You are on page 1of 20

SPE 129972

Wettability Challenges in Carbonate Reservoirs


M.B. Alotaibi, R.A. Nasralla, and H.A. Nasr-El-Din, Texas A&M University, all SPE members

Copyright 2010, Society of Petroleum Engineers

This paper was prepared for presentation at the 2010 SPE Improved Oil Recovery Symposium held in Tulsa, Oklahoma, USA, 24–28 April 2010.

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
Injection of water for pressure maintenance and waterflooding applications is becoming an increasingly important issue.
Historically, the total suspended solids of the water were given more consideration rather than the total dissolved solids. The
water from injection wells physically sweeps the displaced oil to adjacent production wells. Displacement of the oil is highly
affected by many interacting variables. Wettability has been recognized as one of the controlling parameters of the remaining
oil-in-place. Brine salinity, oil chemistry, and rock lithology are believed to play a significant role on altering rock
wettability. Recent studies highlted that low salinity water can increase oil recovery in carbonate reservoirs.
This paper highlights extensive wettability studies to determine the optimum brine salinity, which results in higher oil
recovery. Contact angle and Amott imbibition methods are considered as the most common methods to measure the
preferential affinity of reservoir rocks. Crude oil, formation core samples (dolomite, and calcite), and synthetic brines from
Middle East (formation, aquifer, and seawater) are used to evaluate wettability quantitatively, and qualitatively. All
experiments were conducted at high pressure (up to 2,000 psia), and elevated temperature (up to 270°F).\
Contact angle decreased with pressure, and temperature upto 194°F. The optimum salinity is found that can enhance the
rock wettability toward water-wet. Currently, there is confusion concerning the optimum brine salinity for injection wells.
The results of this study give a new insight on the optimum salinity and provide better understanding of some wettability
challenges in carbonate reservoirs.

Introduction
Carbonate rocks (dolomite and limestone) account for more than half of the world's hydrocarbon reserves. Understanding the
important mechanisms which control recovery and retention in a given reservoir is essential for successful oil production.
Formation wettability is considered as an important factor that controls fluids distribution in the reservoir. The wettability
preference, towards oil or water wet, influences many aspects of reservoir performance. Making incorrect assumption about
reservoir wettability might result in unexpected formation damage. Reservoir rocks are complex structures of a variety of
minerals. For this reason, it is extremely difficult to describe the wetting character of reservoir rocks. Most petrophysical
properties such as capillary pressure, relative permeability, electrical properties, and waterflood behavior are affected by the
wettability.
Most carbonate reservoirs are believed to have mixed wettability or to be oil-wet. It is also recognized that recovery
factors are higher for sandstone reservoirs than for carbonates. Carbonate reservoirs present a number of specific
characteristics posing complex challenges in reservoir characterization, production and management. The contribution of this
paper will be mainly wettability alteration in carbonate rocks using different salinity brines. This study has two main
objectives:
1) Determine wettability behavior of carbonate rocks by using contact angle and spontaneous imbibtion methods at
high pressure and elevated temperature.
2) Forced fluid displacements through limestone and dolomite cores by using core flood apparatus.

Adjusting the injection water ions can impact rock wettability and thus may enhance the oil recovery (Høgnesen et al.
2005). Høgnesen et al. focused an imbibtion study at high temperatures using reservoir limestone, outcrop chalk cores,
seawater and formation water. Increasing sulphate ion concentration at high temperatures proved to be a major factor in
recovering additional oil. Possibility of scale and sour problems is expected as increasing the sulphate concentration.
Moreover, this strategy has limitations with regard to the initial brine salinity and temperature (Webb et al. 2005).
2 SPE 129972

Most of the waterflood studies were conducted on limestone; seawater, also, was recommended as an injection fluid in
chalk formation. Strand et al. (2008) stated in preliminary experimental studies the mechanism for the wettability alteration
in fractured limestone after injecting seawater, sodium chloride brine, and formation water. Synthetic seawater with and
without sulphate ions was used to determine the sulphate effect on wettability. Spontaneous imbibitions results using
seawater at 248°F showed 15% increase in oil compared to seawater free of sulphate ions. Seawater has the lowest TDS
compared to the other brines. The main mechanisms to improve recovery in oil-wet formations are: wettability alteration and
interfacial tension reduction.
Spontaneous imbibition is the most important phenomenon in oil recovery, especially in fractured reservoirs. Imbibition
is also essential in evaluation of the rock wettability (Morrow et al. 1994). The rate of imbibition is usually a function of
porous media and fluid properties such as absolute and relative permeability, viscosity, interfacial tension, and wettability
(Zhang et al. 1996).

Literature Review
Wettability Alteration Mechanism. Wettability alteration is the main challenge for enhancing oil recovery in carbonate
reservoir. Strand et al. (2008) showed the effect of injection water ions (calcium, magnesium and sulphate) on oil recovery.
Activation energy for chemical reactions is important and required for any wettability improvement. In general, bonding
energy between polar components in oil and carbonates is higher than what is observed in sandstones. Carbonate rock is
neutral to preferentially oil-wet, because of adsorbing the carboxylic components in crude oil onto the carbonate surface.
Sulphate ion can act as a wettability modifier alone without other additives, such as surfactants. Sulphate is believed to be a
very strong potential determining ion towards calcium carbonate (Pierre et al. 1990; Strand et al. 2003; Strand and Austad
2008).
Imbibtion tests using seawater showed that the ions concentration (calcium and sulphate) and temperature appeared to be
crucial in wettability modification. Moreover, adsorption of sulphate increased as the concentration of calcium in seawater
increased due to co-adsorption of calcium on the carbonate surface. Adsorption of sulphate onto carbonate decreased the
positive charge on rock surface, which increased the excess of calcium at the surface because of reduced electrostatic
repulsion (Strand et al. 2008; Strand et al. 2006; Austad et al. 2009).
Adsorption of sulphate onto chalk surface facilitates desorption of negatively charged carboxylic materials by changing
the surface charge of the chalk surface (Strand et al. 2003). The imbibition rate and oil recovery increase as the temperature
increases due to a stronger adsorption of sulphate and calcium onto the chalk surface. Magnesium ions adsorbs less strongly
than calcium onto the chalk surface at low temperatures (Zhang and Austad 2006; Zhang et al. 2007). At high temperatures,
magnesium substitutes calcium at the chalk surface and the degree of substitution increases as the temperature increases.
Magnesium becomes more reactive because of dehydration, and that is the reason for substituting calcium in the surface
lattice of chalk. The presence of calcium, magnesium, and sulphate ions is necessary to increase the spontaneous imbibition
of water and changes the rock wettability. Limestone rocks also showed similar interactions behavior with seawater.
Lichaa et al. (1992) studied the wettability changes of carbonate reservoir rocks during stages of preserved as received
from the field, cleaned, and restored cores. USBM, Amott, and contact angle methods were used to evaluate the rock/fluid
interaction. In brine/crude oil/rock system, the salinity and pH can strongly affect the surface charges on the rock surface and
fluid interfaces, which in turn can affect the wettability. The presence of large quantities of multivalent ions, Ca+2, Mg+2, Sr+2,
in the injection seawater and formation brines, and the weak base characteristic of the reservoir rock, suggest that a
preferentially oil-wet system should prevail in the presence of polar compounds in the crude oil. The pH of the brine should
affect the wetting character, particularly where the zeta potential crosses the zero point of charge.

Contact Angle Method. Contact angle is a function of the IFT at the solid/liquid and liquid/liquid interfaces. Wettability
of a reservoir rock is a manifestation of the thermodynamic equilibrium between fluids within the pores and the mineral
surfaces of the pore walls. Therefore, temperature, pressure, and fluids characteristics are strongly believed to have an effect
on the wettability. Roughness and heterogeneity of the solid surface will affect the contact angle and contribute to what is
commonly known by hysteries. A contact angle of 180º implies complete oil-wet characteristics, while 0º means complete
water-wet. Wettability classifications, in terms of contact angle, was well classified by Anderson (1986) as water-wet
(0─75º), intermediate (75─115º), and oil-wet (115─180º). Weakly water-wet and oil-wet conditions can be also represented
by (55─75º) and (115─135º), respectively.
Hjelmeland and Larondo (1986) studied the effect of pressure, temperature, and oil composition on the wettability of
calcium carbonate rocks. Temperature was found to have a large effect on the wetting characteristics. At low temperatures
(72°F), the solid phase exhibited oil-wet behavior, whereas at high temperatures (≥ 140°F), it exhibited water-wet behavior.
At 104°F an intermediate state of wettability seemed to prevail. No pressure effect was found on wettability. The wettability
of calcium carbonate, on the other hand, was not affected by the light fraction of the oil. Advancing and receding equilibrium
contact angles were obtained within 100 to 200 hours.
Saner et al. (1991) studied a carbonate reservoir wettability using USBM, Amott, and contact angle methods. Dead
crude oil and synthetic brines varying in salinity (20 to 200 kppm) were tested under elevated temperature and pressure
conditions. Increasing the temperature from ambient to 158°F lowered the contact angle from neutral-wet to moderately
water-wetting condition. Contact angle at ambient condition, also, decreased from 61° to 42° as the salinity was increased
SPE 129972 3

from 20 to 200 kppm. Yet under elevated temperature conditions, low salinity did not demonstrate any significant variation in
the contact angle (32° versus 28°). Pressure was varied from 20 to 2,800 psia at constant temperature (158°F). Results clearly
indicated that pressure did not affect the wettability. At similar temperature condition, salinity effect was almost negligible.
Saudi Arabian carbonate reservoir wettability was evaluated and compared by Lichaa et al. (1992) using USBM, Amott,
and contact angle techniques. Calcite, marble, and formation rocks were used in the receding contact angle measurements, as
well as synthetic formation brine, seawater, and dead crude oil. A wide temperature range was examined in all experiments
between 77 to 194°F. Test pressure was varied from ambient to 50 psi. Calcite surface tests became preferentially more
water-wet at higher temperatures. Contact angle results for oil/brine/marble system was slightly oil-wet to an intermediate
wettability, and tended to became weakly water-wet at higher temperatures. Formation rock tests, on the other hand, showed
an intermediate wettability or preferentially, slightly oil-wet at room temperature and became preferentially less oil-wet at
higher temperatures.
Wang and Gupta (1995) investigated the influence of temperature and pressure on reservoir rocks wettability. They used
stock-tank crude oil and reservoir brine from a carbonate reservoir. Contact angle for calcite system was not sensitive to
pressure (less than 5% increase for 3,000 psig increase in pressure). Wettability of calcite surfaces changed toward weakly
water-wet as temperature increased from 72.5 to 175°F. This change in wettability with temperature might due to changing
the fluids chemistry at the interface.
The effect of salinity on the contact angle was investigated by Almehaideb et al. (2004). Limestone rock, crude oil, NaCl
solutions were all used in their study. Four runs were examined using distilled water, 1,000, 10,000, and 50,000 ppm. All
experiments were conducted at room temperature. A significant reduction of oil/water contact angle was observed at 10,000
ppm.
Wettability alteration of chalk rocks by sulphate containing water were determined by Yu et al. (2007). They reported
contact angle measurements on both calcite crystal and chalk slices at high temperatures (up to 266°F). Increasing the
temperature up to 194°F changed the calcite crystal wettability towards water-wet due to accelerate desorption of stearic acid
from the calcite surface for all fluid systems investigated. The authors concluded a decrease in the contact angle after
replacing distilled water with sulphate containing water. Moreover, the contact angle decreased more at higher temperatures
(266°F) when exposed to sulphate water.
Yang et al. (2008) presented the wettability of the crude oil−reservoir brine−reservoir rock system at elevated
temperatures, using the axisymmetric drop shape analysis (ADSA) technique. Vuggy limestone rocks were used in this study
with intermediate wettability. The contact angle increased as pressure increased. Slight fluctuation was reported in the contact
angle measurement at pressure and temperature of 29 psi and 80.6°F, respectively. This contact angle fluctuation may be
ascribed to be strong electrostatic interactions between the crude oil and the reservoir brine. The contact angle also decreased
as increasing the fluid temperature.
Hamouda and Karoussi (2008) measured the advance and recede contact angles as a function of temperature for modified
calcite surfaces with 0.005 M stearic acid dissolved in decane. This measurement was conducted in a water medium at a
maximum temperature of 194°F. Contact angle decreased with increasing the temperature; indicating that the calcite surface
is becoming more water-wet as a function of temperature. The authors stated that this phenomenon is based on the total
interaction potential, which consists of van der Waals attractive and short-range Born repulsive and double layer electrostatic
forces. The fluid/rock interaction was shown to be dominated by the repulsive forces above 176°F, hence increased fine
detachment enhancing oil trapping.
The objectives of this study are to: (1) examine wettability of carbonate rocks over a wide range of salinity: formation
brine, seawater, and aquifer water, and (2) assess the impact of sulphate ions on the contact angle. Amott test was used as
another method to describe the optimum salinity fluid, which may enhance the oil recovery at HTHP conditions.

Experimental Studies
Porous Media. We used pink desert limestone, and Silurian dolomite for the spontaneous imbibtion and coreflood
experiments. The limestone rocks were drilled from Edwards Plateau outcrop formation in west-central Texas, USA. These
kinds of rock are soft, porous, and homogenous. It, also, has a permeability range from 90 to 110 md, and 30% of porosity.
Dolomite rock, on the other hand, is very hard, and has some tiny fractures. Dolomite rocks were delivered from Thornton in
Chicago. Permeability is usually less than 200 md, while porosity is below 20 vol%. X-ray powder diffraction technique
(XRD) was utilized in determining the minerarlogyy of the rock samples. The height of the peaks (intensity) for both samples
matched limestone and dolomite structures in the database, Figs. 1&2.
Calcite crystals and limestone samples were utilized for the contact angle tests. Iceland spar calcite crystals were ordered
from WARD’s Natural Science in small specimens (1 in. x 1 in.). The limestone rocks had permeability less than 0.01 md.

Fluids. Synthetic brines, from the Middle East, were used in this study: 230 kppm formation brine, 54 kppm seawater, 4
kppm aquifer water, and deionized water. Table 1 shows the brines composition. Two stock tank sweet (no H2S) and sour
crude oil samples were used in the contact angle study. Amott tests and core flood experiments were conducted utilizing the
sweet oil. Crude oil (sweet) was sampled all from one well at field A. Its density was determined at high temperature (194ºF)
which equal to 0.87 g/cm3. The sour crude oil was received from West Texas field, USA. Oil density was 0.79 at 194 ºF.
Capillary viscometer was used to determine oil viscosity at ambient temperature. The viscosity of the sour crude oil was 3.35
4 SPE 129972

cP and that of the sweet oil was 7.2 cP. Both crude oil samples were filtered through 1 µm filter paper, and then re-filtered
through core plug to remove solid particles that may cause plugging problems.

Apparatus and Procedure


Drop Shape Analysis System (DSA)
The Pendant drop method is well known to determine interfacial property from the drop shape that generated inside a view
chamber (Andreas et al. 1938). The results were subjected to fairly high errors due to analogous photographing images and
an empirical evaluation method. Electronic data processing is developed lately to digitalize drop images and solve the
theoretical equation of a drop profile. Drop Shape Analysis System (DSA) is a technique to determine liquid-fluid interfacial
tensions and contact angles from the shape of axisymmetric menisci, i.e., from sessile as well as pendent drops (Li and
Neumann 1992), Fig. 3.

Interfacial Tension
Drop Shape Analysis technique for the pendant drop is probably the most advanced and accurate method for measuring the
IFT in a large range of pressures and temperatures. In the experiment, a pendant oil drop is formed at the tip of a stainless
steel capillary needle. Then, digital image of the pendant drop is acquired by an image acquisition system. By applying
digital computer image analysis and processing techniques, the accurate interfacial profile of the pendant drop can be
obtained. Finally, the IFT of the oil drop is determined by solving the Laplace equation of capillarity and finding the best fit
of the numerically obtained interfacial profile to the physically observed drop profile.
The maximum operating pressure and temperature are 10,000 psi and 392°F, respectively. Pressure was applied using
compressed nitrogen, and temperature was increased gradually using a digital temperature control. Three drops were usually
evaluated at each temperature. Vibration and temperature variations are the most common sources of errors when dealing
with small drops. Ideally, the pendent drop apparatus should be mounted on a vibration-free base. More details on IFT were
given by Alotaibi and Nasr-El-Din (2009).

Contact Angle
The application of ‘Drop Shape Analysis’ to sessile and captive drop contact angles requires the solid surface to be smooth
and homogenous so as to ensure that the drop is axisymmetric. The angle between the baseline and the tangent at the drop
boundary is measured.

Rock Substrates
Calcite crystals and Laurda limestone were prepared for all contact angle tests using the following procedure:
• The substrates were cut from small rock piece with the dimensions of 0.62 in. x 0.72 in. x 0.25 in.
• One slide of the rock substrate is polished with sand paper (600 meshes and then 300 meshes) to minimize the
contact angle hysteresis due to surface roughness.
• Load all substrates into an empty glass flask, and then apply vacuum pressure for at least two hrs.
• Introduce formation brine (230 kppm) to the substrates and keep them soaked in a vacuum for at least 3 hrs. To
remove the contaminants and surface charges induced by polishing, the substrates left in brine for at least 24 hrs.
• Rock substrates were then placed in crude oil and centrifuged at 3,000 rpm for 30 min.
• Wait for 30 min. before centrifuging the rock substrates again at the same speed and time period.
• The centrifuge step was to displace the water droplets on the rock surface to keep only the irreducible water (Swi)
• Some samples were aged in oil at 194°F for different aging periods
• The above procedure was followed to simulate the in-situ state of the mineral surface under reservoir condition.
• Soak the substrates in toluene for 2 seconds to remove the bulk oil from the rock surface.

A very important step before any experiment is a thorough cleaning of the apparatus, because trace amounts of
contamination can alter the results. Flow line was cleaned first with hexane several times. Acetone was then used to flush the
lines. The line was dried by flowing dry air and was flushed with copious amounts of deionized water. The system was tested
for leakage with deionized water before each experiment. The prepared rock substrate was placed inside the chamber with its
holder. Then, each brine was introduced into the cell until completely covering the rock substrate. Nitrogen gas was slowly
injected into the brine to pressurize the system at 2,000 psi. Temperature was controlled manually to the set point (122, 194,
and 266°F). The rock substrate was equilibrated with brines for at least 20 h. Captive drop of crude oil was formed on the
rock substrate at the previous condition. To an extent, the quality of the measurements relies very much on the skill of the
experimentalist and the heterogeneity of the rock substrates. In this paper, only advanced contact angle is reported.

Modified HTHP Amott Cell


Amott cell was modified for high temperature/high pressure conditions. A steel cell with Teflon coating was utilized for high
salinity fluids up to 230 kppm. The cell was pressurized from 100 to 120 psi inside an electrical oven, after connecting the
inlet and outlet 1/8 in. tubing to the modified cell. The oven temperature was set at 194°F. Temperature was measured using
SPE 129972 5

digital thermometer to ensure accuracy. A syringe pump was operated to inject brines at constant pressure. After that, an
outlet valve was released carefully to collect the fluids in a glass burette. It has an accuracy of 0.1 cm3 and that to record oil
volume more accurately. A schematic diagram of the modified Amott Cell is shown in Fig. 4.

Coreflood Setup
Coreflood equipment mainly consists of three major parts: coreholder, accumulators, and two pumps as shown in Fig. 5. An
oven was utilized to heat the cores and fluids at high pressure condition. The tubing was coiled inside the oven to heat the
fluids before any contact with the core. Fluids displacement was achieved by operating the syringe pump at only constant
rates. Injection fluids were stored in stainless steel accumulators (2 L) at room temperature. A hydraulic pump was utilized to
apply 1,000 psi of overburden pressure. Two back pressure regulators were installed to regulate the overburden pressure and
core outlet flow at 1,000 and 500 psi, respectively. This is to control the flow and avoid undesirable higher pressure after
heating the system. Differential pressure transducer (300 psi) was used to measure pressure drop between core inlet and
outlet.

Results and Discussion


Interfacial Tension (IFT). Interfacial tension of a brine droplet in oil was determined at high pressure and elevated
temperatures. On the other hand, IFT values of an oil droplet in brine solutions were also measured. In the first case, the brine
droplet was injected from top to bottom. For that reason, the shape of the droplet was pendant. In the second system, the
shape of the oil droplet was more spherical. Gravity plays a major role on the equatorial or maximum horizontal diameter as
well as the drop shape factor (Alotaibi and Nasr-El-Din, 2009). However, IFT of both cases were too close (oil/seawter: 9.67
mN/m, and seawater/oil: 12.9 mN/m).
In brine/oil system, we filled a small glass cuvette with crude oil. Pressure was then applied gradually using compressed
N2 gas. The system was heated to 122°F. A few seawater droplets were then released through 1/32 inch capillary tube to
avoid any trapped air bubbles in the tubing. Seawater droplet was suspended in oil phase before recording IFT measurements,
Fig. 6.
Fig. 7 detailed seawater/crude oil IFT as a function of pressure and time. Equibrium at the interface was observed after 15
min. Results also suggested direction relationship between IFT and pressure. Increasing the temperature to 194°F at 200 psi
reduced the IFT from 23.01 to 16.3 mN/m.

IFT of Sour Crude Oil/Brines. Interfacial properties of formation brine, seawater, aquifer water and fresh water were
studied with sour crude oil. Formation brine IFT decreased to 14.6 mN/m as increasing the temperature to 194°F. In crude
oil/sweater system, IFT decreased to 9.7 mN/m at 122°F and 2,000 psi. Interfacial tension decreased by almost 5 mN/m as a
result of decreasing the salinity. Reversing the systems produced very close IFT results (oil/seawater: 9.67 mN/m, and
seawater/oil: 12.9 mN/m).
Interfacial tension of aquifer water system was determined at 77, 194, and 266°F. Pressure was maintained constant at
2,000 psi for the three tests. Interfacial tension decreased from 12.8 to 9.2 mN/m as increasing temperature from 77 to 194°F.
At 266°F, interfacial tension increased slightly to 11.4 mN/m.
Interfacial tension of DI water system was similar to aquifer test where IFT values were 11.4 and 10.3 at 77 and 194°F,
respectively. More details on salinity and brines dilution effect on IFT were mentioned in previous publication (Alotaibi and
Nasr-El-Din 2009). Interfacial tension of sweet crude oil in seawater was studied at high pressure and elevated pressure by
Alotaibi et al. (2010)

Contact Angle of Brine/West Texas Crude Oil/Calcite. Rock substrate was soaked in high salinity formation brine
overnight at 122°F and 200 psi. Right, and left contact angles, aging time, as well as fitting errors are always recorded with
time for all experiments. An oil droplet was placed on the rock surface in the presence of formation brine. The left angle was
stable with time at a mean value of 99.8 ± 0.6°, and that classified the surface as an intermediate wet. Likewise, the mean
value for the right angle was 101.7 ± 3.27°. Maximum contact angles for right (R) and left (L) sides were 100.8 and 110.3°,
respectively. Moreover, the minimum R and L angles were 93.3° and 98.4°. In short, R and L angles results were matching.
In this test, we did not age the rock substrate.
The second test was conducted using formation brine at 194°F and 2,000 psi. Maximum R and L angles were 100.8 and
104.7°, while minimum values were 95 and 97.2°, respectively. After 80 min. contact time, calcite wettability represented
intermediate condition. Oil droplet was aged for 23 h at the same conditions. Contact angles were calculated at different time
intervals during the aging process. Contact angles results fluctuated slightly up to 10° as shown in Fig. 8.
Light and heavy components in crude oil significantly affect reservoir wettability. Aging time also change wettability
toward oil-wet. So, we aged all next substrates at 194°F after finishing the centrifuge step. At the beginning, samples were
aged for 24 h. For comparison, formation brine was examined to determine if contact angles were changing with time or not
after aging. Fig. 9 shows an increase in contact angles up to 120 and 128°. No change in angles with contact time, for that
reason we stopped the experiment after 14 h. In general, adding divalent cations into brine solutions resulted in more positive
charges on the calcite surface. The additional charges acted as attractive sites for the negative ends of polar components in
crude oil (Xu et al. 2006). This mechanism explained the oil-wet characteristics of calcite surface with formation brine.
6 SPE 129972

Injection pressure plays an important role in oil recovery. To determine the effect of pressure on wettability, we decreased
pressure from 2,000 to 1,500 psi. Then, a new oil droplet was introduced to the previous droplet. This oil advancement made
new contact points with calcite surface. Contact angles decreased 8 to 10° as decreasing the pressure from 2,000 to 1,500 psi.
Angles with contact time showed more stability than before as shown in Fig. 10.
Ionic interactions at the oil/water and solid/water affect wettability when the oil is stable (Xu et al. 2006). Brine
composition influences the adsorption asphaltenes on the rock surface. As a result, rock surface charges will be affected.
Seawater test was conducted at 194°F and 2,000 psi. The simulated oilfield in Middle East has intermediate wettability.
Therefore, aging time was fixed at one day for all remaining tests since intermediate and weak oil-wet conditions were
achieved. Mean values for L and R angles were 94.3 ± 0.46° and 98.6 ± 0.31°, respectively. Seawater has lower Ca, Mg, and
Na concentrations and higher concentration of sulphate than formation brine. Seawater altered calcite wettability toward
intermediate condition. Contact angles increased by at least ten degrees to 109 and 111° as decreeing pressure to 1,500 psi.
The surface wettability is still within intermediate range.
Pressure was decreased gradually to 500 psi, and angles were measured with time. Right angles increased significantly
toward oil-wet (130°), while slight change was observed on the left angle. At 122°F and 200 psi, new oil droplet was placed
on the rock surface. Left and right angles suggested identical results where the mean values were 142.7 ± 0.34° and 142.1 ±
0.65°. At low pressure and temperature conditions, seawater altered the surface wettability toward strong oil wet. Increasing
the pressure to 500 psi showed no effect on both sides’ angles. Contact angles maintained stability at 136.3 ± 0.47°, and
144.7 ± 0.51° as increasing pressure to 2,000 psi, Fig. 11. In summary, pressure effect on contact angles is unpredictable, and
gave us unclear trend. For that reason, all other tests were conducted only at 2,000 psi, and different temperature and fluids’
salinity.
Seawater, as discussed in the introduction, altered the carbonate wettability at high temperatures (> 212.5°F). Therefore,
we conducted seawater test at high temperature and pressure conditions (270.5°F and 2,000 psi). Left angle significantly
altered toward oil-wet (132.7 ± 0.35°), but we noticed no change on R angle (106.6 ± 2.45°). In brief, contact angles behavior
is oil-wet at low and high temperatures (122 and 270.5°F), and intermediate at 194°F. This conclusion is in agreement with
those of Hjelmeland and Larondo (1986).
Aquifer water usually re-injected into the oil reservoir to maintain pressure and increase oil recovery. Aquifer water has
lower Ca, Mg, Na, and SO4 concentrations than seawater. Sulphate concentration in the aquifer water was 7 times higher
formation brine. At 122°F, the aquifer water produced weakly water-wet surface. The results of L and R angles were 66.6 ±
1.44°, and 65.8 ± 2.10°. Fluids/rock temperature increased to 194°F at 2,000 psi. Contact angles were measured every 20
min. within time duration of 24 h. Contact angles results of L and R sides were 72.1 ± 1.11°, and 73.8 ± 1.50°. At higher
temperature (266°F), aquifer water showed intermediate wetting characteristics.
To determine the effect of sulphate ion on wettability without any interference from the divalent cations, we conducted
three sodium sulphate tests: 1) 3,560 ppm, which represented SO4 concentration in seawater, 2) 1,780 ppm, and 3) 890 ppm.
Adsorption of sulphate ion onto the carbonate surface will definitely decreased the surface positive charges. Thus, might alter
the rock wettability from oil-wet to either water or intermediate. In the first test, angles were measured at 122°F and 2,000
psi. Surface wettability reported to be water-wet, where L and R angles were 57.6 ± 2.11° and 57.3 ± 2.55°. Increasing the
temperature increased the contact angles toward strong water-wet condition. Contact time for this test was 21 h.
In the second test, we diluted the first sample 50% using DI water. The objective was to determine if lowering sulphate
concentration affects wettability or not. Similar temperature and pressure conditions were selected as we did in the first test.
At 122°F, L and R angles results showed 39.9 ± 2.15° and 38.4 ± 0.70°. Rock wettability remained within water-wet
condition, but higher angles, as we increased temperature to 194 and 266°F. The time duration of that test was 24 h.
The third sample which has sulphate concentration of 890 ppm showed unexpected behavior. The contact angle at 122F
increased within intermediate-wet range. Contact angles of L and R sides were reported to be 74.3 ± 8.29°, and 87.7 ± 9.37°.
Surface wettability was advanced toward intermediate wet even at higher temperatures. In summary, optimum sulphate
concentration does exist to decrease contact angles.
The de-ionized water test showed water-wetting characteristics (47.1 ± 3.41° and 58.0 ± 1.65°) at 122°F and 2,000 psi.
These angles indicated crude oil cannot penetrate the wetting water film without ionic interactions in brine. Identical contact
angles were observed as increasing temperature to 194°F. Results suggested R and L mean angles of 49.5 ± 0.88° and 47.8 ±
1.58°. Thermal gradient effect on contact angles was closely measured as increasing temperature from 194 to 266°F. Results
represented angles at different temperature as shown in Fig. 12. Within 25 min., angles reached equilibrium at 55 to 56°.
All in all, calcite wettability showed intermediate-wet surface before the aging process. We noticed this behavior even at
high temperature condition. Wettability was modified toward oils after one day of aging at 194°F. Contact angles decreased
as decreasing the pressure from 2,000 to 1500 psi. In comparison to formation brine, seawater altered the surface wettability
toward intermediate at 2,000 psi and 194°F. We observed inverse relationship between contact angles and pressure. Seawater
adjusted the wettability toward intermediate condition except at 500 psi. A strongly oil-wet surface was also reported as
increasing the temperature to 270°F.
Aquifer water adjusted rock’s wettability to water-wet at low and high temperatures. In addition, deionized water changed
the wettability towards weakly water-wet. Ions interaction expected to play significant effect on changing wettability to oil or
intermediate conditions. Water film showed stability where free ions water showed no interaction with rock and oil. Calcite
surface has positive charges, and sulphate ions in the examined brines will adsorbed on the rock surface. So, three sulphaet
SPE 129972 7

solutions were prepared to determine if that has some clear trend with concentration or not. Optimum sulphate concentration
showed to be 1,780 ppm which resulted in the lowest contact angles. All angles are shown in Table 3. Images for calcite
crystals at different temperature were also described in Fig. 13.

Contact Angle of Brine/Sweet Crude Oil/Laurda Limestone. We used Laurda limestone rocks and sweet crude oil to
study brines’ ionic strength effect on contact angles and surfaces wettability. In addition to de-ionized water, three brines
were examined at HTHP conditions: 1) formation brine (230 kppm), 2) seawater (54,680 ppm), and 3) aquifer water (5,436
ppm).
Limestone substrates were prepared and aged in crude oil for one day at 194°F and atmospheric pressure. Then, the
substrate was soaked in formation brine for another day to reach ionic equilibrium. At 122°F and 2,000 psi, formation brine
showed intermediate wetting characteristics with L and R angles of 94.7 ± 2.09° and 92.7 ± 3.06°, respectively. Contact
angles dropped toward weak water-wet as increasing temperature to 194 and 266°F. Rock roughness and heterogeneity can
affect contact angles measurements.
Crude oil/seawater/limestone system showed stable contact angles at 122°F and 2,000 psi. Seawater as discussed in
previous test altered rock wettability toward water-wet. Seawater decreased contact angles toward water-wet although we
used different crude oil and rocks samples. Surface contact angles decreased as increasing temperature to 194°F. At 266°F,
slight contact angles were observed. Measurements were continued for 10 h to ensure equilibrium was achieved
Aquifer water and seawater showed similar wettability results at low and high temperature conditions. Contact angles at
122°F were 63.0 ± 0.81° and 60.5 ± 1.12°. Contact angle decreased upto 8° as increasing temperature from 122 to 194°F.
Increasing temperature to 266°F decreased the contact angles few degrees.
De-ionized water run was conducted and the angles results showed weak water-wet condition. Left and right angles
reported to be 68.2 ± 1.25°, and 64.4 ± 0.62° for contact time of 24 h. Advance contact angle was measured by increasing the
oil droplet volume. Identical contact angles to previous results were noticed. Good match was achieved for L and R angles
(65.7 ± 2.24° and 64.1 ± 0.71°). This gave good reproducibility measurements for DSA instrument. Higher temperature test
decreased the contact angles up to 5°.
Using formation brine produced an intermediate wet surface. Aquifer water and seawater altered the wettability toward
water-wet and gave similar contact angle results. De-ionized water improved the wettability to be more water-wet at 194 and
266°F. Therefore, the lowest contact angle results were observed for the aquifer and seawater systems as illustrated in Table
3 and Fig. 14.

Spontaneous Imbibition. Pink desert limestone and dolomite cores were air vacuumed in a closed PVC pipe for at least 8
h. Then, we introduced filtered formation brine (230 kppm) from bottom to top direction under vacuum. Cores weight,
dimensions, porosity, permeability, and all others properties are summarized in Table 2. The base permeability was measured
using formation brine at three different flow rates (0.1, 0.5 and 1 cm3/min.) and ambient temperature condition. Pressure drop
was plotted versus flow rates to determine the base permeability using Darcy’s law, Fig. 15. Formation brine viscosity was
determined using capillary viscometer method at similar temperature condition (1.85 cP).
Sweet oil was used to saturate the cores at slow flow rates (0.1 to 1 cm3/min.) for 3 PV. Irreducible water saturation results
were then determined, Table 2. Reversing the flow direction was applied for the first limestone core and no extra water was
produced. Other cores were flooded with oil only in one direction. Pressure drop during the oil flood showed in Fig. 16.
Dolomite and limestone cores were aged in closed glass jars for 2 weeks at temperature of 194°F. Two dolomite cores
were utilized for Amott test at 194°F and 100 to 120 psi to prevent fluid from vaporization. The outlet flow was collected in a
glass burette with time. Fig. 17 showed the oil recovery from core plug no. 3 versus time. The induction time, where oil
started to be produced, was 48 min. Formation brine spontaneous imbibition test produced 19.5% oil. The cell pressure was
released at ambient temperature. We refilled the cell with seawater. Seawater increased the oil recovery up to 23.8%. The
incremental recovery was 3.8% as switched from formation brine to seawater. After 7 days, we drained seawater and filled
the cell with aquifer water. Spontaneous imbibition using aquifer water increased oil recovery only 1.1%.
Spontaneous imbibition test was conducted for dolomite core plug (no. 4) using formation brine. After five days, oil
recovery was 27.85%. The Amott cell was refilled with aquifer water at ambient conditions. No more oil was recovered for
13 days. It might need more time at ambient conditions to produce more oil. The core sample was returned back to the oven
at 194°F. Aquifer water showed 1.2% increase in oil recovery at 194°F, Fig. 18.

Coreflood Results. We run four coreflood experiments at temperature of 194°F. Overburden and back pressures were
applied at 1,000 and 500 psi, respectively. Pink desert limestone core plug (9) was saturated with formation brine and then
crude oil as described earlier. Oil recovery increased to 63% as a result of injecting one pore volume of formation brine.
Then, we switched to aquifer water accumulator and started injection at rate of 0.3 cm3/min. Aquifer water produced 8.6%
additional oil, Fig. 19. This confirmed the wettability study using contact angles where aquifer water altered wettability
toward water-wet.
Injecting seawater as a secondary mode was examined in limestone core plug (10). Aquifer as well as 1:1 diluted aquifer
water was then injected for at least two pore volumes. Fig. 20 suggested oil recovery of 54% after injecting seawater.
Injecting formation brine similar to the original saturation brine may be explained the extra recovery in core plug (9). Aquifer
8 SPE 129972

water was then injected at the same flow rate 0.3 cm3/min. Only 2.12% additional oil was recovered after we shifted to
aquifer water. Diluted aquifer water (1:1) showed only 0.2% oil recovery. Injecting aquifer water after high salinity formation
brine improved the recovery much better than seawater and then aquifer.
Dolomite cores were saturated and flooded with oil at high temperature. Then, dolomite core plug (5) was water flooded
with aquifer water as well as diluted aquifer water (1:1 and 1:10). Fig. 21 represented the oil recovery as a function of pore
volume. Oil recovery was 21.7% after injecting aquifer water for more than two pore volumes, at slow rate 0.3 cm3/min.
Diluted aquifer water at two different ratios, 1:1 and 1:10, were injected but no additional oil was recovered.
Dolomite core plug (6) was flooded with seawater, aquifer water and then 1:1 diluted aquifer water. Seawater injection as
a secondary mode in dolomite rock showed more recovery than aquifer. In this test, oil recovery was 35.5% after injecting
seawater for 2.5 pore volumes. Then, we switched to aquifer water and 8.9% additional oil was recovered. Moreover, 1:1
diluted aquifer water added 1.7% more oil recovery, Fig. 22.

Conclusions
The effect of salinity on contact angles of calcium carbonate substrates was examined over a wide range of parameters.
Based on the results obtained, the following conclusions can be drawn:

1. Similar to previous studies, the effect of pressure on wettability was not significant.
2. Based on contact angle measurements, calcium carbonate substrates were oil-wet when seawater or formation water
was used. However, the rock because water wet when aquifer or deionzed water was used.
3. Based on core flood experiments, addition oil recovery was obtained when aquifer water was injected after
formation brine. This results confirms those obtained from contact angle measurements.
4. Seawater recovered more oil from limestone that from dolomite cores.

Acknowledgments
The authors would like to thank Petroleum Engineering Department in Texas A&M University for their support. The authors
would like to acknowledge Saudi Aramco Company for funding this work. We thank Dr. Philip Jaeger from Eurotechnica
GmbH Company for useful discussion and ConocoPhillips Company for providing crude oil.

References
Almehaideb, R.A., Ghannam, M.T., and Zekri, A.Y. 2004. Experimental Investigation of Contact Angles of Crude Oil-
Microbial Solution on Carbonate Rocks. Petroleum Science & Technology 22(3–4): 423–438.
Alotaibi, M.B., and Nasr-El-Din, H.A. 2009. Effect of Brine Salinity on Reservoir Fluids Interfacial Tension. Paper SPE
121569 presented at the SPE EUROPEC/EAGE Conference, Amsterdam, The Netherlands, 8–11 June.
Alotaibi, M.B., Nasralla, R.A., and Nasr-El-Din, H.A. 2010. Wettability Study Using Low-Salinity Water in Sandstone
Reservoirs. Paper OTC-20718 will be presented at the 2010 Offshore Technology Conference, Houston, Texas, 3 - 6 May.
Anderson, W.G. 1986. Wettability Literature Survey—Part 1: Rock/Oil/Brine Interactions and the Effects of Core Handling
on Wettability. Journal of Petroleum Technology 38: 1125–1144.
Andreas, J.M., Hauser, E.A., and Tucker, W.B. 1938. Boundary Tension by Pendant Drops. The Journal of Physical
Chemistry 42: 1001–1019.
Austad, T. et al. 2009. Seawater in Chalk: An EOR and Compaction Fluid. SPEREE 11 (4): 648-654.
Hamouda, A.A., and Karoussi, O. 2008. Effect of Temperature, Wettability and Relative Permeability on Oil Recovery from
Oil-wet Chalk. Energies 1 (1): 19–34.
Hjelmeland, O.S., and Larrondo, L.E. 1986. Experimental Investigation of the Effects of Temperature, Pressure, and Crude
Oil Composition on Interfacial Properties. SPERE 1 (4): 321–328.
Høgnesen, E.J., Strand, S., and Austad, T. 2005. Waterflooding of Preferential Oil-Wet Carbonates: Oil Recovery Related to
Reservoir Temperature and Brine Composition. Paper SPE 94166 presented at the SPE Europec/EAGE Annual
Conference, Madrid, Spain, 13–16 June.
Li, D., and Neumann, A.W. 1992. Contact Angles on Hydrophobic Solid Surfaces and Their Interpretation. Journal of
Colloid and Interface Science 148 (1): 190–200.
Lichaa, P. et al. 1992. Wettability Evaluation of a Carbonate Reservoir Rock. Advances in Core Evaluation III Reservoir
Management, reviewed proceedings of the Society of Core Analysis Third European Core Analysis Symposium, P.W.
Worthington and C. Chardaire-Riviere (eds.), Gordon and Breach, Paris.
Morrow, N.R., Ma, S., Zhou, X., and Zhang X. 1994. Characterization of wettability from spontaneous imbibition
measurements. CIM 94-475, 45th Ann. Tech. Meeting.
Pierre, A. et al.1990. Calcium as Potential Determining Ion in Aqueous Calcite Suspensions. Journal of Dispersion Science
and Technology 11(6): 611–635.
Saner, S. et al. 1991. Wettability Study of Saudi-Arabian Carbonate Reservoir Core Samples. The Arabian Journal for
Science & Engineering 16(3): 357–371.
Strand, S. et al. 2008. “Smart Water” for Oil Recovery from Fractured Limestone: A Preliminary Study. Energy & Fuels 22
(5): 3126–3133.
SPE 129972 9

Strand, S., and Austad, T. 2008. Effect of Temperature on Enhanced Oil Recovery from Mixed-wet Chalk Cores by
Spontaneous Imbibition and Forced Displacement using Seawater. Energy & Fuels 22(5): 3222–3225.
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(1–3): 1–10.
Strand, S., Standnes, D.C., and Austad, T. 2003. Spontaneous Imbibition of Aqueous Surfactant Solutions into Neutral to Oil-
wet Carbonate Cores: Effects of Brine Salinity and Composition. Energy & Fuels 17(5): 1133–1144.
Wang, W., and Gupta, A. Investigation of the Effect of Temperature and Pressure on Wettability Using Modified Pendant
Drop Method. Paper SPE 30544 presented at the 1995 SPE Annual Technical Conference & Exhibition, Dallas, Texas,
USA, 22–25 October.
Webb, K.J., Black, C.J.J., and Tjetland, G. 2005. A Laboratory Study Investigating Methods for Improving Oil Recovery in
Carbonates. Paper IPTC 10506 presented at the International Petroleum Technology Conference, Doha, Qatar, 21–23
November.
Xu, W., Ayirala, S.C., and Rao, D.N. 2006. Compositional Dependence of Wetting and Contact Angles in Solid-liquid-liquid
Systems under Realistic Environments. The Canadian Journal of Chemical Engineering, 84(1): 44–51.
Yang, D., Gu, Y., and Tontiwachwuthikul, P. 2008. Wettability Determination of the Crude Oil−Reservoir Brine−Reservoir
Rock System with Dissolution of CO 2 at High Pressures and Elevated Temperatures. Energy & Fuels 22(4): 2362–2371.
Yu, L. et al. 2007. Wettability Alteration of Chalk by Sulphate Containing Water, Monitored by Contact Angle Measurment.
Paper SCA2007-01 presented at the International Symposium of the Society of Core Analysts, Calgary, Canada, 10–12
September.
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: 179–187.
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: Physicocheical and Engineering Aspects 301(1–3): 199–208.
Zhang, X., Morrow, N.R., and Ma, S. 1996. Experimental verification of a modified scaling group for spontaneous
imbibition. SPERE, 280–285.
10 SPE 129972

Table 1: Geochemical analysis of formation, aquifer, and seawater.

Ions Formation water Seawater Aquifer water


Concentration, mg/l
Na+ 51,187 16,877 1,504
Ca2+ 29,760 664 392
Mg2+ 4,264 2,279 66
2+
Ba 10 0 0
2+
Sr 1,035 0 5
Cl− 143,285 31,107 2,577
HCO3− 351 193 192
2−
CO3 0 0 0
2−
SO4 108 3,560 700
TDS, mg/l 230,000 54,680 5,436

Table 2: Properties of limestone and dolomite core plugs.

Core No. Rock Type Average Diameter Length Bulk Volume Porosity Permeability Swi
3
in. in cm vol% md %
9 1.4 3.731 94.21 30.45 96.52 19.0
Limestone
10 1.4 3.732 94.23 32.04 108.20 22.2
3 1.5 4.186 120.68 19.93 198.78 27.5
4 1.5 4.115 118.58 18.51 123.53 32.1
Dolomite
5 1.5 4.985 143.97 19.86 384.87 30.1
6 1.5 4.990 144.95 19.23 217.19 35.4
SPE 129972 11

Table 3: Contact angle at HTHP conditions.

Rock & Oil Aging Period Temperature Pressure Contact angles, °


information Brine Type Wettability mode
h °F psi Left Right
122 200 99.8 ± 0.60 101.7 ± 3.27
Not aged Intermediate–wet
Formation Brine 93.0 ± 2.48 87.8 ± 2.52
2000
( 230 kppm) Oil–wet
117.6 ± 8.17 126.6 ± 1.91
1500 110.2 ± 0.61 118.4 ± 1.16 Intermediate–wet
194 2000 94.3 ± 0.46 98.6 ± 0.31
Intermediate–wet
1500 109.5 ± 0.52 111.4 ± 0.51
1000 111.6 ± 0.99 120.1 ± 0.70 Intermediate and oil–
wet
500 113.4 ± 1.40 129.1 ± 1.01
Seawater
( 54 kppm) 200 142.7 ± 0.34 142.1 ± 0.65
122 500 139.2 ± 0.76 142.7 ± 0.45
Oil–wet
Calcite crystal/West Texas oil

1000 142.9 ± 0.37 148.3 ± 0.74


122 136.3 ± 0.47 144.7 ± 0.51
2000
270.5 132.7 ± 0.35 106.6 ± 2.45
122 58.0 ± 1.65 47.1 ± 3.41
Deionized water 194 47.8 ± 1.58 49.5 ± 0.88
266 54.7 ± 0.65 51.6 ± 0.76
122 57.6 ± 2.11 57.3 ± 2.55
Sodium Sulphate
(3,560 ppm) 194 70.7 ± 2.68 62.1 ± 2.78
266 62.4 ± 3.01 56.7 ± 2.28
122 39.9 ± 2.15 38.4 ± 0.70
Sodium Sulphate 24 h aging 2000 Water–wet
(1,780 ppm) 194 49.9 ± 1.69 44.8 ± 1.31
266 47.2 ± 4.13 45.6 ± 10.27
Sodium Sulphate 122 67.4 ± 1.53 91.7 ± 1.31
(890 ppm)
194 74.3 ± 8.29 87.7 ± 9.37
122 66.6 ± 1.44 65.8 ± 2.10
Aquifer Water
(5,436 ppm) 194 72.1 ± 1.11 73.8 ± 1.50
266 71.0 ± 1.86 74.5 ± 2.10
122 94.7 ± 2.09 92.7 ± 3.06
Formation Brine
194 84.1 ± 2.09 86.2 ± 2.05 Intermediate–wet
( 230 kppm)
Laurda limestone/Sweet crude oil

266 76.4 ± 3.32 64.6 ± 14.05


122 63.6 ± 1.86 60.3 ± 1.62
Seawater
( 54 kppm) 194 48.1 ± 0.99 53.2 ± 1.20
266 2000 52.0 ± 2.55 58.3 ± 2.01
122 63.0 ± 0.81 60.5 ± 1.12
Aquifer Water Water–wet
(5,436 ppm) 194 58.8 ± 5.09 57.3 ± 3.32
266 52.5 ± 0.79 52.2 ± 0.83
194 65.7 ± 2.24 64.1 ± 0.71
Deionized water
266 60.7 ± 0.77 61.4 ± 0.89
12 SPE 129972

1000

900

800

700

600
Lin, Cps

500

400

300

200

100

0
4 10 20 30 40 50 60 70

2-Theta-Scale

Fig. 1: XRD of pink desert rock (limestone)

6000

5000

4000
Lin, Cps

3000

2000

1000

0
4 10 20 30 40 50 60 70

2-Theta-Scale

Fig. 2: XRD of Silurian dolomite rock


SPE 129972 13

PC N2 for Pressure

Light source
Compressed N2
Digital camera

Oil

Accumulator

Syringe pump

Fig. 3: Axisymmetric Drop Shape Analysis Instrument

Electrical oven
Glass burette
Core plug

(accuracy: 0.1 cc)


Amott cell

Brine accumulator

1
1, 2, 3: Ball valves
ISCO syringe pump

Fig. 4: Schematic representation of HTHP modified Amott cell set-up


14 SPE 129972

Pressure transducer
Back pressure  
regulator (2)

Core holder

Brine II accumulator
Brine I accumulator
Oil accumulator
Back pressure
regulator (1)

Hydraulic pump

ISCO syringe pump

Seawater

Fig. 5: Schematic of coreflood set-up

Crude oil

P= 200 psi P= 1,000 psi

Fig. 6: Images of seawater droplet in sour crude oil at 122°F and two different pressures
SPE 129972 15

24

22
200psi
20
500psi

IFT, mN/m
18
1000psi
16
1500psi
14
1950psi
12

10
0 10 20 30 40 50
Time, min.

Fig. 7: IFT of seawater droplet in sour crude oil as a function of pressure and time, (T= 122°F, time interval= 1 min.)

130
120
110 Left angle
100
Contact Angle, º

90 Right angle
80
70
0 200 400 600 800 1000 1200 1400 1600

Time, min.

Fig. 8: Contact angle of formation brine/crude oil/calcite at 194°F and 2000 psi (time interval=20 min., no aging)

130

120
Conta ct Angle ,

110

Left angle
100
Right angle

90
0 5 10 15 20 25
Time, min.

Fig. 9: Contact angle of formation brine/crude oil/calcite at 194°F and 2000 psi (One day aging at 194°F)
16 SPE 129972

130

120

Contact Angle,
110

100 Left a ngle


Right a ngle

90
0 5 10 15 20 25
Time, min.

Fig. 10: Contact angle of formation brine/crude oil/calcite at 194°F and 1500 psi (One day aging at 194°F)

160

150 Left Angle


Right Angle
140
Contact Angle, °

130

120
0 5 10 15 20 25

Time, min.
Fig. 11: Contact angle of formation brine/crude oil/calcite at 122°F and 2000 psi (One day aging at 194°F)

59

Left Angle Right Angle


57

55
Contact Angle,

53

51

49
0 5 10 15 20 25
Time, min.

Fig. 12: Contact angle of DI water/crude oil/calcite at 194-266°F and 2000 psi (One day aging at 194°F)
SPE 129972 17

Na2SO4 solution (3,560 ppm) 

T= 122°F, t= 10 s T= 122°F, t = 20.59 min.  T= 194°F, t = 10 s  T= 194°F, t = 22.30 min. 


Na2SO4 solution (1,780 ppm) Aquifer water (5,436 ppm)

T= 122°F, t= 10 s T= 122°F, t = 28.17 min.  T= 122°F, t= 10 s T= 122°F, t= 27.33 min.


Aquifer water (5,436 ppm)

T= 194°F, t = 10 s T= 194°F, t = 22.17 min. T= 266°F, t = 10 s T= 266°F, t = 25.33 h


 
Fig. 13: Brines/crude oil/calcite angles images at different temperature and time interval (P = 2,000 psi)

Formation brine (230kppm)

   
T= 122°F, t = 10 s  T= 122°F, t = 36.16 min.  T= 266°F, t = 10 s  T= 266°F, t = 22.00 h 
Aquifer water (5,436 ppm) 

     
T= 194°F, t = 10 s  T= 194°F, t = 19.50 h  T= 266°F, t = 10 s  T= 266°F, t = 22.00 h 
 
Fig. 14: Brines/crude oil/Laurda limestone images at different temperature and time interval (P = 2,000 psi)
18 SPE 129972

10
8
6 q= 0.5 cc/min. 10
4 8
q= 1 cc/min.
2 6
ΔP, psi

0 q= 2 cc/min. 4

ΔP, psi
ΔP = 4.4799*q
0 50 100 2 R² = 0.9994
0
Cumulative time, min.
0 0.5 1 1.5 2 2.5 3

q, cm3/min.

Fig. 15: Differential pressure and flow rate data for core plug limestone no. 9

150
120
90 q= 0.5 cc/min.
60
30 q= 1 cc/min.
0
ΔP, psi

q= 2 cc/min.
0 50 100
Cumulative time, min.

Fig. 16: Differential pressure and flow rate data for core plug limestone no. 9

40

35

30

25 Seawater
Oil Recovery, %

20

15

10

5 Formation brine
0
0 50 100 150 200 250 300 350 400

Cumulative time, h

Fig. 17: Salinity effect on oil recovery from core plug 3D at temperature 194F and Pressure 100 psi
SPE 129972 19

  40

35 Formation brine
at 194 F
30

Oil Recovery, %
25
Aquifer brine at 194 F
Aquifer water
20 at ambient condition

15

10

0
0 100 200 300 400 500 600 700
Cumulative time, h

Fig. 18: Salinity effect on oil recovery from dolomite core plug (4) at ambient and high temperature conditions

100

80
Formation
Oil Recovery, %

60

40
Aquiferwater
20

0
0 0.5 1 1.5 2 2.5
Pore Volumes

Fig. 19: Oil recovery from limestone core plug (9) using formation brine and aquifer water

100

80 Seawater
Oil Recovery, %

60

40

20
Aquiferwater
1:1 Diluted aquiferwater
0
0 1 2 3 4 5 6 7 8 9
Pore Volumes

Fig. 20: Oil recovery from limestone core plug (10) using seawater, aquifer water, and 1:1 diluted aquifer water
20 SPE 129972

100

80
Aquiferwater

Oil Recovery, %
1:1 Diluted 1:10 Diluted
60 aquiferwater aquiferwater

40

20

0
0 1 2 3 4 5 6 7
Pore Volumes

Fig. 21: Oil recovery from dolomite core plug (5) using aquifer water, 1:1 and 1:10 diluted aquiferwater

100

80
Oil Recovery, %

60 Seawater Aquiferwater
1:1 Diluted aquiferwater

40

20

0
0 1 2 3 4 5 6 7 8 9

Pore Volumes

Fig. 22: Oil recovery from dolomite core plug (6) using seawater, aquifer water, 1:1 diluted aquiferwater

You might also like