Você está na página 1de 18

Colloids and Surfaces A: Physicochem. Eng.

Aspects 520 (2017) 622639

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


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

Experimental and theoretical study of wettability alteration during


low salinity water ooding-an state of the art review
Hongna Ding, Sheik Rahman
School of Petroleum Engineering, University of New South Wales, UNSW, Sydney, NSW, 2052 Australia

g r a p h i c a l a b s t r a c t

Initially oil-wet carbonate reservoirs as the carbonate rock is positively charged and crude oil is negatively charged at reservoir conditions (pH 7-8,
T 100 C), therefore, the oil adheres strongly onto the carbonate rock that makes it oil-wet (left). Low saline with potential determining ions (pdi) injected
into carbonate reservoirs will increase the negative surface charge/potential due to chemisorption, therefore, enhance the repulsive part of disjoining
pressure and changes the wettability to water-wet. At equilibrium, a stable thin lm will be formed by double layer expansion mechanism (right).

h i g h l i g h t s

Wettability alteration is considered as the main reason for low salinity effect in sandstone and carbonate reservoirs.
The microscopic phenomenon, double layer expansion (DLE), is contributed to the wettability alteration caused by low saline.
Intermolecular interactions in an oil/brine/rock system are quantied collectively by DLVO and non-DLVO theory.
Interactions between crude oil and raw rock sample immersed in low saline are determined by surface force apparatus, especially by AFM.
Chemically force microscopy and force volume technique are able to reveal the intrinsic nature of wettability alteration.

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

Article history: Low salinity waterooding (LSW) has been proved to be a promising technique in improving oil recovery
Received 10 October 2016 for both sandstone and carbonate reservoirs. Sandstone and carbonate are two completely differ-
Received in revised form 24 January 2017 ent minerals but the mechanism of LSW can both be contributed to wettability alteration. Though
Accepted 3 February 2017
many reasons regarding wettability alteration have been proposed, central to the entire subject is the
Available online 6 February 2017

Corresponding author.
E-mail addresses: sheik.rahman@unsw.edu.au, sghasemi@umz.ac.ir (S. Rahman).

http://dx.doi.org/10.1016/j.colsurfa.2017.02.006
0927-7757/Crown Copyright 2017 Published by Elsevier B.V. All rights reserved.
H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639 623

Keywords: double layer expansion (DLE), which can be quantitatively explained by disjoining pressure. Disjoining
Low salinity water ooding pressure is the sum of van der Waals force, electrical double layer force and structural force, which is
Wettability alteration not a new concept, however, it attracted a lot of attention with the growing popularity of LSW. The
Double layer expansion three inter molecular forces are explained by DLVO (Derjaguin-Landau-Verwey-Overbeek) theory and
Disjoining pressure
in an oil/brine/rock system these forces can be accurately measured by atomic force microscopy (AFM).
Atomic force microscopy
The extracted information from force vs distance curves and topographical images are used equally to
evaluate the theoretical model. In petroleum industry, the two theories are linked together to explain the
wettability alteration induced by DLE as a result of an alteration from high salinity water to low salinity
water in sandstone and carbonate reservoirs.
Crown Copyright 2017 Published by Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 623
2. Theoretical background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624
2.1. DLVO forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624
2.1.1. Van der Waals Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624
2.1.2. Electrical double layer force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 626
2.1.3. Modied poisson-Boltzmann . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 627
2.2. Non-DLVO forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 628
2.2.1. Structural force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 628
2.2.2. Hydration force. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .628
2.2.3. Hydrophobic force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 629
2.2.4. Steric force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 629
2.2.5. Depletion force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 629
2.2.6. Born repulsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 629
3. Direct force measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 629
3.1. Interactions of pH-dependent sandstones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630
3.1.1. VDW force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630
3.1.2. EDL force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630
3.1.3. Hydration force. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .631
3.1.4. Structural force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 632
3.1.5. Hydrophobic force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633
3.2. Interactions of deformable oil surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633
3.2.1. Rigid tip-oil droplet and drop-drop interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633
3.2.2. Chemical force microscopy (CFM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634
3.3. Interactions of chemical active carbonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634
4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 636

1. Introduction salinities (30,00050,000 ppm) [139]. LSW actually is an ambigu-


ous concept and the salinity of the injected water is not restricted to
The wettability of an oil reservoir can be categorized into three some specic values. As suggested by some researchers, chemistry-
types: oil-wet, mix-wet and water-wet. It is believed that most optimized water or smart water seems make more sense as it
sandstone reservoirs are mix-wet after some extent of oil migra- emphasizes not only the salinity but also the adjusted composition
tion [1]. About 80% carbonate reservoirs, on the other hand, around of the injected water. Low salinity water with less multivalent ions
the world are oil-wet amongst which 15% is strongly oil-wet [43]. is more protable for sandstone reservoirs, while seawater with
It is estimated that the average oil recovery factor worldwide is more potential determining ions (Ca2+ , Mg2+ and SO4 2 ) is more
only 2040% and more than half is still left underground [149]. Low benecial for carbonate reservoirs.
salinity waterooding (LSW) is, therefore, applied as a low cost sec- The mechanism behind LSW ooding for improving oil recovery
ondary or tertiary enhanced oil recovery (EOR) method to extract is proven to alter wettability of reservoir rocks [17,190]. A number
extra oil from sandstone or carbonate reservoirs. Compared to other of explanations for change of wettability of an oil reservoir from
EOR methods (e.g. chemical ooding, thermal ooding, CO2 ood- oil-wet to mix-wet (carbonates) or from water-wet to mix-wet
ing etc.), LSW is more economically applicable and environmentally (sandstones) are given by numerous authors: pH effect [144], multi-
friendly. component ion exchange (MIE) [134], double layer expansion (DLE)
Despite its success in laboratory core displacement tests and [138], salt in and salt out effect [185] and the chemical mechanisms
some limited scaled eld trials in both sandstone and carbonate proposed by Austad et al. [12]. These mechanisms are primarily pro-
reservoirs, there are two major differences between these two posed based on the low salinity effect observed on sandstones and
kinds of reservoirs. One distinct difference is the salinity of the clays are necessary to observe the low salinity effect. It is not the
injected water. It is concluded that the optimal salinity for LSW case for carbonates but these mechanisms can be extended to car-
in sandstone reservoirs is less than 4,000 ppm [135], but the salin- bonates provided, calcite dissolution takes place in order to trigger
ity reported is sometimes as high as 5,000 to 7,000 ppm. Seawater all the above processes [216] and DLE is the leading mechanism
with a salinity of more than 10, 000-20, 000 ppm is commonly used responsible for wettability alteration both for sandstones Ligthelm
for LSW in carbonate reservoirs, but many reported even higher et al. [138] and carbonates [3].
624 H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639

where, Pc is capillary pressure (Pa) between wetting (water) and


non-wetting phase (oil), (h) is disjoining pressure (Pa),  ow is
interfacial tension (N/m),  is contact angle ( ), r is mean radius of
pores [119]. In most cases, the capillary pressure is larger than the
disjoining pressure in microscope pores, but if the rock surface is
at the capillary pressure equals to disjoining pressure.
Disjoining pressure ((h)) is regarded as the force tends to
separate two phases (oil and water), results from molecular and
interionic interactions between the rock, oil and brine phases [103].
Therefore, three different components of disjoining pressure are
generally considered: electrical, molecular and structural forces
[61]. These contributions are broken into three forces in unit surface
area:

Fig. 1. A conceptual model of a thin water lm formed between oil/brine and


brine/solid interface. (h) = VDW (h) + EDL (h) + STR (h) (2)

DLE is closely correlated with the electrochemical interactions where, (h) is disjoining pressure (Pa) as a function of wetting lm
among mineral surfaces, oil and injected brine. These interactions, thickness h (nm), VDW is the London van der Waals forces (Pa),
namely van der Waals force (vdW), electrical double layer force EDL is the electrostatic forces in the electrical double layer (Pa) and
(edl) and structural force, were introduced by Hirasaki [103]. This STR is the structural force (Pa). The rst two forces are introduced
electrochemical interaction model is further illustrated by study- as DLVO theory [63,64,207], the last one is usually introduced as a
ing the wetting lm stability and contact angle variation as direct part of non-DLVO theory.
indicators of wettability alteration during LSW ooding [119,212]. For oil to be detached from carbonate minerals, low salinity
It is suggested that the wettability of mineral surfaces is highly water should overcome the capillary pressure to get into the throats
dependent on the micro-topography and physical and chemical and disjoining pressure acts to increase the water lm thickness.
properties of minerals, crude oil quality (acidic and basic groups), A schematic representation of DLE which corresponds to disjoin-
brine composition (ionic strength, pH), and ambient conditions ing pressure of oil/brine/rock system is shown in Fig. 2. The lm
(e.g. temperature, pressure and dissolved CO2 ) [199]. These factors thickness of wetting phase (water in this case) is a nal result of
were vastly investigated by laboratory core imbibition/ooding the competition of repulsive and attractive forces. The repulsion
experiments, contact angle and interfacial tension analysis, other- between the charges/ions in brine/oil and brine/rock interfaces pre-
wise by zeta potential measurements [5,152,153,212,139], directly vents the oil from reaching the rock surface and wetting it. In oil
intermolecular forces measurements, and wetting lm thickness wetting reservoirs, however, it is desired to change the state of
measurements [135,148]. wettability from oil-wet to water-wet and to further increase the
In this paper, we will review the theoretical and experimen- thickness of the newly formed water lm as much as possible. This
tal studies corresponding to wettability alteration during LSW helps to increase the mobility of the oil and eventually increase the
ooding process. Accordingly the paper is organized in the fol- amount of recoverable oil.
lowing orders: rstly, it introduces the theoretical background of On the basis of augmented Young-Laplace equation (Eq. (1)), a
wettability alteration induced by DLE. An electrochemical model quantitative relationship of wettability and disjoining pressure can
based on augmented Young-Laplace equation is established and be established [16]. Contact angle is therefore expressed as,
is further used to explain how disjoining pressure controls the
 Pc
stability of wetting lm and therefore the rock wettability. Then 1
a detailed description of disjoining pressure is given in order to cos  = 1 + I Where I= hd (3)
ow 0
fulll the framework of the electrochemical model; secondly, it
overviews the measurement of direct intermolecular forces as an
evaluation of the electrochemical model. We focus on the force a. the water lm is unconditionally stable (Fig. 3a) if I > 0 ( = 0 );
curves and topography images that are obtained from surface force b. the water lm is meta-stable (Fig. 3b) if  ow < I <0, (0 < < 90 ,
apparatus because the experimental methods like core imbibi- water-wet); or if 2 ow < I <- ow (90 < < 180 , oil-wet);
tion/ooding, contact angle/interfacial tension analysis and zeta c. the water lm is unstable (Fig. 3c) if I <-2 ow .
potential measurements are so well developed and broadly used
in petroleum engineering. This part is further divided into three
categoriessandstone, clay and carbonate; at last, it draws a brief A stable water lm on silica/clay surface is proved to be with
conclusion based on the historical theoretical and experimental a thickness of 10 nm from the small angle neutron scattering
studies. Moreover it discusses our further research interests regard- performed by Lee et al. [135]. It is suggested that the water lm
ing to wettability alteration of LSW ooding in carbonate reservoirs. thickness of silica/clay-like particles is thinner at lower salt con-
centrations (e.g., 0.001 M) compared with higher concentrations
2. Theoretical background (e.g., 0.1 M). Experimental data presented also supported divalent
ions have a smaller water layer thickness than monovalent ions.
The residual water on mineral surfaces appears in two kinds of The thickness is mineral different: the water lm thickness for clay
congurations: as pendular rings at grain-to-grain contact points like particles is 12 nm larger than those for silica like particles. Fur-
and as a roughly 10 nm thick lm which cover the mineral surface thermore, the variations of lm thickness with salinity for clay like
[197,148] (Fig. 1). In the oil/brine/rock system, the equilibrium of particles are more signicant than silica like particles, especially in
the interactions between oil, brine and rock can be expressed by the divalent ion cases. Similarly, it is again shown by imaging ellip-
the augmented Young-Laplace equation, sometry technique that the lm thickness at mica surface in 1 M
NaCl solution is about 10 nm but it decreases from 8 nm to <1 nm
Pc = (h) + 2ow cos /r (1) with increase in CaCl2 concentration [148].
H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639 625

Fig. 2. (a) Mix-wet state after high salinity (sea water or formation water) ooding, (b) Low salinity brine increases the repulsive disjoining pressure and changes the
wettability to water-wet, (c) At equilibrium, a stable thin lm will formed by DLE (the electrical potential at oil/brine interface and brine/rock interface is expressed as 1
and 2 , respectively), indicating a fully water-wet state [Reconstructed from Myint and Firoozabadi [15]].

Fig. 3. Disjoining pressure prole and its correspondence to wetting lm stability: (a) unconditionally stable, (b) meta-stable, (c) unstable [Reconstructed from Basu and
Sharma [15]].

Fig. 4. Schematic diagram of (a) permanent dipole and permanent dipole interactions, (b) permanent dipole and induced dipole interactions, (c) temporary dipole and
temporary dipole interactions [from youtube: https://www.youtube.com/watch?v=cERb1d6J4-M].

2.1. DLVO forces which have permanent multiple moments. Debye force arises
from induced polarity by a polar molecule in a nearby neutral
2.1.1. Van der Waals Force molecule and results in an attractive induction force between them.
Van der Waals (vdW) force between any two atoms or molecules Non-polar molecules have nite uctuating dipoles and higher
has been divided into Keesom force, Debye force and London dis- multipole moments at very short intervals, these instantaneous
persion force; also known as orientation force, induction force moments induce polarity in neighboring atoms or molecules to give
and dispersion force, respectively [112]. Keesom force arises only rise to an attractive force called London dispersion force [112] (see
in the case of polar molecules (static dipoles, quadrupoles, etc.) Fig. 4).
626 H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639

Hamaker [94] rst gave the exact calculation of vdW force total vdW energy. Previous measurements of vdW force between
between spherical particles by integrating the point-by-point inter- glass surfaces have conrmed the existence of retardation. vdW
actions. Then the vdW force for two macro-bodies with other force between mica surfaces in air also has shown retardation effect
geometries can be calculated based on the pairwise additivity above 50 nm [112]. Even at smaller distances (h 510 nm) retar-
assumption. One should note that this assumption is met with- dation effect should be considered [116,181] because of a strong
out intervening or interlayer (such as in vacuum) between two absorption of electromagnetic waves in water.
macro-bodies and it may break down otherwise [115]. With the Retarded Hamaker constant becomes h2 -dependent for two
popularity of the quantum electrodynamics, Lifshitz [137] pro- planes and is given by [88] as:,
posed the famous Lifshitzs theory for calculating the vdW force  
Aretardation = A=0 + A>0 / 1 + 5.32h/
lw (7)
between two bodies separated by vacuum. In his theory, vdW is
directly linked to the dielectric constants and optical properties where,
lw is London wavelength (nm) and it is assumed to be
(such as reective index and absorption frequency) of the interlayer 100 nm. The power law n of vdW force increases with separation
and macro-body materials. A generalized vdW force for multi- which indicating the vdW force decreases much lower at long dis-
ple interlayers has been proposed by Parsegian and Ninham [168] tance.
which facilitates the usage of vdW force in colloidal and biochemi- The ionic screening effect usually signicantly reduces the
cal science. An exact calculation of the Hamaker constant based on strength of electrostatic force compared with vdW force and also
Lifshitzs theory is given by Hough and White [109]. However, some there is a limit for ionic strength from where vdW force stops
approximation equations only render slight differences; no more decreasing. However, vdW force becomes less attractive due to the
than 5% compared with the exact result [115]. For a comprehen- combination of retardation effect and ionic screening. It is often the
sive understanding of the history, developments and applications case that in real reservoir condition where high salinity formation
in various areas of vdW force one can reference to Frenchs work water or sea water exists. As a result, the electrical double layer
[80]. force becomes the dominant component of disjoining pressure
In the oil/brine/rock system, vdW force is always present as part which inuences the stable water lm thickness on oil/brine/rock
of disjoining pressure. The vdW force for two at planes can be system.
calculated by,
2.1.2. Electrical double layer force
A
VDW (h) = (4) The electrical double layer (edl) force is derived from the
6h3
Coulombic interactions between charged ions or a charged ion
where, A is Hamaker constant [94] for the oil/brine/rock system and a polar particle. Counterions from the solution balance the
(J). An approximated non-retarded Hamaker constant based on Lif- excess surface charge of the solid particle forming an immobile
shitzs theory is given by Overbeek and Sparnaay [166], layer referred to as the Stern layer. The counterions in the following
3
   layer become more diffuse due to less binding strength. The com-
1 3 2 3
Anon = A=0 + A>0 = kB T bination of Stern layer and diffuse layer is the so-called electrical
4 1 + 3 2 + 3
double layer. The plane from where the counterions concentra-
(n21 n23 )(n22 n23 ) (5) tion begins to obey Poisson-Boltzmann equation is called outer
3pe
+   Helmholtz plane (OHP) or slipping plane. The potential on slipping
8 2 (n2 + n2 ) (n2 + n2 )1/2 (n2 + n2 )1/2 + (n2 + n2 )1/2
1/2
1 3 2 3 1 3 2 3 plane is referred to as the zeta potential (). Formation of edl can
be derived from preferential adsorption of potential-determining
where, A=0 is the zero-frequency energy and includes the Kee- ions and interior crystal imperfections [59]. Adsorption can occur
som and Debye dipolar contribution, A > 0 is the dispersion energy through chemical bonds or physical adsorption (e.g., hydrogen
and includes the London energy contribution. kB is the Boltzmanns bonds and vdW forces). The preferentially adsorbed ions will form
constant (1.38 1023 J K1 ), T is absolute temperature (K), P is the the inner layer and the counter ions will be accumulated on top to
Planks constant (6.63 1034 J s),  e is electronic adsorption or ion- form a second layer. Upon changing the salinity, the surface poten-
ization frequency in the UV range (3.0 1015 s1 ). 1 , 2 and 3 are tial will remain constant whilst the surface charges change. The
the dielectric constant of oil, rock and brine, respectively and n1 , n2 second type of double layer forms due to imperfections in lattice
and n3 are the refractive index of oil, rock and brine, respectively. layers, especially for clay crystals. Upon injecting low salinity water,
The zero-frequency energy, A = 0 only dominates in small dis- though the surface charge cannot be changed, the surface potential
tances (below 2 nm), while at long distance, the dispersion energy will become smaller. This type of edl is regarded as being of more
would take over the zero-frequency energy part and control the importance in low salinity waterooding.
vdW force. A = 0 is related to the temperature and therefore kB T, The edl force can be either repulsive or attractive depending on
dielectric constant and refractive index of the oil/brine varies as chemical structures, suspending medium properties and surface
temperature changes, indicating that the vdW force is affected by charges. The sign of surface charge is determined by isoelectric
the thermal conditions of the reservoir. However, the temperature point (iep), which is dened as the pH at which no net electri-
sensitivity of Hamaker constant is negligible as it only produces cal charge exists on the particle surface. Sandstone is negatively
3.5% difference results when temperature fries from 20 C to charged above pH 23 [76,118]. Crude oil is positively charged at
200 C for ceramic systems [80]. Its fact that A=0 undergoes ionic pH < 4 [29,30]. Carbonate rock (iep 911) is positively charged in
screening at high ionic strength conditions because of its low fre- reservoir condition [211,5,118].
quencies (Mahanty and Ninham [140]. The screening effect at a The mathematical procedures to gain insight into the electro-
limit where h  1 could be given by [23], static forces are derived from PoissonBoltzmann (PB) distribution
Ascreen = A=0 (1 + 2 h) exp (2 h) + A>0 (6) [115] which is used to describe the electrical potential in the direc-
tion normal to a surface where electrolytes are present,
where, is the inverse of the Debye length (nm). Eq. (6) suggests
d2 enb ze
that A=0 decays exponentially in electrolytes and with a decay = e kB T (8)
length equals to Debye length. dx2 z0 r
Butkus and Grasso [31] had experimentally proven the ionic where, 0 is the dielectric constant of vacuum, r is the relative
screening A = 0, but it is not the only contributor to decreasing the dielectric constant of the aqueous medium, kB is the Boltzmanns
H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639 627

constant (1.38 1023 J K1 ), T is absolute temperature (K), e is < 2) and moderately ranges of ionic strength. However, they rep-
the electron charge (1.60 1019 Coulomb), z is ion valence. A resent two extremes with the correct answer lying in between
simple algorithm has been developed by Chan et al. [41] to solve [87]. A linear superposition approximation (LSA) or weak overlap
this non-linear differential equation for identical, parallel planes in approximation between the constant charge and constant potential
symmetrical/unsymmetrical valent electrolytes at various bound- assumptions is given as follows,
ary conditions,    
The surface potential for an isolated surface at a certain distance 1 2
EDL (h) = 64nb kB T tanh tanh exp ( h) (18)
is given by Gouy-Chapman equation, 4 4

2kB T

1 +  exp ( x)

4kB T where, i is zeta potential of each surface (V). The prerequisite for
(x) = ln exp ( x) (9)
ez 1  exp ( x) ez this equation is the lm thickness is relatively large than the char-
acteristic length ( h > 1) far from where the surface potentials for
The parameter is given by,
each interface become additive. The LSA is favored for tting the

e 0 surface force measurement data [115].
 = tanh (10)
4kB T
2.1.3. Modied poisson-Boltzmann
where 0 is surface potential (V). In general the surface potential is
In the classical DLVO theory, ions are regarded as point charges,
experimentally immeasurable, so the zeta potential is often used
the mean electrostatic potential is proportional to the potential
as an estimate [206].
of mean force (PMF) and the medium is assumed as a dielectric
 z

continuum up to the OHP characterized by its dielectric constant
0 = 1+ e z (11)
[136]. Comparisons between the theoretical predictions and exper-
where z is the distance between the surface of the charged particle imental results reveal some inconsistencies which led to attempts
and the slipping plane (z 0.5 nm), is the Stokes radius of the to modify the PB equation and they are summed up as Modied
particles. PoissonBoltzmann (MPB) equations. The modications have been
It is important to note that the zeta potential is a highly sen- investigated to include the following effects:
sitive property that varies with pH, temperature and the molar i. Excluded-volume effect [136,165,22] due to the nite size of
concentration of the solution. Extensive laboratory measurements ions. In Stern-Gouy-Chapman (SGC) model [52], non-specically
are needed to measure the zeta potential for both the oil/brine and adsorbed counterions carry full solvation shells bound to the sur-
brine/solid interfaces. face that form the OHP and specically adsorbed counterions
The inverse of the Debye length is given by, with residual solvation shells chemically bound with charged sur-
face, and their centers form the inner Helmholtz plane (IHP), which
e2 nb makes the plane a few angstroms further from the physical solid-
= (12) liquid interface (vdW plane). The OHP at the mica-water interface
r 0 kB T
is 2.5 nm farther out from each surface [116]. A shift of 0.9 nm
where, nb is the charge density and given by, has been assumed by Pashley and Israelachvili [174] to describe
the high apparent surface potentials on mica surfaces. Shubin and
nb = 2INA (13)
Kkicheff [191] considered 0.75 nm which equals to two hydrated
where NA is conversion coefcient (6.022 1023 m3 ), I is ionic molecules is a more reasonable value, but Kkicheff et al. [124] have
strength given by, chosen 0.54 nm (one hydrated Ca2+ diameter) instead for mica sur-
1 n
faces which leads to a lowering of 10% apparent surface potential.
I= ci zi2 (14) ii. Self-atmosphere-image effect. The mean potential of a given
2 i=1
ion is lowered by the ionic atmosphere of a local opposite ion,
where ci is ionic concentration in the bulk solution (mol L1 ), zi which gives rise to self-atmosphere potential. Image effect is
is ion valence. more signicant at higher salinity [161]. Levine [136] had modeled
Modeling electrostatic double layer force requires proper the self-atmosphere as an asymmetrical sphere in bulk solu-
boundary conditions for the Poisson-Boltzmann equation either tion and discontinuities in the dielectric constant at the interface.
using a constant potential or constant charge assumption. The con- Simulation results suggested that dielectric constant varies little
stant potential condition assumes that there is always attraction therefore dielectric continuum assumption was acceptable with-
between plates with unequal potentials even if they have the same out render seriously error. However, the potential at the middle
charge sign. The contribution inside the electrical double layer plane decreases signicantly and produces a change of 50% in the
under constant potential is given by [87], possibly the upper limit, electrostatic force.
2 2
iii. Variations of dielectric constant in the diffuse layer. A dielec-
2 r1 r2 cosh ( h) r1
r2
EDL (h) = nb kB T (15) tric sandwich the IHP and OHP which separate the target domain
2
(sinh ( h)) into three layers is common practice. Another one is introducing a
Conversely, the constant charge condition assumes a repulsion variable dielectric constant regarding electric led or ionic concen-
force between unequal plates and is given [87], possibly the lower tration into the PB equation [85,194,92,93].
limit, by, iv. Static or dynamic surface roughness. DLVO theory assumes
2 2
perfectly smooth and geometrically regular surfaces, however for
2 r1 r2 cosh ( h) + r1
+ r2 natural particles or surfaces, discrepancies frequently shown when
EDL (h) = nb kB T (16)
(sinh ( h))
2 comparing the interactions predicted by theory and measured by
force apparatus. These discrepancies are largely attributed to het-
where, ri is reduced potential and is given by, erogeneity either in the form of static surface roughness which
is simply modeled as hemispherical asperities [19,209], or non-
ri = ze i /kB T (17)
uniform charge density [67,68]. Dynamic surface roughness is
The above analytical compression approximation (CA) model another contributor to the failure of DLVO theory. A great exam-
is suitable for low to intermediate electrostatic potential (ze /kB T ple is that the negative charges possessed by silica are located at
628 H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639

the ends of silica hairs and they are dynamically mobile and their tration, break the linear connection between surface charge and
values vary with pH or ionic strength [208]. surface electrical eld, bend the differential capacity curves of the
v. Charge regulation is ascribed to the cause of intermolecular double layer and separate the adsorptions of ion species of monova-
interactions intermediate between constant charge and constant lent electrolyte. Another attempt was by incorporating a negative
potential [158]. Thence edl force is modied to include charge adsorption energy of counterions and an increased size of hard-
regulation model in order to describe the association/dissociation sphere ion (including hydration shell 2.7 ) with the anisotropic
of surface charges. Mass action (site binding) model [169], which HNC model to quantitatively evaluate the hydration energy [124].
assumes competitive binding of hydrogen ions and metal ions with viii. Ion penetration. Another issue regarding edl force is
dissociated, negative surface sites, is commonly used to character- the penetration of ions into the surface, especially for bilayers.
ize the regulated surface. It is capable of describing the variations in The structural charges smear perpendicular into the membrane
mica/glass surface potential with an enlarged, hydrated counterion plane and cause an average transverse displacement of the struc-
size without the existence of a Stern layer [169]. The mass action tural charges, which reduces the electrostatic potentials near the
model gives a good agreement with experimental data [39,84], membrane-solution interface. In this area (a zone of relatively small
however, Ducker et al. [71] has discussed that the surface potentials electric eld strength), the edl force will be canceled out by vdW
could not be perfectly predicted by the mass action model and the or hydrophobic interactions [37]. The reduction in edl force will
potentials obtained by tting the experimental data are not con- increase at much higher salinity. Penetration effect has been exten-
rmed by zeta potential measurements. Also, some differences in sively considered by Ohshima and coworkers. Ohshima and Kondo
force behavior happen by using different charge regulation model. [162,163] assumed that the penetrated charges are distributed uni-
For example, hydration forces were not observed at low electrolyte formly within the membrane but the potential remain constant
concentrations (0.0001 M to 0.001 M) at small distances ( < 1.0 nm) (equal to Donnan potential) independent of membrane separations.
by Basu and Sharma [15], but Pashley [170] have reported compara- The edl force at Donnan potential lies between constant charge and
ble hydration forces with mass action model. The charge regulation constant potential; similar phenomenon exists at linear approx-
model used by Basu and Sharma [15] is Yatess model [213] which imation model. Ohshima and Kondo [163] further compared the
takes into account the interfacial ion pairs formed with neutral ion-penetrable model to LSA model and it turned out that penetra-
surface site by potential determining ions (pdi) and coions. But tion effect gives a comparable repulsive edl force to that LSA model
H+ is not considered as pdi and no amphoterism. Charge regula- at low surface potentials.
tion is more important for asymmetric surfaces [81]. Two surfaces
each carrying either positive charges or negative charges main- 2.2. Non-DLVO forces
tain constant charge, for which the total intermolecular force is
always repulsive. If the two surfaces maintain constant potential, At very small distance, DLVO theory often fails to describe some
the repulsion turns into attraction on approach to each other as a experimental phenomena exist in colloid or polymer stability mea-
result of charge reversal at the surface with the lower potential. The sured by surface force microscopy. That is where non-DLVO theory
more extreme the ratio between the two surface charge densities comes into play (see Israelachvili [115] and the references therein).
at innite separation, the larger the separation and at this innite The forces which belong to the category of non-DLVO theory gener-
separation the interaction reversal occurs [81]. ally consist of structural force, hydration force, hydrophobic force,
vi. Ion-ion correlation effect due to the connement of ions steric interaction, depletion interaction and born repulsion.
when two particles approach one another [132]. Ion-ion correla-
tion force is usually attractive, short-range ( < 4 nm) [132] and is 2.2.1. Structural force
characterized with exponentially decaying with separations. It is Structural force arises for two reasons: (i) a thin lm formed
far more larger than vdW force [66]. Within the scope of prim- between smooth surfaces due to molecular structuring and (ii)
itive model [128,129,127,130,131], the ionion correlation force macromolecules like polymers exist in liquid lm owing to volume
becomes stronger as the ionic strength increases. It is assumed that restriction and exclusion [60]. The rst gives rise to solvation force
ionion correlation at low to medium concentrations is a result related to polar solvents, and if the solvent is water it is referred to as
of correlations between counterions but coions at concentrated hydration force [217]. Surfactants, on the other hand, can generate
concentrations [124]. More specically, at concentrated ( > 0.24 M) hydrophobic force by bonding on a surface. However, the aflia-
Ca(NO3 )2 solutions a triple layer forms on mica surface: a Stern tion is not rigorous in aqueous solutions and Horn and Smith [107]
layer consists of excess counterions near each surface, a middle distinguished the structural force, hydration force and hydropho-
layer compacted with coions, and the bulk solution with discrete bic force by their features. Hydration force is usually monotonic
coions. Increasing the ionic strength, the coions are expelled by repulsion and decays exponentially at small distances. Structural
counterions to escape from the gap and results in a strong ionion force is characterized as oscillated and with a periodicity equals to
correlation attraction. At even higher concentrations the interac- the diameter of one hydrated molecule. It arises from the packing
tions turn into oscillatory thus causing each successive layer to be limitation of nite sized liquid molecules which is predominantly
expelled from the gap [124]. an entropic effect. Hydrophobic force is discovered on hydrophobic
vii. Repulsive hydration energy. Intermolecular force measure- surfaces which are modied by adsorption of surfactants. This dis-
ments have proven that there exist repulsive hydration forces, tinction is adopted in the present paper. The second case normally
which brings about an additional energy need to be considered brings about steric force and depletion force, separately.
into traditional DLVO theory [92,93,89,90,47]. It was argued that
hydration force and edl force cannot be considered separately 2.2.2. Hydration force
[159,157] as hydration force is intrinsically another kind of force Hydration force appears ascribed to different mechanisms [86].
arises from Coulombic forces. Gur et al. [92,93] have incorporated More widely recognized mechanisms include physical adsorption
the hydration free energy term for individual ion into the Boltz- of hydrated cations which induced by ion exchange progress with
mann distribution of ions in the double layer. Basu and Sharma hydrogen ions [169] and formation/breakage of hydrogen bonding
[15] extended Gur et al.s model into two overlapping double lay- network [42], which links to secondary hydration force and primary
ers. The ions in the double layer are subjected to both electrostatic hydration force, separately. Primary hydration force is related to the
potential and their own hydrated potential. It was turned out enthalpic energy of adsorbed water layers, thus its decay length is
that the hydration free energy will lower the surface ion concen- associated with the thickness of water layers (0.20.4 nm). Sec-
H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639 629

ondary hydration force comes from the hydrated energy of cations 2.2.5. Depletion force
adsorbed at the interface. The decay length of weaker secondary Depletion force occurs as a result of volume exclusion effect;
hydration is usually equaled to the Debye length [167]. when two large particles approach each other, the smaller
Many mathematical models have been developed for character- molecules will be excluded from the gap between them. It leads
izing the hydration forces [86]. An empirical exponential hydration to a reduction of osmotic pressure between the particles and gives
force is proposed by Molina-Bolivar and Ortega-Vinuesa [147] for rise to an attractive depletion interaction force [86].
hydrophilic systems ( a < 15 , Derjaguin and Churaev [62],

STR (h) = C1 eh/ 1 + C2 eh/ 2 (19) 2.2.6. Born repulsion


Born repulsion is a very short-range molecular interaction that
where, CH is force coefcient, hs is decay length (0.61.1 nm) and originated from the overlap of electron clouds. Many studies have
an estimate hs = 1.0 nm was attained for mica in monovalent taken born repulsion into disjoining pressure to explicate the strong
and divalent electrolytes from 104 to 102 M [114,169,170] and repulsive at short range [125,126,187,123,119]. The born repulsion,
hs = 0.85 nm for glass laments in KCl solutions [181].  a refers to BR is given by [126],
advancing contact angle.
 
A6 8rp + h 6rp h
2.2.3. Hydrophobic force
BR (h) =
7560  7 +
h7
(21)
2rp + h
Although extensive laboratory experiments have shown that
LSW has great potential in improving oil recovery, petroleum engi-
Where, is collision diameter 0.5 nm according to Khilar and
neers are not satised with the incremental recovery from purely
Fogler [125], rp is the radius of sphere particle.
LSW. Surfactant assisted LSW or low salinity surfactant (LSS) has
been performed to further improve the oil recovery [160,188,108].
Adsorption of surfactants modies the rock and oil surface from
3. Direct force measurement
hydrophilic to hydrophobic (or vice versa) and changes the orien-
tation and conguration of water molecules at the interface which
Surface force apparatus (SFA) and atomic force microscope
generates an attractive hydrophobic force [106]. Marcelja and Radic
(AFM) have long been used to measure intermolecular forces
[141] associated the hydrophobic force with a larger density of
[115]. Compared with SFA, which only applicable with molecu-
water molecules next to the surfaces than in the bulk solution and
larly smooth and cylindrical shape surfaces [70], AFM seems to be
expressed it as a double exponential function,
a more powerful tool to quantify atomic forces [84] or to achieve
  true atomic-resolution images of surface topographies [82]. Lots of
6
A 8rp + h 6rp + h materials (mica, clay, silica, glass, sapphire, gold, oil droplets and air
BR (h) =
7560  7 +
h7
(20)
2rp + h bubbles) and geometries (sphere, planar or pyramid) can be han-
dled by AFM technique. Despite the differences in geometry and
where, C1 , C2 are force coefcients determined by boundary con- size of samples, the measured forces by SFA and AFM are all trans-
ditions (Pa), 1 , 2 are decay lengths (nm), 1 is in the range of formed to surface energy between two planar surfaces through
0.30.5 nm and 2 = 3.0 nm for rutile or mica (Yotsumoto and Yoon, Derjaguin approximation; in other words, the forces are compara-
1993). ble for these two methods [71,124]. But Derjaguin approximation
Eq. 19 and Eq. 20 are phenomenological model and hence fails at small separations when very high load is exerted on the
experimental hydration force or hydrophobic force exhibits indis- surface due to elastic deformation in SFA which leads to a progres-
tinguishable behaviors; for example, the hydration force between sive attening and enlarging of the tip curvature. The deformation
silica sheets can be tted by single exponential functions [71], effect is much less important for AFM [84].
double exponential function [170], power law function [84], or Data that can be generated from force-distance curves by SFA or
mixture of exponential function and power law [15,42]. In gen- AFM can be summarized as follows:
eral, hydrophobic force is determined by the hydrophobicity of i. Gaining fundamental knowledge of intermolecular forces.
the surface which is indicated by contact angle measurements. For Besides some basic forces, the hydrodynamic interactions between
less hydrophobic surfaces (45 < < 90 ) the hydrophobic force can deformable surfaces (oil droplet and thin lm) can be measured
be tted by exponential function, while it can be usually tted by [53,55,56,57].
double-exponential function or power law for highly hydrophobic ii. Elucidating the nature and characteristics of each force.
surfaces ( > 90 ). Origins of some forces that not included in DLVO theory
are still controversial. Additionally, the factors (ionic strength
[169,171,172], pH [70], et al.) affect intermolecular forces are vastly
2.2.4. Steric force investigated by SFA and AFM techniques.
Polymers have long been used to enhance oil recoveries as a iii. Estimation of surface charge/potential. Surface
chemical ooding method. Currently, polymer assisted LSW or low charge/potential is acquired by edl force measurements and
salinity polymer ooding (LSP) has been addressed to gain more oil theoretical computations with DLVO theory. The potential d at
recoveries [138,13]. Adsorption of polymers onto suspended par- the boundary of diffuse layer estimated by tting the long-range
ticles introduces two effects: thermal uctuations/protrusions and edl force curve, therefore, is called effective or apparent surface
chain bridging as polymer chains overlap [115]. Polymer molecules potential. SFA or AFM also serves as a quantitative basis for other
are thermally mobile but the movement is constrained by volume. methods, such as electrokinetic studies [96], potentiometric and
This restriction effect causes a reduction on the congurational conductometric titrations, adsorption density measurements,
entropy of polymer chains and gives rise to a repulsive steric force. negative adsorption and colloid stability analyses [191].
Another factor that contributes to steric force is an osmotic inter- iv. Giving indirect clues for mineral chemistry. A monotonic and
action arises from polymer chains protrude too far to attach and short-range repulsion is usually generated between hydrophilic
bind to other particles. This bridging force is usually attractive at surfaces, while hydrophobic surfaces create a monotonic but long-
large separations [115]. range attraction [107].
630 H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639

v. Estimating the thickness of surfactant/polymer adsorption


layers. It is also possible to establish the conformation or orien-
tation of molecules in the adsorbed layer [4446,143,36,113].

3.1. Interactions of pH-dependent sandstones

Sandstone is primarily composed of quartz but may incorporate


small amounts of clays. The structural unit of all silicate minerals is
the silicon tetrahedra. Silicates like quartz or K-feldspar consists of
stacked sheets of silicon tetrahedra with oxygen shared between
two tetrahedra. The mutual repulsion between the highly charged
cations (Si4+ ) in those tetrahedra ensures that the framework of the
silicates is fairly open and easily accommodates cations. The most
common substitution is Al3+ or Fe3+ replaces the Si4+ in the tetrahe-
Fig. 5. Retarded van der Waals force in air between crossed mica cylinders with a
dra, resulting in net negative charges in the crystal lattice. Charge
character that the power law index (n) changed from 2 (non-retarded) to 3 (retarded)
neutrality is maintained either by inserting appropriate amounts [112].
of Ca2+ , Na+ , Li+ , and K+ or adding H+ which binds to O2 to form
OH (Nesse, 2012). The surface charges of silica partly depend on
qualitatively interpreted by multicomponent ionic exchange (MIE)
the pH of the solution correlates to the following interactions:
mechanism [134,204] and the chemical mechanism proposed by
Austad et al. [12]. Quantitatively measurements of these interac-
tions are extensively available in the literature.
At higher pH (7.58), the silica surface becomes increasing
negatively-charged due to the ionization of silanol groups; in the 3.1.1. VDW force
pH range of 3 7.5, the surface charges are slightly pH-dependent; Tabor and Winterton [196] rst successful measured the vdW
however, the surface charges will be positive due to the adsorption forces between two mica surfaces in the non-retarded region
of H+ at pH below 3 [26,27,28]. Reservoir brines are typically at pH (530 nm). The results showed that at separations less than about
values above the iep (pH = 22.5) of silicates. 10 nm the forces were normal vdW forces whilst at distances
Clay minerals contained within sandstones though little, can greater than 20 nm they were retarded vdW forces. Israelachvili
exert great impact on the performance of LSW. Typical clay min- and Tabor [112], based on a series of jump and resonance exper-
erals include montmorillonite, kaolinite, illite, and chlorite. The iments, have observed a gradual transition from non-retarded to
structural unit for clay minerals is octahedral sheet which is made retarded vdW forces and a reduction of Hamaker constant from
up of either Al or Mg atoms in octahedral coordination with eight 12 to 50 nm between curved mica surfaces separated by air (see
oxygen atoms, and tetrahedral sheet with each Si atom being coor- Fig. 5). VdW forces seem to be independent of the cation type and
dinated with four oxygen atoms. Octahedral sheet and one or two concentration as the attractive vdW in 102 and 101 M Ca(NO3 )2
tetrahedral sheet then are tied together by sharing common oxygen were very similar to those measured in 101 and 1 M KNO3
atoms and leaving oxygen network or hydroxyls on basal surfaces solutions [116]. Also, Israelachvili and Adams [116] observed a
[33]. It is proved that three kinds of surface charges present on shorter retarded separation (6.5 nm) for mica surfaces immersed
clay minerals: (i) Permanent negative charges from isomorphous in mono/divalent electrolytes.
substitution of other atoms in crystal lattice. For example, Al3+ in Previous force measurements performed on glass surfaces
the octahedral sheet is replaced by Mg2+ and one charge deciency shown signicant retarded vdW forces in a range of 25 nm to
results. In the presence of other ions in brine solutions, the adsorbed 1200 nm (see Israelachvili and Tabor [112] and the references
cations (Mg2+ in the above example) undergo ion exchange process. therein). Rouweler and Overbeek [186] rst successfully measured
The permanent negative charges mostly occur on the surface of the dispersion forces between a fused glass lens and a at glass at
clay minerals. (ii) Variable negative charges which varies with pH. 25350 nm. A d3 -dependent retarded vdW forces at separations
Two reasons may contribute to the existence of variable negative greater than 50 nm were observed and an estimate of a transi-
charges: rstly, the protons in the hydroxyls, which are attached to tion region at 20 nm between retarded and non-retarded forces
Al atoms on crystal edge, dissociate in alkaline or neutral solutions, characterized with d2 was expected.
resulting in a negatively charged crystal edge; secondly, adsorp- VdW force is assumed to be very weak or even not exist for gold
tion of OH , SiO3 2 anions, or organic anions makes the edge of the surfaces. This is convinced by trying to reproduce the measured
clay crystals negatively charged. (iii) Positive charges on the crystal forces between a gold-coated sphere and a gold surface immersed
edge when the pH of the solution is lower than 9. This is because in 104 M NaCl [71] that a reasonable t was achieved by eliminat-
the hydroxyls which bind to Al atoms on the edge of clay crystals ing the vdW force. Similar conclusions also have been reached by
dissolute in more acid conditions. Since the negative charges on the other researchers [20,183,122,74,73]. Giesbers et al. [81] illustrated
clay surface overwhelm the positive charges on the edge, the clay that vdW forces between dissimilar surfaces, a silica sphere and a
crystals present a negatively charged property in nature [58]. thin gold lm, are negligible as well.
The interactions between brine/oil and brine/sandstone inter-
faces include: (i) electrostatic repulsions between two double 3.1.2. EDL force
layers formed at the charged brine/oil interface and brine/rock Extensive studies have proven that edl force at low salt concen-
interface, separately; (ii) hydrogen bonding between polar func- tration (mostly at 104 102 M) could be adequately characterized
tional groups in the crude oil and polar groups on clays; (iii) Lewis by classical DLVO theory; the magnitude, decay length and decrease
acid/base interactions between basic groups (e.g. NH4 + ) in the oil in surface potential/charge with increased ionic strength or pH3
and negatively charged groups on rock surface; (iv) chemical reac- [116,181,104]. Discrepancies happen at higher concentrations, for
tions between acidic groups (e.g. COO) in oil and divalent cations example at a concentration of 102 101 M the edl forces for K+
(usually Ca2+ and Mg2+ ) adsorbed on rock surface which forms solution still decay roughly exponentially with distance but the
organometallic complexes [24,151]. The last three interactions are mean decay lengths are 25% higher than the theoretical values
H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639 631

[116,181]. Above 1 M NH4 Cl concentrations, however, the edl forces It keeps increasing with concentration (e.g., 1 M in NaCl) until
cannot be described by DLVO theory [69]. Toprakcioglu et al. [200] approaching a maximum. The maximal adhesion corresponds to
have reported an even lower concentration ranges ( > 4 102 M) a full neutralization of surface charges. Then the adhesion starts
at which the diffuse potentials dropped to 30 mV that were not to decrease with much higher concentration which implies a re-
expected by theory for mica surfaces immersed in KNO3 solu- emergence of edl repulsion due to the charge reversal caused by
tions. In particular, the edl forces are much reduced in divalent excess adsorption of cations [66], see Fig. 6). An unanticipated
solutions than those in KNO3 solutions as the mean decay lengths nding was observed in 3:1 electrolytes (LaCl3 , [Co(NH3 )6 ]Cl3 ,
are lower by 2045% than the theoretical values even in diluted and AlCl3 ). Maximal adhesions, which are twice larger than those
(104 103 M) Ca(NO3 )2 and BaCl2 solutions [116]. observed in 1:1 and 2:1 electrolytes, are observed at 1 mM con-
Possible explanations for the above inconsistencies could be centration for all three solutions. This critical concentration is 3
the limitation of DLVO theory and surface contaminations. For orders of magnitude smaller than those of monovalent and divalent
instance, Shubin and Kkicheff [191] have reported a shorter decay electrolytes (see Fig. 6). Reasons behind the unexpected behavior
length 0.25 nm than the expected length 0.36 nm for mica sur- of trivalent ions need to be further explored but may be governed
faces immersed in 7.3 101 M LiNO3 solution. The difference is by ionion correlations. Subsequently, the adhesions vanish above
attributed to a decrease of effective dielectric constant and thus, a 2 mM owing to overcharging [66]. Charge reversal is also observed
non-continuum theory should be used. Additionally, the edl plane on mica surfaces through zeta potential measurements in the pres-
may be shifted by approximately one hydrated ion diameter which ence of Ca2+ during LSW. At concentrations beyond 0.05 M due to
indicates that additional hydration force arises at very short dis- a strong adsorption of Ca2+ onto mica surface, the surface charge
tances. Moreover, as suggested by Pashley [169], a continual ionic turns from negative to positive. Accordingly it will enhance the oil
exchange happens between the mineral lattice and the aqueous production by reducing or removing the Ca2+ ions in low salinity
solution. Differences between surface preparation and treatment water [148].
methods have great inuences on force behaviors. Ducker et al.
[70] noticed that their apparent surface potentials are lower than
those obtained from streaming potential measurements on silica 3.1.3. Hydration force
capillaries but higher than those from tting force curves on large Generally, at relatively higher electrolyte concentrations
silica sheets by Horn et al. [105]. Actually, there are great differences (102 101 M) and at separations less than 5 nm hydration force
between densities of silanol groups on these surfaces due to differ- arises [116]. The hydration force between hydrophilic surfaces is
ent surface preparation processes. The edl force for silica surfaces monotonically repulsive and has been measured by AFM at a sepa-
at a given concentration and separation is increased in the order ration of 1.55 nm. In Pashleys serial AFM measurements between
of amed <steamed < ammonia-exposed [84]. Even worse is that mica surfaces in monovalent ions (Li+ , Na+ , K+ and Cs+ ) [170],
the adsorption of other materials may totally change the surface divalent ions (Mg2+ and Ca2+ ) [172], and trivalent ion (Li3+ and
properties and make the DLVO theory to fail [107,81].Other expla- Cr3+ ) [171], all detected additional short-range ( < 2.5 nm) repulsive
nations like pH variations due to the dissolution of atmospheric forces at higher ionic concentrations. However, the critical con-
CO2 and not strictly following the cleaning procedures, cannot be centration varies with ion type: in monovalent electrolytes, the
ruled out either [107]. concentrations from which the hydration forces to be observed
The electrostatic nature determines that the edl is rather are in the order Cs+ <K+ < Li+ Na+ . Divalent ions show a similar
sensitive to the ionic strength, which is characterized by ion con- trend with monovalent ions that is in the order Ca2+ < Mg2+ . The
centrations and valences. Furthermore, the edl also depends largely overall trend is given as Cs+ <K+ <Li+ Na+ <Ca2+ <Mg2+ <La3+ and
on the ionic specicity. This concept is highlighted as specic ion is interpreted as the higher the hydrated degree, the higher the
effect which supposes that ions adsorb specically to the particle concentration required for that cation to displace hydrogen ions
surface and modify its surface charge. It further species the size originally adsorbed onto the mica surface [173]. In other words,
and hydration of ions which follows the Hofmeister series [164]. hydration forces are originated from the physically adsorbing of
As discussed before, at low concentration range (e.g., <103 M), the hydrated cations onto mica surface which are subjected to a cation
colloids tend to be more stable with decrease in ionic strength exchange process with hydrogen ions. The difference between
as theoretically predicted. It is the theoretical basis of LSW in divalent ions and monovalent ions is that monovalent cations are
petroleum engineering. However, stability is disturbed because of more strongly hydrated and therefore do not easily shed their
ionic screening by further increasing the concentration. Yotsumoto hydration layers, instead, they keep the residual hydration shells
and Yoon [215] have observed a reduction in stability of silica sus- in order to adhere onto the mica surface [173].This rule is also
pensions at iep with the increase in NaCl concentration up to 2 M. reasonable for rutile surfaces [215].
The repulsive edl and hydration force for silica surfaces can be com- Hydration forces between mica surfaces in K+ and Na+ solu-
pletely removed by raising the pH to 10.3 or adding Ca2+ to 1 M tions (above 104 M) have shown to be exponentially repulsive
concentration [145]. Same conclusion was reached but half mag- with decay lengths of 1.0 0.2 nm [169,170]. The decay lengths of
nitude of values were given by Valle-Delgado et al. Delgado et al. Mg2+ and Ca2+ solutions are much larger (1.8-2.0 nm) which are
(2005). EDL repulsion becomes negligible at 1 M NaCl or 0.5 M CaCl2 enough to remove the hydration minimum as a result of the balance
solution, which highlights the hydration repulsion and vdW attrac- between repulsive hydration force and attractive vdW force. Hydra-
tion at around 2 nm. This inconsistency may be due to the different tion forces in La3+ electrolytes, on the other hand, are extended to
silica materials used by the two authors: the use of a glass sphere much longer distance (ca. 30 nm) and are quantitatively described
and a silica wafer by Meagher [145] and a nonporous silica sphere by a double exponential function with decay lengths of 0.5 nm and
and a very smooth silica plane by Valle-Delgado et al. [205] may 3.0 nm, respectively [172]. Inuences of multivalent ions on the
have contributed to different densities of silanol surface groups and range and decay length of hydration force are assumed to be specic
therefore, different numbers of negative sites on silica surface. adsorption and ionic screening. The rst one indicates that divalent
At rather high concentrations, the colloids have been observed cations (e.g., Ca2+ ) have a favorable adsorption ability compared to
to restable. AFM adhesion measurements for silica surfaces at 1:1 monovalent cations (e.g., Na+ ). Given same concentration divalent
(NaCl, KCl and CsCl) and 2:1 (MgCl2 , CaCl2 and SrCl2 ) aqueous solu- cations as well, have a larger screening effect on reducing the edl
tions suggested that adhesion appears until a specic concentration force; Anisotropic Hypernetted Chain Approximation (AHNC) cal-
is reached, e.g., 0.10.2 M in NaCl, 0.050.1 M for CaCl2 and MgCl2 . culation results indicated that the negative charges on each mica
632 H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639

Fig. 6. (a) Adhesion (full-off force) between a silica bead and a at silica surface as a function of ionic strength in 3:1, 2:1, and 1:1 electrolytes at pH 5.5 (AlCl3 at pH 3.8). (b)
Force vs separation curves in different salt solutions. The ion specic concentration is chosen at which edl was negligible [66].

surface were totally compensated by counterions at intermediate


concentration (0.05-0.24 M) of Ca(NO3 )2 solutions [124].
Unlike molecularly smooth mica, the origin of the additional
short-range repulsion on amorphous silica is still in debate. Ducker
et al. [71] postulated that this additional repulsive force was orig-
inated from a residue water lm or hydrated ions adsorbed on the
glass surface which results in a shift of charge plane. Vigil et al.
[208] proposed another explanation regarding a silica gel-layer
(10 thick of silanol and silicilic acid groups) that shifts the OHP
0.5 nm further out from each surface which effectively enhances
double-layer repulsion at relatively large separations (up to 5 nm),
moreover, the silica hairs along protrude into aqueous solution
which adds a monotonic steric repulsion by overlapping gel-layers
at small separations. They argued that the range of hydration force
decreases with an increase of ionic strength, a decrease of pH or
adding of Ca2+ shrinks the anionic silica gel-layer and hence, gives
rise to a shorter range steric repulsion. It has been assumed that the
overlapping of water layers on silica surface gives rise to the short-
range repulsion. Nevertheless, more studies seem to agree with
the hypothesis that hydration force which arises from the over-
Fig. 7. Schematics of hydration force for mica/silica surface. (a) Ion exchange
lapping of water layers on silica surface that is responsible for the
between the fully hydrated counterions on mica surface. (b) Small counterions
short-range repulsion. with a higher afnity for water penetrate deeper into the hydrated layer than large
Hydration forces between silica surfaces at monovalent solu- counterions on silica surface [51].
tions (104 101 M) are found to be monotonic, exponentially
decaying with a decay length of 0.51.0 nm below 23 nm. The
ter [51]. Because the rst adsorbed layer of hydrated counterions
hydration force is insensitive to ionic strength which is opposed to
is completely neutralized at high concentrations, it is assumed
the secondary hydration force on mica surfaces [181,70,71,84,42].
that the structured water layer is more favorable to accommo-
The more hydrated cations produce the weaker is the hydration
date smaller ions with a higher afnity for water. Larger ions are
forces for silica surfaces, that is Li+ <Na+ <K+ <Cs+ , which is also
more likely to reside outside the hydrated layer (see Fig. 7). Direct
opposite to the general trend for mica surfaces. Moreover, the
verication includes intermolecular force measurements [65,66]
short-range force is always present and is the strongest between
between silica surfaces. The adsorption capacity of monovalent and
silica surfaces in deionized water [42]. Unlike the simple the case
divalent cation onto silica surface is ranked as Mg2+ < Ca2+ < Na+ <
of mica, the hydration force for silica surfaces in fact comes from the
Sr2+ < K+ < Cs+ according to the size of cations as well as solvation
breakdown of hydrogen bonding network built on silanol groups;
energy but not correlating with hydrated radius.
because each cation exchanges a H+ which reduces the extent of
network and in turn diminishes the repulsion. Li+ , for example, is
3.1.4. Structural force
the most hydrated cation and it has the strongest power to break the
Oscillatory structural force arises from closely packed nite
network and therefore the lowest hydration force. Li+ is structure
sized molecules with restricted space between the two surfaces. It
maker in mica, but a structure breaker in silica [42]. A similar
usually extends to about 1.5 nm with a periodicity of 0.25 0.05 nm
conclusion has already been extracted by Peschel et al. [175] on
[117]. Israelachvili and Pashley [117] argued that measurement
studying the interactions between two polished silica plates sepa-
of short-range forces from the deection of a single cantilever is
rated by thin aqueous LiCl, NaCl and KCl layers. They found that the
not enough to disclose the high repulsive forces. In the follow-up
decay lengths increased in the order Li+ <Na+ <K+ which is in agree-
studies, they used a variable stiffness double-spring cantilever and
ment with Chapels [42] work. Conversely, the trend at higher ionic
measured over six orders of forces. Structural forces between mica
strengths for silica is opposite to the ion adsorption sequence at
surfaces in monovalent KCl electrolytes were found to be oscillated
low ionic strength (104 -101 M). An indirect clue showed that the
for several steps with a periodicity 0.24 0.2 nm (almost equals to
highest viscosity, which corresponds to the lowest hydration force,
the diameter of one water molecule) below 1.5 nm. The structural
of silica slurries (1 M and 4 M) was obtained for Li+ ion. The general
force between mica sheets immersed in 1 M NaCl solution reported
order of hydration force hence is in the order of Li+ <Na+ <K+ < Cs+
by McGuiggan and Pashley [143] also showed jagged forces super-
which correlates to the bare size rather than the hydrated diame-
imposed on a strong monotonically repulsive force. The width of
H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639 633

these steps was found to be 0.30 0.05 nm, which is slightly larger found in heavy fraction of crude oils like resin and asphaltene and
than those measured in KCl solutions (0.25 0.03 nm) [173], but the it is suggested that the carboxyl group plays a more important role
number of steps is same as observed in KCl solutions. In 0.1 M NaCl on the recovery of LSW compared with base number [195]. Negative
solution, in contrast, an oscillatory force with similar periodicity but COO binds to the positively charged carbonate surface or adsorbs
a shorter range was observed. Ducker and Pashley [69] conrmed on the clay surface on sandstone rock could turn these two surfaces
the oscillatory nature of structural force by measuring the inter- more oil-wet. Positive NH3 + can also adsorb on the clay/silica sur-
actions between two mica crystals in aqueous NH4 Cl solutions. A face. The adsorption of acidic and basic groups onto clay surface is
short range ( < 1.5 nm) repulsion appeared at concentration above very pH-dependent [11].
103 M and oscillated with a periodicity of about 0.4 nm. This peri-
odicity is regarded as an integral of adsorbed hydrated NH4 + and
water molecular packing in thin lm. Overall, it has been observed 3.2.1. Rigid tip-oil droplet and drop-drop interactions
that varieties of ions give the similar oscillatory force which sug- To date, interactions between a rigid probe and a hard sur-
gests that structural force is caused by water molecules layering face are well emphasized. Many experiments are also carried
in thin lms between molecularly smooth, rigid surfaces. Never- out for a rigid probe approaches using a single air bubble
theless specic ion-water interaction may vary the periodicity of [198,32,72,78,176,177,18,156,21,83,110,6,75,4]. Models to quan-
these oscillations [69]. The oscillatory structural force likewise has tify the probe-bubble interactions are not well developed to
been observed in divalent electrolytes. At concentrated CaCl2 solu- elucidate the deformation characteristics of air bubble. Instead an
tion (0.12 M), short-range oscillatory force added to a monotonic air bubble is just treated as a normal hydrophobic surface. In this
tail appeared for two muscovite mica surfaces due to adsorp- paper the probe-bubble interactions in terms of intermolecular
tion of hydrated Ca2+ ions [131]. Upon detailed examination it interactions related to oil droplet is of interest. For rigid surfaces,
was concluded that the strength of structural repulsion is one to a complete force curve can be captured in the compliance range of
four orders higher than the hydration force observed at 0.12 M the cantilever, while a much longer compliance range is required
Ca2+ concentration [172]. Note that in divalent ions like Ca2+ , the to achieve constant compliance for oil droplets due to the deforma-
oscillated structural force is also found on the dynamic rough sil- tion. Long range force, such as the edl force is impossible to detect
ica surfaces (ever changing/dynamic). Indeed, Fielden et al. [79] since a weak deformation of the oil droplet shades the deection of
have observed an oscillatory structural force with the periodic- the cantilever. Weak vdW force may also be masked for the same
ity associated with expulsion of water layers in the direct force reason [55].
measurements between silica surfaces in 1 and 5 M CaCl2 solutions, A large and growing body of studies has focus on this area
both from experimental and theoretical aspects, e.g., direct force
3.1.5. Hydrophobic force measurements between a rigid probe and a single oil droplet
Strong monotonically attraction emerges between two [150,193,95,18,7,10,210], and two deformable oil drops immersed
hydrophobic surfaces and it is referred to hydrophobic force. in another immiscible uid [91,34,57,55]. Quantitative description
Hydrophobic force originates from the expulsion of entropically of forces act on deformable oil droplet relies on a coupling of the
unfavorable water molecules from the surface [143] and may static interactions (edl force, vdW force, and hydrophobic force)
be 10100 times stronger than the vdW attraction [111,107]. and dynamic drainage. The hydrodynamic force involved here is
Hydrophobic force is commonly of long-range (several tens of referred as the force that exerted on the particles by liquid lm
nanometers) and has a decay length of 12 nm [182]. This char- before a hydrophobized glass colloid approaches an oil drop in
acteristic force has been observed on hydrophobic mica surfaces aqueous solution. For soft materials (oil drop) the deformation will
[48,50,49,203], 4hydrophobic glass surfaces [180,182,214] or prevent the thinning and rupture of the lm, therefore the drainage
dissimilar surfaces [201,202]. Alternative explanation has been of the thin lm is a function of oil droplet radius and external driving
given in terms of ionion correlation by [9] which indicates that velocity of the scanner [7,8]. Several models have been proposed to
hydrophobic force arises from an enhanced electrostatic response determine the dynamic interactions between deformable surfaces
by large charge/dipole uctuation. Therefore the long-range [133,38,14,54,154,56,35,83,40].
attraction decays exponentially with a decay length equals to Drop-drop interactions are measured by mounting one oil
half the Debye length at high salt concentration and remains droplet onto a tipless cantilever and the other attaching to a at
constant independent of salt concentration in particular at low salt base (see Fig. 8a). A thin lm will be formed along the at sur-
concentrations. face in the surfactant (usually SDS) or protein solutions [91,55].
Hydrophobic force has been measured between hydropho- The presence of thin aqueous lm between oil droplets causes a
bized quartz laments in 102 M KCl solution. Different degrees of change in drainage velocity: the drainage is quickly and evenly for
hydrophobicity are achieved by treating the glass surfaces in 10% surfactant system with highly mobile interfacial lm, on the con-
DMOCS solution or DMOCS vapor for different times. The advanc- trary, the protein stabilized interfaces are immobile and drainage
ing contact angle ( a = 100 ) was independent of time and types of is low and unevenly. Drainage is faster at the edges of two oil
surface treatment (in vapor or solution). The receding angle var- droplets than in the center, resulting in a dimple with trapped water
ied from 55 to 70 depending on both time and kind of treatment. [91]. Force curves at pure anionic surfactant (SDS) shows that the
The hydrophobic attraction that appears at 15 nm increases with droplets remain stable due to signicant edl force, but in the mix-
the growth of contact angle, in other words, the hydrophobicity. ture of SDS and 1 mM NaCl, the oil droplets tend to approach each
The strength of the attraction is almost 104 times greater than vdW other because of ionic screening by counterions. The oil droplets are
force and extends to 200300 nm [181]. forced to go together in a mixture of SDS and 40 mM NaCl due to
a greater screening effect [91]. For large droplets (above 100 m),
3.2. Interactions of deformable oil surfaces the interfacial deformation and hydrodynamic drainage effect are
obvious due to an equilibrium result with surface forces. However,
It is believed that the oil composition relates to the causes and there will be no dimple in the interfacial prole for small droplets
consequences of LSW. Crude oil basically contains two kinds of with low deformation. At velocities similar to Brownian motion, the
functional groups: carboxyl group ( COOH) and amide ( NH2 ) contributions of hydrodynamic and surface forces are strongly cou-
derived from carboxylic acids, which are determined by acid num- pled and the relative length scales will determine which component
ber and basic number (mg KOH/g), separately. COOH is mostly dominates (see Fig. 8b). At higher velocities (about twice Brown-
634 H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639

Fig. 8. (a) A schematic of AFM force measurement between two oil droplets (one immobilized on the cantilever and the other immobilized on the substrate). (b) Dynamic
interaction force F versus piezo drive motion X between two decane droplets in 0.1 mM NaCl solution with SDS at a series of driving velocities (green circles, 2 m/s; blue
triangles, 9.3 m/s; red diamonds, 28 m/s). Open symbols (approach) and lled symbols (retract) refer to the experimental data, and the solid lines are numerical simulation
results based on Stokes-Reynolds-Young-Laplace Model (Dagastine et al., 2006). (For interpretation of the references to colour in this gure legend, the reader is referred to
the web version of this article.).

ian motion), hydrodynamic drainage dominates the total pressure in oil on wettability. As shown by Mugele et al. [148] that even small
[55]. amount of polar hydrocarbon and stearic acid was able to enhance
the hydrophobicity of mica surface. In fact as proved by AFM imag-
3.2.2. Chemical force microscopy (CFM) ing, bare mica surface was seen with occasional groups of mono
Recently, Chemical Force Microscopy (CFM) with different stearate layer and this layer is much thicker in the vicinity of the
chemical modied tips has been used to gain insights into LSW original contact line of the droplet with occasional holes. Meaning
mechanism for sandstones [99,98,142,101,97,120,121] and chalk that Ca2+ and S.A., a typical surfactant synergistically transform the
[100,192] based on adhesion measurement. The basic idea is to mica surface to a more oil-wet state by promoting the self-assembly
modify the gold-coated tip with polar/nonpolar or basic/acidic of hydrophobic Ca-stearate monolayers.
groups in order to investigate how different oil composition, rock
minerology, and interface physics/chemistry affect LSW. More 3.3. Interactions of chemical active carbonates
specically:
Carbonate minerals are divided into calcite, dolomite and lime-
i. CH3 modied tip is used to simulate hydrophobic surface or stone, amongst these minerals calcite (CaCO3 ) is the most abundant
nonpolar oil droplet, carbonate minerals [155]. In this paper calcite is taken as an exam-
ii. CF3 modied tip is used to simulate hydrophobic and polar ple in order to introduce the crystal structure of carbonate minerals.
surface, The structural element of calcite is the CO3 2 anionic group. The
iii. NH2 modied tip is used to simulate basic group NH3 + of oil, central Ca2+ coordinates to six oxygens from distinctive CO3 2
iv. COOH modied tip to simulate acidic group COO of oil, group, forming a slightly distort octahedra. Adjacent octahedra are
v. OH modied tip is used to simulate polar groups of oil and also linked both by CO3 2 groups and sharing corners of octahedra from
to detect the effect of hydrogen bonding. layers above and below. Ca2+ layers alternate with CO3 2 layers
along the c axis and CO3 2 groups reverse orientations in successive
layers [184].
Amongst the above, CH3 and COOH terminated tips are the
Carbonate minerals are chemically active with dissolu-
most favorable ones to explore the relationship between wetta-
tion/precipitation, adsorption and substitution happens in an
bility and salinity. The extracted force volume maps along with
oil/brine/carbonate system. Calcite can dissolve in brine solutions
captured topography images provide us sufcient information
to produce Ca2+ , CaH2 O+ , CO3 2 , HCO3 , and H2 CO3 [102]. If pH is
about nanoscale interactions: the adhesion between CH3 tip and
below a certain value (68), the dissolution tends to produce more
quartz grain is higher (more oil-wet) in articial seawater than
Ca2+ and the calcite surface is more positively charged. On the other
that in dilute seawater (see Fig. 9). Additionally, adhesion is more
hand, if the pH is high enough ( > 8.7), the reactions produce more
signicant in the presence of heterogeneous distributed organic
CO3 2 and the calcite surface is more negatively charged. In the pH
material as it offers very sticky anchor points for oil molecules
range of formation water (6.57.5), however, the surface charge of
to adhere onto the quartz surface [142] (see Fig. 9). Hilner et al.
calcite is positive. Anhydrite dissolution and Ca/Mg substitution are
[101] has quantied the adhesion between a nonpolar CH3 mod-
expressed by the following equations:
ied tip (which served as oil droplet) and sand grains taken from
core plugs in articial formation water and its diluted solutions CaSO4 (s) Ca2+ +SO4 2
as a function of ionic strength (salinity ranges from 28,500 ppm to
1,500 ppm). The results show that the adhesion reduces as salin- CaCaCO3 (s)+Mg2+ MgCaCO3 (s)+Ca2+
ity decreases below a threshold of 50008000 ppm, indicating a
less oil-wet transition. Clay is necessary in increasing edl expan- The adsorption of sulfate ions which produced by anhydrite dis-
sion either through increasing pore surface charges or providing solution [189] also enhances the negative charges on the carbonate
higher surface areas and anchor points for the oil compounds. As surface. This leads to co-adsorption of Ca2+ and Mg2+ onto the rock.
a result, clay will enhance wettability alteration from oil-wet or The Ca2+ ions then react with carboxylic groups in oil and adhere
mix-wet to more water-wet in sandstone reservoirs. COOH ter- to the carbonate surface. The reaction breaks the bonds between
minated tip can be used to elucidate the effect of polar components the brine/carbonate and brine/oil interfaces and releases the car-
H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639 635

Fig. 9. Topography and adhesion map created by CH3 modied tip reacts with a quartz grain during articial seawater (ASW) and low ASW ooding. (a) A height map with
a at area enclosed by red line (clean quartz) and several topographically higher features that enclosed by green lines (adsorbed organic material). (be) Adhesion maps after
two cycles of ASW to ASW-low ooding. Note: high adhesion (red spots) is observed on the at area but near the topographically higher parts. Average adhesion corresponds
to different areas as listed in the table [142]. (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article.)

tions are activated and make brine/carbonate interface to become


more negatively charged and repelled by the negatively charged
brine/oil interface. Though it is still debated on the reasons for
wettability alteration in carbonate reservoirs, double layer expan-
sion is believed to be the main process responsible for improving
production by LSW ooding [2].

4. Conclusion

This paper gives a review of some theoretical and experimental


studies on wettability alteration mechanism due to double layer
expansion during low salinity waterooding on both sandstone
and carbonate reservoirs. A few points that addressed in this paper
Fig. 10. Schematic of wettability alteration induced by seawater in chalk. (a) Ca2+ mainly include:
and SO4 2 involved reactions at lower and high temperatures. (b) Mg2+ and SO4 2
acts in releasing oil at high temperatures [218].
i. Low salinity waterooding is an efcient way to improve oil
recovery for sandstone and carbonate reservoirs as proved by
boxylic groups, resulting in improving oil recovery (see Fig. 10). eld test and laboratory experiments. However, the content of
Another benet of anhydrite dissolution is that it may increase the low salinity waterooding is different for these two kinds of
connectivity of the pores so that the reservoir becomes more per- reservoirs.
meable to ow. At sufciently high temperatures ( > 100 C), Mg2+ , ii. Wettability alteration is regarded as the primary mechanism
either brings by low salinity brine or dissolves from dolomite, sub- of low salinity waterooding in sandstone and carbonate
stitutes Ca2+ from the carbonate surface and thereby displace Ca2+ reservoirs. Wettability alteration is triggered by double layer
ions that are bridged to carboxylic groups, thus the Ca/Mg sub- expansion mechanism which is correlated to the electrical and
stitution further improves oil recovery [185,218] (see Fig. 10). To chemical properties of the oil/brine/rock system.
conclude that Ca2+ , Mg2+ and SO4 2 are the pdi for carbonate reser- iii. Intermolecular interactions in the oil/brine/rock system can be
voirs. It should be noted that either Ca2+ or Mg2+ together with qualitatively and quantitatively described by an electrochemical
SO4 2 are able to change the wettability of carbonate reservoirs but model. Wettability is directly associated with disjoining pressure
SO4 2 alone is not capable to achieve the same outcome [185,146]. which is responsible for the stability of water lm. Disjoining
It is much favorable to reduce or remove the Na+ ions in injected pressure is usually composed of van der Waals force, electrical
brine as the Na+ ions will prevent the pdi from reaching and react- force and structural force which belong to the domain of DLVO
ing with the oil or carbonate minerals [77]. The chemical response and non-DLVO theory. Note that it is hard to quantify the dis-
in carbonate reservoirs during LSW can be simulated with designed joining pressure in a real oil/brine/rock system because lots of
geochemical software like PHREEEQC [59,25], in-house simulator factors (pH, temperature, oil composition, surface heterogeneity
[178,179,211] which incorporates ow simulation to gain insights etc.) may ruin the predicted results. Future trials should be done
into which chemical process that dominates in LSW. to assess the impact of these factors.
When low salinity brine is injected into carbonate reservoirs, iv. Surface force apparatus like atomic force microscopy provides
the thermodynamic equilibrium among the ions dissolved in the us a powerful way to handle nanoscale intermolecular forces
brine lm, ions adsorbed onto the rock surface and the species among oil/brine/rock system. Along with topography imaging
incorporated into the rock matrix is disturbed. All the above reac- and tip modied technique, we can qualitatively investigate how
636 H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639

oil or brine composition and mineral properties affect the per- [25] P.V. Brady, J.L. Krumhansl, P.E. Mariner, Surface complexation modeling for
formance of low salinity waterooding. Considerably more work improved oil recovery. SPE improved oil recovery symposium, Soc.
Petroleum Eng. (2012).
will need to be done with reservoir rock samples and crude oil [26] P.V. Brady, J.V. Walther, Controls on silicate dissolution rates in neutral and
in real reservoirs conditions (relatively high temperature and basic pH solutions at 25C, Geochim. Cosmochim. Acta 53 (1989) 28232830.
pressure). [27] P.V. Brady, J.V. Walther, Kinetics of quartz dissolution at low temperatures,
Chem. Geol. 82 (1990) 253264.
v. The precise mechanism of wettability alteration in carbonates [28] P.V. Brady, J.V. Walther, Surface chemistry and silicate dissolution at
remains to be elucidated. Although calcite and anhydrite dis- elevated temperatures, Am. J. Sci. 292 (1992) 639658.
solution are highly empathized, the nal result is a change of [29] J. Buckley, K. Takamura, N. Morrow, Inuence of electrical surface charges on
the wetting properties of crude oils, SPE Reservoir Eng. 4 (1989) 332340.
surface charge/potential which rooted in the territory of double
[30] J.S. Buckley, Chemistry of the crude oil/brine interface. proc. 3 rd
layer expansion. This paper therefore holds the view that the international symposium on evaluation of reservoir wettability and its,
wettability alteration caused by double layer expansion is the Effect Oil Recovery (1994) 3338.
[31] M.A. Butkus, D. Grasso, Impact of aqueous electrolytes on interfacial energy,
nal ending of low salinity waterooding. But the processes (pH
J. Colloid Interface Sci. 200 (1998) 172181.
increase, MIE, Austad et al.s chemical mechanism or else may [32] H.-J. Butt, A technique for measuring the force between a colloidal particle
operate simultaneously) that led to this result are still possible in water and a bubble, J. Colloid Interface Sci. 166 (1994) 109117.
area of future research. [33] R. Caenn, H.C. Darley, G.R. Gray, Composition and Properties of Drilling and
Completion Fluids, Gulf professional publishing, 2011.
[34] S.L. Carnie, D.Y. Chan, C. Lewis, R. Manica, R.R. Dagastine, Measurement of
dynamical forces between deformable drops using the atomic force
microscope. I. Theory, Langmuir 21 (2005) 29122922.
References [35] S.L. Carnie, D.Y. Chan, R. Manica, Modelling dropdrop interactions in an
atomic force microscope, ANZIAM J. 46 (2005) 805819.
[1] W. Abdallah, J.S. Buckley, A. Carnegie, J. Edwards, B. Herold, E. Fordham, A. [36] G. Cevc, Hydration force and the interfacial structure of the polar surface, J.
Graue, T. Habashy, N. Seleznev, C. Signer, Fundamentals of wettability, Chem. Soc. Faraday Trans. 87 (1991) 27332739.
Technology 38 (1986) 268. [37] G. Cevc, A. Watts, D. Marsh, Titration of the phase transition of
[2] E.W. Al-Shalabi, K. Sepehrnoori, G. Pope, Geochemical interpretation of phosphatidylserine bilayer membranes. Effects of pH, surface electrostatics,
low-salinity-water injection in carbonate oil reservoirs, SPE J. (2015). ion binding, and head-group hydration, Biochemistry 20 (1981) 49554965.
[3] E. Al Shalabi, M. Delshad, K. Sepehrnoori, Does the double layer expansion [38] D. Chan, R. Dagastine, L. White, Forces between a rigid probe particle and a
mechanism contribute to the LSWI effect on hydrocarbon recovery from liquid interface: i. The repulsive case, J. Colloid Interface Sci. 236 (2001)
carbonate rocks? SPE reservoir characterization and simulation conference 141154.
and exhibition, Soc. Petroleum Eng. (2013). [39] D. Chan, J. Henry, L. White, The interaction of colloidal particles collected at
[4] J. Ally, M. Kappl, H.-J.R. Butt, A. Amirfazli, Detachment force of particles from uid interfaces, J. Colloid Interface Sci. 79 (1981) 410418.
air- liquid interfaces of lms and bubbles, Langmuir 26 (2010) 1813518143. [40] D.Y. Chan, E. Klaseboer, R. Manica, Film drainage and coalescence between
[5] M.B. Alotaibi, R.A. Nasralla, H.A. Nasr-El-Din, Wettability studies using deformable drops and bubbles, Soft Matter 7 (2011) 22352264.
low-salinity water in sandstone reservoirs, SPE Reservoir Eval. Eng. 14 [41] D.Y. Chan, R.M. Pashley, L.R. White, A simple algorithm for the calculation of
(2011) 713725. the electrostatic repulsion between identical charged surfaces in electrolyte,
[6] S. Assemi, A.V. Nguyen, J.D. Miller, Direct measurement of particlebubble J. Colloid Interface Sci. 77 (1980) 283285.
interaction forces using atomic force microscopy, Int. J. Miner. Process. 89 [42] J.-P. Chapel, Electrolyte species dependent hydration forces between silica
(2008) 6570. surfaces, Langmuir 10 (1994) 42374243.
[7] D.E. Aston, J.C. Berg, Quantitative analysis of uid interfaceatomic force [43] G.V. Chilingar, T. Yen, Some notes on wettability and relative permeabilities
microscopy, J. Colloid Interface Sci. 235 (2001) 162169. of carbonate reservoir rocks, II, Energy Sources 7 (1983) 6775.
[8] D.E. Aston, J.C. Berg, Thin-lm hydrodynamics in uid interface-atomic force [44] H.K. Christenson, Experimental measurements of solvation forces in
microscopy, Ind. Eng. Chem. Res. 41 (2002) 389396. nonpolar liquids, J. Chem. Phys. 78 (1983) 69066913.
[9] P. Attard, Long-range attraction between hydrophobic surfaces, J. Phys. [45] H.K. Christenson, Non-DLVO forces between surfaces-solvation: hydration
Chem. 93 (1989) 64416444. and capillary effects, J. Dispersion Sci. Technol. 9 (1988) 171206.
[10] P. Attard, S.J. Miklavcic, Effective spring description of a bubble or a droplet [46] H.K. Christenson, P.M. Claesson, R.M. Pashley, The hydrophobic interaction
interacting with a particle, J. Colloid Interface Sci. 247 (2002) 255257. between macroscopic surfaces, in: Proceedings of the Indian Academy of
[11] T. Austad, Water-based EOR in carbonates and sandstones: new chemical Sciences-Chemical Sciences, Springer, 1987, pp. 379389.
understanding of the EOR potential using smart water, Enhanced Oil [47] N. Churaev, B. Derjaguin, Inclusion of structural forces in the theory of
Recovery Field Case Stud. (2013) 301335. stability of colloids and lms, J. Colloid Interface Sci. 103 (1985) 542553.
[12] T. Austad, A. Rezaeidoust, T. Puntervold, Chemical mechanism of low [48] P. Claesson, R. Horn, R. Pashley, Measurement of surface forces between
salinity water ooding in sandstone reservoirs. SPE improved oil recovery mica sheets immersed in aqueous quaternary ammonium ion solutions, J.
symposium, Soc. Petroleum Eng. (2010). Colloid Interface Sci. 100 (1984) 250263.
[13] S.C. Ayirala, E. Uehara-Nagamine, A.N. Matzakos, R.W. Chin, P.H. Doe, P.J. [49] P.M. Claesson, H.K. Christenson, Very long range attractive forces between
Van Den Hoek, A designer water process for offshore low salinity and uncharged hydrocarbon and uorocarbon surfaces in water, J. Phys. Chem.
polymer ooding applications. SPE Improved Oil Recovery Symposium, Soc. 92 (1988) 16501655.
Petroleum Eng. (2010). [50] P.M. Claesson, R. Kjellander, P. Stenius, H.K. Christenson, Direct
[14] D. Bardos, Contact angle dependence of solid probe?liquid drop forces in measurement of temperature-dependent interactions between non-ionic
AFM measurements, Surf. Sci. 517 (2002) 157176. surfactant layers, J. Chem. Soc. Faraday Trans. 1: Phys. Chem. Condensed
[15] S. Basu, M.M. Sharma, Effect of dielectric saturation on disjoining pressure in Phases 82 (1986) 27352746.
thin lms of aqueous electrolytes, J. Colloid Interface Sci. 165 (1994) [51] M. Colic, M.L. Fisher, G.V. Franks, Inuence of ion size on short-range
355366. repulsive forces between silica surfaces, Langmuir 14 (1998) 61076112.
[16] S. Basu, M.M. Sharma, Measurement of critical disjoining pressure for [52] T. Cosgrove, Colloid Science: Principles, Methods and Applications, John
dewetting of solid surfaces, J. Colloid Interface Sci. 181 (1996) 443455. Wiley & Sons, 2010.
[17] S. Berg, A. Cense, E. Jansen, K. Bakker, Direct experimental evidence of [53] R. Dagastine, Forces between a rigid probe particle and a liquid interface: II.
wettability modication by low salinity, Petrophysics (2010) 51. The general case, J. Colloid Interface Sci. 247 (2002) 310320.
[18] D. Bhatt, J. Newman, C. Radke, Equilibrium force isotherms of a deformable [54] R. Dagastine, L. White, Forces between a rigid probe particle and a liquid
bubble/drop interacting with a solid particle across a thin liquid lm, interface: II. The general case, J. Colloid Interface Sci. 247 (2002) 310320.
Langmuir 17 (2001) 116130. [55] R.R. Dagastine, R. Manica, S.L. Carnie, D. Chan, G.W. Stevens, F. Grieser,
[19] S. Bhattacharjee, C.-H. Ko, M. Elimelech, DLVO interaction between rough Dynamic forces between two deformable oil droplets in water, Science 313
surfaces, Langmuir 14 (1998) 33653375. (2006) 210213.
[20] S. Biggs, P. Mulvaney, C. Zukoski, F. Grieser, Study of anion adsorption at the [56] R.R. Dagastine, D.C. Prieve, L.R. White, Forces between a rigid probe particle
gold-aqueous solution interface by atomic force microscopy, J. Am. Chem. and a liquid interface: III. Extraction of the planar half-space interaction
Soc. 116 (1994) 91509157. energy E (D), J. Colloid Interface Sci. 269 (2004) 8496.
[21] E. Bonaccurso, G. Gillies, Revealing contamination on AFM cantilevers by [57] R.R. Dagastine, G.W. Stevens, D. Chan, F. Grieser, Forces between two oil
microdrops and microbubbles, Langmuir 20 (2004) 1182411827. drops in aqueous solution measured by AFM, J. Colloid Interface Sci. 273
[22] I. Borukhov, D. Andelman, H. Orland, Steric effects in electrolytes: a (2004) 339342.
modied poisson-boltzmann equation, Phys. Rev. Lett. 79 (1997) 435. [58] H.C. Darley, G.R. Gray, Composition and Properties of Drilling and
[23] W.R. Bowen, F. Jenner, The calculation of dispersion forces for engineering Completion Fluids, Gulf Professional Publishing., 1988.
applications, Adv. Colloid Interface Sci. 56 (1995) 201243. [59] DEBRUINBSCWJ, Simulation of geochemical processes during low salinity
[24] P.V. Brady, J.L. Krumhansl, A surface complexation model of water ooding by coupling multiphase buckley-Leverett, Flow to the
oil?brine?sandstone interfaces at 100 C: Low salinity waterooding, J. Geochemical Package PHREEQC. (2012).
Petrol. Sci. Eng. 81 (2012) 171176.
H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639 637

[60] N. Denkov, P. Kralchevsky, Colloid structural forces in thin liquid lms, [93] Y. Gur, I. Ravina, A.J. Babchin, On the electrical double layer theory. II. The
Trends Colloid Interface Sci. IX (1995) (Springer). PoissonBoltzmann equation including hydration forces, J. Colloid Interface
[61] B. Derjaguin, N. Churaev, Structural component of disjoining pressure, J. Sci. 64 (1978) 333341.
Colloid Interface Sci. 49 (1974) 249255. [94] H. Hamaker, The Londonvan der Waals attraction between spherical
[62] B. Derjaguin, N. Churaev, The current state of the theory of long-range particles, Physica 4 (1937) 10581072.
surface forces, Colloids Surf. 41 (1989) 223237. [95] P.G. Hartley, F. Grieser, P. Mulvaney, G.W. Stevens, Surface forces and
[63] B. Derjaguin, L. Landau, Theory of the stability of strongly charged lyophobic deformation at the oil-water interface probed using AFM force
sols and of the adhesion of strongly charged particles in solutions of measurement, Langmuir 15 (1999) 72827289.
electrolytes, Acta Physicochim. URSS 14 (1941) 633662. [96] P.G. Hartley, I. Larson, P.J. Scales, Electrokinetic and direct force
[64] B. Derjaguin, L. Landau, Theory of stability of highly charged liophobic sols measurements between silica and mica surfaces in dilute electrolyte
and adhesion of highly charged particles in solutions of electrolytes, Zhurnal solutions, Langmuir 13 (1997) 22072214.
Eksperimentalnoi I Teoreticheskoi Fiziki 15 (1945) 663682. [97] T. Hassenkam, M.P. Andersson, E. Hilner, J. Matthiesen, S. Dobberschtz, K.N.
[65] M. Dishon, O. Zohar, U. Sivan, From repulsion to attraction and back to Dalby, N. Bovet, S.L. Stipp, P. Salino, C. Reddick, Could atomic-force
repulsion: the effect of NaCl, KCl, and CsCl on the force between silica microscopy force mapping be a fast alternative to core-plug tests for
surfaces in aqueous solution, Langmuir 25 (2009) 28312836. optimizing injection-water salinity for enhanced oil recovery in sandstone?
[66] M. Dishon, O. Zohar, U. Sivan, Effect of cation size and charge on the Spe J. (2015).
interaction between silica surfaces in 1: 1, 2: 1, and 3: 1 aqueous [98] T. Hassenkam, A.C. Mitchell, C.S. Pedersen, L. Skovbjerg, N. Bovet, S.L.S. Stipp,
electrolytes, Langmuir 27 (2011) 1297712984. The low salinity effect observed on sandstone model surfaces, Coll. Surf. A
[67] J. Drelich, Y.U. Wang, Charge heterogeneity of surfaces: mapping and effects 403 (2012) 7986.
on surface forces, Adv. Colloid Interface Sci. 165 (2011) 91101. [99] T. Hassenkam, C.S. Pedersen, K. Dalby, T. Austad, S.L.S. Stipp, Pore scale
[68] J. Drelich, X. Yin, Mapping charge-mosaic surfaces in electrolyte solutions observation of low salinity effects on outcrop and oil reservoir sandstone,
using surface charge microscopy, Appl. Surf. Sci. 256 (2010) 53815387. Coll. Surf. A 390 (2011) 179188.
[69] W. Ducker, R. Pashley, The forces between mica surfaces in ammonium [100] T. Hassenkam, L.L. Skovbjerg, S.L.S. Stipp, Probing the intrinsically oil-wet
chloride solutions, J. Colloid Interface Sci. 131 (1989) 433439. surfaces of pores in North Sea chalk at subpore resolution, Proc. Natl. Acad.
[70] Ducker, W.A., Senden, T.J. & Pashley, R.M., 1991. Direct measurement of Sci. 106 (2009) 60716076.
colloidal forces using an atomic force microscope. [101] E. Hilner, M.P. Andersson, T. Hassenkam, J. Matthiesen, P. Salino, S.L.S. Stipp,
[71] W.A. Ducker, T.J. Senden, R.M. Pashley, Measurement of forces in liquids The effect of ionic strength on oil adhesion in sandstone?the search for the
using a force microscope, Langmuir 8 (1992) 18311836. low salinity mechanism, Sci. Rep. (2015) 5.
[72] W.A. Ducker, Z. Xu, J.N. Israelachvili, Measurements of hydrophobic and [102] A. Hiorth, L. Cathles, M. Madland, The impact of pore water chemistry on
DLVO forces in bubble-surface interactions in aqueous solutions, Langmuir carbonate surface charge and oil wettability, Transport Porous Media 85
10 (1994) 32793289. (2010) 121.
[73] T. Ederth, Template-stripped gold surfaces with 0.4-nm rms roughness [103] G. Hirasaki, Wettability: fundamentals and surface forces, SPE Form. Eval. 6
suitable for force measurements: application to the Casimir force in the (1991) 217226.
20100-nm range, Phys. Rev. A 62 (2000) 062104. [104] R. Horn, D. Clarke, M. Clarkson, Direct measurement of surface forces
[74] T. Ederth, P. Claesson, B. Liedberg, Self-assembled monolayers of between sapphire crystals in aqueous solutions, J. Mater. Res. 3 (1988)
alkanethiolates on thin gold lms as substrates for surface force 413416.
measurements. Long-range hydrophobic interactions and electrostatic [105] R. Horn, D. Smith, W. Haller, Surface forces and viscosity of water measured
double-layer interactions, Langmuir 14 (1998) 47824789. between silica sheets, Chem. Phys. Lett. 162 (1989) 404408.
[75] A.H. Englert, M. Krasowska, D. Fornasiero, J. Ralston, J. Rubio, Interaction [106] R.G. Horn, Surface forces and their action in ceramic materials, J. Am. Ceram.
force between an air bubble and a hydrophilic spherical particle in water: Soc. 73 (1990) 11171135.
measured by the colloid probe technique, Int. J. Miner. Process. 92 (2009) [107] R.G. Horn, D.T. Smith, Measuring surface forces to explore surface
121127. chemistry: mica: sapphire and silica, J. Non-Cryst. Solids 120 (1990) 7281.
[76] U. Farooq, M.T. Tweheyo, J. Sjblom, G. ye, Surface characterization of [108] H. Hosseinzade Khanamiri, I. Baltzersen Enge, M. Nourani, J.. Stensen, O.
model, outcrop, and reservoir samples in low salinity aqueous solutions, J. Torster, N. Hadia, EOR by low salinity water and surfactant at low
Dispers. Sci. Technol. 32 (2011) 519531. concentration: impact of injection and in situ brine composition, Energy
[77] S.J. Fathi, T. Austad, S. Strand, Smart water as a wettability modier in chalk: Fuels 30 (2016) 27052713.
the effect of salinity and ionic composition, Energy Fuels 24 (2010) [109] D.B. Hough, L.R. White, The calculation of Hamaker constants from Liftshitz
25142519. theory with applications to wetting phenomena, Adv. Colloid Interface Sci.
[78] M.L. Fielden, R.A. Hayes, J. Ralston, Surface and capillary forces affecting air 14 (1980) 341.
bubble-particle interactions in aqueous electrolyte, Langmuir 12 (1996) [110] N. Ishida, Direct measurement of hydrophobic particle?bubble interactions
37213727. in aqueous solutions by atomic force microscopy: effect of particle
[79] M.L. Fielden, R.A. Hayes, J. Ralston, Oscillatory and ion-correlation forces hydrophobicity, Coll. Surf. A 300 (2007) 293299.
observed in direct force measurements between silica surfaces in [111] J. Israelachvili, R. Pashley, Measurement of the hydrophobic interaction
concentrated CaCl 2 solutions, Phys. Chem. Chem. Phys. 2 (2000) 26232628. between two hydrophobic surfaces in aqueous electrolyte solutions, J.
[80] R.H. French, Origins and applications of London dispersion forces and Colloid Interface Sci. 98 (1984) 500514.
Hamaker constants in ceramics, J. Am. Ceram. Soc. 83 (2000) 21172146. [112] J. Israelachvili, D. Tabor, The measurement of van der Waals dispersion
[81] M. Giesbers, J.M. Kleijn, M.A.C. Stuart, The electrical double layer on gold forces in the range 1.5130 nm, in: Proceedings of the Royal Society of
probed by electrokinetic and surface force measurements, J. Colloid London A: Mathematical, Physical and Engineering Sciences, 1972, The
Interface Sci. 248 (2002) 8895. Royal Society, 1972, pp. 1938.
[82] F.J. Giessibl, Advances in atomic force microscopy, Rev. Modern Phys. 75 [113] J. Israelachvili, H. Wennerstrm, Role of hydration and water structure in
(2003) 949. biological and, Colloidal Interactions (1996).
[83] G. Gillies, M. Kappl, H.-J. Butt, Direct measurements of particlebubble [114] J.N. Israelachvili, Measurement of forces between surfaces immersed in
interactions, Adv. Colloid Interface Sci. 114 (2005) 165172. electrolyte solutions, Faraday Discuss. Chem. Soc. 65 (1978) 2024.
[84] A. Grabbe, R.G. Horn, Double-layer and hydration forces measured between [115] J.N. Israelachvili, Intermolecular and Surface Forces, revised third edition,
silica sheets subjected to various surface treatments, J. Colloid Interface Sci. Academic press, 2011.
157 (1993) 375383. [116] J.N. Israelachvili, G.E. Adams, Measurement of forces between two mica
[85] D.C. Grahame, Effects of dielectric saturation upon the diffuse double layer surfaces in aqueous electrolyte solutions in the range 0100 nm, J. Chem.
and the free energy of hydration of ions, J. Chem. Phys. 18 (1950) 903909. Soc. Faraday Trans. 1: Phys. Chem. Condensed Phases 74 (1978) 9751001.
[86] D. Grasso, K. Subramaniam, M. Butkus, K. Strevett, J. Bergendahl, A review of [117] Israelachvili, J. N. & Pashley, R. M., 1983. Molecular layering of water at
non-DLVO interactions in environmental colloidal systems, Rev. Environ. surfaces and origin of repulsive hydration forces.
Sci. Biotechnol. 1 (2002) 1738. [118] M. Jaafar, A.M. Nasir, M. Hamid, Measurement of isoelectric point of
[87] J. Gregory, Interaction of unequal double layers at constant charge, J. Colloid sandstone and carbonate rock for monitoring water encroachment, J. Appl.
Interface Sci. 51 (1975) 4451. Sci. 14 (2014) 3349.
[88] J. Gregory, Approximate expressions for retarded van der Waals interaction, [119] Z. Jalili, V.A. Tabrizy, Mechanistic study of the wettability modication in
J. Colloid Interface Sci. 83 (1981) 138145. carbonate and sandstone reservoirs during water/low salinity water
[89] D. Gruen, S. Marcelja, B.A. Pailthorpe, Theory of polarization proles and the ooding, Energy Environ. Res. 4 (2014) p78.
hydration force, Chem. Phys. Lett. 82 (1981) 315320. [120] K. Juhl, N. Bovet, T. Hassenkam, K. Dideriksen, C.S. Pedersen, C.M. Jensen, D.
[90] D.W. Gruen, S. Marcelja, Spatially varying polarization in water. A model for Okhrimenko, S.L.S. Stipp, Change in organic molecule adhesion on
the electric double layer and the hydration force, J. Chem. Soc. Faraday -Alumina (Sapphire) with change in NaCl and CaCl2 solution salinity,
Trans. 2 79 (1983) 225242. Langmuir 30 (2014) 87418750.
[91] A. Gunning, A. Mackie, P. Wilde, V. Morris, Atomic force microscopy of [121] K. Juhl, C.S. Pedersen, N. Bovet, K.N. Dalby, T. Hassenkam, M.P. Andersson, D.
emulsion droplets: probing dropletdroplet interactions, Langmuir 20 Okhrimenko, S.L.S. Stipp, Adhesion of alkane as a functional group on
(2004) 116122. muscovite and quartz: dependence on pH and contact time, Langmuir 30
[92] Y. Gur, I. Ravina, A.J. Babchin, On the electrical double layer theory. I. A (2014) 1447614485.
numerical method for solving a generalized PoissonBoltzmann equation, J.
Colloid Interface Sci. 64 (1978) 326332.
638 H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639

[122] V. Kane, P. Mulvaney, Double-layer interactions between self-assembled low-salinity waterood to improve oil recovery in carbonate reservoirs by
monolayers of -mercaptoundecanoic acid on gold surfaces, Langmuir 14 qualitative coreood. Abu Dhabi International Petroleum Exhibition and
(1998) 33033311. Conference, Soc. Petroleum Eng. (2014).
[123] O. Karoussi, A.A. Hamouda, Imbibition of sulfate and magnesium ions into [154] S.A. Nespolo, D.Y. Chan, F. Grieser, P.G. Hartley, G.W. Stevens, Forces
carbonate rocks at elevated temperatures and their inuence on wettability between a rigid probe particle and a liquid interface: comparison between
alteration and oil recovery, Energy fuels 21 (2007) 21382146. experiment and theory, Langmuir 19 (2003) 21242133.
[124] P. Kkicheff, S. Marcelja, T. Senden, V. Shubin, Charge reversal seen in [155] W.D.W.D. Nesse, Introduction to Mineralogy (2012).
electrical double layer interaction of surfaces immersed in 2: 1 calcium [156] A. Nguyen, J. Nalaskowski, J.D. Miller, A study of bubbleparticle interaction
electrolyte, J. Chem. Phys. 99 (1993) 60986113. using atomic force microscopy, Miner. Eng. 16 (2003) 11731181.
[125] K.C. Khilar, H.S. Fogler, Colloidally induced nes migration in porous media, [157] B. Ninham, On progress in forces since the DLVO theory, Adv. Colloid
Rev. Chem. Eng. 4 (1987) 41108. Interface Sci. 83 (1999) 117.
[126] S. Kia, H.S. Fogler, M. Reed, Effect of salt composition on clay release in Berea [158] B.W. Ninham, V.A. Parsegian, Electrostatic potential between surfaces
sandstones. SPE International Symposium on Oileld Chemistry, 1987, Soc. bearing ionizable groups in ionic equilibrium with physiologic saline
Petroleum Eng. (1987). solution, J. Theor. Biol. 31 (1971) 405428.
[127] R. Kjellander, S. Marcelja, Double-layer interaction in the primitive model [159] B.W. Ninham, V. Yaminsky, Ion binding and ion specicity: the Hofmeister
and the corresponding Poisson-Boltzmann description, J. Phys. Chem. 90 effect and Onsager and Lifshitz theories, Langmuir 13 (1997) 20972108.
(1986) 12301232. [160] M. Nourani, T. Tichelkamp, B. Gawe, G. ye, Method for determining the
[128] R. Kjellander, S. Marcelja, Inhomogeneous Coulomb uids with image amount of crude oil desorbed from silica and aluminosilica surfaces upon
interactions between planar surfaces. I, J. Chem. Phys. 82 (1985) 21222135. exposure to combined low-salinity water and surfactant solutions, Energy
[129] R. Kjellander, S. Marcelja, Interaction of charged surfaces in electrolyte Fuels 28 (2014) 18841889.
solutions, Chem. Phys. Lett. 127 (1986) 402407. [161] H. Ohshima, Theory Of Colloid And Interfacial Electric Phenomena,
[130] R. Kjellander, S. Marcelja, R. Pashley, J. Quirk, Double-layer ion correlation Academic Press, 2006.
forces restrict calcium-clay swelling, J. Phys. Chem. 92 (1988) 64896492. [162] H. Ohshima, T. Kondo, Electrostatic repulsion of ion penetrable charged
[131] R. Kjellander, S. Marcelja, R. Pashley, J. Quirk, A theoretical and experimental membranes: role of Donnan potential, J. Theor. Biol. 128 (1987) 187194.
study of forces between charged mica surfaces in aqueous CaCl2 solutions, J. [163] H. Ohshima, T. Kondo, Comparison of three models on double layer
Chem. Phys. 92 (1990) 43994407. interaction, J. Colloid Interface Sci. 126 (1988) 382383.
[132] R. Kjellander, D.J. Mitchell, An exact but linear and PoissonBoltzmann-like [164] T. Oncsik, G. Trefalt, M. Borkovec, I. Szilagyi, Specic ion effects on particle
theory for electrolytes and colloid dispersions in the primitive model, Chem. aggregation induced by monovalent salts within the Hofmeister series,
Phys. Lett. 200 (1992) 7682. Langmuir 31 (2015) 37993807.
[133] E. Klaseboer, J.P. Chevaillier, C. Gourdon, O. Masbernat, Film drainage [165] C.W. Outhwaite, L.B. Bhuiyan, S. Levine, Theory of the electric double layer
between colliding drops at constant approach velocity: experiments and using a modied poissonboltzman equation, J. Chem. Soc., Faraday Trans. 2
modeling, J. Colloid Interface Sci. 229 (2000) 274285. 76 (1980) 13881408.
[134] A. Lager, K. Webb, C. Black, M. Singleton, K. Sorbie, Low salinity oil [166] J.T.G. Overbeek, M. Sparnaay, Classical coagulation. London-van der Waals
recovery-An experimental investigation1, Petrophysics (2008) 49. attraction between macroscopic objects, Discuss. Faraday Soc. 18 (1954)
[135] S.Y. Lee, K.J. Webb, I. Collins, A. Lager, S. Clarke, M. Osullivan, A. Routh, X. 1224.
Wang, Low salinity oil recovery: increasing understanding of the underlying [167] V. Parsegian, T. ZemB, Hydration forces: observations, explanations,
mechanisms. SPE Improved Oil Recovery Symposium, Soc. Petroleum Eng. expectations, questions, Curr. Opin. Colloid Interface Sci. 16 (2011) 618624.
(2010). [168] V.A. Parsegian, B.W. Ninham, Van der Waals forces in many-layered
[136] P. Levine, The solution of a modied PoissonBoltzmann equation for structures: generalizations of the Lifshitz result for two semi-innite media,
colloidal particles in electrolyte solutions, J. Colloid Interface Sci. 51 (1975) J. Theor. Biol. 38 (1973) 101109.
7286. [169] R. Pashley, DLVO and hydration forces between mica surfaces in Li+, Na+, K+,
[137] E. Lifshitz, The theory of molecular attractive forces, Between Solids. (1956). and Cs+ electrolyte solutions: a correlation of double-layer and hydration
[138] D.J. Ligthelm, J. Gronsveld, J. Hofman, N. Brussee, F. Marcelis, H. Van Der forces with surface cation exchange properties, J. Colloid Interface Sci. 83
Linde, Novel waterooding strategy by manipulation of injection brine (1981) 531546.
composition. EUROPEC/EAGE conference and exhibition, Soc. Petroleum [170] R. Pashley, Hydration forces between mica surfaces in electrolyte solutions,
Eng. (2009). Adv. Colloid Interface Sci. 16 (1982) 5762.
[139] H. Mahani, A.L. Keya, S. Berg, W.-B. Bartels, R. Nasralla, W.R. Rossen, Insights [171] R. Pashley, Forces between mica surfaces in La 3+ and Cr 3+ electrolyte
into the mechanism of wettability alteration by low-salinity ooding (LSF) solutions, J. Colloid Interface Sci. 102 (1984) 2335.
in carbonates, Energy Fuels 29 (2015) 13521367. [172] R. Pashley, J. Israelachvili, DLVO and hydration forces between mica surfaces
[140] J. Mahanty, B. Ninham, Dispersion forces academic, London,. 16 JN in Mg 2+, Ca 2+, Sr 2+, and Ba 2+ chloride solutions, J. Colloid Interface Sci.
israelachvili, Intermol. Surf. Forces (1991) 1991. 97 (1984) 446455.
[141] S. Marcelja, N. Radic, Repulsion of interfaces due to boundary water, Chem. [173] R. Pashley, J. Quirk, The effect of cation valency on DLVO and hydration
Phys. Lett. 42 (1976) 129130. forces between macroscopic sheets of muscovite mica in relation to clay
[142] J. Matthiesen, N. Bovet, E. Hilner, M.P. Andersson, D. Schmidt, K. Webb, K.N. swelling, Colloids Surf. 9 (1984) 117.
Dalby, T. Hassenkam, J. Crouch, I. Collins, How naturally adsorbed material [174] R.M. Pashley, J.N. Israelachvili, Molecular layering of water in thin lms
on minerals affects low salinity enhanced oil recovery, Energy & Fuels 28 between mica surfaces and its relation to hydration forces, J. Colloid
(2014) 48494858. Interface Sci. 101 (1984) 511523.
[143] P. Mcguiggan, R. Pashley, Molecular layering in thin aqueous lms, J. Phys. [175] G. Peschel, P. Belouschek, M. Mller, M. Mller, R. Knig, The interaction of
Chem. 92 (1988) 12351239. solid surfaces in aqueous systems, Colloid. Polym. Sci. 260 (1982) 444451.
[144] P. Mcguire, J. Chatham, F. Paskvan, D. Sommer, F. Carini, Low salinity oil [176] M. Preuss, H.-J. Butt, Direct measurement of particle-bubble interactions in
recovery: an exciting new EOR opportunity for alaskas north slope. SPE aqueous electrolyte: dependence on surfactant, Langmuir 14 (1998)
western regional meeting , 2005, Soc. Petroleum Eng. (2005). 31643174.
[145] L. Meagher, Direct measurement of forces between silica surfaces in [177] M. Preuss, H.-J. Butt, Direct measurement of forces between particles and
aqueous CaCl 2 solutions using an atomic force microscope, J. Colloid bubbles, Int. J. Miner. Process. 56 (1999) 99115.
Interface Sci. 152 (1992) 293295. [178] C. Qiao, R. Johns, L. Li, J. Xu, Modeling low salinity waterooding in
[146] K.K. Mohanty, S. Chandrasekhar, Wettability alteration with brine mineralogically different carbonates. SPE annual technical conference and
composition in high temperature carbonate reservoirs. SPE Annual exhibition, Soc. Petroleum Eng. (2015).
Technical Conference and Exhibition, Soc. Petroleum Eng. (2013). [179] C. Qiao, L. Li, R.T. Johns, J. Xu, A mechanistic model for wettability alteration
[147] J. Molina-Bolivar, J. Ortega-Vinuesa, How proteins stabilize colloidal by chemically tuned waterooding in carbonate reservoirs, SPE J. (2015).
particles by means of hydration forces, Langmuir 15 (1999) 26442653. [180] Y.I. Rabinovich, B. Derjaguin, Interaction of hydrophobized laments in
[148] F. Mugele, B. Bera, A. Cavalli, I. Siretanu, A. Maestro, M. Duits, M. aqueous electrolyte solutions, Colloids Surf. 30 (1988) 243251.
Cohen-Stuart, D. Van Den Ende, I. Stocker, I. Collins, Ion adsorption-induced [181] Y.I. Rabinovich, B. Derjaguin, N. Churaev, Direct measurements of
wetting transition in oil-water-mineral systems, Sci. Rep. 5 (2015). long-range surface forces in gas and liquid media, Adv. Colloid Interface Sci.
[149] A. Muggeridge, A. Cockin, K. Webb, H. Frampton, I. Collins, T. Moulds, P. 16 (1982) 6378.
Salino, Recovery rates, enhanced oil recovery and technological limits, Phil. [182] Y.I. Rabinovich, R.-H. Yoon, Use of atomic force microscope for the
Trans. R. Soc. A 372 (2014) 20120320. measurements of hydrophobic forces between silanated silica plate and
[150] P. Mulvaney, J. Perera, S. Biggs, F. Grieser, G. Stevens, The direct glass sphere, Langmuir 10 (1994) 19031909.
measurement of the forces of interaction between a colloid particle and an [183] R. Raiteri, M. Preuss, M. Grattarola, H.-J. Butt, Preliminary results on the
oil droplet, J. Colloid Interface Sci. 183 (1996) 614616. electrostatic double-layer force between two surfaces with high surface
[151] P.C. Myint, A. Firoozabadi, Thin liquid lms in improved oil recovery from potentials, Coll. Surf. A 136 (1998) 191197.
low-salinity brine, Curr. Opin. Colloid Interface Sci. 20 (2015) 105114. [184] R.J. Reeder, D.J. Barber, Carbonates: Mineralogy and Chemistry,
[152] R.A. Nasralla, M.A. Bataweel, H.A. Nasr-El-Din, Investigation of wettability Mineralogical Society of America, Washington DC, 1983.
alteration and oil-recovery improvement by low-salinity water in sandstone [185] A. Rezaeidoust, T. Puntervold, S. Strand, T. Austad, Smart water as wettability
rock, J. Can. Pet. Technol. 52 (2013) 144154. modier in carbonate and sandstone: a discussion of similarities/differences
[153] R.A. Nasralla, E. Sergienko, S.K. Masalmeh, H.A. Van Der Linde, N.J. Brussee, in the chemical mechanisms, Energy Fuels 23 (2009) 44794485.
H. Mahani, B. Suijkerbuijk, I. Alqarshubi, Demonstrating the potential of
H. Ding, S. Rahman / Colloids and Surfaces A: Physicochem. Eng. Aspects 520 (2017) 622639 639

[186] G.C.J. Rouweler, J.T.G. Overbeek, Dispersion forces between fused silica [204] T. Underwood, V. Erastova, P. Cubillas, H.C. Greenwell, Molecular dynamic
objects at distances between 25 and 350 nm, Trans. Faraday Soc. 67 (1971) simulations of Montmorillonite?Organic interactions under varying salinity:
21172121. an insight into enhanced oil recovery, J. Phys. Chem. C 119 (2015)
[187] J. Schembre, A. Kovscek, Mechanism of formation damage at elevated 72827294.
temperature, J. Energy Res. Technol. 127 (2005) 171180. [205] J. Valle-Delgado, J. Molina-Bolivar, F. Galisteo-Gonzalez, M. Galvez-Ruiz, A.
[188] S. Shaddel, S. Tabatabae-Nejad, Alkali/Surfactant improved low-Salinity Feiler, M.W. Rutland, Hydration forces between silica surfaces:
waterooding, Transp. Porous Media 106 (2015) 621642. experimental data and predictions from different theories, J. Chem. Phys.
[189] S.F. Shariatpanahi, S. Strand, T. Austad, Initial wetting properties of 123 (2005) 034708.
carbonate oil reservoirs: effect of the temperature and presence of sulfate in [206] C. Van Oss, R. Giese, P.M. Costanzo, DLVO and non-DLVO interactions in
formation water, Energy Fuels 25 (2011) 30213028. hectorite, Clays Clay Miner. 38 (1990) 151159.
[190] J. ShENG, Critical review of low-salinity waterooding, J. Petrol. Sci. Eng. 120 [207] E. Verwey, J.T.G. Overbeek, Theory of Stability of Lyophobic Solids, Elsevier
(2014) 216224. Amsterdam, 1948.
[191] V.E. Shubin, P. Kkicheff, Electrical double layer structure revisited via a [208] G. VIGIL, Z. XU, S. Steinberg, J. Israelachvili, Interactions of silica surfaces, J.
surface force apparatus: mica interfaces in lithium nitrate solutions, J. Colloid Interface Sci. 165 (1994) 367385.
Colloid Interface Sci. 155 (1993) 108123. [209] J.Y. Walz, The effect of surface heterogeneities on colloidal forces, Adv.
[192] L. Skovbjerg, T. Hassenkam, E. Makovicky, C.P. Hem, M. Yang, N. Bovet, S.L.S. Colloid Interface Sci. 74 (1998) 119168.
Stipp, Nano sized clay detected on chalk particle surfaces, Geochim. [210] G.B. Webber, R. Manica, S.A. Edwards, S.L. Carnie, G.W. Stevens, F. Grieser,
Cosmochim. Acta 99 (2012) 5770. R.R. Dagastine, D.Y. Chan, Dynamic forces between a moving particle and a
[193] B.A. Snyder, D.E. Aston, J.C. Berg, Particle-drop interactions examined with deformable drop, J. Phys. Chem. C 112 (2008) 567574.
an atomic force microscope, Langmuir 13 (1997) 590593. [211] M. Wolthers, L. Charlet, P. Van Cappellen, The surface chemistry of divalent
[194] M. Sparnaay, Corrections of the theory of the at diffuse double layer, Recl. metal carbonate minerals; a critical assessment of surface charge and
Trav. Chim. Pays-Bas 77 (1958) 872888. potential data using the charge distribution multi-site ion complexation
[195] D.C. Standnes, T. Austad, Wettability alteration in chalk: 1. Preparation of model, Am. J. Sci. 308 (2008) 905941.
core material and oil properties, J. Petrol. Sci. Eng. 28 (2000) 111121. [212] Q. Xie, Y. Liu, J. Wu, Q. Liu, Ions tuning water ooding experiments and
[196] D. Tabor, R. Winterton, The direct measurement of normal and retarded van interpretation by thermodynamics of wettability, J. Petrol. Sci. Eng. 124
der Waals forces Proceedings of the Royal Society of London A: (2014) 350358.
Mathematical, Physical and Engineering Sciences, R. Soc. (1969) 435450. [213] D.E. Yates, S. Levine, T.W. Healy, Site-binding model of the electrical double
[197] K. Takamura, R.S. Chow, A mechanism for initiation of bitumen layer at the oxide/water interface, Journal of the Chemical Society, Faraday
displacement from oil sand, J. Can. Pet. Technol. 22 (1983). Transactions 1: Physical Chemistry in Condensed Phases 70 (1974)
[198] S. Tchaliovska, P. Herder, R. Pugh, P. Stenius, J.C. Eriksson, Studies of the 18071818.
contact interaction between an air bubble and a mica surface submerged in [214] R.-H. Yoon, D.H. Flinn, Y.I. Rabinovich, Hydrophobic interactions between
dodecylammonium chloride solution, Langmuir 6 (1990) 15351543. dissimilar surfaces, J. Colloid Interface Sci. 185 (1997) 363370.
[199] T.K. Tokunaga, DLVO-based estimates of adsorbed water lm thicknesses in [215] H. Yotsumoto, R.-H. Yoon, Application of extended DLVO theory: i. Stability
geologic CO2 reservoirs, Langmuir 28 (2012) 80018009. of rutile suspensions, J. Colloid Interface Sci. 157 (1993) 426433.
[200] C. Toprakcioglu, J. Klein, P.F. Luckham, Forces between mica surfaces in [216] A. Zahid, A.A. Shapiro, A. Skauge, Experimental studies of low salinity water
aqueous KNO 3 solution in the range 10-4-10-1 mol dm-3, showing ooding carbonate: a new promising approach, SPE EOR Conference at Oil
long-range attraction at high electrolyte concentration, J. Chem. Soc. and Gas West Asia,. Society of Petroleum Engineers (2012).
Faraday Trans. Phys. Chem. Condensed Phases 83 (1987) 17031709. [217] C. Zhang, Hydration force, Encyclopedia Of Tribology (2013) (Springer).
[201] Y.-H. Tsao, D.F. Evans, H. Wennerstrom, Long-range attractive force between [218] P. Zhang, M.T. Tweheyo, T. Austad, Wettability alteration and improved oil
hydrophobic surfaces observed by atomic force microscopy, Science 262 recovery by spontaneous imbibition of seawater into chalk: impact of the
(1993) 547550. potential determining ions Ca 2+, Mg 2+, and SO 4 2-, Coll. Surf. A 301 (2007)
[202] Y.H. Tsao, D.F. Evans, H. Wennerstroem, Long-range attraction between a 199208.
hydrophobic surface and a polar surface is stronger than that between two
hydrophobic surfaces, Langmuir 9 (1993) 779785.
[203] Y.H. Tsao, S. Yang, D.F. Evans, H. Wennerstroem, Interactions between
hydrophobic surfaces. Dependence on temperature and alkyl chain length,
Langmuir 7 (1991) 31543159.

Você também pode gostar