You are on page 1of 14

SPE-171134-MS

Kinetics of Wet In-Situ Combustion: A Review of Kinetic Models


E.A. CAVANZO and S.F. MUOZ, Universidad Industrial de Santander; A. ORDOEZ and H. BOTTIA,
ECOPETROL S.A.

Copyright 2014, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Heavy and Extra Heavy Oil Conference - Latin America held in Medellin, Colombia, 24 26 September 2014.

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

Abstract
In Situ Combustion is an enhanced oil recovery method which consists on injecting air to the reservoir,
generating a series of oxidation reactions at different temperature ranges by chemical interaction between
oil and oxygen, the high temperature oxidation reactions are highly exothermic; the oxygen reacts with
a coke like material formed by thermal cracking, they are responsible of generating the heat necessary to
sustain and propagate the combustion front, sweeping the heavy oil and upgrading it due to the high
temperatures. Wet in situ combustion is variant of the process, in which water is injected simultaneously
or alternated with air, taking advantage of its high heat capacity, so the steam can transport heat more
efficiently forward the combustion front due to the latent heat of vaporization.
A representative model of the in situ combustion process is constituted by a static model, a dynamic
model and a kinetic model. The kinetic model represents the oxidative behavior and the compositional
changes of the crude oil; it is integrated by the most representative reactions of the process and the
corresponding kinetic parameters of each reaction. Frequently, the kinetic model for a dry combustion
process has Low Temperature Oxidation reactions (LTO), thermal cracking reactions and the combustion
reaction. For the case of wet combustion, additional aquathermolysis reactions take place.
This article presents a full review of the kinetic models of the wet in situ combustion process taking
into account aquathermolysis reactions. These are hydrogen addition reactions due to the chemical
interaction between crude oil and steam. The mechanism begins with desulphurization reactions and
subsequent decarboxylation reactions, which are responsible of carbon monoxide production, which reacts
with steam producing carbon dioxide and hydrogen; this is the water and gas shift reaction. Finally, during
hydrocracking and hydrodesulphurization reactions, hydrogen sulfide is generated and the crude oil is
upgraded.
An additional upgrading mechanism during the wet in situ combustion process can be explained by the
aquathermolysis theory, also hydrogen sulphide and hydrogen production can be estimated by a suitable
kinetic model that takes into account the most representative reactions involved during the combustion
process.
2 SPE-171134-MS

Figure 1Scheme of the zones presented in the reservoir during an in situ combustion process (Green, 1998).

Introduction
In situ combustion (ISC) is a thermal enhanced oil recovery method in which an oxygen-containing gas
in injected into the reservoir where it reacts with the crude oil to create a high temperature combustion
front that propagates through the reservoir. In most cases, the injected gas is air, although the use of 100%
oxygen has been reported [1]. Oil is displaced by the direct action of the resulting thermal front, as well
as the drive provided by the combustion gases, hot water and steam. The process is fuelled by a small
fraction of the oil (between 5-10% of the oil in place) [2]. Many of the ISC applications have failed and
led to the misconception that the method is highly risky and unpromising, but most of these failures
resulted from inappropriate application of the process [3]. Figure 1 shows the characteristic zones
presented in the reservoir during an ISC process.
Modeling ISC phenomena is an extremely complex job and requires a full understanding of the
behavior of different physical factors such as phase changes, chemical reactivity, heat and mass transfer
[4]. Historically, many authors [5,6,7] have defined three groups of chemical reactions involved in the
process: Low Temperature Oxidation or addition reactions (LTO), Medium Temperature Oxidation
(MTO) or fuel deposition reactions and High Temperature Oxidation (HTO). Modeling these reactions is
important for understanding the oxidative behavior and compositional changes occurring during the
process. The temperature range in which the reactions occur depends on the oil reactivity. Initially, the air
spreads into the porous medium near the well bore leading the LTO reactions which are exothermic and
occur due to the dissolution of the oxygen in the oil at reservoir temperature, they are characterized for
the production of partially oxygenated compounds such as aldehydes, ketones, carboxylic acid, alcohols
and hydroperoxides, also in these reactions the viscosity, density, boiling temperature and the volume of
asphaltenes in the oil increase and the volume of saturated and aromatic compounds decrease, for this
reason this reaction is not desirable during the process, however it permits initiate more easily the HTO
combustion reactions, because is the first exothermic reaction that occurs [8].
SPE-171134-MS 3

As the volume of air in the reservoir and its temperature increase considerably, the MTO or fuel
deposition reactions take place. These reactions occur in three different and consecutive mechanisms:
Dehydrogenation, cracking and condensation, on the first mechanism the hydrogen atoms are stripped
from the hydrocarbon molecules, carbon atoms are not altered, on cracking reactions the C-C bonds are
broken and finally a certain number of carbon atoms join to form a carbon rich compound or coke like
material, which is the fuel for the heterogeneous combustion reaction or HTO, on this reaction the coke
reacts with the oxygen to produce carbon oxides and water, the heat released here provides the energy
necessary to sustain and propagate the combustion front [9]. The injected air flux in a heavy oil project
should be maintained at a value well above the value needed to maintain the reactions in the HTO regime.
Among the factors influencing fuel deposition, the addition of clay and the increase of initial oil
saturation, viscosity, API gravity, as well as the surface area of porous media increase the amount of
deposited fuel (Poettmann et al, 1958). There are many controversies in the literature regarding the main
mechanism controlling fuel deposition. Bousaid and Ramey [10] considered the fuel to be a solid
resembling carbon formed by the pyrolysis of crude oil. Others researchers believed that fuel mainly
consists of heavy low-volatility hydrocarbon fractions left behind by distillation [11]. Globally, many
parameters affect the performance of an ISC process, such as the addition of clay or metallic salt additives
to the porous medium.

Kinetics of in situ combustion


Within the burning or combustion front, four known transport processes occur: 1) oxygen diffuses from
the bulk gas stream to the fuel interface; then perhaps, 2) the oxygen absorbs and reacts with the fuel; 3)
the combustion products desorb; and 4) products finally transfer into the bulk gas stream. If any of these
steps is inherently much slower than the remaining ones, the overall rate will be controlled by that step.
Considering the highly nonlinear equations that represent these phenomena, a huge amount of ISC studies
have focused on describing the kinetics of the combustion reaction [12, 13]. In general, the combustion
rate, Rc, of crude oil in a porous medium can be described as:
(1)

The reaction constant, k, is expressed by the velocity law, as:


(2)

Where PO2 is the partial pressure of oxygen, Cm is the instantaneous concentration of fuel, m and n are
the reaction orders with respect to PO2 and Cm respectively, A is the pre-exponential factor, E is the
activation energy, R is the universal gas constant and T is the temperature
In the above equation, A, E, m and n are unknown and are called the kinetic parameters. This means
that if a combustion model has three reactions (one for each regime), there are twelve variables that need
to be estimated, which results in an infinite set of solutions. Nonetheless, the orders of reaction are
normally assumed to be unity, hence the pre exponential factor and the activation energy are the
parameters which are normally adjusted [14]. Laboratory tests are necessary to quantify the reaction
kinetics of the process. Most of these analyses are based on EGA at either isothermal or non-isothermal
conditions. Most non-isothermal kinetic experiments are conducted by imposing a linear heating rate also
referred to as Ramped Temperature Oxidation (RTO). Additionally, DTA and TGA are used. The term
reaction regime is associated with the oxygen uptake velocity at which the oxidation reaction occurs
(LTO, MTO or HTO). For heavy oils, LTO reactions are dominant at temperatures below 300C and HTO
are the dominant reaction mechanism at temperatures above 350C, between these two ranges there is the
region known as the negative temperature gradient region where the rate of reaction actually decreases
4 SPE-171134-MS

Figure 2Reaction regimes for heavy and light oils during in situ combustion (D. Gutierrez et al, 2009).

with increasing temperature [15]. For light oils, HTO reactions are dominant at most temperature levels,
and LTO only seem to dominate at temperatures below 150C (Figure 2).
The kinetic model is composed by the reaction model and the kinetic parameters associated to each
reaction. The importance of this model lies in its ability to represent the oxidative behavior and the
chemical changes of a specific crude oil, so it can be implemented in a thermal numerical simulator to
model a combustion tube test, or even the performance of an in situ combustion process at field scale. The
development of a meaningful kinetic model that represents the transition of one oxidative reaction regime
to another and predicts the conditions at which the combustion process would be ineffective in mobilizing
oil is the main challenge of modeling the technique.
Normally, authors group the reaction model into pseudocomponents of similar properties, in traditional
compositional processes, the selection of pseudocomponents must allow proper modeling of the phase
behavior of the original hydrocarbons. This usually results in the oil being characterized in terms of
maltenes and asphaltenes or in terms of saturates aromatics, resins and asphaltenes (SARA) fractions.
Adegbasan [16] was the first who proposed a LTO reaction model by pseudocomponents (maltenes,
asphaltenes and coke) for the Athabasca Bitumen. The reaction mechanism of Adegbasan model is:

Sequera [17] proposed a model for LTO using SARA components, including the carbon oxides
formation reactions by oxygen addition and an intermediate product Resin 1 (Hydroperoxide). The
model presents an acceptable match and predicts adequately the operating conditions of the reactions.
SPE-171134-MS 5

Hayashitani [18] proposed a MTO reaction scheme by pseudocomponents for modeling the fuel
deposition and gas production during this regime:

Benham and Poettmann describe the stoichiometry of the coke combustion, which is:

Where is the CO2/CO % molar relationship and is the H/C atomic relationship. Thomas [19]
developed the first coke oxidation reaction during the bitumen combustion; it groups carbon oxides
production and found the CO2/CO relationship. The reaction heat of the coke combustion is approximately
4.278 105 kJ/kmol.

Belgrave [20] proposed a model for LTO, MTO and HTO reaction regimes, and obtained the kinetic
parameters by the Gauss Newton numerical method, obtaining a good match with the numerical
simulator. Belgrave reaction scheme model is:

Belgraves model described a pseudomechanistic reaction model for mathematical modeling of in situ
combustion of Athabasca bitumen. The model represented a consolidation of individual experimental
kinetic studies on thermal cracking, LTO and HTO.
For dry in situ combustion, in which only air is injected to the reservoir, it is comprehensible to take
into account in the kinetic model the LTO, MTO and HTO reactions. But for the case of wet in situ
combustion, in which water in injected alternately or simultaneously with air, chemical interaction
between steam and crude oil occurs [21].
Wet combustion
In the conventional dry in situ combustion process, a great amount of heat released during combustion is
stored in the burned sand behind the combustion front and is not used for the oil displacement. The heat
capacity of dry air is very small and air can not transfer heat from rock matrix as rapidly as is released.
Water, on the other hand, can absorb and transport heat much more efficiently than air. If water is injected
simultaneously or alternately with air, the process is called wet combustion. The vaporized water
condenses downstream the combustion front and helps with the oil displacement by a more pronounced
steam plateau and condensation front [22].
Dietz and Weidjema [23] studied the process in its early by combustion tube tests and established that
there is a better heat distribution downstream the combustion front, the coke volume formed during
thermal cracking is less compared with dry combustion, so the air requirement to burn this fuel is less, this
represent an economic advantage of the process in terms of air compression costs. At very low water
injection rates (incomplete wet combustion), water is vaporized before reaching the combustion front
(figure 3), at low and intermediate water injection rates (complete wet combustion) almost all heat from
the burned zone is carried by steam (figure 4), at high injection rates (super wet combustion) the
combustion front is partially quenched (figure 5). Smith and Perkins [24] were the first which observed
the effect of water on kinetics of wet combustion. They found that in wet combustion, the cooling effect
6 SPE-171134-MS

Figure 3Schematic of temperature profile for complete wet combustion (Sarathi, 1999).

Figure 4 Schematic of temperature profile for incomplete wet combustion (Sarathi, 1999).

Figure 5Schematic of temperature profile for super wet combustion (Sarathi, 1999).

generated by the injected water can promote a high LTO regime and low HTO. Moore et al. [25] observed
from wet and dry combustion tube tests that wet runs showed temperature peaks higher than dry runs, due
to the additional heat input to the combustion zone, also they observed that for wet runs in all cases was
registered an excess in CO2 production. The air requirement, fuel consumption, air-oil relationship and
air-fuel relationship decrease with a WAR (water air relationship) increase, while the front velocity and
the oxygen uptake increase.
Joseph and Pusch [26] showed the results of a wet combustion project in the Bellevue field and
observed that the volumetric sweep efficiency increase compared with the dry case, also the air required
SPE-171134-MS 7

to burn a volume of fuel was reduced by 63%. Bagci and Okandan [27] design dry and wet combustion
tube tests and observe an excess in CO2 production due to conversion of CO to CO2 in the presence of
steam at high temperatures or to the reaction between steam and coke in the presence of some rock matrix
minerals which can act as catalysts. The coke consumption decreases for the wet runs until reaching a
minimum, then it starts to increase as the WAR increase. Bagci [28] analyzed the effect of water on certain
performance parameters for crude oils of different API gravity at low and high WAR values; the air
requirement and fuel consumption decrease with a WAR increase, but it tends to reach a critical value as
the WAR increase continuously, also the oil recovery factor reach an optimal value for an specific WAR
value.
The main advantage of wet combustion is to reduce air requirement and improve the sweep efficiency
by expanding the steam plateau region. Since most of the oil is displaced ahead of the burning front, the
steam plateau becomes the primary driving mechanism for oil production. Wet combustion at high WAR
will eventually lead to quenching of the front. The high WAR caused the water bank velocity behind the
front to accelerate and eventually quench the burning zone, thus reducing the front to steam region.
Kinetic studies in wet combustion
Several studies have been published about the thermal effect of water on the combustion process,
concluding that it decrease the fuel deposited on the rock matrix, increase the combustion front velocity
and the volume of recovered oil, due to an improvement in the oil mobility through the heat transferred
by steam, Sarathi states that the mechanism causing the fuel deposit to be decreased during wet
combustion is believed to be the increased availability of hydrogen in the combustion zone. Few works
have been published about the effects of water on heavy oil combustion reactions and kinetics of the
process. Wet ISC numerical simulations are always performed without taking into account steam in
oxidation reactions, and results are not in good agreement with experimental data comparing to dry ISC
simulations. Reactive mechanisms, including water effects improve matching. Lee and Noureldin [29]
studied the effect of steam on the LTO reaction of heavy oil by looking at the evolution of the composition
of the sample. When LTO is conducted in presence of water while all other conditions are held constant,
there is a sharp decrease in the amount of tetrahydrofuran solubles and coke formed. Furthermore,
liberation of carbon dioxide, with a consequent decrease in both the viscosity and acidity of the produced
oil, is much greater in the presence of water.
Urban and Udell [30] analyzed the influence of steam on the combustion of oil in a wet combustion
test and found that the LTO exothermicity is similar to the dry case, but the oxygen consumption is
dramatically reduced and, unlike the dry case, all consumed oxygen appears in the effluent carbon oxides.
Smith and Perkins observed a substantial increase in oxidation rate when water was injected into the
reaction tube at pressures above the boiling level. Greaves el al. [31] in their combustion tube experiments,
found oxygen diffusion to be an important controlling parameter for wet combustion. The reaction order
with respect to oxygen partial pressure was found to be unity for dry combustion and less than 0.5 for wet
combustion.
Bagci [32] conducted a study to investigate the effect of water on combustion reaction kinetics in
limestone packs containing heavy oils where simultaneous injection of air and water was realized for
different crude oils for Turkish oil fields. The results indicated that for wet combustion the amount of CO
decrease while the amount of CO2 increased, this indicates that oxygen was efficiently used during the
combustion reaction when water was present.
At higher temperatures where fuel combustion reaction occurs, activation energies were higher when
water was present, for low temperature reactions activation energies were lower. In general, activation
energies for only air injection were greater than simultaneous air-water injection runs.
Lapene et al. [33] conducted a complete study of the kinetics of dry and wet combustion by RTO tests
for a specific crude oil (figure 6). The authors conclude that steam acts on LTO reaction, drastically
8 SPE-171134-MS

Figure 6 Effect of water on reaction regimes (Lapene et al, 2009).

reducing oxygen consumption; this reduction has an indirect effect on HTO reaction, increasing the
temperature range during which oxygen is consumed. Steam also acts directly on HTO reactions
increasing oxygen consumption period and producing more carbon dioxide, and reducing carbon mon-
oxide production caused by coke combustion. Oil upgrading can be explained by LTO weakening caused
by steam.
The evidences suggest that water should be included in the reactive mechanism during modeling
because it can affect strongly combustion and indirectly other mechanisms, such as transport. Laboratory
studies and field applications of thermal recovery processes have demonstrated that appreciable amounts
of gaseous environmental contaminants such as H2S and CO2 are created by the aquathermolysis
(steam/oil reactions) of heavy oils. It has also been demonstrated that H2 and light saturated hydrocarbons
are also produced by steam/oil reactions over the temperature range 200 to 300C.

Aquathemolysis theory
Hyne [34] presented the aquathermolysis of oil sands and/or heavy oils, which means the breaking down
of the components of oil through chemical reactions brought about by contact with hot water either in the
vapour or liquid phase or a combination of both. The aquathermolysis window concept means that
aquathermolysis appears to be the predominant form of chemical reaction between the thermal cracking
that is occurring above 350C and the very slow chemical reactions that characterize slow thermal
maturation of the reservoir and its contents below 200C.
The heart of the overview is contained in the reaction schematics presented in figure 7, in this figure
the authors have attempted to identify the principal chemical reactions that together make up the overall
aquathermolysis process. The reactive species produced upon aquathermolysis, probably of organosulphur
compounds, can either polymerize (step 2 of figure 7) or react with water (Step 1) to yield smaller
fragments that participate in a subsequent series of reactions. The polymerization leads to undesirable
viscosity increases.
SPE-171134-MS 9

Figure 7Reaction mechanism of aquathermolysis of heavy oil (Hyne, 1986).

Acid polymerization
The effect of aquathermolysis on viscosity is of critical importance. Hyne studied this effect for Athabasca
oil sand. Aquathermolysis at 200C results in steady increase in viscosity. At 300C, however, the
opposite effect was observed and substantial decrease in viscosity over a 30 day aquathermolysis period
resulted. This behavior is explained by the interplay between polymerization reactions and molecular
cleavage. The polymerization reactions, catalyzed by acid conditions caused by initial fast release of CO2,
are dominant at lower temperatures probably because of the higher activation energies of the chemical
reactions that form the subsequent steps of the overall aquathermolysis process. At higher temperatures
or longer aquathermolysis times at intermediate temperatures, the molecular weight cleavage reactions
become dominant and viscosities decrease (Hyne, 1986).
The viscosity changes may also be more the result of changes in molecular shape or conformation than
of molecular weight. The opening and reclosing of sulphur linkages in high molecular weight species may
10 SPE-171134-MS

well result in considerable reorganization of the special arrangement of asphaltenic and other large
molecular structures.

Reactions with water


This is the initial reaction between the high temperature water and the organosulphur compounds which
are the most sensitive components of the heavy oils as far as aquathermolysis is concerned. Aquather-
molysis takes the form of hydrolysis of a carbon-shulphur bond and may or may not involve release of
H2S. There are two important features of this step. The first is that the bond scission counters the negative
effect of the polymerization step either by reducing molecular weight or by changing the shape of the
molecule which can also markedly affect viscosity. The second important feature of this step is that it
produces intermediates (aldehydes) which can decarboxylates to produce carbon monoxide. The transient
appearance of CO as a component of the gas phase immediately suggests that it is being reconsumed in
the overall process. This is known as the water gas shift reaction (WGSR). The equilibrium WGSR is
(Hyne, 1986):

This reaction relates to aquathermolysis if the mineral component of the oil sand as opposed to the oil
component. It can be a dominant source of the CO2 in certain cases when reactive mineral carbonates, as
siderite, are present. Copious amounts of CO2 can be produced which provide additional in situ pressure
drive but also tend to acidify the aqueous system thus encouraging acid polymerization. Large amounts
of CO2 from mineral carbonates can also limit the WGSR by displacing the equilibrium back to the CO
and H2O side thus reducing hydrogen production. Since components of heavy oils and their associated
minerals can catalyze the WGSR, it would be expected that CO2 should predominate over CO in the
produced gases, in some cases only CO2 is observed (Hyne, 1986).
The in situ production of CO2 is known to have beneficial effects on viscosity reduction of the heavy
oils and the availability of H2 suggests the possibility of hydrogenation upgrading in situ.

Hydrodesulphurization
The generation of H2S during thermal enhanced oil recovery processes in heavy oils appears to be the
consequence of chemical reactions of the organosulphur components of the heavy oil. While metal
sulphides in the sand phase may, on occasion, be a source of H2S such components of the porous medium
are likely to be more important as catalysts in H2S generation.
The hydrogen produced by WGSR is of little value if it cannot be made to react further with the heavy
oil and thus upgrade the fluid. Here, the role of added metal ion catalysts can be critical. This is classical
catalyzed hydrocracking at low temperatures, in aqueous medium, in situ. It reduces viscosity, adds
hydrogen to the oil, and removes sulphur (as H2S).
Temperature and time can be used to control the amount of H2S produced during a thermal recovery
process, although the use of time is likely to be severely circumscribed by other operating factors. If
satisfactory production can be achieved at lower temperatures, however, it might be economically
preferable to do so rather than seek higher production at higher temperature at the cost of having to handle
more produced H2S if the reservoir fluid was one which more readily produced H2S. Control can also be
affected to some extent by use of added metal salts which form sulphides and trap H2S (Hyne, 1986).
1(a) Metal complexation with thioether sulphur in complex between Runthenium salts and thiolane.
Evidences suggests intracomplex interactions.
1(b) Intermediate thiol production detected by potentiometric titration of SH.
1(c) Enol rearrangement can yield aldehyde. Aldehydes detected by IR and formation of 2, 4 DNPH
complex.
SPE-171134-MS 11

1(d) Decarboxylation to yield CO shown possible by aquathermolyzing n-butyaldehyde under same


conditions to yield CO.
Gases in boxes are principal components of gas phase produced from aquathermolysis of both heavy
oil sands and thiolane and thiophene. Discontinuous lines represent metal ions from mineral phase or
intentional addition act as catalysts for both WGSR and hydrodesulphurization possibly in sulphide
form through reaction with produced H2S.
(3) Intermediate nature of CO explains transient appearance in produced gas.
Kapadia et al. [35] proposed a kinetic model for hydrogen generation from Athabasca Bitumen during
in situ combustion, which includes thermal cracking, low and high temperature oxidation, gasification,
water gas shift reaction, methanation, and hydrogen and carbon monoxide combustion
Thermal Cracking Reaction (Hayashitani, 1978):

Low Temperature Oxidation Reaction (Adegbasan 1982):

High Temperature Coke Oxidation Reaction (Adegbasan 1982):

High Temperature Gas Oxidation Reaction (Yang and Gates 2009):

Hydrogen generation reactions:

Hydrogen consumption reactions:

Carbon Monoxide Combustion Reaction (Babushok and Dakdancha 1993):

Reaction were simulated using a thermal numerical simulator (CMG 2008) analyzing the effect of
pressure and temperature on the volume of hydrogen generated per volume of Athabasca bitumen reacted,
revealing that hydrogen generation depends more strongly on temperature than pressure, kinetic param-
eters were determined by matching in the simulator with the experimental results. In general, the reactions
proceed as follows. First, in the zone upstream of the oxygenated combustion zone which is sufficiently
hot due to heat conduction, since there is an absence of oxygen, bitumen undergoes thermal cracking. This
leads to the production of asphaltene, coke and gas. After coke has formed, it can react with water to
generate carbon monoxide and hydrogen. This generated hydrogen can then react further with coke to
generate methane. Also, hydrogen can react according to water-gas shift reaction. After oxygen reaches
this zone, since temperature is still relatively low, maltenes and asphaltenes are converted to coke under
12 SPE-171134-MS

the low temperature oxidation reactions and methane and hydrogen, if present, will combust with oxygen.
Next, the coke starts to combust under high temperature providing substantial amounts of energy for
downstream heat conduction. If the coke consumption is large, hydrogen generation by gasification of
coke can suffer. If hydrogen gas is in the presence of oxygen, it will be consumed (Kapadia et al, 2009).

Conclusions
Durint the wet in situ combustion process, water can absorb and transport heat much more efficiently than
air because of its higher heat capacity. The heat stored in the burned sand can be recovered and transported
downstream the combustion front, enhancing the heat distribution and upgrading the oil.
Coke volume formed during wet in situ combustion is low compared with dry combustion, due to an
improvement in the oil mobility for the heat transferred by steam forward the combustion front, this
reduces the air requirement, increases the oil recovery and the air compression costs.
Kinetic studies on wet in situ combustion suggests that steam acts on LTO reaction, drastically
reducing oxygen consumption; this reduction has an indirect effect on HTO reaction, increasing the
temperature range during which oxygen is consumed. Steam also acts directly on HTO reactions
increasing oxygen consumption period and producing more carbon dioxide, and reducing carbon mon-
oxide production caused by coke combustion. Oil upgrading can be explained by LTO weakening caused
by steam.
The evidences suggest that water should be included in the reactive mechanism during modeling a wet
in situ combustion process because it can affect strongly combustion and indirectly other mechanisms,
such as transport. Laboratory studies and field applications of thermal recovery processes have demon-
strated that appreciable amounts of gaseous environmental contaminants such as H2S and CO2 are created
by the aquathermolysis (steam/oil reactions) of heavy oils.
Wet in situ combustion numerical simulations are always performed without taking into account steam
in oxidation reactions, and results are not in good agreement with experimental data comparing to dry in
situ combustion simulations. Reactive mechanisms, including water effects improve matching (Kapadia,
2009).
Chemical interaction between steam and heavy oil can be explained by aquathermolysis reactions; the
mechanism begins with the breakup of some reactive species of the heavy oil, especially large chains
containing sulphur compounds, forming reactive fragments which conduct to decarboxylation, the CO
produced react with steam (Water gas shift reaction WGSR) generating CO2 and H2, Subsequently, the
H2 react and hydrogenate certain compounds of the oil obtaining an enhanced crude oil and non-
hydrocarbon gases such as CO2, CO, H2 and H2S.

Acknowledgements
We gratefully thank ECOPETROL S.A. for suggesting the research in wet combustion and financial
support. Ph.D. Louis Castanier from Stanford University and Ph.D. Claude Gadelle from Xytel Corp. are
acknowledged gratefully for their technical support.

Nomenclature
C
m Instantaneous concentration of fuel
K Reaction rate, unit depends on reaction model
PO2 Partial pressure of oxygen
m,n Reaction orders
Relacin en %molar de CO2/CO.
x Relacin atmica H/C
E Activation energy, J/gmol
SPE-171134-MS 13

A Pre-exponential factor
T Temperature, K
R Universal gas constant, Pa m3/gmol K
COX Carbon oxydes

References
1. D.W. GREEN, G.P. WILLHITE. Enhanced Oil Recovery. Henry L. Dohery Memorial Fund of
AIME, Society of Petroleum Engineers. Richardson, Texas. 1998.
2. D. GUTIERREZ, R.G. MOORE, S.A. MEHTA, M.G. URSENBACH, F. SKOREYKO. The
challenge of predicting field performance of air injection projects based on laboratory and
numerical modeling. Journal of Canadian Petroleum Technology JCPT, Volume 48, No. 4. 2009.
3. P. SARATHI. In Situ Combustion Handbook Principles and Practices, Report: DOE/PC/91008-
0374, OSTI_ID:3174, National Petroleum Technology Office, U.S. D.O.E., Tulsa, Oklahoma,
1999.
4. M. CINAR, L.M. CASTANIER, A.R. KOVSCEK. Improved Analysis of the kinetics of crude oil
in situ combustion. Society of Petroleum Engineers, Western Regional and Pacific Section AAPG,
Bakersfield, California. 2008.
5. M.R. FASSIHI. Analysis of fuel oxidation in in situ combustion oil recovery. Ph. D. Thesis,
Stanford University. 1981.
6. BURGER, J.G, SAHUQUET, B.C. Chemical Aspects of In-situ Combustion-Heat of Combustion
and Kinetics, Society of Petroleum Engineers. 41O-422; Trans., Ale, Vol. 253. 1972.
7. BOUSAID, S, RAMEY, H.J. JR. Oxidation of Crude Oils in Porous media. Society of Petroleum
Engineers. pp 137148. 1968.
8. M.K. DABBOUS, P.F. FULTON. Low temperature oxidation reaction kinetics and effects on the
in situ combustion process. Society of Petroleum Engineers, SPE-AIME 47th Annual fall
meeting, San Antonio, Texas. 1972.
9. J.D. ALEXANDER, W.L. MARTIN, J.N. DEW. Factors Affecting Fuel Availability and Com-
position during In-Situ Combustion. J. Pet. Tech. 1154 1164; Trans., AIME, Vol 225. 1962.
10. BOUSAID. I.S, RAMEY. H.J. Oxidation of crude oil in porous media. SPEJ 8 (2): 137148;
Trans, AIME, 243. Spe-1937-PA. DOI: 10.2118/1937-PA. 1968.
11. BAE. J.H. Characterization of crude oil for fireflooding using thermal analysis methods. SPEJ 17
(3): 211218. SPE-6173-PA- DOI: 10.2118/6173-PA.
12. M.R. FASSIHI, W.E. BRINGHAM, H.J. RAMEY. Reactions kinetics of in situ combustion: Part
1 observations. Society of Petroleum Engineers of AIME. 1984.
13. M.R. FASSIHI, W.E. BRINGHAM, H.J. RAMEY. Reactions kinetics of in situ combustion: Part
2 modeling. Society of Petroleum Engineers of AIME. 1984.
14. R.G. MOORE, M.G. URSENBACH, C.J. LAURESHEN, J.D.M. BELGRAVE, S.A. MEHTA.
Ramped Temperature Oxidation Analysis of Athabasca Oil Sands Bitumen. University of Calgary.
Paper JPTC.
15. M.R. FASSIHI. Analysis of fuel oxidation in in situ combustion oil recovery. Ph. D. Thesis,
Stanford University. 1981.
16. ADEGBASAN, DONELLY, R.G MOORE. Liquid phase oxidation kinetics of oil sands bitumen:
Models for in situ combustion numerical simulators. SPE-12004-PA. 1987.
17. SEQUERA, R.G. MOORE, S.A. MEHTA, M.G. URSENBACH. Numerical simulation of in situ
combustion experiments operated under low temperature conditions. University of Calgary. 2007.
J. Cdn Petroleum Tech., Vol. 49, No 1, p 5564. 2010.
18. HAYASHITANI, M. Thermal Cracking of Athabasca Bitumen. Ph.D. Thesis. The University of
Calgary, Alberta. 1978.
14 SPE-171134-MS

19. F.B. THOMAS, R.G. MOORE, D.W. BENNION. Kinetic parameters for the high temperature
oxidation of in situ combustion coke. PETSOC-85-06-05. Vol 24. 1985.
20. J.D.M. BELGRAVE, R.G. MOORE, M.G. URSENBACH, D.W. BENNION. A comprehensive
approach to in situ combustion modeling. University of Calgary. SPE 20250. 1994.
21. A. LAPENE, L.M. CASTANIER, G. DEBENEST, M. QUINTARD, A.M. KAMP, B. CORRE.
Effects of water on kinetics of wet in situ combustin. SPE western regional meeting, San Jose,
California. SPE-121180-MS. 2009.
22. A.M GARON, R.J WYGAL, JR. A Laboratory Investigation of Fire-Water Flooding. SPE-4762-
PA, Vol 14, pp 537544. 1974.
23. D.N DIETZ, J. WEIJDEMA. Wet and Partially Quenched Combustion. SPE 1899-PA, Vol 20, pp
411415. 1968.
24. F.W. SMITH, T.K. PERKINS. Experimental and numerical simulation studies of wet combustion
recovery process. PETSOC-73-03-05. Journal of Canadian Petroleum Technology, Vol 12. 1973.
25. R.G. MOORE, J.D.M. BELGRAVE, R. MEHTA, M.G. URSENBACH, K. XI. Some insights into
the low temperature and high temperature in situ combustion experiments. Pp 179 190 in
SPE/DOE Eight Symposium on Enhanced Oil Recovery, Tulsa, Oklahoma. 1992.
26. C. JOSEPH, W.H. PUSCH. A field comparison of wet and dry combustion. SPE-7992-PA,
Journal of Petroleum Technology, Vol 32, pp 15231538. 1980.
27. S. BAGCI, E. OKANDAN. Dry and wet combustion studies of different API gravity crude oils
from Turkish oil fields. PETSOC-88-39-58, Annual Technical Meeting, Calgary, Alberta. 1988.
28. S. BAGCI, M.V. KOK. In situ combustion laboratory studies of Turkish heavy oil reservoirs. Fuel
processing technology 74 pp 6579, ELSERVIER. Department of petroleum and natural gas
engineering, Middle East Technical University, 06531, Ankara, Turkey. 2001.
29. D.G. LEE, N.A. NOURELDIN. Effect of water on the low temperature oxidation of heavy oil.
Energy and Fuels 3(6): 713715. DOI: 10.1021/ef00018a009. 1989.
30. D.L. URBAN, K.S. UDELL. The effects of steam on the combustion of oil on sand. SPERE 5 (2):
170 176; Trans, AIME, 289. SPE-18073-PA. DOI: 10.2118/18073-PA. 1990.
31. M. GREAVES, R.W. FIELD, V.A. ADEWUSI. In situ combustion for oil recovery. Chem. Eng.
Res. Design 65:2328. 1988.
32. S. BAGCI. Reaction kinetics of wet combustion of crude oils. Energy sources, Part A: Recovery,
Utilization, and Enviromental Effects 28(3): 233244. DOI: 10.1080/009083190889997. 2006.
33. A. LAPENE, L.M. CASTANIER, G. DEBENEST, M. QUINTARD, A.M. KAMP, B. CORRE.
Effects of Steam on Heavy Oil Combustion. SPE-118800-PA, Vol 12, pp 508 517. 2009.
34. J.B. HYNE. Aquathermolysis of heavy oils. AOSTRA 11, 103, 103B/C. ID-OL19361916M, The
University of Tulsa. 1986.
35. P.R. KAPADIA, M.S. KALLOS, C. LESKIW, L.D. GATES. Potential for hydrogen generation
during in situ combustion of Bitumen. SPE-122028-MS. EUROPEC/EAGE Conference and
Exhibition, Amsterdam, The Netherlands. 2009.

You might also like