You are on page 1of 13

SPE 113993

Investigation into the Processes Responsible for Heavy Oil Recovery by


Alkali-Surfactant Flooding
J. Bryan, University of Calgary and TIPM Laboratory; A. Mai, University of Calgary and TIPM Laboratory, Laricina
Energy Ltd.; A. Kantzas, University of Calgary and TIPM Laboratory
Copyright 2008, Society of Petroleum Engineers

This paper was prepared for presentation at the 2008 SPE/DOE Improved Oil Recovery Symposium held in Tulsa, Oklahoma, U.S.A., 1923April2008.

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
This paper describes a suite of alkali-surfactant (AS) floods that were performed in systems containing viscous heavy oil
(11,500 mPas). The study investigates how AS injection can be used to generate oil and water emulsions, which can in turn
lead to improved sweep efficiencies and oil recovery. Data is obtained from core flooding, with in-situ saturation
measurements made using low field NMR. This work is applicable to the many heavy oil reservoirs in countries like Canada
and Venezuela that contain viscous oil that still has some limited mobility under reservoir conditions. In previous studies,
improved oil recovery compared to waterflooding was observed. This work provides additional information that can be used
to better understand how chemical injection can lead to oil recovery.
The core floods in this study indicate that emulsification is most efficient when used to block pre-formed water channels
and improve the sweep efficiency of the flood. Both O/W and W/O emulsions may form in the same system, even under
controlled salinity conditions. The re-distribution of water from the flooded channels into emulsified droplets in the oil is at
least partially responsible for the pressure increase seen in these systems. W/O emulsification is accompanied by wettability
alteration, as evidenced by the NMR spectra obtained. After the chemical flood is completed, it may be possible to restore
the original water wet condition of the rock, which can provide potential for future non-thermal improved oil recovery.

Introduction
Several countries in the world, notably Canada and Venezuela, contain significant deposits of heavy oil and bitumen. These
oil sands are characterized as unconsolidated, high porosity and permeability reservoirs. The viscosity of the oil in place may
range from tens to millions of mPas (cP) at reservoir conditions, and the oil densities will approach or even be higher than
that of water. Due to the significant size of these oil sand deposits, coupled with rising oil prices and demand, international
interest is now shifting towards recovery of this heavy oil and bitumen.
Heavy oil reservoirs are a special subset of the oil sands, whereby the oil viscosity at reservoir temperature and pressure
ranges from around 50 50,000 mPas. This oil, while still highly viscous, has some limited mobility at reservoir conditions.
The oil may also contain some dissolved solution gas, and initially it may be possible to recover a fraction of the oil through
primary production. Recent estimates by the AEUB
1
put the primary recovery of heavy oil at an average of around 5%
OOIP, meaning that at the end of primary production there are still significant oil resources remaining as potential for
secondary and tertiary recovery. Unfortunately, many of the reservoirs in Canada are relatively small or thin, and were
possibly disturbed during primary production. As a result, reservoirs these are not prime candidates for expensive thermal or
hydrocarbon solvent enhanced oil recovery technologies. Therefore, in-expensive (non-thermal) methods of recovering the
oil have to be considered.
Previous research in our group
2-5
has focused on improved heavy oil recovery through the application of waterflooding
and alkali-surfactant (AS) flooding. In this work, several conclusions have been reached. First, during a heavy oil
waterflood, water will break through very early in the life of the flood due viscous instabilities that are present as a result of
the adverse mobility ratio between oil and water
6-10
. After water breakthrough, however, continuous channels of water are
present throughout the reservoir. At this time, theories of instability and viscous fingering
9,10
can no longer be used to
explain the mechanism of oil recovery through waterflooding. Darcys law for two-phase flow predicts that in the situation
of a reservoir containing viscous oil and water, further injected water should simply channel through the low-resistance water
2 SPE 113993
pathways, and there is little remaining driving force for additional oil production. However, experience has proven that
significant oil can still be recovered after breakthrough in a waterflood.
It has been shown
2,3
that at later stages in a heavy oil waterflood, capillary forces and water imbibition are the dominant
mechanism for heavy oil recovery during waterflooding. At low injection rates, significant volumes of heavy oil can be
recovered after water breakthrough, albeit at a high water cut. AS injection could either be considered as a
primary/secondary recovery process on its own, or could alternatively be performed after the conclusion of a waterflood. Our
previous experience
4,5
has shown that AS injection after waterflooding tends to be a more efficient process. Therefore, the
proposed injection scheme would be to first waterflood the reservoir and use capillary forces to aid in recovering as much oil
as economically possible. At the same time, laboratory and pilot studies could be considered for AS flooding, since these
processes are considerably more complex and depend strongly on the rock-fluid interactions for each specific reservoir.
Overall, a combination of waterflooding and AS flooding, if properly designed and controlled, can lead to significant
improved oil recovery beyond that of primary production. In this study, data is provided for several AS floods performed at
different rates into an unconsolidated sand pack. The pressure and recovery data are analyzed to infer how the chemical
flood is working, and NMR spectra of the fluids in the sand pack are monitored to understand the wettability of the core as
the flood progresses.

Chemical Injection in Heavy Oil Systems
Surfactants are a special class of molecules that are both hydrophobic and hydrophilic. This unique property allows them to
position stably at the oil-water interface, thus displacing molecules of immiscible oil and water away from the interface with
the overall result being a decrease in the oil-water interfacial tension
11,12
. Alkaline flooding is a subset of surfactant flooding,
whereby the surfactants in question are generated in-situ through the reaction between organic acids in the oil and the injected
alkaline solution
13
. During immiscible displacement in conventional oil waterflooding, residual oil exists as discontinuous
ganglia, trapped by capillary forces. Reduction in the oil-water interfacial tension can lead to higher capillary numbers and
improved oil recovery
14
. The influence of surfactants (either formed in-situ or preformed) can also lead to emulsification of
the fluids and possible alterations in wettability
15
. In conventional oil systems, wettability alteration from water wet to oil
wet conditions can lead to changes in the oil relative permeability behavior and oil production at very low saturations through
film drainage
16
. Challenges for field applications of chemical injection usually revolve around loss of chemical to the rock
surface
17,18
, but in general the oil recovery mechanisms have been well-defined.
There have been several successful laboratory studies performed for chemical injection (alkali, surfactant or AS) into
heavy oil systems. Some of the recovery mechanisms proposed have been reduction in the oil-water interfacial tension
19,20
,
the generation of water-in-oil (W/O) emulsions
5,21,22
and the generation of oil-in-water (O/W) emulsions
4,23-26
. The fact that
various researchers have identified different oil recovery mechanisms points to the complexity of interpreting these chemical
floods.
The key to understanding how heavy oil can be recovered through chemical injection is to consider the state of these
reservoirs at the time of injection. In heavy oil reservoirs, the oil that is un-recovered at the end of waterflooding was by-
passed due to the adverse mobility ratio between oil and water. As a result, this oil is still continuous and capable of flow,
although the flow rates will depend on the applied pressure gradients and the permeability of the rock. A simple reduction in
oil-water interfacial tension, similar to the mechanism proposed for conventional oil enhanced oil recovery, will not be
effective in displacing the oil. Rather, the injected chemical must somehow be improving the mobility ratio between the oil
and water, thus giving a more stable displacement of oil to the production wells.

Emulsions and Heavy Oil Recovery
Under conditions of high temperature (i.e. steam condensation)
27,28
or high shear rates
29
, emulsions may form in heavy oil
systems, even in the absence of surfactants and ultra-low interfacial tensions. Due to the high oil viscosity compared to
water, these surfactants will generally tend to be W/O emulsions. The viscosity of these emulsions is related to the
continuous (oil) phase viscosity and the emulsified water fraction
30
. In this manner, W/O emulsions are more viscous than
the constituent heavy oil. While these viscous emulsions can lead to significant problems for production and transport of
heavy oil, several researchers have investigated their application in improved heavy oil recovery. Dranchuk et al.
31
and
Farouq Ali et al.
22
both observed that when caustic (alkaline) solutions were injected into sand packs containing viscous
heavy oil, W/O emulsions would form and this corresponded to an improvement in the oil production rates and recovery
factor. It was proposed
22
that the mechanism responsible for the improved oil recovery is that the more viscous W/O
emulsion is able to displace the heavy oil in a more stable fashion.
Under certain conditions, O/W emulsions may also form with the addition of surfactant and/or alkali mixtures in the
water. These emulsions, having water as the continuous phase, are expected to be less viscous than the constituent oil.
Therefore, emulsification and entrainment of oil has been suggested as a recovery mechanism
15,25,26
, whereby oil would be
produced as a low viscosity O/W emulsion. This emulsification was observed in micro-model studies of AS injection into
heavy oil
25
, but no actual O/W emulsions were observed in the produced fluids from core floods. Therefore, the suggested
mechanism for flowing O/W emulsions is that they will be first dispersed and carried in the aqueous phase, and then will re-
coalesce into an oil bank
26
. In only a few select core flood studies
4,5,24
were the actual effects of entrainment and flow of
emulsions in porous observed and measured.
SPE 113993 3
When O/W (or W/O) emulsions form, another mechanism that can be responsible for improved oil recovery is that of
emulsification and entrapment of droplets
4,5,15,23,24
. Even under low interfacial tension conditions, these droplets may be
trapped due to capillary forces in porous media, and the pressure required to displace these droplets can be significant
20,23,24
.
If these emulsion droplets can plug off the water channels formed during viscous fingering, then improved sweep efficiency
of the reservoir can be achieved. Several authors
23,24
have noted that in viscous oil systems, improved sweep efficiency will
be a much more important effect than microscopic residual oil saturation reduction, as would be achieved in entrainment.
Jamaluddin and Butler
32
performed an analysis considering oil and water in a capillary tube of constant radius R. The net
work required to emulsify a droplet of water (i.e. to form a single W/O emulsion droplet) was found to be
32
:

= cos
3
2
r
R
R
3
P
ow
emul
(1)

Where P
emul
= the net work per unit volume of the water droplet

ow
= the oil-water interfacial tension
= the contact angle, measured through the water
r = the radius of the water droplet
R = the radius of the tube.

Equation 1 shows that if a water droplet was emulsified, the minimum net work requirement corresponds to the point where
the droplet size is a maximum (i.e. r = R). Therefore, emulsion droplets will most likely be in the same range as the pore
sizes. In this situation, discussion of the emulsion as low or high viscosity compared to the oil phase does not hold true.
Such definitions are only applicable in conditions where the emulsified droplets are much smaller than the radius of the
pores. Therefore, our previous work
5
has shown that emulsification can lead to improved heavy oil recovery, but that the
recovery mechanism is that entrapment and re-direction of flow, as was also previously discussed by other authors
22-24
. Since
improved oil recovery is due to plugging of water channels and improved sweep efficiencies, both W/O and O/W emulsions
can be generated in AS injection, and can be used to achieve improved oil recovery.

Wettability and Heavy Oil Recovery
Another consequence of alkali and/or surfactant injection into porous media is the potential for them to adsorb to the rock,
thus altering its wettability. Several works
18,33,34
provide an excellent overview of the mechanisms by which this can occur.
In general, surfactant adsorption will be enhanced in reservoirs where the rock and surfactant have opposite charges
18
,
however even with similarly charged surfactants and rocks some adsorption and wettability alteration may still occur
34
. In
conventional oil, altering the reservoir to an oil wet state can lead to oil film drainage and improved flow
16
, however this is
not the case in heavy oil systems. In these reservoirs, since oil recovery after water breakthrough is due to water
imbibition
2,3
, the reservoir is required to be water wet. If chemical injection then alters the wettability, oil will adhere to the
rock surfaces and further water injection after AS flooding may no longer be viable.
An understanding of reservoir wettability is fundamental to interpreting the response from a chemical flood. However,
this parameter is not easy to measure. Often the assumption is made that because sand is naturally hydrophilic, reservoirs
will also tend to be water wet
35
. Although evidence for water wetting can be found through theoretical calculations of film
stability
35,36
, there is actually very little experimental evidence that sand will always be water wet. Methods for monitoring
wettability and wettability alteration are needed in EOR studies.
If emulsification is responsible for improved oil recovery in heavy oil systems, then the effect of the rock wettability on
the emulsions formed should also be considered. Equation 1 shows that the net work required to emulsify a water droplet
into oil is related to the contact angle, . The net work to form the W/O emulsion is minimized for the situation where =
180C (i.e. the rock is oil wet). This indicates that W/O emulsions will more easily form in oil wet porous media. Likewise,
if the net work required to form an O/W emulsion droplet were calculated, this would also be minimized for a strongly water
wet rock. Equation 1 therefore indicates that the emulsion types that are formed may be dependent on the rock wettability.
Rock wettability inferred from contact angle is measured on flat solid surfaces through Youngs Equation
34,37
:

ws os ow
cos = (2)

Where
os
= the oil-solid interfacial free energy

ws
= the water-solid interfacial free energy.

Youngs equation expresses the nature of the solid to be wetted by either oil or water. The solid wettability is defined by the
difference in interfacial free energy of the solid between the two liquids (i.e.
os
-
ws
). Since these parameters cannot be
measured directly, Youngs Equation allows them to be determined based on easily measured parameters:
ow
and . In
chemical flooding systems, however, oil-water interfacial tension may be reduced by several orders of magnitude. In this
4 SPE 113993
situation, the right hand side of Equation 2 becomes very small and it becomes unclear whether the contact angle can still be
used as an indicator of wettability. Contact angle analysis also is not easy to translate for fluids in porous media, where the
pores are not flat surfaces. Therefore, in this study a new technique is applied (low field nuclear magnetic resonance) to
obtain an understanding of how wettability changes during the AS flood.

Materials and Experiments
Core floods were previously done in systems with varying parameters. Details of these floods are provided elsewhere
4,5
, but
a summary of the rock and fluid parameter is provided in Table 1. These floods were performed on unconsolidated sand
packs, ranging from 0.9 19 D in permeability, and for two oils of 11,500 and 15,000 mPas viscosity. All the oil was
obtained from a single field in Saskatchewan, Canada at different times. Differences in handling or possibly geological
heterogeneities could have been responsible for the different measured viscosity values.
Originally, the goal was to perform AS floods at rates that are representative or even slightly higher than field rates. The
reason for this is that emulsification is governed by the applied shear, and higher injection rates into any sand will lead
correspondingly to higher shear. In the data collected in this study, floods were performed using the 11,500 mPas oil and a
relatively low permeability sand pack. The average value of the permeability and porosity measured for the three
experiments are provided in Table 1. The injection rates studied were also considerably lower for this sand, which was a
reflection of the fact that in a lower permeability sand, high shear rates can be obtained even at lower injection rates.
The sand used for all the unconsolidated cores was Lane Mountain (LM 70) sand, which has an average particle sizes
ranging from 150 250 m. Due to the nature of some of these sand packs, overburden pressure could not be applied to the
sand, thus in order to obtain lower permeability sand a portion of the LM 70 (< 100 m) was sieved. This sieved portion
provided sand packs that had permeability varying around 1 D, as shown in Table 1. The sieved sand was used for the three
additional AS floods performed in this study.
The pre-formed surfactant used was a commercial anionic surfactant (Bio-Terge PAS-8S) that was supplied by Stepan
Company. This is a sodium alkane sulfonate, and was chosen because its negative head group should repel the negatively
charged sand grains, thus reducing surfactant loss due to adsorption
18
. The alkaline agent used was sodium carbonate,
Na
2
CO
3
, which was chosen over other alkali agents because of its buffering capabilities, and because it has been shown to be
less sensitive to variations in water chemistry than NaOH
38
. The alkali will react with organic acids in the oil phase and
should generate monovalent anionic surfactants. In this manner, both the preformed and in-situ surfactants were anionic, and
should not strongly adsorb to the sand grains.
The core floods completed in this study were performed on sieved sand that was packed into a core holder made of
TECAPEEK
TM
PVX, which is a high performance plastic that can be used for high-pressure applications. This material is
also NMR transparent, which allowed for measurements to be taken of the in-situ fluids. All NMR measurements were
obtained using an Ecotek FT low field NMR relaxometer, which operates at a frequency of around 1.2 MHz. NMR
measurements were obtained using an echo spacing of 0.16 ms, 5000 echoes, a re-polarization time of 5000 ms, and 16
trains. This relatively fast measurement was required in order to capture high-resolution spectra from fluids that are flowing
while being measured.
Fluids were injected at constant rates using a digital ISCO pump, and pressures were electronically logged into a data
acquisition system. The produced fluids were measured for water cut using NMR, and then were separated when possible,
using high ratios of toluene. All waterflood samples could be separated and used to tune the NMR predictions.
In previous work
4,5
bulk liquid studies were performed in order to identify the proper concentration of alkali to be used in
these floods. The results of the bulk liquid studies indicated that a composition of 0.1 wt% surfactant and 0.5 wt% alkali
would lead to ultra-low interfacial tensions and stable emulsions. This was therefore the composition used for all the cores
floods. In the bulk liquid study, it was observed
4,5
that when the chemical was present in de-ionized (DI) water, O/W
emulsions would form. However, when the aqueous phase consisted of water and 2 wt% NaCl (Brn), only viscous W/O
emulsions would form. Therefore, by injecting DI AS solutions, it was expected that O/W emulsification would occur,
whereas Brn AS injection should lead to W/O emulsions. In this study, this was verified both in the produced fluids and also
in the NMR spectra of the in-situ oil and water.

Oil Recovery through DI AS Injection at Various Rates
Figure 1 shows an example of the injection pressure profile and the oil production profile for a waterflood followed by an AS
flood. At early times in the waterflood, water was injected at a constant rate but viscous oil could not be produced at this
same rate, so pressure built up considerably in the core. Eventually water would finger through to the outlet, after which
point water could quickly travel through the water channels that had been formed, and pressure declined quickly. After
approximately five pore volumes of water had been injected, AS injection (0.1% surfactant and 0.5% Na
2
CO
3
in DI water)
was commenced at the same rate. Upon injection of AS solution, the core responded with pressure building considerably.
As pressure increased in the core, the fluid that was produced was mainly water. After pressure reached a maximum, the
produced water cuts decreased and the oil recovery improved.
Previous experience in the multiple cores flooded with AS solution
4,5
has shown that, in general, AS flooding does not
reduce the residual oil saturation down to very low levels. In fact, the optimal oil recoveries were around 70% of OOIP, and
SPE 113993 5
most floods actually recovered less oil. Under very low rate waterflooding for many pore volumes
3,39
, it is also possible to
achieve similar residual oil saturation by allowing water imbibition to slowly sweep oil from the core. Therefore, as also
observed previously by Jennings et al.
24
, the improved oil response is one of improved sweep and accelerated oil production.
Figure 2 shows the oil recovery profile for a longer range of pore volumes injected, and plots a dashed trend line to estimate
the waterflood recovery profile. Eventually the waterflood could recover a similar oil fraction, however this would take over
twenty pore volumes of injected fluid, which is a slow process and would also require significant water handling and
recycling. The use of the AS solution allows for this same oil recovery to be achieved much faster.
Previously
4,5
experience was obtained regarding the efficiency of AS floods under varying injection rates. At higher rates
or during injection into lower permeability porous media, the fluids are subjected to much higher ranges of shear. Under
higher shear, the emulsion droplets formed are smaller, and may actually be capable of flowing through the rock pores. In
these conditions, actual produced O/W emulsions have been measured
4,5
. When oil is produced as an O/W emulsion, the oil
fraction in the emulsion is relatively small. In contrast, under lower shear when oil droplets cannot be forced through the
rock pores, the oil plugs off the water channels and this leads to re-distribution of flow pathways and additional oil recovery.
The efficiency of the AS flood is expressed as the oil recovery per pore volume of AS fluid injected. In previous work
5
a
relationship was observed between the flood efficiency and the dimensionless shear rate of the flood:

P
L
k
o
dim

= (3)

Where
dim
= the dimensionless shear rate
= the frontal (Darcy) velocity of the AS flood

o
= the oil viscosity (mPas)
k = the absolute permeability of the sand
P/L = the maximum pressure gradient that is generated in the AS flood.

Figure 3 plots the efficiencies and dimensionless shear from the previous core floods
4,5
, along with the three AS floods
performed in this study. At the maximum injection rate, the fluids produced were a black, low viscosity O/W emulsion.
Correspondingly, the AS flood recovery efficiency was low. The high rate AS flood, even at the injection of 0.09 m/day,
appears to be in the range of the entrainment floods measured previously. By reducing the rate further, to 0.045 m/day, the
flood was performing the entrapment regime, and the AS flood efficiency improved as well.
It was also previously observed
5
that more efficient AS floods corresponding to a situation whereby the effective mobility
ratio between the AS solution and the viscous oil could be reduced. Figure 4 shows that as the AS injection rate is reduced,
this leads to floods that have lower effective mobility ratios, and will therefore be more efficient. The results from Figures 3
and 4 indicate that lower values of dimensionless shear correspond to lower effective mobility ratios and overall improved
flood performance. The results of these three additional core floods demonstrate that optimal improved oil recovery is due to
emulsification and entrapment of droplets, compared to entrainment and flow.

Wettability Changes During AS Flooding
In addition to evaluating the effect of reducing the AS injection rate on the dimensionless shear and the flood efficiency, a
secondary goal was to monitor the location of oil and water in-situ using low field NMR. For this reason, the core floods
were performed in the NMR-transparent PEEK core holder. Sand does not contain any hydrogen, so the NMR signal will
come only from the oil and water in the core. However, the fluid relaxation times will be governed by a combination of the
fluid properties (i.e. the viscosity) and the surface-to-volume ratio of the pore space in which the fluid is located
40,41
:

V
S
T
1
T
1
T
1
s
S 2 B 2 2
+ + = (4)

Where 1/T
2
= the measured relaxation rate
1/T
2B
= the bulk relaxation rate of a fluid
1/T
2S
= the surface relaxation rate of the fluid
= the fluid viscosity

s
= the surface relaxivity of the rock
S/V = the surface-to-volume ratio of the rock.

The total relaxation rate is the summation of the bulk and surface terms, as shown in Equation 4. The bulk relaxation rate
is proportional to the fluid viscosity: higher viscosity fluids will relax faster than low viscosity fluids. Surface relaxation is
proportional to the surface-to-volume ratio of the rock pores that contain the fluid being measured. Smaller pores have larger
values of S/V, which will in turn lead to faster surface relaxation. For heavy oil and water in porous media, two extreme
6 SPE 113993
conditions can be measured. First, for low viscosity fluids such as water, the bulk relaxation rate is considerably slower than
the surface relaxation, so the NMR T
2
distribution measured is analogous to a pore size distribution, where peaks at lower T
2

values correspond to smaller pores
40,41
. The other extreme is for high viscosity heavy oil, where the bulk relaxation occurs so
quickly that the relaxation times become essentially independent of the influence of the porous medium. In this case, oil
signals relax at roughly the same times whether they are measured in bulk or in porous media
42
.
Figure 5 shows the spectra obtained at four different length fractions in the core at the stage where it was fully saturated
with water. At this condition, the NMR peaks provide a measure of the average pore sizes in this sand pack. This sand pack
has a small fraction of pores that give a signal under 10 ms, but the majority of the signal is between 10 200 ms. All the
peaks are similar, indicating that the sand pack is relatively homogeneous with length, and the mean relaxation of the water-
saturated sand is approximately 53 ms.
When oil was injected at a fixed rate into the sand pack, the signal in the T
2
range between 10 200 ms mostly
disappeared, as can be seen in Figure 6. This is an indication that oil had displaced water out of the largest pores, or that the
displacement was a drainage process. This points to the water-wet nature of the sand
43
. The small remaining signal in the
original range of T
2
values are water-filled pores that were by-passed by the oil. The oil itself is a viscous fluid, therefore
even though the oil is in the large pores (as evidenced by the disappearance of the water signal at later T
2
values) the oil still
appears as a fast-relaxing peak under 10 ms.
The spectra for the core after being waterflooded for five pore volumes are shown in Figure 7. Several observations can
be made from comparing these spectra to the water-saturated spectra in Figure 5 and the spectra of the core saturated with oil
and irreducible water in Figure 6. First, the amplitude in the first peak (under 10 ms) has decreased. From Figure 6, a large
component of this first peak was oil, therefore a lower amplitude is equivalent to a lower oil saturation, which is expected
from the production data. Additionally, amplitude peaks have grown in the T
2
range between 10 200 ms. This is water that
is in the same range of the original pore sizes, indicating that some water has displaced oil completely out of some rock
pores. Finally, there is some additional signal that occurs at later T
2
values, after 200 ms. This is a slow relaxing fluid, so it
must also be water, but somehow the water is relaxing more slowly than the signal in the core that was saturated with water
(Figure 5). The water is still present in the porous media, so physically the water cannot be located in anything larger than
the size of a pore. The only possible mechanism that could be responsible for this slow-relaxing water is that the water is
present as a W/O emulsion, whereby the surface relaxivity,
s
, is lower for the oil/water interface than it is for the sand/water
interface. Emulsion droplets are likely to be in the same range of the pore sizes, as discussed previously in Equation 1.
However, the lower surface relaxivity will lead to longer relaxation times for the emulsified water.
Figure 7 shows that water can form emulsions, even in the absence of any injected surfactant. This could be partially due
to the release of natural surfactants from the heavy oil, or could be due to the high shear encountered during the significant
pressure gradient that was present prior to water breakthrough. Despite the fact that some W/O emulsions were shown to
have formed, and these are often indicative of an oil wet condition (Equation 1), the total water geometric mean T
2
values
(T
2gm
) plotted in Figure 8 are still lower than the T
2gm
of the original water-saturated core. This indicates that the rock was
still water wet, and was verified through oil recovery via water imbibition
39
.
Figure 9 shows the NMR spectra at the same four length locations after flooding one pore volume of DI AS solution
through the core at a relatively high rate (0.51 m/day). At this velocity, one pore volume of AS solution was flooded in
approximately 11 hrs, so the core was in contact with AS solution for less than one day. At this high injection velocity, the
core was subjected to a high value of dimensionless shear, and the fluid that was produced was a black, low viscosity O/W
emulsion. These emulsions were actually observed, so they must also have formed and flowed within the rock pores. Figure
9 shows, however, that a higher degree of W/O emulsification is also observed to have occurred. The NMR is therefore
providing evidence for the opposite emulsion type as what was observed, which indicates that during chemical flooding the
emulsion systems that are present may be quite complicated. Bulk liquid studies
4,5
and the knowledge of the nature of the
anionic surfactants, predicted that O/W emulsions should have formed, and these were indeed observed. However, the NMR
also provides evidence for additional W/O emulsions that formed as well. In Figure 9, the W/O emulsion peak is larger for
the first half of the core (length fractions of 0.09 and 0.35), indicating that the W/O emulsions formed more in the inlet half
of the core.
After flooding with one pore volume of DI AS solution, a 2% NaCl salinity buffer was injected, followed by a Brn AS
flood at the same high rate. Figure 10 shows the NMR spectra at the same four length locations after the conclusion of the
Brn AS flood. The W/O emulsion peak is large and consistent all along the length of the core. As expected, Brn AS solution
led to the formation of W/O emulsions. In the produced fluids, no more O/W emulsions were observed so this matched the
experience gained in the bulk liquid systems.
The results of the water T
2gm
values at different flooding conditions are summarized in Figure 11. The water-saturated
core has water T
2gm
values constantly close to 50 ms. The values of the water T
2gm
at S
wi
and after waterflooding are similar.
Significant differences are observed after injection of AS solution. During the DI AS flood, due to higher fraction of W/O
emulsions at the inlet of the core, the water T
2gm
values at these locations are higher than the values of the core fully saturated
with water. The fact that the overall water T
2gm
value is higher than the core when it was saturated just with water is a
possible indication of wettability alteration to a more oil wet state at the inlet of the core. In the latter half of the core the
water T
2gm
values are similar to the waterflood data. After Brn AS injection, where significant W/O emulsions formed all
along the core, the water T
2gm
values are high all along the length.
SPE 113993 7
The pressure response of the core to DI AS and Brn AS flooding are shown in Figure 12. DI AS injection began at the
first vertical line, after approximately five pore volumes of water had already been injected. After approximately six pore
volumes of injection, another vertical line indicates the point where the DI AS flood ended, and the salinity buffer was
injected. Finally, after approximately seven pore volumes of fluid, the third vertical line indicates the start of the Brn AS
flood. The pressure response in the system can possibly be related to the emulsions shown in Figure 12. During the DI AS
flood, W/O emulsions formed in the inlet half of the core, but not close to the outlet. This was shown in the NMR spectra,
where the emulsified water signal was more prominent in the first two length fractions measured. Correspondingly, pressure
builds up and then declines again during the flood. It is possible that while these W/O emulsions were forming, this was
leading to improved sweep efficiency of the core, as water was re-directed into new pathways and once the water broke
through once more, pressure declined. O/W emulsions could also possibly be leading to plugging of the water channels,
however the fact that these emulsions were observed in the produced fluid indicates that in this core, the O/W emulsions
seemed to be too small to generate entrapment and improved sweep. Conversely, during the Brn AS flood W/O emulsions
formed all along the length of the core. The pressure correspondingly stays high in this system, indicating that over time
W/O emulsions were forming at different parts of the core, and keeping the pressure gradient high. Therefore, it appears that
the pressure drops that have been observed in previous studies in DI AS flooding
4,25,26
may be due to the formation of W/O
emulsions even in the absence of salt.
The results from Figure 9 12 indicate that W/O emulsions are constantly being formed during chemical flooding, and
these may be at least partially responsible for the pressure response measured during these floods. Equation 1 showed that
W/O emulsions will preferentially form in oil wet media, however the analysis of the waterflooding spectra showed that the
rock remained water wet, even though a small degree of W/O emulsification was observed. Proper understanding of the
emulsion response and how it related to improved oil recovery also involves an analysis of the rock wettability, specifically
determining if during the AS flooding the rock became more oil wet over time. Surfactant adsorption is also a time-
dependent process
18
, thus in slower AS flooding or in field conditions where the rock is in contact with surfactant for a much
longer time, the effect of any possible wettability alteration will be more severe. Therefore, more detailed analyses were
performed for the lowest rate (0.045 m/day) DI AS flood. At this injection rate, one pore volume of fluid injection required
approximately 120 hrs, or five days. Therefore, longer times were provided for surfactant adsorption and possible wettability
alteration.
The interpretation of wettability from NMR spectra has been developed previously in the analysis of imbibition during
heavy oil waterflooding
39,43
. Figure 13 shows the range of T
2
values for water located in various different physical locations
in the porous media. Surface water is water in the same range of T
2
values as the original water-saturated signal, so this is
water in pores that are governed by surface relaxation. W/O emulsions were previously shown as water that relaxes more
slowly than the surface water. Finally, water signal may also be present at early T
2
values, under 10 ms. Figure 5 shows that
there is very little water signal corresponding to such small pores. However, in past waterflooding studies
39,43
it was
observed that if only the signal after the first peak was attributed to water, and all the signal within the first peak was
attributed to oil, the NMR would consistently under-estimate the water saturations and the oil recovery that were measured in
the produced fluids. The only way that NMR could match the production data was to assign some of the amplitude under 10
ms to water. Since there are no small pores physically present to result in such fast-relaxing water signal, this water must be
located in regions of very high surface-to-volume ratio: this is therefore the signal from water pendular rings between sand
grains, and thin films of water covering the sand. For simplicity, this water is all referred to as films in Figure 13.
There is no way to determine a priori how much of the signal under 10 ms is due to this constrained water amplitude.
Therefore, the water signal in films was calculated as the difference between the actual water saturation (measured from mass
balance of the produced fluids) and the water saturations predicted by the NMR signal after 10 ms. This methodology was
developed for the interpretation of imbibition in heavy oil waterflooding
39
and has been applied in this work to the analysis of
the AS floods. Figure 14 shows the water saturation in films, pores (surface) and W/O emulsions as a function of the pore
volumes of DI AS solution injected. These values were calculated as the average over the four length locations measured at
each time.
During a low rate heavy oil waterflood, where imbibition was responsible for oil recovery, the water saturation in the
films grew over time
39,43
. Water film thickening was evidence not only of the water wet nature of the rock, but also was part
of the imbibition process. In Figure 14, it is evident that as DI AS fluid was injected and pressure built up in the core, the
surface water decreased and the W/O emulsion correspondingly increased. Therefore, water that existed in the pores was
emulsified into oil. Further insight into the wettability of the system was gained through an examination of the film water
saturation in Figure 14. In direct contrast to the observations made during waterflooding
39,43
, the water film saturation
remained constant through the flood. Therefore, none of the injected fluid was traveling through water films.
Figure 15 plots the total water T
2gm
over the time of the DI AS flood. Again, in contrast to Figure 8, the water relaxation
times became much higher than that of the original water-saturated core. This is an indication that water exists mainly in an
emulsified state in the pores. W/O emulsions were present all along the length of the core, similarly to the state of the core
after flooding at a high rate with a Brn AS solution, which had been designed to yield W/O emulsification. The high water
T
2gm
values, compared to the values during waterflooding, are evidence of wettability alteration to a more oil wet state during
the DI AS flood. Higher T
2gm
values are indicative of more water existing in an emulsified state, instead of being in contact
with the pore walls. More evidence of this is provided in Figure 16, which plots the water saturation in films for each of the
8 SPE 113993
four length locations during the DI AS flood. Although from the average of these values it appeared that the water film
saturation was a constant, in fact the water film saturation dropped abruptly for the length fractions of 0.09 and 0.35. This is
again evidence of stronger wettability alteration at the inlet of the core. Conversely, the water film saturation actually grew
in the latter half of the core, which indicated that as wettability was altered near to the inlet of the core, water was displaced
through films to the outlet. Therefore, the outlet actually became more water wet at early times. Once again, the pressure
buildup observed appears to be at least partially due to plugging of water channels through wettability reversal from water in
channels changing into discontinuous W/O emulsions. At later times, as W/O emulsions were also formed in the latter half
of the sand pack, the water film saturation decreased at these length locations as well.
The result from Figure 16 provides evidence that even in systems with similarly charged mineral surfaces and polar
surfactant head groups, wettability alteration can still occur. Rosen
34
suggested several mechanisms by which this could be
possible; however in general wettability alteration of sand using anionic surfactants is expected to be minimal. The fact that
this was not the case in the low rate DI AS flood speaks to the nature of the surfactants in the system. Surfactants generated
from the reaction of heavy oils with alkaline solution will tend to have relatively long, or oil soluble, hydrocarbon chains.
Even with low levels of surfactant adsorption onto similarly charged sand, the surfactant tails must be highly attracted to the
hydrocarbon, which will lead to significant oil deposition on the rock surface.
At the end of the DI AS flood, therefore, if the rock is already water wet and W/O emulsions have already formed, then
there will be no benefit in performing additional Brn AS flooding or waterflooding. Brn AS floods would be designed to
generate an oil wet core with a higher degree of W/O emulsions, but this situation is already present after a low rate DI AS
flood. Likewise, if oil is recovered due to water imbibition in heavy oil waterflooding, this mechanism will no longer be
present in an oil wet core, or a core with low interfacial tensions. In order to investigate this further, a waterflood was
performed after the DI AS flood. The recovery and pressure profile is shown in Figure 17, and the water saturations in
emulsions, pores and films are shown in Figure 18.
At first, the core still contained surfactant solution, so the actual waterflood response could only be determined after the
surfactant was flushed out of the system. In order to expel the surfactant quickly, DI water was injected at a relatively fast
frontal velocity (0.27 m/day) for almost one pore volume. During this time, pressure built up considerably in the low
permeability core, and this led to high values of dimensionless shear. Accordingly, oil was produced as an O/W emulsion,
and this led to the production of an additional 2% OOIP. This is region A in Figure 17. At the first dashed line, the
waterflood rate was reduced back to 0.045 m/day. At this time, the pressure in the core was still high, but it quickly declined.
While pressure in the core was elevated (and there was still some remaining surfactant), an additional 2% OOIP was
produced. After approximately two pore volumes of water had been injected, the system pressure was lower and the
surfactant had been flushed out of the core. At this point, the waterflood response could be determined.
Over the next 2.5 pore volumes of water injection, only an additional 1% of oil was recovered. Therefore, the waterflood
was considerably less efficient than before the DI AS flood, again pointing to a more oil wet condition. After close to five
pore volumes of water injection, a surfactant flood (0.1 wt% Bio-Terge PAS-8S) was performed, still at the same injection
velocity. The surfactant alone was not able to yield low interfacial tensions, but in the bulk liquid studies it was seen to make
the glass more water wet
4,5
, therefore the surfactant flood was performed in an attempt to restore the original wettability of
the rock. The surfactant flood is region C in Figure 17. Finally, at the third dashed line, the core was shut in for a period of
two weeks, and then Brn AS injection commenced at the same rate. At this point, pressure fell in the model, so the Brn AS
solution was not successful in recovering any additional heavy oil.
When the water saturation in different locations (emulsions, pores and films) are considered, a better understanding of the
flood response can be achieved. During the waterflood, the W/O emulsion saturation was slowly decreasing, and the surface
(pore) water saturation was increasing. Therefore, the low rate waterflood was breaking the emulsions and re-forming the
water channels, even though this did not lead to any additional oil recovery. Upon injection of 0.1% surfactant solution, the
same trend was observed. However, after shutting in the core and then re-commencing injection, a startling difference was
observed. The surface water saturation increased sharply, and the emulsion saturation decreased down to zero. Likewise, an
increase in the water film saturation was also measured. Figure 19 shows the spectra at the four length locations during this
final stage of the flood. The W/O emulsion peak has completely disappeared, and all the water is in the film and surface
locations.
It is not known at this point why the injection of the Brn AS solution did not then revert wettability back to oil wet
conditions. Instead, this solution seemed to simply cycle through the water channels, and essentially no additional oil was
recovered. However, the results of Figures 17 19 clearly indicate that with surfactant injection and soaking, or potentially
with prolonged low rate waterflooding, the core can be restored back to its water wet state again. Therefore, additional
waterflooding recovery may be possible, and will be the focus of future investigations.

Conclusions
In this work, a study was performed whereby de-ionized alkali-surfactant fluid was injected into an unconsolidated sand pack
at multiple rates. As expected, as the injection rate was decreased, this led to a higher degree of plugging in the water
channels, which translated to improved chemical flooding efficiency and higher oil recovery. Therefore, this work
demonstrates that low rate chemical flooding, which utilizes the mechanism of emulsification and entrapment, can be
successful in heavy oil systems.
SPE 113993 9
NMR spectra were acquired to investigate the nature of the emulsions formed. It was observed that even under DI AS
injection, where O/W emulsification was expected and was in fact even observed at higher shear rates, W/O emulsions were
also forming in the rock pores. Under high rate AS injection, these emulsions were only prevalent at the inlet of the core, but
under Brn AS injection W/O emulsions formed everywhere. Under low rate emulsion flooding, where significantly more
time was provided for surfactant adsorption to occur, water in the rock pores was converted to emulsified water all along the
length of the core. The pressure response and plugging of water channels is at least in part due to the formation of these W/O
emulsions.
But analyzing the water saturation in films (i.e. water signal relaxing in regions that are smaller than the pores) it was seen
that the core wettability was altered to an oil wet state during the course of the waterflood. Although wettability alteration
occurred at different times at the inlet vs. the outlet of the core, it was evident that increased W/O emulsification was indeed
related to more oil wet conditions. By later flooding the core with low rate water and a low concentration of water-soluble
anionic surfactant, the original water wet condition of the core can be recovered.

Acknowledgements
The authors wish to acknowledge the technical contributions of Xiao Dong Ji, Jun Gao and Michael Erath of TIPM
Laboratory in helping to set up the experimental systems in this study. The oil used in these core floods was generously
supplied by Nexen Inc., and the surfactant was donated by Stepan Company. Financial support for this project was received
from NSERC, COURSE, ISEEE and the Canada Research Chair in Energy and Imaging and its Industrial Affiliates (Shell,
Nexen, Devon, PetroCanada, Canadian Natural, ET Energy, Suncor, Schlumberger, CMG, Laricina, Paramount and
ConocoPhillips).

Nomenclature
AS = alkali-surfactant solution
Brn = brine
DI = de-ionized water
K = the absolute permeability of the sand
L = the length of the sand pack
O/W = an oil-in-water emulsion
P
emul
= the net work per unit volume of the water droplet
r = the radius of the water droplet
R = the radius of the tube.
S/V = the surface-to-volume ratio of the rock
T
2
= the measured relaxation time
T
2B
= the bulk relaxation time of a fluid
T
2gm
= the geometric mean relaxation time of a fluid
T
2S
= the surface relaxation rate of the fluid
W/O = a water-in-oil emulsion

dim
= the dimensionless shear rate
= the contact angle, measured through the water
P = the maximum pressure drop that is generated in the AS flood

o
= the oil viscosity (mPas)

s
= the surface relaxivity of the rock

os
= the oil-solid interfacial free energy

ow
= the oil-water interfacial tension

ws
= the water-solid interfacial free energy
= the frontal (darcy) velocity of the AS flood

References
1. Alberta Energy and Utilities Board ST98-2007: Albertas Energy Reserves 2006 and Supply/Demand Outlook 2007-2016, AEUB ISSN
1910-4235, Farhood Rahnama, coord., Alberta Energy and Utilities Board, Calgary, AB Canada, Jun 2007.
2. Mai, A., Bryan, J., Goodarzi, N. and Kantzas, A.: Insights Into Non-Thermal Recovery of Heavy Oil, WHOC 2006-533, 1
st
World
Heavy Oil Conference, Beijing China, Nov 12 15 2006.
3. Mai, A. and Kantzas, A.: Heavy Oil Waterflooding: Effects of Flow Rates and Oil Viscosity, PS 2007-144, 58
th
Annual Technical
Meeting of the Petroleum Society, Calgary, AB Canada, Jun 12 14 2007.
4. Bryan, J. and Kantzas, A.: Potential for Alkali-Surfactant Flooding in Heavy Oil Reservoirs Through Oil-in-Water Emulsification, PS
2007-134, 58
th
Annual Technical Meeting of the Petroleum Society, Calgary, AB Canada, Jun 12 14 2007.
5. Bryan, J. and Katnzas, A.: Enhanced Heavy Oil Recovery by Alkali-Surfactant Flooding, SPE 110738, SPE Annual Technical
Conference and Exhibition, Anaheim, CA USA, Nov 11 14, 2007.
10 SPE 113993
6. Jennings, H.Y.: Waterflood Behaviour of High Viscosity Crudes in Preserved Soft and Unconsolidated Cores, SPE 1202, J. Pet.
Tech., 116 120, Jan 1966.
7. Kumar, M., Hoang, V. and Satik, C.: High Mobility Ratio Waterflood Performance Prediction: Challenges and New Insights, SPE
97671, SPE International Improved Oil Recovery Conference, Kuala Lumpur, Malaysia, Dec 5 6, 2005.
8. Miller, K.A.: Improving the State of the Art of Western Canadian Heavy Oil Waterflood Technology, J. Can. Pet. Tech., 45 (4), 7
11, 2006.
9. Peters, E.J. and Flock, D.L.: The Onset of Instability During Two-Phase Immiscible Displacement in Porous Media, SPE Journal,
248 258, Apr 1981.
10. Bentsen, R.G.: A New Approach to Instability Theory in Porous Media, SPE Journal, 765 779, Oct 1985.
11. Davis, H.T. and Scriven, L.E.: Stress and Structure in Fluid Interfaces, Adv. Chem. Phys., 49, 357 454, 1982.
12. Green, D.W. and Willhite, G.P.: Enhanced Oil Recovery, SPE Textbook Series Vol. 6, Society of Petroleum Engineers Inc., Richardson,
TX USA, 1998.
13. Nutting, P.G.: Chemical Problems in the Water-Driving of Petroleum Reservoirs, Ind. And Eng. Chem., 17, Oct 1925.
14. Moore, T.F. and Slobod, R.L.: The Effect of Viscosity and Capillarity on the Displacement of Oil by Water, Prod. Monthly, 20 30,
Aug 1956.
15. Johnson, C.E. Jr.: Status of Caustic and Emulsion Methods, SPE 5561, J. Pet. Tech., 85 92, January 1976.
16. Cooke, C.E. Jr., Williams, R.E. and Kolodzie, P.A.: Oil Recovery by Alkaline Waterflooding, J. Pet. Tech., 16, December 1974.
17. Farouq Ali, S.M. and Thomas, S.: Field Experience with Chemical Oil Recovery Methods, CIM 94-055, 45
th
Annual Technical
Meeting of the Petroleum Society of CIM, Calgary, AB CANADA, June 12 15, 1994.
18. Wesson, L.L. and Harwell, J.H.: Surfactant Adsorption in Porous Media, Surfactants: Fundamentals and Applications in the
Petroleum Industry, L.L. Schramm ed., Cambridge University Press, Cambridge, UK, 121 158, 2000.
19. Scott, G.R., Collins, H.N. and Flock, D.L.: Improving Waterflood Recovery of Viscous Crude Oils by Chemical Control, J. Can. Pet.
Tech., 4 (10), October-December, 1965.
20. Pitts, M.J., Wyatt, K., and Surkalo, H.: Alkaline-Polymer Flooding of the David Pool, Lloydminster Alberta, SPE 89386, 2004
SPE/DOE 14
th
Symposium on Improved Oil Recovery, Tulsa, Oklahoma, USA, April 17 21, 2004.
21. Dranchuk, P.M., Scott, J.D. and Flock, D.L.: Effect of the Addition of Certain Chemicals On Oil Recovery During Waterflooding, J.
Can. Pet. Tech., Jul Sept 1974.
22. Farouq Ali, S.M., Figueroa, J.M., Azuaje, E.A., and Farquharson, R.G.: Recovery of Lloydminster and Morichal crudes by caustic,
acid and emulsion floods, J. Can. Pet. Tech., 53 59, Jan Mar 1979.
23. McAuliffe, C.D.: Oil-in-Water Emulsions and Their Flow Properties in Porous Media, J. Pet. Tech., 727-733, Jun 1973.
24. Jennings, H.Y. Jr., Johnson, C.E. Jr. and McAuliffe, C.D.: A Caustic Waterflooding Process for Heavy Oils, SPE 4741, J. Pet. Tech.,
1344-1352, December 1974.
25. Liu, Q., Dong, M. and Ma, S.: Alkaline/Surfactant Flood Potential in Western Canadian Heavy Oil Reservoirs, SPE 99791, 2006
SPE/DOE Symposium on Improved Oil Recovery, Tulsa, OK USA, Apr 22 26, 2006.
26. Liu, Q., Dong, M., Ma, S. and Tu, Y.: Surfactant enhanced alkaline flooding for Western Canadian heavy oil recovery, Coll. & Surf.
A: Physicochem. Eng. Aspects, 293, 63 71, 2007.
27. Chung, K.H. and Butler, R.M.: In situ Emulsification by the Condensation of Steam in Contact With Bitumen, J. Can. Pet. Tech., 28
(1), 48 55, Jan Feb 1989.
28. Vittoratos, E.: Flow Regimes During Cyclic Steam Stimulation at Cold Lake, J. Can. Pet. Tech., 30 (1), 82 86, Jan-Feb 1991.
29. Cuthiell, D., Green, K., Chow, R., Kissel, G., and McCarthy, C.: The In Situ Formation of Heavy Oil Emulsions, SPE 30319, 1995
International Heavy Oil Symposium, Calgary, AB Canada, June 19 21, 1995.
30. Pal, R., Yan, Y and Masliyah, J.: Rheology of Emulsions, Emulsions Fundamentals and Applications in the Petroleum Industry, L.L.
Shcramm ed., Advances in Chemistry Series 231, 131 170, American Chemical Society, 1992.
31. Dranchuk, P.M., Scott, J.D. and Flock, D.L.: Effect of the Addition of Certain Chemicals On Oil Recovery During Waterflooding, J.
Can. Pet. Tech., Jul Sept 1974.
32. Jamaluddin, A.K.M. and Butler, R.M.: Factors Affecting the Formation of Water-in-Oil Emulsions During Thermal Recovery,
AOSTRA J. Research., 4 (2), 109 116, 1988.
33. Spinler, E.A. and Baldwin, B.A.: Surfactant Induced Wettability Alteration in Porous Media, Surfactants: Fundamentals and
Applications in the Petroleum Industry, L.L. Schramm, ed., Cambridge University Press, Cambridge, UK, 159 202, 2000.
34. Rosen, M.J.: Surfactants and Interfacial Phenomena, 2
nd
Edition, John Wiley & Sons, Inc., New York, NY USA, 1989.
35. Czarnecki, J., Radoev, B., Schramm, L.L. and Slavchev, R.: On the nature of Athabasca Oil Sands, Adv. Coll. Int. Sci., 114 115, 53
60, 2005.
36. Takamura, K.: Microscopic Structure of Athabasca Oil Sand, Can. J. Chem. Eng., 60, pg. 538 545, Aug 1982.
37. Wagner, O.R and Leach, R.O.: Improving Oil Displacement Efficiency by Wettability Adjustment, SPE 1101-G, Pet. Trans. AIME,
216, 65 72, 1959.
38. Martin, F.D. and Oxley, J.C.: Effect of Various Alkaline Chemicals on Phase Behavior of Surfactant/Brine/Oil Mixtures, SPE Paper
13575, International Symposium on Oilfield and Geothermal Chemistry, Pheonix, AZ USA, Apr 9 11, 1985.
39. Mai, A.: Mechanisms of heavy Oil Recovery by Waterflooding, PhD Thesis, Department of Chemical & Petroleum Engineering,
University of Calgary, Calgary, AB Canada, 2008.
40. Straley, C., Rossini, D., Vinegar, H., Tutunjian, P. and Morriss, C.: Core Analysis by Low Field NMR, The Log Analyst, 38 (2), 1997.
41. Coates, G., Xiao, L. and Prammer, M.: NMR Logging Principles and Applications, Halliburton Energy Services, 1999.
42. Bryan, J., Kantzas, A., Badry, R., Emmerson, J. and Hancsicsak, T.: In Situ Viscosity of Heavy Oil: Core and Log Calibrations, J.
Can. Pet. Tech., 46 (11), 47 55, Nov 2007.
43. Mai, A., Bryan, J. and Kantzas, A.: Investigation into the Mechanisms of Heavy Oil Recovery by Waterflooding and Alkali-Surfactant
Flooding, SCA 2007-24, International Symposium of the Society of Core Analysts, Calgary, AB Canada, Sept 10 12, 2007.

SPE 113993 11

Table 1: Properties of Core Systems Used in These
Experiments
System studied Rock/fluid
Property
Value
Porosity (fraction) 0.45 0.49
Data from
previous core
floods
Permeability (D) 0.92 19.22
Oil viscosity
(mPas)
11,000/15,000
AS injection rates
(m/day)
0.26 2.05
Porosity (fraction) 0.41
Core floods
added in this
work
Permeability (D) 0.77
Oil viscosity
(mPas)
11,000
AS injection rates
(m/day)
0.51, 0.09,
0.045


Table 2: Reduction in Dimensionless Shear and Minimum
Effective Mobility Ratio through Rate Reduction
AS flood velocity
(m/day)

dim
Minimum M
eff

0.51 21.96 12.21
0.09 17.75 6.60
0.045 2.82 3.07

0
1000
2000
3000
4000
5000
6000
7000
8000
0 2 4 6 8
PV Injected
P
r
e
s
s
u
r
e

(
k
P
a
)
0
0.1
0.2
0.3
0.4
0.5
0.6
O
i
l

R
e
c
o
v
e
r
y

(
f
r
a
c
t
i
o
n
)
Pressure
Oil RF

Figure 1: Pressure and recovery profile for a waterflood
followed by an AS flood

0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0 5 10 15 20 25 30
PV Injected
O
i
l

R
e
c
o
v
e
r
y

(
f
r
a
c
t
i
o
n
)

Figure 2: Enhancement in oil recovery rate through AS
flooding

0.01
0.1
1
1 10 100
Dimensionless Shear Rate
T
o
t
a
l

O
i
l

R
e
c
o
v
e
r
y
/
P
V

I
n
j
e
c
t
e
d
Secondary AS
Entrain
0.51 m/day
0.09 m/day
0.045 m/day

Figure 3: Relationship between dimensionless shear rate
and AS flood recovery efficiency

0
0.05
0.1
0.15
0.2
0.25
0.3
0.01 0.1 1 10 100 1000
Minimum Effective Mobility Ratio
T
o
t
a
l

O
i
l

R
F
/
P
V

I
n
j
Secondary AS Floods
0.51 m/day
0.09 m/day
0.045 m/day

Figure 4: Relationship between the minimum effective
mobility ratio and AS flood recovery efficiency

0
0.1
0.2
0.3
0.4
0.5
0.1 1 10 100 1000 10000
Relaxation Time T2 (ms)
A
m
p
l
i
t
u
d
e
Length = 0.09
Length = 0.35
Length = 0.65
Length = 0.78

Figure 5: NMR spectra of the sand pack at four locations,
fully saturated with water

0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.1 1 10 100 1000 10000
Relaxation Time T2 (ms)
A
m
p
l
i
t
u
d
e
Length = 0.09
Length = 0.35
Length = 0.65
Length = 0.78

Figure 6: NMR spectra of the sand pack at four locations,
flooded to oil and irreducible water

12 SPE 113993
0
0.02
0.04
0.06
0.08
0.1
0.12
0.1 1 10 100 1000 10000
Relaxation Time T2 (ms)
A
m
p
l
i
t
u
d
e
Length = 0.09
Length = 0.35
Length = 0.65
Length = 0.78

Figure 7: NMR spectra after constant rate waterflooding for
five pore volumes

10
20
30
40
50
60
0 1 2 3 4 5 6
PV injected
T
o
t
a
l

w
a
t
e
r

T
2
g
m

(
m
s
)

Figure 8: Total water relaxation times during a constant low
rate waterflood

0
0.02
0.04
0.06
0.08
0.1
0.12
0.1 1 10 100 1000 10000
Relaxation Time T2 (ms)
A
m
p
l
i
t
u
d
e
Length = 0.09
Length = 0.35
Length = 0.65
Length = 0.78

Figure 9: NMR spectra after flooding with DI AS solution at
0.51 m/day injection velocity

0
0.02
0.04
0.06
0.08
0.1
0.1 1 10 100 1000 10000
Relaxation Time T2 (ms)
A
m
p
l
i
t
u
d
e
Length = 0.09
Length = 0.35
Length = 0.65
Length = 0.78

Figure 10: NMR spectra after flooding with Brn AS solution
at 0.51 m/day injection velocity

1
51
101
151
201
0 0.2 0.4 0.6 0.8 1
Length fraction
W
a
t
e
r

T
2
g
m

(
m
s
)
Sw = 100% Sw = Swi Post WF
Post DI AS Post Brn AS

Figure 11: Water NMR relaxation times after different
flooding states

0
2000
4000
6000
8000
10000
12000
14000
0 2 4 6 8 10
PV Injected
P
r
e
s
s
u
r
e

(
k
P
a
)
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
O
i
l

R
e
c
o
v
e
r
y

F
a
c
t
o
r
Pressure
Oil RF

Figure 12: Pressure and production profile for waterflood, DI
AS flood and Brn AS flood

0
0.02
0.04
0.06
0.08
0.1
0.1 1 10 100 1000 10000
Relaxation Time T2 (ms)
A
m
p
l
i
t
u
d
e
0.09 0.35 0.65 0.78
Film
Surface
W/O emulsion

Figure 13: Analysis of different water relaxation times in
waterflood or AS flood spectra

0
0.1
0.2
0.3
0.4
0.5
0.6
0 0.5 1 1.5 2
PV Injected
S
a
t
u
r
a
t
i
o
n
0
500
1000
1500
2000
2500
3000
P
r
e
s
s
u
r
e

(
k
P
a
)
Surface water W/O emulsion Film water Pressure

Figure 14: Change in water saturation in different locations
during the low rate DI AS flood

SPE 113993 13
0
20
40
60
80
100
120
140
160
0 0.5 1 1.5 2
PV Injected
W
a
t
e
r

T
2
g
m

(
m
s
)

Figure 15: Total water relaxation times during the low rate DI
AS flood

0
0.05
0.1
0.15
0.2
0.25
0.3
0 0.5 1 1.5 2
PV Injected
W
a
t
e
r

F
i
l
m

S
a
t
u
r
a
t
i
o
n
0
500
1000
1500
2000
2500
3000
P
r
e
s
s
u
r
e

(
k
P
a
)
0.09 0.35 0.65 0.78 Pressure

Figure 16: Water film saturation at different length locations
during the low rate DI AS flood

1
10
100
1000
10000
0 1 2 3 4 5 6 7 8
PV Injected
P
r
e
s
s
u
r
e

(
k
P
a
)
0.5
0.52
0.54
0.56
0.58
0.6
O
i
l

R
e
c
o
v
e
r
y

F
a
c
t
o
r
Pressure
Oil RF
A
B
C D

Figure 17: Oil pressure and recovery profile for post DI AS
waterflooding

0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
0 1 2 3 4 5 6 7 8
PV Injected
S
a
t
u
r
a
t
i
o
n
1
10
100
1000
10000
P
r
e
s
s
u
r
e

(
k
P
a
)
Surface water W/O water Film water Pressure

Figure 18: Water saturation in different locations during
waterflooding and surfactant flooding after the DI AS flood

0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0.1 1 10 100 1000 10000
Relaxation Time T2 (ms)
A
m
p
l
i
t
u
d
e
0.09
0.35
0.65
0.78

Figure 19: Water spectra at different lengths after soaking in
low concentration anionic surfactant

You might also like