You are on page 1of 22

Subscriber access provided by UNIV OF NEBRASKA - LINCOLN

Article
Impact of Gas-Condensate Composition and Interfacial
Tension on Oil-repellency Strength of Wettability Modifiers
Jalal Fahimpour, and Mahmoud Jamiolahmady
Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/ef5007098 • Publication Date (Web): 03 Oct 2014
Downloaded from http://pubs.acs.org on October 18, 2014

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Energy & Fuels is published by the American Chemical Society. 1155 Sixteenth
Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the
course of their duties.
Page 1 of 21 Energy & Fuels

1
2
3
4 1 Impact of Gas-Condensate Composition and Interfacial Tension on Oil-
5
6 2 repellency Strength of Wettability Modifiers
7
8 3 J. Fahimpour, M. Jamiolahmady, Heriot-Watt University
9
10
11 4 Abstract: Productivity of gas-condensate wells can significantly be declined as condensate
12
13 5 bank evolves around the wellbore. Wettability alteration of formation minerals from strongly
14
6 liquid-wet to intermediate gas-wet conditions using liquid-repellent fluorinated chemicals has
15
16 7 shown promising results to mitigate such liquid-blockage issues. This paper, for the first time,
17
18 8 presents the results of unique contact angle measurements conducted on the treated carbonate
19
20 9 substrates using synthetic gas-condensate fluids. The impact of hydrocarbon composition and
21
22 10 interfacial tension (IFT) on the condensate wetting tendency was investigated using various
23 11 binary and multi-component mixtures at ambient temperature. A high-pressure high-
24
25 12 temperature (HPHT) drop shape analysis setup was customized to perform the static contact
26
27 13 angle tests. The generality of the new findings from these measurements was also confirmed
28
29 14 using dynamic unsteady-state flow tests. The results demonstrated that the wetting tendency
30
15 of the condensate is significantly decreased as IFT increases and/or the hydrocarbon
31
32 16 composition becomes heavier. The C1-nC4 condensate drop, for instance, completely wetted
33
34 17 out the treated surface at both low and high-IFT limits of 1 and 10 mN/m, whereas the C1-
35
36 18 nC10 sample resulted in contact angles of 30° and 65° at similar IFT conditions, respectively.
37
38
19 The fluorochemical’s oil-repellency strength was also evaluated at high temperatures of 50
39 20 and 100°C with promising results. In this paper, the initial investigations on the impact of
40
41 21 surface roughness on the solid surface wettability was carried out, demonstrating the positive
42
43 22 impact of surface roughness on increasing the desired level of oil-repellency achieved after
44
45 23 the wettability alteration. The new findings from this work provide important guidelines in
46
24 design and application of wettability modifiers for gas-condensate reservoirs.
47
48
49 25 1. Introduction
50
51 26 A gas-condensate fluid is mainly composed of methane with decreasing amounts of ethane
52
53 27 and propane and smaller amounts of butane, pentane and hexane.1 The proportion of heptanes
54
55
28 plus hydrocarbons in the systems is also generally less than 12.5 mole percent.2 The presence
56 29 of heavier hydrocarbons in such fluid systems, compared to dry and wet gases, causes the
57
58 30 liquid condensate to drop out from the gas phase in the reservoir as well as surface.3 The
59
60 31 retrograde condensate is a light liquid with an API gravity ranging from 40° to 60°, which,
32 compared to crude oil, is considered as a lucrative hydrocarbon for finished products.

1
ACS Paragon Plus Environment
Energy & Fuels Page 2 of 21

1
2
3
1 Surface forces between rock and fluid system significantly impact the fluids' distribution and
4
5 2 their flow characteristics within the porous media.4 In gas-condensate systems when the
6
7 3 pressure near the wellbore falls below the dew-point, the dropped out condensate quickly
8
9 4 wets the pore surfaces. That is, the condensate exists as a thin film throughout the pores while
10
11
5 non-wetting gas occupies the centre of them.5-8 As pressure continues to decline further and
12 6 more condensation occurs, the low-mobility condensate accumulates and blocks the narrower
13
14 7 pores restricting the gas phase to flow freely. Substantial productivity losses have been
15
16 8 reported due to such condensate blockage issues, even in lean gas-condensate reservoirs.11,12
17
18 9 The positive coupling (increase of gas condensate relative permeability as velocity increases
19 10 and/or IFT decreases) and negative inertia (decrease of relative permeability (kr) as velocity
20
21 11 increases), have also been reported to affect the gas and condensate flow behaviour.6,9,10
22
23 12 These two effects, which are governed by the competition between capillary (surface) and
24
25 13 viscous forces, further highlights the unique characteristics of such low-IFT systems.
26
27
14 The chemical treatment of reservoir rocks using liquid-repellent agents whereby the rock
28 15 surface energy is reduced was first proposed by Li and Firoozabadi, 2000.13 This technique is
29
30 16 basically aimed at altering the rock wettability from strongly liquid- to intermediate gas-wet
31
32 17 conditions. As a result, the fluids’ mobility increases and the condensate blockage is
33
34 18 alleviated. For several years, a broad range of studies have been dedicated to this subject,
35 19 investigating a number of pertinent parameters to the chemical treatment process such as
36
37 20 chemical type versus reservoir rock type, solvent composition, high-temperature treatment,
38
39 21 chemicals' stability in brine and treatment longevity.14-17 All these studies, however, suffer
40
41 22 from lack of correct and thorough procedures for evaluating the ultimate performance of
42
23 tested wettability modifiers in gas-condensate systems. That is, in all these investigations, the
43
44 24 conventional oil and gas fluids at ambient conditions, e.g. decane/air or heptane/air,14,19 have
45
46 25 been used to perform the contact angle measurements on the treated substrates assuming that
47
48 26 such tests are representative for gas-condensate systems. On the other hand, there are few
49
50 27 cases18 that have employed gas-condensate fluids during pseudo-steady state displacement
51 28 tests but not during contact angle measurements. Such flow tests, however, have been
52
53 29 conducted at IFTs above 4 mN/m but not at lower IFTs and under steady-state conditions,
54
55 30 which are more representative of such low-IFT condensing systems. Accordingly, to the best
56
57 31 of authors’ knowledge, there is no experimental evidence reported on the alteration of contact
58
32 angle data on the treated surfaces in presence of gas-condensate fluids.
59
60 33 As mentioned above the distribution and coupled flow of gas and condensate in porous media
34 at low-IFT values is known to be fundamentally different from that of conventional gas-oil
2
ACS Paragon Plus Environment
Page 3 of 21 Energy & Fuels

1
2
3
1 flow systems. This further emphasizes the importance of more realistic evaluations on the
4
5 2 performance of wettability modifiers in such low-IFT conditions, which has been underrated
6
7 3 in all previous investigations in the literature. This topic has been addressed in this study by
8
9 4 performing a series of contact angle measurements and displacement tests using synthetic
10
11
5 gas-condensate mixtures at high- and low-IFT limits. The experimental design, procedures
12 6 and results of this study are presented in the following sections.
13
14
15 7 2. Design of Experiments Performed
16
17 8 The wettability modifier sought in this work was specifically designed for carbonate
18
19 9 minerals, which compared to sandstone rock types, have received less attention. Accordingly,
20
10 in this study, carbonate substrates of Texas-Cream (TC) limestone and Baker-Dolomite (BD)
21
22 11 outcrop rocks were mainly used for the wettability alteration purposes. It should be
23
24 12 mentioned that these two rock types have strongly liquid-wet surfaces at their virgin state
25
26 13 (before treatment), i.e. θliquid/air=0° (Figure 1a). To find the appropriate wettability modifier
27
28
14 for these carbonate substrates, sixteen fluorinated chemicals with various ionic functionalities
29 15 were initially screened, amongst which A1 anionic chemical compound was selected. It
30
31 16 should be noted that during this screening stage, the conventional decane/air system was
32
33 17 employed to perform the contact angle measurements, i.e. the decane droplet was deposited
34
35 18 on the treated surface surrounded by air at ambient conditions. This provided a simple and
36
19 quick method for examining the oil-repellency strength of the chemicals. The performance of
37
38 20 the selected chemical was then evaluated in presence of more complicated gas-condensate
39
40 21 systems. For this purpose, the binary-component mixtures of C1-nC10, C1-nC6 and C1-nC4
41
42 22 were primarily employed at different pressures and at ambient temperature of 20°C, where
43
44
23 the impacts of hydrocarbon composition and interfacial tension on wetting tendency of the
45 24 condensate were investigated. It should be mentioned that conducting the experiments at
46
47 25 room rather than high temperatures, makes the test procedures considerably easier and more
48
49 26 straightforward. These measurements were further extended to more complex multi-
50
51 27 component gas-condensate systems using C1-nC5-nC8-nC10 and C1-nC5-nC8-nC15 fluid
52 28 mixtures. Next, to explore the impact of temperature on the oil-repellency strength of the
53
54 29 treated surface (or chemical performance), a number of measurements were repeated for C1-
55
56 30 nC10 binary mixtures at higher temperatures of 50 and 100°C. Finally, to further explore the
57
58 31 generality of the new findings under dynamic flow conditions, a number of unsteady-state
59
32 displacement tests were carried out using different binary mixtures.
60

3
ACS Paragon Plus Environment
Energy & Fuels Page 4 of 21

1
2
3
1 2.1. Chemical Solutions and Rock Treatments
4
5 2 Based on initial screening tests performed, A1 anionic fluorosurfactant was selected for
6
7 3 treatment of the carbonate substrates. Chemical A1 is composed of 14% solids in water and at
8
9 4 22°C, has a pH, CMC (critical micelle concentration) and relative density of 7-9, 0.5% and
10
11
5 1.1, respectively. This chemical compound does not flash, has a boiling point of about 100°C
12 6 and its thermal decomposition occurs at temperatures above 200°C. The final optimized
13
14 7 treatment solution employed here has 2 wt% A1 diluted in methanol. To treat the rock surface
15
16 8 the carbonate substrate was soaked in the chemical solution (about 30 cc) held in a piston cell
17
18 9 for 18 hours under 400 psig and 130°C. The treated rock sample was then dried for two hours
19 10 inside the oven at 100°C before conducting the post-treatment contact angle measurements.
20
21 11 Figure 1 shows the oil-wetting state of the Texas-Cream and Baker-Dolomite substrates
22
23 12 before (θdecane/air=0°) and after (θdecane/air~90°-95°) the wettability alteration process. It should
24
25 13 be mentioned that in this study, the treated Texas-Cream rock substrates were mainly used for
26
27
14 the contact angle measurement tests whilst the Baker-Dolomite core sample with higher
28 15 permeability was used for the dynamic flow tests.
29
30 16
31
32
33
34
35
36
37
38
39
40
41 17
42
18 Figure 1: The oil-wetting state of the Texas-Cream and Baker-Dolomite carbonate substrates
43 19 before and after the wettability alteration: complete spreading of the decane (nC10) drop
44 20 before the treatment (a) and effective oil-repellency achieved after the treatment on treated
45 21 Texas-Cream (b) and Baker-Dolomite (c) surfaces.
46 22
47
48
49 23 2.2. Preparation of Gas-Condensate Fluids
50 24 Three binary-component mixtures of C1-nC4, C1-nC6 and C1-nC10 and two multi-
51
52 25 component mixtures of C1-nC5-nC8-nC10 and C1-nC5-nC8-nC15 were used to perform the
53
54 26 contact angle measurements. The PTVi module of Eclipse commercial reservoir simulator23
55
56 27 was employed to model the thermodynamic phase behaviour of each hydrocarbon mixture
57
28 and subsequently to determine the pressures corresponding to the selected IFT values. To
58
59 29 simplify the experimental procedures, e.g. mixing the fluids and transferring them between
60
30 the accumulators and the analysis apparatus (DSA), the high-pressure hydrocarbon mixtures

4
ACS Paragon Plus Environment
Page 5 of 21 Energy & Fuels

1
2
3
1 were prepared and the corresponding contact angle measurements were conducted at ambient
4
5 2 temperature of 20°C, unless otherwise specified. Table 1 lists the total molar composition of
6
7 3 all fluid mixtures used in our measurements at 20°C and their corresponding pressure and
8
9 4 IFT ranges. The molar distribution of each hydrocarbon component in the condensate phase,
10
11
5 as the main element prevailing the contact angle behaviour, has also been shown in Table 2.
12 6 The reported IFT values in Table 1 are either estimated by PVTi or measured during the
13
14 7 course of the tests by the conventional pendant-drop technique.
15
16 8 To prepare the binary-component gas-condensate mixture, initially the required volume of
17
18 9 methane (C1) gas was transferred from the methane cylinder into the mixing cell at the
19 10 desired test pressure. Subsequently, the required volume of butane (nC4), hexane (nC6) or
20
21 11 decane (nC10) (depending on the mixture total molar composition) was injected from a
22
23 12 reserve piston cell into the mixing cell to be combined with methane. The mixing was then
24
25 13 continued for three to four days, while the mixing cell was connected to a displacement pump
26
27
14 to keep its pressure constant at the test pressure. The equilibrated gas and condensate fluids
28 15 were finally separated by transferring into two accumulator cells to be used for contact angle
29
30 16 measurements. The same procedures were also followed to prepare multi-component
31
32 17 mixtures with heavier components added one at a time.
33
34 18
35 19 Table 1: Total molar composition of hydrocarbon mixtures used in contact angle
36
37 20 measurements at 20°Cand their corresponding pressures and IFTs (measured and estimated).
38 IFT
Composition IFT
39 Test Gas-Condensate Test Pressure (Estimated by
(mole percent) (Measured)
40 Index Mixture (psig) PVTi)
(%) (mN/m)
41 (mN/m)
42 1 C1-nC10 70-30 1500 10.8 10.02
43 2 C1-nC10 70-30 2800 3.15 3.86
44
45 3 C1-nC10 70-30 3900 0.85 1.42
46 4 C1-nC6 50-50 200 15.80 15.04
47
48 5 C1-nC4 74.4-25.6 400 8.8 10.57
49 6 C1-nC4 74.4-25.6 1450 1.16 1.18
50
51 7 C1-nC5-nC8-nC10 87-7-3-3 1100 10.40 9.71
52 8 C1-nC5-nC8-nC10 87-7-3-3 2800 0.89 1.12
53
54 9 C1-nC5-nC8-nC15 87-7-3-3 1100 10.16 9.68
55 10 C1-nC5-nC8-nC15 87-7-3-3 3000 0.96 1.90
56
57 21
58 22
59 23
60

5
ACS Paragon Plus Environment
Energy & Fuels Page 6 of 21

1
2
3
1 Table 2: Molar distribution of hydrocarbon components in the condensate phase at different
4
5 2 pressures and at 20°C (corresponding to the different IFT values, Table1) based on the
6 3 thermodynamic fluid models available in PVTi.
7 Molar Distribution (%)
Composition
8 Test (of condensate phase)
Test Gas-Condensate (mole
9 Pressure
Index Mixture percent)
10 (psig) C1 nC4 nC5 nC6 nC8 nC10 nC15
(%)
11
12 1 C1-nC10 70-30 1500 37.7 - - - - 62.3 -
13 -
2 C1-nC10 70-30 2800 56.5 - - - - 43.5
14
15 3 C1-nC10 70-30 3900 68.1 - - - - 31.9 -
16 -
4 C1-nC6 50-50 200 7.6 - - 92.4 - -
17
18 5 C1-nC4 74.4-25.6 400 13.8 86.2 - - - - -
19 -
6 C1-nC4 74.4-25.6 1450 51.9 48.1 - - - -
20
21 7 C1-nC5-nC8-nC10 87-7-3-3 1100 33.1 - 32.8 - 16.9 17.2 -
22
23 8 C1-nC5-nC8-nC10 87-7-3-3 2800 64.3 - 16.3 - 9.3 10.1 -
24
25 9 C1-nC5-nC8-nC15 87-7-3-3 1100 32.6 - 33.0 - 17.0 - 17.4
26
27 10 C1-nC5-nC8-nC15 87-7-3-3 3000 62.4 - 16.5 - 9.7 - 11.4
28
29 4
30
31 5 2.3. Contact Angle Measurements
32
33 6 High-pressure high-temperature (HPHT) compartment of the drop shape analysis (DSA)
34
35 7 setup, shown in Figure 2, was used to perform the contact angle measurements. For this
36 8 purpose, the treated carbonate substrate was positioned inside the DSA optical cell
37
38 9 horizontally, where the sample needle was mounted on the chamber ceiling, whose tip was
39
40 10 about 3 mm above the rock surface. After this and in case of high-temperature tests, the cell
41
42 11 temperature was raised to the test temperature in incremental stages. The chamber was then
43
12 filled and pressurised with dry methane gas slowly. This was followed by injecting the
44
45 13 desired gas-mixture at the test pressure into the chamber, purging it of the methane gas. This
46
47 14 stage was continued for three times the chamber volume (30cc) to ensure the chamber is fully
48
49 15 saturated with the gas-mixture. It should be mentioned that pressurising the chamber with dry
50
51
16 methane prior to gas-mixture was necessary to prevent the optical cell misting up due to
52 17 sudden exposure of the gas-mixture to the atmospheric pressure.
53
54 18 To carry out the contact angle measurement, the condensate-mixture was injected through the
55
56 19 needle very slowly till it was detached and deposited on the rock substrate. The shaped angle
57
58 20 between the condensate/gas/rock phases was then analysed and traced to evaluate its stability
59 21 versus time. Analysis of the condensate pendant drop, right before running away from the
60
22 needle tip, was also used to measure the interfacial tension between gas and condensate

6
ACS Paragon Plus Environment
Page 7 of 21 Energy & Fuels

1
2
3
1 phases. The calculated IFTs were later compared to those estimated based on the PVTi
4
5 2 thermodynamic model demonstrating a reasonable agreement between the measured and
6
7 3 estimated IFT values. The DSA polynomial or tangent method-2, which adapts a curve to the
8
9 4 left and right regions of the three-phase contact point of condensate/gas/rock, irrespective of
10
11
5 the drop shape symmetry, was employed to calculate the corresponding contact angles
12 6 measured through the liquid phase. Accordingly, all contact angles reported here are the
13
14 7 average value of the left and right angles measured by this method. The experimental error
15
16 8 pertinent to this measurement procedure is about 5°. Here, it should be mentioned that to
17
18 9 check the accuracy and generality of the observed contact angles during this study each
19 10 measurement was repeated for two to three times at different spots on the treated surface.
20
21 11 This was facilitated by using a moving device incorporated into the DSA view chamber
22
23 12 (Figure 2) enabling the rock substrate to be shifted horizontally under high pressure and
24
25 13 temperature conditions for multiple sample dosing. Almost similar observations were
26
27
14 obtained during these repeatability tests in all cases with the observed differences to be within
28 15 the range of experimental errors. Accordingly, the average values of the results obtained for
29
30 16 each case has been reported in this work. Here, it should be highlighted that the main focus of
31
32 17 these unique contact angle measurements was to investigate the performance of wettability
33
34 18 modifiers at low-IFT conditions, which are more prevalent in gas-condensate systems and
35 19 dominate the flow behaviour of such systems. In addition, the performance of such chemicals
36
37 20 was also evaluated at high IFT-limits, which become more relevant as the reservoir depletion
38
39 21 continues and pressure declines further. It will be shown that if the treated surface indicates
40
41 22 poor oil-repellency at high-IFT conditions, it will also be ineffective at low-IFTs, whereas the
42
23 converse does not apply.
43
44 24
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

7
ACS Paragon Plus Environment
Energy & Fuels Page 8 of 21

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20 1
21
22 2 Figure 2: Custom-designed high-pressure high-temperature module of the drop shape analysis
23 3 (DSA) setup equipped with HPHT mixing cells for making and employing gas condensate
24 4 fluid mixtures in place.
25 5
26
27
6 3. Results and discussion
28
29 7 In this section, the results of contact angle measurements performed on treated carbonate
30
31 8 rocks at ambient temperature of 20°C are initially presented. Following this, the impacts of
32
33 9 temperature and surface roughness on the contact angle data are briefly discussed. At the end,
34
10 the results of the dynamic flow tests will be presented.
35
36
37 11 3.1. Binary-Component Gas-Condensate Mixtures
38
39 12 To simplify the experimental procedures and minimize the number of variables involved in
40
41 13 the tests, we decided to initiate our investigations using binary-component, instead of more
42
14 complex multi-component, systems. Three binary-mixtures of C1-nC4, C1-nC6 and C1-nC10
43
44 15 were used at different pressures corresponding to different IFT values.
45
46
47 16 3.1.1. C1-nC10 Mixture
48
49 17 Contact angle measurements were performed on the treated substrate at three different IFTs
50
51 18 of 10.0, 3.9 and 1.4 mN/m, corresponding to test pressures of 1500, 2800 and 3900 psig,
52 19 respectively. The pendant condensate drops as shown in Figure 3 were initially used to
53
54 20 calculate the corresponding IFTs (Table 1) between gas and condensate before depositing the
55
56 21 liquid droplet on the treated substrate. Figure 4 depicts the contact angles corresponding to
57
58 22 IFT of 10.0 mN/m. Contact angles of 65°, 65° and 63° were recorded right after the first
59
23 contact of the condensate drop with the substrate (θ0) and then after 1 and 3 hours,
60
24 respectively. This demonstrates the effective oil-repellency strength of the treated surface

8
ACS Paragon Plus Environment
Page 9 of 21 Energy & Fuels

1
2
3
1 with respect to this high-IFT hydrocarbon mixture. Moreover, it should be noted that in
4
5 2 contrast to the high contact angle value of about 95° obtained for conventional oil/gas
6
7 3 systems (i.e. decane-air), the chemical’s effectiveness is less pronounced for C1-nC10
8
9 4 hydrocarbon mixture as θ=65° was obtained. This trend is either linked to the IFT reduction
10
11
5 or fluid composition change when gas-condensate, instead of conventional, fluid has been
12 6 employed. These effects will be further highlighted as the remaining results of this study are
13
14 7 discussed in the following.
15
16 8 The performance of chemical A1 was further examined at the lower IFT values of 3.9 and 1.4
17
18 9 mN/m as depicted in Figure 5. Accordingly, stable contact angles of 44° and 30° were
19 10 observed corresponding to these two low-IFT conditions. These results, compared to higher
20
21 11 contact angle of 65° observed at high IFT of 10 mN/m, demonstrates the decreasing level of
22
23 12 oil-repellency of the treated surface with decreasing IFT. In other words, the condensate
24
25 13 wetting tendency increases as the interfacial tension between gas and condensate decreases.
26
27
14
28 15
29
30 16
31
32
33
34
35
36
37
38
39
40
41
42
17
43
44 18 Figure 3: Pendant drops of C1-nC10 used for IFT measurements at three different pressures.
45 19
46
47 20
48
49
50
51
52
53 21
54
55 22 Figure 4: Status of a C1-nC10 condensate drop on the treated Texas-Cream substrate at
56 23 T=20°C and IFT=10.02 mN/m, recorded versus time.
57 24
58
59 25
60

9
ACS Paragon Plus Environment
Energy & Fuels Page 10 of 21

1
2
3
4
5
6
7
8 1
9 2 Figure 5: Status of C1-nC10 condensate drops on the treated Texas-Cream substrate at
10
11 3 T=20°C and IFT=3.86 mN/m (left) and IFT=1.42 mN/m (right).
12 4
13
14
15 5 3.1.2. C1-nC6 Mixture
16 6 The mixture of C1-nC6 was prepared at 200 psig corresponding to an IFT of 15.0 mN/m.
17
18 7 Figure 6 shows the observed contact angles at different time steps on the treated substrate.
19
20 8 Although θ0=45° was recorded at the zero time step, the condensate drop imbibed very
21
22 9 quickly into the rock surface within 10 seconds.
23
10 It is interesting to note that compared to the first test conducted with C1-nC10 mixture at
24
25 11 IFT=10 mN/m, here higher IFT limit of 15 mN/m was considered but smaller contact angles
26
27 12 and more importantly rapid spreading of condensate on the treated surface were observed.
28
29 13 These findings underlined that in addition to the fluids interfacial effects, observed in the C1-
30
31
14 nC10 mixture case, the fluid mixture composition, i.e. the type of the hydrocarbon molecules
32 15 present in the system, has also a crucial impact on the wetting characteristics of the treated
33
34 16 surface.
35
36 17
37
38
39
40
41 18
42
19 Figure 6: Status of a C1-nC6 condensate drop on the treated Texas-Cream substrate at
43
44 20 T=20°C and IFT=15.04 mN/m, recorded versus time.
45 21
46
22
47
48
49 23 3.1.3. C1-nC4 Mixture
50
51 24 Contact angle measurements by C1-nC4 mixture were conducted on the treated substrate at
52
53 25 two high- and low-IFT values of 10.6 and 1.2 mN/m, corresponding to the test pressures of
54
55 26 400 and 1450 psi, respectively. Figure 7 and Figure 8 show the condensate behaviour on the
56 27 treated substrate versus time at these two high- and low-IFT limits, respectively. It was
57
58 28 observed that at both IFT conditions, condensate drop imbibed into the rock surface very
59
60 29 quickly, within 1 second. In other words, no evidence of the oil-repellency, even at the first
30 contact of the condensate with the substrate at high-IFT conditions, was observed, i.e. θ0=0°.

10
ACS Paragon Plus Environment
Page 11 of 21 Energy & Fuels

1
2
3
1 Comparison of the chemical performances at the high-IFT conditions above 10 mN/m for C1-
4
5 2 nC4 (θ0=0°), C1-nC6 (θ0=45°) and C1-nC10 (θ0=65°) mixtures, provides more credit to our
6
7 3 previous findings, expressing the reduced oil-repellency strength of the treated surface as the
8
9 4 gas-condensate composition becomes lighter. In other words, the tendency of the condensate
10
11
5 liquid phase to wet the treated solid surface increases, at a constant interfacial tension, as the
12 6 number of carbon atoms of the heavy-end components in the mixture decreases.
13
14 7
15
16 8
17 9
18
19
20
21
22
23
24 10
25
26 11 Figure 7: Quick spreading of the C1-nC4 condensate drop on the treated Texas-Cream
27 12 substrate at T=20°C and IFT=10.57 mN/m.
28 13
29
14
30
31 15
32
33
34
35
36
37
38 16
39
17 Figure 8: Quick spreading of the C1-nC4 condensate drop on the treated Texas-Cream
40
41 18 substrate at T=20°C and IFT=1.18 mN/m.
42 19
43
44 20 3.2. Multi-component Gas-Condensate Mixtures
45
46 21 The previous contact angle measurements, using binary gas-condensate systems, proved
47
48 22 substantial dependency of the treated surface oil-wetting characteristics to the gas/condensate
49
23 interfacial tension and molar composition. Following these interesting results we decided to
50
51 24 further investigate the impact of IFT and composition on the contact angles in more complex
52
53 25 multi-component systems. Accordingly, two multi-component gas-condensate mixtures,
54
55 26 composed of C1, nC5 (as the lighter proportions) and C8 and nC10/C15 (as the heavier
56
57
27 proportions), were prepared at ambient temperature (20°C). The mole fraction of C1, nC5,
58 28 nC8 and nC10 in the first fluid model was 87, 7, 3 and 3%, respectively. In the second
59
60 29 mixture, to explore the sensitivity of the chemical’s oil-repellency to the heavy-end
30 component, nC10 part (in the first model) was replaced by nC15 under similar proportion.

11
ACS Paragon Plus Environment
Energy & Fuels Page 12 of 21

1
2
3
1 Like binary-mixtures, two high- and low-IFT limits of about 10 and 1 mN/m (Table 1) were
4
5 2 also considered for both multi-component systems.
6
7
8 3 3.2.1. C1-nC5-nC8-nC10 Mixture
9
10 4 Contact angle measurements by C1-nC5-nC8-nC10 mixture were conducted at two IFT
11
12 5 values of 9.7 and 1.1 mN/m, corresponding to the test pressures of 1100 and 2800 psig,
13 6 respectively.
14
15 7 It was noted that at the high-IFT value of 9.7 mN/m, the initial contact angle at the zero time
16
17 8 step was θ0=65°. However, the condensate drop imbibed rapidly into the treated surface
18
19 9 within 1 minute. Figure 9 depicts these contact angle data. Comparing this result with that of
20
10 the C1-nC10 binary-mixture at nearly similar IFT of 10.0 mN/m, where a sustainable contact
21
22 11 angle of about 65º was obtained, highlights the negative impact of the intermediate
23
24 12 components, i.e. nC5 and nC8, in the gas-condensate composition, on the treated surface oil-
25
26 13 repellency strength. It should be noted that the total molar distribution of the nC10
27
28
14 component in the condensate phase of C1-nC10 mixture, i.e. 62%, has been split between
29 15 nC5 (33%), nC8 (17%) and nC10 (17%) in the corresponding liquid phase of the multi-
30
31 16 component mixture (Table 2).
32
33 17 The chemical performance became even less effective when IFT of the system was reduced
34
35 18 to 1.1 mN/m, as complete spreading of the condensate on treated substrate was observed.
36
19 Figure 10 demonstrates the imbibition of the condensate that occurred within 5 seconds.
37
38 20 These observations, in line with our previous understandings from binary-mixtures, further
39
40 21 highlights the importance of fluid composition and interfacial tension on condensate wetting
41
42 22 tendency characteristics.
43
23
44
45 24
46
47
48
49
50 25
51
52 26 Figure 9: Status of a C1-nC5-nC8-nC10 condensate drop on the treated Texas-Cream
53 27 substrate at T=20°C and IFT=9.71 mN/m, recorded versus time.
54 28
55
56 29
57
58 30
59
60

12
ACS Paragon Plus Environment
Page 13 of 21 Energy & Fuels

1
2
3
4
5
6
7
8
9
1
2 Figure 10: Quick spreading of the C1-nC5-nC8-nC10 condensate drop on the treated Texas-
10
11 3 Cream substrate at T=20°C and IFT=1.12 mN/m.
12 4
13
14
15 5 3.2.2. C1-nC5-nC8-nC15 Mixture
16 6 The impact of the heavy-end component on the treated surface oil-repellency strength was
17
18 7 investigated using the C1-nC5-nC8-nC15 fluid system, whereby nC10 proportion in the first
19
20 8 fluid model, discussed above, was replaced by nC15. The fluids were prepared at two
21
22 9 pressures of 1100 and 3000 psig, corresponding to IFT values of 9.7 and 1.9 mN/m,
23
10 respectively.
24
25 11 Figure 11 shows the contact angles of the condensate drop on the treated TC substrate
26
27 12 measured at high IFT=9.7 mN/m. A promising contact angle of θ0=71° was measured at zero
28
29 13 time step. More importantly, the treated surface showed sustainable oil-repellency for a long
30
31 14 period, i.e. θ=70° was recorded after 1 hour. More interesting result was observed at the low-
32 15 IFT of 1.9 mN/m, at which θ0=79° was obtained. Furthermore, the condensate drop was
33
34 16 reasonably stable over time as contact angle of 76º was recorded after 1 hour (Figure 12). It
35
36 17 should also be noted that none of the previous gas-condensate mixtures performed such high
37
38 18 and stable contact angles at low-IFT conditions of about 1 mN/m observed here. These
39
19 results at both IFTs, compared to those reported for the C1-nC5-nC8-nC10 mixture, highlight
40
41 20 the dominant impact of the nC15 heavy-end component, overcoming the adverse impact of
42
43 21 the nC5 and nC8 intermediate molecules on the condensate wetting tendency. These findings
44
45 22 place more emphasis on the prevailing impact of the heavy-end components in the gas-
46
47 23 condensate composition on improving the oil-repellency strength of the wettability modifiers,
48 24 even at low-IFT conditions.
49
50 25
51
52 26
53
54 θ = 71° 0 sec θ = 69° 1800 sec θ = 70° 3600 sec
55
56
57
58 27
59 28 Figure 11: Status of a C1-nC5-nC8-nC15 condensate drop on the treated Texas-Cream
60 29 substrate at T=20°C and IFT=9.68 mN/m, recorded versus time.
30

13
ACS Paragon Plus Environment
Energy & Fuels Page 14 of 21

1
2
3
4 θ = 79° 0 sec θ = 77° 1800 sec θ = 76° 3600 sec
5
6
7
8 1
9 2 Figure 12: Status of a C1-nC5-nC8-nC15 condensate drop on the treated Texas-Cream
10 3 substrate at T=20°C and IFT=1.90 mN/m, recorded versus time.
11 4
12
13
14 5 3.3. Effect of Temperature on Contact Angle Data
15 6 To explore the impact of temperature on oil-repellency strength of the treated surface, two
16
17 7 tests were conducted at 50 and 100°C using the C1-nC10 binary mixture. The measurements
18
19 8 were carried out at two high- and low-IFT limits of about 10 and 1 mN/m on the treated
20
21 9 Texas-Cream carbonate substrate. Promising contact angles of 71° and 74° were obtained at
22
10 50 and 100°C, respectively, at IFT of 10 mN/m. Figure 13 shows the corresponding measured
23
24 11 contact angles. It should be noted that the difference between the contact angle observed at
25
26 12 these two high temperatures and that obtained previously at T=20°C (i.e. θ=65°, Figure 4) are
27
28 13 within the range of the experimental errors. At the low-IFT limit of 1 mN/m, almost the same
29
30 14 level of oil-repellency observed at ambient temperature (i.e. θ=30°, Figure 5), was also
31 15 achieved when temperature was increased to 50 and 100°C giving contact angles of 40° and
32
33 16 35° , respectively. Figure 14 shows the corresponding measured contact angles at these two
34
35 17 high temperature values. These observations first of all demonstrate that the chemical oil-
36
37 18 repellency strength or the condensate wetting tendency was almost maintained constant when
38
19 temperature was increased. Furthermore, in line with the general trend observed previously
39
40 20 during the tests conducted at ambient temperature of 20°C, the condensate wetting tendency
41
42 21 at high temperature conditions was also increased with decreasing IFT.
43
44 22
45
46
47
48
49
23
50 24 Figure 13: C1-nC10 condensate drops on the treated Texas-Cream substrates at IFT~10
51
52 25 mN/m and T=50°C (left) and 100°C (right).
53 26
54
55
56
57
58
59 27
60 28 Figure 14: C1-nC10 condensate drops on the treated Texas-Cream substrates at IFT~1 mN/m
29 and T=50°C (left) and 100°C (right).

14
ACS Paragon Plus Environment
Page 15 of 21 Energy & Fuels

1
2
3
1 3.4. Effect of Surface Roughness on Contact Angle Data
4
5 2 All previous contact angle tests discussed above were conducted on treated carbonate rock
6
7 3 substrates with non-uniform rough surfaces. Generally speaking, it is believed that in addition
8
9 4 to the surface energy, surface roughness has also a significant impact on wetting
10
11
5 characteristics of solid substrates.20-22 Here, to provide a preliminary understanding of the
12 6 impact of surface roughness on the performance of wettability modifiers and evaluate the
13
14 7 generality of the new findings from the previous contact angle measurements on rock
15
16 8 samples, further contact angle tests were performed on smooth surfaces of calcite. For this
17
18 9 purpose, highly smoothed single-crystals of calcite (CaCO3) were sourced from MTI Corp.
19 10 The rhombus shape crystals were supremely polished bringing them to an average roughness
20
21 11 (Ra) less than 5 Angstroms. Figure 15 compares the smoothed surface of a CaCO3 crystal
22
23 12 used here with the rough surface of the Texas-Cream carbonate rock samples. The calcite
24
25 13 crystals were treated using A1 chemical solution according to the same procedures explained
26
27
14 previously for treating the carbonate rock samples. The C1-nC10 and C1-nC4 binary gas-
28 15 condensate mixtures were employed to repeat two contact angle measurement tests
29
30 16 previously performed on carbonate rock substrates at ambient temperature (20°C) and IFT of
31
32 17 10 mN/m. Here, it should be noted that the smooth surface like the rough surface of the
33
34 18 carbonate rock (Figure 1a) was strongly oil-wet before the wettability alteration, i.e. a contact
35 19 angle of about 5° was measured through the decane drop deposited on the untreated calcite at
36
37 20 ambient conditions.
38
39 21 Figure 16 depicts the oil-wetting state of the treated smooth surface of calcite with respect to
40
41 22 two condensing systems of C1-nC4 and C1-nC10. It is noted that while the C1-nC4
42
43
23 condensate drop has almost wetted the treated surface (i.e. θ=7°), the C1-nC10 mixture has
44 24 shown an acceptable level of oil repellency (i.e. θ=50°). The observed trend, in line with our
45
46 25 previous findings from the contact angle tests on rough rock surfaces, further highlights the
47
48 26 significant impact of fluid composition on condensate wetting tendency, i.e. the lighter
49
50 27 molecular composition the more oil-wetting characteristics. Furthermore, comparing the
51 28 wetting state of the treated smooth surface of calcite with respect to the C1-nC10 mixture
52
53 29 with that of the Texas-Cream carbonate rock surface (θ=65°, Figure 4), demonstrates that
54
55 30 lower level of oil-repellency has been achieved on the smooth surface after the wettability
56
57 31 alteration. In other words, the surface roughness inherent in the rock surfaces has promoted
58
32 the wetting tendency of the treated surface towards less oil-wet conditions. This is also in
59
60 33 agreement with the general understanding of the impact of roughness on surface wettability,
34 i.e. when a surface is roughened, its wetting state is increased.24,25
15
ACS Paragon Plus Environment
Energy & Fuels Page 16 of 21

1
2
3
1 The underlying mechanisms for the effect of molecular composition on oil-repellency
4
5 2 strength of the treated surface, reported in this paper, have not thoroughly been understood. It
6
7 3 is believed that the interaction of intermolecular forces is one of the main driving forces
8
9 4 behind this effect. Such molecular interactions are partly captured by the measurable physical
10
11
5 properties such as viscosity and surface tension. Further investigation on this subject is
12 6 currently under way.
13
14 7
15
16 8
17
18
19
20
21
22
23
24
9
25
26 10 Figure 15: ESEM images of the rough surface of the Texas-Cream carbonate rock (left) and
27 11 uniform smooth surface of CaCO3 calcite crystal (right).
28 12
29
30 13
31
32
33 (a) θ=7° (b) θ=50°
34
35
36
14
37 15 Figure 16: C1-nC4 (a) and C1-nC10 (b) condensate drops on the treated smooth surface of
38
39
16 calcite at 20°C and IFT of 10 mN/m.
40 17
41
42 18 3.5. Unsteady-State Displacement Tests
43
44 19 Following the interesting results of the static contact angle measurements, demonstrating the
45
20 significant dependency of the treated surface oil-repellency on the gas-condensate
46
47 21 composition and IFT, it was decided to verify whether these findings can be extended to the
48
49 22 dynamic flow tests. For this purpose, a Baker-Dolomite (BD-13a) core sample (L=11.3 cm,
50
51 23 D=2.56 cm, ϕ=25% and K=54 md) was treated using the optimised A1 chemical solution, i.e.
52
53
24 2 wt% A1 diluted in methanol. Two binary-mixtures of C1-nC10 (at low- and high-IFT limits
54 25 of 1 and 10 mN/m) and C1-nC4 (at high-IFT limit of 10 mN/m) were employed to perform
55
56 26 the displacement tests at ambient temperature (i.e. 20°C). It should be reminded that the
57
58 27 results discussed above demonstrated promising contact angles for C1-nC10 at high-IFT (i.e.
59
60 28 θ=65º) and to lesser extent at low-IFT values (i.e. θ=30º), whereas for C1-nC4 there was
29 complete wetting of the treated surface (i.e. θ=0°).

16
ACS Paragon Plus Environment
Page 17 of 21 Energy & Fuels

1
2
3
1 3.4.1. Conditions of the Experiments Performed
4
5 2 The unsteady-state tests corresponding to the fluid mixtures mentioned above were carried
6
7 3 out at the pressures pertinent to the IFT condition of each fluid system listed in Table 1. Here
8
9 4 it should be mentioned that, similar to the majority of contact angle measurements performed
10
11
5 in this study, all the fluids were mixed and employed at room temperature, i.e. 20ºC, while
12 6 the treatment stage was conducted at 130ºC.
13
14 7 To carry out the pre-treatment displacement tests, the dry core was initially saturated with the
15
16 8 dry methane gas at the test pressure. Subsequently, the gas-mixture was injected into the core
17
18 9 for several pore volumes to displace the resident methane gas completely. Then the oil cell
19 10 was opened to the core inlet and the oil was injected at a constant flow rate of 10 cc/hr,
20
21 11 equivalent to the actual pore velocity of 1.9 m/d. The differential pressure between the core
22
23 12 inlet and outlet and produced volumes of the gas and oil at ambient conditions were recorded
24
25 13 subsequently. After each displacement test was accomplished, the core was dried inside the
26
27
14 oven to be prepared for the next test using a different fluid mixture.
28 15 Following the pre-treatment displacement tests, the dry BD-13a core sample was treated by
29
30 16 injecting about 14 PV (205 cc) chemical solution at a constant flow rate of 8 cc/hr through
31
32 17 the core pre-saturated with N2 at 900 psig for approximately one day, after which the core
33
34 18 was soaked in solution for an additional 14 hours by shutting the core holder inlet and outlet
35 19 valves. The unabsorbed chemicals were subsequently removed by flushing the core with N2
36
37 20 gas. The core was then dried to be used for the post-treatment displacement tests following
38
39 21 the same procedures explained for pre-treatment tests. It should be mentioned that the
40
41 22 absolute permeability of the core to the single phase gas was measured at high pressure
42
23 before and after the treatment. The permeability of 54 mD was recorded in both cases,
43
44 24 demonstrating that the rock permeability has remained unaffected by the adsorption of
45
46 25 chemicals on the rock surface.
47
48
49 26 3.4.2. Results and Discussion
50
51 27 Figure 17 shows the differential pressure (DP) versus time before (BT) and after treatment
52 28 (AT) for the two fluid systems used here. Significant decrease in DP is evident for both C1-
53
54 29 nC10 mixtures at high- and low-IFT values, confirming the effective level of wettability
55
56 30 alteration occurred with respect to this fluid system. Table 3 also shows the calculated
57
58 31 trapped gas saturation at the breakthrough time for each test before and after the treatment. It
59
32 is noted that the pre-treatment trapped gas saturation at the breakthrough point for two cases
60
33 using C1-nC10 (at IFT~10 and 1 mN/m) is 43% and 60%, respectively. These values have

17
ACS Paragon Plus Environment
Energy & Fuels Page 18 of 21

1
2
3
1 noticeably been reduced after the wettability alteration process to 30% and 49%, respectively.
4
5 2 The observed trends confirm that after the treatment the rock wettability has been altered
6
7 3 effectively to intermediate gas-wet conditions whereby the oil phase with less wetting
8
9 4 characteristics occupies the larger pores and flows with the less difficulty, resulting in smaller
10
11
5 trapped gas volume and less differential pressures during the oil injection. The results of the
12 6 experiments performed using the C1-nC4 mixture, contrary to those with the C1-nC10
13
14 7 system, demonstrated the minimal impact of the wettability alteration on improving the oil
15
16 8 mobility as almost similar DPs were observed before and after the treatment (Figure 17).
17
18 9 Here, it should also be mentioned that the durability of the treatment was also investigated by
19 10 flowing large volume of gas (about 1000 pore volumes) through the core following the above
20
21 11 mentioned displacement tests. Subsequently, the C1-nC10 mixture was employed again to
22
23 12 repeat the corresponding displacement test to IFT of 10 mN/m. Pressure drops similar to
24
25 13 those obtained initially after the treatment was observed, demonstrating the acceptable level
26
27
14 of chemical's durability during the gas production period.
28 15 The general trends observed during the unsteady-state displacement tests performed here are
29
30 16 in close agreement with our previous findings from the contact angle measurement tests. That
31
32 17 is, the complete spreading of the C1-nC4 condensate phase on the treated carbonate surface
33
34 18 observed during the course of contact angle measurements, reflected in the displacement test
35 19 in the form of lack improvement in the oil mobility. On the other hand, promising wettability
36
37 20 alterations were observed in both static contact angle and dynamic flow tests using the C1-
38
39 21 nC10 fluid system. Furthermore, comparison between the results of displacement tests
40
41 22 performed by C1-nC10 and C1-nC4, both at IFT=10 mN/m, strongly confirms the significant
42
23 impact of the heavy-end components, e.g. nC10, in the gas-condensate composition on
43
44 24 reducing the wetting tendency of the condensate phase on the treated surface.
45
46 25
47
48 26 Table 3: Gas saturation calculated at the breakthrough time corresponding to the unsteady-
49 27 state displacement tests before and after treatment for three different fluid systems.
50 Test Conditions Gas Saturation at Breakthrough (%)
51 Mixture, Pressure (psi),IFT (mN/m) Before Treatment After Treatment
52
53 C1-nC10, 1500, 10 43 % 30 %
54 C1-nC10, 3900, 1 60 % 49 %
55 28
56 29
57 30
58
59
60

18
ACS Paragon Plus Environment
Page 19 of 21 Energy & Fuels

1
2
3 16
4 Pre-Treatment, C1-C10, IFT=10 mN/m
5 14 Pre-Treatment, C1-C10, IFT=1 mN/m
6 Pre-Treatment, C1-C4, IFT=10 mN/m
7 12 Post-Treatment, C1-C10, IFT=10 mN/m
8 Pressure Drop (psi)
Post-Treatment, C1-C10, IFT=1 mN/m
9 10
10 Post-Treatment, C1-C4, IFT=10 mN/m
11 8
12
13 6
14
15 4
16
17 2
18
19 0
20 0 2 4 6
21
22 Number of Pore Volume Injected
1
23 2 Figure 17: Pre- and post-treatment differential pressures measured across the BD-13a
24
25 3 carbonate rock during unsteady-state injection of oil with a velocity of 1.9 m/d into the core
26 4 sample saturated with gas-mixture using different fluid mixtures of C1-nC10 and C1-nC4.
27 5
28
29
30 6 4. Conclusions
31
32
7 This paper for the first time presented the results of a unique investigation into the
33 8 performance of wettability modifiers using gas-condensate fluids. In this exercise a series of
34
35 9 static contact angle measurements and dynamic unsteady-state flow tests were performed.
36
37 10 The impacts of fluid composition and interfacial tension (IFT) on the performance of
38
39 11 chemical treatment were studied with the following main conclusions:
40 12 - The performed contact angle tests demonstrated that the performance of wettability
41
42 13 alteration is highly dependent on the gas-condensate molecular composition and the
43
44 14 thermodynamic conditions. Overall, the condensate wetting tendency on the treated
45
46 15 surface considerably increased as IFT decreased. Furthermore, the presence of the
47
16 heavy-end components, e.g. nC10 and nC15, in the fluid system showed a significant
48
49 17 impact on decreasing the condensate wetting characteristics even at low-IFT values.
50
51 18 The dynamic flow tests qualitatively confirmed these trends.
52
53 19 - Increasing the temperature to 50 and 100°C had minimal impact on the oil-repellency
54
55 20 strength of the fluorinated chemical at both low- and high-IFT limits.
56 21 - The same wettability dependency of the treated rock surface to fluid composition (i.e.
57
58 22 heavy-end component) was also observed on smooth surfaces. However, the value of
59
60 23 contact angle on the rough rock surface was higher indicating the positive impact of

19
ACS Paragon Plus Environment
Energy & Fuels Page 20 of 21

1
2
3
1 surface roughness on the rock wettability, a subject that requires further investigation
4
5 2 currently under way.
6
7 3
8
9
10 4 Acknowledgments
11
12 5 The above study was conducted as a part of the Gas-condensate Recovery Studies at Heriot-
13 6 Watt University. This research project is sponsored by: DongEnergy, Ecopetrol-Equion,
14 7 ExxonMobil, GDF, INPEX, Nippon, Petrobras, RWE, Saudi-Aramco and TOTAL, which are
15 8 gratefully acknowledged.
16
17 9 Nomenclature
18
19 10 A1: anionic fluorosurfactant
20 11 BD: Baker-Dolomite
21
12 DP: differential pressure
22
23 13 DSA: drop shape analysis
24 14 HPHT: high-pressure high-temperature
25 15 IFT: interfacial tension
26 16 TC: Texas-Cream
27 17
28
29 18 θ: contact angle measured through the liquid phase
30 19 θ0: contact angle measured immediately after liquid deposition
31 20
32
33 21 References
34
35 22 (1) Thornton, O. F. (1946). Gas-Condensate Reservoirs-A Review, American Petroleum Institute.
36 23 (2) Moses, P. L. (1986). "Engineering Applications of Phase Behavior of Crude Oil and Condensate Systems"
37 24 Journal of Petroleum Technology 38(07): 9.
38 25 (3) Danesh, A. (1998). "PVT and Phase Behaviour of Petroleum Reservoir Fluids." Elsevier Science and
39 26 Technology Books.
40 27 (4) Danesh, Ali, Henderson, G.D., Peden, J.M. 1991. Experimental Investigation of Critical Condensate
41 28 Saturation and Its Dependence on Interstitial Water Saturation in Water-Wet Rocks. SPE Reservoir
42 29 Engineering Journal 6 (3): 336-342.
43 30 (5) Danesh, A., Henderson, G.D., Krinis, D., and Peden, J.M. 1988. Experimental of Retrograde Condensation
44 31 in Porous Media at Reservoir Conditions. SPE Annual Technical Conference, 2-5 October 1988.
45 32 (6) Henderson, G.D., Danesh, A., Tehrani, D. H., Al-Shaidi, S., Peden, J.M. 1998. Measurement and
46 33 Correlation of Gas Condensate Relative Permeability by the Steady-State Method. SPE Reservoir
47 34 Evaluation & Engineering Journal 1 (2):134-140.
48 35 (7) Haniff, M.S., and Ali, J.K. 1990. Relative Permeability and Low Tension Fluid Flow in Gas Condensate
49
36 Systems. SPE, Europec 90, The Hague, 22-24 October 1990.
50
37 (8) Jamiolahmady, M., Danesh, A., Tehrani, D. H., and Duncan, D. B., M. 2000. "A Mechanistic Model of
51
38 Gas-Condensate Flow in Pores." Transport in Porous Media 41(1): 17-46
52
39 (9) Jamiolahmady, M., Danesh, A., Tehrani, D.H., Duncan, D.B. 2003. Positive effect of flow velocity on gas-
40 condensate relative permeability: network modeling and comparison with experimental results. Transport
53
41 in Porous Media 52 (2): 159-183.
54
42 (10) Jamiolahmady, M., Sohrabi, M., Ireland, S., Ghahri, P. 2009. A generalized correlation for predicting gas–
55
43 condensate relative permeability at near wellbore conditions. Journal of Petroleum Science and
56
44 Engineering 66 (2009): 98–110.
57
45 (11) Afidick, Deddy, Kaczorowski, N.J., Bette, Srinivas. 1994. Production Performance of a Retrograde Gas
58
46 Reservoir: A Case Study of the Arun Field. Paper SPE 28749 presented at the SPE Asia Pacific Oil and
59 47 Gas Conference, Melbourne, Australia.
60 48 (12) Engineer, R., 1985. Cal Canal Field, California: Case History of a Tight and Abnormally Pressured Gas
49 Condensate Reservoir. Paper SPE 13650 presented at the SPE California Regional Meeting.

20
ACS Paragon Plus Environment
Page 21 of 21 Energy & Fuels

1
2
3 1 (13) Li, K., Firoozabadi, A. 2000. Experimental Study of Wettability Alteration to Preferential Gas-Wetting in
4 2 Porous Media and Its Effects. SPE Reservoir Evaluation & Engineering Journal 3 (2): 139-149.
5 3 (14) Fahes, M., Firoozabadi, A. 2007. Wettability Alteration to Intermediate Gas-Wetting in Gas-Condensate
6 4 Reservoirs at High Temperatures. SPE Journal 12 (4): 397-407.
7 5 (15) Adibhatla, B., Mohanty, K.K., Berger, P., Lee, C. 2006. Effect of surfactants on wettability of near-
8 6 wellbore regions of gas reservoirs. Journal of Petroleum Science and Engineering 52 (2006) 227–236.
9 7 (16) Bang, V. Pope, G.A., Sharma, M.M., Baran, J.R., Ahmadi, M. 2008. A New Solution to Restore
10 8 Productivity of Gas Wells with Condensate and Water Blocks. Paper SPE 116711 presented at the SPE
11 9 Annual Technical Conference and Exhibition, 21-24 September 2008, Denver, Colorado, USA.
12 10 (17) Fahimpour, J., Jamiolahmady, M., Severac, R., Sohrabi, M., 2012. Optimization of Fluorinated Wettability
13 11 Modifiers for Gas-Condensate Carbonate Reservoirs. Paper SPE 154522 presented at EAGE Annual
14 12 Conference & Exhibition incorporating SPE Europec, 4–7 June 2012, Copenhagen, Denmark.
15 13 (18) Kumar. V., Pope, G.A., Sharma, M.M., 2006. Improving the Gas and Condensate Relative Permeability
16 14 Using Chemical Treatments. Paper SPE 100529 presented at the Gas Technology Symposium, 15-17 May
17 15 2006, Calgary, Alberta, Canada.
18 16 (19) Al-Anazi, H. A., J. Xiao, A. A. Al-Eidan, I. M. Buhidma, M. S. Ahmed, M. Al-Faifi and W. J. Assiri
19 17 (2007). Gas Productivity Enhancement by Wettability Alteration of Gas-Condensate Reservoirs. SPE-
20 18 107493-MS, European Formation Damage Conference, 30 May-1 June, Scheveningen, The Netherlands.
21 19 (20) Norman R. Morrow. 1975. The Effects of Surface Roughness on Contact Angle With Special Reference to
22 20 Petroleum Recovery. Journal of Canadian Petroleum Technology, 14 (4): 42-53.
23 21 (21) Wenzel, R. N. 1936. Resistance of Slid Surfaces to Wetting by Water. Industrial and Engineering
24 22 Chemistry, 28 (8): 988-994.
25 23 (22) Tamai, Y., Aratani, K. 1972. Experimental Study of the Relation between Contact Angle and Surface
26 24 Roughness. The Journal of Physical Chemistry, 76 (22): 3267-3271.
27 25 (23) Schlumberger, ECLIPSE reservoir simulation software. PVTi Reference Manual. Version 2013.1.
28 26 (24) Anderson, W.G.1987. Wettability Literature Survey-Part 4: Effects of Wettability on Capillary Pressure.
29 27 Journal of Petroleum Technology, 39(10): 1283-1300.
30 28 (25) Gao, N. and Yan, Y. 2012. Characterisation of Surface Wettability Based on Nanoparticles. Nanoscale,
31 29 4(7): 2202-2218.
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

21
ACS Paragon Plus Environment

You might also like