Вы находитесь на странице: 1из 22

Journal of Petroleum Science and Engineering 39 (2003) 137 158 www.elsevier.

com/locate/jpetscieng

The role of interfacial rheology in reservoir mixed wettability


E.M. Freer, T. Svitova, C.J. Radke *
Chemical Engineering Department, University of California at Berkeley, Berkeley, CA 94720-1462, USA Received 18 March 2002; accepted 29 September 2002

Abstract Since the early 1950s, industrial researchers have recognized that asphaltenic crude oil/water interfaces form so-called rigid skins. This work emphasizes the role that such oil/water interfacial microstructures play in establishing the mixed-wet state of reservoirs. We utilize a new oscillating-drop dynamic tensiometer that sinusoidally and infinitesimally expands and contracts a crude-oil droplet immersed in brine at a fixed frequency and measures the resulting dynamic interfacial stress from image analysis and axisymmetric drop-shape analysis. Linear viscoelastic theory permits evaluation of the dilatational interfacial elastic storage and viscous loss moduli. We find that for two crude oils, designated as Crude AS and Crude AH, immersed in synthetic sea water, the interface behaves primarily elastically and that the more asphaltenic the oil the stronger is the interfacial elasticity. Moreover, interfacial elasticity grows slowly in time over days and is clearly manifest even when rigid skins are not visible to the eye. Apparently, macroscopic, networked asphaltenic structures slowly evolve in time at the interface. Advancing and receding contact angles are also measured on smooth mica surfaces for the same crude oil/brine systems. We find that water advancing and receding contact angles when measured within hours are about equal (i.e., there is little hysteresis). However, aging of the drop over days dramatically alters the subsequent advancing and receding contact angles. Water receding angles grow somewhat in time, but the corresponding advancing angles increase over days towards 180j or towards complete pinning. Interestingly, the advancing contact angles for both crude oils do not depend on whether the drop is aged in the brine or in contact with the mica surface. Also, the measured, receding contact angles for both crude oils are much higher than those commonly assumed in the literature. Fascinatingly, aging kinetics of the contact angles correlates directly with the aging of interfacial elasticities and interfacial tensions. Based on in situ AFM studies of the asphaltene-coated mica surfaces, we explain why this happens. Upon rupture of the protective water film and adhesion of the oil droplet to the mica substrate, the surface underneath the oil droplet is pockmarked with water-wet patches in a Dalmatian microwetting pattern. To our knowledge the crucial role of oil/water interface aging in controlling wettability changes has not previously been recognized. Finally, by sketching various primary drainage and imbibition pore-level events, we emphasize the importance of the observed changes in contact angles towards the evolution of mixed-wet oil reservoirs. D 2003 Elsevier Science B.V. All rights reserved.
Keywords: Crude oil/brine interfaces; Dilatational surface rheology; Advancing and receding contact angles; Mica surface; Asphaltene deposits; Atomic force microscopy; Mixed wettability

1. Introduction
* Corresponding author. Tel.: +1-510-642-5204; fax: +1-510642-4778. E-mail address: radke@cchem.berkeley.edu (C.J. Radke).

In 1973, Salathiel (1973) demonstrated that waterflooding of crude oils from reservoir cores reached

0920-4105/03/$ - see front matter D 2003 Elsevier Science B.V. All rights reserved. doi:10.1016/S0920-4105(03)00045-7

138

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

very low oil saturations, but after many pore volumes of water throughput. For the proposed recovery mechanism, Salathiel (1973) pictured interconnected oil seeping along pore walls dragged by water flow. He termed the wettability of the core as mixed, whereby portions of the reservoir rock are oil wet and others water wet, but each is continuous in the pore space. Although low water saturations exacerbated mixed wettability, asphaltenic oils were a prerequisite. Waterflooding of clean or de-asphalted crude oils yielded water-wet cores with normal high residual oil saturations. Since most crude oils contain asphaltene components, understanding the mixed-wet reservoir state is paramount. Accordingly, much effort has been directed towards elucidating both the chemical and physical origins of mixed wettability, particularly the work of Melrose (1982) and Morrow et al. (Buckley et al., 1989; Jia et al., 1991; Ma et al., 1996; Yildiz et al., 1999; Zhou et al., 2000) and Buckley (1993, 2001; Buckley and Liu, 1998). Based on this body of knowledge, Kovscek et al. (1993) outlined a pore-level scenario of how mixed wettability might evolve in a nascent reservoir. These authors added two important features to the original insights of Salathiel: pore corners and rupture of protective water films. Kovscek et al. (1993) argued that upon initial invasion of oil into a water-filled pore, thin water films separate the oil from the pore walls. The stability of the water cushions arises from repulsive forces in the thin films called disjoining pressures. The water films remain stable as long as the capillary pressure in the porous medium does not

exceed a maximum value, Pmax (Basu and Sharma, c 1996; Kovscek et al., 1993). A great deal of effort has now been directed to water-film stability, especially using the so-called adhesion test (Buckley et al., 1997; Liu and Buckley, 1996, 1999; Milter, 1996; Morrow, 1990), although there has been at least one attempt to measure directly the maximum capillary pressure or equivalently the rupture disjoining pressure directly (Basu and Sharma, 1996). As oil migrates into a reservoir, the capillary pressure eventually rises to Pmax c , and the protective water films rupture. In this case, the oil is defined to be adhered to the surface. According to the picture of Kovscek et al. (1993) the walls of the pores where the thin films break become oil wet, whereas the corners of the pores are water filled and remain water wet. Mixed wettability, as defined by Salathiel (1973), evolves in this manner. The pore-level scenario of Kovscek et al. (1993) correctly predicts realistic capillary-pressure curves in mixed-wet rock using a bundle of star-shaped capillaries. More recently, the underlying framework of Kovscek et al. (1993) is adopted in a number of detailed network simulators to predict very successfully two- and three-phase capillary pressures and relative permeabilities and electrical resistivity indices (Blunt, 1998, 2001; Man and Jing, 1999, 2000; Oren et al., 1998). Unfortunately, there are several deficiencies in the original picture of Kovscek et al. (1993). First, these authors did not consider carefully how the pore walls become oil wet upon rupture of the protective water films. Fig. 1, taken from the study of Kaminsky and

Fig. 1. Deposition of asphaltene film on mineral surface (b) after rupture of the protective water film in (a). Trapped water pockets are exaggerated in size.

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

139

Radke (1997), focuses attention on this issue. Fig. 1a accentuates the region near the pore wall early in the drainage process when the thin water layer is present. In this schematic, various proposed asphaltenic species are pictured in the oil phase (Agrawala and Yarranton, 2001; Murgich et al., 1999; Stausz et al., 2002). Agreement on exactly what these species are and how they complex in the bulk oil phase remains elusive. Even more complicated is how the asphaltene components configure at the oil/brine interface. Since the early work of Bartell and Neiderhauser (1949) and others (Kimbler et al., 1966; Reisberg and Doscher, 1956; Strassner, 1968), interfacial films have been noted at the asphaltenic crude oil/water interface, such that upon retraction of an oil droplet in water, wrinkled skins are visible. Simple reversible adsorption at the oil/water interface of the smaller and more polar of the asphaltenic components is an unlikely explanation. More likely, the interfacial region consists of an irreversibly congealed, macroscopic film instead of a reversibly adsorbed monolayer of amphipathic surfactant molecules. Indeed, Neustadter et al. (1979) and Mohammed et al. (1993) demonstrate that crude oil/ water interfaces, especially those from asphaltenic oils, exhibit substantial elastic mechanical strength. Among others, this is one reason why crude oil/water emulsions can be difficult to break (McLean and Kilpatrick, 1997a,b; Strassner, 1968). Fig. 1a diagrams the smaller, more polar components from the oil phase adsorbed onto the rock surface. Kaminsky and Radke (1997) argue that components in the oil phase with even a miniscule amount of water solubility can readily diffuse through the water layer to adsorb at the solid surface on laboratory time scales. Thus, water cushions between the oil and the rock do not protect against solute adsorption from the water phase. Since rock surface adsorption of water-solubilized polar-oil species is permitted on all surfaces of a pore, any subsequent wettability alteration must occur homogeneously. Accordingly, the type of heterogeneous mixed wettability envisioned by Salathiel is precluded. Altenatively, Fig. 1b illustrates the wettability alteration process after water-film rupture. The asphaltenic interfacial film, initially confined to the oil/water interface, now deposits directly onto the rock surface (Reisberg and Doscher, 1956). It is this asphaltenic deposit or coating that apparently leads to the alter-

ation of wettability where the thin water films rupture and no where else along the pore wall (Kaminsky and Radke, 1997; Salathiel, 1973). During the deposition process, some water is inevitably trapped in the asphaltene coating yielding a dalmation pattern of water patches on the solid surface (Kaminsky et al., 1994). During aging, some of this trapped water may migrate from the surface deposit (Liu and Buckley, 1996). It follows from Fig. 1b that the coherent asphaltene-rich film born at the crude oil/water interface controls, in large part, the resulting oil-wetting behavior of the subsequently asphaltene-coated solid surface. In this paper, we investigate the mechanical and, in particular, the aging behavior of the crude oil/water interface and its impact on wettability alteration of the rock surface. A second and major deficiency of the Kovscek et al. (1993) theory of mixed wettability is the imposition of a zero water receding contact angle of the three-phase contact line after film rupture in Fig. 1b and complete pinning of the advancing oil/water interface on the solid surface (i.e., a water advancing contact angle of 180j). In actuality, after breakage of water films, a range of receding and advancing contact angles is expected for various crude oils in different mineral-content and permeability reservoirs, rather than one asymptotic case. Commonly, water receding angles are small, usually less than 30j (Ma et al., 1996; Yang et al., 2002). However, we show later that, when appropriately measured with aged interfaces, water receding angles for adhered oil range well beyond 30j. Values of advancing (hA) and receding (hR) contact angles dramatically control drainage and imbibition pore-level events, thereby dictating oil-recovery behavior. This point is amplified in the detailed discussion in Appendix A where various primary drainage and imbibition pore-level events are categorized, depending on the values of hA and hR after oil adhesion. Fascinatingly, different events occur beyond those enunciated by Kovscek et al. (1993) and Ma et al. (1996), depending on the receding and advancing contact angles that the oil/water interface makes with the pore surface and on the morphology of the pore cross section. Hence, understanding mixed wettability of oil reservoirs demands investigation of what controls advancing and receding contact angles for

140

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158 Table 1 Properties of crude oils Property API gravity Sulfur (wt.%) Nitrogen (ppm) Acid number (mg KOH/g) Kinematic viscosity at 40 jC (cSt) Saturates (wt.%) Aromatics (wt.%) Resins (wt.%) Asphaltenes, n-C7 insoluble (wt.%) Crude AS 22.2 0.52 4306 1.75 37.3 49.7 23.7 24 2.6 Crude AH 24.1 0.75 4806 1.25 38 37.5 34.3 24.3 3.9

asphaltenic crude oils after adhesion of the oil to the solid surface. This is a second goal of the present work. Thus, as opposed to others, we exclusively study contact angles after oil adhesion. We do not focus primarily on water-film rupture and adhesion physics. Both crude oils in this study exhibit adherence. We attempt, as far as possible, measurements of water advancing and receding angles reflective of reservoir processes. Mica is chosen as the solid surface because it is an alumino-silicate mineral and because it is smooth permitting optical visualization of the contact angles (Liu and Buckley, 1999; Yang et al., 1999). The brine is simulated sea water (SSW) containing both calcium and magnesium hardness. Images from in situ atomic force microscopy (AFM) permit study of the deposited asphaltene coatings and their changes during aging. Due to the strong role that oil/water interfacial skins are expected to play in wettability alteration, we also measure dynamic interfacial tensions and, for the first time, the dilatational elastic and viscous moduli of the crude oil/water interface. Formation of rigid skins at the oil/water interface with significant mechanical strength demands interconnection of and growth into large-scale network structures. Such structures are expected to evolve slowly. Therefore, in this study, we also focus on aging of both the oil/water interface and the asphaltene-coated solid surface.

2. Experiment 2.1. Materials Two different crude oils, designated as Crude AS and Crude AH, are used in the experiments. Their physical properties are listed in Table 1, as determined by ChevronTexaco Exploration Production. Resin contents of both crude oils are about the same, but Crude AH contains significantly more asphaltenes in comparison to Crude AS. Remaining properties vary somewhat between the two oils. Simulated-sea-water (SSW) brine solutions are made with distilled water further purified using a Milli-Q filtration unit (greater than 18.2 MV cm resistivity). A liter of synthetic brine contains 24.0047 g

of NaCl, 1.4673 g of both CaCl2 (2H2O) and MgCl2 (6H 2O), 3.9163 g of Na2 SO 4 , and 0.0382 g of NaHCO3 (Liu and Buckley, 1996). All salts are from J.T. Baker Chemical (Phillipsburg, NY) and are of analytic grade. They are used as received. The pH of the prepared synthetic brine is 8.0 F 0.1. All brine solutions are pre-contacted with the oil in a 6:1 volume ratio for at least 8 h to permit equilibration of the brine with the crude oils. The pH of the oil-equilibrated SSW brine remains close to 8. In all experiments described below, the SSW brine is always pre-equilibrated with the crude oil under study. Pure muscovite mica from Ted Pella (Redding, CA) serves as the solid substrate. For each experiment, it is freshly cleaved from the supplied sample using scotch tape and cut into 10 20 mm rectangular slides. The mica slides are then equilibrated with the aqueous phase (that has been previously equilibrated with oil) for at least 3 h before any contact-angle or atomic-force-microscopy (AFM) studies. All experiments are performed at ambient temperature. 2.2. Interfacial tension To determine the dynamic interfacial tension of the crude oil/water interface we use pendant-drop tensiometry, with the less dense oil drop formed upwards at the tip of a U-bent stainless-steel needle (3.2 mm in diameter) immersed in the aqueous brine. The homebuilt apparatus combines both the interfacial tension and interfacial rheology measurements, as illustrated in Fig. 2. The imaging system includes a video camera manufactured by Rame Hart, a Cole-Parmer fiber-optic illuminator, and two polarizers. The polarizers eliminate stray light reflections and also permit fine tuning of the light

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

141

Fig. 2. Oscillatory pendant-drop tensiometer.

intensity. Positions of the camera, sample holder, and drop-dispensing capillary are adjustable in three directions by means of multi-movement Oriel Instruments translation stages. Likewise, fine positioning of the optical glass cell (HellmaR Model 700.00) is obtained using an Oriel Instruments vertical and horizontal translation stage. The optical cell is filled with 30 ml of brine and covered with a 5-ml layer of the crude oil under study to maintain saturation of the water phase with any soluble oil components. The cell is sealed with a TeflonR lid to prevent water evaporation and compositional changes of the oil phase. The entire apparatus is mounted on a pressurized vibration isolation table from Newport (Model VW-3046OPT-2). After forming a fresh oil drop at the needle tip, the dynamic tension is followed in time using axisymmetric drop-shape analysis (Rusanov and Prokhorov, 1996). Image acquisition and regression of the interfacial tension is performed with commercially available Dropimagen software by fitting the Laplace equation to the drop shape. Dropimagen software also controls an automatic pipetting system (manufac -Hart) that maintains constant drop tured by Rame volume for the very long time periods (3 days) over

which dynamic tensions are measured. Typical precision in tension is F 1%. 2.3. Interfacial rheology Simple visual observation of rigid skins when brine-immersed crude oil droplets are retracted provides no quantitative information on their strength. Hence, we measure the surface dilatational storage modulus, EV , and the surface dilatational loss modulus, EW by subjecting the oil/water interface to an infinitesimal periodic expansion and contraction. The surface dilatational modulus is defined as E dr EV iE W dlnA 1

where A is the oil-drop interfacial area and r is the oil/ water interfacial stress. Since the drop area periodically oscillates, the dilatational modulus exhibits two contributions: an elastic part accounting for the recoverable energy stored in the interface (storage modulus, EV ) and the dissipative part accounting for energy lost through relaxation processes (loss modulus, EW). The interfacial storage and loss moduli correspond to the real and imaginary components of the dilatational modulus (Edwards et al., 1991).

142

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

In this work, we apply a periodic strain by differentially oscillating the drop area, and we measure the periodic stress response using pendant-drop tensiometry and axisymmetric drop-shape analysis. Since the drop oscillates, the resulting transient Laplace shapes measure the interfacial stress which includes both isotropic (interfacial tension) and viscous contributions (Edwards et al., 1991). For sinusoidal variations in drop surface area at a given oscillation frequency, EVand EW are independently determined from the following relations (Tschoegl, 1989): EV Dr and EW Dr Ao sinu DA 3 Ao cosu DA 2

surface area and interfacial stress are fit to the following functions A Ao DAsinxt and r ro Drsinxt u 5 4

where Dr is the amplitude of periodic interfacialstress variation, Ao is the unperturbed interfacial area of the drop, DA is the amplitude of periodic interfacial area variation, and u is the phase angle between the periodic stress and strain curves. Results for a typical drop-oscillation experiment are shown in Fig. 3 for Crude AH at an oscillation frequency of x/2p = 0.025 Hz. Experimental data for the measured surface area and interfacial stress are shown as circles and triangles, respectively. To determine the surface storage and loss moduli from Eqs. (2) and (3) above, the

where the unknown parameters ro, Dr, Ao, DA, and u are regressed using a least squares method. Fits of Eqs. (4) and (5) are shown in Fig. 3 as solid and dashed lines, respectively. Once the fitting procedure is complete, the surface storage and loss moduli follow from Eqs. (2) and (3). Miller et al. (1996) provide a comprehensive review of oscillatory pendant-drop tensiometry. Modification of the pendant-drop tensiometer in Fig. 2 enables sinusoidal variations in the drop surface area. Oscillation hardware consists of a 50-ml Hamilton gas-tight syringe (Model 1050) mechanically coupled to a linear piezoelectric actuator manufactured by Physik Instrumente (Model P-840.3). Actuator motion is forced using a Hewlett-Packard function generator (Model 3325A) that is computer controlled with National Instruments LabView software. The piezoelectric actuator is capable of subnanometer resolution ensuring the smoothest possible drop-volume oscillation. Similar to the dynamic-tension measurements above, surface rheological behavior is followed over long time frames. To avoid continually oscillating the drops for such long times, fresh drops are formed for each experiment and aged for the desired amount of time prior to imposing periodic oscillation. Eqs. (2) and (3) demand small strains so that the interface lies in the linear viscoelastic regime. We set DA/Ao at 2.5%, since above a relative strain of about 4.0%, nonlinear effects are seen. Below this value, we find that the surface dilatational moduli are independent of strain. In order to maintain a Laplacian shape for the oscillating drop, we restrict attention to drops that are not highly viscous (Wong et al., 1998). 2.4. Contact angles A second homebuilt apparatus is used to measure the advancing and receding contact angles, as illustrated in Fig. 4. This apparatus is also mounted on a

Fig. 3. Stress response (interfacial stress) to oscillatory strain (surface area) for Crude AH in SSW at x/2p = 0.025 Hz.

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

143

Fig. 4. Contact-angle apparatus.

pressurized vibration isolation table from Newport (Model VH-3036-OPT), and the video-system includes a Pulnix video camera, a Cole-Parmer fiber-optic light source, and two polarizers (see description of the tensiometer above). The positions of the camera, sample holder, and drop dispenser (Gilmont micro syringe with a U-bent stainless steel needle of 0.5 mm diameter) are adjustable in three directions by means of multi-movement optical stages. The position of the 100-ml optical glass cell, filled with the aqueous phase, is also adjustable by movement of a support plate attached to an optical support column. A 90j-bent glass rod serves as the solid-substrate holder. Mica slides are attached to the flattened end of the bent glass rod by melted Paraffin, and the rod is then adjusted to fix the mica slide in the horizontal plane. A drop of oil, usually 1 3 mm3 in volume, is formed underneath the water-immersed mica slide and then is slowly brought into a contact with the solid substrate by the syringe needle. Drop images are captured by an IMAQ frame-grabber and interpreted by an in-house software program (virtual instrument, VI) written in LabView (National Instruments). The VI determines drop edge coordinates, drop height, diameter of the drop-solid contact, and left, right, and average contact angles with a maximum speed of eight measurements per second. 2nd order polynomial fitting of 25 50 points nearest to

the drop edges are used to calculate the contact angles. To check for consistency, a commercial sub-VI provided by National Instruments is also used for contactangle determination. Agreement between the two angle determinations is always within F 2j. For some systems, we compare contact-angle measurements made using our homebuilt setup with those from a Kruss DSA-10 apparatus. Good agreement is found for angles less than 90j (F 0.5j) and reasonable agreement for angles greater than 90j (F 3j). All contact angles in this work are measured through the water phase. Two types of contact angles are measured. While on the syringe capillary, the oil droplet is brought into contact with the mica surface and then slowly increased in volume until the contact line moves outward. This gives the water receded contact angle. After this, the oil droplet is decreased in volume until the contact line now moves inward. This exercise yields the water advanced contact angle. Extreme care must be taken to change the oil-drop volume very slowly to avoid influence of viscous forces on the contact angle. We use flow rates in the syringe of less than 0.2 mm3/min. Advanced and receded angles are studied as function of aging both of the oil/water interface and of the oil droplet adhered to the solid. In addition, we report some measurements of relaxing water advanced contact angles. Here, after initial oil-

144

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

drop attachment and expansion on the mica surface, the drop is left to age for a specified period. Oil is then withdrawn through the syringe needle until an oil neck forms and ruptures. The resulting remnant drop is allowed to self relax as a function of time. We designate this contact angle as a secondary-relaxed advancing angle. 2.5. Atomic force microscopy Morphology of the surface of clean and asphaltenecoated mica substrates is studied using AFM. For in situ AFM imaging, a Digital Instruments (DI) MutliMode SPM Nanoscope II is used in the tapping mode. This mode is convenient for adsorption-layer studies, giving stable and reproducible images (Svitova et al., 2001). Probes are oxide-sharpened silicon-nitride cantilevers (Model NP-S) with nominal spring constants of 0.3 N/m. All studies are performed in the aqueous medium at ambient temperature using the fluid cell supplied by DI. Besides common flattening along scan lines, no other image filtering is performed. Scan rate is usually 1 1.5 Hz, and the driving frequency is in the range of 30 130 kHz. The mica slides, prepared as above, are immersed into the oil-saturated aqueous brine for equilibration with the solid surface. A large (0.5 ml) oil drop is carefully attached to a pre-marked area on the brineimmersed mica surface, aged on the surface for a desired period of time, and then slowly retracted (at less than 1 mm3/min) until the oil column breaks from the capillary tip leaving a remnant oil patch on the mica surface. With the mica slide still immersed in the aqueous brine, excess oil is removed from the asphaltene surface deposit by ultrasonication under constant refreshing of the aqueous phase with distilled/deionized water. During mica sample preparation for subsequent AFM, we take special precautions not to move the mica slide through the water/air interface, thus avoiding deposition of a thin oil film that may have previously spread on the water surface. We choose not to wash the remnant oil left on the mica surface with any solvents (Buckley et al., 1997; Lord and Buckley, 2002; Xie and Morrow, 1998; Yang et al., 1999), as this process likely alters the morphology of the asphaltenic coating. AFM images are taken inside and close to the border of oil-drop/mica contact area. We assert that this sample-preparation method

provides an adequate picture of a natural oil-retraction event.

3. Results and discussion 3.1. Interfacial tensions Fig. 5 reports the dynamic interfacial tensions, c(t), on a semi-logarithmic scale for the two crude oils immersed in the oil-equilibrated SSW. A significant difference in the tension lowering is evident with Crude AS providing more tension reduction than Crude AH. Another important feature of Fig. 5 is the significant aging of the two oil/water interfaces. Crude AS apparently achieves a nominally steady tension value after about 3 4 h, whereas the Crude AH tension continues to fall for up to 3 days, at which time a finite slope remains but the experiment was terminated. When the oil-equilibrated brine in Fig. 5 is replaced by fresh brine not contacted by oil, no rise in tension is evidenced. Hence, the material causing tension reduction in Fig. 5 does not desorb into the aqueous phase. Surface-active species in Fig. 5 are irreversibly attached to the oil/water interface, at least with respect to exchange with the water phase. The long time scales for tension lowering in Fig. 5 are reminiscent of those of large molecular weight mol-

Fig. 5. Dynamic interfacial tension of Crude AH and AS.

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

145

ecules that require long relaxation times for adsorption and reconfiguration at interfaces (Beverung et al., oz et al., 2000). Apparently, asphaltenes 1998; Mun and resins in the oil phase produce network surface structures that slowly evolve at the oil/water interface. If the oil droplet in Fig. 5 is retracted only slightly for Crude AH, a rigid skin is clearly visible. However, for the same retraction experiment, Crude AS does not produce a visible film until the oil is almost completely retracted into the capillary. This difference and the difference in time scales for relaxation of the tension for the two crude oils in Fig. 5 are surely due to the larger asphaltene concentration for Crude AH (see Table 1). The ability of Crude oil AS to lower tension somewhat more effectively may be due to the relatively large ratio of resin to asphaltene concentration. The molecular mechanisms by which tension is lowered when macroscopic skins form are not understood. Indeed, tensions reported in Fig. 5 at the very long times may not arise from molecular-scale phenomena, but rather from a macroscopic elastic interphase that obeys Laplaces equation for drop shape. 3.2. Interfacial rheology Fig. 6 displays the oil/water dilatational surface moduli for the two crude oils as a function of time on

Fig. 6. Dynamic dilatational elasticity of Crude AH and AS.

a semi-logarithmic scale. Lines drawn on this figure simply guide the eye; aging times up to 3 days are investigated. We note that the interfacial loss modulus, EW, is considerably smaller than the interfacial storage modulus, EV , for each crude oil. Thus, the asphaltene films growing at the oil/water interface are primarily elastic in nature. Consistent with the interfacial tensions shown in Fig. 5, the dilatational moduli evolve over very long time periods indicative of interfacial structure development. In particular, Crude AH, which is more asphaltenic, slowly builds surface elasticities that surpass those of Crude AS and that continue to increase in time, just as the interfacial tension of Crude AH continues to fall in time. Similar to the dynamic-tension evolution of Crude AS in Fig. 5, the elasticity of this crude oil/water interface rises more quickly than that of Crude AH, but then levels off. Clearly, this difference in behavior of Crude AH and AS reflects the larger asphaltene content of Crude AH causing slow growth into a strong skin. For a model heptane/xylene oil containing asphaltenes and resins, Mohammed et al. (1993) measured the compressional modulus of the oil/water interface using a Langmuir trough. Similar to our findings, these authors conclude that the rigidity of the oil/water interface arises from the formation of an asphaltenic network structure that strengthens with aging. Identical findings have recently been reported for long-time interfacial network formation of asphaltenes adsorbed at the oil/air interface (Bauget et al., 2001). Note that Crude AS barely displays a visible skin upon drop retraction, but nevertheless, does exhibit substantial dynamic surface dilatational moduli. Hence, reliance only on the appearance of rigid interfacial skins may be misleading, since clearly interfacial elasticity is evident even when skins are not visible to the eye. Figs. 5 and 6 also suggest that the relaxation time scales (Lucassen and van den Temple, 1972) for the Crude AH/water interface are much greater than those of the Crude AS/water interface, a point that is elucidated further when we compare secondary-relaxed advancing water contact angles for Crude oils AH and AS. The most important point from Fig. 6, and also from Fig. 5, is the long time aging of the oil/water interface characteristic of surface asphaltenic networks (McLean and Kilpatrick, 1997a,b; Mohammed et al., 1993; Neustadter et al., 1979). As noted in

146

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

Section 1, we expect that when protective water films rupture, the oil/water asphaltenic film deposits directly onto the reservoir rock. Accordingly, it is the physicochemical characteristics of the oil/water film that initially control wettability alteration. 3.3. Oil adhesion As part of the contact-angle studies, standard adhesion tests were performed (Buckley et al., 1997; Liu and Buckley, 1996, 1999; Milter, 1996; Morrow, 1990). In the synthetic seawater, both crude oils adhere to mica over the pH range from 4.5 to 9.5. Crude AH does exhibit a transition to nonadhesion near pH = 9.5, but we are unable to examine higher pH values because of aqueous hardness precipitation. Thus, for the natural pH = 8 conditions in this study, neither crude oil is protected by a stable water film. This means that upon oil entry into a pristine reservoir, wettability alteration to the mixed-wet state occurs at relatively high water saturations. The receding contact angle and pore shape now determine how mixed wettability evolves during drainage per Fig. A1 of Appendix A. 3.4. Contact angles Figs. 7 and 8 give the main wettability results of this study. Here we graph advanced and receded water

Fig. 8. Receding and advancing contact angles for Crude AS.

Fig. 7. Receding and advancing contact angles for Crude AH.

contact angles of the two crude oils on mica as a function of aging time on semi-logarithmic scales. Crude AH is shown in Fig. 7 and Crude AS in Fig. 8. Lines drawn in the figures again merely aid the eye. Aging time in these plots represents two different physical processes. First, the oil droplet is aged in the oil-equilibrated SSW and then brought into contact with the mica surface for contact-angle determination (per the procedures described above in Section 2). These data are represented by open symbols. Second, the oil droplet is brought into contact with the mica surface immersed in the oil-equilibrated SSW and left for the aging time before contact-angle measurements. These data are exhibited as closed symbols in the two figures. The fascinating result is that both types of aging give identical results for the advanced and receded contact angles of both crude oils for aging times up to 5 days. Hence, the dominant aging process over this time period is that of elastic skin development at the oil/ water interface, a rather unexpected result. Apparently, upon first rupture of the aqueous layer between the oil drop and the solid substrate, water is left on the solid substrate making aging of the deposited asphaltene coating somewhat akin to the aging of a drop immersed in bulk water. Dalmation water patches (cf. Fig. 1b) are repeatedly observed underneath oil drop-

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

147

lets adhered to solid substrates (Buckley et al., 1997; Ese et al., 2000; Kaminsky et al., 1994; Milter, 1996; Yang et al., 1999) and reconfirmed later here using AFM. The second striking feature from Figs. 7 and 8 is the very large growth in time of the advanced angle. After over 100 h, both crude oils approach hA = 180j, the complete pinning case originally discussed by Kovscek et al. (1993) and illustrated in Fig. A2e and f of Appendix A. However, if not aged, the advanced angle is low, less than 90j, giving totally different pore-level drainage and imbibition events (cf. Figs. A1 and A2). Indeed, if the advanced contact angle after oil attachment is less than the critical pore corner angle, water displacement of oil occurs as if the pore is water-wet yielding high residual oil saturations (see Fig. A2b). Thus, it is crucial, and not well recognized, to age the system before contact angles are assessed. Note that hA for Crude AS ages much more quickly than that for Crude AH. This result emerges directly from the aging of the oil/water interface, as predicted from the dynamic interfacial tension in Fig. 5 and, most strikingly, from the dilatational interfacial elasticities in Fig. 6. The higher asphaltene concentration in the AH oil develops stronger interfacial structures that take longer to form. However, the long-term advanced angles do not seem to differ that much between the two crude oils. Each approaches 180j. Receding angles in Figs. 7 and 8 do not demonstrate as dramatic effects upon aging. However, aging cannot be ignored. Receding angles increase in time, again at a rate dictated by the asphaltene content of the crude oil. Somewhat surprising is the rather large values of hR. As highlighted in the introduction, receding contact angles are normally thought to be less than 30j (Ma et al., 1996; Yang et al., 2002). In our work, receding angles up to 50j are found for the aged AS oil/brine system. This observation has important implications for reservoir wettability alteration, as now oil may directly enter pores whose critical corner angles are less than hR, as illustrated in Figs. A1d and A4 of Appendix A. Perhaps one reason why the receded angles in Figs. 7 and 8 are higher than normally reported is because of the slow contact-line displacement rates employed in our work. By increasing this rate, we find smaller

values for hR (and higher values for hA). Thus, to obtain meaningful receding contact angles, it is important to minimize the rate of contact-line motion, in addition to aging sufficiently long. We also measure secondary-relaxed advancing water contact angles, which reflect the relaxation of a crude-oil droplet over a previous asphaltene surface coating. These are reported in Figs. 9 and 10 for the AH and AS crude oils, respectively. In Fig. 9, we find that after sufficient aging of the AH-crude-oil drop on the mica surface, the advancing angle is high (see Fig. 7) but there is no relaxation of the remnant drop. Rather, once the drop is isolated by needle removal, its configuration remains unchanged. Apparently, the strong elastic skin of Crude AH freezes the drop not permitting relaxation on the time scales investigated. Conversely, the AS crude oil in Fig. 10 does relax. Indeed, Fig. 10 reveals that the secondary-relaxed AS contact angle falls below 90j over the experimental time scale if the drop is aged on the mica surface for less than about 3 h. The results in Figs. 9 and 10 are in accord with the interfacial rheology experiments in Fig. 6 that reveal long elastic relaxation times for Crude AH compared to Crude AS. The importance of secondary-relaxed advancing contact angles lies in the behavior of rivulets that can form on pore walls during forced imbibition, as

Fig. 9. Secondary-relaxed contact angles for Crude AH.

148

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

described by Salathiel (1973). In the converse case, where the secondary-relaxed advanced contact angles are less than 90j, the rivulets now become unstable and breakup in the axial direction leaving trapped oil droplets along the asphaltene-coated pore walls. 3.5. Atomic force microscopy Figs. 11 and 12 show, respectively, AMF images of what is left on the SSW-immersed mica surface after contact with Crude AS and Crude AH oil droplets, aging for 3 days, and subsequent ultrasonic washing, as described in the experimental section. In Fig. 11 for Crude AS, we observe the asphaltene deposit near the drop edge. Thus, the smooth upper right portion of this picture shows the mica surface beyond drop contact. Here the mica remains water wet. Only directly beneath the drop is there any asphaltenic material that changes surface wettability. In the lower left part of Fig. 11, we see the asphaltenic oily deposit. It is composed of oil microdroplets protruding from bare mica regions that originally correspond to trapped water droplets, as pictured in Fig. 1B. This leads

Fig. 10. Secondary-relaxed contact angles for Crude AS.

described in Fig. A2f. For secondary-relaxed advanced contact angles greater than 90j, rivulets are unconditionally stable and can slowly produce oil as

Fig. 11. AFM of asphaltene/oil deposits on mica: Crude AS aged for 3 days.

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

149

Fig. 12. AFM of asphaltene/oil deposits on mica: Crude AH aged for 3 days.

Fig. 13. AFM of asphaltene/oil deposits on mica: Crude AS aged for 3 weeks.

150

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

to the so-called dalmation wetting patterns previously reported by Kaminsky et al. (1994) and others (Buckley et al., 1997; Buckley, 2001; Ese et al., 2000; Yang et al., 1999). Note that for Crude AH in Fig. 12, the oil microdroplets are somewhat larger, but rather sparsely distributed compared to Crude AS after 3 days of contact. Apparently, the stronger rigid oil/water films for Crude AH initially trap more water adjacent to the mica surface. Microscopic water/oil contact angles for the surface droplets, as estimated from AFM image analysis, are f 150j for Crude AS and f 155 160j for Crude AH. These values are in reasonable agreement with the macroscopic, water advancing angles (cf. Figs. 7 and 8) of these oil drops on mica. There is not a large difference between Crude AS and Crude AH microdroplets after 3 days of contact with mica. Next, Figs. 13 and 14 display AFM images of Crude AS and Crude AH oil remnants deposited on the mica surface after 3 weeks of aging in SSW brine. For Crude AS in Fig. 13, the microdroplets appear

quite similar to those after aging for 3 days in Fig. 11. However, there are fewer trapped-water domains indicating water escape during prolonged aging at the mica surface. More dramatic aging behavior is observed for Crude AH in Fig. 14. Here there is a much larger amount of asphaltenic deposit with considerably smaller microdroplets that merge into each other. This coating has a distinctive scaly appearance, especially when compared to that for Crude AH reported in Fig. 12. A possible reason is that when the aged oil/water film above the trapped water pockets collapses, a much finer textured coating is deposited on the mica surface. Figs. 11 14 confirm many of the ideas presented in the mixed-wettability picture outlined in the introduction. Wettability alteration arises mainly from an asphaltene coating deposited on the rock surface where intervening water films rupture. Hence, the process is one of deposition of the oil/ water asphaltenic film. Aging at the oil/water interface and aging at the rock surface are both important.

Fig. 14. AFM of asphaltene/oil deposits on mica: Crude AH aged for 3 weeks.

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

151

4. Conclusions Advancing and receding contact angles that emerge after the rupture of protective water films between invading crude oil and reservoir rock are paramount to the evolution of the mixed-wet reservoir state. Configurations of the oil/water interface in cornered pores depend strongly on the pore crosssection shape and on the advancing and receding contact angles. A rich variety of behaviors may arise during drainage and imbibition processes including mixed-wet pores where the pore walls are oil-wet and the pore corners water wet, complete oil-wet pores, and lens, rivulet and oil-globule formation depending on the relative magnitudes of the advancing and receding contact angles and the critical pore-corner angle. When crude oil first invades into a pristine reservoir, asphaltenic material accumulates at the oil/water interface. Depending on how asphaltenic the oil is, rigid skins develop at the oil/water boundary. For the first time, we measure the dilatational strength of these skins using a periodically-oscillating pendant oil drop. For the two oils studied, Crude AH slowly evolves a strong elastic oil/water film. Crude AS, with lower asphaltene content, shows a more quickly developing dilatational storage elasticity. However, leveling off of the dilatational storage elasticity after this initial increase indicates weak network formation. Advancing and receding contact angles after adherence of these two crude oils to mica exhibit dramatic aging behavior with both angles increasing in time over days. Fascinatingly, we find that aged oil/water interfaces exhibit water receding angles that are much larger than the commonly expected value of 30j. Contact-angle maturation parallels that seen in the elasticities of the oil/water interface and indicates that the age of the oil/water interface when protective water films rupture is a critical parameter in the development of mixed wettability. Atomic force microscopy of the asphaltene coating confirms a deposition process whereby asphaltenic material originally at oil/water interface coats directly onto the solid surface once the water film ruptures. The subsequent advancing and receding contact angles of the asphaltene-coated solid surface then control pore-level drainage and imbibition events. Aging of the asphaltene deposit on the

solid surface expels trapped water giving a more coherent and finer textured coating depending on the asphaltene content of the crude oil. Thus, aging of the oil/water interface and the asphaltene-coated surface are both important in the evolution of mixed wettability. Only two crude oils were studied in this work with only one solid surface (mica) and one brine composition (synthetic sea water at a natural pH of 8) and at ambient temperature. Examination of a wider range of oils, solids, aqueous-solution compositions, and temperatures is necessary before the aging behaviors observed here at both the oil/water and solid interfaces can be generalized. Nomenclature A surface area, m2 Ao unperturbed surface area, m2 C mean curvature of oil/water interface, m 1 E dilatational modulus, N/m EV storage modulus, N/m EW loss modulus, N/m g(h) = p 3h + 6cosh sin(p/3 h) i imaginary number Pc capillary pressure, Pa Pmax disjoining pressure of thick wetting-film c collapse, Pa Po oil-phase pressure, Pa Pw water-phase pressure, Pa R radius of largest inscribed circle in pore cross-section, m t time, s x distance from pore corner to arc meniscus, m a half corner angle u phase angle between periodic stress and strain c interfacial tension, N/m r interfacial stress, N/m ro unperturbed interfacial stress (i.e., interfacial tension), N/m hA advancing contact angle hC critical contact angle hL limiting contact angle hR receding contact angle DA amplitude of periodic drop area change, m2 Dr amplitude of periodic interfacial stress change, N/m x oscillation frequency, s 1

152

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

Acknowledgements This work was supported by the U.S. Department of Energy under Contract No. DC03-76SF00098 to the Lawrence Berkeley Laboratory of the University of California. We thank Drs. E. deZabala, J. Creek, and S. Subramanian of the ChevronTexaco Exploration Production for supplying the crude oil samples.

Appendix A . Role of advancing and receding contact angles in pore-level events To illustrate the importance of water advancing and receding angles on the development of mixed wettability in reservoir rock, consider an equilateral trian-

gular pore cross-section with smooth walls, as illustrated in Fig. A1. Such a pore shape is highly idealized, but is sufficient to document the roles that hA and hR play in oil-recovery behavior (Ma et al., 1996). Initially, the pore is filled completely with water, and the solid walls consist of oxide minerals that are naturally water wet to a non-asphaltenic, clean oil. As oil migrates into the reservoir during primary drainage, the capillary pressure, Pc, rises. Since the concepts of wetting and nonwetting phases become ambiguous in what follows, we define the capillary pressure as the difference between the oil and water-phase pressures Pc Po Pw cC A1

where P is pressure with the subscripts o and w denoting the oil and water phases. c is the interfacial

Fig. A1. Primary drainage in an equilateral triangular pore. Shading represents the oil phase and asphaltene deposition on the mineral surface is shown as thick solid lines.

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

153

tension between the oil and water, and C is the mean curvature of the oil/water interface in the pore. Oil first enters the larger pores when the capillary pressure exceeds the entry curvature value corresponding to a zero water receding contact angle (Ransohoff et al., 1987), and drainage commences at C 1:77=R A2 where R is the inscribed-circle radius of the pore. This oil configuration consists of circlular arcs in the projected prospective of Fig. A1. Throughout, positive capillary pressures (i.e., Po>Pw) correspond to oil/water interfaces that are convex to the oil phase. Following classical Deryagin Frumkin theory (Dergaguin, 1955; Hirasaki, 1991), a zero contact angle is equivalent to having a thin water film sandwiched between the crude oil and the pore wall and stabilized by repulsive disjoining forces. This situation is portrayed in Fig. A1 by thin black lines along the solid walls. As oil accumulates in the reservoir, the capillary pressure rises, water recedes at zero contact angle toward the pore corners, and eventually the thin protective water films rupture depositing asphaltenic material, originally formed at the oil/water interface, directly onto the solid walls (Kaminsky and Radke, 1997; Kaminsky et al., 1994; Kovscek et al., 1993). Fig. A1b portrays this series of events. The solid walls in Fig. A1b are shown as dark heavy lines depicting adherence of the oil to the pore surfaces through an asphaltene coating. Because the thin protective water films are now broken, finite receding contact angles emerge in the water-drainage problem. Thin water films rupture at a positive capillary pressure commonly designated as Pmax (Basu and c Sharma, 1996; Kovscek et al., 1993). Smaller pores whose entry capillary pressures lie above Pmax are c invaded and drain slightly differently than those in Fig. A1b. Protective water films never form, and asphaltene material deposits onto the pore walls upon initial oil invasion into the pore. The entry capillary curvature no longer follows Eq. (A2) but corresponds to that of the water receding angle for the asphaltenecoated surface (Ransohoff et al., 1987) g hR =R q C h p i p 3 3coshR F 27cos2 hR 3 3ghR for a p=6 A3

where g(hR) = p 3hR + 6coshRsin(p/3 hR) and a is the half corner angle (i.e., a = 30j for equilateral triangular pores). (Note in this appendix we use angles in radians when they appear in formulae and in degrees otherwise). The entry configuration and subsequent further water drainage is pictured in Fig. A1c. Configurations in this diagram are identical to those occuring later in Fig. A1b after Pmax is exceeded and c the water films break. Thus, after water-film rupture, Figs. A1b,c are identical. Every pore corner has a characteristic critical angle given by hC = 90j a (Concus and Finn, 1974; Ma et al., 1996; Wong et al., 1992). If the water receding angle after water-film rupture is less than the critical angle, the oil/water interfaces advance into the pore corners at hR( < hC) depositing an asphaltene coating underneath until the connate water configuration is established. This scenario is pictured in Fig. A1b,c. Conversely, if the water receding is greater than the pore-corner critical angle, then crude oil completely fills the pore, coating the pore walls everywhere with asphaltenes, including in the pore corners. Fig. A1d illustrates this situation. Fascinatingly, in this case, the pore may be considered as oil-wet even though the receding water contact angle is less than 90j. As noted in the introduction, it is often thought that water receding angles are small, usually less than 30j (Ma et al., 1996; Yang et al., 2002). Thus, common perception is that complete oil filling in Fig. A1d does not happen (unless, of course, hR>90j). Accordingly, Figs. A1b,c represent the expected behavior. However, our contact-angle measurements in Fig. 8 demonstrate that water receding angles can be significantly greater 30j and can realistically exceed the critical pore-corner angle, hC. Hence, it is possible for the configuration in Fig. A1d to emerge. Such a case is characterized by heterogeneous wettability with some pores completely oil filled and others that are small and completely water filled. This type of mixed wettability is contrasted to that in Figs. A1b,c where the wettability is different within the same pore. Our discussion of primary drainage thus emphasizes the need for understanding hR once water films break and the crude oil adheres to the rock surfaces. We turn our attention now to the primary imbibition process where the role of the water-advancing angle is emphasized. A number of subcases arise depending on the magnitude of hA and on whether the connate-water

154

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

saturation corresponds to configurations in Figs. A1a d. We discuss each of these cases in turn. Fig. A2a reflects the first and simplest case where connate-water saturation is at a capillary pressure below that of water-film rupture (i.e., below Pmax as c in Fig. A1a). Since the water advancing angle with water-film cushions remains zero, the water/oil inter-

face advances, upon spontaneous imbibition, toward the pore inscribed circle in Fig. A2a, a configuration that is unstable (Kovscek et al., 1993). The resulting axial liquid thread in the pore undergoes capillary snap-off resulting in trapped oil blobs (Chambers and Radke, 1990; Gauglitz et al., 1987; Kovscek and Radke, 1996; Ransohoff et al., 1987).

Fig. A2. Primary imbibition in an equilateral triangular pore. Shading represents the oil phase, asphaltene deposition on the mineral surface is shown as thick solid lines, and pinning of the three-phase contact line is represented by small open circles.

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

155

An almost identical case emerges when the initial state is that in either Fig. A1b or Fig. A1c where the pore walls away from the corners are coated by asphaltene deposits yielding a finite water advancing contact angle, hA ( < hC). Since advancing angles are greater than (or equal to) receding angles, the oil/ water interface remains pinned, but bows until the advancing angle is attained. Once this happens and since the water-advancing angle is less than the critical angle of the pore corners, the oil/water interfaces advance until the contact lines touch yielding yet another unstable configuration and trapped residual oil. Fig. A2b illustrates this sequence. The amount of trapped residual oil is less than that for the completely water-wet case in Fig. A2a, but marginally so. A second imbibition behavior emerges for the initial state in Figs. A1b,c whenever hA is greater than hC. In this case, upon spontaneous water imbibition into the pore of Fig. A1b or Fig. A1c, the contact line remains pinned, but the contact angle increases until hC is attained. At this point, the capillary pressure falls to zero, and the oil/water interface is flat. Further water imbibition now occurs under forced conditions where Pc is negative (i. e., the oil/water interface flexes toward the oil phase). The three-phase contact line hinges allowing negative curvatures, but remains pinned until hA is attained. Now once hA is reached and provided that hA is greater than hC but less than 90j, the contact lines translate along the pore walls and away from the corners until they encounter one another, as shown in Fig. A2c. Yet again, this particular configuration is unstable to fluctuations in the interface shape, and the oil core breaks to form trapped oil globules. The amount of residual oil is, however, smaller than either of the two cases described above. The maximum advancing contact angle at which the translating arc menisci meet at along the pore wall is 90j, which is the largest advancing angle for which snap-off can occur. If hA is greater than 90j but less than or equal to 180j hC, another scenario emerges that is shown in Fig. A2d. Here the arc menisci advance toward the pore center and approach closest away from the walls forming oil lenses within the pore. Continued expansion of the oil/water interface leads to lens rupture leaving small oil globules trapped within the pore and oil rivulets on the centers of the pore walls.

When hA exceeds 180j hC, a more complicated but important behavior emerges, as portrayed in Fig. A2e. Here the corner-water contact angle increases until it reaches a limiting value that we designate as hL. The limiting angle corresponds to the particular arc-meniscus curvature that just equals the water-entry curvature into an oil-filled pore at the specified water advancing angle hA. C in Eq. (A3) is now defined by (Ransohoff et al., 1987) replacing the receding angle with the advancing angle g hA =R q i C h p p 3 3coshA F 27cos2 hA 3 3g hA for a p=6 A4

Once the arc meniscus bows to attain the curvature C in Eq. (A4), the angle that water makes with the pore wall defines hL. As opposed to hC, the limiting angle is not a purely geometric quantity, but depends on how far the water/oil interface penetrates into the pore corners at the end of primary drainage (i.e., it depends on the connate-water saturation). Fig. A3 graphs the limiting contact angle as a function of the reduced distance from the pore corner, x/R, where x is the distance along the pore wall from the corner. Clearly, x gauges the water content in the pore (Ma et al., 1996). The limiting contact angle increases almost linearly with x/R. Note that hL is considerably less

Fig. A3. Limiting contact angle for an equilateral triangular pore.

156

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158

than the corresponding advancing contact angle and always lies below 90j. It is larger, however, the larger the value of hA. When the oil/water interfaces in Fig. A2e distend away from the corner such that the contact angle reaches hL, water enters into the center of the pore, and oil lenses form (Kovscek et al., 1993). Eq. (A4) gives the water entry curvature for this event. Thus, the advancing contact angle in the center of the pore (i.e., hA) is not the same as the pinned contact angle in the corners (i.e., hL), but the curvatures of the two oil/water interfaces match. Additional forced imbibition proceeds by the central core of water advancing toward the pore corners with the corner-water contact line remaining pinned but with increasing corner angles greater than hL. Provided oil is everywhere continuous, very low oil saturations can be reached in this manner, but at very slow rates. The configurations in Fig. A2e are those originally of Kovscek et al. (1993) in their description of the origin of mixed wettability in reservoirs. However, the picture of Kovscek et al. (1993) employed only the case of hR = 0 and hA = 180j (complete pinning). Eventually, the expanding oil/water interface in the pore center encounters that of the pore corner leading to lens rupture. This event corresponds to the termination points of hL versus x/R lines in Fig. A3. Oil rivulets appear on the pore walls near the corners, as illustrated in Fig. A2f. Since the water advancing contact angle is greater than 120j, these rivulets are stable (i.e., rivulets with oil contact angles less than 90j are stable to axial breakup whereas those with angles great than 90j are unstable) (Davis, 1980). The stable rivulets slowly empty under the imposed water pressure gradient into oil-continuous regions in surrounding downstream pores. The final imbibition event to describe reflects the connate-water state in Fig. A1d. Here, the water receding angle is greater than the corner critical angle, hC. Water imbibition can only occur under forced conditions and then only for advancing angles greater than 180j hC. As illustrated in Fig. A4, the process is identical to water entry in Fig. A2e, except that oil completely fills the pore corners. As water continues to invade the pore, oil drains from the corners provided there is continuity with oil in other nearby pores. Again, the process is very slow because of the large hydrodynamic resistance for fluid flow in corners (Ransohoff and Radke, 1988).

Fig. A4. Water entry into an oil-filled pore (forced primary imbibition): hR>hC.

The picture painted above is highly oversimplified. Pore cross-sections are taken as triangular, and the pore walls are smooth. There is no recognition of the distinction between pore bodies and pore throats, and no interconnectedness is accounted for. A pore-size distribution is implicit in the discussion but not utilized. Although the rich drainage and imbibition events described above serve as rules for later network simulations, our purpose here is to emphasize the role of the water advancing and receding angles in understanding waterflooding from mixed-wet reservoirs. Rupture of water films and oil adherence to pore walls is a critical foundation for establishing mixed wettability, but the subsequent values of hA and hR, characteristic of oily asphaltene-coated rock surfaces, dramatically control the course of water imbibition and oil recovery.

References
Agrawala, M., Yarranton, H.W., 2001. An asphaltene association model analogous to linear polymerization. Ind. Eng. Chem. Res. 40, 4664 4672. Bartell, F.E., Neiderhauser, D.O., 1949. Film forming constituents of crude petroleum oils. Fundamental Research on Occurrence and Recovery of Petroleum. API, vol. 57, pp. 1946 1947. Basu, S., Sharma, M.M., 1996. Measurement of critical disjoining pressure for dewetting of solid surfaces. J. Colloid Interface Sci. 181, 443 455. Bauget, F., Langevin, D., Lenormand, R., 2001. Dynamic surface properties of asphaltenes and resins at the oil air interface. J. Colloid Interface Sci. 239, 501 508.

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158 Beverung, C.J., Radke, C.J., Blanch, H.W., 1998. Adsorption dynamics of L-glutamic acid copolymers at a heptane/water interface. Biophys. Chem. 70, 121 132. Blunt, M.J., 1998. Physically-based network modeling of multiphase flow in intermediate-wet porous media. J. Pet. Sci. Eng. 20, 117 125. Blunt, M.J., 2001. Flow in porous media-pore-network models and multiphase flow. Curr. Opin. Colloid Interface Sci. 6, 197 207. Buckley, J.S., 1993. Asphaltene precipitation and crude oil wetting. SPE Adv. Tech. Ser. 3 (1), 53 59. Buckley, J.S., 2001. Effective wettability of minerals exposed to crude oil. Curr. Opin. Colloid Interface Sci. 6, 191 196. Buckley, J.S., Liu, Y., 1998. Some mechanisms of crude oil/brine/ solid interactions. J. Pet. Sci. Eng. 20, 155 160. Buckley, J.S., Takamura, K., Morrow, N.J., 1989. Influence of electrical surface charges on the wetting properties of crude oils. SPE Reserv. Eng., 332 340. Buckley, J.S., Liu, Y., Xie, X., Morrow, N.R., 1997. Asphaltenes and crude oil wettingthe effect of crude oil composition. SPE J. 2, 107 119. Chambers, K.T., Radke, C.J., 1990. Capillary phenomena in foam flow through porous media. In: Morrow, N.R. (Ed.), Interfacial Phenomena in Petroleum Recovery. Marcel Dekker, New York, pp. 191 255. Concus, P., Finn, R., 1974. Capillary free surfaces in absence of gravity. Acta Math. 132 (3 4), 177 198. Davis, S.H., 1980. Moving contact lines and rivulet instabilities: Part 1. The static rivulet. J. Fluid Mech. 98, 225 242. Dergaguin, B.V., 1955. Definition of the concept of and magnitude of the disjoining pressure and its role in the statics and kinetics of thin layers of liquids. Colloid J. 17 (3), 191 197. Edwards, D.A., Brenner, H., Wasan, D.T., 1991. Interfacial Transport Processes and Rheology. Butterworth-Heinemann, Boston. Ese, M.-H., Sjo blom, J., Djuve, J., Pugh, R., 2000. An atomic force microscopy study of asphaltenes on mica surface: influence of added resins and demulsifiers. Colloid Polym. Sci. 278, 532 538. Gauglitz, P.A., St. Laurent, C.M., Radke, C.J., 1987. An experimental investigation of gas-bubble breakup in constricted square capillaries. J. Pet. Technol. 39 (9), 1137 1146. Hirasaki, G.J., 1991. Wettability: fundamentals and surface forces. SPE Form. Eval. 6, 217 226. Jia, D., Buckley, J.S., Morrow, N.J., 1991. Control of core wettability with crude oil. SPE 21041 SPE International Symposium on Oilfield Chemistry, Anaheim, CA, Feb. 20 22. Kaminsky, R., Radke, C.J., 1997. Asphaltenes, water films, and wettability reversal. SPE J. 2, 485 493. Kaminsky, R., Milter, J., Radke, C.J., 1994. Thin films and wettability alteration of silica and calcite surfaces. 3rd International Symposium on Evaluation of Reservoir Wettability and its Effect on Oil Recovery, Laramie, WY, Sept. 21 23. Kimbler, O.K., Reed, R.L., Silberberg, I.H., 1966. Physical characteristics of natural Films formed at crude oil water interfaces. SPE J. 6, 153 165. Kovscek, A.R., Radke, C.J., 1996. Gas bubble snap-off under pressure-driven flow in constricted noncircular capillaries. Colloids Surf. Physiochem. Eng. Aspects 117, 55 76.

157

Kovscek, A.R., Wong, H., Radke, C.J., 1993. A pore-level scenario for the development of mixed wettability in oil reservoirs. AIChE J. 39 (6), 1072 1085. Liu, L., Buckley, J.S., 1996. Evolution of wetting alteration by adsorption from crude oil. SPE Form. Eval. 12, 5 11. Liu, L., Buckley, J.S., 1999. Alteration of wetting of mica surfaces. J. Pet. Sci. Eng. 24, 75 83. Lord, D.L., Buckley, J.S., 2002. An AFM study of the morphological features that effect wetting at crude oil water mica interfaces. Colloids Surf. Physiochem. Eng. Aspects 206, 531 546. Lucassen, J., van den Temple, M., 1972. Dynamic measurements of dilatational properties of a liquid interface. Chem. Eng. Sci. 27, 1283 1291. Ma, S., Mason, G., Morrow, N.R., 1996. Effect of contact angle on drainage and imbibition in regular polygonal tubes. Colloids Surf. Physiochem. Eng. Aspects 117, 273 291. Man, H.N., Jing, X.D., 1999. Network modelling of wettability and pore geometry effects on electrical resistivity and capillary pressure. J. Pet. Sci. Eng. 24, 255 267. Man, H.N., Jing, X.D., 2000. Pore network modelling of electrical resistivity and capillary pressure characteristics. Transport Porous Media 41, 263 286. McLean, J.D., Kilpatrick, P.K., 1997a. Effects of asphaltene solvency on stability of water-in-crude-oil emulsions. J. Colloid Interface Sci. 189, 242 253. McLean, J.D., Kilpatrick, P.K., 1997b. Effects of asphaltene aggregation in model heptane toluene mixtures on stability of waterin-oil emulsions. J. Colloid Interface Sci. 196, 23 34. Melrose, J.C., 1982. Interpretation of mixed wetttability states in reservoir rocks. SPE 10971 SPE Annual Conference and Exhibition of the Society of Petroleum Engineers, New Orleans, LA, Sept. 25 29. Miller, R., Wu gel, J., Kretzschmar, G., 1996. Dila stneck, R., Kra tional and shear rheology of adsorption layers at liquid interfaces. Colloids Surf. Physiochem. Eng. Aspects 111, 75 118. Milter, J., 1996. Improved oil recovery in chalk: spontaneous imbibition affected by wettability, rock framework and interfacial tension. PhD Thesis, University of Bergen, Bergen, Norway. Morrow, N.R., 1990. Wettability and its effect on oil recovery. JPT, 1476 1484 (Dec.). Mohammed, R.A., Bailey, A.I., Luckham, P.F., Taylor, S.E., 1993. Dewatering of crude oil emulsions: 1. Rheological behavior of the crude oil water interface. Colloids Surf. Physiochem. Eng. Aspects 80, 223 235. oz, M.G., Monroy, F., Ortega, F., Rubio, R.G., Langevin, D., Mun 2000. Monolayers of symmetric triblock copolymers at the air water interface: 2. Adsorption kinetics. Langmiur 16, 1094 1101. Murgich, J., Abanero, J.A., Strausz, O.P., 1999. Molecular recognition in aggregates formed by asphaltene and resin molecules from athabasca oil sand. Energy Fuels 13, 278 286. Neustadter, E.L., Whittingham, K.P., Graham, D.E., 1979. Interfacial rheological properties of crude oil/water systems. In: Shah, D.O. (Ed.), Surface Phenomena in Enhanced Oil Recovery. Plenum Press, New York, NY, pp. 307 326. Oren, P.-E., Bakke, S., Arntzen, O.J., 1998. Extending predictive capabilities to network models. SPE J., 324 336.

158

E.M. Freer et al. / Journal of Petroleum Science and Engineering 39 (2003) 137158 Wong, H., Morris, S., Radke, C.J., 1992. Three-dimensional menisci in polygonal capillaries. J. Colloid Interface Sci. 148 (2), 317 336. Wong, H., Rumschitzki, D., Maldarelli, C., 1998. Theory and experiment on the low-Reynolds-number expansion and contraction of a bubble pinned at a submerged tube tip. J. Fluid Mech. 356, 93 124. Xie, X., Morrow, N.R., 1998. Wetting of quartz by oleic/aqueous liquids and adsorption from crude oil. Colloids Surf. Physiochem. Eng. Aspects 138, 97 108. Yang, S.-Y., Hirasaki, G.J., Basu, S., Vaidya, R., 1999. Mechanisms for contact angle hysteresis and advancing contact angles. J. Pet. Sci. Eng. 24, 63 73. Yang, S.-Y., Hirasaki, G.J., Basu, S., Vaidya, R., 2002. Statistical analysis on parameters that affect wetting for the crude oil/brine/ mica system. J. Pet. Sci. Eng. 33 (1 3), 203 215. Yildiz, H.O., Valat, M., Morrow, N.R., 1999. Effect of brine composition on wettability and oil recovery of a prudhoe bay crude oil. J. Can. Pet. Technol. 38 (1), 26 31. Zhou, X., Morrow, N.R., Ma, S., 2000. Interrelationship of wettability, initial water saturation, aging time, and oil recovery by spontaneous imbibition and waterflooding. SPE J. 5 (2), 199 207.

Ransohoff, T.C., Radke, C.J., 1988. Laminar flow of a wetting liquid along the corners of a predominantly gas-occupied noncircular pore. J. Colloid Interface Sci. 121 (2), 392 401. Ransohoff, T.C., Gauglitz, P.A., Radke, C.J., 1987. Snap-off of gas bubbles in smoothly constricted noncircular capillaries. AIChE J. 33 (5), 753 765. Reisberg, J., Doscher, T.M., 1956. Interfacial phenomena in crudeoil water systems. Prod. Mon. 21 (1), 43 50. Rusanov, A.I., Prokhorov, V.A., 1996. Interfacial Tensiometry. Elsevier, Amsterdam. Salathiel, R.A., 1973. Oil recovery by surface film drainage in mixed-wettability rocks. J. Pet. Technol. 25, 1216 1224. Stausz, O.P., Peng, P., Murgich, J., 2002. About the colloidal nature of asphaltenes and the mw of covalent monomeric units. Energy Fuels 16, 809 822. Strassner, J.E., 1968. Effect of pH on interfacial films and stability of crude oil water emulsions. JPT 20, 303 312. Svitova, T., Hill, R.M., Radke, C.J., 2001. Adsorption layer structures and spreading behavior of aqueous non-ionic surfactants on graphite. Colloids Surf. Physiochem. Eng. Aspects 183 185, 607 620. Tschoegl, N.W., 1989. The Phenomenological Theory of Linear Viscoelastic Behavior: An Introduction. Springer-Verlag, New York.

Вам также может понравиться