Вы находитесь на странице: 1из 11

Colloids and Surfaces A: Physicochem. Eng.

Aspects 402 (2012) 13–23

Contents lists available at SciVerse ScienceDirect

Colloids and Surfaces A: Physicochemical and


Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa

Removal of crude oil from kaolinite by water flushing at varying salinity and pH
Andrew Fogden ∗
Department of Applied Mathematics, Research School of Physics and Engineering, Australian National University, Canberra ACT 0200, Australia

a r t i c l e i n f o a b s t r a c t

Article history: The ability of crude oil to adhere to kaolinite in an aqueous environment by adsorbing or depositing its
Received 25 October 2011 polar components (asphaltenes and resins), and the ability of subsequent water flushing to remove this
Received in revised form 4 March 2012 bound oil, were compared for two oils and varying salinity and pH of the flushing solution. A scanning
Accepted 6 March 2012
electron microscopy technique was used to image the locations of oil residues on and between kaolinite
Available online 15 March 2012
platelets, while fluorescence spectroscopy yielded the corresponding amount of asphaltenics. Consid-
eration of the mechanisms of electrostatic interfacial attraction and surface precipitation giving rise to
Keywords:
the adsorption/deposition of asphaltenics on kaolinite were used to interpret the extent of their removal
Kaolinite
Crude oil
by flushing. The oil exhibiting greater surface precipitation was fairly unresponsive to flushing, with
Asphaltene displacement of the bulk oil leaving residues substantially lining the kaolinite platelets and filling the
Wettability pores between them. The oil which adsorbed by electrostatic attraction was more amenable to removal
Oil recovery of residues, to partially reinstate the water-wetness of the pristine kaolinite, by flushing with salt solu-
Environmental remediation tions which weakened this attraction. In particular, higher salinity of the flush minimized the attractions
between oppositely charged sites on the heterogeneous interfaces, and pH shift to higher or lower levels
rendered the interfacial charges more similar and mutually repulsive.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction Crude oils and asphaltic petroleum products contain a


high molecular weight, polar fraction called asphaltenes. These
Wettability plays a key role in two-phase flow through porous comprise polycondensed aromatic sheets substituted with hetero-
media, in applications such as formation of crude oil reservoirs and atoms in acid and base groups, and naturally associate with lower
oil recovery by water flooding [1,2], or contamination of subsur- molecular weight analogs termed resins [10]. At locations where
face environments by dense non-aqueous phase liquid (DNAPL) the oil has drained the aqueous phase to a thin film, this film can
and its remediation by aqueous flushing [3]. Mineral surfaces of rupture over time if the interfacial interactions are insufficiently
pores originally saturated with water (brine) are generally intrinsi- repulsive. This allows the asphaltenics to adsorb or surface precip-
cally hydrophilic. Partial displacement of water by the immiscible itate onto the mineral to locally alter its surface chemistry toward
oleic phase during reservoir formation, by upwardly accumulating oil-wetting [1,2]. The tendency for local wettability alteration has
crude oil, or soil contamination, by downwardly migrating DNAPL, been studied for single mineral surfaces in isolation, idealized as
is a drainage process driven by the capillary pressure arising from smooth flat substrates of quartz, mica or glass, by contact angle
the density difference. Displacement mechanisms at the pore scale measurement of oil in aqueous phase or atomic force microscope
have been investigated in idealized porous media, such as glass imaging of the absorbed/deposited asphaltenics [3,11–16]. The
bead or sand grain packs [4–6], etched glass micromodels [4,7] and mixed-wet state of locally adhering or non-contacting oil in pores
epoxy replicas of rock fractures [8]. These surrogate rocks and soils is thought to pertain to the majority of reservoirs. This state has
are amenable to video microscopy of multiphase flow, which has been directly evidenced in glass bead packs drained with crude oil
guided development of theoretical and computational models of and aged [17].
meniscus configurations in pore networks [9]. During drainage, it The wettability and its pore-scale distribution impact the mech-
has been well established that the non-wetting oleic phase enters anisms and efficiency of subsequent oil removal by flooding or
the interiors of sufficiently large pores, with the wetting aqueous flushing. The above-mentioned idealized media have been used to
phase thinning to a film lining their surfaces and filling smaller visualize and quantify the oil displacement and residuals, mainly in
pores. the extreme cases of uniformly water-wetting or oil-wetting pores
[4,5,7]. For the former, slow flow readily removes some of the oil
in larger pores, but isolates the remainder as snapped-off blobs or
∗ Tel.: +61 261254823; fax: +61 261250732. bypassed clusters. In the oil-wet case, oil lining larger pores and
E-mail address: andrew.fogden@anu.edu.au filling smaller ones will tend to be retained.

0927-7757/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfa.2012.03.005
14 A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23

Many reservoir sandstones and soils contain a sizable fraction density 0.9062 g cm−3 , viscosity 77.2 mPa s, n-C7 asphaltene con-
of clays, of which one of the most common minerals is kaolinite. tent 9.0 wt%, and acid and base numbers 0.17 and 2.29 mg KOH/g
The presence of kaolinite reduces pore sizes and increases surface [13]. Both were filtered at 5 psi through three Whatman no. 4 filter
area, also increasing structural heterogeneity due to the variable papers before use. The kaolinite used (KGa-1b; Washington County,
aggregates of its individual platelets and their variable distribu- Georgia; Clay Minerals Society) comprises 96% kaolinite and trace
tion as pore-filling, pore-spanning and/or grain-lining. Extensions dickite, 3% anatase and 1% crandallite plus mica and/or illite, and has
of bead or sand pack studies to include kaolinite are scarce cation exchange capacity 3.0 mequiv./100 g and BET surface area
[6]. Most pore-scale investigations have instead used the natural 11.7 m2 /g [34–36]. Salt solutions were made by dissolving NaCl and
kaolinite-containing rock or soil, in which the distributions of oleic CaCl2 ·2H2 O (both AnalaR) in deionized water (Millipore Milli-Q) at
and aqueous phases after drainage or after flooding or flushing the fixed salt ratio of 90 wt% NaCl and 10 wt% CaCl2 , to the two
have been imaged by cryogenic or environmental scanning elec- total concentrations of 50 g/l (i.e. 770 mM NaCl + 45 mM CaCl2 ) and
tron microscopy (cryo-SEM or ESEM) or confocal laser scanning 1 g/l (15 mM NaCl + 0.90 mM CaCl2 ). Each was vacuum degassed and
microscopy [18–21]. These studies have shown that the pore-scale prepared at three pH values, namely unadjusted (pH 5.9 ± 0.2) or
textural and chemical influences of kaolinite are often at odds. The adjusted to 4.0 ± 0.1 with HCl or 9.0 ± 0.2 with NaOH.
capillarity of its tight pores tends to lock in the original aqueous
phase and limit drainage by oil. On the other hand, kaolinite has a
2.2. Substrate preparation and treatment
high chemical affinity for most crude oils and related DNAPL, thus
during prolonged aging it can alter wettability to become a sink for
A kaolinite substrate comprising a relatively smooth and uni-
oil, limiting its removal by flooding or flushing.
form, porous coat of chemically unmodified platelets on a clean
Over the past decade, experiments mimicking crude oil recovery
microscope glass slide (Knittel Gläser) was prepared by casting a
from clay-rich sandstone reservoirs by draining brine-filled rock
7 wt% aqueous suspension of stable, fine kaolinite using a published
cores with crude oil and aging, followed by flooding with a second
method [28,29]. These thin coats, of ∼0.5 ␮m thickness, remained
brine, have often demonstrated improved recovery using flooding
robust, without platelet movement or loss, throughout all ensuing
brines of reduced salinity [22–26]. The boost is thought to be due to
treatment steps. The sequence of steps was based on procedures
removal of oil associated with kaolinite, as suggested by pore-scale
developed and tested in a recent publication [33]. The first step
cryo-ESEM imaging [25]. However, the intermolecular interactions
entailed exposure of the kaolinite-coated glass in salt solution to
giving rise to this oil liberation remain unclear. One proposed mech-
crude oil, to allow its original water-wet state to be altered by
anism is desorption of divalent cations from kaolinite in response
adsorption/deposition of asphaltenics. This defined the initial state,
to their lowered bulk concentration, causing detachment of asphal-
pertaining to an established oil reservoir or contaminated zone
tenics bonded by these cation bridges [24]. Another hypothesis is
prior to recovery or remediation. The initial state was created by
that adsorption of protons on kaolinite causes an upward shift in
immersing a rectangular piece (7 mm × 26 mm) of the substrate in
pH, increasing the negative charge on both oil and mineral inter-
the initial salt solution (50 g/l at pH 6) in a 3 ml vial, with Oil A or B
faces to favor their mutual repulsion [23,26].
then added on top, and equilibrated for 6 h at 22 ◦ C. The salt solu-
Improved design and predictability of recovery or remediation
tion was then withdrawn by pipette from the vial bottom, and the
strategies in kaolinite-rich rocks or soils requires that macro-scale
aqueous film enveloping the oil-immersed substrate was further
flow experiments and pore-scale imaging are complemented by
drained by centrifuging the vial at 1000 × g for 10 min. The sealed
analyses at length scales approaching the molecular level. A body
vial was then aged for 30 ± 2 days at 60 ◦ C. In the second step, the
of the literature deals with asphaltene adsorption and desorp-
bulk oil was removed by flushing with a second salt solution, as in
tion from solution (usually toluene) onto kaolinite particles [27].
recovery or remediation. With the vial partly immersed in a 60 ◦ C
However, for relevance to flooding or flushing, kaolinite should be
water bath, slow flushing was performed using a syringe pump to
exposed to aqueous and oleic phases in a manner closely imitat-
inject one of the six salt solutions into the vial bottom at 2.5 cm3 /h
ing the real processes. Recent work has prepared model substrates
for 1 h, after which the oil level had risen above the substrate. The
comprising a thin, smooth coat of kaolinite on glass [28,29]. Con-
vial was then centrifuged twice at 1000 × g for 10 min to strip any
tact angles of crude oil drops in brine were measured to probe the
larger oil drops adhering to the substrate. The oil in the upper phase
development of oil–kaolinite adhesion during the early stages of
was decanted and any slick at the air interface was removed using
drainage [28]. Later stages were studied by aging the model sub-
lint-free tissue. In the third step, the substrate was taken from its
strate in oil to analyze the adsorbed/deposited asphaltenics by field
vial and immersed in organic solvents to facilitate analysis of the
emission SEM (FESEM) and fluorescence spectroscopy [29]. The
micro/nano-scale oil residues on the substrate. The solvents used
current study extends this work by performing slow flushing of the
depended on the analysis method, described below.
aged state with brines of varying salinity and pH to investigate the
amount and distribution of oil residues remaining on and between
individual kaolinite platelets. This provides insight into the intrin- 2.3. Analysis of adhering oil residues
sic contribution of oil–brine–kaolinite interactions to the efficacy
of oil removal. For each flushing salt solution, a replicate pair of substrate pieces
was thus treated; one for imaging of adhering oil residues by FESEM
and the other to determine the amount of asphaltenic residues by
2. Materials and methods fluorescence spectroscopy. For the former, the substrate after the
second step was immersed in methanol (99.8%; Sigma–Aldrich) for
2.1. Materials 10 min to remove salt and water, and dried ambiently. FESEM (Zeiss
UltraPlus Analytical) was performed under high vacuum in sec-
Two asphaltic crude oils commonly studied in the literature ondary electron mode at 1 kV, prior to which the substrate was
were used, namely WP (Alaska) [30,31] and Minnelusa (Gibbs Field, lightly sputter coated with platinum. For spectroscopy, the treated
Wyoming) [13,17,28,29,32,33], referred to here as Oil A and B, substrate was instead immersed in decalin (decahydronaphtha-
respectively. At room temperature, Oil A has density 0.9125 g cm−3 , lene, 98%; Sigma–Aldrich) to dissolve bulk oil, first for 20 min and
viscosity 111 mPa s, n-C6 asphaltene content 6.3 wt%, and acid then for 2 h in fresh decalin, which remained colorless. This was fol-
and base numbers 1.46 and 2.49 mg KOH/g [30]. Oil B has lowed by the above-mentioned methanol immersion, and vacuum
A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23 15

drying for 3 h at 22 ◦ C. After cleaning of its uncoated backside, the established in the high salinity solution at pH 6, with subsequent
piece was soaked in 3.1 g of the azeotropic blend of 87.3 wt% chloro- flushing performed using the high or low salinity solution at pH
form (99%; Sigma–Aldrich) and 12.7 wt% methanol, in a sealed flask 4, 6 or 9. Figs. 2 and 3 show representative micrographs of these
at 50 ◦ C for 24 h, to extract the asphaltenics. Their total dissolved six flushed states for Oil A and B, respectively. It must be empha-
amount was determined using a FluoroMax-3 spectrofluorometer sized that these FESEM images cannot be directly compared to
(Horiba Jobin Yvon), utilizing the natural fluorescence of these pol- those in the corresponding initial state (Fig. 1a and c). For the latter,
yaromatic asphaltenics [29]. The solution was analyzed in a quartz decalin rinsing to dissolve the surrounding oil is necessary for imag-
cuvette (10 mm × 10 mm × 40 mm; Starna Cells) to record emis- ing of the substrate, while the displacement of this surrounding
sion spectra over 380–560 nm from 340 nm excitation. In separate oil during salt solution flushing and ensuing centrifugation allows
experiments, the asphaltene fraction of the two crude oils was iso- imaging without the artificial step of decalin post-rinsing. In both
lated by heptane precipitation, from which asphaltene solutions states, methanol post-rinsing is required prior to drying to avoid
were prepared in this same solvent blend. Their fluorescence was salt crystallization. While the initial-state samples bear only the
analyzed similarly to calibrate intensities of sample extracts to con- adsorbed/deposited asphaltenics, the flushed-state counterparts in
centration. Figs. 2 and 3 bear all adhering oil components, aside from the more
Other related control experiments were performed to exam- volatile fraction evaporated during drying and in the vacuum of the
ine the initial state, in the absence of salt solution flushing, by FESEM chamber.
omitting this second step. In these cases, the substrate after the The oil residues in Figs. 2 and 3 appear fluidic, as was also the
first step of oil drainage of the initial salt solution (50 or 1 g/l case in a recent study [33] of smooth glass substrates exposed to
at pH 6) and aging was directly immersed in decalin, and then Oil B and flushed with various NaCl solutions using an analogous
methanol, and vacuum dried, as per the above-mentioned pro- procedure. There, the remaining oil typically took the form of a
cedures. The adsorbed/deposited asphaltenics were analyzed by sparse scattering of nano-droplets, implying that the oil interfacial
FESEM as above. film at the substrate in the initial state becomes increasingly per-
forated by the advancing water during flushing, partly by swelling
of the undrained trapped water. The oil film retracts and detaches
3. Results and discussion with the bulk oil, decomposing to a foam-like network of connected
struts, which decompose further to nano-droplets via Rayleigh
3.1. Morphology of oil residues on kaolinite instability. Similar mechanisms apparently occur during flushing
of the kaolinite coats, although the oil retraction and detachment is
For each treated kaolinite substrate, 20 FESEM images at 100k× less fully developed than for glass. As oil substantially occupies the
magnification were acquired at a fixed pattern of locations over the inter-platelet pores within the coat in the initial state, the amount
piece. Fig. 1 shows representative micrographs of the initial wet- of surface oil and its anchoring are clearly greater than for a smooth
tability state, resulting from displacement of the bulk salt solution impervious substrate. The platelet edges and gaps between them
by the crude oil and adsorption/deposition of its asphaltenes and also provide sites where oil retraction can terminate or become
resins on exposed platelets during the 1 month of aging. The subse- snagged, leading to its disconnection from the bulk oil. Accord-
quent decalin rinse removes all bulk oil to leave only this adhering ingly, the oil residues on and in the kaolinite coats generally take
asphaltenic footprint on kaolinite. For both oils and both concen- the form of more irregularly shaped nano-droplets, together with
trations of the pH 6 salt solution, the thin films of water sheathing larger blobs and pools.
the platelets rupture to allow a significant alteration of their wetta- By the same token, Figs. 2 and 3 demonstrate that the asphalten-
bility. The adsorbate/deposit takes the form of primary-aggregated ics adsorbed/deposited on kaolinite in the initial state can locally
asphaltenic nanoparticles which often build irregular, secondary- detach with the oil during flushing. Further, this response can be
aggregated chains. All visible platelets, at the coat surface or one or quite rapid, as the average rate of advance of the bulk water front
more layers down, bear this same texture. Moreover, in the occa- over the substrate during injection is 10 ␮m/s, amounting to traver-
sional locations at which the underlying glass can be seen through sal of the length of the images of Figs. 2 and 3 in ∼0.3 s. The lesser
inter-kaolinite pores, it too bears this nanoparticulate deposit. As residues after flushing of Oil A in Fig. 2 compared to B in Fig. 3 tally
the capillary pressure applied during the centrifugal drainage is far with its deposits in the initial state of Fig. 1a being thinner and of
below that required to force oil into such tight pores, the oil infiltra- lower coverage than those of Fig. 1c. Thus for Oil A, somewhat less
tion occurs spontaneously during aging, driven by the concomitant oil is initially present in coat pores, its bonding area with kaolin-
alteration of wettability to progressively favor expulsion of water. ite is smaller, and the more connected nano-domains of undrained
From Fig. 1 it is apparent that Oil B produces thicker deposits, water give greater access pathways during flushing. Note that the
which more completely cover platelet faces and extend to lining solidified deposits after decalin rinsing in Fig. 1 are not indicative
their edges. A previous study [29] of the initial wettability state cre- of their fluidic, responsive nature in the oil. In particular, analyses
ated by Oil B on kaolinite showed that this substantial and pervasive [11,13,29] in which such rinsing of the initial state is performed
deposition was manifested for most of the salt solutions tested. For for subsequent measurement of contact angles of a probe oil (e.g.
Oil A, the coexistence of deposit with uncovered nano-domains, due decane) in salt solutions on the footprint-bearing substrate will
to undrained trapped water, on platelets is far more pronounced. under-represent the in situ changes in wettability during flushing.
In comparison, the influence of salinity in Fig. 1 is less apparent, For Oil A, its residues strongly depend on the salinity and pH
although the 1 g/l solution leads to somewhat greater deposition of the flushing solution. The 50 g/l flush at pH 4 in Fig. 2 exhibits
from Oil B than its high salinity counterpart, which is also in line isolated nano-droplets on platelet faces, and also tendril-shaped
with the trend from the previous study of this oil [29]. Some stud- analogs lining edges, while the inter-kaolinite pores are mainly
ies [21] have inferred that asphaltic DNAPL only penetrates pores vacated. For the 50 g/l flush at pH 9, residues are somewhat more
larger than ∼50 nm in clay-rich soil aggregates. As most pores in prevalent, as nano-droplets on faces and larger blobs stranded in
the kaolinite coats are slightly wider than this, it is not possible to pores. The high salinity flush at intermediate pH 6 shows oil films
determine whether such a lower bound applies here. sheathing many platelets and substantially filling pores, smoth-
The remainder of this sub-section addresses FESEM images of ering much of the angular features of the kaolinite coat. The
the oil residues remaining on and in the kaolinite coats after salt corresponding images in Fig. 2 for the 1 g/l flush show similar
solution flushing of their aged, initial state. This initial state is trends in pH, although in each case with systematically greater
16 A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23

Fig. 1. FESEM images of asphaltenics bonded to kaolinite coats after drainage of the initial salt solution of 50 or 1 g/l at pH 6 by Oil A or B and aging, without flushing prior
to decalin–methanol post-rinsing. Images are 3.0 ␮m × 2.0 ␮m; the scale bar in the bottom left image is 500 nm and applies to all.

Fig. 2. FESEM images of Oil A residues on/in kaolinite coats after flushing with high (left column) or low (right column) salinity solution at the three pH values. Samples were
methanol post-rinsed. Images are 3.0 ␮m × 2.0 ␮m; the scale bar in the bottom left image is 500 nm and applies to all.
A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23 17

Fig. 3. FESEM images of Oil B residues on/in kaolinite coats after flushing with high (left column) or low (right column) salinity solution at the three pH values. Samples were
methanol post-rinsed. Images are 3.0 ␮m × 2.0 ␮m; the scale bar in the bottom left image is 500 nm and applies to all.

residues than for 50 g/l flushing. For Oil B in Fig. 3, its residues 3.2. Amount of asphaltenics on kaolinite
are much greater, in the form of more stranded oil lining platelet
edges, sheathing faces and filling pores. Its thicker and more com- The residues, such as in Figs. 2 and 3, after displacement of
pletely covering layer of adsorbed/deposited asphaltenics in the oil from a kaolinite-bearing rock or soil pore are indicative of the
initial state (Fig. 1c) presumably limits the dynamics and extent adhesion which resisted this removal. Further, for bulk oil which
of retraction and detachment during flushing. Even for the 50 g/l subsequently enters the pore during the course of continued flush-
flush at pH 4 which best removes Oil B, the residue is high by the ing, the residues will influence its ability to adhere to kaolinite and
standards of Oil A. Its low salinity counterpart at pH 4 appears to resist displacement. Even in the idealized case of these relatively
be least capable of removing Oil B from kaolinite. smooth, coated, planar substrates, occasional larger oil droplets or
In a recent study [33], the Oil B residues on glass after flushing puddles (of micron to millimeter size) remain after flushing and
could be quantified by segmentation and analysis of FESEM images, centrifugation. Wetting and adhesion of newly arriving oil is thus
yielding measures of area coverage of the substrate and lateral dictated by the contribution of nano-scale residuals to the surface
size of nano-droplets. For kaolinite-coated glass, the FESEM surface chemistry and larger-scale residuals to bulk coalescence. Contact
images do not reveal all residues, and those that are visible are dif- angle measurements of a macroscopic pendant drop of oil con-
ficult to segment from the locally varying brightness of the stacked tacted with the treated substrate in the salt solution would thus
platelets. However, one measure was able to be obtained, as shown be problematic to perform reliably and to interpret.
in Fig. 4. The edge-finding tool of the ImageJ software was applied to Accordingly, the sister samples of those analyzed above by
each FESEM image to identify and count the white edge pixels, i.e. FESEM were instead characterized here by their total amount of
of grayscale value (GSV) 255. This was compared to the correspond- asphaltenics in the residues. As mentioned in Section 2.3, this
ing average edge pixel count for micrographs of untreated kaolinite necessitated decalin post-rinsing of the flushed substrates to dis-
coats to determine the percentage of platelet edge pixels obscured solve the other oil components. Recent work [33] showed that such
by residues. This measure was termed the SEM image edge loss, and rinsing of flushed states retains the substrate-bound asphalten-
its average and standard deviation over the 20 FESEM images for ics, but causes asphaltenics at the oil–water interface of droplet
each sample were calculated, and are plotted in Fig. 5. The trends residues to also be deposited on the substrate. Thus the amounts
agree with the above-mentioned hierarchy of remaining oil from determined by their chloroform/methanol extraction and solution-
inspection of Figs. 2 and 3. Edge loss for Oil A peaks at flush pH 6, phase fluorescence spectroscopy are best regarded as including
and decreases with salinity, in line with the observed extent of its both bonded and non-bonded asphaltenes. FESEM images in Fig. 6
residues. Edge loss for Oil B is substantially greater, and approaches of kaolinite substrates after extraction, for the Oil A and B sam-
100%, i.e. with residues almost totally obscuring platelets. ples with most residue after flushing, verify that the extraction of
18 A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23

Fig. 4. FESEM images of oil residues on/in kaolinite coats after flushing with (a) 50 g/l salt solution at pH 4 for Oil A, (b) 50 g/l at pH 6 for Oil A, and (c) 1 g/l at pH 4 for Oil
B. Samples were methanol post-rinsed. Images are 3.0 ␮m × 2.0 ␮m; the scale bar in (a) is 500 nm and applies to all. Images at right highlight the edges (bright) within the
micrographs at left, from which the edge loss metric is calculated.

asphaltenics is very close to complete, as also seen in a previous Fig. 7 plots the difference in fluorescence emission intensity
study [29]. The slight residue persisting after extraction corre- between solution and background solvent for the calibration solu-
sponds to the occasional lighter dots in Fig. 6a and faint trails in tions of pure asphaltenes of the crude oils and the extracts from the
Fig. 6b. treated kaolinite samples. The calibration solutions for Oil A and B
display a broad peak (dotted line) at 499 and 492 nm in Fig. 7a and
b, respectively. In both cases the peak intensity is very well fitted
100 to the power law in asphaltene concentration shown in Fig. 7c and
d, with exponent close to unity. The highest concentration point is
90
SEM image edge loss, %

not included in the fit, as its slight downturn marks the onset of re-
80
absorption of emitted fluorescence. For the spectra of the extracted
70 residues from the six flushed substrates, the average peak position
60 is shifted substantially for Oil A (to 476 nm in Fig. 7e) and slightly
50 for Oil B (to 494 nm in Fig. 7f). Moreover, in both graphs the max-
40 ima for the pH 6 and 9 flushings lie systematically to the left of
30
these dotted line averages, while those for pH 4 lie to the right.
These shifts relative to Fig. 7a and b are due to the added pres-
20 Oil A, 50 g/l Oil B, 50 g/l ence of resins in the extracts, together with a possible preference
10 Oil A, 1 g/l Oil B, 1 g/l for the flush pH to remove polar components bearing acid or base
0 groups. However, without calibration spectra for these subclasses
3 4 5 6 7 8 9 10 in isolation, no further interpretation of the shifts can be made.
Flush pH From the intensities at 499 and 492 nm in Fig. 7e and f for Oil
A and B, respectively, the equivalent concentrations of extracted
Fig. 5. Percentage of kaolinite platelet edges obscured by oil residues in FESEM
images for the two oils flushed with salt solutions of varying concentration and pH,
asphaltenics were back-calculated from Fig. 7c and d, and from the
and post-rinsed with methanol. known mass of solvent used, the solute mass was determined. The
A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23 19

Fig. 6. FESEM images of kaolinite coats after flushing with 1 g/l salt solution at (a) pH 6 for Oil A and (b) pH 4 for Oil B, with decalin–methanol post-rinsing, followed by
chloroform/methanol extraction of asphaltenic residues. Images are 3.0 ␮m × 2.0 ␮m; the scale bar in (a) is 500 nm and applies to both.

masses are presented in Fig. 8a, normalized by the total planar area calcium ions are also potential determining; their binding to kaoli-
of the underlying glass piece (as distinct from the kaolinite particle nite basal planes is evidenced by the shift in its isoelectric point
surface area). The overall trends between the samples are not influ- (IEP) from pH ∼2 in the absence of calcium [28] to pH ∼5 in the
enced by whether the wavelengths corresponding to the peaks of 1 g/l solution of Fig. 11b. Calcium also binds to acid groups at the
the calibration or extract solutions are used. The plots are very sim- oil interface, apparent from the corresponding increase in IEP of
ilar to their analogs for the edge loss measure from FESEM image Oil B from pH ∼4 [28] to ∼5. In spite of its substantially higher acid
analysis in Fig. 5, as shown by the correlation in Fig. 8b. The agree- number, Oil A has similar, but slightly higher, zeta potential to Oil
ment suggests that the FESEM images such as in Figs. 2 and 3 are B in Fig. 11, and thus presumably also binds calcium to boost its
indeed reasonably representative of the entire sample. Compared positive charge.
at the same flushing conditions, the residue amount for Oil B in In addition to these considerations of the charges on the oil
Fig. 8a is on average 2.4 times greater than for Oil A. The decrease and kaolinite interfaces and their electrostatic interactions, the
in flush salinity at fixed pH increases residue amount for Oil A on likelihood and extent of asphaltene deposition on the substrate
average by a factor of 1.4, while for increase in flush pH from 4 to will increase if they are less stably dispersed in the oil [15,16].
6 at fixed salinity, the corresponding averaged factor is 1.9. The oil’s robustness against asphaltene flocculation can be quan-
A previous study [29] of the initial state of wettability alteration tified by the decrease in its refractive index, RI, required to reach
of kaolinite substrates by Oil B showed that the relative amount the onset of flocculation, PRI , by the addition of precipitant (n-
of adsorbed/deposited asphaltenics could be estimated by a simple heptane) [16]. From the CO-Wet database [15], the value of RI − PRI
non-destructive technique in which the sample pieces were digi- is 1.5184 − 1.4469 = 0.0715 for Oil A, with 57% heptane in the onset
tally scanned with a document scanner (Konica Minolta C252) for mixture, while that for Oil B is 1.5144 − 1.4831 = 0.0313, with 26%
image analysis. The reduction in average GSV over each piece due heptane at onset. This difference for Oil B is very low relative to the
to the brownish tinge of the deposits (compared to an untreated average of ∼0.06 over all crude oils in the database [15], hence it is
coated substrate) correlated well with their absolute amount from prone to asphaltene precipitation.
extraction and fluorescence spectroscopy. The scanned images of During establishment of the initial state in Fig. 1a and c at pH 6,
the flushed samples for Oil B are displayed in Fig. 9, and Fig. 10 it appears that the kaolinite and oil interfaces (for Oil A and B) are
shows that the average GSV loss due to the asphaltenic residues is sufficiently close to their IEP’s for a substantial fraction of positive
reasonably correlated to their amount in Fig. 8a. This lends further and negative charges to coexist on each, leading to local attraction
support to the trends described above. For Oil A, the residues were of oppositely charged interfacial patches to instigate wettability
too slight to be clearly distinguished by scanning. alteration. As the zeta potentials of the two oils are quite similar, the
heavier deposition for Oil B is presumably the result of asphaltene
3.3. Interpretation of mechanisms precipitation accompanying the adsorption.
The lesser and more variable amount of residues of Oil A after
The differences in initial-state deposits and flushed-state flushing in Fig. 8a thus probably reflects both the lesser oil amount
residues for the varying oils and salt solutions are dictated by the on and in the kaolinite coat in the initial state and the greater
interactions of the asphaltenes and resins with the other compo- reversibility of electrostatically driven adsorption. For the 50 g/l
nents of the oil, with the aqueous phase at its interface, and in turn flush, the residue is maximized at flush pH 6, matching the pH of
with the kaolinite aqueous interface. The salinity and pH effects the initial state. For flush pH 4, the zeta potentials of both inter-
derive from the electrostatic interactions of the two interfaces. The faces in Fig. 11a become weaker and more positive, while at pH 9
zeta potentials of these interfaces in isolation were measured in both potentials become stronger and more negative. In either case,
all salt solutions with a Zetasizer Nano-ZS (Malvern Instruments), the electrostatic interactions between kaolinite and oil become less
using the Smoluchowski equation, and are plotted in Fig. 11. attractive, favoring removal of oil to give the observed lowering of
Kaolinite and the oil’s polar molecules are amphoteric, so zeta the residue. Alternatively, the flush may proceed too rapidly for
potential transitions from more positive at pH 4, due to base group re-equilibration of the interfacial interactions to the new pH of 4
protonation, to more negative at pH 9, due to acid group deproto- or 9, in which case disturbance of the acid–base adhesive inter-
nation. The magnitude of the potential naturally increases for the actions from their pre-equilibrated initial state at pH 6 would be
lower salinity in Fig. 11b, due to the increased Debye screening responsible for the greater oil removal.
length. This increase in the range of interactions also more clearly For the 1 g/l flush in Fig. 8a, the Oil A residue is also maxi-
reveals the positive surface potential at pH 4 by reducing the con- mized at flush pH 6, and the amounts are consistently higher than
centration of counterbalancing anions in the diffuse double layer for the high salinity counterparts. This latter trend may appear
at the shear plane where zeta potential is measured. The dissolved somewhat incongruous, as reduced salinity increases the strength
20 A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23

a 1.E+06 b 1.E+06

Intensity difference, cps


Intensity difference, cps

1.E+05 1.E+05
9.61E-04
8.32E-04
3.28E-04
2.09E-04
2.70E-04
8.31E-05
1.E+04 1.17E-04 1.E+04 2.01E-05
3.77E-05
7.32E-06
1.46E-05
1.88E-06
4.83E-06
1.E+03 1.E+03

1.E+02 1.E+02
380 400 420 440 460 480 500 520 540 560 380 400 420 440 460 480 500 520 540 560
Emission wavelength, nm Emission wavelength, nm
c d
1.E+06 1.E+06
Intensity difference (499 nm), cps

Intensity difference (492 nm), cps


1.E+05 1.E+05

0.898
y = 301401351.609x
2
R = 0.999 0.935
y = 260068680.044x
1.E+04 1.E+04 2
R = 1.000

1.E+03 1.E+03
1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02

Asphaltene concentration, wt% Asphaltene concentration, wt%

e 1.E+05 f 6.E+04

9.E+04
5.E+04
8.E+04
Intensity difference, cps

Intensity difference, cps

7.E+04 50_4 50_4


4.E+04
50_6 50_6
6.E+04
50_9 50_9
5.E+04 1_4 3.E+04 1_4

4.E+04 1_6 1_6


1_9 2.E+04 1_9
3.E+04

2.E+04
1.E+04
1.E+04

0.E+00 0.E+00
380 400 420 440 460 480 500 520 540 560 380 400 420 440 460 480 500 520 540 560
Emission wavelength, nm Emission wavelength, nm

Fig. 7. Fluorescence emission spectra of the solutions of asphaltenes from (a) Oil A and (b) Oil B, of the wt% concentrations in the legend, yielding the calibration relation
between these concentrations and the peak intensity difference for (c) Oil A and (d) Oil B. Spectra of asphaltenic residues for (e) Oil A and (f) Oil B after flushing the kaolinite
substrates with salt solutions of concentration (g/l) and pH in the legend, followed by decalin–methanol post-rinsing and chloroform/methanol extraction.

of the similarly signed interfacial charges in equilibrium at pH 4 induces an extra osmotic pressure contribution serving to dilute
and 9 in Fig. 11b, and also increases the disturbance of the initial the high salinity solution trapped by oil in the confines of the kaoli-
equilibrated state. However, previous studies [28,29] showed that nite coat in its initial state. This film swelling may disconnect oil
decreased salinity of sodium–calcium brines typically increased in/on the coat from bulk oil ahead of the advancing meniscus of the
short-time adhesion and longer-time adsorption of oil on kaolinite, flush, to thus hinder its removal.
in agreement with Fig. 8a, by aiding local attraction of oppositely As the zeta potential behavior of the two oils is comparable,
charged patches on the two interfaces, even if both net zeta poten- the dependence of the residue amount for Oil B on flush salin-
tials were of the same sign. Flushing with the 1 g/l solution also ity and pH in Fig. 8a resembles that for Oil A, although with the
A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23 21

18
a 30
16

Residue intensity, GSV


2
Residue mass/area, mg/m

25 14
y = 2.9156e0.0552x
12
20 R2 = 0.8146
10

15 8
6
10 4
2
5 Oil A, 50 g/l Oil B, 50 g/l
Oil A, 1 g/l Oil B, 1 g/l 0
0 5 10 15 20 25 30 35
0
2
3 4 5 6 7 8 9 10 Residue mass/area, mg/m
Flush pH
Fig. 10. Correlation between the mass of asphaltenic residue per planar slide area,
from Fig. 8a, and the reduction in average grayscale value of the scanned images in
b 30 Fig. 9 due to the discoloration by these residues, for the six samples treated using
Oil A Oil B.
2
Residue mass/area, mg/m

25 Oil B

20

15 to allow water ingress to prematurely disconnect oil in the coat


from that in the bulk.
10 The results enable some assessment of the likely mechanisms for
improved oil recovery from kaolinite-rich sandstone reservoirs by
5 low salinity flooding. Along with the necessary conditions proposed
from core-scale experiments [22–26], the current study demon-
0 strates that crude oils in which their asphaltenes are relatively
0 10 20 30 40 50 60 70 80 90 100 stable (at reservoir temperatures and pressures) should gener-
SEM image edge loss, % ally be more conducive to wettability change during flooding. For
our model systems, in which the flood exposes kaolinite to a vast
Fig. 8. (a) Total mass of asphaltenic residues per planar area of kaolinite-coated excess of brine, the direct effect of its low salinity at fixed pH is to
substrate, for the two oils flushed with salt solutions of varying concentration and
favor the retention of oil by kaolinite, relative to the high salinity
pH, and post-rinsed with decalin–methanol, followed by extraction. (b) Correlation
of the two measures of oil residues in Figs. 5 and 8a. flood. In kaolinite-rich sandstone pores, the kaolinite/water ratio
is sufficiently high that cation exchange in response to low salin-
ity flooding can temporally perturb bulk cation concentrations and
handicap of limited scope to reverse the non-electrostatic precip- pH [22–26]. It appears that increased water-wetness of kaolinite
itation of asphaltenes in the initial state. The recent study [33] of at the molecular scale cannot occur without the aid of such associ-
Oil B residues on glass showed that removal could be extensive, ated pore-scale effects. The shift in flood pH imposed in our model
even in the presence of such precipitation, however the kaolinite systems to mimic such a spontaneous perturbation can favor water-
coats can support a complex network of precipitate with far greater wetness. For example, in Fig. 8a, the Oil A residue on kaolinite after
anchoring ability. The only condition for which significant removal the 1 g/l flood at pH 9 is slightly lower than for the 50 g/l flood at pH
occurs is for the 50 g/l flush at pH 4, which apparently weakens the 6. More substantial removal of residue, if occurring, must thus be
oil–kaolinite attraction without inducing such strong repulsion as due to pore-scale perturbation of divalent cation concentrations.

Fig. 9. Scanned images of the kaolinite-coated glass pieces (of average size 7 mm × 26 mm) exposed to Oil B and flushed with the six salt solutions, followed by
decalin–methanol post-rinsing.
22 A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23

a 5
Oil A, 50 g/l
b 40
Oil A, 1 g/l
30
Oil B, 50 g/l Oil B, 1 g/l

Zeta potential, mV

Zeta potential, mV
0
Kaolinite, 50 g/l 20 Kaolinite, 1 g/l

-5 10

0
-10 -10

-20
-15
-30

-20 -40
3 4 5 6 7 8 9 10 3 4 5 6 7 8 9 10
Brine pH Brine pH

Fig. 11. Average and standard deviation (error bars) of zeta potential of the crude oils and kaolinite in the salt solutions at their two concentrations of (a) 50 g/l and (b) 1 g/l,
each at three pH values.

4. Conclusions [10] O.C. Mullins, E.Y. Sheu, Structure and Dynamics of Asphaltenes, Springer, New
York, 1998.
[11] J.S. Buckley, Y. Liu, X. Xie, N.R. Morrow, Asphaltenes and crude oil wetting – the
Smooth, thin and robust coats of kaolinite on glass were used effect of oil composition, SPE J. 2 (1997) 107–119.
to analyze the propensity for crude oil to spontaneously displace [12] S.-Y. Yang, G.J. Hirasaki, S. Basu, R. Vaidya, Mechanisms for contact
an initially surrounding salt solution (as during formation of an oil angle hysteresis and advancing contact angles, J. Pet. Sci. Eng. 24 (1999)
63–73.
reservoir or subsurface DNAPL contamination), and the ability of [13] X. Xie, N.R. Morrow, J.S. Buckley, Contact angle hysteresis and the stability of
salt solutions of varying salinity and pH to subsequently displace wetting changes induced by adsorption from crude oil, J. Pet. Sci. Eng. 33 (2002)
the infiltrated, adhering oil (as during recovery or remediation). 147–159.
[14] E.M. Freer, T. Svitova, C.J. Radke, The role of interfacial rheology in reservoir
Both of the tested crude oils altered the wettability of kaolinite
mixed wettability, J. Pet. Sci. Eng. 39 (2003) 137–158.
by adsorbing/depositing their asphaltenics to penetrate all pores [15] J.S. Buckley, J.X. Wang, Crude oil and asphaltene characterization for prediction
between platelets. However, the oil with less stable asphaltenes of wetting alteration, J. Pet. Sci. Eng. 33 (2002) 195–202.
[16] J.S. Buckley, J.X. Wang, J.L. Creek, Solubility of the least soluble asphaltenes, in:
yielded a much thicker and more completely covering deposit due
O.C. Mullins, E.Y. Sheu, A. Hammami, A.G. Marshall (Eds.), Asphaltenes, Heavy
to surface precipitation. Accordingly, flushing of this oil with salt Oils and Petroleomics, Springer, New York, 2007, pp. 401–437.
solutions largely failed to remove it from the surface and pores [17] M. Kumar, A. Fogden, Patterned wettability of oil and water in porous media,
of the kaolinite coat. The oil with stable asphaltenes was more Langmuir 26 (2010) 4036–4047.
[18] M. Robin, E. Rosenberg, O. Fassi-Fihri, Wettability studies at the pore level – a
responsive to flushing, which yielded lesser residues and partially new approach by use of cryo-SEM, SPE Form. Eval. 10 (1995) 11–19.
reinstated the original water-wetness of the substrate. Its residues [19] C. Durand, E. Rosenberg, Fluid distribution in kaolinite- or illite-bearing cores:
also depended strongly on the salinity and pH of the flushing solu- cryo-SEM observations versus bulk measurements, J. Pet. Sci. Eng. 19 (1998)
65–72.
tion, displaying more complete removal at conditions conducive [20] M. Robin, Interfacial phenomena: reservoir wettability in oil recovery, Oil Gas
to weakening of the electrostatic local attraction between oil and Sci. Technol. 56 (2001) 55–62.
kaolinite interfaces, namely at higher salinity and/or by shifting the [21] S. Karimi-Lotfabad, M.R. Gray, Characterization of contaminated soils
using confocal laser scanning microscopy and cryogenic-scanning electron
initial intermediate pH to higher or lower levels. microscopy, Environ. Sci. Technol. 34 (2000) 3408–3414.
[22] G.Q. Tang, N.R. Morrow, Influence of brine composition and fines migration
on crude oil/brine/rock interactions and oil recovery, J. Pet. Sci. Eng. 24 (1999)
Acknowledgments 99–111.
[23] P.L. McGuire, J.R. Chatham, F.K. Paskvan, D.M. Sommer, F.H. Carini, Low salinity
This study was funded by an ARC Discovery Grant and the mem- oil recovery: an exciting new EOR opportunity for Alaska’s North Slope, in:
Proceedings of the SPE Western Regional Meeting, Irvine, USA, 2005, Paper SPE
ber companies of the Digital Core Consortium Wettability Satellite. 93903.
Warwick Hillier (RSB at ANU) is thanked for access to the FluoroMax [24] A. Lager, K.J. Webb, C.J.J. Black, M. Singleton, K.S. Sorbie, Low salin-
instrument. ity oil recovery – an experimental investigation, Petrophysics 49 (2008)
28–35.
[25] H.C. Wideroee, H. Rueslaatten, T. Boassen, C.M. Crescente, M. Raphaug, G.H.
References Soerland, H. Urkedal, Investigation of low salinity water flooding by NMR and
cryo-ESEM, in: Proceedings of the International Symposium of the Society of
Core Analysts, Halifax, Canada, 2010, Paper 26.
[1] N.R. Morrow, Wettability and its effect on oil recovery, J. Pet. Technol. 42 (1990) [26] A. RezaeiDoust, T. Puntervold, T. Austad, Chemical verification of the EOR mech-
1476–1484. anism by using low saline/smart water in sandstone, Energy Fuels 25 (2011)
[2] G.J. Hirasaki, Wettability: fundamentals and surface forces, SPE Form. Eval. 6 2151–2162.
(1991) 217–226. [27] D. Dudasova, S. Simon, P.V. Hemmingsen, J. Sjoblom, Study of asphaltenes
[3] J. Zheng, S.H. Behrens, M. Borkovec, S.E. Powers, Predicting the wettability adsorption onto different minerals and clays. Part 1. Experimental adsorption
of quartz surfaces exposed to dense nonaqueous phase liquids, Environ. Sci. with UV depletion detection, Colloids Surf. A 317 (2008) 1–9.
Technol. 35 (2001) 2207–2213. [28] E.V. Lebedeva, A. Fogden, Adhesion of oil to kaolinite in water, Environ. Sci.
[4] I. Chatzis, N.R. Morrow, H.T. Lim, Magnitude and detailed structure of residual Technol. 44 (2010) 9470–9475.
oil saturation, SPE J. 23 (1983) 311–326. [29] E.V. Lebedeva, A. Fogden, Wettability alteration of kaolinite exposed to crude
[5] N.R. Morrow, I. Chatzis, J.J. Taber, Entrapment and mobilization of residual oil oil in salt solutions, Colloids Surf. A 377 (2011) 115–122.
in bead packs, SPE Res. Eng. 3 (1988) 927–934. [30] N. Loahardjo, X. Xie, N.R. Morrow, Oil recovery by sequential waterflooding of
[6] D. Matmon, N.J. Hayden, Pore space analysis of NAPL distribution in sand–clay mixed-wet sandstone and limestone, Energy Fuels 24 (2010) 5073–5080.
media, Adv. Water Resour. 26 (2003) 773–785. [31] S. Wickramathilaka, N.R. Morrow, J. Howard, Effect of salinity on oil recovery
[7] R. Lenormand, C. Zarcone, A. Sarr, Mechanisms of the displacement of one fluid by spontaneous imbibition, in: Proceedings of the International Symposium of
by another in a network of capillary ducts, J. Fluid Mech. 135 (1983) 337–353. the Society of Core Analysts, Halifax, Canada, 2010, Paper 12.
[8] E. Bergslien, J. Fountain, The effect of changes in surface wettability on two- [32] H.G. Tie, Z.X. Tong, N.R. Morrow, The effect of different crude oil/brine/rock
phase saturated flow in horizontal replicas of single natural fractures, J. Contam. combinations on wettability through spontaneous imbibition, in: Proceedings
Hydrol. 88 (2006) 153–180. of the International Symposium of the Society of Core Analysts, Pau, France,
[9] P.-E. Øren, S. Bakke, Reconstruction of Berea sandstone and pore-scale mod- 2003, Paper 2.
elling of wettability effects, J. Pet. Sci. Eng. 39 (2003) 177–199.
A. Fogden / Colloids and Surfaces A: Physicochem. Eng. Aspects 402 (2012) 13–23 23

[33] E.V. Lebedeva, A. Fogden, Nano-scale structure of crude oil deposits on water- [35] D. Borden, R.F. Giese, Baseline studies of The Clay Minerals Society source clays:
wet substrates: dependence on aqueous phase and organic solvents, Colloids cation exchange capacity measurements by the ammonia-electrode method,
Surf. A 380 (2011) 280–291. Clays Clay Miner. 49 (2001) 444–445.
[34] S.J. Chipera, D.L. Bish, Baseline studies of The Clay Minerals Society source clays: [36] R.J. Pruett, H.L. Webb, Sampling and analysis of KGa-1b well-crystallized kaolin
powder X-ray diffraction analyses, Clays Clay Miner. 49 (2001) 398–409. source clay, Clays Clay Miner. 41 (1993) 514–519.

Вам также может понравиться