematical model for CSP. The
model highlighted the impact of significant convection on enhancing solvent and heavy oil mixture and
showed that the larger solvent injection rates lead more efficient mixing processes. Yadali Jamaloei et
al. (2012) discussed the limitation of the conventional CSP introduced by Dong et al. (2006). The main
limitations of CSP come from that during production the pressure drops greatly, and the light particles of
oil phase change into gas phase. Hence, the oil viscosity will increase again and it will be resulted with the
limited oil production. To overcome these limitations, A new process called ECSP (enhanced cyclic solvent
process) proposed by Yadali Jamaloei et al. (2012), In the process, two types of hydrocarbon solvents are
cyclically injected as two separate slugs as the first slug is a more voliant material such as methane, and the
other is more miscible material like propane. ECSP shows a better performance in low reservoir thickness
cases compared with CSP for the heavy oil recovery at the range of low-to-intermediate pressures. The two
solvents cooperate to support the pressure and reduce the viscosity. Methane expansion during the pressure
reduction that provides drive mechanism while propane mixes with oil to reduce the oil viscosity (Figure A7)
Another promising mechanism that increases the production of heavy oil is the flow of oil foam with cold
processes in a solution gas drive. Anomalous behaviors were observed in some reservoirs contain heavy
oil when the reservoir pressure drops. In some cases like saturated reservoirs, the reservoir pressure does
not drop rapidly and producing GOR does not show a significant increment. Sampling oil at the wellhead
from such anomalous reservoirs shows a form of an oil-continuous foam. Smith (1988) first described this
10 SPE-193646-MS

type of anomalous production. He noted that conventional solution gas drive models could not fit the cold
production from several Lloydminster heavy oil reservoirs. Sheng et al. (1999) provided a critical review of
foamy flow in primary heavy oil production with solution gas drive mechanism. During reservoir pressure
decline, gas released from oil forms small bubbles dispersed within the pores. As the pressure, continuous
to drop the discrete bubbles size increases. However, instead of the gas forms a continuous phase, it forms a
gas-in-oil dispersion. This kind of dispersion is what we call "foamy oil flow". Maini and Busahmin (2010)
discussed the field practices and observations related to foamy oil production in Canada. They also discussed
the influence of various process parameters. Several pore scale processes contribute to the formation of
foamy oil e.g. nucleation, growth, mobilization, trapping, breakup and coalescence of bubble. Lu et al.
(2015) conducted an experimental study using micro model to sort out the effect of temperature on foamy
oil behavior. This study confirms that the application for the solution gas drive mechanism under 120°C
shows a performance enhancement on heavy oil production sourced by foamy oil.
Cold heavy oil production with sand (CHOPS) is widely preferred in heavy oil reservoirs in Western
Canada. It's responsible for about 500,000 bbl/day of oil production (Maini and Busahmin 2010).
Unconsolidated sand heavy oil reservoirs are usually a good candidate for CHOPs. In these reservoirs,
sand production can be initiated easily to create wormholes that facilitate heavy oil production. The major
mechanisms that contribute to the success of CHOPS are wormhole generation and foamy oil flow (Xiao
and Zhao, 2017).

Thermal Recovery Methods


They have been widely proven to be effective in reducing heavy oil viscosity. This facilitates the flow of oil
to the wellbore. Thermal recovery methods involve inducing heat into the reservoir to improve the recovery
factor.
One of the most widely applied thermal methods is cyclic steam stimulation which is also usually called
Huff-n-Puff steam injection process. In CSS, steam injection and production of oil are operated in a single
well, where steam is injected then shut down the wellbore for some time for steam condensation. Following
this soaking process, the wellbore is opened for production of oil with reduced viscosity. CSS is usually
operated through vertical wells at a pressure higher than the formation pressure to create fractures that
increase the connectivity with the reservoir. Although CSS is a well-developed technology, the recovery
factor by CSS is usually less than 30 % with steam-to-oil ratios (SOR) of 3 to 5.22. However, CSS is still
the favored method for heavy oil reservoirs with strong overburden that can contain high-pressure steam.
CSS requires a minimum depth of 1000 foot for efficient operation, with thick pay zones (>10 m) with high
porosity sands (>30%).
Unlike CSS, steamflooding involves injection-production well pairs. Steam is injected through the
injection well increases the mobility ratio and mobilize the heavy oil phase to the production well (Speight
2013). Usually, steamflooding follow cyclic steam stimulation process (see Figure A8) Sufficient amount of
steam should be injected to increase the reservoir temperature to the limit that would achieve enough heavy
oil viscosity reduction. This requires reservoirs with sufficiently high effective permeability. However,
steamflooding is usually accompanied by two problems; steam override and reservoir plugging. Steam has
a lower density than the cold reservoir fluids. Hence, the steam shows a tendency to migrate upward to the
upper sections of reservoir overriding reservoir fluids. In order to resolve this issue, steam can be injected
at high injection rate at the lower sections of the reservoir. The successful steamflooding requires at least
porosity of 20% and permeability larger than 100 millidarcies. Meanwhile, the saturation of oil phase should
be greater than 40% with reservoir depth less than 3000 ft, and the thickness of at least 30 ft (Speight 2013).
In situ combustion is another thermal process, in which dry or wet air (air mixed with water) is injected
to the reservoir (Speight 2013). Then the air is ignited to create a fire that uniformly propagates through to
the production well with the oil moving ahead of the combustion front. The temperature in the combustion
front can be hundreds of degrees centigrade which allow breaking up the heavy oil components lowering
SPE-193646-MS 11

the boiling point of the product components. These light product components can achieve miscibility with
heavy oil ahead of the front lowering the heavy oil viscosity. The coke remained behind the front provides
the fuel to maintain the fire front (see Figure A9). However, the front propagation is hard to control and may
propagate haphazardly to the production well causing damage with gases escaping to the surface. Similar to
steamflooding process, air injected to start the in situ combustion process can override the reservoirs fluids.
The produced fluid may contain hard to break emulsions that contain heavy metal compounds. Despite these
limitations, in situ combustion process has been proven to be efficient for thick reservoirs with effective
gravity drainage.
Steam Assisted Gravity Drainage (SAGD) is another form of steamflooding that relies on gravity drainage
to enhance the steam injection performance. SAGD involves two horizontal wells in the same vertical plane.
Steam injected into the top well creating a steam chamber propagate upward and horizontally in the reservoir
(See Figure A10 and A11) Steam condensation at the edge of steam chamber interface provides an extensive
latent heat that decreases the bitumen viscosity. Then the thinned oil drain under gravity forces to the liquid
sump (or liquid pool) at the base of steam chamber. Production well produces the oil from liquid sump
with one of threee methods: natural flow, gas-lifting or electrical submersible pump (ESP). The typical
distance between the injection and production wells tried to be at 15 feet but undulation is about 4 feets in
most SAGD drillings. SAGD process is the most efficient thermal recovery method with recovery factor
around 50-80%. During SAGD process several gases are produced such as methane (due to oil cracking and
exsolution of methane), some carbon dioxide (due to aquathermolysis and water reaction with rock) and
traces of hydrogen sulfide (due to aquathermolysis). The noncondensable gases (NCG) stay at the top of
the reservoir filling the oil-depleted pores. It may also form an insulating blanket that reduces vertical heat
losses (Al Bahlani and Babadagli 2008). In most SAGD operations NGC is injected in blow-down stage
to keep the pressure and reduce the SOR.
Electromagnetic heating was proposed as a good stimulation technique for injectivity reservoirs with thin
pay zones and moderate viscosity of the heavy oil (Sahni et al. 2000). Meanwhile, electromagnetic heating
can be combined with other thermal and nonthermal recovery methods to provide preferential pathways
for the injected fluid especially in thin reservoirs to maintain the propagating front within the payzone. As
an example, Effective Solvent Extraction Incorporating Electromagnetic Heating (ESEIEH™) technology
used RF-heating and solvent injection to deploy bitumen from Athabascan reservoirs (Ghannadi et al.,
2016). One of the common methods of EM-heating is the ohmic heating. The electric heater vendors are
providing two criteria for electric heaters: target temperature on the sand face and power provided. As
illustrated by Irani and Ghannadi (2017) electrical heaters provide a large temperature but cannot support the
required flux or power. In Figure A12, the temperature distribution for electrical heater that only provides
50W/ft vs. steam circulation. The temperature distribution would be similar if power of electrical heaters
rise to four times.
Steam injection, in some cases, can be inefficient due unfavorable reservoir properties such as oil
wettability and low permeability. Alternative injection of the proper hydrocarbon solvent and steam can
help to change the reservoir and oil properties to enhance the performance of the steamflooding process.
This method is called steam-over-solvent injection (SOS). Al-Bahlani and Babadagli (2012) conducted
an experimental study using different solvents to show the potential of SOS for heavy oil recovery from
fractured carbonate reservoirs. They found that lighter molecular weight solvents produced a higher amount
of oil. Meanwhile, the amount of asphaltene precipitation is significant. Higher injection rates can improve
the oil recovery. However, the solvent to oil ration would increase significantly.
Improving the performance of steam injection process becomes critical and requires understanding the
nature of the interaction of steam with heavy oil. According to Hyne et al (Technology et al. 1986) decreasing
heavy oil viscosity during steam stimulation is a combination of a physical process and chemical reactions
(aquathermolysis) between steam, oil, and sand. These chemical reactions can induce in situ upgrading
the heavy oil. This is accompanied by increasing the content of saturates and aromatics in the heavy oil
12 SPE-193646-MS

while decreasing the asphaltene and resins contents. This would induce higher H/C ratio and an irreversible
lowering of heavy oil viscosity. At the end of aquathermolysis, CO2 and H2 formed by the reaction of water-
gas shift (Clark and Hyne 1990). The temperature window for the aquathermolysis reactions ranges from
200 °C to 300 °C. Several researchers investigated the ways to catalyze the aquathermolysis reactions to
enhance the performance of steamflooding methods. Hamedi Shokrlu and Babadagli (2010) had investigated
the effect of nanoparticles as a catalysis for the aquathermolysis reactions. They used copper and iron
oxides micron-sized and nano-sized particles. Nanoparticles have higher specific surface area compared to
microparticles. Nanoparticles were found to be more effective in reducing the viscosity of the heavy oil due
to the higher surface area and more reactivity.

Completion Design in Heavy Oil Reservoirs


Cavender (2004) multilaterally summarized the completion methods that used in heavy reservoir
development. Multilateral completions drastically increase the total contact area of wellbore(s) with the
reservoir by using a single surface location. This subsequently reduces the overall field development costs.
Heavy oil reservoirs are commonly found in unconsolidated sandstones formations. This may require
precautions regarding sand control and wellbore stability. Slotted and screened liners and gravel packing
can be used to enhance sand control from heavy oil reservoirs.
Jain et al. (2013) presented a comprehensive and integrated workflow for completion design in heavy
oil reservoirs that involves injection and production from the same well. They were mainly focused on
cyclic fluid injection like VAPEX through horizontal wells in reservoirs not efficient for thermal recovery
implementation. This kind of processes involves the injection of viscosity reducing solvent into the reservoir
followed by producing heavy oil and solvent mixture from the same well. This process is more effectively
implemented through horizontal wells to increase the contact area with the reservoir.
Heavy oil is generally produced from shallow reservoirs. The primary recovery method carried out at
reservoir temperature is called "cold production". A cold production is driven by the solution gas drive.
During the process of solution gas drive, oil expands and oil viscosity decreases. Abnormally high primary
recovery factor as high as 20% is obtained in many heavy oil reservoirs characterized by high oil-gas ratio
and high oil flow rate. Two mechanisms attribute to this production behavior: the abnormal solution gas
drive in foamy oil flow and wormhole formulation (Chen 2006a, b).
Solution gas drive occurs during the primary production when the oil with dissolved gas depletes and
generates bubbles slowing down reservoir pressure decline. The pressure at which the first bubble comes
out is defined as the bubble point. Chemical potentials of gas and liquid equal to each other in the oil with
dissolved gas (Eq. 3).
[3]
Eq. 4 describes the pressure relationship between the gas phase and liquid phase.
[4]
where γ represents the interfacial tension between liquid and gas phases and r represents the bubble radius.
Eq. (4) shows that the liquid phase pressure is higher than the gas pressure, and their difference is defined
as the supersaturation (Bauget and Lenormand 2002) which equals to the capillary pressure.
Homogeneous and heterogeneous models have been used to describe bubble nucleation in oil. Actually
both of them cannot properly describe the this phenomenon. The homogeneous model cannot predict a
threshold supersaturation for bubble growth, which is experimental observed in the micromodel experiment.
The heterogeneous model is also contradictory to the micromodel experiment because the bubble nucleation
occurs only with certain degree angles of the cavity. The relative permeability for gas and oil of foamy
heavy oil are defined based on Corey model as:
SPE-193646-MS 13

[5]
[6]
[7]
where S gives the dimensionless oil saturation, So is oil saturation, Sorg is residual saturation of oil, Krg is the
relative permeability of gas phase and Kro is the relative permeability of oil phase, no and ng are exponents
for the oil and gas phase, respectively.
In conventional reservoirs, gas oil ratio (GOR) reaches a critical value after a period of primary
production. After that,the curves of relative permeability for gas and oil phases can be used to describe the
two-phase flow. In most of the cases, lower GOR values are observed in heavy oil reservoirs which might
contribute to the higher oil recovery. Two possible reasons are provided in the literature. One reason is that
the gas saturation is critical high for the gas to form the continuous flow. This reason is difficult to validate
in the laboratory. The other reason is that in the foamy oil flow, gas flows in the porous media in the form of
gas-in-oil dispersion instead of single phase flow. Foamy heavy oil is different from conventional oil are the
larger bubbles breaking up into dispersions migrating with the oil (Sheng 1999; Maini 1999). Moreover, the
pressure gradient in the solution gas drive is larger than that in the conventional oil to be able to mobilize
gas clusters. However, mechanisms of the solution gas drive by foamy heavy oil remain poorly understood.
Wormholes might be formed during the production of heavy oil reservoirs. Initially, they form near the
wellbore region then propagate into the formation. If the dragging force is large enough to overcome the
force holding the grains in the unconsolidated formations, fluids will be transported in the porous media
carrying the detached grains. As a result, whether or not the dragging force exceeds a critical value larger
than the grain cohesive strength can be used as a criterion of wormhole formation. Wormholes network
are characterized with high-permeability within the reservoir that can reach as high as 13 Darcies that the
communication measurement between the injection well and the production well can be obtained within
several hours compared to several months in reservoirs not wormholed. The formation of wormholes can
be detected by the seismic signals in the field by the change of the fluid bulk modulus which is reflected by
the ratio between the P-wave velocity and S-wave velocity (Tatham 1982; Tatham and Stoffa 1976). This
ratio decreases as the reservoir fluid is depleted and therefore, the movement of the fluid can be monitored
by the time-lapse seismic survey (Chopra et al. 2010).
Sand production is also a major concern for heavy oil reservoirs. Issues come along with sand production
are the equipment erosion, casing collapse, reservoir subsidence, facility plug, wellbore blockage and
reservoir conductivity decrease. Actually, thirty years ago sand production was considered to be unfavorable
to cold production. For example, Kimmel and Laviolette (1982) concluded that the low primary recovery
in the Lloydminster heavy oil reservoirs is because of the oil high viscosity and scarce solution gas drive
and high sand production (Loughead and Saltuklaroglu 1992). Sanding flow into the standing and floating
valves of the artificial lift system causing serious wear and sands settle out between the moving plunger and
the standing valves, leading to high water production rate. In the 1990s, because of the development of the
progressive cavity (PC) pump and the employment of more resistant materials used for the rubber rotor to
allow for the reduced back pressure with large-size tubing and large-torque pump rod, the oil production
is significantly increased.
Operators have noticed that it is necessary to produce high-viscosity oil along with large quantities of
sand. Sand production induces complex processes including the stress redistribution, solid/fluid coupled
fluid transport, stress redistribution and sand failure propagation. These processes are very complex that are
not fully understood yet. Chemically and physically sand production control have been performed to control
the mobility of sand. The chemical method includes resin injection; and the physical methods include the
application of slotted liners and gravel packs for the sand cleaning. In practice, these methods might not
work efficiently due to the complex mechanisms. A better completion design and optimization design are
14 SPE-193646-MS

important to improve the economics. Sand production can be favorable for oil production by increasing the
permeability and porosity of the reservoir, and the rock compaction by the reduced compressibility also
contributes to more oil production.
The high oil production in heavy oil reservoirs can be attributed to the large drawdown pressure during the
cold production process caused by the heavy oil high viscosity caused by the breakup of bubbles. Whether
or not the bubbles breakup is estimated by the capillary number which is the ratio between the viscous
dragging force and the interfacial tension, expressed as:
[8]
Where, Uo is oil velocity and μo is oil viscosity and σ is the interfacial tension. Oil recovery factor
increases with the increase of depletion rate by the solution gas drive due to the presence of a large number
of bubbles and the high gas saturation caused by the high supersaturation of oil. The reason is that a low
depletion rate gas has sufficient time to diffuse into the bubble while the gas diffusion is limited when the
depletion rate is high.

Modeling and Simulation of Heavy Oil Reservoirs


Over the past decade, upgrading investments and increasing volume of cheaper heavy oil sourced by the
development of refinery equipment which can process poorer-quality heavier crude resulted in maintaining
the demand of heavy oil and boosting the development of heavy oil projects worldwide. Thus, the feasibility
of these heavy oil projects became the driven-mechanism that evaluates whether the heavy oil reservoirs
are worth to produce, or not. Heavy and Super heavy oilfield characteristics are mainly depending on
the variations in fluid characteristics such as viscosity, density etc…and these characteristics dominates
the production behavior. On the other hand, porosity and permeability distributions, as well as reservoir
heterogeneities and fluid saturation assessments, are also controlling factors that affect heavy oil and tar
sand exploration and production (Larter, 2006). In most of the cases, exploration, production, and processing
lower-quality crude need the use of secondary and tertiary recovery methods to produce more oil from these
reservoirs. The most common problem of these processes is the mobility differences between displaced and
displacing fluids resulted in the poor reservoir volumetric sweep efficiency sourced by viscous fingering
and channeling (Mansoori, 1996).
Herein, geomechanical behavior of porous media by geomechanical modeling (computation of stress/
strain modeling, reservoir simulation (modeling multiphase flow and heat transfer in porous media)
(Gutierrez, 2001), as well as fracture mechanics (crack propagation and geometry), become crucial to
describe and support the decision making of oil and gas FDP (Settari, 2001). In SAGD projects, it is
important to capture the heterogeneity at relatively small scales to have a better understanding of its impact
on the project performance. If the well located in a place that the reservoir parameters vary rapidly, and
the model did not use a scale small enough, these changes possibly will not be taken into account and the
predictions for steam, pressure, as well as reserves, will be inaccurate. Therefore, it is critical to perform
thermal simulations which require more iterations, more memory and more computing power with accurate
models.

Geomechanical Modeling
The first study that includes a simple mechanism and analyses fluid flow problems sourced by rock
deformation introduced by Terzaghi in 1925. He used 1D consolidation theory to explain the problem.
Although the theory followed by many scientists, Biot firstly extended the phenomena to a 3D case by
including Terzaghi's effective stress principle. Biot described all the basic properties of soil by making
assumptions (isotropy of the material, linearity of stress/strain relations, water phase may contain bubbles
or air, water is incompressible inside the pores) that has to be taken into account by assuming that the water
SPE-193646-MS 15

flows through the porous media according to Darcy's Law. In a porous media that has a fluid flow of two
immiscible fluids, Darcy's Law can be written as,

[9]

Biot described the final equation for mass balance as;

[10]

Whereas,
Bπ – Formation Volume Factor for phase π
Dijkl – Constitutive Tensor
Ks – Grain Compressive Modulus
Sπ – Saturation of phase π
 qπ – Fluid Source or Sink
 vπi – Fluid Velocity for phase π
 xi – Spatial Coordinate
δij – Kronecker Delta Function
εkl – Strain Tensor
Lewis explained the contributions of the Biot's Equation as sub-terms of 4 different physical parameters
(Lewis, 1993). Each phase (π) fluid flow and saturation changing rate,

[11]

Rock Volumetric Change Rate;


[12]
Change of Solid Particle Volumetric Rate (Depending on pore pressure change);

[13]

Solid-Particle Volume change (Sourced by the change in mean effective stress)

[14]

Reservoir Simulation
Fluid movements are controlled by both microscopic (Viscosity, IFT etc.) and macroscopic (Reservoir
heterogeneity, Mobility differences etc.) effects in the reservoir. In heavy oil reservoirs, oil viscosity is
higher compared with other conventional reservoir types and mostly they have values comparable to that
of tar. Also, Interfacial Tension (IFT) of two immiscible reservoir fluids can sometimes create barriers
that prevent continuous flow paths and creates disconnection within the pores of the fluid that has lower
saturation. In other words, one of the reservoir fluids can inhibit the flow of the other fluid. Therefore,
characterization of the Interfacial Geometry becomes crucial. On the other hand, reservoir heterogeneity is
a variable parameter that sourced from the effects such as, fracturing, faulting etc… that changes in time and
space. Also, mobility differences could affect the performance of secondary and tertiary recovery methods.
Thermal effects also need to be included in thermal recovery methods (in-situ combustion, steam flooding
etc.) that are planned to perform in the heavy oil reservoir. In a heavy oil reservoir, all the effects mentioned
above need to be described carefully, a correct model needs to be chosen and reservoir simulation has to be
performed by taking account of each parameter (Gerritsen, 2005). Today, it is possible to combine modeling
16 SPE-193646-MS

of multi-component three-phase fluids with complicated PVT relations and relative permeability data by
using advanced simulators.
The aim of reservoir simulation is to build the correct model that works with correct correlations for both
single-phase and multiphase fluid flow as well as heat transfer in porous media (Peaceman, 2000; Aziz and
Settari, 1979; Munka and Pápay, 2001; Ertekin, et al., 2001; Carlson, 2003). Most of the reservoir fluid flow
problems require multiphase flow simulation rather than the single-phase. In multiphase flow simulation,
both water/oil and gas/oil interactions and movements have to be described (Mattax and Dalton, 1990).
One-dimensional fluid flow of two immiscible fluids in porous media can be written as; (Gerritsen, 2005).

[15]

In cases that the displacement of water/oil includes gas-cap expansion, gas injection and solution-gas
evolution, then three-phase simulation needs to be used. In most of the problems that involve two or three
phase flow, Black-oil simulators can bring the solution to Reservoir Engineers and Geoscientists. However,
due to the low shrinkage factor values and little amount of gas phase, Geoscientists mostly have to deal with
water/oil interactions in heavy oil reservoir simulation. In many cases, it is acceptable to use gas solubility in
oil phase only as the function of pressure change and to use fixed oil & gas compositions (Gerritsen, 2005).
In reservoirs that have a thin layer with large areal extent, the pressure gradient and z-directional flow
can be neglected. (Aziz and Settari, 1979). However, it is possible to add the effects of vertical thickness
by showing the parameters such as; block thickness, ∆z and elevation (h) as the functions of x and y. Thus,
the typical flow equation can be written in two-dimensional form as (See Figure A14);

[16]

Herein,
If we adopt that to a petroleum reservoir where the properties of a reservoir described as averages over
the thickness, reservoir parameters can be described as;

[17]

In three-dimensional case, single-phase flow can be described by using the general partial differential
equation in cartesian coordinates as;

[18]

In cases that a detailed study regarding on three-dimensional flow around a wellbore need to be
investigated, the flow equation can be written in cylindrical coordinates as well. Furthermore, Finite-
difference approximations with discretization in space and time as well as the discretization of errors,
and other explicit/implicit as well as iterative methods in the literature are developed and based on above
equations (Fanchi, 2005).
Challenges and Limitations. In general, computing a reservoir simulation accompanies limitations such
as; processing the simulation in a reasonable time, handling a large number of grid blocks without having
a problem with the software and hardware, and develop sufficient models that can capture the spatial
variations of the reservoir with fine scale grids. Scientists explored the ways to accomplish these limitations
by including the usage of LGRs (Local Grid Refinements) or maintaining the matrix size that has been
discretized near the steam chamber boundary and keeping the size of grid blocks larger inside and outside
of this chamber.
In heavy oil reservoir simulation studies, it is also important to consider the effects of operational
limitations in a heavy oil reservoir, especially in water injection and production rates. For instance,
SPE-193646-MS 17

secondary and tertiary recovery methods are best solutions to perform in heavy oil reservoirs. Description of
the rate limitations are so important due to oil viscosity is so variable in the reservoir. Exceeding the limits on
rates can decrease sweeping effect and increase fingering. Therefore, it is crucial to apply the limitations in
Reservoir Simulation to avoid any miscalculations which can dramatically affect the economic performance
of a chemical EOR process (Fortenberry, 2015).
Another difficulty in a reservoir simulation is the quality of input data which resulted in a data gap
in describing a reservoir. In a case that the input data points could not increase in direct proportion,
the prediction becomes more arbitrary by including the larger number of grid blocks. Under these
circumstances, its inevitably to reach a lower accuracy level in fractured reservoir cases (Branets et al.,
2009).
Additionally, Representative Elemental Volume (REV) of grid blocks are larger compared with the
volumes of reservoir cores. This difference also increased when dealing with fractured reservoirs that have
higher REV values. In fractured formations, the results could make no sense of instability due to the grid
block is too large to capture the physics or fluctuating and shows instability (too small grid blocks) (Lee
et al., 2001).
Latest Advances. Reservoir Simulation became one of the driven research topics for decades. Various
computer techniques applied in different scenarios (Zhang, 2009; Liu, 2015). For instance, in 1991, Rutledge
and his group developed a Parallel computer compositional simulator which uses an implicit pressure-
explicit saturation (IMPES) method. The simulator solves the unknown pressure parameter and derives
the other parameters by using existing pressure data. Shiralkar et al. implemented a parallel reservoir
simulator that able to run in different parallel systems in 1997. Compositional, Black Oil and Thermal
models described by using both Finite-element methods and Finite-difference Methods by Chen et al. in
2006.
Some other black oil simulators are developed by various scientists such as; Guan et al. (a model that
hundreds millions of grid cells), Wheeler25 (a model that uses parallel computing techniques and advanced
numerical methods), Also, a highly efficient parallel black oil simulator which uses both structured and
unstructured grids and has an ability to perform a simulation with billions of grid cells developed by Dogru
et al.26 Latest developments in Reservoir Simulation can also be viewed in topics such as; the simulation
Speed and the accuracy of the model, new fluid flow equations, Coupled Fluid Flow and Geo-mechanical
stress modeling, the modeling of fluid flow with thermal stress.
Speed and Accuracy. In recent years, the speed of computers, as well as memories, increased by doubling
in every 12 to 18 months as mentioned in Moore's law (Schaller, 1997). It paved the way of using a great
number of grid blocks and increased the accuracy by combining with the use of higher order terms in Taylor
series approximation.
New Fluid Flow Equations. In fractured reservoirs, description of the physical behavior of fractures,
fractured area, and spacing, aperture and orientation of the fractures are the driven mechanisms that
influence the movement of fluid through to the rock. In addition, these parameters also affect the
heterogeneity and anisotropy of the flow as well as the conductivity of rock and connectivity of the network
in a fractured reservoir. For this reason, dual-porosity and dual-permeability models were strongly needed
to have an accurate flow simulation through fractures. In reservoirs that has a complex geology, usage
of mechanisms such as; single-porosity single-permeability (SPSP), dual-porosity single-permeability
(DPSP), dual-porosity dual-permeability (DPDP) can vary by region and layers. Different mechanisms
such as; Dual-porosity for single phase flow (Gerke, 1993; Zimmerman, 1993), Dual-permeability (Vogel,
2000; Larsbo 2005), Dual-porosity single-permeability (Reeves, 2001), Dual-Porosity dual-permeability for
single-phase flow (Pereira et al., 2006), Dual Porosity dual-permeability for multiphase flow (Pride, 2003;
Thararoop, 2012), Triple-porosity dual-permeability (Wei, 2005; Zhou, 2015) and also Discrete fracture
(Karimi-Fard et al., 2003; Gong et al., 2008) models are studied by many petro-technical experts.
18 SPE-193646-MS

Coupled Fluid Flow and Geomechanical Stress Model


Preliminary studies of coupling Reservoir fluid flow and fracturing simulations were proposed by Hagoort
et al., Settari, and Nghiem et al. Additionally, First Hybrid grid model which was coupling Cartesian and
Cylindrical grid blocks introduced to the industry by Pedrosa and Aziz in 1986. The aim of the coupled
model is to use two different sets of grid blocks for fracture fluid flow as well as in the reservoir matrix (See
Figure A15). It uses fluid continuity equation for fluid flow in fractures and partial differential equations
(such as finite difference method) for the fluid flow in reservoir matrix. These works became a keystone for
other coupling models (Settari, 1998; Rutqvist et al., 2002; Minkoff, 2003; Ji et al., 2009).
Apart from coupling geomechanics [commonly modeled in finite element analysis (FEA)] and reservoir
simulations (in FDM) by using iterative or sequential scheme, the flow/stress coupling is mostly used for
porosity and absolute permeability as a function of effective stress (which is related to pore pressure and the
total stress) and plastic strain history. One is big challenges in thermal operations is to evaluate an increase
in permeability vs. dilation (Irani, 2018) and common correlation such Kozeny–Carman cannot be helpful,
due to complexity and combination of shear and volumetric strains. As Irani (2018) explained the pre- and
post-heating dilation results were comparable, suggesting that one of the shortcomings of the current form
of this dilative model. In some applications analytical methods such as the one presented by Irani (2017)
can show the trend of deformations accurately without any need of complex numerical simulations.

Fluid Flow Modeling with Thermal Stress


Thermal variance of reservoir rocks is also a critical parameter that affects the fracture structure. Depending
on the amount of thermo-elastic stress, new fractures can be formed, or alteration occurs in the shapes of
fractures previously existed. In heavy oil reservoirs, the main reason of thermal stress is the temperature
difference between the injected fluid and reservoir fluids during the EOR processes. Although some earliest
studies neglected the thermal stress and only mechanical stresses are taken into account (Chalaturnyk
and Scott, 1995). However, recent studies investigated the effects of thermal stress on the propagation
of fractures by performing numerical simulations on different categories like; water and gas injection/
production. (Zekri and Chaalal., 2001). Additionally, the obligation of deriving new mathematical models
which combines the thermal effects with mechanical (Hossain et al., 2009).

What Kind of Models are Hard to Model Accurately


Different reservoir types are modeled by different simulators in oil and gas industry. The simplest and most
common model to use known as "Black Oil Model". It assumes that only the gas phase dissolves in the oil
and water phase. Herein, Gas solubility is described as Pressure dependant and incorporated Gas solubility
factors (Rso and Rsw). BOM can model Capillary imbibition processes, immiscible gas injection, recovery by
fluid expansion, (Solution gas drive), waterflooding (viscous, capillary and gravity forces are calculated),
and WAG (water-alternating gas) processes. However, some cases could not succeed to give accurate results
with BOM. They require more complex reservoir simulation models which are harder to model due to
extra data needed to describe. These are; Compositional Models, Dual Porosity Models (Fractured systems),
Chemical Flood Models and Thermal Models.
Compositional Models. Compositional simulation models are needed when significant inter-phase mass
transfer effects occur in displacement processes of different fluid phases. In this model, simulation is
possible by the description of separate components of oil and gas phases. As distinct from BOM, the mass
conversation is applied to each component rather than just to "oil", "water" and "gas" phase. Compositional
models are used in recovery processes of heavy oil reservoirs by reason of it enables to model Gas injection
into near critical reservoirs, Gas recycling processes in condensate reservoirs, Gas injection with oil
mobilization by first contact or multi-contact (CO2 flooding) accurately (Mattax, 1990).
SPE-193646-MS 19

Chemical Flood Models. These models are developed to model Surfactant and Polymer Flooding in
displacement processes. Chemical Flood models can be used in Polymer flooding cases that mobility ratio
and microscopic sweep efficiency need to be improved, and Polymer-surfactant flooding that IFT between
oil and water phases needs to be lower. Also, Low-tension polymer flooding, Alkali flooding and Foam
Flooding are the other application areas.
Thermal Models. In steam injection or combustion processes that aims to reduce the viscosity of a heavy
oil reservoir, thermal models are most effective and accurate tools to perform. Cases can be classified into
three different groups such as; Steam Soaks (Huff n’ Puff Process), Steam Drive and In-situ combustion.
In Steam Soaks (Huff n Puff Process) following the steam injection, the well left in shut-in position for a
time to allow heat transmission in the reservoir to increase the mobility of oil phase. Then, the well opens
back to produce oil with less viscosity (Myhill, 1978).
In Steam "driv"e processes, there is an injection well that inject the steam continuously to the reservoir
and production wells that continue to produce. The aim is same with Huff n Puff process (Myhill, 1978).
In-situ Combustion. Cases that oxygen or air injected into the reservoir called "In-Situ Combustion".
Herein, a part of the oil phase is burned to produce combustion gases and heat which aims to mobilize the
heavy components and increase the oil production. (Barron, 1964).
Dual-Porosity Models for Fractured Systems. These models are widely used to simulate fractured systems
that have multiphase flow. They are effectively model flows both inside fracture and matrix parts as well
as the exchange of fluids between the fractured part and rock matrix.

References
1. IEA, 2005. Resources to Reserves: Oil & Gas Technologies for the Energy Markets of the Future.
International Energy Agency, Paris, France
2. Ancheyta, J., Speight, J.G., 2007. Hydroprocessing of Heavy Oils and Residua. CRC-Taylor and
Francis Group, Boca Raton, Florida.
3. Speight, J.G., 2016. Introduction to Enhanced Recovery Methods for Heavy Oil and Tar Sands.
Elsevier, New York.
4. June 2017. BP Statistical Review of World Energy. 66th Edition
5. Hein, F.J., 2006. Heavy oil and oil (tar) sands in North America: an overview & summary of
contributions. Natural Resources Research, 15(2), pp.67–84.
6. Utah Heavy Oil Program, 2007. Bauman, Jacob & Burian, Steven & Deo, Milind & Eddings,
Eric & Gani, M & Goel, R & Huang, Chih-Kun & Hogue, M & Keiter, R & Li, L & Ruple, John
& Ring, Terry. Unconventional Oils Research Report.
7. International Energy Agency, 2005. Resources to Reserves: Oil & Gas Technologies for the
Energy Markets of the Future. International Energy Agency, Paris, France
8. Speight, J.G., 2013. Enhanced recovery methods for heavy oil and tar sands. Elsevier.
9. BP Statistical Review of World Energy, 2017. 66th Edition
10. California Department of Conservation; Division of Oil, Gas & Geothermal
11. Resources. Monthly Production and Injection Databases. http://www.conservation.
12. ca.gov/DOG/prod_injection_db/index.htm (accessed December 21, 2017).
13. Canadian Association of Petroleum Producers (CAPP). Canadian Crude Oil
14. Production and Supply Forecast, 2006-2020, CAPP: Calgary, Alberta, May 2006.
15. Cornelius, C.D., 1987. Classification of Natural Bitumen: A Physical and Chemical Approach:
Section II. Characterization, Maturation, and Degradation.
16. Head I. M., Jones D. M., Larter S. R., Biological activity in the deep subsurface and the origin of
heavy oil, Nature, Vol: 426, November 2003il)
20 SPE-193646-MS

17. Larter S., Adams J., Gates I.D., Bennett B., Huang H., The Origin, Prediction and Impact of Oil
Viscosity Heterogeneity on the Production Characteristics of Tar Sand and Heavy Oil Reservoirs,
Journal of Canadian Petroleum Technology, January 2008, Volume 47, No. 1
18. Shu W. R., A Viscosity Correlation for Mixtures of Heavy Oil, Bitumen, and Petroleum Fractions,
SPE, JUNE 1984
19. Santos R. G., Loh W., Bannwart A. C., Trevisan O. V., an overview of heavy oil properties and its
recovery and transportation methods, Vol. 31, No. 03, pp. 571 – 590, July – September, 2014
20. Lashkarbolooki M., Ayatollahi S., Effect of asphaltene and resin on interfacial tension of acidic
crude oil/ sulfate aqueous solution: Experimental study, Fluid Phase Equilibria 414 (2016)
149–155
21. Mandal P. C., Wahyudiono, Sasaki M.,Goto M., Reduction of total acid number (TAN) of
naphthenic acid (NA) using supercritical water for reducing corrosion problems of oil refineries,
Fuel 94 (2012) 620–623
22. Mandal P. C., Abdullah A. B., Rahman M. M., Total Acid Number Reduction of 2, 6-
Naphthalenedicarboxylic Acid Using Subcritical Methanol for Reducing Acidity of Heavy Oil: A
Kinetic Study, Procedia Engineering 148 (2016) 1213 – 1219
23. Ramirez-Corredores M. M., The Science and Technology of Unconventional Oils, Finding
Refining Opportunities, 2017, Pages 295–385
24. BaiBing Y., ChunMing X., SuoQi Z., Samuel H. C., Keng H.C., Quan S., Thermal
transformation of acid compounds in high TAN crude oil, Chemistry of Heavy Petroleum
Fractions and Its Impacts on Refining Processes, July 2013 Vol.56 No.7: 848–855
25. Barth T., Hoiland S., Fotland P., Askvik K. M., Pedersen B. S., Borgund A. E., Acidic
compounds in biodegraded petroleum, Organic Geochemistry 35 (2004) 1513–1525
26. Davudov D., Moghanloo R. G., A systematic comparison of various upgrading techniques for
heavy oil, Journal of Petroleum Science and Engineering 156 (2017) 623–632
27. Pevneva G.C., Golovko A.K., Korneev D.S., Levashova A.I, Thermal Conversion of Heavy
Oil Systems and Analysis of Structural Changes of their High Components with PMR Method,
Procedia Chemistry 10 (2014) 15 – 19.
28. Speight J. G., Introduction to Enhanced Recovery Methods for Heavy Oil and Tar Sands, Second
Edition, 2016
29. Lin C. Y., Tjeerdema R. S., Crude Oil, Oil, Gasoline and Petrol, Ecotoxicology, 2008
30. Ilyin S., Arinina M., Polyakova M., Bondarenko G., Konstantinov I., Kulichikhin V., Malkin A.,
Asphaltenes in heavy crude oil: Designation, precipitation, solutions, and effects on viscosity,
Journal of Petroleum Science and Engineering 147 (2016) 211–217
31. Powers D. P., Sadeghi H., Yarranton H. W., van den Berg F. G. A., Regular solution based
approach to modeling asphaltene precipitation from native and reacted oils: Part 1, molecular
weight, density, and solubility parameter distributions of asphaltenes, Fuel 178 (2016) 218–233
32. Arciniegas L. M., Babadagli T., Quantitative and visual characterization of asphaltenic
components of heavy-oil after solvent interaction at different temperatures and pressures, Fluid
Phase Equilibria 366 (2014) 74–87
33. Chopra S., Lines L., Schmitt D. R., Batzle M., Heavy Oils: Reservoir Characterization and
Production Monitoring, Geophysical Developments Series, 2010
34. Ilyin S. O., Arinina M. P., Polyakova M. Y., Kulichikhin V. G., Malkin A. Y., Rheological
comparison of light and heavy crude oils, Fuel 186 (2016) 157–167
35. Bartle K. D., Jones J. M., Lea-Langton A. R., Pourkashanian M., Ross A. B., Thillaimuthu J. S.,
Waller P.R., Williams A., The combustion of droplets of high-asphaltene heavy oils, Fuel 103
(2013) 835–842
SPE-193646-MS 21

36. Arciniegas L. M., Babadagli T., Quantitative and visual characterization of asphaltenic
components of heavy-oil after solvent interaction at different temperatures and pressures, Fluid
Phase Equilibria 366 (2014) 74–87
37. Langevin D., Argillier J-F., Interfacial behavior of asphaltenes, Advances in Colloid and
Interface Science 233 (2016) 83–93
38. Mousavi M., Abdollahi T., Pahlavan F., Fini E. H., The influence of asphaltene-resin molecular
interactions on the colloidal stability of crude oil, Fuel 183 (2016) 262–271
39. Zhang J., Tian D., Lin M., Yang Z., Dong Z., Effect of resins, waxes and asphaltenes on water-oil
interfacialproperties and emulsion stability, Colloids and Surfaces A: Physicochem. Eng. Aspects
507 (2016) 1–6
40. Zendehboudi S., Shafiei A., Bahadori A., James L. A., Elkamel A., Lohi A., Asphaltene
precipitation and deposition in oil reservoirs –Technical aspects, experimental and hybrid neural
networkpredictive tools, Chemical Engineering Research and Design 92 (2014) 857–875
41. Arciniegas L. M., Babadagli T., Asphaltene precipitation, flocculation and deposition during
solvent injection at elevated temperatures for heavy oil recovery, Fuel 124 (2014) 202–211
42. Luo P., Gu Y., Effects of asphaltene content on the heavy oil viscosity at different temperatures,
Fuel 86 (2007) 1069–1078
43. Ali M. F., Abbas S., A review of methods for the demetallization of residual fuel oils, Fuel
Processing Technology 87 (2006) 573–584
44. Flahaut D., Minvielle M., Sambou A., Lecour P., Legens C., Barbier J., Identification of sulphur,
oxygen and nitrogen species in heavy oils by X-ray photoelectron spectroscopy, Fuel 202 (2017)
307–317
45. Khuhawar M.Y., Mirza M. A., Jahangir T. M., Determination of Metal Ions in Crude Oils, Crude
Oil Emulsions- Composition Stability and Characterization, 2012 Prof. Manar El-Sayed Abdul-
Raouf (Ed.)
46. Javadli R., Klerk A., Desulfurization of heavy oil, Appl Petrochem Res (2012) 1:3–19
47. Chen X. Y., Gao J. J., Lu Y. Z., Meng H., Li C. X., Acylation desulfurization of heavy cracking
oil as a supplementary oil upgrading pathway, Fuel Processing Technology 130 (2015) 7–11
48. Zhu G., Wang M., Zhang T., Identification of polycyclic sulfides hexahydrodibenzothiophenes
and their implications for heavy oil accumulation in ultra-deep strata in Tarim Basin, Marine and
Petroleum Geology 78 (2016) 439–447
49. Martinez-Palou R., Mosqueira M. L., Zapata-Rendon B., Mar-Juarez E., Bernal-Huicochea C.,
Clavel-Lopez J. C., Aburto J., Transportation of heavy and extra-heavy crude oil by pipeline: A
review, Journal of Petroleum Science and Engineering 75 (2011) 274–282
50. Speight, J., G., Introduction to Enhanced Recovery Methods for Heavy Oil and Tar Sands, Heavy
Oil and Tar Sand Bitumen, 2016, 3–48
51. Coal, Oil Shale, Natural Bitumen, Heavy Oil And Peat – Vol. II – Natural Bitumen (Tar Sands)
and Heavy Oil -James G. Speight, 2009
52. Guo, K., Li H., Yu Z., In-situ heavy and extra-heavy oil recovery: A review, Fuel 185 (2016)
886–902
53. Hashemi R., Nassar N. N., Almao P. P., Nanoparticle technology for heavy oil in-situ upgrading
and recovery enhancement: Opportunities and challenges, Applied Energy 133 (2014) 374–387
54. Redelius P., Soenen H., Relation between bitumen chemistry and performance, Fuel 140 (2015)
34–43
55. Speight J. G., Heavy and Extra-heavy Oil Upgrading Technologies, Refining Heavy Oil and
Extra-heavy Oil, 2013, 1–13
22 SPE-193646-MS

56. Attanasi, E. D.,Meyer, R. F., Natural bitumen and extra-heavy oil, in 2010 Survey of Energy
Resources, eds., J. Trinnaman and A. Clarke: World Energy Council, 2010, p. 123–150.
57. Gray M. R., Fundamentals of Bitumen Coking Processes Analogous to Granulations: A Critical
Review, The Canadian Journal of Chemical Engineering, Volume 80, June 2002
58. Muraza O., Hydrous pyrolysis of heavy oil using solid acid minerals for viscosity reduction,
Journal of Analytical and Applied Pyrolysis 114 (2015) 1–10
59. Hasan S. W., Ghannam M. T., Esmail N., Heavy crude oil viscosity reduction and rheology for
pipeline transportation, Fuel 89 (2010) 1095–1100
60. N.H. Abdurahman a,n, Y.M. Rosli a, N.H. Azhari b, B.A. Hayder, Pipeline transportation
of viscous crudes as concentrated oil-in-water emulsions, Journal of Petroleum Science and
Engineering 90-91 (2012) 139–144
61. Ahmadi, Mohammad Ali, Sohrab Zendehboudi, Alireza Bahadori, Lesley James, Ali Lohi, Ali
Elkamel, Ioannis Chatzis. 2014. Recovery Rate of Vapor Extraction in Heavy Oil Reservoirs—
Experimental, Statistical, and Modeling Studies. Industrial & Engineering Chemistry Research
53 (41): 16091–16106. 10.1021/ie502475t.
62. Al-Bahlani, Al-Muatasim, Tayfun Babadagli. 2012. Laboratory scale experimental analysis of
Steam-Over-Solvent injection in Fractured Reservoirs (SOS-FR) for heavy-oil recovery. Journal
of Petroleum Science and Engineering 84: 42–56. http://dx.doi.org/10.1016/j.petrol.2012.01.021.
63. Al Bahlani, As Muatasim Mohammad, Tayfun Babadagli. 2008. A Critical Review of the
Status of SAGD: Where Are We and What Is Next? Presented at the SPE Western Regional and
Pacific Section AAPG Joint Meeting, Bakersfield, California, USA, 29 March-4 April, https://
doi.org/10.2118/113283-MS.
64. Ali, S. M. Farouq. 1976. Non-Thermal Heavy Oil Recovery Methods. Presented at the SPE
Rocky Mountain Regional Meeting, Casper, Wyoming, 11-12 May, https://doi.org/10.2118/5893-
MS.
65. Brook, G., A. Kantzas. 1998. Evaluation of Non-thermal EOR Techniques For Heavy Oil
Production. Presented at the Annual Technical Meeting, Calgary, Alberta, June 8 – 10, https://
doi.org/10.2118/98-45.
66. Butler, R. M., I. J. Mokrys. 1991. A New Process (VAPEX) For Recovering Heavy Oils Using
Hot Water And Hydrocarbon Vapour. Journal of Canadian Petroleum Technology 30 (01).
https://doi.org/10.2118/91-01-09.
67. Cavender, Travis. 2004. Summary of Multilateral Completion Strategies Used in Heavy Oil
Field Development. Presented at the SPE International Thermal Operations and Heavy Oil
Symposium and Western Regional Meeting, Bakersfield, California, 16-18 March, https://
doi.org/10.2118/86926-MS.
68. Clark, PD, JB Hyne. 1990. Studies on the chemical reactions of heavy oils under steam
stimulation condition. Aostra J Res 29 (6): 29–39.
69. Clark, Peter D, Robert A Clarke, James B Hyne, Kevin L Lesage. 1990. Studies on the effect of
metal species on oil sands undergoing steam treatments. Aostra J Res 6 (1): 53–64.
70. Dong, Mingzhe, S. S. Sam Huang, Keith Hutchence. 2006. Methane Pressure-Cycling Process
With Horizontal Wells for Thin Heavy-Oil Reservoirs. SPE Reservoir Evaluation & Engineering
9 (02). https://doi.org/10.2118/88500-PA.
71. Greff, John, Tayfun Babadagli. 2011. Catalytic Effects of Nano-Size Metal Ions in Breaking
Asphaltene Molecules during Thermal Recovery of Heavy-Oil. Presented at the SPE Annual
Technical Conference and Exhibition, Denver, Colorado, USA, 30 October-2 November https://
doi.org/10.2118/146604-MS.
SPE-193646-MS 23

72. Hamedi Shokrlu, Yousef, Tayfun Babadagli. 2010. Effects of Nano-Sized Metals on Viscosity
Reduction of Heavy Oil/Bitumen During Thermal Applications. Presented at the Canadian
Unconventional Resources and International Petroleum Conference, 19-21 October, Calgary,
Alberta, Canada, https://doi.org/10.2118/137540-MS.
73. Huang, Tuo, Xiang Zhou, Huaijun Yang, Guangzhi Liao, Fanhua Zeng. 2017. CO2 flooding
strategy to enhance heavy oil recovery. Petroleum 3 (1): 68–78. http://dx.doi.org/10.1016/
j.petlm.2016.11.005.
74. Jain, Rachna, Shivani Syal, Ted A. Long, Robert C. Wattenbarger, Ivan J. Kosik. 2013.
An Integrated Approach To Design Completions for Horizontal Wells for Unconventional
Reservoirs. SPE Journal 18 (06). https://doi.org/10.2118/147120-PA.
75. Jia, Xinfeng, Jianli Li, Zhangxin Chen. 2015. Mathermatical Modeling of Dynamic Mass
Transfer in Cyclic Solvent Injection. Presented at the SPE Canada Heavy Oil Technical
Conference, Calgary, Alberta, Canada, 9-11 June, https://doi.org/10.2118/174519-MS.
76. Lu, Teng, Zhaomin Li, Songyan Li, Shangqi Liu, Xingmin Li, Peng Wang, Zhuangzhuang Wang.
2015. Behaviors of Foamy Oil Flow in Solution Gas Drive at Different Temperatures. Transport
in Porous Media 109 (1): 25–42. https://doi.org/10.1007/s11242-015-0499-4.
77. Mai, A., J. Bryan, N. Goodarzi, A. Kantzas. 2009. Insights Into Non-Thermal Recovery of Heavy
Oil. Journal of Canadian Petroleum Technology 48 (03). https://doi.org/10.2118/09-03-27.
78. Maini, Brij B., Bashir Busahmin. 2010. Foamy Oil Flow and its Role in Heavy Oil Production.
AIP Conference Proceedings 1254 (1): 103–108. http://dx.doi.org/10.1063/1.3453794.
79. Meyer, R. F. 1997. World Heavy Crude Oil Resources. Presented at the 15th World Petroleum
Congress, Beijing, China, 12-17 October.
80. Nasehi Araghi, Majid, Koorosh Asghari. 2010. Use of CO2 in Heavy-Oil Waterflooding.
Presented at the SPE International Conference on CO2 Capture, Storage, and Utilization, New
Orleans, Louisiana, USA, 10-12 November, https://doi.org/10.2118/139672-MS.
81. Rivas, O. R., R. E. Campos, L. G. Borges. 1988. Experimental Evaluation of Transition Metals
Salt Solutions as Additives in Steam Recovery Processes. Presented at the SPE Annual Technical
Conference and Exhibition, Houston, Texas, 2-5 October, https://doi.org/10.2118/18076-MS.
82. Sahni, Akshay, Mridul Kumar, Richard B. Knapp. 2000. Electromagnetic Heating Methods for
Heavy Oil Reservoirs. Presented at the SPE/AAPG Western Regional Meeting, Long Beach,
California, 19-22 June, https://doi.org/10.2118/62550-MS.
83. Selby, Rawya, A. A. Alikhan, S. M. Farouq Ali. 1989. Potential Of Non-Thermal Methods
For Heavy Oil Recovery. Journal of Canadian Petroleum Technology 28 (04). https://
doi.org/10.2118/89-04-02.
84. Sheng, J. J., B. B. Maini, R. E. Hayes, W. S. Tortike. 1999. Critical Review of Foamy Oil Flow.
Transport in Porous Media 35 (2): 157–187. https://doi.org/10.1023/a:1006575510872.
85. Shokrlu, Yousef Hamedi, Tayfun Babadagli. 2011. Transportation and Interaction of Nano and
Micro Size Metal Particles Injected to Improve Thermal Recovery of Heavy-Oil. Presented at
the SPE Annual Technical Conference and Exhibition, Denver, Colorado, USA, 30 October-2
November, https://doi.org/10.2118/146661-MS.
86. Smith, Gerald E. 1988. Fluid Flow and Sand Production in Heavy-Oil Reservoirs Under
Solution-Gas Drive. SPE Production Engineering 3 (02). https://doi.org/10.2118/15094-PA.
87. Speight, James G. 2013. Enhanced recovery methods for heavy oil and tar sands, Elsevier
(Reprint).
88. Technology, Alberta Oil Sands, Research Authority, JB Hyne. 1986. Aquathermolysis: A Synopsis
of Work on the Chemical Reaction Between Water (Steam) and Heavy Oil Sands during Simulated
Steam Stimulation, Alberta Oil Sands Technology & Research Authority (Reprint).
24 SPE-193646-MS

89. Xiao, Lei, Gang Zhao. 2017. Estimation of CHOPS Wormhole Coverage from Rate/Time Flow
Behaviors. SPE Journal Paper. https://doi.org/10.2118/157935-PA.
90. Yadali Jamaloei, Benyamin, Mingzhe Dong, Nader Mahinpey, Brij B. Maini. 2012. Enhanced
Cyclic Solvent Process (ECSP) for Heavy Oil and Bitumen Recovery in Thin Reservoirs. Energy
& Fuels 26 (5): 2865–2874. 10.1021/ef300152b.
91. Yadali Jamaloei, Benyamin, Mingzhe Dong, Ping Yang, Daoyong Yang, Nader Mahinpey. 2013.
Impact of solvent type and injection sequence on Enhanced Cyclic Solvent Process (ECSP) for
thin heavy oil reservoirs. Journal of Petroleum Science and Engineering 110: 169–183. http://
dx.doi.org/10.1016/j.petrol.2013.08.028.
92. Bauget, F., and Lenormand, R. 2002. Mechanisms of Bubble Formation by Pressure Decline
in Porous Media: a Critical Review. SPE Annual Technical Conference and Exhibition, San
Antonio, Texas, 29 September-2 October. SPE-77457-MS. http://doi.org/10.2118/77457-MS.
93. Chen, Z. 2006a. Heavy oils, Part I. SIAM News 39 (3)
94. Chen, Z. 2006b. Heavy oils, Part II. SIAM News 39 (4)
95. Chopra, S., Lines, L., Schmitt, D. et al. 2010. Heavy Oils: Reservoir Characterization
and Production Monitoring, Society of Exploration Geophysicists. https://
doi.org/10.1190/1.9781560802235.
96. Kimmel, T. B., and Laviolette, D. J. 1982. Heavy Oil Production Problems in the Lloydminster
Area. Thirty-Third Annual Meeting of the CIM, Calgary, Alberta, June 6-9.
97. Loughead, D. J., and Saltuklaroglu, M. 1992. Lloydminster heavy oil-why so unusual? Ninth
Annual Heavy Oil and Oil Sand Technology Symposium, Calgary, Alberta, 11 March.
98. Maini, B. B. 1999. Foamy Oil Flow in Primary Production of Heavy Oil under Solution
Gas Drive. SPE Annual Technical Conference and Exhibition, Houston, Texas, 3-6 October.
SPE-56541-MS. http://doi.org/10.2118/56541-MS.
99. Sheng, J. J. 1999. Critical Review of Foamy Oil Flow. Transport in Porous Media 35 (2):
157–187. http://doi.org/doi.org/10.1023/A:1006575510872.
100. Tatham, R. 1982. Vp/Vs and lithology. GEOPHYSICS 47 (3): 336–344. http://
doi.org/10.1190/1.1441339.
101. Tatham, R., and Stoffa, P. 1976. Vp/Vs—A POTENTIAL HYDROCARBON INDICATOR.
GEOPHYSICS 41 (5): 837–849. http://doi.org/10.1190/1.1440668.
102. Larter, S. R., Adams, J., Gates, I. D., Bennett, B., & Huang, H. (2006, January). The origin,
prediction and impact of oil viscosity heterogeneity on the production characteristics of tar
sand and heavy oil reservoirs. In Canadian International Petroleum Conference. Petroleum
Society of Canada.
103. Mansoori, G. A. (1997). Modeling of asphaltene and other heavy organic depositions. Journal
of petroleum science and engineering, 17(1-2), 101–111.
104. Gutierrez, M., Lewis, R. W., & Masters, I. (2001). Petroleum reservoir simulation coupling
fluid flow and geomechanics. SPE Reservoir Evaluation & Engineering, 4(03), 164–172.
105. Settari, A., & Walters, D. A. (2001). Advances in coupled geomechanical and reservoir
modeling with applications to reservoir compaction. Spe Journal, 6(03), 334–342.
106. Terzaghi, K. (1925). Principles of soil mechanics, IV—Settlement and consolidation of clay.
Engineering News-Record, 95(3), 874–878.
107. Biot, M. A. (1941). General theory of three-dimensional consolidation. Journal of applied
physics, 12(2), 155–164.
108. Lewis, R. W., & Sukirman, Y. (1993). Finite element modelling of three-phase flow in
deforming saturated oil reservoirs. International Journal for Numerical and Analytical
Methods in Geomechanics, 17(8), 577–598.
SPE-193646-MS 25

109. Gerritsen, M. G., & Durlofsky, L. J. (2005). Modeling fluid flow in oil reservoirs. Annu. Rev.
Fluid Mech., 37, 211–238.
110. Peaceman, D. W. (2000). Fundamentals of numerical reservoir simulation (Vol. 6). Elsevier.
111. Aziz, K., & Settari, A. (1979). Petroleum reservoir simulation. Chapman & Hall.
112. Munka, M., & Pápay, J. (2001). 4D numerical modeling of petroleum reservoir recovery.
Akadémiai Kiadó.
113. Ertekin, T., Abou-Kassen, J. H., & King, G. R. (2001). Basic Applied Reservoir Simulations.
Society of Petroleum Engineers.
114. Carlson, M. R. (2003). Practical reservoir simulation: using, assessing, and developing
results. PennWell Books.
115. Mattax, C. C., & Dalton, R. L. (1990). Reservoir Simulation (includes associated papers 21606
and 21620). Journal of Petroleum Technology, 42(06), 692–695.
116. Fanchi, J. R. (2005). Principles of applied reservoir simulation. Gulf Professional Publishing.
117. Fortenberry, R., Suniga, P., Mothersele, S., Delshad, M., Lashgari, H., & Pope, G. A. (2015,
June). Selection of a Chemical EOR Strategy in a Heavy Oil Reservoir Using Laboratory
Data and Reservoir Simulation. In SPE Canada Heavy Oil Technical Conference. Society of
Petroleum Engineers.
118. Branets, L. V., Ghai, S. S., Lyons, S. L., & Wu, X. H. (2009). Challenges and technologies in
reservoir modeling. Communications in Computational Physics, 6(1), 1.
119. Lee, S. H., Lough, M. F., & Jensen, C. L. (2001). Hierarchical modeling of flow in naturally
fractured formations with multiple length scales. Water Resources Research, 37(3), 443–455.
120. Zhang, L. B. (2009). A parallel algorithm for adaptive local refinement of tetrahedral meshes
using bisection. Numer. Math.: Theory, Methods and Applications, 2, 65–89
121. Liu, H., Wang, K., Chen, Z., Jordan, K. E., Luo, J., & Deng, H. (2015, October). A parallel
framework for reservoir simulators on distributed-memory supercomputers. In SPE/IATMI
Asia Pacific Oil & Gas Conference and Exhibition. Society of Petroleum Engineers.
122. Rutledge, J. M., Jones, D. R., Chen, W. H., & Chung, E. Y. (1992). The use of a massively
parallel SIMD computer for reservoir simulation. SPE Computer Applications, 4(04), 16–21.
123. Shiralkar, G. S., Stephenson, R. E., Joubert, W., Lubeck, O., & van Bloemen Waanders,
B. (1998). FALCON: A production quality distributed memory reservoir simulator. SPE
Reservoir Evaluation & Engineering, 1(05), 400–407.
124. Chen, Z., Huan, G., & Ma, Y. (2006). Computational methods for multiphase flows in porous
media. Society for Industrial and Applied Mathematics.
125. Guan, W., Qiao, C., Zhang, H., Zhang, C. S., Zhi, M., Zhu, Z., Li, Z. (2015, September).
On Robust and Efficient Parallel Reservoir Simulation on Tianhe-2. In SPE Reservoir
Characterisation and Simulation Conference and Exhibition. Society of Petroleum Engineers.
126. Wheeler, M. F. (2002). Advanced techniques and algorithms for reservoir simulation, II:
The multiblock approach in the integrated parallel accurate reservoir simulator (IPARS). In
Resource Recovery, Confinement, and Remediation of Environmental Hazards (pp. 9–19).
Springer, New York, NY.
127. Dogru, A. H., Fung, L. S., Middya, U., Al-Shaalan, T. M., Pita, J. A., HemanthKumar, K.,
… & Hahn, W. A. (2009, October). A next-generation parallel reservoir simulator for giant
reservoirs. In SPE/EAGE Reservoir Characterization & Simulation Conference.
128. Schaller, R. R. (1997). Moore's law: past, present and future. IEEE spectrum, 34(6), 52–59.
129. Gerke, H. H., & Genuchten, M. V. (1993). A dual-porosity model for simulating the
preferential movement of water and solutes in structured porous media. Water resources
research, 29(2), 305–319.
26 SPE-193646-MS

130. Zimmerman, R. W., Chen, G., Hadgu, T., & Bodvarsson, G. S. (1993). A numerical dual-
porosity model with semianalytical treatment of fracture/matrix flow. Water resources
research, 29(7), 2127–2137.
131. Vogel, T., Gerke, H. H., Zhang, R., & Van Genuchten, M. T. (2000). Modeling flow and
transport in a two-dimensional dual-permeability system with spatially variable hydraulic
properties. Journal of Hydrology, 238(1), 78–89.
132. Larsbo, M., Roulier, S., Stenemo, F., Kasteel, R., & Jarvis, N. (2005). An improved dual-
permeability model of water flow and solute transport in the vadose zone. Vadose Zone
Journal, 4(2), 398–406.
133. Reeves, S., & Pekot, L. (2001, January). Advanced reservoir modeling in desorption-
controlled reservoirs. In SPE Rocky Mountain Petroleum Technology Conference. Society of
Petroleum Engineers.
134. Pereira, C. A., Kazemi, H., & Ozkan, E. (2006). Combined effect of non-Darcy flow and
formation damage on gas well performance of dual-porosity and dual-permeability reservoirs.
SPE Reservoir Evaluation & Engineering, 9(05), 543–552.
135. Pride, S. R., & Berryman, J. G. (2003). Linear dynamics of double-porosity dual-permeability
materials. I. Governing equations and acoustic attenuation. Physical Review E, 68(3), 036603.
136. Thararoop, P., Karpyn, Z. T., & Ertekin, T. (2012). Development of a multi-mechanistic, dual-
porosity, dual-permeability, numerical flow model for coalbed methane reservoirs. Journal of
Natural Gas Science and Engineering, 8, 121–131.
137. Wei, X., Wang, G., & Massarotto, P. (2005, January). A review on recent advances in the
numerical simulation for coalbed methane recovery process. In SPE Asia Pacific Oil and Gas
Conference and Exhibition. Society of Petroleum Engineers.
138. Zou, M., Wei, C., Yu, H., & Song, L. (2015). Modeling and application of coalbed methane
recovery performance based on a triple porosity/dual permeability model. Journal of Natural
Gas Science and Engineering, 22, 679–688.
139. Karimi-Fard, M., Durlofsky, L. J., & Aziz, K. (2003, January). An efficient discrete fracture
model applicable for general purpose reservoir simulators. In SPE Reservoir Simulation
Symposium. Society of Petroleum Engineers.
140. Gong, B., Karimi-Fard, M., & Durlofsky, L. J. (2008). Upscaling discrete fracture
characterizations to dual-porosity, dual-permeability models for efficient simulation of flow
with strong gravitational effects. SPE Journal, 13(01), 58–67.
141. Settari, A., Kry, P. R., & Yee, C. T. (1989). Coupling of fluid flow and soil behaviour to model
injection into uncemented oil sands. Journal of Canadian Petroleum Technology, 28(01).
142. Hagoort, J., Weatherill, B. D., & Settari, A. (1980). Modeling the propagation of waterflood-
induced hydraulic fractures. Society of Petroleum Engineers Journal, 20(04), 293–303.
143. Nghiem, L. X., Forsyth, P. A.Jr, & Behie, A. (1984). A fully implicit hydraulic fracture model.
Journal of petroleum technology, 36(07), 1–191.
144. Settari, A., & Walters, D. A. (2001). Advances in coupled geomechanical and reservoir
modeling with applications to reservoir compaction. Spe Journal, 6(03), 334–342.
145. Pedrosa, O. A.Jr, & Aziz, K. (1986). Use of a hybrid grid in reservoir simulation. SPE
Reservoir Engineering, 1(06), 611–621.
146. Ji, L., Settari, A., & Sullivan, R. B. (2009). A novel hydraulic fracturing model fully coupled
with geomechanics and reservoir simulation. Spe Journal, 14(03), 423–430.
147. Settari, A., & Mourits, F. M. (1998). A coupled reservoir and geomechanical simulation
system. Spe Journal, 3(03), 219–226.
SPE-193646-MS 27

148. Rutqvist, J., Wu, Y. S., Tsang, C. F., & Bodvarsson, G. (2002). A modeling approach for
analysis of coupled multiphase fluid flow, heat transfer, and deformation in fractured porous
rock. International Journal of Rock Mechanics and Mining Sciences, 39(4), 429–442.
149. Minkoff, S. E., Stone, C. M., Bryant, S., Peszynska, M., & Wheeler, M. F. (2003). Coupled
fluid flow and geomechanical deformation modeling. Journal of Petroleum Science and
Engineering, 38(1), 37–56.
150. Huang, H., Wattenbarger, R., Gai, X., Brown, W. P., Hehmeyer, O. J., Wang, J., & Long, T. A.
(2013). Using a fully coupled flow and geomechanical simulator to model injection into heavy
oil reservoirs. International Journal for Numerical Methods in Fluids, 71(6), 671–686.
151. Agarwal, A., & Kovscek, A. R. (2013, April). Solar-generated steam for heavy-oil recovery: A
coupled geomechanical and reservoir modeling analysis. In SPE Western Regional & AAPG
Pacific Section Meeting 2013 Joint Technical Conference. Society of Petroleum Engineers.
152. Chalaturnyk, R., & Scott, J. D. (1995, January). Geomechanics issues of steam assisted gravity
drainage. In SPE International Heavy Oil Symposium. Society of Petroleum Engineers.
153. Zekri, A. Y., & Chaalal, O. (2001, January). Thermal stress of carbonate rocks: an
experimental approach. In SPE Western Regional Meeting. Society of Petroleum Engineers.
154. Hossain, M. E., Mousavizadegan, S. H., & Islam, M. R. (2009). Effects of memory on the
complex rock-fluid properties of a reservoir stress-strain model. Petroleum Science and
Technology, 27(10), 1109–1123.
155. Myhill, N. A., & Stegemeier, G. L. (1978). Steam-drive correlation and prediction. Journal of
Petroleum Technology, 30(02), 173–182.
156. Barron, J. M. (1964). U.S. Patent No. 3,120,264 A. Washington, DC: U.S. Patent and
Trademark Office.
157. Irani, M., 2018. Predicting Geomechanical Dynamics of Steam-Assisted Gravity-Drainage
Process. Part I: Mohr-Coulomb (MC) Dilative Model, SPE J. 23(4): 1223—1247. http://
dx.doi.org/10.2118/SPE-189976-PA
158. Ghannadi, S., Irani, M. and Chalaturnyk, R., 2016. Overview of Performance and Analytical
Modeling Techniques for Electromagnetic Heating and Applications to Steam-Assisted-
Gravity-Drainage Process Startup, SPE J. 21 (2): 311—333. http://dx.doi.org/10.2118/178427-
PA.
159. Irani, M. and Gates, I., 2017. On Subcool Control in Steam-Assisted-Gravity-Drainage
Producers—Part III: Efficiency of Subcool Trapping in Nsolv Process, SPE J. Preprint. http://
dx.doi.org/10.2118/SPE-191355-PA.
160. Irani, M. and Ghannadi, S., 2017. On Temperature Fall-Off Interpretation in Circulation Phase
of Steam-Assisted Gravity Drainage Process, SPE Canada Heavy Oil Technical Conference,
Calgary, Alberta, Canada, 13-14 March. http://dx.doi.org/10.2118/189740-MS.

Authors
Cenk Temizel is a reservoir engineer at Aera Energy LLC (a Shell-ExxonMobil Affiliate) in Bakersfield,
California, USA. He has more than 10 years of experience in the industry working on reservoir simulation,
smart fields, heavy oil, optimization, field development, EOR with Schlumberger, Halliburton and Aera
Energy LLC in the Middle East, the US and the UK. He holds a BS degree (Honors) from Middle East
Technical University – Ankara (2003) and an MS degree (2005) from University of Southern California
(USC), Los Angeles, CA both in petroleum engineering. He served as a teaching/research assistant at USC
and Stanford University before joining the industry. He received Halliburton Innovation Award in 2012
along with 2 issued US patents in the area of reservoir management and smart field workflows and 2nd place
28 SPE-193646-MS

at SPE Global R&D Competition at SPE ATCE 2014 in Amsterdam and 22th WPC 2017 in Istanbul. He has
several publications in SPE and serves as a technical reviewer for journals.
E-mail: ctemizel@aeraenergy.com

Celal Hakan Canbaz is a Senior Reservoir Engineer with 15 years of Reservoir Engineering and
Institutional Research experience. He worked as Reservoir Domain Champion of Schlumberger Testing
Services in Iraq Geomarket and previously as Researcher in ADNOC group in United Arab Emirates.
He holds a Bachelor and Master of Science degrees in Petroleum and Natural Gas Engineering
from Istanbul Technical University, Turkey. He is an expert in SCAL analysis, Reservoir Wettability
Characterization, Well Testing Analysis, Perforation and Testing Design, Multiphase Flow Meters and PVT
Data Interpretation. As a Reservoir Lead, he involved more than 30 projects with Shell, BP, ExxonMobil
and several oil companies. He has several publications and a patent in his area of expertise. He is also
the winner of several paper contests including the SPE Middle East Young Professionals Paper Contest in
2010. Currently, He is a PhD candidate in Energy Engineering department of Ege University Solar Energy
Institute.

Zumra Peksaglam is a PhD student at Chemical engineering department of University of Southern


California. Her major focus is photo-responsive surfactants effects on macromolecule (e.g. proteins,
enzymes, nucleic acids) structure and functions under different light wavelength. She also volunteers for
the projects centered on renewable energy sources as well as enhanced oil recovery. She holds Chemical
engineering MS degree from University of Southern California in 2016 and BS degree from Gazi University
in 2012.
Email: peksagla@usc.edu

Minh Tran received a B.S. degree with summa cum laude in Petroleum Engineering and a minor in Geology
at the University of Oklahoma, Norman, USA in 2012. In 2014, he received a M.S. degree in Petroleum
Engineering at the Stanford University, Stanford, California, USA. He worked as a production engineer in
Belridge and Lost Hills oil fields for Aera Energy LLC (Bakersfield, California, USA) from 2014 to 2015.
After that, he served as a reservoir engineer for PetroVietnam between 2015 and 2016, working in block
B multinational project in the Gulf of Thailand. Currently, under a full Provost fellowship funding, he is a
Ph.D. candidate in Petroleum Engineering at the University of Southern California, Los Angeles, California,
USA. His research interest focuses on quantitative characterization of stimulated reservoir volume in
unconventional reservoir by the integration of geomechanically-coupled fluid flow, micro-seismicity and
soft computing applications. He can be reached by phone at 213-235-6696 or email tranmt@usc.edu

Karthik Balaji is a doctoral candidate at the University of North Dakota in the field of Petroleum
Engineering. His major focus is on development of data mining technologies and artificial intelligence
in unconventional plays. Prior to his present position, he was a researcher at the University of Southern
California focusing on Pressure Transient Analysis of mature wells in California and the integration of
artificial intelligence in waterflooding. He has a Master degree from the University of Southern California
and a Bachelors degree in Chemical Engneering from the University of Mumbai.

Elsayed Abdelfatah is a postdoctoral research fellow with Canada Excellence Research Chair (CERC) in
materials engineering for unconventional oil reservoirs at the university of Calgary. He currently works on
developing new ionic liquids based fluid systems for bitumen extraction and well stimulation. He also works
on Viscosification of scCO2 using nanoparticles. Before joining CERC, Elsayed worked as a postdoctoral
research fellow at Petroleum engineering department at university of Oklahoma. He worked on developing
nanoparticles-based acid diversion systems for carbonate reservoirs. He has several technical conferences
SPE-193646-MS 29

and peer-reviewed journal papers. Elsayed holds a B.Sc. and M.Sc. degrees in petroleum engineering from
Suez University in Egypt and Ph.D. in petroleum engineering from University of Oklahoma in USA.

Bao Jia is a doctoral candidate at the University of Kansas. His research focuses on numerical modeling
and experimental work of CO2 huff-n-puff process in oil shale reservoirs and investigating shale gas flow
mechanisms during the Ph.D. period. He holds a BS degree from China University of Petroleum (East
China) in Oil & Gas Transportation and Storage, and a Masters degree from New Mexico Tech in Petroleum
Engineering. He worked on nanosilica-stabilized CO2 foam for mobility control and enhanced oil recovery
(EOR) with and without the presence of surfactant during his Master period.

Dr. Fred Aminzadeh is Professor of Petroleum Engineering and Electrical Engineering, and Executive
Director of its Global Energy Network (GEN.usc.edu). He also leads USC Reservoir Monitoring
Consortium (RMC.usc.edu), Induced Seismicity Consortium (ISC.usc.edu) and Center for Geothermal
Studies (CGS.usc.edu). He previously worked in technical and management positions, including manager
of geophysical technology at Unocal (now Chevron). He was also president and CEO of dGB-USA
(dgbes.com and member of technical staff at Bell Laboratories. He also consulted at several National
Laboratories including Lawrence Berkeley (LBNL), Lawrence Livermore (LLNL), Los Alamos (LANL),
Oak Ridge (ORNL) and National Energy (NETL.) He teaches graduate level courses on Intelligent and
Collaborative Oilfield Systems Characterization & Management, and Advanced Oilfield Operations with
Remote Visualization and Control. He holds 3 US patents, with another 2 pending. He has authored 14
books and over 350 publications spanning wide areas. He holds a Ph.D. degree from USC.

Mazda Irani is the director of Ashaw Energy. He has over 15 years of experience to a wide range of thermal
activities. He is currently engaged in the design and optimization of different solvent/steam processes and
helping different operators to verify and optimize their operation. Irani was previously employed in technical
and supervisory roles with Suncor Energy, RPS Energy and C-FER Technologies. He has published and
presented more than 40 technical papers on different aspects of SAGD operation. Irani holds two PhD
degrees in petroleum and geomechanics and three MSc degrees.
30 SPE-193646-MS

Appendix A

Figure A1—Diagrammatic representation of heavy crude oil proposed by UNITAR/


UNDP Information Centre for Heavy Crude and Tar Sands (Cornelius, 1987)

Figure A2—Locations of heavy oil deposits in North America (Utah Heavy Oil Program, 2007)
SPE-193646-MS 31

Figure A3—Canadian oil production (CAPP 2017)

Figure A4—WTI and heavy crude price (Source: IEA 2005)


32 SPE-193646-MS

Figure A5—CO2 Flooding Process

Figure A6—VAPEX Process

Figure A7—Cyclic Steam Stimulation Process


SPE-193646-MS 33

Figure A8—Steamflooding Process

Figure A9—In-Situ Combustion Process

Figure A10—Steam-Assisted Gravity Drainage (SAGD)


34 SPE-193646-MS

Figure A11—Steam Chamber formed during SAGD

Figure A12—Comparison of temperature distribution of steam circulation and electrical heater


SPE-193646-MS 35

Figure A13—Reservoir Simulation Process

Figure A14—The Basic block of a Two-dimensional model.

Figure A15—Coupled Geomechanical Fracture Modeling System (Ji et al.,2009)


36 SPE-193646-MS

Appendix B
Table B1—Crude types and associated API values (Source – Speight, 2016)

Country Crude Oil API Sulphur (%)

Canada Bow River Heavy 26.7 2.1


China Daqing 32.6 0.09
India Bombay High 39.2 0.15
Iran Iran Heavy 33.8 1.35
Kuwait Kuwait Export 31.4 2.52
Nigeria Bonny Medium 25.2 0.23
North Sea Ekofisk 39.2 0.169
Saudi Arabia Arab Heavy 27.4 2.8
West Texas
USA 40.8 0.34
Intermediate
Venezuela Temblador 21 0.83

Table B2—Crude Classification on API gravity

Type API (°) Recovery Methods

Conventional Primary and


> 25
Crude Oil Secondary Recovery
Primary and Secondary
Tight Oil >25 recovery with
Hydraulic Fracturing
Medium Primary and
20-25
Crude Oil Secondary Recovery
Tertiary Recovery
Heavy Crude Oil 10-20
(Enhanced Oil Recovery)
Tertiary Recovery
Extra Heavy Oil <10
(Enhanced Oil Recovery)
Tar Sand
<10 Mining, SAGD, VapEX, etc.
Bitumen

Table B3—most important recovery methods for heavy oil reservoirs.

Thermal Recovery Methods Nonthermal Recovery Methods

• Cyclic Steam Injection (CSC) • Waterflooding

• Steam Injection • Polymer Injection

• In Situ Combustion • Vapor-Assisted Petroleum Extraction


(VAPEX)
• Steam Assisted Gravity Drainage (SAGD)
• Alkali-Surfactant Injection
• Electromagnetic Heating
• CO2 Injection
• Steam-Over-Solvent Injection
• Cold Heavy Oil Production with Sand
• Nanocatalyst (CHOPS)
• Foamy Heavy Oil Production

You might also like