Вы находитесь на странице: 1из 10

A Comprehensive Approach to In-Situ Combustion Modeling

J.D.M. Belgrave, R.G. Moore, M.G. Ursenbach, D.W. Bennion


The University of Calgary, Department of Chemical and Petroleum Engineering
ABSTRACT
Low temperature oxidation (LTO) has long been recognized as one of
the dominant mechanisms controlling fuel availability in in-situ
combustion. Its effect on the physical properties of crude oils is also
well known. In spite of these fmdings, the prevailing conceptual model
of in-situ combustion still hinges on thermal cracking (in isolation)
ahead of the firefront, to provide sufficient fuel (coke) for propagation
of the reaction zone. Previous simulation studies, which purported to
include LTO as part of the reaction scheme, have unrealistically
specified the reaction products as carbon oxides and water.
Furthermore, oil compositional changes due to oxidation have been
completely neglected.
This paper describes a unified pseudo-mechanistic reaction
model for mathematical modeling of in-situ combustion of Athabasca
bitumen. The model represents a consolidation of individual
experimental kinetic studies on thermal cracking and low temperature
oxidation of Athabasca bitumen, and reported data for the high
temperature oxidation of coke. The formulation is comprehensive in that
it allows bitumen to undergo density and viscosity increases, as well as
reduced reactivity to oxidation, with increased oxidation extent.
Hydrocarbon bypassing due to quenching of the combustion front is also
permitted with the proposed kinetic model.
The paper includes application of the reaction model in
numerical simulations of adiabatic combustion tube tests performed on
Athabasca bitumen. A significant feature of the model is its ability to
predict the dual oxidation uptake peaks associated with ramped
temperature oxidation experiments.

INTRODUCTION
The oil sands of Alberta, Canada collectively represent one of the
largest hydrocarbon deposits in the world l . Cyclic steam stimulation, to
date, has been the most widely used technique for exploiting these
deposits. This technique is capable of recovering only 15 to 20 % of the
oil-in-place, and a follow-up process is required to improve recovery2.
Laboratory investigations of in situ combustion as a'
post-steaming process have been sufficiently encouraging to warrant
implementation of pilot studies by some lease operatorg3. However, the
transition from experimental and pilot stages to commercial
development has been virtually non-existent. This stems, in part, from
the fact that in situ combustion is not well understood, mechanistically.
A great deal of laboratory work has shown that frontal
advance and air requirement arc controlled by the kinetics of the
reactions in the vicinity of the burning front. Several studies4,5,6 have
reported on three major reactions which occur during fireflooding: (1)
thermal cracking, (2) liquid phase low temperature oxidation (L TO),
and (3) high temperature oxidation (HTO) of a solid hydrocarbon
residue. In their pioneering experimental effort, Alexander et aU
concluded that of all the process variables which they studied, LTO
prior to high temperature burning had the greatest effect on fuel
availability. Poettmann et al. 8 estimated that L TO could increase the
fuel deposition by as much as 100 %, and Lerner et al. 9 emphasized the
need to consider the effects of LTO in numerical simulations of
combustion processes.
In spite of these findings, published conceptual profiles of in
situ combustion still adhere to thermal cracking as the sole means of
98

fuel generation ahead of the reaction zone. In addition, most simulation


studies1o,lI, 12 which considered LTO, have neglected associated chemical
changes such as increased viscosity and density of the oil as well as its
reduced reactivity to oxidation with increased oxidation extent.
The main focus of this work was the development of a pseudo
component reaction model that is able to produce the oxidation related
phenomenon mentioned above, in combustion simulation studies of
Athabasca oil sands. Individual thermal cracking and LTO kinetic
studies on Athabasca bitumen, and reported data for coke combustion
have been consolidated into such a model. In this paper, the
performance of the proposed reaction scheme is examined in numerical
simulations of a differential reactor undergoing an imposed ramped
temperature history, and two combustion tube tests (dry and superwet)
performed on Athabasca bitumen. Through this analysis an improved
quantitative description, and therefore understanding, of in situ
combustion has emerged.

. EXPERIMENTAL BASIS FOR REACTION MODEL


For an explicitly correct kinetic representation of hydrocarbon cracking
and oxidation, an inordinately large number of chemical species would
have to be considered. Such a system would not be practical as it would
impose prohibitive computational demands on thermal reservoir
simulation. A pseudo component model offers the only useful
alternative. Furthermore, unification of the three classes of reactions
into a comprehensive model can only be achieved if kinetic studies on
the reactions are consistent in their fractionation or characterization of
the oil.
As regards LTO and thermal cracking of Athabasca bitumen,
Adegbesad 3 and Hayashitani et al. 4 have reported pseudo component
kinetic data for these reactions. Both studies used bitumen which was
free of water and minerals.
Hayashitani thermally cracked Athabasca bitumen in a closed
system at 651 to 828 OF [344 to 442 0c] under an inert atmosphere.
The cracked liquid products were first separated into maltenes and an
asphaltenes-coke residue by filtration, using n-pentane as solvent.
Asphaltenes were next recovered from the residue by solution in
benzene. Hayashitani further fractionated the maltenes into light oils,
middle oils, and heavy oils by vacuum distillation.
Adegbesan used a stirred semiflow batch reactor to investigate
LTO of Athabasca bitumen in the 140 to 302 OF [60 to 150 0c]
temperature range, and at oxygen partial pressures of 7 to 324 psig [50
to 2233 kPa]. His characterization technique for the reaction products
duplicated Hayashitani's as far as separation of the maItenes and
asphaltenes-coke in n-pentane. Coke was defined as the bitumen fraction
insoluble in toluene. The maltenes were then separated in saturates,
aromatics, oils and resins by. a combination of solvent extraction and
chromatographic techniques.
In view of the difference in the methods used to analyze the
maltenes in the two studies (thermal as opposed to
solvent/chromatographic), consolidation of the kinetics data was
restricted to coke, asphaltenes, maltenes, and gas, as pseudo
components.

SPE Advanced Technology Series, Vol. 1, No.1

100

25

397 C

397 C
coke

maltenes
asphaltenes
80

60

If

40

:.
if
:IE

~
:IE

20

15

10

gas

20

10

o~~~----~--~----~--~--~

12

Time (hours)

10

12

Time (hours)

Fig. la - Cracking model vs. Hayashitani's data for maltenes and


asphaltenes.

Fig. Ib - Cracking model vs. Hayashitani's data for gas and coke.
these parameters are:

Reaction Kinetics
Based on Hayashitani's data, the thermal cracking scheme
proposed in this paper assumes three fIrst order reactions:

MALTENES -

ASPHALTENES

ASPHALTENES ASPHALTENES -

(1)

COKE

(2)

GAS

(3)

If kl> k2' and k3 are the rate constants respectively for the
above reactions, then differential equations describing the material
conversion can be written as

dCasp

(4)

dt

Al
EI

9.092 x 1012
2.347 X 105

sec- I
kJ/kmol

A2
E2

4.064 x 109
1.772 x lOS

sec- I
kJ/kmol

A3
E3

1.362 x 109
1.763 X 105

sec- I
kJ/kmol

Figure 1 compares this cracking model versus Hayashitani's


data at 747 OF [397C]. The percent of each component is on a mass
basis. At this and higher temperatures the agreement is quite good.
However, the initial increase in maltenes concentration above the
predicted curve becomes more prominent at lower temperatures,
indicating the need for another pseudo component.
For the LTO data reported by Adegbesan, the following
.reactions are proposed:

MALTENES
(5)

dCgas

(6)

dt
with the temperature dependence of the rate constants being described
by the Arrhenius equation
~exp( -EJRT)

(7)

Using the technique described by Kalogerakis and Luus l4 ,


kinetic parameters were estimated for the thermal cracking scheme
specifIed above. In the order in which the reactions have been specifIed,
SPE Advanced Technology Series, Vol. I, No.1

OXYGEN -

ASPHALTENES

ASPHALTENES

OXYGEN -

COKE

(8)

(9)

The kinetic parameters for this system were estimated from rate
equations similar to that described for the thermal cracking reactions.
However, oxygen concentration was specifIed in terms of its partial
pressure and the reaction order with respect to oxygen partial pressure
was allowed to vary in an unconstrained fashion. With respect to
hydrocarbon mass fraction the reactions were kept as fIrst order.
Respectively, the Arrhenius constants are:
6.819
8.673

X
X

103
104

2.133 X 10- 10
1.856 X 105

sec-I Pa- 0.4246


kJ/kmol
sec-I Pa-4.7627
kJ/kmol

99

100

eo

...c:

8...
f

:E

100

135 C
maltenes
asphaltenes
... coke

150 C
maltenes
asphaltenes
... coke

80

60

60

40

40

20

20

o l..-::::::::t::::~===--1-~

oL-~~~~==~==~~
2
o
1
3
4
6
5

0.0

0.5

1.0

for the conversion of the hydrocarbon reactants. Note that the frequency
factors inherently state reaction orders, with respect to oxygen partial
pressure, of 0.4246 for maltenes oxidation and 4.7627 for asphaltenes
forming coke. Figures 2 and 3 show the experimental data at two
temperature levels, along with the predicted bitumen compositional
changes. Generally it was found that at temperatures of 275 OF [135
C], or less, the predicted mass percentages 0 f the components agreed
very well with the experimental data. However, at higher temperatures
(Figure 3) it was not possible to predict the reduction in asphaltenes
content and increased coke synthesis which occurred at later reaction
times.
It is important to note here that we have not specified the
release of oxygen in the cracked product streams of reactions 1-3. This
stems from our experimental findings (unreported) which have shown
that thermal cracking of preoxidized bitumen does not regenerate the
molecular oxygen which was consumed in the additive LTO reactions.
Kinetic parameters for coke combustion were obtained from
the work of Thomas et al. 15. Using integral analysis, these researchers
studied the oxygen uptake of coke combustion with coke derived from
an oxidized Athabasca bitumen-water-sand mixture. From this
reference, coke combustion is first order with respect to both reactants,
and the reaction rate was given as:
X

10-6
(10)

x exp( -34763/RT) Cook..

P02

with Ceoa in kg coke/m 3 bulk volume.


The in-situ combustion kinetic model offered above is
considered to be preliminary in nature, and there is much latitude for
refinement and optimization. As more consistent experimental data
becomes available, more intermediate reaction pathways and pseudo
components may be specified for thermal cracking as well as LTO.
The thermal cracking scheme we have proposed is based
solely on the experimental observation that the asphaltenes
concentration monotonically decreases (Figure 1) which facilitates the
determination of rate equations for predicting the monotonic increase in
coke and gas formation. The change in maltenes concentration on the
100

2.5

Fig. 3 - LTO model vs. Adegbesan's data at 150C.

Fig. 2 - LTO model vs. Adegbesan's data at 135C.

3.612

2.0

Time (hours)

Time (hours)

gmolOzfhr

1.5

other hand is not monotonic; it first increases and then decreases.


Maltenes involvement was therefore limited to that of a buffer supplier
of asphaltenes.
With respect to high temperature oxidation, it is expected that
inclusion of the complete oxidation of maltenes and asphaltenes will
improve the representation of the combustion process. However, the
experimental data needed to furnish the stoichiometric coefficients and
kinetic parameters for these reactions remain unavailable.
Stoichiometry
Estimates for the molecular weights of the pseudo components
must be obtained if stoichiometric coefficients for the preceding
reactions are to be specified.
Bishnoi et al. I6 presented characterization data for oil sands
bitumen, which included specific gravities and molecular weights of the
reported pseudo components. These data were compared with measured
specific gravities of the original bitumen and maltenes fraction, and
molecular weights were inferred for the oil components (see Table 1).
The gaseous products from Hayashitani's cracking experiments
indicated an average molecular weight of 29.0, and for coke,
Adegbesan reported a measured hydrogen/carbon ratio of 1.13, which
gives a coke molecular weight of 13.13.
In thermal cracking, a unit mass of reactant produces a unit
mass of product. The stoichiometric coefficients for these reactions are
therefore determined from the ratios of the molecular weights of
reactants and products. Thus the thermal cracking reactions, on a molar
basis, can be written as:

MALTENES -

0.372 ASPHALTENES

ASPHALTENES -

ASPHALTENES

(11)

83.223 COKE

(12)

37.683 GAS

(13)

The amounts of oxygen which react with unit mass of


maltenes and asphaltenes were determined from Adegbesan' s LTO data
by parameter estimation l4 with the following oxygen uptake rate
SPE Advanced Technology Series, Vol. 1, No.1

simulation runs are discussed in this paper. For these runs, a constant
value 8.96 was specified for the CO2 /CO molar ratio, as the
stoichiometry is not appreciably affected by small changes to this
parameter. On a molar basis, and lumping the carbon oxides into CO..
the coke burning reaction becomes:

3.0
0.22 g 0/g Asphaltenes

0.27 g 0/g MaJtenee

CH1.13

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

Measured Oxygen Uptake Rate (g/hr)


Fig. 4 - Measured vs. predicted oxygen uptake rates.

equation:

dr00

dt

dtn"wt
dt

dmc:oke
+ -b
---(1 +b)

(14)

dt

(15)

0.4726 ASPHALTENES

ASPHALTENES + 7.513 OXYGEN -

(16)

101.539 COKE
The general form of the coke combustion reaction is well
documented in the literature l7 :

n]

2m+l
CH + - - + - 0 [ 2m+2
n
4 2
+

-CO
m+l
2

1
--CO
m+l

+ 0.565~0

(18)

Bitumen Chemical Changes


It is evident that the reaction scheme specified above allows
oil components to be synthesized and/or consumed as a result of
oxidation. This was taken one step further in that an attempt was made
. to determine the viscosity and density of the oil components, based on
laboratory measurements.
The viscosity of the original bitumen and its maltenes fraction
were independently measured and the data curve fitted to give the
following viscosity-temperature relationships:
'-bitumen

MALTENES + 3.431 OXYGEN

COx

Heats of Reaction
There is little evidence in the literature which suggests that
thermal cracking of hydrocarbons is accompanied by any release or
absorption of heat which significantly affects the combustion process .
Therefore the enthalpies of these reactions were assumed to be zero. To
obtain an estimate of the heat liberated by L TO and coke combustion,
we referred to the publication by Burger and Sahuqud. For the
stoichiometries of the oxidation reactions, they suggested heats of
reactions for LTO and coke burning of the order of 5.567 x lOS, 1.228
X 106, and 1.893 X lOS Btu/Ibm mol [1.295 x 106, 2.857 X 106, and
4.278 x 105 kJ/kmol] of maltenes, asphaltenes and coke respectively.

1/

where a and b are the masses of oxygen that combines with maltenes
and asphaltenes respectively. Equation 14 reflects the fact that the total
oxygen uptake rate is due to both maltenes and asphaltenes conversion,
and that the total mass of oil in the system is increased by the additive
oxidation reactions. Figure 4 shows the estimated oxygen uptake ratios
for the two LTO reactions, as well as their ability to reproduce the
experimental data. Some scatter is evident, but the trend has been
adequately duplicated. Since the asphaltenes represent a partially
oxidized state they should have fewer oxygen addition sites than the
maltenes. This fact is observed in the lower oxygen uptake ratio.
Based on the L TO uptake ratios, and the assumed molecular
weights of the oil components, the molecularity of the reactions may
now be written:

1.23202

0.48267

0.19359

10-6 exp(7685.2{f)

10-4 exp(5369.2{f)

(19)

(20)

Based on the assumed molecular weights, the asphaltenes viscositytemperature relationship was then inferred from the mixing rule

to give
~asp

4.892

10-25 exp(33147{f)

(22)

Bitumen viscosity at 212 OF [100C] as a function of


asphaltenes mole fraction is shown plotted in Figure 5. Asphaltenes by
themselves are observed to be an essentially immobile component.
Furthermore, any increase in the asphaltenes content of the bitumen is
accompanied by a significant increase in the overall bitumen viscosity.
Similarly, density increases due to oxidation are
accommodated by the model simply by assigning different densities to
the oil pseudo components. The determined specific gravities for the
asphaltenes and maltenes were respectively 1.1580 and 0.9832.
Extending this concept, reduced volatility with increased oxidation
extent is effected by giving the asphaltenes a zero gas/oil equilibrium
K-Value.

(17)
n
-HzO
2

where m is the molar ratio of carbon dioxide to carbon monoxide


produced, and n is the hydrogen/carbon of the coke burned. Three
SPE Advanced Technology Series. Vol. 1. No. I

101

TABLE 1- DATA FOR SYSTEM COMPONENTS


Molecular
Weight

(K)

18.0

373

647

22107

1092.8

1011

1177

792

Gas

43.2

189

295

7176

Coke

13.13

Maltenes

406.7

703

892

1478

Oxygen

32.0

91

154

5046

Component
Water
Asphaltenes

Tb

Pc
(kPa)

Initial Bitumen Mole Fraction


~

= 0.9151

x..

= 0.0849

k-values

Ie; = 0.133 Ci {exp [-DJ(T+Ei)]}/P


Water

Maltenes

Asphaltenes

8.89513E+07

1.41960E+08

O.OOOOOE+OO

3.81644E+03

6.56230E+03

O.OOOOOE+OO

-4.61300E+Ol

-1.93020E+02

O.OOOOOE+OO

Constants

Heat Capacities

Component

CPa

CPb

Asphaltenes

6.010E+02

Gas

4.728E+00

CPd

O.OOOE+OO

CPc
O.OOOE+OO

O.OOOE+OO

1.754E-02

-1.338E-05

4.097E-09

7.950E-Ol

1. 287E-02

-1.042E-05

3.503E-09

Maltenes

2.376E+02

O.OOOE+OO

O.OOOE+OO

O.OOOE+OO

Oxygen

6.713E+00

-0. 879E-06

4. 170E-06

-2.544E-09

Coke

Thermal Conductivity of Fluid Phases


Water = 0.673 W/(m.K)
Oil = 0.155 W/(m.K)
Gas = 0.050 W/(m.K)

Coke density = 1380 kg/m 3

TABLE 2 - ROCK PROPERTIES


Porosity (fraction)
Permeability
Thermal conductivity
Heat capacity
Compressibility

102

0.412
12/lm2

7.0 W/(m.K)
2280 kJ/(m3.K)
6
3 x 10. kPa 1

SPE Advanced Technology Series, Vol. 1, No.1

TABLE 3 - RELATIVE PERMEABILITY DATA FOR RUNS 1 AND 2


Oil/Water

Gas/Oil

krow

S.+Sw

krg

krog

0.00000

1.00000

0.07000

0.10000

0.00000

0.00039

0.88581

0.16000

0.08615

0.00316

0.00156

0.77855

0.21000

0.06632

0.01262

Sw

krw

0.05000
0.10000
0.15000
0.25000

0.00625

0.58478

0.31000

0.03711

0.05050

0.35000

0.01406

0.41869

0.41000

0.01881

0.11362

0.45000

0.02500

0.28028

0.51000

0.00829

0.20199

0.65000

0.03906

0.08651

0.61000

0.00296

0.31562

0.75000

0.05625

0.03114

0.71000

0.00073

0.45449

0.85000

0.07656

0.00346

0.80000

0.00011

0.60106

0.95000

0.10000

0.00000

0.95000

0.00000

0.89080

1.00000

0.12656

0.00000

1.00000

0.00000

1.00000

TABLE 4 - RELATIVE PERMEABILITY DATA FOR RUN 3


Oil/Water

Gas/Oil

Sw

krw

krow

S.+Sw

Krg

krog

0.11000

0.00000

1.00000

0.12000

0.10000

0.00000

0.15000

0.00013

0.87891

0.16000

0.08615

0.00316

0.20000

0.00052

0.76562

0.21000

0.06632

0.01262

0.30000

0.00208

0.56250

0.31000

0.03711

0.05050

0.40000

0.00468

0.39063

0.41000

0.01881

0.11362

0.50000

0.00833

0.25000

0.51000

0.01029

0.20199

0.60000

0.01302

0.14063

0.61000

0.00400

0.31562

0.70000

0.01875

0.06250

0.71000

0.00105

0.45449

0.80000

0.02552

0.01562

0.80000

0.00021

0.60106

0.85000

0.02929

0.00391

0.95000

0.00000

0.89080

1.00000

0.00000

1.00000

0.90000

0.03333

0.00000

1.00000

0.04218

0.00000

TABLE 5 - MODEL INITIALZATION FOR SIMULATIONS


Run #1

Run #2

Run #3

Sand pack length (m)

0.250

1.83

1.83

Sand pack diameter (m)

0.0508

0.0994

0.0994

Number of axial grid blocks

36

36

Initial water saturation

0.0500

0.1180

0.1130

Initial oil saturation

0.2500

0.8820

0.8870

Initial gas saturation

0.7000

0.0000

0.0000

4100

4100

5500

Pressure (kPa)

SPE Advanced Technology Series, Vol. I, No. I

103

600

c::

100

1.0

0.25 '2

500

5'

U
CD
400

0.20

m
... 300

0.15

II)

Q.

l-

II)

II)

E
II)

...

Sa.

j ...... ../

200

::l

I,

100

/~

0.10 e:
II)

"
\

oxygen

2!

-a

0.6

1:.

a.
(/)

0.4

0.00

:J
CD

60

0.2

0.0
0

OJ

eII)

..loI:

40

I..

coke

\.......----

oil sat.

e:

1
I

.,e

/ \

(/)

II)

80

"\

:::iE

~,

0.8

en

" .....
2

II)

~...

I \I

CD

0.05 0

II)
~

n
u:

asph. mole
fraction

()

to

:J

:2
(/)

20

\
~

0
6

Time (hours)

Time (hours)

Fig. 6a - Calculated oxidation behaviour of Athabasca bitumen:


oxygen uptake.

Fig. 6b - Calculated oxidation characteristics of Athabasca bitumen:


oil composition/saturation and residual coke.

APPLICATION OF THE REACTION SCHEME


Combustion Tube Simulator
A general purpose mathematical model of combustion tube
reactors 18 which rigorously includes the annular region surrounding the
sand pack was used in the simulations described below. Our experience
has been that radial heat transfers to and from the sand pack can
significantly affect experimentally observed temperature levels, and
must be accounted for during simulations of combustion tube
experiments. The coded mathematical treatment of in-situ combustion
in the sand pack is essentially that described by Coats 19
Combustion Simulations
Three simulation runs were performed to investigate the
oxygen uptake characteristics of the reaction scheme, and to evaluate
its feasibility as a predictor of fireflooding frontal advance. Rock and
fluid properties are given in Tables 1 to 4, and model initialization data
are presented in Table 5. Note that all non-condensible gas components
other than oxygen have been lumped and referred to in Table 1 as gas.
Run 1 was a simulation of a differential reactor undergoing
ramped temperature oxidation, of the kind reported by Burger and
Sahuquet5 . The combustion tube model was programmed to raise the
temperature of the tube wall by 180 F/hr [100 C/hr] while air (21.0%
02) was injected at a rate of 2.828 scf/hr [0.081 std-m3/hr]. With the
small axial length of the grid specified (see Table 5) temperature and
oxygen concentration gradients were still present. Therefore only the
conditions in the first grid block are reported, and are plotted in
Figure 6.
Two successive oxidation peaks are observed in Figure 6a.
The first occurs at about 1.25 hrs and becomes a maximum at 2.0 hrs.
During this period maltenes are converted to asphaltenes which can be
seen to increase in Figure 6b. As the maltenes become almost depleted,
the oxygen uptake rate drops to a minimum around 2.25 hrs, indicating
that the bitumen has become less reactive. With further increases in
temperature, and reduced maltenes content, asphaltenes are oxidized to
produce a substantial amount of coke (whose initial appearance is
delayed). The thermal cracking reactions also assist in the conversion
of asphaltenes to coke. Finally, coke is entirely consumed around 1110
OF [600C]. The second oxygen uptake peak is therefore due to the
oxidation of asphaltenes and coke combustion. It is worthwhile
mentioning that these dual oxidation peaks have been experimentally
observed 5 , and that the temperature at which the calculated first peak
104

reaches its maximum (518 OF [270 CD agrees very well with those
published data.
This simulation was repeated without any LTO reactions, and
the results plotted in Figure 7. There is a single oxidation peak, and
only a small quantity of coke is deposited. It does not appear feasible
that this amount coke can support an advancing combustion front. It is
also important to note that coke is formed at a much higher temperature
without LTO.
Run 2 simulated a dry enriched air (94.78 % 02) combustion
tube test which has been reported as Test 206 by Belgrave18 . The
stabilized air injection rate was 1.955 scf/hr [0.056 std-m3/hr]. This run
was performed as a test of the ability of the proposed reactions to
duplicate the combustion front velocity as well as fuel and oxygen
consumption.
As was also the case with run 3, the model initialization
procedures closely followed the experiment, since at ignition fluid
saturations are not uniformly distributed. The simulations started from
uniform water and oil saturation and zero gas saturation (Table 5). Gas
was the injected at the top of the vertically oriented sand pack until the
model reached the experimental operating pressure given in Table 5.
Next, gas was flowed through the core until a continuous gas phase
saturation had been established. At this point the injection end of the
core was heated to 752 OF [400C] in run 2 and 572 OF [300 0c] in
run 3. Enriched air injection was then started. Temperature profiles for
run 2 at model times of 12.0 and 15.9 hrs are shown in Figure 8.
Model run time at ignition was 8.20 hrs. The experimental and
calculated (solid lines) leading edges of the combustion front are in
good agreement. Behind the fronts, however, the calculated temperature
profiles were higher than those obtained by experiment. This
discrepancy in radial heat losses is due to a less than adequate
description of the average thermophysical properties of the annular
region surrounding the combustion tube. These properties, which are
very dependent on experimental conditions 18 and are also are affected
by tube operation 21 , were not manipulated during these simulations.
Up to 15.9 hrs, 44 % of the tube had been traversed in the
simulation, and at this time the calculated average coke and oxygen
consumed were 1.44 Ibm/ft3 [23.09 kg/m3] and 64 scf/ft3 [64
std-m3/m3], respectively. The stabilized experimental values for the test
were 1.47lbm/ft3 and 56.4 to 62.9 scf/ft 3 [23.6 kg/m3 and 56.4 to 62.9
std-m3 /m 3]. Figure 9 shows that the produced fluids were also in good
agreement.
Run 3 simulated a superwet combustion tube test that has been
SPE Advanced Technology Series, Vol. 1, No.1

600

c:

100

1.0

:0:;

<U
....

0.25 '2

500

u. 0.8

t;

.--6 400
....
e 300

0.20
,

CD

::l

0.15

temperature

CD

...

CD

CD
~

CD

c:

$
(jj
.c:

CD

<I(

0.10 c:
oxygen
~/"

0
0

" '"

0.4

....
::l 0.2
10
en

{)

0.00
6

0.0

40

oil sat.

0)
Q)

='
III

0.05 0

c:

CD

100

60

0.
f/)

'"E

0.6

:::J

200

asph. mole
I
fraction - - .

f/)

S0.

0.

.::

' - - core

80

~
:::

<U

()
(jj
::l

32
f/)

20

coke

- -

.. .. ..

,/

- -

0
2

Time (hours)

Time (hours)

Fig. 7a - Calculated oxidation characteristics of Athabasca bitumen


without LTO: oxygen uptake.

Fig. 7b - Calculated oxidation characteristics of Athabasca bitumen


without LTO: oil composition/saturation and residual coke.

described as Test 41 by Moore et a1. 20 Enriched air (94% O2) was


injected at a stabilized rate of 1.955 scf/hr [0.056 std-m3/hr] and water
at 1.85 lbm/hr [0.84 kg/hr]. This run was performed to observe the
spatial oxidation behaviour at superwet temperature levels. Figure 10
shows the temperature profile match at 3.0 hrs after ignition. The
simulation approximates the experiment fairly well. For that run time,
the axial distributions of the grid block variables are plotted in Figure
11. The occurrence of asphaltenes mole fractions at locations without
any apparent oil saturation is due to imperceptibly low oil saturations
which are obscured by the scale of the vertical axis. The model has
determined an elongated burning zone with about 1 ft [0.3 m] of coke
saturation, and an increase in the asphaltenes content of the fmite oil
saturation of about 0.5 ft [0.15 m] axial penetration.
More needs to be said about these simulations. The shape of
the relative permeability data presented in Tables 3 and 4 were derived,
for the most part, from history matching of the initial gas flooding and
pre-ignition heating stages of the respective experiments. Very low
values of gas and water relative permeabilities were required to
respectively obtain high enough oil production and to limit the
calculated water produced during this displacement period. Also, for the
sake of reducing the mathematical nonlinearity of the flow equations,
the water relativity permeability curves were extrapolated to values less
than unity at maximum water saturation. This did not significantly
impact the calculated results.
Additionally, the residual oil saturation to gas had to be set at
unusually low values of 0.07 for the dry test (run 2) and 0.12 in the
superwet test (run 3) to match fuel consumptions and firefront
velocities. This suggests a need to include temperature dependency in
the relative permeabilities.

exothermally increases the viscosity and density of the bitumen and


reduces its volatility because of an increase in asphaltenes content.
These factors all serve to reduce the tendency for the oxidized bitumen
. to move away from the firefront, thus increasing the quantity of
hydrocarbon available as fuel for combustion. As the combustion front
approaches the oxidized bitumen, oxygen continues to consume any
unreacted or undistilled maltenes while beginning to consume coke.
With increasing temperature, the asphaltenes are thermally cracked to
produce more coke which further fuels the combustion process. This
description of the combustion process is consistent with the conclusions
of Moore et a1. 22
The main difficulty in the numerical simulation of in situ
combustion, to date, has been the representation of the reaction zone.
Therefore the successful development of a reaction scheme which can
predict combustion performance over a broad range of operating
conditions would be of great significance in predicting the performance
of field tests. A complete description of the actual reaction mechanism
would be very complex and impossible to determine. The unified
pseudo component approach proposed here is both practical and
tractable.
With respect to field scale studies, this reaction model offers
an important feature in its ability to realistically delay the onset of coke
formation; Figure 6 shows that no coke was formed up to a time of 2.0
hours and a corresponding temperature of 392 OF [200C]. Apart from
this attribute the inherent problems of in-situ combustion simulation,
arising from the averaging of firefront temperatures over large grid
blocks, remain unsolved. It is more than likely that apart from dynamic
grid refinement the solution to this problem will come from the
development of pseudo functions for the thermodynamic and mobility
relationships as a function of grid block temperature.

Mechanism of In-Situ Combustion


The reaction model developed in this paper suggests the
following mechanism for in situ combustion of oil sands.
Unreacted oxygen moving through the combustion front
oxidizes the bitumen ahead of it at the lower temperatures downstream,
depositing coke, and at the same time increasing the asphaltenes
fraction of the bitumen, all at the expense of the maltenes. An
important factor here is that the maltenes are by far the more reactive
component. Thus any reduction in maltenes content makes the bitumen
less reactive to oxidation at low temperatures. This reduced reactivity
with increased oxidation extent is consistent with the observations of
Smith and Perkins 21 Oxidation of the bitumen ahead of the firefront
SPE Advanced Technology Series, Vol. I, No. I

CONCLUSIONS
(1)

(2)

(3)

The in-situ combustion reaction scheme described in this paper


represents a consistent consolidation of individual thermal
cracking and oxidation kinetic studics.
It's formulation is comprehensive in that it is able to reproduce
well known oil chemical changes such as increased viscosity
and density as well as reduced reactivity to oxidation, with
increased oxidation extent.
Application of the reaction scheme has confirmed LTO to be
105

1-----;:::::::======;1

800

~ r-------------~--------~

experiment
-model

12.0 hours
.15.9 hours

.-6

sao

-.

sao

-~

-~

as 400

CD
Q.

400

CD
Q.

{!

{!
200

200

~~~

__

~~~

0.4

0.0

__

0.8

~~~~~-J

1.2

0
0.0

2.0

1.6

0.4

Distance (meters)

i
~

"8
It

...

"-

c: 1.0
.2
'0

gas

oil
water
0.9

~$

CD

~CD

:8as

...
.a

12

16

20

Time (hours)
Fig. 9 - Experimental and numerical produced fluids for dry enrichedair combustion test.

the dominant mechanism controlling fuel availability for the in-situ


combustion process. Thermal cracking, in isolation, does not generate
sufficient fuel for high temperature combustion propagation.
(5)
The dual oxidation uptake peaks, associated with ramped
temperature oxidation tests, and the delay in coke formation have
been reproduced by the reaction model, and attributed to
significant differences in reactivity between oxygen and
individual components which make up the oil. This implies that
in-situ combustion cannot be adequately simulated using a single
component oil system.
(6)
The reaction scheme presented here is capable of predicting
experimentally determined frontal velocity, and oxygen and fuel
requirements.

106

0.4

asph. mole
fraction

-----,

15

/ \

I \

10

0.0
0.2

~
.~

~
1:

CD

0.2

0.0
4

I\

c:

0.3 ~

20

\./\

0.6

as

1000

coke

=:

0.6 Q:
2000

25

0.8

:::!E

"8

2.0

30

15

3000

=a

!.1lI

1.6

Fig. 10 - Experimental and numerical temperature profiles of


superwet test at 3.0 hrs after ignition.

1.2

5000

4000

1.2

Distance (meters)

Fig. 8 - Experimental and numerical temperature profiles for dry


enriched-air combustion test.

&

0.8

0.4

0.6

0.8

0
1.0

Distance (meters)
Fig. 11 - Spatial variation of grid block variables for superwet test at
3.0 hrs after ignition.
NOMENCLATURE
a
b

mass of oxygen that combines with unit


mass of maltenes
mass of oxygen that combines with unit
mass of asphaltenes
frequency factor for reaction r
frequency factors for reactions 1, 2, and
3
mass fraction
activation energy for reaction r
activation energies for reactions 1, 2, and
3
rate constant for reaction r
gas relative permeability
SPE Advanced Technology Series, Vol. I, No. 1

kl,k2,k3
m
Pc
POl

Sw
So
t
T
Tb
Tc
X

Jl.

oil relative permeability in gas/oil system


oil relative permeability in oil/water
system
water relative permeability
rate constants for reactions 1, 2, and 3
mass
critical pressure, kPa
oxygen partial pressure, Pa
universal gas constant
water saturation, fraction
oil saturation
time
temperature, K
normal boiling point, K
critical temperature, K
mole fraction
viscosity, mPa.s

8.

9.

10.

11.

12.

SUBSCRIPTS
asp
bitumen
coke
malt
n

r
x

asphaltenes
bitumen
coke
maltenes
hydrogen/carbon ratio of coke
reaction r
oxygen/carbon ratio in carbon oxides

13.

14.

15.

ACKNOWLEDGEMENTS
The authors
gratefully acknowledge the significant financial support of the Alberta
Oil Sands Technology and Research Authority (AOSTRA). Also
acknowledged is the support of NSERC, AMOCO Canada, and the
valuable contributions of the technical staff of the University of
Calgary. The authors also wish to thank Dr. N. Kalogerakis for his
expert assistance in setting up the parameter estimation models.

16.

17.

18.

REFERENCES
1.

2.

3.

4.

5.

6.
7.

Wightman, D., Rottenfusser, B., Kramers, J., and Harrison,


R.: "Geology of the Alberta Oil Sands Deposits," AOS1RA
Technical Handbook on Oil Sands, Bitumens and Heavy Oils,
(1989). Loren G. Hepler and Chu Hsi Eds. AOSTRA Technical
Publication Series #06, 3-9.
Hallam, R.J. and Donnelly, J.K.: "Pressure-Up Blowdown
Combustion: A Channelled Reservoir Reservoir Recovery
Process," SPE 18071 presented at the 63rd Annual Technical
Conference and Exhibition of the SPE, Houston, TX, Oct. 2-5,
1988.
Marchesin, L.A.: "Gregoire Lake Block 1 Pilot, 1976 - 1981" ,
Paper No. 82-33-62 presented at the 33rd Annual Technical
Meeting of the Petroleum Society of CIM, Calgary, AB. June
6-9, 1982.
Hayashitani, M.: Thennal Cracking of Athabasca Bitumen,
Ph.D. Thesis, The University of Calgary, Alberta (1978).
Burger, J.G. and Sahuquet, B.C.: "Chemical Aspects ofIn-Situ
Combustion--Heat of Combustion and Kinetics," Soc. Pet. Eng.
J. (Oct. 1972) 410-422; Trans., AIME, Vol. 253.
Bousaid, I.S. and Ramey, H.J. Jr.: "Oxidation of Crude Oils in
Porous media," Soc. Pet. Eng. J. (June 1968) 137-148.
Alexandcr, J.D., Martin, W.L., and Dew, J.N.: "Factors

SPE Advanced Technology Series, Vol. I, No. I

19.
20.

21.

22.

Affecting Fuel Availability and Composition During In-Situ


Combustion," J. Pet. Tech. (Oct. 1962) 1154-1164; Trans.,
AIME, Vol. 225.
Poettman, F.H., Schilson, R.E., and Surlako, H.: "Philosophy
and Technology of In-Situ Combustion in Light Oil
Reservoirs," Proc., Seventh World Pet. Cong., Mexico City
(1976) Vol.3, 487.
Lerner, S.L., Fleming, G.C., and Lara, P.P.: "Dominant
Processes in In-Situ Combustion of Light Oil Reservoirs," J.
Pet. Tech. (May 1985) 889-900.
Kumar, M.: "Simulation of Laboratory In-Situ Combustion
Data and Effect of Process Variations," SPE 16027 presented
at the 9th SPE Symposium on Reservoir Simulation, San
Antonio, TX, Feb. 1-4, 1987.
Lin, C.Y., Chen, W.H., Lee, S.T., and Culham, W.E.:
"Numerical Simulation of Combustion Tube Experiments and
the Associated Kinetics of In-Situ Combustion Process," Soc.
Pet. Eng. J. (Dec. 1984) 657-666; Trans., AIME, Vol. 277.
Rubin, B. and Buchanan, W.L.: "A General Purpose Thermal
Model," Soc. Pet. Eng. J. (April 1985) 202-214; Trans.,
AIME, Vol 279.
Adegbesan, K.O.: Kinetic Study ofLow Temperature Oxidation
of Athabasca Bitumen, Ph.D. Thesis, The University of
Calgary, Alberta (1982).
Kalogerakis, N. and Luus, R.: "Improvement of Gauss-Newton
Method for Parameter Estimation through the Use of
Information Index," Ind. Eng. Chern. Fundam., (Nov. 1983)
Vol. 22, 436-445.
Thomas, F.B., Moore, R.G., and Bennion, D.W.: "Kinetics
Parameters for the High-Temperature Oxidation of In-Situ
Combustion Coke," J. Can. Pet. Tech. (Nov.lDec. 1985)
60-67.
Bishnoi, P.R., Heidemann, R.A., and Shah, M.K.:
"Calculation of Thermodynamic Properties of Bitumen
Systems", The Oil Sands of Canada-Venezuela (1977). D.A.
Redford, A.G. Winestock Eds. CIM Special Volume 17,
248-255.
Benham, A.L. and Poettmann, F.H.: "The Thermal Recovery
Process - An analysis of Laboratory Combustion Data," Trans.,
AIME (1958), Vol. 213, 406.
Belgrave, J.D.M.: An Experimental and Numerical1nJlestigation
ofIn-Situ Combustion Tube Tests, Ph.D. Thesis, The University
of Calgary, Alberta (1987).
Coats, H.H.: "In-Situ Combustion Model", Soc. Pet. Eng. 1.
(Dec. 1980) 533-554.
Moore, R.G., Bennion, D.W., Belgrave, 1.D.M., Gie, D.N.,
and Ursenbach, M.G.: "New Insights into Enriched Air In Situ
Combustion," SPE 16740 presented at the 62nd Annual
Technical Conference and Exhibition of SPE, Dallas, TX, Sept.
27-30, 1987.
Smith, F.W. and Perkins, T.K.: "Experimental and Numerical
Simulation Studies of the Wet Combustion Recovery Process,"
J. Can. Pet. Tech. (July/Sept. 1973) 44-54.
Moore, R.G., Bennion, D.W., Millour, J.P., Ursenbach,
M.G., and Ursenbach, M.G.: "Comparison of the Effect of
Thermal Cracking and Low Temperature Oxidation on Fuel
Dcposition During In Situ Combuation," Proceedings of the
1986 Tarsands Symposium DOE/METC 87/6073, held in
Jackson, Wyoming, July 1986.

(SPE 20250)

\07

Вам также может понравиться