You are on page 1of 57

“Numerical simulation of low salinity/smart waterflooding in

carbonate reservoirs”

EG5911 Individual Project in Petroleum Engineering

By

Museyib Huseynov, B.Sc. Petroleum Engineering

51771179

A dissertation submitted in partial fulfilment of the requirements of the award of Master


of Science in Petroleum Engineering at the University of Aberdeen

(August, 2018)
Acknowledgements

I would like to emphasize my endless gratitude to my supervisor, Dr Prashant


Jadhawar, for his great assistance and patience to answer all my questions. I was lucky
enough to have an opportunity to work with such a knowledgeable person.
Abstract

Although the structure of the carbonate reservoirs is very complex due to


heterogeneity, more than half of the total hydrocarbon resources are contained in
carbonate reservoirs. Therefore, it is very important to probe the methods that can increase
the recovery factor through those complex carbonate reservoirs. Injection of low salinity
water (LSWI) into the formation has been very popular method for the last two decades,
which can increase the hydrocarbon recovery from the reservoirs. Although a lot of
experiments have been conducted both in laboratory and field scale to comprehend the
underlying mechanism that leads to incremental oil recovery, no agreement on a single
mechanism has been reached that explains the reason for the incremental oil recovery due
to low salinity water injection. This study focuses on investigating the low salinity
waterflooding method with an objective to understand the underlying mechanisms
responsible for incremental oil recovery through numerical simulation studies.

There are not many numerical studies on carbonate reservoirs in the literature,
compared to the numerical studies on sandstone reservoirs. Therefore, the aim of this
project is to investigate the driving mechanisms at coreflood scale using CMG (Computer
Modelling Group) software for the LSWI in carbonate reservoirs. The reservoir model
has been developed using CMG’s compositional simulator GEM. Impact of multi-
component ionic exchange and mineral dissolution on incremental oil recovery due to
low salinity waterflooding has been investigated using the diluted seawater. The impact
of temperature of the diluted sweater on oil recovery has also been investigated in this
study. The comparison of the results of different cases of sensitivity in this study has
shown that Multi-component ionic exchange (MIE) is the responsible underlying
mechanism of the low salinity waterflooding and temperature is the contributing factor to
the incremental oil recovery.
Table of Contents
List of Tables ..................................................................................................................... i

List of Figures ................................................................................................................... ii

Chapter 1: Introduction ......................................................................................................1

1.1 Challenges in Carbonate Reservoirs ........................................................................1

1.2 Low Salinity Water Injection (LSWI) .....................................................................1

1.3 Project Objectives ....................................................................................................2

Chapter 2: Literature Review ............................................................................................4

2.1 Low Salinity Waterflooding in Carbonate Reservoirs .............................................4

2.1.1 Individual Role of Reservoir and Injected Fluid Parameters on Improved Oil
Recovery.....................................................................................................................4

2.1.2 LSWI at laboratory scale .................................................................................12

2.1.3 LSWI at the field scale ....................................................................................14

2.1.4 Suggested Mechanisms ...................................................................................15

2.1.5 Numerical studies of low salinity smart waterflooding ..................................21

2.2 Description of CMG software................................................................................23

Chapter 3: Model development and analysis of LSWI ...................................................24

3.1 Building a model of core flooding experiment using CMG software ...................24

3.1.1 Reservoir section .............................................................................................24

3.1.2 Components section ........................................................................................25

3.1.3 Rock-fluid section ...........................................................................................25

3.1.4 Initial conditions section .................................................................................26

3.1.5 Wells section ...................................................................................................27

3.2 Chemical reactions during low salinity waterflooding ..........................................27

3.3 Calculations of aqueous-mineral reactions in GEM ..............................................29

3.4 Methodology ..........................................................................................................30


3.4.1 LSWI case for carbonates with SO4 simple relative permeability interpolation
..................................................................................................................................30

3.4.2 LSWI case for carbonates with mineral relative permeability interpolation ..31

3.4.3 LSWI case with temperature effects ...............................................................32

3.4.4 LSWI case with ion exchange relative permeability interpolation .................32

Chapter 4: Results and discussion ...................................................................................33

4.1 Validation of the model .........................................................................................33

4.2 LSWI case for carbonates with SO4 simple relative permeability interpolation ...33

4.2.1 Oil Saturation ..................................................................................................36

4.3 LSWI case for carbonates with mineral relative permeability interpolation .........39

4.4 LSWI case with temperature effects ......................................................................41

4.5 LSWI case with ion exchange relative permeability interpolation ........................42

Chapter 5: Conclusions....................................................................................................43
List of Tables

Table 3.1. Grid properties used for simulation taken from the experiment [40] ............ 24
Table 3.2. The composition of the crude oil taken from CMG tutorial [41] .................. 25
Table 3.3. End-point relative permeability data from experiment [40] .......................... 25
Table 3.4. Initial values of the properties for the initialization section [40] .................. 26
Table 3.5. Concentrations of initial aqueous and diluted seawater compositions and initial
mineral volume fractions (values are taken from the experiment [40]) ......................... 30

i
List of Figures

Figure 2.1: Imbibition test of different crude oils with various AN into chalk core [8] . 5
Figure 2.2: Imbibition of seawater and formation water into reservoir limestone at 130°C
[4] ..................................................................................................................................... 7
Figure 2.3: Spontaneous imbibition of increased sulfate ions concentration in seawater
into chalk cores [3] ........................................................................................................... 8
Figure 2.4: Spontaneous imbibition of increased calcium ions concentration in seawater
into chalk cores [16] ......................................................................................................... 8
Figure 2.5: Combined effect of an increase in temperature and concentration of calcium
ions on oil recovery from the calk core [16] .................................................................... 9
Figure 2.6: Combined effect of modifying concentrations of PDIs (Ca2+, Mg2+, SO42-) in
different temperatures (70°C, 100°C, 130°C) on oil recovery from a chalk core [17] .. 10
Figure 2.7: Increased oil recovery using sweater depleted in NaCl [18] ...................... 10
Figure 2.8: Impact of diluted seawater on oil recovery [19] ......................................... 11
Figure 2.9: Impact of adjusting salinity of the injected brine in spontaneous imbibition
test [23] ........................................................................................................................... 13
Figure 2.10: Impact of injection of softened water on oil recovery [24] ...................... 14
Figure 2.11: Visualisation of the proposed underlying mechanism (MIE) for the
wettability alteration of the carbonate rock surface by seawater [17] ............................ 16
Figure 2.12: Schematic of the attachment of fine particles to the rock surface [29] ..... 18
Figure 2.13: Impact of salinity decrease on interfacial tension [19] ............................. 18
Figure 2.14: A schematic of expansion of electric double layer: a) initial wetting
condition b) impact of NaCl depleted seawater [34] ..................................................... 19
Figure 2.15: Reaction of surface active components to high and low salinity [36] ...... 20
Figure 3.1: Relative permeability curve of the core represents initial wetting condition
[40] ................................................................................................................................. 26
Figure 3.2: Schematic of the 1D model of the simulated core formation...................... 27
Figure 3.3: Aqueous and mineral reactions in the process wizard ................................ 31
Figure 3.4: One more added chemical reaction to the process wizard .......................... 32
Figure 3.5: An ion exchange reaction with ion exchange parameters ........................... 32
Figure 4.1: Validating of the model by matching the simulation data with the
experimental data............................................................................................................ 33
Figure 4.2: Impact of dilution of seawater on oil recovery ........................................... 34

ii
Figure 4.3: An aqueous component mole of SO42- ions against pore volume injected . 34
Figure 4.4: An aqueous component mole of Ca2+ ions against pore volume injected... 35
Figure 4.5: Mineral moles changes against pore volume .............................................. 35
Figure 4.6: Wettability alteration from oil wet towards more water wet by injecting
diluted seawater .............................................................................................................. 36
Figure 4.7: Oil saturation versus distance plots after 25, 55, 96, and 276 minutes of
injection of seawater ....................................................................................................... 37
Figure 4.8: Oil saturation versus distance plots after 25, 55, 96, and 276 minutes of
injection of 5 times diluted seawater .............................................................................. 38
Figure 4.9: 2D visualisation of oil saturation across the reservoir ................................ 39
Figure 4.10: Impact of dilution of seawater on oil recovery (with mineral relative
permeability interpolation) ............................................................................................. 40
Figure 4.11: Mineral moles changes against pore volume (with mineral relative
permeability interpolation) ............................................................................................. 40
Figure 4.12: Effect of injection fluid temperature on oil recovery ................................ 41
Figure 4.13: Effect of temperature on aqueous component mole of SO42- ions ............ 41
Figure 4.14: Aqueous component moles of CH3COO-X and SO4-X ............................ 42
Figure 4.15: Impact of dilution of seawater on oil recovery (with ion exchange relative
permeability interpolation) ............................................................................................. 42

iii
Chapter 1: Introduction

1.1 Challenges in Carbonate Reservoirs


Carbonate reservoirs contain approximately 60% of oil, and 40% of gas reserves
out of the world’s total hydrocarbon reserves and most of these reservoirs are placed in
the Middle East. Compare to the sandstone reservoirs, carbonate reservoirs are very
complicated in structure in terms of heterogeneity such as natural fractures and a matrix
with very low permeability. All of these complexities are sourced from their depositional
history and later diagenesis [1]. Another factor causing carbonate reservoirs to be
heterogeneous is wettability of the rock-fluid system, which in turn it is a very crucial
factor controlling the fluid distribution in the reservoir. Although most of the sandstone
reservoirs are water wet, carbonate reservoirs are ranged from mixed wet to oil wet.
Because the carbonate rocks are mostly oil wet, because of the adsorbed oil on the rock
surface, it is harder to produce oil from matrix. Hence, the recovery factor of carbonate
reservoirs is usually way below than 30% [2]. Another challenge of these carbonate
reservoirs are the fluids containing hydrogen sulfide and carbon dioxide, which is very
toxic and extremely corrosive.

The reason of carbonate reservoirs to be more oil wet is related to the endurance
of the water layer between the rock surface and oil phase. Because of the attractive forces
between positively charged carbonate surface and negatively charged carboxylic group
in the crude oil are powerful, the water layer between rock surface and oil phase turns out
to be unstable, and the system becomes more oil wet [3].

1.2 Low Salinity Water Injection (LSWI)


In order to produce more fraction of the hydrocarbon reserves from carbonate
reservoirs, alteration of wettability from oil wet towards more water wet is required. Low
salinity water injection (LSWI) technique serves a right purpose to fulfill this function.
This technique is termed as Smart Waterflooding (Saudi Aramco), Advanced Ion
Management (AIMSM – Exxon Mobil), Designer Waterflooding (Shell), LoSal (BP) etc.
[4]. It has several following advantages compared to the other chemical enhanced oil
recovery methods:

(a) Ability to displace light to medium crude oils;


(b) Water is available and affordable;

1
(c) It requires less operating costs compare to the other expensive chemical EOR
methods.

Sourcing and disposal of water might be counted as the only problem associated
with low salinity water injection. In overall, there are two significant reasons to conduct
waterflooding in the reservoirs:

 In order to maintain the pressure in the reservoir to be more than bubble point
pressure;
 Oil is swept by water with the aid of viscous forces.

However, low salinity water injection differs from average water injection
technique due to the reason that it can change the chemical equilibrium between crude
oil-brine-rock system [4]. Basically, crude oil-brine rock system is being in chemical
equilibrium in the formation. If injection fluid composition has not been altered, in other
words, if the compositions of injected fluid and initial formation fluid are the same, then
chemical equilibrium will not be affected, only relative saturation of the fluids in the
formation will be changed. Therefore, the composition of the injected fluid is needed to
be different from initial formation fluid to cause an alteration in chemical equilibrium.
Hindering side of low salinity water injection is the speed of the geochemical reactions
which is very slow and can be affected by several parameters such as temperature,
pressure, etc. [4].

1.3 Project Objectives


Although low salinity water injection technique has gained popularity over the
last two decades, the responsible underlying mechanism is still under question.
Comprehending the underlying mechanism is very crucial in terms of production
forecasting. For sandstone reservoirs, number of laboratory experiments and field trials
have been conducted to understand the mechanism responsible for incremental oil
recovery [4]. However, the research on the study of mechanisms of LSWI for carbonate
reservoirs is limited because of the complex structure of the carbonate rocks. The primary
objective of this project is to investigate the reasons behind the incremental oil recovery
due to low salinity waterflooding in carbonate reservoirs by numerical modeling of a core-
flooding experiment using the GEM software of CMG (Computer Modelling Group)
suite. This research is mainly interested in scrutinizing the effect of multi-component

2
ionic exchange and mineral dissolution on incremental oil recovery due to low salinity
waterflooding.

3
Chapter 2: Literature Review

2.1 Low Salinity Waterflooding in Carbonate Reservoirs


Studies have verified that the residual oil saturation might be further reduced by
implementing Low Salinity Water Injection (LSWI) technique. However, the
mechanisms responsible for the additional oil recovery due to LSWI are still in question.
The complex structure of the carbonate rocks adds extra complexity to understand and
affirm these underlying mechanisms [5].

This chapter helps to comprehend the individual role of both reservoir and injected
fluid parameters on improved oil recovery. The applications of the LSWI in laboratory
and field scale on carbonate rocks, suggested mechanisms for improved oil recovery,
numerical studies of LSWI and description of GEM software of CMG suite have been
reviewed in this chapter.

2.1.1 Individual Role of Reservoir and Injected Fluid Parameters on


Improved Oil Recovery
The basic idea in injecting low salinity water is to change the initial chemical
equilibrium between crude oil-brine-rock system [4]. In order to understand this behavior,
the individual role of the reservoir and injected fluid parameters needs to be understood.

2.1.1.1 Composition of the Connate Water


Potential determining ions (PDIs) present in the connate water determine the zeta
potential (surface charge) of the rock-water interface. Connate water is known to be
highly saline and comprises of high concentration of Ca2+ ions, but the very low
concentration of Mg2+ [3]. High concentration of sulfate in the connate water, particularly
in high temperatures, leads to the precipitation of anhydrite (CaSO4). Therefore, the
concentration of sulfate ions in the connate water is very low [6]. According to the studies
of Nasralla et al.[7], if the injected fluid contains a high concentration of sulfate ions, it
can cause to form sulfate scales due to the presence of Ca2+, Sr2+, and Ba2+ ions in the
connate water, which in turn clogs the pore throats.

Therefore, in order to obtain maximum recovery factor from a given carbonate


reservoir, it is very crucial to determine the concentration of PDIs in the connate water
before optimizing the composition and concentration of injected fluid [5].

4
2.1.1.2 Composition of the Crude Oil
The number of carboxylic groups (-COOH) in the crude oil is the determining
factor regarding increased residual oil saturation because of the strong bond between the
negatively charged carboxylic group and positively charged carbonate rock surface. The
number of the carboxylic group is defined by the acid number (mg KOH/g), which can
be quantified as the required mass of KOH in milligrams to counteract with one gram of
crude oil [4]. The relationship between the acid number and oil recovery has been
investigated by Standnes and Austad [8], and the result is shown in Figure 2.1. Different
crude oils with various acid numbers have used in this experiment, and the high the acid
number, the low the recovery factor and low imbibition rate have been observed.

Figure 2.1: Imbibition test of different crude oils with various AN into chalk core [8]

2.1.1.3 Effect of temperature and pressure


Although the idea of the smart waterflooding is to change initial chemical
equilibrium between the crude oil-brine-rock system, most of the geochemical reactions
happen in a very long-time frame compare to the production time of the field. The
temperature in the reservoir is a contributing factor in terms of increasing the speed of the
geochemical reactions [4]. It has been found by Shimoyama and Johnes [9] that there is
a relationship between the temperature and the acid number of crude oil. As temperature

5
increases in the reservoir, the acid number of the crude oil decreases, thereby oil recovery
factor is increased.

Adsorption of potential determining ions (SO42-, Ca2+, Mg2+) on rock surface


increases as the temperature increases. However, the reactivity of divalent cations (Ca2+,
Mg2+) toward the carbonate rock surface is different. At low temperatures, Ca2+ ions
appear to be more adsorbed on rock surface than Mg2+ ions, but at high temperatures,
Mg2+ ions are being more reactive toward the rock surface [3, 10].

The presence of anhydrite (CaSO4) in the reservoir is essential to observe low


salinity effect. But anhydrite can only be dissolved in formation water at lower
temperatures [11].

According to the report of Buckley [12], pressure is also a contributing factor to


change the wetting conditions in the reservoir. He observed that the solubility of
asphaltene in crude oil increases as the pressure increases.

2.1.1.4 Effect of rock mineralogy

The impact of low salinity water injection for various carbonate rocks is not the
same [13]. The experiments of Romanuka et al.[14] indicated that LSWI did not have an
impact on chalk core samples, but dolomite and limestone responded to the diluted
seawater. Because of the limestone is less positively charged compared to the dolomite,
the bond between the dolomite surface and the carboxylic group is stronger than the bond
between limestone surface and carboxylic group. Therefore, the effect of low salinity
waterflooding is stronger for limestone [13].

Limestone can be classified as reservoir and outcrop; which outcrop limestone


shows utterly different behavior compare to the reservoir limestone. Experiments have
shown that the surface of the outcrop limestone is not sensitive to the potential
determining ions. Therefore, outcrop limestone cannot be used for the modeling of the
reservoir rock to simulate the future performance of the reservoir [4].

6
2.1.1.5 Ionic composition of the injected fluid
In order to achieve an extra oil production by waterflooding in a tertiary oil
recovery process, the injected fluid composition needs to be modified so as to not to be
in the same composition with initial connate water. Otherwise, the chemical equilibrium
between crude oil-brine-rock system will not be exposed to any alteration, only relative
saturations will be changed [4]. This is illustrated in the below Figure 2.2 to show the
difference in oil recovery when formation water and seawater are used as an injection
fluid. As it can be seen when seawater replaced formation water the recovery factor
increased from less than 10% to more than 35%.

Figure 2.2: Imbibition of seawater and formation water into reservoir limestone at 130°C [4]

The presence of potential determining ions in seawater is an explanation for the


increased oil recovery compare to the formation water. Potential determining ions are
capable of altering the zeta potential (surface charge) of the carbonate surface [5]. The
influence of each of PDIs on oil recovery has been investigated by several authors and is
reviewed below:

Influence of SO42-: Research conducted by Zhang and Austad [3] has concluded
that increasing concentration of SO42- in seawater from 0 to 4 times the concentration of
normal seawater can increase the oil recovery substantially. Another author [15] has also
reported that seawater with four times increased concentration is an optimum
concentration that should be used in LSWI. As it can be seen from Figure 2.3 that

7
increasing SO42- ions concentration increased oil recovery from 10% to almost 50%
OOIP.

Figure 2.3: Spontaneous imbibition of increased sulfate ions concentration in seawater into
chalk cores [3]

Influence of Ca2+: Another research by Zhang et al. [16] confirmed that increasing
Ca2+ ions concentration in seawater from 0 to 4 times the concentration of ordinary
seawater can increase the oil recovery factor by almost 30%. In this experiment, sulfate
ions concentration is not modified and remained as same as in the ordinary seawater.

Figure 2.4: Spontaneous imbibition of increased calcium ions concentration in seawater into
chalk cores [16]

8
However, increasing Ca2+ ions concentration in the presence of excess sulfate ions
decreases the adsorption of rock surface due to the precipitation of anhydrite (CaSO4).
Therefore, the ratio of Ca2+/SO42- needs to be in consideration, especially in high
temperatures [10].

In another experiment, Zhang et al. [16] investigated the combined effect of an


increase in temperature and the concentration of calcium ions, and the result is in the
following Figure 2.5.

Figure 2.5: Combined effect of an increase in temperature and concentration of calcium ions on
oil recovery from the calk core [16]

Influence of Mg2+: At low temperatures, Mg2+ ions are not as reactive as Ca2+
ions towards the rock surface. However, at high temperatures, Mg2+ ions turn to be more
adsorbed on rock surface than Ca2+ and can even displace complex calcium-carboxylic
group from the surface of the carbonate rock and alter the wettability towards more water
wet [4].

Zhang et al. [17] have investigated the combined effect of modifying the
concentration of three potential determining ions (Ca2+, Mg2+, SO42-) in different
temperatures (70°C, 100°C, 130°C) on oil recovery from a chalk core. As it can be easily
seen from the Figure 2.6 that at low temperatures, modifications in the concentrations of
PDIs do not affect oil recovery. Therefore, it can be concluded that temperature is an

9
essential determining factor which needs to be considered when implementing low
salinity waterflooding.

Figure 2.6: Combined effect of modifying concentrations of PDIs (Ca2+, Mg2+, SO42-) in
different temperatures (70°C, 100°C, 130°C) on oil recovery from a chalk core [17]

NaCl depleted seawater: “Smart water” can be made by decreasing the


concentration of NaCl in the seawater due to the reason that NaCl is an inactive ion in
seawater which hinders the access of the active ions (Ca2+, Mg2+, SO42-) [11]. Fathi et al.
[18] conducted an experiment using a different concentration of NaCl. Results in Figure
2.7 shows that decreasing the concentration of non-active ion increased the oil recovery
factor.

Figure 2.7: Increased oil recovery using sweater depleted in NaCl [18]

10
2.1.1.6 Ionic concentration of the injected fluid
Wettability alteration can be achieved by using diluted seawater instead of
modifying ionic composition by changing concentrations of potential determining ions.
Yousef et al. [19] conducted an experiment using 2, 10, 20 and 100 times diluted seawater
to observe the effect on oil recovery. It can be observed from the Figure 2.8 that there is
no increase in oil recovery after injection of 100 times diluted seawater. It can be
concluded that there is a salinity threshold which when designing the injection fluid, the
salinity should not be lower than the threshold value.

Figure 2.8: Impact of diluted seawater on oil recovery [19]

2.1.1.7 Addition of Surfactant into Seawater


Another way of changing wettability towards more water-wet conditions is the
addition of positively charged surfactant monomers into seawater. The underlying
mechanism of this process is that of the bond between positively charged surfactant
monomers and negatively charged carboxylic group in the crude oil is so strong that it
can be desorbed from the surface of the carbonate rock, hence rock surface becomes more
water-wet. The experiment of Standnes and Austad [8], also confirms that the addition of
cationic surfactant monomers can increase the oil recovery factor by up to 70%. The only
drawback of this method is that it is economically expensive, hence a cheaper method of
injecting low salinity water needs to be fully understood.

11
2.1.2 LSWI at laboratory scale

Spontaneous imbibition tests:

Because of the combined effect of capillary forces and interfacial tension, wetting
phase displaces the non-wetting phase in spontaneous imbibition process. One of the
drawbacks of spontaneous imbibition test is that of the long-time period needs to be
waited to get the results, as it can take up to several months [20].

Strand et al. [21] conducted an experiment to verify seawater as a “Smart water”


to cause incremental oil recovery. They investigated the reaction of the potential
determining ions toward the rock surface at different temperatures and result of this
experiment concluded that carbonate rock surface is susceptible to the relative
concentrations of PDIs. One of the observations was that the adsorption of Ca2+ and Mg2+
ions onto the rock surface was kind of similar at low temperatures. However, adsorption
of Mg2+ ions onto the rock surface was very strong at high temperatures.

In another spontaneous imbibition experiment [22], the impact of changing


salinity of the injecting brine has been investigated, and the result of this experiment
confirmed that incremental oil recovery could be obtained just lowering the salinity of
the injected brine, even at 70°C.

Zaeri et al. [23] conducted an experiment to observe the effect of LSWI on


wettability alteration process. They used 5, 10, 20, 40 times diluted seawater and distilled
water to see the impact on oil recovery by spontaneous imbibition. As it can be seen from
Figure 2.9, distilled water as an imbibition fluid showed no incremental oil recovery. The
maximum oil recovery obtained from the test that 20 times diluted seawater used as an
imbibition fluid and 40 times diluted seawater did not show maximum oil recovery, even
lower recovery factor than ten times diluted seawater. The conclusion can be made from
this experiment that there is a salinity threshold value that no further reduction below than
that value can cause incremental oil recovery.

12
Figure 2.9: Impact of adjusting salinity of the injected brine in spontaneous imbibition test [23]

Coreflooding tests:

Although the impact of low salinity water injection can be observed by


implementing spontaneous imbibition tests, the direct application of low salinity water
injection on the field scale cannot be conducted, as the recovery factor obtained from
spontaneous imbibition test is more than that of obtained on the field scale. For that
purpose, core-flooding tests are used to comprehend the underlying mechanism of
increased oil recovery due to low salinity water injection [5].

According to the core-flooding experiment conducted by Gupta et al. [24],


modifying injecting water composition can substantially increase the oil recovery factor.
The result of this experiment showed that 5-9% incremental oil recovery can be obtained
just using seawater increased with sulfate ions instead of formation water. They also
added borate and phosphate ions to the injected water which resulted in 15-20%
incremental oil recovery. As can be seen from Figure 2.10, an incremental oil recovery
can be obtained using softened water (Ca2+ and Mg2+ ions eliminated from formation
water) as an injecting fluid.

13
Figure 2.10: Impact of injection of softened water on oil recovery [24]

2.1.3 LSWI at the field scale


In comparison with sandstone reservoirs, the application of low salinity water
injection in carbonates on the field scale has not been widespread. The application of low
salinity water injection in carbonates on the field scale was firstly reported by Yousef et
al. [25], which 2 single well chemical tracer tests were conducted to estimate the impact
of LSWI on oil recovery. Adjusting the ionic concentration of Qurayyah seawater to inject
into Upper Jurassic carbonate reservoir has concluded 7 saturation units decrease in the
residual oil saturation and the underlying mechanism for IOR due to low salinity water
injection was verified to be wettability alteration. It was argued that application of LSWI
could impact differently for various carbonate reservoirs due to the difference in reservoir
properties, such as reservoir temperature, the different chemical composition of the crude
oil and connate water and heterogeneity in the reservoir [25].

Moreover, substantial incremental oil recovery was also reported by the injection
of seawater with the pipelines from the Persian Gulf into a limestone reservoir (with no
fracture) located in Saudi Arabia [4].

14
2.1.4 Suggested Mechanisms
In the literature, various mechanisms suggested on increased oil recovery due to
low salinity water injection. These mechanisms depend on the chemical equilibrium
between crude oil – formation brine – rock system. Altering wettability of the carbonate
rock surface from oil wet toward more water-wet condition has been the most common
agreed primary mechanism for incremental oil recovery. However, some studies suggest
other mechanisms, which the underlying primary mechanism is still in debate. But, all of
the proposed mechanisms cause the alteration in the wettability of the rock surface.
Therefore, some can say that the change in wettability of the carbonate surface is a result
of low salinity waterflooding rather than being the reason for it [5].

2.1.4.1 Multi-component Ionic Exchange (MIE)


This underlying mechanism was first suggested by Zhang and coworkers [17].
According to them, this mechanism is the result of fluid-rock interactions, which the
reduction of attractive forces between rock surface and adsorbed carboxylic groups of the
crude oil is deemed to be the primary mechanism in charge of the wettability alteration.
The visualization of this mechanism is shown in Figure 2.11, which indicates that multi-
component ionic exchange is about exchanging ions between potential determining ions
and the ions on the rock surface.

Because of the lower pH (less than 9), carbonate surface was initially positively
charged [4]. It is known that the concentration of sulfate ions in seawater is about twice
the concentration of calcium ions [3]. However, the amount of sulfate can be present in
three forms in a reservoir:

- CaSO4 (solid) as a precipitated anhydrite


- SO42- (in the aqueous phase)
- SO42- (adsorbed on rock surface)

But, the sulfate ions only in the aqueous phase can lead to the wettability alteration of
rock surface [11]. As the seawater is injected into the carbonate reservoir, the sulfate ions

15
Figure 2.11: Visualisation of the proposed underlying mechanism (MIE) for the wettability
alteration of the carbonate rock surface by seawater [17]

present in the aqueous phase commence to absorb onto the positively charged carbonate
rock surface and makes the rock surface less positively charged. In the meantime, the
concentration of Ca2+ ions close to the carbonate surface increases due to the less
repulsive forces between positively charged calcium ions and less positively charged rock
surface. As the concentration of calcium ions close to rock surface increases, the chances
of desorbing crude oil from the surface increases due to the bond between calcium ions
and negatively charged carboxylic group in the crude oil. Adsorption on rock surface is
very sensitive to the temperature, as the temperature increases, the concentration of Ca2+
and SO42- adsorbed on rock surface increases and the system turns to be more water wet
[17].

However, the reactivity of Mg2+ toward the carbonate surface increases as the
temperature increases. At high temperatures, Mg2+ ions can displace Ca2+ ions adsorbed
on the rock surface. It can even eliminate the complex Ca2+ - carboxylic bonding from
the carbonate surface and lead to more water wetness [4].

A study by Austad et al. [11] resulted that if the injected brine is lack of sulfate
ions, the presence of anhydrite in the rock formation is enough for wettability alteration.
In this case, sulfate ions are created by the dissolution of anhydrite in the reservoir.
It can be concluded that in order to change the wettability of the rock system,
reservoir temperature and the concentrations of the potential determining ions need to be
in consideration.

16
2.1.4.2 Mineral Dissolution
According to the report of Hiorth et al. [26], the multi-component ionic exchange
is not a responsible mechanism for increased oil recovery. Instead, mineral dissolution is
in charge of the increase in oil recovery factor. Calcium ions are not extremely reactive
towards the rock surface at low temperatures. However, at high temperatures, because of
the presence of sulfate ions in the formation water, calcium ions react with sulfate ions,
and precipitation of anhydrite (CaSO4) happens. Therefore, the concentration of calcium
ions in the aqueous phase decreases. In order to keep the chemical equilibrium between
the aqueous phase and carbonate rock surface, the dissolution of anhydrite into the
aqueous phase happens. Hence desorption of crude oil with calcium from the carbonate
surface increases oil recovery by decreasing residual oil saturation.

However, another experiment by Nasralla et al. [27] showed that mineral


dissolution could not happen with all kind of brines used, so in that case, calcite
dissolution cannot be regarded as a primary mechanism. Mahani et al. [28] conducted an
experiment to observe the impact of seawater and 25 times diluted seawater on oil
recovery. The oil recovery factor increased in both case, although no mineral dissolution
was expected by the injection of SW and 25dSW. Hence they concluded that mineral
dissolution is not the underlying mechanism of low salinity water injection.

2.1.4.3 Fines detachment


Investigation of Schembre and Kovscek [29] confirmed that migration of fines is
the primary mechanism of low salinity water injection. As it is shown in Figure 2.12, a
carboxylic group in the crude oil is attached to fines in the rock formation. At elevated
temperatures fines are released from the carbonate rock surface with its attached heavy
end fraction of the crude oil and altering the wettability of system toward more water
wetness.

Another experiment by Tang and Morrow [30] resulted that as the fine particles
detached from the rock surface at elevated temperatures, they can clog some pore throats
and divert the fluid flow, thereby increasing the microscopic sweep efficiency with the
decrease in permeability.

17
Figure 2.12: Schematic of the attachment of fine particles to the rock surface [29]

2.1.4.5 Increase in pH
Increase in pH can lead to the generation of in-situ surfactants which in turn gives
rise to a decrease in interfacial tension. Hence, as the value of interfacial tension
decreases, the value of capillary number increases, consequently residual oil saturation
decreases and oil recovery increases [5]. According to the experiment of Meng et al. [31],
seawater increased with the concentration of phosphate can lead to a decrease in
interfacial tension and improve the wettability toward more water wetness. In another
experiment, Yousef et al. [19] replaced connate water with seawater to observe the impact
on interfacial tension. As can be seen from Figure 2.13, the substantial decrease in the
value of IFT was only observed when connate water was replaced by seawater. However,
the subsequent dilution of seawater did not have a significant effect on IFT. From this
experiment, it can be concluded that low salinity waterflooding has more effect on rock-
brine interaction rather than oil-brine interaction

Figure 2.13: Impact of salinity decrease on interfacial tension [19]

18
2.1.4.6 Expansion of Electric Double Layer
This mechanism was first ever experimented on sandstone rocks by Ligthelm et
al. [32], and the results showed that presence of high concentration of multivalent cations
in the high saline seawater could lead to wettability to be more oil-wet due to the bond
between positively charged multivalent cations and negatively charged oil surface. In this
case, compression of the electric double layer is experienced [5].

In another experiment by Fathi et al. [33], it has been reported that the impact of
the presence of non-active NaCl ions on wettability alteration is as crucial as the presence
of potential determining ions in seawater. They observed that as NaCl depleted seawater
used as an injecting fluid, oil recovery factor increased significantly. This can be tied up
to the reason that non-active NaCl ions hinder the reactivity of the active potential
determining ions in the double layer. A schematic of expansion of an electric double layer
is illustrated in the following Figure 2.14.

Figure 2.14: A schematic of expansion of electric double layer: a) initial wetting condition b)
impact of NaCl depleted seawater [34]

19
2.1.4.7 Formation of Micro-Dispersions
All of the mechanisms that have been suggested so far were only focused on rock-
fluid interaction rather than fluid-fluid interaction. Experimental investigation by
Mohamed et al. [35] verified that incremental oil recovery could be obtained by low
salinity water injection due to fluid-fluid interaction. They attributed this increase in oil
recovery to the presence of surface-active components of crude oil, which is called micro-
dispersions. The reaction of surface active components to high and low salinity water is
shown in below Figure 2.15.

Figure 2.15: Reaction of surface active components to high and low salinity [36]

They conducted another test to observe the role of micro-dispersion in crude oil,
so they removed the surface-active components of crude oil, and no incremental oil
recovery was observed. Hence, they concluded that there should be a relationship between
improved oil recovery and the tendency of oil to form micro-dispersions [35].

20
2.1.5 Numerical studies of low salinity smart waterflooding
Laboratory and field scale applications of low salinity waterflooding have
indicated an incremental oil recovery compared to the basic waterflooding. However, so
as to estimate the feasibility of the project, numerical modeling of the field is required in
order to be able to evaluate the future performance of the reservoir. Hence, different
numerical studies have been conducted by several authors as described in the following
subsections.

2.1.5.1 Studies of C. Esene et.al


In this study [37], numerical modeling technique has been deployed to explore the
responsible mechanism for the incremental oil recovery due to low salinity water
injection. The authors have investigated the crucial aspects of LSWI such as a change in
the wettability of the rock-fluid system, alteration in the pH of the formation water and
mineral chemical reactions. In order to look into the impact of multi-component ionic
exchange (MIE), they developed a composition model using CMG software and all of the
required laboratory data was supplied by Computer Modelling Group. The key results of
this study are as follows:

- Alteration of the wettability from oil wet towards more water wet was considered
as the main mechanism in charge of the incremental oil recovery.
- For both carbonate and sandstone reservoirs, there is a threshold value for the
salinity of the injected fluid, which is regarded as an optimum salinity gives rise
to the maximum oil recovery. Further decrease in salinity beyond the threshold
value does not give rise to a substantial change in oil recovery.
- Sandstone reservoirs are more sensitive to the change in pH of the formation water
compared to the carbonate reservoirs. It is quite obvious that as pH of the
formation water increases, interfacial tension decreases, hence residual oil
saturation decreases, which in turn leads to greater oil recovery factor.
- Calcite dissolution and dolomite precipitation are found to be the significant
mineral reactions happening during LSWI, which leads to increased oil recovery
in the presence of HCO3- and other catalytic ions.

21
2.1.5.2 Studies of C. Qiao et.al
In this study [38], authors have built a mechanistic model in which all
geochemical interactions between CBR (crude oil-brine-rock) system that can change
rock wettability have been included. They improved the model by adding mineral
dissolution reactions to the dataset, and low salinity waterflooding was conducted on a
limestone and chalk cores with and without anhydrite to observe the importance of the
presence of the anhydrite in the formation. All reactions and flow equations are solved in
PennSim simulator at the same time. The key results of the experiment are as following:

- The injection of low salinity water into the chalk cores without anhydrite can lead
to up to 6% incremental oil recovery.
- However, the injection of low salinity water into limestone cores without
anhydrite did not give any incremental oil recovery.
- If anhydrite is present in the formation, it plays a role as a source of natural sulfate.
When diluted formation water is injected into the formation, anhydrite dissolution
leads to the creation of in-situ sulfate ions.

2.1.5.3 Studies of E.W. Al-Shalabi et al.


The impact of low salinity water injection on increased oil recovery has been
proved by the laboratory tests. In this study [39], a mechanistic model was developed by
authors to observe the impact of geochemical reactions on oil recovery due to low salinity
waterflooding. They used that model to history match the published core-flooding
experiments using UTCHEM simulator. This model in the UTCHEM simulator was tuned
to estimate the molar Gibbs free energy of brine, which relative permeability curves are
a function of this free energy. Some of the key results of this study are as follows:

- Modeling the impact of low salinity water injection on oil recovery was
conducted, and this model was successfully matched with the core-flooding
experiment of Yousef et al. [40].
- From the model, the conclusion can be drawn that residual oil saturation and oil
rel-perm curves are functions of Gibbs free energy of brine.
- This model has shown that change in the wettability of the rock, mineral
dissolution and migration of the fines are responsible mechanisms for the
incremental oil recovery.

22
2.2 Description of CMG software
CMG suite of software contains three simulators (IMEX, GEM, STARS). All
three simulators can be launched by Builder which is a tool used to establish input files
for the simulators. Builder gives access to validate the data prior to running the simulation
by creating or importing reservoir properties, such as grid and its properties, fluid model
properties, rock-fluid data, well locations, and initial conditions. WinProp is used for to
create fluid models. [41]. In this project, the a multidimensional, equation-of-state (EOS)
compositional simulator GEM is used to simulate the impact of low salinity (smart)
waterflooding effect on the oil recovery from the carbonate rocks.

23
Chapter 3: Model development and analysis of LSWI

3.1 Building a model of core flooding experiment using CMG software


The model was developed using the experimental data conducted by Yousef et al.
[40] which the impact of the dilution of injection fluid on oil recovery has been
investigated. The objective of numerical modeling of this experiment is to examine the
underlying mechanism responsible for incremental oil recovery, which comprehending
the mechanism is very crucial in terms of production forecasting. The procedure that has
been used to build a model using CMG software is shown in the below subsections.

3.1.1 Reservoir section


The compositional model was constructed using the Builder in CMG software.
Reservoir section in Builder is used for creating simulation grid structure and feed the
grid blocks with the array properties such as porosity, permeability, initial water
saturation, grid top and bottom depth, grid thickness, rock compressibility, etc.
1D model is consists of a Cartesian grid system with 40 x 1 x 1 grid blocks; length,
width, and thickness of the grid blocks are constant in I, J and K directions which 0.0041
m, 0.0337 m, 0.0337 m respectively. Data needed to feed the Reservoir section is taken
from the experiment [40], and are compiled in Table 3.1.

Table 3.1. Grid properties used for simulation taken from the experiment [40]

Property Value
1D model (40 x 1 x 1)
I direction: ∆x = 0.0041 m
Dimensions of the model
J direction: ∆y = 0.0337 m
K direction: ∆z = 0.0337 m

Porosity 0.251
Permeability 39.6 mD
Initial water saturation 0.1

24
3.1.2 Components section
The oil composition used in this simulation is taken from the CMG tutorial [41].
Peng-Robinson equation of state is used to build the fluid model. Crude oil composition
is shown in Table 3.2.

Table 3.2. The composition of the crude oil taken from CMG tutorial [41]

Component Composition fraction


N2+C1 0.298274
C2 - C5 0.263297
C6 - C9 0.181483
C10 - C15 0.160805
C21 - C26 0.0633269
C32 - C34 0.0238583
Asphaltene 0.0089552

3.1.3 Rock-fluid section


The initial wetting condition of the rock-fluid system has been represented by
relative permeability curves obtained from the experiment (Figure 3.1). Relative
permeability curves were generated using the end-point relative permeability data as
listed in the Table 3.3.

Table 3.3. End-point relative permeability data from experiment [40]

Parameter Value
Endpoint saturation: connate water 0.1
Endpoint saturation: critical water 0.1
Endpoint saturation: irreducible oil 0.3
Endpoint saturation: residual oil 0.3
Relative permeability at connate water 0.5
Relative permeability at irreducible oil 0.4
Exponent for calculating relative permeability at connate water 4
Exponent for calculating relative permeability at irreducible oil 2

25
0.6

0.5
Relative permeability

Krw Kro
0.4

0.3

0.2

0.1

0
0.10 0.14 0.18 0.21 0.25 0.29 0.33 0.36 0.40 0.44 0.48 0.51 0.55 0.59 0.63 0.66 0.70

Water Saturation (Sw)

Figure 3.1: Relative permeability curve of the core represents initial wetting condition [40]

3.1.4 Initial conditions section


Initial conditions section specifies the state of the reservoir at the beginning of the
simulation process. Initial properties specified were: concentrations of the primary
aqueous components, reservoir pressure, water saturation, the volume fraction of the
minerals and composition of the oil components.

Table 3.4. Initial values of the properties for the initialization section [40]

Parameter Value
Initial reservoir pressure (kPa) 20000
Initial reservoir temperature (°C) 100
Initial water saturation 0.1
Ca++ initial aqueous concentration (ppm) 19040
S04- initial aqueous concentration (ppm) 350
Mg++ initial aqueous concentration (ppm) 2439
HCO3- initial aqueous concentration (ppm) 354
Na+ initial aqueous concentration (ppm) 59491
Calcite initial volume fraction 0.64
Dolomite initial volume fraction 0.1
Anhydrite initial volume fraction 0.02

26
3.1.5 Wells section
Reservoir model utilizes two wells: 1 injector to inject the low salinity water and
1 producer to recover the oil discalced by the injection water. The injection and producer
wells were drilled and perforated through blocks of (1 1 1) and (40 1 1) respectively. The
constraint for the producer is to maintain a minimum bottom-hole pressure of 20000 kPa.
Two constraints were set for the injection well by an injection flow rate of 0.00024151
bbl/day and by a maximum BHP of 27600 kPa. The schematic of the constructed 1D
model is shown in Figure 3.2.

Figure 3.2: Schematic of the 1D model of the simulated core formation

3.2 Chemical reactions during low salinity waterflooding


Modeling of low salinity waterflooding requires several chemical reactions
expressing aqueous, mineral, ion exchange, adsorption reactions and hydrocarbon
solubility [41]. Those reactions are as following;

Hydrocarbon solubility:

𝐶𝑂2(𝑜) ↔ 𝐶𝑂2(𝑔) ↔ 𝐶𝑂2(𝑎𝑞) (1)

𝐶𝑂2(𝑎𝑞) + H2O ↔ H+ + HCO3- (2)

Aqueous reactions:

CaSO4 ↔ Ca2+ + SO42- (3)

27
MgSO4 ↔ Mg2+ + SO42- (4)

NaSO4- ↔ Na+ + SO42- (5)

CaCO3 + H+ ↔ Ca2+ + HCO3- (6)

MgCO3 + H+ ↔ Mg2+ HCO3- (7)

CaHCO3+ ↔ Ca2+ + HCO3- (8)

MgHCO3+ ↔ Mg2+ + HCO3- (9)

NaHCO3+ ↔ Na+ + HCO3- (10)

Mineral reactions:

Calcite + H+ ↔ Ca2+ + HCO3- (11)

Dolomite + 2H+ ↔ Ca2+ + 2HCO3- + Mg2+ (12)

Anhydrite + H+ ↔ Ca2+ + SO42- (13)

Ion exchange reactions:

1 1
Na+ + 2 (Ca – X2) ↔ 2 Ca2+ + (Na – X) (14)

1 1
Na+ + 2 (Mg – X2) ↔ 2 Mg2+ + (Na – X) (15)

Adsorption reactions:

X+ + SO42- ↔ XSO4- (16)

28
3.3 Calculations of aqueous-mineral reactions in GEM
In order to observe the low salinity effect on incremental oil recovery, the
geochemical equilibrium between crude oil-brine-rock system needs to be changed. This
means that aqueous and mineral species are not homogenous and the speed of the reaction
between aqueous and mineral species is slower than aqueous species reactions. It can be
concluded that if there is no equilibrium between aqueous and mineral species, minerals
can deposit or dissolve. The rate of the mineral reactions is calculated using the following
equations in GEM [41].

𝑄𝛽
𝑟𝛽 = Ā𝛽 𝑘𝛽 (1- ), β = 1,…,Rmn (17)
𝐾𝑒𝑞,𝛽

𝐸𝑎𝛽 1 1
𝑘𝛽 = 𝑘0𝛽 exp [ - ( - )] (18)
𝑅 𝑇 𝑇0

Where, 𝑟𝛽 - is the rate of the mineral reaction;

Ā𝛽 - is the reactive surface area;

𝑘𝛽 - is the rate constant of mineral reaction β;

𝑄𝛽 - is analogous to the activity product for aqueous reactions (i.e. related to


molality);

𝑘0𝛽 - is the reaction rate constant;

𝐸𝑎𝛽 - is the activation energy;

𝑄𝛽
- is the saturation index.
𝐾𝑒𝑞,𝛽

𝑄𝛽 𝑄𝛽
If > 1, mineral dissolution occurs; If < 1, precipitation occurs.
𝐾𝑒𝑞,𝛽 𝐾𝑒𝑞,𝛽

29
3.4 Methodology
The impact of smart waterflooding on oil recovery has been investigated by
implementing different methods using “Process wizard” function, which enables us to
model ion exchange that is initiated by a decrease in salt concentration of the injected
fluid. Winprop can be used to define aqueous components with its associated reactions
and ion exchange parameters.

3.4.1 LSWI case for carbonates with SO4 simple relative permeability
interpolation
In this first scenario, the low salinity waterflood is investigated using SO4 simple
relative permeability interpolation in carbonates with respect to change in the aqueous or
mineral reaction structure. In order to understand the mechanism responsible for
wettability alteration, this model is developed to observe the impact of the dilution of
seawater on oil recovery. The concentrations of initial aqueous and diluted seawater
compositions, initial mineral volume fractions that are used in this simulation study are
shown in Table 3.5.
Table 3.5. Concentrations of initial aqueous and diluted seawater compositions and initial mineral
volume fractions (values are taken from the experiment [40])

Composition (ppm) Na+ Ca2+ Mg2+ SO42- HCO3- CL-


Formation water 59491 19040 2439 350 354 132060
Seawater 1464 52 168 342 10 2576
Five times diluted seawater 293 10 34 68 2 515
Ten times diluted seawater 146 5 17 34 1 258

Initial mineral volume fractions


Calcite 0.64
Dolomite 0.1
Anhydrite 0.02

The process wizard adds 5 aqueous and 3 mineral reactions to the existing model.
The added reactions are shown in the following Figure 3.3.

30
Figure 3.3: Aqueous and mineral reactions in the process wizard

3.4.2 LSWI case for carbonates with mineral relative permeability


interpolation
In this second scenario seawater and five times diluted seawater are used in this
scenario to observe the impact on oil recovery. The concentrations used for the injection
fluids are as same as shown in Table 3.5. Sorw and Krw reduction values for the relative
permeability interpolation has been set to 0.482 and 0.7 respectively in the process
wizard. The process wizard adds 5 aqueous and 3 mineral reactions to the existing model
as shown in Figure 3.3.

31
3.4.3 LSWI case with temperature effects
In this scenario, the impact of temperature of injection fluid on oil recovery due
to low salinity water injection has been investigated. However, altering the temperature
of the injection fluid is not available through Builder. Therefore, in order to switch on the
thermal calculations in GEM simulator, several keywords need to be added to the dataset
through the aid of text editor.

First of all, thermal calculations need to be turned on in GEM; hence “THERMAL


ON” keyword needs to be added to the dataset. Next step is adding the injection fluid
temperature, which after the “OPERATE” keyword for the injection well “INJ-TEMP
‘INJ’” keyword needs to be added and the temperature of the injection fluid is specified
in the next row. In this case, three injection fluid temperatures (100 °C; 150 °C; 200 °C)
were compared to observe the impact on oil recovery.

3.4.4 LSWI case with ion exchange relative permeability interpolation


In this fourth scenario, Sorw and Krw reduction values for the relative
permeability interpolation has been set to 0.482 and 0.7 respectively in the process
wizard. The initial concentration of CH3COO- ion was defined as the value of 5000 ppm.
Seawater and five times diluted seawater were used in this case, and the concentrations
of the ions in the injected fluid were specified as shown in Table 3.5.

Process wizard added one more aqueous reaction as shown in Figure 3.4, which
overall, 6 aqueous and 3 mineral reactions were added to the dataset. An ion exchange
reaction with ion exchange parameters is shown in Figure 3.5. The constant value of 50
for the cation exchange capacity (CEC) was included in the dataset.

Figure 3.4: One more added chemical reaction to the process wizard

Figure 3.5: An ion exchange reaction with ion exchange parameters

32
Chapter 4: Results and discussion

4.1 Validation of the model


Prior to analyzing the impact of different scenarios, validation of the existing
model by matching the simulation data with the experimental data is required. The
experimental result of the oil recovery factor has been extracted from the paper [42] to
match with the simulation data. As it can be seen from Figure 4.1, simulation data pretty
much matches with the experimental data. Therefore, this model regarded as a validated
model and can be further investigated.

70

60

50
Oil recovery (%)

40
Simulated

30 Experimental

20

10

0
0 5 10 15 20 25 30 35 40
Pore volume injected

Figure 4.1: Validating of the model by matching the simulation data with the experimental data

4.2 LSWI case for carbonates with SO4 simple relative permeability
interpolation
According to the experiment of Yousef et al. [40], the effect of dilution of the
injected fluid on carbonates has been investigated. In the first experiment core was
flooded with seawater and then followed by different dilutions of seawater. In this project,
we consider three injection cycles, which firstly with seawater, then followed by five and
ten times diluted seawater. As it can be seen from Figure 4.2, injection of seawater gives
the oil recovery factor of 62 % and the second cycle of injection of five times diluted
seawater increased oil recovery up to 75%. However, the third cycle of injection of 10
times diluted seawater did not show up a substantial incremental oil recovery, just a 2 %
increase from the second cycle of injection. Hence, it can be concluded that there is a

33
threshold value for the maximum dilution of seawater, if exceeded, no substantial
incremental oil recovery can be observed.

90
80
70
Oil recovery factor (%)

60
50
10 times diluted seawater
40
5 times diluted seawater
30
Seawater
20
10
0
0 500 1000 1500 2000 2500
Pore volume injected

Figure 4.2: Impact of dilution of seawater on oil recovery

As can be seen from Figure 4.3, as the dilution of seawater increased, the
component mole of SO42- ions in the aqueous phase decreases meaning that more of the
SO42- ions adsorbed on rock surface and causing desorption of oil from rock surface,
hence causing to change the wettability from oil wet towards more water wet.

9.00E-05
Aqueous component mole of SO42-

8.00E-05

7.00E-05

6.00E-05
ions (gmole)

10 times diluted seawater


5.00E-05
5 times diluted seawater
4.00E-05
Seawater
3.00E-05

2.00E-05

1.00E-05

0.00E+00
0 500 1000 1500 2000 2500
Pore volume injected

Figure 4.3: An aqueous component mole of SO42- ions against pore volume injected

34
The component mole of Ca2+ ions in the aqueous phase also shows a decrease
when the diluted seawater injected, means that as the adsorbed mole of SO42- ions on rock
surface increases, because of the attractive forces between positively charged calcium and
negatively charged sulfate ions, the adsorbed mole of calcium ions also increases. Hence,
it also contributes to the desorption of heavy end fraction of the crude oil components,
leading to the wettability alteration towards more water wet. Due to the reason that Ca2+
ions in the aqueous phase have decreased, calcite and anhydrite dissolution happened in
the formation to keep the chemical equilibrium in the formation (Figure 4.5).

0.0007
Aqueous component mole of Ca2+

0.0006

0.0005
ions (gmole)

10 times diluted seawater


0.0004
5 times diluted seawater
Seawater
0.0003

0.0002

0.0001

0
0 500 1000 1500 2000 2500
Pore volume injected

Figure 4.4: An aqueous component mole of Ca2+ ions against pore volume injected

0
0 500 1000 1500 2000 2500
Mineral moles changes (gmole)

-0.00005

-0.0001
Dolomite

-0.00015 Calcite
Anhydrite

-0.0002

-0.00025
Pore volume injected
Figure 4.5: Mineral moles changes against pore volume

35
Initially, the carbonate core was in oil wetting state, and as we inject diluted
seawater into the formation, the wettability of the rock-fluid system turned out to be more
water wet. As it can be seen from Figure 4.6, the final relative permeability curve shifted
to the right compared to the initial relative permeability curve, indicating that residual oil
saturation decreased, and the system is more water wet.

1
0.9
Relative permeability

0.8
0.7
0.6 Kro, initial
0.5
Kro, final
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Water Saturation
Figure 4.6: Wettability alteration from oil wet towards more water wet by injecting diluted
seawater

4.2.1 Oil Saturation


It is worthwhile to analyse the oil saturation versus distance plots in order to
observe the impact of dilution of seawater as the displacement front progresses from
injection well towards production well. Since this case of the sensitivity gives the
optimum recovery amongst all the sensitivity runs in this section of the simulation studies,
oil saturation (So) changes with respect to distance away from the injection well towards
the production well has been reported of only this case. As it can be seen from Figure 4.7
that displacement front moves further away from injection well as we inject seawater into
the formation. Initially, So in the reservoir is uniform which is 0.9 (90%). As the low
salinity water injection is continued, it is reduced to 45% after 25 minutes. in the
formation next to the injection well. With the continued seawater injection, So is reduced
to 0.42, 0.4, 0.37 after 55 mins, 96 mins, 276 mins of injection time respectively. This
means that more and more oil is displaced away from the injection well towards the
production well from the reservoir, as depicted by the moving saturation front (black line).
After 276 minutes of injection of seawater, oil saturation in the production well reduced
to 50%.
36
After 25 mins

Displacement front

After 55 mins

After 96 mins

After 276 mins

Figure 4.7: Oil saturation versus distance plots after 25, 55, 96, and 276 minutes of injection of
seawater

37
After 25 mins
Displacement front

After 55 mins

After 96 mins

After 276 mins

Figure 4.8: Oil saturation versus distance plots after 25, 55, 96, and 276 minutes of injection of
5 times diluted seawater

38
Comparison of Figure 4.7 and 4.8 shows that the speed of the displacement front
of seawater and 5 times diluted seawater is almost same. However, oil saturation in the
formation near the injection well is 0.38 after 25 minutes of injection of 5 times diluted
seawater, which this value is 0.45 when seawater is used as an injection fluid. After 276
minutes of injection of 5 times diluted seawater, oil saturation reduced to 0.25.
Comparatively 0.38 So is left behind when seawater is used as an injection fluid. It can
be concluded that dilution of seawater increases the sweep performance with less residual
oil saturation is left behind. In order to better visualise the comparison of sweep
performance of seawater and 5 times diluted seawater, 2D plot of oil saturation across the
reservoir is shown in Figure 4.9.

Figure 4.9: 2D visualisation of oil saturation across the reservoir

4.3 LSWI case for carbonates with mineral relative permeability


interpolation
It can be seen from Figure 4.10 that there is not as much incremental oil recovery
observed in this case when compared with the low salinity water injection case for
carbonates with SO4 simple relative permeability interpolation. However, mineral moles

39
changes against pore volume injected graph as shown in Figure 4.11 shows that calcite
and anhydrite dissolution process happened in this case. Although mineral dissolution
happened in the formation, no substantial incremental oil recovery observed. Hence, it
can be concluded that the mineral dissolution process has no considerable impact while
injecting low saline water into the formation.

70

60
Oil recovery factor (%)

50
5 times diluted seawater
40
seawater
30

20

10

0
0 5 10 15 20 25 30 35 40
Pore volume injected

Figure 4.10: Impact of dilution of seawater on oil recovery (with mineral relative permeability
interpolation)

0
0 5 10 15 20 25 30 35 40
-0.00002

-0.00004
Mineral moles changes (gmole)

-0.00006

-0.00008

-0.0001 Dolomite
-0.00012 Calcite

-0.00014 Anhydrite

-0.00016

-0.00018

-0.0002
Pore volume injected

Figure 4.11: Mineral moles changes against pore volume (with mineral relative permeability
interpolation)

40
4.4 LSWI case with temperature effects
As it can be easily seen from Figure 4.12, as the temperature of the injection fluid
increases, the oil recovery increases. In this case, the oil recovery factor has increased
from 59 % to 65 %, as we increased the temperature of the injected fluid from 100 °C to
200 °C. As it can be seen from Figure 4.13 that concentration of SO42- ions in the aqueous
phase decreases as the temperature increases, implying that more of the SO42- ions are
adsorbed on rock surface, and lead to desorption of oil from rock surface.

70

60
Oil recovery factor (%)

50

40 100 °C
150 °C
30
200 °C
20

10

0
0 5 10 15 20 25 30 35 40
Pore volume injected

Figure 4.12: Effect of injection fluid temperature on oil recovery

9.00E-05
Aqueous component mole of SO42-

8.00E-05

7.00E-05

6.00E-05
ions (gmole)

5.00E-05

4.00E-05 200 °C

3.00E-05 100 °C

2.00E-05 150 °C

1.00E-05

0.00E+00
0 10 20 30 40
Pore volume injected

Figure 4.13: Effect of temperature on aqueous component mole of SO42- ions

41
4.5 LSWI case with ion exchange relative permeability interpolation
Because of CaCH3COO+ is present in the aqueous phase, the following reaction
happens in the aqueous phase:

CH3COO-X + 0.5 SO42- = CH3COO- + SO4-X

It is implying that an increase in the concentration of SO4-X and a decrease in the


concentration of CH3COO-X are expected (Figure 4.14), which means that more of the
sulfate ions are adsorbed on rock surface leading to desorption of carboxylic group from
rock surface, and the system turns out to be more water wet.
0.0006
Aqueous component moles (gmole)

0.0005

0.0004

CH3COO-X
0.0003
SO4-X
0.0002

0.0001

0
0 5 10 15 20 25 30 35 40
Pore volume injected

Figure 4.14: Aqueous component moles of CH3COO-X and SO4-X

As can be seen from Figure 4.15, when ion exchange relative permeability interpolation
was chosen as a low salinity modeling method in the process wizard, the oil recovery factor has
increased up to 80%.

90
80
Oil recovery factor (%)

70
60
50
40 seawater
30 5 times diluted seawater
20
10
0
0 5 10 15 20 25 30 35 40
Pore volume injected
Figure 4.15: Impact of dilution of seawater on oil recovery (with ion exchange relative
permeability interpolation)

42
Chapter 5: Conclusions

Low salinity waterflooding is a crucial enhanced oil recovery technique because


of the benefits it can bring compared to the chemical enhanced oil recovery techniques
regarding the cost of the injected chemicals and the environmental effects of it. In this
project, various low salinity modeling methods were deployed on the basis of brine
dilution to examine the impact on oil recovery and comprehend the underlying
responsible mechanism of low salinity water injection. The simulation model was
developed on the basis of a core-flooding experiment and validated for compatibility with
the experimental data.

In the first scenario, simple relative permeability interpolation of SO42- ions was
chosen as a low salinity modeling method, and three injection cycles were considered,
which firstly with seawater, then followed by five and ten times diluted seawater. Result
in Figure 4.2 shows that five times diluted seawater increased oil recovery factor up to
75%, however, following ten times diluted seawater did not increase oil recovery
substantially. Conclusion can be drawn that there is a threshold value of the salinity of
injected fluid which further decrease beyond that value does not give any considerable
incremental oil recovery. Decrease in the concentrations of SO42- and Ca2+ in the aqueous
phase means that more of the sulfate and calcium ions were adsorbed on the rock surface
and causing desorption of oil from rock surface (Figure 4.3 and Figure 4.4). Hence, we
can say that ion exchange is a responsible mechanism for incremental oil recovery due to
dilution of seawater and wettability alteration from oil wetness towards more water
wetness is a result of this process (Figure 4.6). Although calcite and anhydrite dissolution
happened in the formation (Figure 4.5), it cannot be concluded that mineral dissolution is
the responsible mechanism. Hence another scenario needs to be conducted to observe the
impact of mineral dissolution.

In the second scenario, mineral volume fraction relative permeability interpolation


was chosen as a low salinity modeling method. Although calcite and anhydrite dissolution
happened in the formation as shown in Figure 4.11, no considerable incremental oil
recovery was observed by implementing this scenario (Figure 4.10). It can be concluded
that mineral dissolution is not a responsible mechanism for low salinity waterflooding.

In the third scenario, the impact of the temperature of the injected fluid on oil
recovery was investigated, and the result has shown that as the temperature of the injected

43
fluid increase, the oil recovery factor increases (Figure 4.12). It can be concluded that the
temperature of the injected fluid is a contributing factor in low salinity waterflooding.

In the fourth scenario, ion exchange relative permeability interpolation was


chosen as a low salinity modeling method, and the result has shown an incremental oil
recovery factor of up to 80% by injecting five times diluted seawater (Figure 4.15).

Summary of the key results of this project is as follows:

• Multicomponent ionic exchange (MIE) is a responsible underlying mechanism of


low salinity water injection, and wettability alteration from oil wetness towards
more water wetness is a result of this process rather than being the reason of it.
• There is a threshold value for the salinity of the injected fluid which is regarded
as an optimum salinity, which further decrease in salinity does not give a
considerable incremental oil recovery.
• Mineral dissolution is not a responsible mechanism for incremental oil recovery
due to low salinity water injection.
• The temperature of the injected fluid is a contributing factor in low salinity
waterflooding.

44
References

[1] Schlumberger, Carbonate Reservoirs, Available at:


https://www.slb.com/~/media/Files/industry_challenges/carbonates/brochures/cb_c
arbonate_reservoirs_07os003.pdf

[2] Alotaibi, M.B., Azmy, R. and Nasr-El-Din, H.A., (2010). Wettability challenges in
carbonate reservoirs, SPE Improved Oil Recovery Symposium,

[3] Zhang, P. and Austad, T., (2006). Wettability and oil recovery from carbonates:
Effects of temperature and potential determining ions, Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 279(1), pp.179-187. Available at:
http://www.sciencedirect.com/science/article/pii/S0927775706000148

[4] Sheng, J., (2013). Enhanced Oil Recovery Field Case Studies. Amsterdam: London :
Elsevier/GPU.

[5] Derkani, M., Fletcher, A., Abdallah, W., Sauerer, B., Anderson, J. and Zhang, Z.,
(2018). Low salinity waterflooding in carbonate reservoirs: Review of interfacial
mechanisms, Colloids and Interfaces, 2(2),

[6] Shariatpanahi, S.F., Strand, S. and Austad, T., (2011). Initial Wetting Properties of
Carbonate Oil Reservoirs: Effect of the Temperature and Presence of Sulfate in
Formation Water, Energy Fuels, 25 (7), pp.3021-3028.

[7] Nasralla, R.A., Sergienko, E., Masalmeh, S.K., van, d.L., Brussee, N.J., Mahani, H.,
Suijkerbuijk, B. and Alqarshubi, I., (2014). Demonstrating the potential of low-
salinity waterflood to improve oil recovery in carbonate reservoirs by qualitative
coreflood, Abu Dhabi International Petroleum Exhibition and Conference,

[8] Standnes, D.C. and Austad, T., (2000). Wettability alteration in chalk: 1. preparation
of core material and oil properties, Journal of Petroleum Science and Engineering,
28(3), pp.111-121. Available at:
http://www.sciencedirect.com/science/article/pii/S0920410500000838
[9] Shimoyama, A. and Johns, W.D., (1972). Formation of alkanes from fatty acids in
the presence of CaCO3, Geochimica Et Cosmochimica Acta, 36(1), pp.87-91.
Available at: http://www.sciencedirect.com/science/article/pii/0016703772901226

[10] Strand, S., Høgnesen, E.J. and Austad, T., (2006). Wettability alteration of
carbonates—Effects of potential determining ions (Ca2+ and SO42−) and
temperature, Colloids and Surfaces A: Physicochemical and Engineering Aspects,
275(1), pp.1-10. Available at:
http://www.sciencedirect.com/science/article/pii/S092777570500796X

[11] Austad, T., Shariatpanahi, S.F., Strand, S., Black, C.J.J. and Webb, K.J., (2012).
Conditions for a low-salinity Enhanced Oil Recovery (EOR) effect in carbonate oil
reservoirs, Energy and Fuels, 26 (1), pp.569-575.

[12] Buckley, J.S., (1995). Asphaltene Precipitation and Crude Oil Wetting,

[13] Mahani, H., Keya, A.L., Berg, S., Bartels, W., Nasralla, R. and Rossen, W., (2015).
Driving mechanism of low salinity flooding in carbonate rocks, Europec 2015,

[14] Romanuka, J., Hofman, J., Ligthelm, D.J., Suijkerbuijk, B., Marcelis, F., Oedai, S.,
Brussee, N., van, d.L., Aksulu, H. and Austad, T., (2012). Low salinity EOR in
carbonates, SPE Improved Oil Recovery Symposium,

[15] Al-Attar, H., Mahmoud, M.Y., Zekri, A.Y., Almehaideb, R.A. and Ghannam,
M.T., (2013). Low salinity flooding in a selected carbonate reservoir: Experimental
approach, EAGE Annual Conference & Exhibition Incorporating SPE Europec,

[16] Zhang, P., Tweheyo, M.T. and Austad, T., (2006). Wettability Alteration and
Improved Oil Recovery in Chalk:  The Effect of Calcium in the Presence of
Sulfate, Energy Fuels, 20 (5), pp.2056-2062.

[17] Zhang, P., Tweheyo, M.T. and Austad, T., (2007). Wettability alteration and
improved oil recovery by spontaneous imbibition of seawater into chalk: Impact of
the potential determining ions Ca2+, Mg2+, and SO42−, Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 301(1), pp.199-208. Available at:
http://www.sciencedirect.com/science/article/pii/S0927775706009915
[18] Fathi, S.J., Austad, T. and Strand, S., (2010). “Smart Water― as a Wettability
Modifier in Chalk: The Effect of Salinity and Ionic Composition, Energy Fuels, 24
(4), pp.2514-2519.

[19] Yousef, A.A., Al-Saleh, S., Al-Kaabi, A. and Al-Jawfi, M., (2010). Laboratory
investigation of novel oil recovery method for carbonate reservoirs, Canadian
Unconventional Resources and International Petroleum Conference,

[20] Morrow, N.R., Ma, S., Zhou, X. and Zhang, X., (1994). Characterization of
wettability from spontaneous imbibition measurements, Annual Technical Meeting,

[21] Strand, S., Austad, T., Puntervold, T., Høgnesen, E.J., Olsen, M. and Barstad,
S.M.F., (2008). “Smart Water― for Oil Recovery from Fractured Limestone:
A Preliminary Study, Energy Fuels, 22 (5), pp.3126-3133.

[22] Yi, Z. and Sarma, H.K., (2012). Improving waterflood recovery efficiency in
carbonate reservoirs through salinity variations and ionic exchanges: A promising
low-cost "smart-waterflood" approach, Abu Dhabi International Petroleum
Conference and Exhibition,

[23] Zaeri, M.R., Hashemi, R., Shahverdi, H. and Sadeghi, M., (2018). Enhanced oil
recovery from carbonate reservoirs by spontaneous imbibition of low salinity
water, Petroleum Science,

[24] Gupta, R., Smith, G.G., Hu, L., Willingham, T., Lo Cascio, M., Shyeh, J.J. and
Harris, C.R., (2011). Enhanced waterflood for carbonate reservoirs - impact of
injection water composition, SPE Middle East Oil and Gas show and Conference,

[25] Yousef, A.A., Al-Saleh, S. and Al-Jawfi, M., (2012). Improved/enhanced oil
recovery from carbonate reservoirs by tuning injection water salinity and ionic
content, SPE Improved Oil Recovery Symposium,

[26] A. Hiorth, L. M. Cathles, J. Kolnes, O. Vikane, A. Lohne, R. I. Korsnes, M. V.


Madland, (2008). A CHEMICAL MODEL FOR THE SEAWATER-CO2-
CARBONATE SYSTEM-AQUEOUS AND SURFACE CHEMISTRY,
International Symposium of the Society of Core Analysts Held in Abu Dhabi, 12 p
[27] Nasralla, R.A., Sergienko, E., Masalmeh, S.K., van, d.L., Brussee, N.J., Mahani,
H., Suijkerbuijk, B. and Alqarshubi, I., (2014). Demonstrating the potential of low-
salinity waterflood to improve oil recovery in carbonate reservoirs by qualitative
coreflood, Abu Dhabi International Petroleum Exhibition and Conference,

[28] Mahani, H., Keya, A.L., Berg, S., Bartels, W., Nasralla, R. and Rossen, W.R.,
(2015). Insights into the Mechanism of Wettability Alteration by Low-Salinity
Flooding (LSF) in Carbonates, Energy Fuels, 29 (3), pp.1352-1367.

[29] Schembre, J.M., Tang, G.-. and Kovscek, A.R., (2006). Wettability alteration and
oil recovery by water imbibition at elevated temperatures, Journal of Petroleum
Science and Engineering, 52(1), pp.131-148. Available at:
http://www.sciencedirect.com/science/article/pii/S0920410506000623

[30] Tang, G. and Morrow, N.R., (1999). Influence of brine composition and fines
migration on crude oil/brine/rock interactions and oil recovery, Journal of
Petroleum Science and Engineering, 24(2), pp.99-111. Available at:
http://www.sciencedirect.com/science/article/pii/S0920410599000340

[31] Meng, W., Haroun, M.R., Sarma, H.K., Adeoye, J.T., Aras, P., Punjabi, S.,
Rahman, M.M. and Al Kobaisi, M., (2015). A novel approach of using phosphate-
spiked smart brines to alter wettability in mixed oil-wet carbonate reservoirs, Abu
Dhabi International Petroleum Exhibition and Conference,

[32] Ligthelm, D.J., Gronsveld, J., Hofman, J., Brussee, N., Marcelis, F. and van, d.L.,
(2009). Novel waterflooding strategy by manipulation of injection brine
composition, EUROPEC/EAGE Conference and Exhibition,

[33] Fathi, S.J., Austad, T. and Strand, S., (2011). Water-Based Enhanced Oil Recovery
(EOR) by “Smart Water―: Optimal Ionic Composition for EOR in
Carbonates, Energy Fuels, 25 (11), pp.5173-5179.

[34] Sohal, M.A., Thyne, G. and Søgaard, E.G., (2016). Review of Recovery
Mechanisms of Ionically Modified Waterflood in Carbonate Reservoirs, Energy
Fuels, 30 (3), pp.1904-1914.
[35] Alhammadi, M., Mahzari, P. and Sohrabi, M., (2017). Experimental investigation
of the underlying mechanism behind improved oil recovery by low salinity water
injection in carbonate reservoir rocks, Abu Dhabi International Petroleum
Exhibition & Conference,

[36] Sohrabi, M., Mahzari, P., Farzaneh, S.A., Mills, J.R., Tsolis, P. and Ireland, S.,
(2015). Novel insights into mechanisms of oil recovery by low salinity water
injection, SPE Middle East Oil & Gas show and Conference,

[37] Esene, C., Onalo, D., Zendehboudi, S., James, L., Aborig, A. and Butt, S., (2018).
Modeling investigation of low salinity water injection in sandstones and
carbonates: Effect of na+ and SO42−, Fuel, 232pp.362-373. Available at:
http://www.sciencedirect.com/science/article/pii/S0016236118310068

[38] Qiao, C., Johns, R. and Li, L., (2016). Modeling Low-Salinity Waterflooding in
Chalk and Limestone Reservoirs, Energy Fuels, 30 (2), pp.884-895.

[39] Al-Shalabi, E., Sepehrnoori, K. and Pope, G., (2015). Mechanistic modeling of oil
recovery due to low salinity water injection in oil reservoirs, SPE Middle East Oil
& Gas show and Conference,

[40] Yousef, A.A., Al-Saleh, S., Al-Kaabi, A. and Al-Jawfi, M., (2011). Laboratory
Investigation of the Impact of Injection-Water Salinity and Ionic Content on Oil
Recovery From Carbonate Reservoirs,

[41] Anonymous CMG (computer modelling group) tutorial,

[42] Awolayo, A.N., Sarma, H.K. and Nghiem, L.X., (2017). A comprehensive
geochemical-based approach at modeling and interpreting brine dilution in
carbonate reservoirs, SPE Reservoir Simulation Conference,

You might also like