You are on page 1of 8

Journal of Petroleum Science and Engineering 108 (2013) 259266

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: www.elsevier.com/locate/petrol

The effect of light gas on miscible CO2 ooding to enhance oil recovery
from sandstone and chalk reservoirs
A.A. Hamouda n, V. Alipour Tabrizy
Department of Petroleum Engineering, University of Stavanger, 4036 Stavanger, Norway

art ic l e i nf o

a b s t r a c t

Article history:
Received 23 January 2012
Accepted 26 April 2013
Available online 29 May 2013

Lower oil recovery is observed from sandstone cores than from chalk for synthetic dead and live oils with
miscible CO2 ooding. This was attributed to the greater likelihood of developing ngers in sandstone
cores and lower exposed surface areas compared to chalk. This was assessed by addressing the effect of
the transverse dispersion on viscous instability.
Recombined oil with C1 in chalk and sandstone cores and ooding with CO2 shows higher recovery
than the case of recombination with C1/C3. A ternary diagram was constructed for the two cases in order
to understand the CO2 ooding mechanisms at the different ooding conditions.
& 2013 Elsevier B.V. All rights reserved.

Keywords:
CO2 EOR
live oil
diffusion
dispersion
viscous instability
ternary diagram

1. Introduction
An extensive survey published by Oil and Gas Journal showed
that CO2 ooding has been the most widely used EOR recovery
method for medium and light oil production in sandstone and
carbonate reservoirs during last decades (Moritis, 2006). In the
past ve decades, there have been extensive laboratory studies,
numerical simulations, and eld applications of CO2 EOR processes
(Haskett and Tartera, 1965; Simon et al., 1978; Burke et al., 1990;
De Boer et al., 1995; Farouq Ali and Thomas, 1996; Werner et al.,
1996; Grigg and Schecter, 1997; Idem and Ibrahim, 2002;
Chukwudeme and Hamouda, 2009). In the 1950s, when carbon
dioxide injection was rst used for oil recovery, the emphasis was
on the immiscible process, which offers an alternative recovery
scheme for reservoirs where water-based EOR is impractical or
ineffective. In recent years, interest has shifted to the miscible CO2
process, which can recover oil that is inaccessible by the immiscible method (Holm and Josendel, 1974).
The mechanisms affecting the displacement of oil by CO2
injection include oil swelling, IFT, and viscosity reduction as well
as increasing the injectivity index due to solubility of CO2 in water
and subsequent reaction of carbonic acid with the minerals (Orr
et al., 1982; Jarrell et al., 2002). Displacement efciency for CO2
ooding is also inuenced by wetting properties of the rock,
injection and production rates, density difference between oil
and gas, viscosity ratio of uids, and oil/CO2 relative permabilities

Corresponding author. Tel.: +47 518 32 2 71; fax: +47 518 317 50.
E-mail address: aly.hamouda@uis.no (A.A. Hamouda).

0920-4105/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.petrol.2013.04.013

(Rojas et al. 1991). Al-Wahaibi et al. (2006) investigated the


behavior of two-phase drainage and imbibition-relative permeabilities at near miscible conditions and concluded that as the
interfacial tension decreases, the relative permeability increases
more rapidly in the non-wetting phase than in the wetting phase.
Investigation both in laboratories and in the eld suggests the
importance of gasoil gravity segregation in CO2 EOR (Hagoort,
1980; Bangia et al., 1993; Rao et al., 2004; Verlaan and Boerrigter,
2006; Jadhawar and Sarma, 2008). Viscous ngering may arise due
to the instability of the interface between CO2 and oil that is
caused by the difference in viscosities. This takes place either in
the front or in the rear of a sample plug when there is sufcient
viscosity contrast between the displacing and displaced uid
(Shalliker et al., 2007; Shalliker and Guiochon, 2009). Viscous
ngering and dispersive bypassing increase with oil viscosity
contrast (Shalliker et al., 2007).
Miscible CO2 EOR reduces viscosity and enhances action by CO2
to vaporize and extract hydrocarbons from reservoir oil and to
develop miscibility under certain pressures and temperatures.
Holm and Josendel (1974) measured the relative oil volume for a
mixture of CO2/oil at different pressures and temperatures. They
observed that at constant temperature, when the pressure
increases to the saturation pressure, the oil volume increases
due to oil swelling (CO2 condenses into the oil). If the pressure
increases further, the oil volume decreases due to extraction
(vaporization) of crude oil fractions into a CO2-rich phase. Zick
(1986) showed that a condensing/vaporizing mechanism is occurring both at low pressures when CO2 is a gas and for pressures
above 73 bar, the critical pressure of CO2. Stalkup (1990) showed
that multi-component oil displacements could be by both

260

A.A. Hamouda, V.A. Tabrizy / Journal of Petroleum Science and Engineering 108 (2013) 259266

vaporizing and condensing drive mechanisms and that the minimum miscibility enrichment (MME) and minimum miscibility
pressure (MMP) indicated by ternary diagrams could differ signicantly from those obtained by slim tube tests. Razak et al.
(2009) conducted multiple contact miscibility tests on live reservoir oil to simulate the continuous multiple contact process when
CO2 is injected into the reservoir uid. They demonstrated that the
amount of methane in vapour phase increased signicantly, as the
number of interactions increased between oil and CO2 while the
intermediate fraction became progressively lighter as contact
progressed.
Alavian and Whitson (2011) and Karimaie and Torster (2010)
reported a numerical simulation and experimental investigation of
miscible CO2 injection in a chalk core initially saturated with
reservoir synthetic oil consisting of C1 and n-C7 at a temperature of
85 1C and pressure of 220 bar. The numerical model and experimental study indicated that the CO2 recovery mechanism was
dominated by viscous force at near-rst-contact miscible displacement, with little impact of capillary/gravity displacement.
The work presented in this paper has two main objectives. The
rst is to shed light on the effect of CO2 miscible ooding
conditions on the recovery from chalk and sandstone rocks for
model dead oil (without light components) at minimum miscible
ooding conditions of 50 1C & 90 bar, 70 1C & 120 bar, and 80 1C &
140 bar. The second objective is to address the effect of light gas
components in the oil on recovery by CO2. Synthetic model oil is
used in this study since it is simpler in composition, however
contains various components that could be abundant in reservoir
uids and recognized to have large inuence on the recovery.
These components are fatty acids and asphaltene.

2. Experimental section
2.1. Materials
2.1.1. Solid phase
Core ood experiments were conducted using outcrop Stevens
klint chalks near Copenhagen and outcrop Benthiemer sandstones.
The chalk and sandstone cores were 67 cm and 1113 cm in length,
respectively and 3.8 cm in diameter. The out crop chalks have
approximate porosity of 4048% and absolute permeability of 2.5
6.5 mD. Sandstone cores have lower porosity near 2025% and higher
permeability of 600900 mD. Table 1 depicts core characteristics in
detail, associated uid content and ooding conditions.
2.1.2. Synthetic dead and live oil preparation procedure
The investigation was done for synthetic dead oil and two types
of live oils (A and B). The compositions of dead oil, live oil A, and
live oil B are shown in Table 2. The synthetic dead (model) oil is
composed of asphaltene dissolved in toluene, 0.01 M stearic acid
(SA) dissolved in n-decane (95% purity) for saturating the chalk
cores, and 0.01 M N,N-dimethyldodecylamine (NN-DMDA) dissolved in n-decane for saturating the sandstone cores. Stearic acid
(SA) and N,N-dimethyldodecylamine (NN-DMDA) are polar natural
fatty acid and amine, respectively. They have been used to modify
the wettability of chalk and sandstone cores toward oil-wet.
The synthetic live oils are prepared by recombination of gas (C1
and C3) with the dead oil at a specic gas oil solubility ratio
(Rs 280 SCF/STB). To prepare the live oil, the required mass of
methane (C1) or propane (C3) was injected from gas cylinders near
bubble point pressure to the oil cell containing synthetic dead oil,
using a Gilson pump. The live oil cell was, then, connected to the
rotating live oil mixer at 50 rpm. The pressure of the cell was
checked every 30 min. Then, the pressure was adjusted as needed
using a Gilson pump, to maintain a pressure higher than the

Table 1
Core characteristics and associated uid content and ooding conditions.
Exp no. Core type

Porosity

Oil type

Flooding pressure and


temperature
(bar) and (1C)

Dead oil
Dead oil
Dead oil
Dead oil
Dead oil
Dead oil
Live oil A
Live oil A
Live oil A
Live oil A
Live oil A
Live oil A
Live oil B
Live oil B
Live oil B
Live oil B
Live oil B
Live oil B

90 and 50
120 and 70
140 and 80
90 and 50
120 and 70
140 and 80
90 and 50
120 and 70
140 and 80
90 and 50
120 and 70
140 and 80
90 and 50
120 and 70
140 and 80
90 and 50
120 and 70
140 and 80

(fraction) (cm) (mD)


1a
2a
3a
1b
2b
3b
1c
2c
3c
1d
2d
3d
1e
2e
3e
1f
2f
3f

Chalk
Chalk
Chalk
Sandstone
Sandstone
Sandstone
Chalk
Chalk
Chalk
Sandstone
Sandstone
Sandstone
Chalk
Chalk
Chalk
Sandstone
Sandstone
Sandstone

0.44
0.46
0.48
0.25
0.23
0.22
0.47
0.40
0.44
0.22
0.22
0.24
0.42
0.40
0.43
0.20
0.21
0.23

6
7
6.5
12
13
11
7
6.5
6.5
12
13
13
7
7
6.5
12
13
13

2.5
4
6.5
820
770
650
5
4
6.5
820
770
650
4
2.5
5.5
920
850
700

Table 2
Composition of dead oil, live oil A and live oil B used for CO2 ooding experiments.
Components

n-C10
Stearic acid (or N,Ndimethyldodecylamine)
Toluene
Asphaltene
C1
C3
Total

Dead oil
composition
(mol%)

Live oil A
composition
(mol%)

Live oil B
composition
(mol%)

50.88
0.04

40.63
0.04

39.81
0.04

49.01
0.06
0.00
0.00
100

39.14
0.05
20.14
0.00
100

38.34
0.05
9.87
11.90
100

calculated bubble point pressure of the uid. The mixing process


continued until a stable pressure was attained for at least 24 h.
2.1.3. Asphaltene preparation procedure
The synthetic dead (model) oil system was prepared from
asphaltene precipitated from crude oil in excess of n-heptane
(1:40). The mixture was shaken at least twice-daily for 2 days and
left for 48 h to equilibrate. The mixture solution was then centrifuged
and ltered through a 0.22 m lter (Millipore) and dried for 1 day
using a vacuum oven at room temperature. The dried asphaltene was
then dissolved in toluene.
2.1.4. Experimental procedure
A schematic ow diagram of the experimental setup used for CO2
ooding is presented in Fig. 1. Its major components consist of a core
holder, pressure regulators, two gas ow meters, pressure manometers, Gilson pump, CO2 piston cell, graduated gas/oil separator,
and a lab-view version 7.1 connected to a digital data acquisition
system. The CO2 ooding experiments were done for saturated cores
with synthetic dead oil and synthetic live oils. All cores were aged
with dead oil (containing polar components) for at least 2 weeks, to
change the wettability of cores toward oil-wet. Two weeks of aging
has been shown sufcient for laboratory core ooding experiments
(Hamouda et al., 2009). The CO2 injection was implemented directly
after aging of cores saturated by synthetic dead oil, using a core
holder that consists of a steel cylindrical body and rubber/nylon

A.A. Hamouda, V.A. Tabrizy / Journal of Petroleum Science and Engineering 108 (2013) 259266

261

Fig. 1. Schematic of CO2 ooding setup with graduated gas/oil separator.

sleeve. For CO2 ooding of saturated cores with live oil, the core was
rst saturated with dead oil and then displaced by (fresh) live oil
near one pore volume under pressure of at least 20 bar higher than
the bubble point pressure.
In all experiments, CO2 injection was carried out at constant
pressure and in miscible mode. The oil-saturated core (with live or
dead oil) was ooded with CO2 at minimum miscibility pressure. CO2
was injected into the core at constant pressures of 9070.2, 12070.2,
and 14070.2 bar for corresponding temperatures of 50, 70, and
80 1C. Constant pressure of CO2 was provided by a Gilson pump. CO2
was injected into a core passing mass ow meter, which recorded the
inow properties of CO2 (mass ow rate, density, and total mass). A
back-pressure regulator was connected downstream of the core to
control the pressure during CO2 ooding. The produced uid from
the core was collected in a graduated gas/oil separator, as shown in
Fig. 1. The outow properties (mass ow rate, density, and total
mass) of the evolved gas were also recorded using a ow meter
connected to the separator. The CO2 injection continued for at least
four pore volumes until oil production stopped. When the injection
was terminated, the core was then removed from the core holder
and dried in a vacuum oven at a temperature of 120 1C until a
constant weight was obtained. It was attempted to determine the gas
composition at different stage, however, it was difcult to handle and
the sample had to be diluted with large volume of helium so it was
discontinued.

expressed as diffusion number, Eq. (1):


D
vCO2 l

Nd

where D is the diffusivity coefcient between CO2 and oil (m2/s)


according to Ficks law, is the porosity; l is the length of capillary
tube, and vCO2 is the linear velocity of the injected CO2 through
porous medium. Diffusivity coefcient (D) between CO2 and oil
can be estimated from the following equation (Renner, 1983):
M 0:6898
V 1:706
P 1:831 T 4:524
D 109 0:4562
CO2
oil
CO2

On a microscopic scale, CO2 diffuses/disperses and mixes with


the surrounding resident oil, creating concentration gradients. The
relative magnitude of diffusion to the convective dispersion
(inverse of Peclet number) during miscibility processes may be

where CO2 is the viscosity of CO2 (cp), M oil is the molecular weight
of oil, V CO2 is the molar volume of CO2 (cm3/mol), P (psia) is the
pressure of the CO2 and oil system at equilibrium, and T (Kelvin) is
the temperature of the CO2 and oil system at equilibrium. Both
diffusion and convective dispersion have profound inuence on
miscible processes. At low injection (or production) rate, diffusion
tends to dominate. Convective dispersion is enhanced by high
injection (or production) rate.
Dispersion is characterized by longitudinal dispersion coefcient (m2/s), KL (in the direction of gross uid ow), and transverse
dispersion coefcient (m2/s), KT (transverse to the direction of
gross uid ow) is suggested by Perkins and Johnston (1963) and
Alkindi et al. (2011).
KL
1
vdP
0:5

3
F
D
D
KT
1
vdP
0:0157

F
D
D

3. Theory

where F is the Archie resistivity factor, v is Darcy velocity (m/s),


and dP is particle diameter (m). The Archie resistivity factor can be
estimated from Eqs. (5) and (6) for sandstone and carbonate rocks,
respectively (Tiab and Donaldson, 2004).
0:62
F 2:15
5

262

1
2

100.00

For consolidated porous media, particle diameter (dp) can be


estimated from the CarmanKozeny correlation according to
Eq. (7) (Tiab and Donaldson, 2004):
!
1
3
K
7
2
5SV gr 12

80.00

Oil Recovery (%)

A.A. Hamouda, V.A. Tabrizy / Journal of Petroleum Science and Engineering 108 (2013) 259266

60.00
40.00
20.00

where K is the permeability of porous media (m2), SV gr is the total


area exposed within the pore space per unit of grain volume, and
dP is particle diameter (m).
Perkins and Johnston (1963) and Alkindi et al. (2011) dened
low velocities where 0:5vdP =D0:1, the molecular diffusion,
dominates in longitudinal mixing. For moderate velocities
0:10:5vdP =D4, both molecular diffusion and convective mixing
are important transport mechanisms and at higher ow rates
0:5v dP =D4, convective dispersion dominates the mixing process. Similarly in the case of transverse dispersion, molecular diffusion
dominates the transverse mixing process when 0:5v dP =D50,
while convective dispersion is found to be the dominant mechanism
in transverse mixing when 0:5v dP =D300.
In miscible gas ooding processes, the transverse dispersion
tends to merge the formed ngers by increasing ngers width;
hence reducing the viscous instability. This depends on time and
the rate of transverse dispersion. Marle (1981) suggested the
following equation to address the effect of transverse dispersion
on viscous instability (dispersivity ratio).
N FV

K T L
vw2

where L is displacement length (m) and w is the width over which


displacement takes place (m). The transverse number is the ratio
between the convective transport time L=v to dispersive transport time w2 =K T . Larger values of NV F indicate that when the
convection time is longer than the dispersion time, the displacement tends to be more stable.

4. Minimum miscibility pressure (MMP) estimation


The minimum miscibility pressures at different temperatures
(50 1C, 70 1C, and 80 1C) are calculated according to the empirical
correlations dened by Alston et al. (1985), Yelling and Metcalfe
(1980), Glas (1985), and Johnson and Pollin (1981) as well as
using SRK EOS, version 17 of PVTsim. Hamouda et al. (2009)
showed that calculated MMP values using PVTsim lay between the
above mentioned empirical correlations and are used in this work.

5. Results and discussion


5.1. Oil recovery from chalk and sandstone by CO2 miscible ooding
Fig. 2a and b shows that after 2PV injection of CO2 under the
same ooding conditions, more oil was recovered from chalk cores
than from sandstone for saturated cores with dead or live oil.
These gures also demonstrate that the higher the miscible
ooding is, the higher the oil recovery. This is in agreement with
Gardner and Ypma (1984), who explained that at higher pressure,
the CO2-rich phase becomes sufciently enriched to form a single
phase with the oil. However, it does not explain the observed
higher recovery from chalk than from sandstone.

0.00

90 bar-50 C

120 bar-70C

140 bar-80 C

100.00
80.00

Oil Recovery (%)

6
SV gr
dp

60.00
40.00
20.00
0.00

90 bar-50C

120 bar-70C

140 bar-80C

Fig. 2. Comparison between oil recovery from chalk and from sandstone cores
saturated with (a) dead oil and (b) live oils (A and B) as a function of ooding
conditions.

The oil recovery from live oil, under the same ooding conditions,
was lower than from dead oil. Moreover, live oil A (recombined
model oil with about 20 mol% methane) showed higher recovery
than did live oil B (recombined model oil with about 9 and 12 mol%
of methane and propane, respectively). The content of light components in both oils was kept almost the same (2072 mol%).
In summary, there are two main observations. First, oil recovery
from chalk was higher than that from sandstone. Second, recovery by
CO2 injection into live oil was lower than that into dead oil. Moreover, the recovery from chalk increases as the miscible ooding
conditions increase, with a difference of about 9% between high and
low conditions. A similar trend is observed with sandstone, but with
a difference of about 6%. The experimental error was within about
2%. The rst observation is addressed in the next section based on
possible development of ooding front instability/by passing phenomenon that may have caused the difference in the oil recovery.
5.2. Development of stable/instable front with CO2 ooding
In order to explain the difference in oil recovery from chalk
versus sandstone, the stability of CO2 front is addressed in this
section. CO2 is in a direct contact with the original oil at the
entrance near where the streamlines are developed, which represent the average paths of the CO2.
Microscopically, CO2 moves along random paths, which are not
uniform in cross section and hence diffuse/disperse into the
surrounding resident oil, creating concentration gradients. In cases
of high permeability contrast, ngers could develop where certain
paths with lower pressure drops would be the preferential
ow path.
Factors that inuence initiation of ngers have been discussed
extensively by Collins (1976). Finger propagation rate is mainly
dependent on the mobility of the displacing uid and local
pressure drop as well as porosity. High mobility and pressure
drop promote nger growth, whereas porosity has the opposite
effect. The present work found both sandstone and chalk outcrops
to have a reasonable natural homogeneity compared to reservoir

A.A. Hamouda, V.A. Tabrizy / Journal of Petroleum Science and Engineering 108 (2013) 259266

rock cores. Pressure drop was kept constant at about 2 bar in all
ooding experiments. The porosity of sandstone cores was about
50% less than that of chalk. In addition, the permeability of the
sandstone cores was two orders of magnitude larger than that of
chalk. Chalk cores have smaller pore throat sizes and higher
surface areas than sandstone cores. Calculation based on the
Kozeny correlation showed that for the tested chalk cores, the
surface area varied between 1.1 and 1.7 m2/g; whereas for sandstone cores, the surface area varied between 0.02 and 0.04 m2/g.
The higher surface area of chalk cores exposed to the owing CO2
could enhance the extraction efciency. In addition, the probability of permeability contrast in sandstone within a core is expected
to be larger than in chalk.
The above arguments suggest that development of unstable
displacement in uids within sandstone cores is more likely than

0.1
0.09

Chalk core saturated with dead oil

0.08

Sandstone core saturated with dead oil

N FV

0.07
0.06

0.051

0.05

0.044

0.04
0.03

0.034
0.027

0.023

263

within chalk. In order to compare the development of viscous


ngering during CO2 ooding of chalk and sandstone with dead and
live oils, the equation suggested by Alkindi et al. (2011) was used.
Figs. 3 and 4 show the viscous ngering number (NFV), which is
indication of the viscous instability for dead oil and for live oil A and B,
respectively. Fig. 3 shows the effect of CO2 miscible ooding conditions
on the viscous instability of chalk and sandstone saturated with dead
oil. The convective transport time's being higher than the dispersive
transport time, indicates that ooding at elevated conditions (140 bar
and 80oC) tends to enhance viscous stability. This may suggest that
sandstone is more likely to develop viscous instability than chalk. This
supports the above argument regarding factors that promote ngering. It is also in agreement with the visually inspected cores after
ooding. It is worth mentioning that gravity segregation occurs in
both sandstone and chalk cores, which is caused by the density
difference between oil and CO2. The density of oil was estimated to be
between 0.4 and 0.6 g/cc, and the density of CO2 was estimated to be
between 0.25 and 0.3 g/cc depending on pressure and temperature.
Fig. 4a and b compares NVF for the two live oils. Similar trends are
observed as for dead oil, where the higher the ooding conditions, the
more stable the ooding. It also shows that sandstone is more likely to
form an unstable ooding front. The pictures of cores after ooding
shown in Fig. 5 indicate that the CO2/oil front is more stable in chalk
core. In contrast, for sandstone cores, clear ngers developed and
penetrated the core, causing an early breakthrough hence lower

0.018

0.02
0.01
0

90 bar-50 C

120 bar-70 C

140 bar-80 C

Fig. 3. Viscous ngering number (NFV) as a function of ooding conditions of chalk


and sandstone saturated with dead oil.

0.1
0.09

Chalk core saturated with live oil A

0.08
0.07

Sandstone core saturated with live oil A

NFV

0.06
0.05
0.04

0.042
0.036

0.034

0.031

0.03

0.024

0.02

0.015

0.01
0

90 bar-50C

120 bar-70C

140 bar-80C

Fig. 5. Pictures of CO2oil front instability/ngering after miscible ooding of


(a) chalk and (b) sandstone cores. (The interface between oil swept by CO2 and the
less swept oil is traced by dots).

0.1
0.09

Chalk core saturated with live oil B


100

0.08
Sandstone core saturated with live oil B

80

0.06
0.05

0.040

0.04
0.03
0.02

0.032
0.024

0.030
0.021

0.012

0.01

Oil Recovery (%)

NFV

0.07

Chalk cores saturated with dead oil


Sandstone cores saturated with dead oil

60

40

20

0
90 bar-50C

120 bar-70C

140 bar-80C

Fig. 4. Comparison of viscous ngering number (NFV) as a function of ooding


conditions for (a) chalk and sandstone cores saturated with live oil type A (with C1
as a recombined gas and GOR 280 SCF/STB) and (b) chalk and sandstone cores
saturated with live oil type B (with C1 and C3 as a recombined gas and GOR 280
SCF/STB).

0
1.0E-03

1.0E-02

1.0E-01

Nd

Fig. 6. Oil recovery from saturated chalk and sandstone cores with model oil, as a
function of diffusion number.

264

A.A. Hamouda, V.A. Tabrizy / Journal of Petroleum Science and Engineering 108 (2013) 259266

sweep efciency than for chalk cores. Fig. 4b shows that introducing
heavier gas component such as propane (C3) reduces the stability of
the ooding front as indicated by the calculated NVF. In chalk cores,
viscous ngering could be initiated due to viscosity contrast between
oil and CO2. The higher the viscosity ratio the more likely viscous
ngering develops; however due to higher transverse dispersion and
higher surface area in chalk cores compared to that in sandstone cores,
ngers could merge which results in a better front displacement
stability, hence higher oil recovery. It is worth mentioning that for
chalk cores, both molecular diffusion and convective mixing are
important transport mechanisms compared to sandstone cores, where
only convective dispersion dominates the mixing process.

5.3. Diffusion/convective dispersion


The physical properties of CO2 that make it widely used in
extraction processes are low surface tension and viscosity and high
diffusivity. The last permits a rapid mass transfer, hence higher
extraction rate. Fig. 6 shows oil recovery as a function of the
diffusion number (Nd) for sandstone and chalk cores saturated
with dead oil. All the cases show higher oil recovery as diffusion
number increases. Fig. 7 compares all the cases for live oil and
dead oil. The diffusion number for live oil A was slightly higher
than for live oil B, for both sandstone and chalk. However, recovery
was less for live oil A.

100

80

Oil Recovery (%)

In summary, sandstone cores show a less stable front that is


more likely to develop ngers compared to the chalk cores. Also,
when the ooding is performed at elevated miscible ooding
conditions, the ooding front is more stable. However, recombining the model dead oil with light components enhanced the
probability of developing an unstable ooding front. This may
help to explain the effect of the lighter components on ooding
front instability and oil recovery compared to dead oil. However,
the small difference in NFV of live oil with heavier gas (C3) could be
considered to be within the experimental error.

60

40

20

0
1.0E-03

1.0E-02

1.0E-01

Nd
Chalk cores saturated with dead oil

Sandstone cores saturated with dead oil

Chalk cores saturated with live oil A

Sandstone cores saturated with live oil A

Chalk cores saturated with live oil B

Sandstone core satuared with live oil B

Fig. 7. Comparison of oil recovery from all the studied cases, as a function of Nd for
sandstone and chalk cores saturated with dead and live oils.

5.4. Phase behavior of CO2/oil system


Ternary diagrams of the CO2oil system at the two miscible
ooding conditions were generated using a software program (PVTi
module of Eclipse version 2010.2). The two miscible conditions
selected here were the largest ooding conditions, i.e., 80 1C and
140 bar for live oils A and B. As CO2 dissolves in the oil, it reduces the
viscosity and interfacial tension. It also vaporizes and extracts hydrocarbons from the oil. As a result of this process, composition gradient

Fig. 8. Pictures of produced uids during CO2 ooding of dead model oil at ooding conditions of (a) T 50 1C and P 90 bar and (b) T 80 1C and P 140 bar of chalk cores.

A.A. Hamouda, V.A. Tabrizy / Journal of Petroleum Science and Engineering 108 (2013) 259266

265

develops within the ooded region. The gradient depends on the


initial composition of the oil and the miscible ooding conditions.
The composition gradient was visually observed during CO2 ooding. As an example, two pictures were taken from the produced uid,
representing low and high miscible ooding conditions, 50 1C/90 bar
and 80 1C/140 bar, respectively. Fig. 8 shows the different color
gradations of the produced uids before and after CO2 breakthrough
for experiments 1a and 3a, respectively. Fig. 8a shows two production samples with different colors that correspond to experiment 1a
at (50 1C and 90 bar), while Fig. 8b shows three distinct color
graduations that correspond to experiment 3a (80 1C and 140 bar).
The lighter color samples were obtained after the breakthrough. In
other words, at higher ooding pressure and temperature, the color
shows more distinct gradations compared to that for lower ooding
pressure and temperature. This qualitatively demonstrates the gradual
composition that developed in the CO2 vaporizing/condensing region
(s) as it contacts the residual oil, which is behind the original oil.
The phase behavior of CO2/live oil A and CO2/live oil B systems,
ooded at 80 1C and 140 bar, is perhaps best represented by pseudoternary diagrams where each apex of the equilateral triangle represents 100% of a single (or grouped) component. Fig. 9 shows the phase
behavior of the two uid systems (oil A and oil B). For live oil A/CO2
system, the lightest component (C1) is plotted at the top apex,
intermediate component (CO2) at the lower right apex, and heaviest

component (C7+) at the lower left apex. For the live oil B/CO2 system,
at the top apex is CO2, at the right apex is C1C3, and at the left one is
C7+. For live oil B (Fig. 9b), the two-phase region does not exist,
whereas live oil A does exhibit a two-phase diagram. In both cases, the
phase diagram (Fig. 9) indicates CO2 ooding at 80 1C and 140 bar, to
be a rst contact miscibility ooding. This may suggest higher oil
recovery is to be expected for oil B, but this is not the case. Again, the
ooding pressure and the pressure drop during the experiment were
kept constant at 140 bar and 1 bar, respectively. If it is assumed that
the pressure drop, during the ooding, had reached 5 bar, at constant
temperature (80 1C), i.e. the pressure would have been 135 bar instead
of 140 bar and CO2 would have contacted live oil A, causing enriched
liquid composition and condensation. In contrast, in the case of live oil
B, vaporization is the dominant process as shown in Fig. 10. The
monitored pressure drop hardly exceeded 2 bars for any of our
experiments, which would produce a dominant ooding mechanism
similar to that at 135 bar (calculated but not included). Oil recovery
may then be slightly more efcient for live oil A than for live oil B,
which is opposite the expected ooding efciency from the phase
diagram at 140 bar. This is in agreement with the obtained recovery
trend shown in Fig. 2. As pressure drop occurs across the core, the
mechanism will be dominant by CO2 condensation for live oil A and oil
vaporization for live oil B. The lower CO2 diffusivity in live oil B and
free vapour (gas) due to vaporization mechanism may explain the
lower oil recovery for live oil B compared to live oil A and dead oil.

Fig. 9. Ternary diagrams for CO2 ooding conditions of 80 1C and 120 bar for
(a) live oil A/CO2 and (b) live oil B/CO2 systems.

Fig. 10. Ternary diagrams for CO2 ooding conditions of 80 1C and 135 bar for (a)
live oil A/CO2 and (b) live oil B/CO2 systems.

266

A.A. Hamouda, V.A. Tabrizy / Journal of Petroleum Science and Engineering 108 (2013) 259266

The uid phase behavior and diffusion numbers indicate that live
oil A and live oil B should yield higher recovery than the dead oil. The
used oil contains 0.35 wt% of asphaltene. When the oil is recombined
with the light gas components (C1 and C1/C3) and ooded with CO2,
asphaltene deposition may occur. Hamouda et al. (2009) studied CO2
enhanced oil recovery using the same source of asphaltene, concentration (0.35 wt%), and CO2 ooding conditions as in the present work.
They concluded that the deposited asphaltene enhanced the sweep
efciency (Hamouda et al., 2009). If asphaltene is precipitated in larger
rock pores during miscible CO2 ooding, the divergence of solvent
(CO2) into other pores will be enhanced; hence the sweep efciency
may increase.
In summary, the phase diagrams for live oil A and live oil B
demonstrate that at 80 1C and 140 bar, rst contact miscibility is
developed, which anticipates efcient oil recovery. If one assumes
138 bars due to a 2 bar pressure drop in the core, higher oil
recovery from oil A than from oil B is anticipated. This is in
agreement with the obtained recovery results.
6. Conclusions
This work showed lower oil recovery from sandstone cores
than from chalk. This is explained by the greater likelihood of
developing ngers in sandstone cores due to the probability of
their having larger permeability contrast than chalk cores. Visual
observation of the cores after ooding, conrmed nger development and propagation in sandstone core.
It is shown that recombination of methane (C1) and methane
and propane (C1 and C3) with the oil affected the recovery by CO2
miscible ooding and C1/C3 gas system reduced the recovery more
than in the case of recombined oil with C1. Two possible explanations are given. First is the presence of light component (C1)
reduces the CO2 extraction efciency; with further addition of C3,
further reduction of the CO2 efciency occurred. The second is that
during the ooding, pressure reduction within the core induced
multiple contact miscibility rather than rst-contact miscibility.
Pressure drop occasionally exceeded 1 bar and reached an average
of 2 bar. Examination of the ternary diagram indicated that
evaporation dominated the C1/C3 system and condensation dominated the C1 system.
Acknowledgments
The authors wish to thank the University of Stavanger for its
nancial support, Krzysztof Piotr Dziadosz for his undeniable
technical support, Svein Myrhen for Lab view support, and Inger
Johanne for her enthusiasm and great assistance in getting the
chemicals used in this work.
References
Alavian, S.A., Whitson, C.H., 2011. Numerical modeling CO2 injection in a fractured
chalk experiment. J. Pet. Sci. Eng. 77, 172182.
Alkindi, A., Al-Wahaibi, Y., Bijeljic, B., Muggeridge, A., 2011. Investigation of
longitudinal and transverse dispersion in stable displacements with a high
viscosity and density contrast between the uids. J. Contaminant Hydrol. 120121, 170183.
Alston, R.B., Kokoris, G.P., Jams, C.F., 1985. CO2 minimum miscibility pressure: a
correlation for impure CO2 stream and live oil systems. SPE J. 25, 268274.
Al-Wahaibi, Y.M., Grattoni, C.A., Muggeridge, A.H., 2006. Drainage and imbibition
relative permeabilities at near miscible conditions. J. Pet. Sci. Eng. 53, 239253.
Bangia, VK., Yau, F., Hendricks, G.R., 1993. Reservoir performance of a gravity-stable,
vertical CO2 miscible ood: Wolfcamp reef reservoir, Wellman unit. SPE Res.
Eng. 8, 261269.

Burke, N.E., Hoobbs, R.E., Kashon, S.F., 1990. Measurement and modeling of
asphaltene precipitation. J. Pet. Tech. 42, 14401446.
Chukwudeme, E.A., Hamouda, A.A., 2009. Enhanced oil recovery (EOR) by miscible
CO2 and water ooding of asphaltenic and non-asphaltenic oils. Energies 2,
714737.
Collins, R., 1976. Flow of Fluids Through Porous Materials. Petroleum Publishing
Company, Tulsa, USA.
De Boer, R.B., Leerlooyer, K., Eigner, M.R., Van Bergen, A.R.D., 1995. Screening of
crude oils for asphaltene precipitation: theory, practice and the selection of
inhibitors. SPE Prod. Facilities 10, 5561.
Farouq Ali, S.M., Thomas, S., 1996. The promise and problems of enhanced oil
recovery methods. J. Can. Pet. Technol. 7, 5763.
Gardner, J.W., Ypma, J.G.J., 1984. An investigation of phase behavior/macroscopicbypassing interaction in CO2 ooding. SPE J. 24, 508520.
Glas, ., 1985. Minimum miscibility pressure correlation. SPE J. 25, 927935.
Grigg, R.B., Schecter, D.S., 1997. State of the industry in CO2 oods. Paper SPE 38849
Presented at SPE Annual Technical Conference and Exhibition, Texas, USA,
October 58 1997.
Hagoort, J., 1980. Oil recovery by gravity drainage. SPE J. 20, 139150.
Hamouda, A.A., Chukwudeme, E.A., Mirza, D., 2009. Investigating the effect of CO2
ooding on asphaltenic oil recovery and reservoir wettability. Energy Fuel 23,
11181127.
Haskett, C.E., Tartera, M., 1965. A practical solution to the problem of asphaltene
deposits-Hassi Messaoud eld. J. Pet. Tech. 4, 387391.
Holm, L.W., Josendel, V.A., 1974. Mechanisms of oil displacement by carbon dioxide.
J. Pet. Tech., 14271436.
Idem, R.O., Ibrahim, H.H., 2002. Kinetics of CO2-induced asphaltene precipitation
from various Saskatchewan crude oils. J. Pet. Sci. Eng. 35, 233246.
Jadhawar, P.S., Sarma, H.K., 2008. Scaling and sensitivity analysis of gasoil gravity
drainage EOR. Paper SPE 115065 Presented at SPE Asia Pacic Oil and Gas
Conference and Exhibition, Perth, Australia. October 2022, 2008.
Jarrell, P.M., Fox, C.E., Stein, M.H., Webb, S.L., 2002. Practical aspects of CO2 ooding.
SPE Monogr., 22.
Johnson, J.P., Pollin, J.S., 1981. Measurement and correlation of CO2 miscibility
pressure. Paper SPE 9790 Presented at SPE/ODE Enhanced Oil Recovery
Symposium, Tulsa, Oklahoma, April 58, 1981.
Karimaie, H., Torster, O., 2010. Low IFT gasoil gravity drainage in fractured
carbonate porous media. J. Pet. Sci. Eng. 70, 6773.
Marle, C.M., 1981. Multiphase Flow in Porous Media. Editions Technip, Paris.
Moritis, G., 2006. EOR survey. Oil and Gas Journal 104 (15), 3757.
Orr Jr., F.M., Silva, M.K., Lien, C.L., Pelletier, M.T., 1982. Laboratory experiments to
evaluate eld prospects for CO2 ooding. J. Pet. Tech. 34, 888898.
Perkins, T.K., Johnston, O.C., 1963. Review of diffusion and dispersion in porous
media. SPE J. 3, 7084.
Rao, D.N., Ayirala, S.C., Kulkarni, M.M., Sharma, A.P., 2004. Development of gas
assisted gravity (GAGD) process for improved light oil recovery. Paper SPE
89357 Presented at SPE/DOE Symposium on Improved Oil Recovery, Tulsa,
Oklahoma, April 1721, 2004.
Razak, W., Daud, W., Faisal, A.H., Carigali, S.D., Zakaria, N.A., 2009. Multi component
mass transfer in multiple contact miscibility test; forward and backward
method. Paper SPE 125219 Presented at SPE/EAGE Reservoir Characterization
and Simulation Conference, Abu Dhabi, October 1921, 2009.
Renner, T.A., 1983. Measurement and correlation of diffusion coefcients for CO2
and rich-gas applications. SPE. Res. Eng. 3, 517523.
Rojas, G.A., Zhu, T., Dyer, S.B., Thomas, S., Farouq Ali, S.M., 1991. Scaled model
studies of CO2 oods. SPE Res. Eng. 6, 169178.
Shalliker, R.A., Catchpoole, H.J., Dennis, G.R., Guiochon, G., 2007. Visualizing viscous
ngering in chromatography columns: high viscosity solute plug. J. Chromatogr.
A 48, 1142.
Shalliker, R.A., Guiochon, G., 2009. Understanding the importance of the viscosity
contrast between sample solvent plug and the mobile phase and its potential
consequence in two-dimensional high-performance liquid chromatography.
J. Chromatogr. A 1216, 787793.
Simon, R., Rosman, A., Zana, E., 1978. Phase behavior properties of CO2-reservoir oil
systems. SPE J. 18, 2026.
Stalkup, F.L., 1990. Effect of gas enrichment and numerical dispersion on enrichedgas-drive predictions. SPE Res. Eng. 4, 647655.
Tiab, D., Donaldson, E., 2004. Theory and practice of measuring reservoir rock and
uid transport properties, 2nd ed. Elsevier. (Online version available at: http://
www.knovel.com/web/portal/browse/display?_).
Verlaan, M., Boerrigter, P., 2006. Miscible gasoil gravity drainage. Paper SPE 10399
Presented First International Oil Conference and Exhibition, Cancun, Mexico,
August 31September 2, 2006.
Werner, A., Behar, F., De Hemptinne, J.C., Behar, E., 1996. Thermodynamic properties of petroleum uids during expulsion and migration from source rocks. Org.
Geochem. 24, 10791095.
Yelling, W.F., Metcalfe, R.S., 1980. Determination and prediction of CO2 minimum
miscibility pressure. J. Pet. Technol. 32, 160168.
Zick, A., 1986. Combined condensing vaporizing mechanisms in the displacement of
oil by enriched gases. Paper SPE 15493 Presented at SPE Annual Technical
Conference and Exhibition, New Orleans, October 58, 1986.

You might also like