You are on page 1of 94

Accepted Manuscript

Review

Demulsification techniques of water-in-oil and oil-in-water emulsions in petro-


leum industry

Reza Zolfaghari, Ahmadun Fakhru’l-Razi, Luqman C. Abdullah, Said S.E.H.


Elnashaie, Alireza Pendashteh

PII: S1383-5866(16)30719-5
DOI: http://dx.doi.org/10.1016/j.seppur.2016.06.026
Reference: SEPPUR 13082

To appear in: Separation and Purification Technology

Received Date: 4 October 2015


Revised Date: 11 June 2016
Accepted Date: 13 June 2016

Please cite this article as: R. Zolfaghari, A. Fakhru’l-Razi, L.C. Abdullah, S.S.E. Elnashaie, A. Pendashteh,
Demulsification techniques of water-in-oil and oil-in-water emulsions in petroleum industry, Separation and
Purification Technology (2016), doi: http://dx.doi.org/10.1016/j.seppur.2016.06.026

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Demulsification techniques of water-in-oil and oil-in-water emulsions in
petroleum industry
a,*
Reza Zolfaghari , Ahmadun Fakhru’l-Razi a, Luqman C. Abdullah a, Said S.E.H.
Elnashaie b, Alireza Pendashteh c

a
Department of Chemical and Environmental Engineering, Faculty of Engineering,
Universiti Putra Malaysia (UPM), Serdang, Malaysia

b
Department of Chemical and Biological Engineering, Faculty of Applied Science, The
University of British Columbia (UBC), Vancouver, Canada

c
The Caspian Sea Basin Research Center, University of Guilan, Rasht, Iran

*
Corresponding author. Tel.: +60 17 694 5846

Email address: petro.eng27@gmail.com (Reza Zolfaghari)

Abbreviations1

1
EOR: enhanced oil recovery; SP: surfactant-polymer; ASP: alkaline-surfactant-polymer; IFT: interfacial
tension; HLB: hydrophilic-lipophilic balance; PEO: poly(ethylene oxide); PPO: poly(propylene oxide); PIT:
phase inversion temperature; HLD: hydrophilic-lipophilic deviation; SAD: surfactant affinity difference; WOR:
water-to-oil volumetric ratio; CMC: critical micellization concentration; PDMS: poly(dimethyl)siloxane; IL:
ionic liquid; CMW: chloroform-methanol-water; TMP: transmembrane pressure; UF: ultrafiltration; MF:
microfiltration; RO: reverse osmosis; TOC: total organic carbon; TSC: total surface charge; COD: chemical
oxygen demand; CFR: cross-flow rate; AC: alternating current; DC: direct current
2

Abstract

The difficulties associated with transportation and refining of crude oil emulsions as well as
produced water discharge limitations are among the conspicuous clues that have led the
oilfield researchers to probe into practical demulsification methods for many decades.
Inconsistent research outcomes observed in the literature for a particular demulsification
method of a typical emulsion (i.e., water-in-oil or oil-in-water) arise not only from the varied
influential parameters associated (such as salinity, temperature, pH, dispersed phase content,
emulsifier/demulsifier concentration, droplet size, etc.), but also from the diverse types of
emulsion constituents (namely oil, surfactant, salt, alkali, polymer, fine solids, and/or other
chemicals/impurities). Being the main component in formation of stabilizing interfacial film
surrounding the dispersed phase droplets, surfactant is the most predominant contributor to
emulsion stability, extent of which depends on its nature (being ionic or nonionic, and its
degree of hydrophilicity/lipophilicity), concentration and interaction with other surface-active
agents in the emulsion, as well as on the salinity, temperature, and pH of the system. In this
paper, it is endeavored to overview some of the most commonly exploited demulsification
techniques (i.e., chemical, biological, membrane, electrical, and microwave irradiation) on
both the oilfield and synthetic emulsions, taking into account the emulsion-stabilizing and -
destabilizing effects with regard to the dominant parameters plus the emulsion composition.
Further, the variations occurring in interfacial properties of emulsions by demulsification
process are discussed. Finally, the mechanism(s) involved in emulsions resolution achieved
by each method is elucidated. Clearly, the most efficient demulsification approach is the one
able to attain desirable separation efficiency while complying with the environmental
regulations and imposing the least economic burden on the petroleum industry.

Keywords: Water-in-oil; Oil-in-water; Separation efficiency; Emulsion stability;


Demulsification mechanism; Surfactants
3

Contents

1. Introduction
1.1. Water-in-oil (W/O) emulsions
1.2. Oil-in-water (O/W) emulsions
2. Chemical demulsification
2.1. Hydrophilic-lipophilic balance
2.1.1. Relative solubility number
2.1.2. Effect of position isomerism on HLB
2.2. R ratio
2.3. Phase inversion temperature
2.4. Hydrophilic-lipophilic deviation
2.5. Influential parameters
2.5.1. Salinity
2.5.2. Temperature
2.5.3. pH
2.5.4. Water content
2.5.5. Oil content
2.5.6. Agitation speed and duration
2.5.7. Emulsifier/demulsifier concentration
2.5.8. Droplet size and distribution
2.6. Demulsifying materials
2.6.1. Nonionic surfactants
2.6.1.1.Molecular structure
2.6.1.2.Critical micellization concentration
2.6.1.3.Steric effects
2.6.1.4.Surface pressure-molecular area
2.6.2. Ionic surfactants
2.6.3. Ionic liquids
2.7. Oil-water interfacial properties
2.8. Zeta potential
2.9. Demulsifier partitioning
2.10. Aging of W/O emulsions
2.11. Demulsification mechanisms
2.11.1. Nonionic demulsifiers
2.11.2. Ionic demulsifiers
2.11.3. Conclusion
3. Biological demulsification
3.1. Cell concentration
3.2. Chemical treatment
3.3. Substrate
3.4. Nitrogen source
3.5. pH
3.6. Thermal treatment
4

3.7. Culture age


3.8. Mechanism
4. Membrane demulsification
4.1. Separation efficiency
4.2. Permeate flux
4.3. Membrane fouling
4.4. Membrane cleaning
4.5. Mechanism
5. Electrical demulsification
5.1. Influential features
5.2. Mechanism
6. Microwave irradiation
6.1. Effectual factors
6.2. Lateral and radial irradiations
6.3. Mechanism
7. Summary

References
5

1. Introduction

Demulsification is an intricate process as oilfield emulsions are, in principal, complex, stable


liquid-liquid colloidal suspensions comprising of dispersed/internal phase,
continuous/external phase, and emulsifying agents basically present at oil-water interface. An
effectual demulsifying route must thus be capable of annulling/minimizing the stability of
targeted emulsion, leading to separation of immiscible phases. Demulsification techniques are
generally classified into three categories, i.e., chemical, biological, and physical treatments.
Physical demulsification includes gravitational settling, centrifugation, pH adjustment,
thermal treatment (conventional heating, microwave irradiation, and freeze/thaw), flotation,
filtration (adsorption and coalescing filters), electrical demulsification (electro-coalescence),
membrane separation, ultrasonic, inertial, orthokinetic (shear flow), etc. However, physical
treatments of emulsions have often been employed in conjunction with other (physical,
chemical, or biological) separation methods to establish hybrid systems in order to improve
the destabilization of emulsions up to satisfactory levels [1-5].

In this paper, it is attempted to overview the most commonly employed demulsification


methods (i.e., chemical, biological, membrane, electrical, and microwave irradiation) by
discussing the effective factors involved in contribution to improving their efficiency for
separation of emulsions. The rest of demulsification techniques mentioned above are not
included in this article.

At first stance, it is useful to introduce the main three phases of crude oil extraction, namely,
primary, secondary, and tertiary oil recovery processes [6,7]: in primary oil recovery, the
reservoir’s natural energy sources, including fluid and rock expansion, solution gas, water
influx, gas cap, and gravity drainage, produce oil; artificial lift devices like pumps are usually
used to aid in oil extraction. However, only about 10% of the reservoir’s original oil in place
is extracted during primary recovery. The extraction efficiency can be boosted to 20-40% of
the original oil in place during secondary recovery phase in which water and/or gas are
injected into the well to displace and push the oil to the ground (production well). To further
improve the crude oil production yield, several tertiary, or enhanced oil recovery (EOR),
processes have been proposed resulting in an ultimate extraction of 30-60% of the reservoir’s
original oil. Three major EOR techniques which have been recognized commercially thriving
to varying extents are briefly described here:
6

i. Thermal recovery, which entails the introduction of heat through steam injection into the
well so as to decrease the viscosity of oil and improve its flow rate through the reservoir.
ii. Gas injection, which utilizes natural gas, nitrogen, or carbon dioxide (CO2) to drive
additional oil to the ground, or employs other gases dissolvable in the oil to reduce its
viscosity and enhance its flow rate. The CO2-EOR technique has recently attracted the
most market interest in the United States.
iii. Chemical injection (or chemical flooding), which involves the injection of chemical(s)
such as surfactant, alkali, and/or polymer into the oil well. Surfactant improves the
microscopic displacement efficiency by lowering the interfacial tension (IFT) between oil
and water which inhibits the flow of oil droplets through the reservoir rock. Alkali reacts
with organic acids in crude oil and generates some in situ surfactant which reduces the
IFT as well. Besides, alkali lessens the surfactant adsorption on the reservoir rock.
Polymer is added as the mobility control agent to diminish the water-to-oil mobility ratio
by increasing the viscosity of injection water, thus improving the volumetric sweep
efficiency. There might be synergistic effects between either of these materials when
injected together. For example, both surfactant and alkali synergistically reduce the IFT,
and they may change the rock wettability toward more hydrophilicity. In modern
chemical EOR, the main goals are to reduce the amount of injected chemicals and to
thoroughly investigate the synergy of varied processes involved.

Subsequently, generation of emulsions in oil extraction process of petroleum industry is due


to the following reasons [8,9]:

- Flow through porous media such as reservoir pores


- Shearing within choke valves
- Pressure drop arising from pumping process in oil wells
- Deliberate injection of water into the oil well

And the post-extraction emulsions which are formed intentionally are as a result of:

- Addition of water to heavy or extra-heavy (super-heavy) crude oil for the ease of
transportation through pipelines (heavy and extra-heavy crude oils do not flow easily at
room temperature due to their high viscosity, and they have larger density or specific
gravity than light crude oil which has low viscosity. Heavy crude oil has an API gravity
less than 20º, whereas the API gravity of extra-heavy oil is below 10º. Light crude oil is
more expensive than its heavy counterpart in merchandise market because of production
7

of higher quantities of gasoline and diesel fuels after it is converted to products by the oil
refinery [10])
- Introduction of water into crude oil at desalting unit

A variety of difficulties may arise when the water is coproduced with crude oil, including
(but is not confined to) [11]:

 The cost of pumping or transporting the water through pipeline or by tanker, respectively
 The corrosion of pumps, pipes, production facilities, and downstream distillation columns
(due to the existence of chloride ions in aqueous phase)
 The necessity for installation of supplementary equipment to produce export-quality
crude oil
 The poisoning of downstream refinery catalysts, and
 The problems pertaining to improved crude oil viscosity due to small dispersed water
drops in the crude oil.

There are mainly two types of emulsions with which petroleum industries deal frequently,
namely water-in-oil emulsions (where the dispersed phase is water, and the continuous phase
is oil) and oil-in-water emulsions (where the dispersed phase is oil, and the continuous phase
is water) which are the main focus of this report and are discussed below. Although these
types of emulsions are thermodynamically unstable (unlike microemulsions), they are
kinetically stable and may remain unchanged for a long period (occasionally for decades)
[12]. However, water-in-oil-in-water (W/O/W) multiple emulsions (ternary emulsions),
where the inner aqueous phase is the dispersed phase of oil phase and the outer aqueous
phase is the continuous phase of oil droplets, (as well as oil-in-water-in-oil (O/W/O)
emulsions) have also potential applications in food, cosmetics, pharmaceutics, and
wastewater treatment [13,14].

1.1. Water-in-oil (W/O) emulsions:

Crude oil is nearly always generated as importunate W/O emulsions that must be broken into
two separate phases before the crude oil can be allowed for transportation or sent to refinery
[15,16]. Oily sludge waste is also a sophisticated W/O emulsion produced during crude oil
conveyance, storage, and refining process, containing 30-50 wt.% oil, 30-50 wt.% water, and
10-12 wt.% solid content, which is mainly stabilized by fine solids adsorbed at the droplet
surface and hamper the drops coalescence, and is deemed hazardous to surroundings as well
8

as human health because of its high concentration of petroleum hydrocarbons, which thus
makes it essential to undergo the oil recovery process prior to the sludge disposal [2,17,18].
Oily sludge treatment pathways have been reviewed elsewhere [17]. Density, viscosity,
chemical composition, and water content of crude oil, and hence the extent of its stability
may vary remarkably from field to field and even within the same production well at different
times [19,20].

Surface-active compounds typically found in the crude oil, including asphaltenes, resins,
waxes, inorganic solids, and oil-soluble organic acids, are among the substances composing
the stabilizing interfacial film which shields the water droplets from coalescence and is
basically developed by hydrogen bonding of the N-, O-, and S-containing groups present in
natural crude as well as Si-OH- and Si-O- in dispersed minerals at the oil-water interface
[11,15,21-23]. However, asphaltenes and resins play the most striking role in the stabilization
of W/O emulsions. Asphaltenes are solvable in toluene, but insolvable in alkanes, such as n-
pentane or n-heptane, while resins are dissolvable in both aliphatic and aromatic solvents.
Both asphaltenes and resins are surface-active materials and have the potential to build up on
the oil-water interface due to possession of hydrophilic functional groups [24]. Rigid
interfacial films principally contain asphaltenic species (and this rigidity may be promoted by
co-adsorption of resins at the interface), whereas weak films predominantly consist of
resinous materials and would form unstable W/O emulsions. In contrast to weak or mobile
interfacial films, rigid or immobile interfacial films are stable and not readily drained out, and
an appropriate demulsification technique has to be deployed to weaken, mobilize, and drain
them; thereby the droplets coalesce once they get closer to one another. The elasticity and
perfection of interfacial films, and consequently the stability of W/O emulsions are
susceptible to several factors related to crude oil, including aromatic degree, resin-to-
asphaltene ratio, concentration of polar functional groups, and the dispersion state of
asphaltenes and resins (molecular vs. colloidal) [16,21,25]. At high resin-to-asphaltene ratios,
resin molecules may even act as a destabilizing agent by disruption of π-π and polar bonding
interactions between asphaltene monomers [26-28]. When the hydrocarbon phase is other
than crude oil in synthetic or model W/O emulsions, oil-soluble nonionic surfactants are
normally introduced for emulsion stabilization, which functionally correspond to asphaltenic
species in crude oil emulsions. Examples of oil-soluble nonionic surfactants are alcohol
ethoxylates, alkylphenols, and alkylolamides, with an average hydrocarbon radical length of
C12 – C16, the chemical compounds comprising up to 4 moles of ethylene-oxide (EO) [29].
9

Spans (i.e., sorbitan esters) are the nonionics which are broadly used as lipophilic emulsifiers,
and confer grander lipophilicity by increasing the esterification degree. Conspicuously,
oilfield emulsions are further recalcitrant against demulsification than model emulsions,
owing to the presence of supplementary chemicals (such as scale inhibitor, corrosion
inhibitor, biocide, H2S scavengers, clay stabilizer, etc. [30,31]) added into the oil well during
the extraction operation.

1.2. Oil-in-water (O/W) emulsions

As it is brought up to the wellhead, the crude oil is accompanied by produced water or oily
wastewater, which is the hugest waste stream produced in the oil and gas industries and
contains a combination of organic and inorganic compounds, discharge of which may
contaminate surface and underground water as well as soil [32]. Produced water is an O/W
emulsion type.

Because of the synergistic impact and manifold functions of mixed chemicals in chemical
EOR, such as alkaline-surfactant, surfactant-polymer (SP), or alkaline-surfactant-polymer
(ASP), chemical-combination flooding yields higher oil recovery than single-component
chemical flooding (such as alkaline flooding, surfactant flooding, or polymer flooding), and
thus has been perceived as a promising pathway for EOR targets [33,34]. Consequently, it is
expected that the corresponding emulsions are more stable than those generated from single-
chemical flooding.

Formation of solids-stabilized O/W emulsions, known as O/W “Pickering emulsions,” as a


result of the presence of finely divided solids has also been one of the major concerns for
economic and environmental reasons in the petroleum and allied industries, which can be
addressed by minimizing the oil content in the water phase through breakup of the solids-
stabilized oil globules. Both the size and the contact angle of fine solids influence on the
related emulsion stability. The particle size is important in the stabilization of emulsions
because it affects the equilibrium position of a particle at an oil-water interface. Particles
must be much smaller than the dispersed drops to create sufficient stability [35-37]. The
contact angle of solid particles is indicative of their wettability by either oil phase or water
phase or both. The philicity and phobicity of both water and oil phases interact in complex
ways. The two phases compete with each other at the solid surface, and the relative philicity
or phobicity tends to control the emulsion stability. For the water drops submerged in an oil
phase, it has been discovered that hydrophilic solids with a contact angle somewhat lower
10

than 90˚ stabilize O/W emulsions, whereas hydrophobic solids with a contact angle slightly
larger than 90˚ stabilize W/O emulsions, and the highest emulsion stability is induced at the
contact angle of 90˚; if the fine solids are thoroughly wetted by either water or oil, they
remain dispersed in the respective phase, and thus no stable emulsion is formed [38,39].

Discharging the oily wastewater or its re-injection to the extraction process in either case
necessitates its treatment for the separation of oil phase, particularly when most stable
emulsions such as those produced from chemical EOR are handled, and also when notable,
bulky quantity of dispersed oil droplets exists. After the breakup of emulsions, the oil phase
obtained can be sent to refinery. Fakhru’l-Razi et al. [32] have lately reviewed different
technologies of oil and gas produced water treatment.

Synthetic or model O/W emulsions are emulsified by a nonionic surfactant, anionic


surfactant, or mixture of both [40]. The nonionics are the same as those aforementioned in
Section 1.1, but with different number of EO (containing over than 10 EO moles) marking
them as water-soluble surfactants [29]. Distinctive water-soluble anionic surfactants include,
depending on the sort of their polar group, alkali metal salts having hydrophobic tail of the
length of C12-C18. Magnesium salts of alkyl ethoxysulphates are an example of partly soluble
anionics [29]. Tweens (i.e., ethoxylated sorbitan esters) are the nonionic surfactants which are
hydrophilic in nature and immensely applied for emulsion preparation, and their solubility in
aqueous solutions is promoted with the degree of ethoxylation. They may be used alone or in
combination with Spans (i.e., sorbitan esters), depending on the aimed application.

2. Chemical demulsification

Stable emulsions are formed only when emulsifying materials are present. Immiscible liquids
can be separated by omission, variation, or neutralization of the emulsifying agents. In this
respect, appropriate chemicals with demulsifying characteristics particular to the crude oil to
be treated are added to deliver rapid, cost-effective, and adaptable separation of emulsions
[41]. This concept is correspondingly applicable for the treatment of O/W emulsions.
Thousands of patents and articles have focused on the feasibility of numerous chemical
substances for destabilization of various emulsions in petroleum industry.

Attainment of efficient chemical demulsification approaches for the separation of W/O


emulsions relies on the implementation of the following steps [41]:
11

1) Sufficient amount of an aptly chosen chemical must be added to the emulsion.


2) Complete mixing of the chemical with the emulsion must be done.
3) Enough heating may be necessary to accelerate the separation of or completely break the
emulsion.
4) Adequate residence time in containers must be allocated to enable settling of demulsified
water droplets.

Some parameters repeatedly encountered in evaluation of demulsification processes are


introduced as follows:

2.1. Hydrophilic-lipophilic balance

In 1949, Griffin [42] described the relative propensity of a surfactant for water and oil by
introducing the concept of hydrophilic-lipophilic balance (HLB), a dimensionless numerical
term on the scale from 0 to 20; surfactants with low HLB number are liable to produce W/O
emulsion (< 10), whereas those with high HLB (> 10) produce O/W emulsions. This
quantification underpins the rule of thumb founded by Bancroft [43], implying that a
surfactant preferentially solubilized in the oil phase favors the formation of W/O emulsions
and that a surfactant preferentially solubilized in the water phase favors the development of
O/W emulsions. The HLB concept is simple, but it has several weaknesses; for instance, it
does not consider the impacts of temperature, salinity, co-surfactant concentration, and the
nature of the hydrophilic group [44,45]. Moreover, it is imprecise when assessing different
surfactant families [45], and practically is merely applicable to nonionic surfactants. This is
essentially because the HLB formula that makes use of the composition data of hydrophilic
portion of surfactants does not reflect the ionic surfactants’ hydrophilicity arising from their
ionization which indeed renders them more hydrophilic (i.e., their HLB value may in some
cases be more than 20); consequently, according to Griffin, an experimental procedure must
be used for these ionic products [46]. However, the experimental method is laborious, time-
consuming, and expensive, and hence the HLB values of only a few ionic surfactants have
been reported [46].
12

2.1.1. Relative solubility number

Similar to HLB, relative solubility number (RSN) is one of the features of nonionic
surfactants, which measures the combined inclination of the polar (hydrophilic group) and the
non-polar (lipophilic group) components of a surfactant to oil or aqueous phase. RSN has
been broadly employed by the surfactant manufacturers due to simplicity and objectivity of
its measurement, while HLB is hard to experimentally determine. The greater the RSN value,
the more hydrophilic the surfactant [47-49].

2.1.2. Effect of position isomerism on HLB

Wu et al. [50] investigated the effect of ethylene oxide (EO) and propylene oxide (PO)
positions in 20 block copolymers present in four various surfactant molecules, namely
Poloxamer (which is the generic term for nonionic triblock copolymers of poly(ethylene
oxide)-poly(propylene oxide)-poly(ethylene oxide) or PEO-PPO-PEO), PoloxamerR,
Poloxamine, and PoloxamineR, on surfactant properties and demulsification efficiency. They
observed dissimilar HLBs of surfactant for different positions of EO and PO, and this
difference in HLB values was more significant in surfactants with larger molecular weight,
reflecting that the influence of “position isomerism” is not taken into account when
calculating HLB using classical equations. There were two groups of block copolymers
examined, i.e., sequential surfactants and reverse-sequential surfactants. The sequential
surfactants are Poloxamer and Poloxamine where the acceptor initially reacts with PO and
then EO (i.e., EO moieties are at both ends, and PO segments in the center of copolymers),
while the acceptor of reverse-sequential surfactants (PoloxamerR and PoloxamineR)
primarily reacts with EO and subsequently with PO (i.e., PO moieties are at both ends, and
EO segments in the center of copolymers), as schematically displayed in Fig. 1. It was found
that more hydrophilicity was gained on sequential block copolymers than the reverse-
sequential ones with equivalent chemical structure, particularly for bulky molecules. In
general, it was deduced that sequential copolymers had superior demulsification effectiveness
for W/O emulsion than reverse-sequential ones [50].
13

Fig. 1. Simplified structures of the molecules of Poloxamer, PoloxamerR, Poloxamine, and


PoloxamineR [50].

2.2. R ratio

R ratio (R) is another parameter that was introduced by Winsor [51] to describe the ratio of
the net molecular interaction energy (per unit area of the interface) between the surfactant (C)
adsorbed at the interface and the oil phase (O) to that between the surfactant and the water
phase (W), as defined in Eq. (1).

(1)

Where,
ACO: interaction energy between surfactant and oil phase molecules
ACW: interaction energy between surfactant and water phase molecules
AOO: interaction energy between two oil molecules
AWW: interaction energy between two water molecules
14

The above definition of Winsor’s R ratio was later redefined to illustrate the overall net
interaction energy between the surfactant and its physicochemical environment (Eq. (2)).
However, by either definition, it does not make much difference in argumentation.

(2)

Where, the subscripts H and L respectively correspond to hydrophilic and lipophilic groups
of the surfactant molecule whose interactions have been completely considered in the latter
expression.

Consequently, for R < 1, representing Winsor type I phase behavior, the surfactant-rich phase
is the water phase (formation of O/W emulsion) which is in equilibrium with almost pure oil
phase, or in other words, the surfactant layer becomes convex toward the aqueous phase; and,
for R > 1, exhibiting Winsor type II phase behavior, the surfactant-rich phase is the oil phase
(formation of W/O emulsion) which is in equilibrium with nearly pure water phase, or the
convexity of the surfactant layer is toward the oil phase. Many authors have shown that the
highest demulsification performance or the least emulsion stability is related to a
physicochemical condition where R = 1 (Winsor type III phase behavior) [52,53], meaning
that the interactions of surfactant with both oil and water phases are balanced; in this case, a
bicontinuous microemulsion (also called middle- or surfactant-phase microemulsion) is
formed, which is in equilibrium with virtually pure oil and pure water excess phases.
Nonetheless, the R ratio is a qualitative factor, and is imperfect since the molecular
interaction energies cannot be experimentally determined.

2.3. Phase inversion temperature

In 1964, Shinoda and Arai [54] established an approach based on determining the phase
inversion temperature (PIT) of a surfactant-oil-water blend heated under stirring. At
temperatures below PIT, a surfactant is preferentially distributed into the aqueous phase as
oil-swollen micelles, whereas at those above PIT a surfactant is preferentially distributed into
the hydrocarbon phase as water-swollen inverted micelles. This approach is more trustworthy
than prior methods as it takes into consideration a range of variables including oil, salinity,
surfactant, and co-surfactant that influence the PIT. Nevertheless, the PIT method is only
appropriate for nonionic surfactants since ionic surfactants are less sensitive to the
temperature [44,55,56].
15

2.4. Hydrophilic-lipophilic deviation

Salager [57,58] introduced the notion of the generalized physicochemical formulation by the
term “surfactant affinity difference (SAD)” or its equivalent dimensionless expression, i.e.,
hydrophilic-lipophilic deviation (HLD), for both ethoxylated nonionic systems (Eq. (3)) and
ionic systems (Eq. (4)) which are linear equations encompassing all the variables involved,
such as surfactant nature, oil nature, salinity nature and concentration, temperature, pressure,
and the existence of co-surfactant (e.g., alcohol), and thus appear to be the most
comprehensive formulations presented so far, which are applicable to a variety of emulsion
systems:

(3)

Where,
α: linear function of the number of carbon atoms in the alkyl group
EON: average number of EO groups per nonionic surfactant molecule
ACN: number of carbon atoms in the alkane oil, or its equivalent if the oil is not an alkane
S: salinity of the water phase expressed in wt.% of NaCl
Φ(A): function of both the alcohol type and concentration
T: temperature (ºC)
k, b, CT: equation coefficients

The values of α, Φ(A), k, b, and CT can be found in the related publications [57,59,60].

(4)

Where,
aT: temperature coefficient of optimum salinity expressed in units of lnS per ºC
ACN: number of carbon atoms in an alkane
f(A): function of alcohol type and concentration
K: slope of the logarithm of the optimum salinity as a function of ACN
S: salinity, grams NaCl per 100 cm3 of aqueous phase
σ: characteristic parameter of surfactant
T: temperature (ºC)
16

The values of K, f(A), σ, and aT can be found in the relevant publications [57,61,62].

When the HLD (or SAD) value is zero or very close to zero, an “optimal formulation” is
achieved, suggesting that the emulsion is very unstable and can be simply ruptured. On the
other hand, positive or negative value of HLD corresponds to W/O or O/W emulsion,
respectively, provided that the water-to-oil volumetric ratio (WOR) is almost close to unity
[45]. At high WORs (3.5-10) and low WORs (0.1-0.25), O/W and W/O emulsions are present
at HLD > 0 and HLD < 0, respectively (contrary to the above explanation suggesting the
presence of the W/O and O/W systems, instead), which are associated with “abnormal”
emulsions holding a very low stability, due to insufficient external phase volume in each
case, although the surfactant is preferentially distributed in the external phase [63,64]. Lack
of compatibility for a wide range of WOR is therefore a shortcoming of the HLD
formulation. Based on the HLD expression, the salinity of the water phase for the emulsions
consisting of ionic surfactants, and the degree of ethoxylation (EON) or temperature for the
emulsions comprising ethoxylated nonionics are the variables that can direct the relevant
systems to or depart them from the optimal state (HLD = 0) [58]. To gain an optimum
formulation with an O/W system containing anionic surfactant for the purpose of chemical
EOR where HLD is negative, a demulsifier must be added to ensure the combination of
anionic surfactant and demulsifier leads to HLD = 0 or close to zero, associated with the least
emulsion stability [55]. By qualitative employment of the HLD concept, Nguyen et al. [55]
used some cationic demulsifiers and realized that they accelerated the breakup of synthetic
O/W emulsions emulsified by anionic surfactants, and attained oil and water phases with
enhanced quality compared with the primary (untreated) emulsion. The concentration of
demulsifier when mixed with anionic surfactants must generate HLD = 0, which is consistent
with the evidence that a small variation in demulsifier concentration (below or above
optimum value) may switch the emulsion from being unstable to stable, or vice versa.

2.5. Influential parameters

There are diverse parameters involved in chemical demulsification that alteration of each can
either decrease or increase the emulsion stability; the contribution of each of these features is
discussed as follows:
17

2.5.1. Salinity

Increasing the water salinity causes the interactions of emulsion-stabilizing surfactants with
water to be gradually decreased (i.e., reduced hydrophilicity of surfactants, known as
“salting-out” effect) [52,65]. For example, it has been observed that raising the salinity
(NaCl) of O/W emulsions reduces their IFT, and expands the range of optimum demulsifier
concentration at which minimum emulsion stability is achieved, implying that HLD can be
shifted from negative to almost zero by increasing the emulsion salinity, thereby enhancing
the demulsification effectiveness [55,66,67]. On the contrary, adding up to the aqueous phase
salinity of W/O emulsions is expected to render them more stable, owing to lessened affinity
of native surfactants to dispersed water phase [68]. In this direction, researchers found out
that when various resins were individually added into a typical W/O system, the more
hydrophilic additive would require higher emulsion brine salinity to reach a maximum
separation rate of dispersed aqueous phase [52]. Nonetheless, the salinity tolerance of
nonionic surfactants is significantly greater than that of anionic surfactants [40,60].
However, contradictory effects have been reported suggesting declined stability of W/O
emulsions at higher salinities (≥ 10%) [69,70]. Perles and colleagues [71] applied the
rheology method and explained this contrast by concluding that the presence of salt improves
W/O emulsion stability up to a critical ionic strength value (~ 0.1-0.3 mol L-1) above which
its stability is diminished. On the other hand, electro-rheological investigations proved that
W/O emulsions with water phase containing divalent cation (Ca 2+ or Ba2+) are more stable
than those with aqueous phase composed of monovalent cation (Na+ or K+) [70,71]. On the
contrary, the presence of salts with divalent cation raises demulsification competence of O/W
emulsion to a greater extent than that with monovalent cation, which is probably assigned to
the lesser hydrophilicity of the former [3,14,72].

In an oil-brine-anionic surfactant-alcohol system, the salinity at which approximately


identical volumes of brine and oil are solubilized in the so-called middle- or surfactant-phase
microemulsion (i.e., surfactant’s water-soluble tendency is the same as its oil-soluble
tendency), and the IFT between surfactant and brine is equal to that between surfactant and
oil is termed as optimal salinity whose value for the corresponding system can be switched to
a desired amount by changing the concentration and structure of alcohol component [73,74].
At the optimum point, it occurs a minimum in phase separation time (i.e., the fastest
coalescence rate), a minimum in IFT, a minimum in surfactant loss, and a minimum in
18

apparent viscosity, and the oil recovery is maximum [58,73]. The IFT ranges associated with
varied salinity regions (low, intermediate, and high) are stated in Section 2.7.

2.5.2. Temperature

Generally, increment of temperature is envisaged to have several advantages; it lowers the oil
phase viscosity, enhances the discrepancy between densities of the phases, increases the
number of droplets collisions, and typically weakens the stabilizing film surrounding the
droplets [75,76]. However, the diffusion rate of asphaltenes to the interface is augmented
when raising the temperature as a consequence of a decline in the bulk viscosity of crude oil
[25]. Basically, temperature diminishes the relative solubility of nonionic surfactants for
water, whilst the effect is opposite for anionic surfactants. In other words, solubility of
anionic surfactants in water is directly proportional to temperature, in contradiction of
nonionics [60]. As a result, it is expected that the interfacial adsorption of anionic surfactant
in an O/W emulsion is reduced upon elevation of temperature, contributing to lower emulsion
stability [77]. Nevertheless, conflicting finding has been published for the increase of
temperature in produced water from ASP flooding containing anionic surfactant, representing
enhanced emulsion stability [55], which might have been due to the opposing impact of
salinity in the relevant system, giving rise to the improved adsorption of anionic surfactant at
the oil-water interface. As pointed out in Section 2.4., stability of ionic surfactant-stabilized
emulsions is often controlled by salinity (rather than temperature).

2.5.3. pH

The effect of water pH on emulsion stability relies on the oil phase and brine compositions in
the emulsion. Also, oil-soluble organic acids or bases may influence the demulsification of
W/O emulsions. The hydrophilicity of surfactants commonly tends to increase upon increase
of pH value of emulsion. Consistent with this argument, the W/O emulsions are liable to be
generated in acidic environment (low pH), while the O/W emulsions are prone to be
developed in basic medium (high pH). Correspondingly, asphaltenic interfacial films are
most rigid in acidic medium, and gradually get weaker with the growth of pH, till they switch
to mobile/unstable films in alkaline environment. Conversely, the resinous films are most
robust in alkaline and weakest in acidic medium. Mechanistically, the dispersed phase
droplets acquire positive or negative surface charge in acidic or basic media, respectively,
due to the ionization of surfactants adsorbed at the oil-water interface induced by variation of
aqueous (or oil) phase pH [68,78-83]. The proposed “ionization effect” is promoted by the
19

addition of salts into the water phase, as a result of interaction/association of ions present in
the brine with surface-active species at the interface [78,82]. In the presence of demulsifier,
the optimal pH (at which separation efficiency is maximum) for the oil-brine systems
stabilized by asphaltenes, which are classified as amphoteric surfactants, has mostly been
reported to be neutral or near-neutral [78,84,85]. Even specific nonionics, which are the
surfactants with no charge, may uncover the behavior of ionic surfactants when placed in
acidic or basic settings. For example, polyethylene oxides exist in cationic form under acidic
conditions, and long-chain carboxylic acids are anionics under basic atmospheres [86]. As
such, it can be concluded that when the effect of studied water pH value or range is in line
with the presence of an interfacially active demulsifier in an emulsion, grander
demulsification efficiency may be accomplished. However, cautions have to be exercised
when the optimum pH for breaking emulsion is sufficiently acidic, to refrain from corrosion
incidents in the treatment facilities.

2.5.4. Water content

Water content in W/O emulsion is one of the principal parameters having effect on the
efficiency of demulsification or the stability of emulsion. The tenacious crude oil emulsions
depend not only on asphaltene, resin, and paraffin contents, but also on water content [87].
Ordinarily, increase of water content of emulsion in presence of demulsifier promotes
demulsification efficiency while decreasing the time needed for water separation as well as
the required demulsifier doses, provided that the water content is in the range of 0.3-0.7
volumetric fraction (i.e., WOR in the range of 30:70-70:30) [88,89]. In other words, the
emulsions containing higher water content are more readily broken than those with lower
water content [90], notwithstanding that increased water content renders crude oil emulsions
more viscous [11].

2.5.5. Oil content

Generally, emulsion separation is expected to take place faster and the rate of coalescence is
improved when the volume of dispersed phase increases due to the proliferation in entropy
for efficient collisions between the dispersed droplets [91]. In contrast, Zaki [92] reported
that decreasing the oil content from 90 vol.% to 60 vol.% in a surfactant-stabilized crude
O/W emulsion (in the absence of demulsifier) made some water separation occur and more
decrease of the oil content to 50 vol.% caused water phase to separate immediately. This is
perhaps because increasing the oil content in an O/W emulsion beyond 70 vol.% may result
20

in shifting the system to a viscous W/O type of emulsion [40], which is extremely stable due
to the excessive amount of oil phase.

The effect of oil content of an emulsion on its pour point is of paramount importance to
ensure that the pour point of a prepared emulsion does not increase to make transportation
difficulties in pipelines at low temperatures [92]. This argument must be taken into account
when O/W emulsions are intentionally formed to facilitate the delivery of viscous crude oils
by reducing their viscosity.

2.5.6. Agitation speed and duration

Increase of the speed or duration of mixing of O/W emulsions leads to reduction of average
size distribution of dispersed oil droplets, augmentation of both emulsion viscosity and
stability due to improved particle-particle interactions as a consequence of greater interfacial
area [40,76]. More specifically, in systems containing small dispersed droplets (diameter < 1
µm), the colloidal surface forces and Brownian motion forces prevail over the hydrodynamic
forces [40].

2.5.7. Emulsifier/demulsifier concentration

Evidently, an emulsion with higher emulsifier concentration requires larger concentration of


chemical demulsifier for treatment. Appropriately, the emulsions formed during the chemical
EOR as a tertiary oil recovery process need a higher amount of demulsifier (normally
hundreds of ppm, and even more in extreme circumstances) than those gained from primary
and secondary oil recoveries as clarified in Section 1. As a result, too little demulsifier may
not break the emulsion produced from tertiary oil extraction, while overdosing demulsifier
would result in the emulsion re-stabilization [55,93,94]. However, the hydrophilicity or
lipophilicity, and molecular weight of a demulsifier are the determining factors in
identification of its optimum dosage for breaking an emulsion. Equally, if an alkaline
additive, for instance, increases the hydrophilicity of asphaltenes in a W/O emulsion through
ionization effect, lesser amount of an efficient hydrophilic demulsifier for achievement of an
optimum formulation in which separation rapidly occurs is required (the relevant mechanism
is described in Section 2.11); likewise, demulsifying surfactants with too low molecular
weight (MW < 4000) are unable to resolve the W/O emulsion with low dosages (300-400
ppm) regardless of their hydrophilic/lipophilic affinities [45,47,83]. Additionally, separation
vessel dimensions may influence the optimal demulsifier dosage, owing to their impact on the
21

contact time between emulsion and separator [95,96], and/or on fluid dynamics. For example,
it was evidenced that larger bed height of a coalescing separator improved the oil removal
efficiency of an O/W emulsion due to the provision of longer contact/retention time [95]. In
terms of fluid dynamics effects, regions of high shearing, for instance, can cause drops to
come closer together and enhance their collision probability, resulting in decrease of the
demulsifier concentration required to achieve the minimum emulsion stability.

2.5.8. Droplet size and distribution

Oilfield emulsions (both W/O and O/W) are mostly classified as macroemulsions, which
have droplet diameters larger than 0.1 µm, and may even surpass 100 µm. Both droplet size
and distribution influence the emulsion viscosity; the smaller size droplets as well as narrow
drop size distribution of a generic emulsion results in an emulsion with higher viscosity and
thus more stability [81,97,98]. This is obvious since it takes longer for the smaller drops to
coalesce and eventually sediment (water globules) or float (oil droplets). An efficient
demulsifier facilitates the coalescence of dispersed drops.

2.6. Demulsifying materials

Surfactants have been most abundantly used as demulsifying materials due to their
amphiphilic nature and tunable chemical structure making them conducive to separation of
various emulsions.

2.6.1. Nonionic surfactants

Some of the characteristics of the nonionic surfactant block copolymers are explained below
in order to give insight into their functionality once added to emulsions. However, it should
be noted that the following features might be analogously construed when ionic surfactants
are deployed as emulsion-breaking substances.

2.6.1.1. Molecular structure

As stated above, asphaltene molecules are crude oil natural surfactants that preclude the
coalescence of water droplets in W/O emulsion systems by shaping a viscoelastic layer
between them. Le Follotec et al. [99] compared the demulsifying capability of four different
triblock copolymer surfactants; the copolymers were produced by two-lateral blocks of
poly(ethylene)oxide (i.e., PEO8 and PEO12) connected to a central block of
poly(dimethyl)siloxane (PDMS) consisting of 13, 24, and 46 monomers. They observed that
22

the polymer with extensive hydrophobic chain length (46 DMS) was not effective, while the
polymers with smaller hydrophobic chain lengths (13 or 24 DMS) were found efficient
demulsifiers. However, the addition of demulsifiers, either efficient or inefficient, affected
the structure in asphaltene network. Furthermore, it was realized that the demulsification-
effective copolymers for W/O emulsion were able to stabilize the model silicon-O/W
emulsions. In contrast, ineffective copolymers stabilized the model water-in-silicon-oil
emulsion. Similarly, Wu et al. [50] recognized that W/O emulsion rupture rate increased by
raising the EO proportion in sequential surfactants. On the contrary, Zhang et al. [41]
concluded from demulsification trials that decreased content of PEO increased the
dehydration speed and demulsification efficiency of W/O emulsions. Overall, all of this
evidence elucidates how significantly the hydrophobic or hydrophilic chain length of triblock
copolymers can influence their destabilization performance in O/W or W/O emulsion
systems.

2.6.1.2. Critical micellization concentration

A substantial, fundamental parameter is the critical micellization concentration (CMC), the


copolymer concentration at which micelles formation commences [100]. Briefly, surface
tension of Poloxamer copolymer solutions decreases with increasing copolymer
concentration and the resultant curve (i.e., surface tension vs. copolymer concentration)
exhibits two slope breaks at two different concentrations where the higher concentration
break corresponds to CMC value after which the surface tension remains constant. The
micellization of block copolymers is intrinsically more complicated than that of conventional,
low-molecular-weight surfactants [101]. A considerable difference is frequently noticed
between the CMC values obtained from varied approaches since their sensitivity toward the
number of individual copolymer molecules may differ [102]. Moreover, composition
polydispersity, batch alterations, discrepancies in the concentration ranges used in CMC trials
or inadequate temperature control may account for the observed differences [103].

Alexandridis and coworkers [101] examined the effect of PEO and PPO blocks as well as
molecular weight of PEO-PPO-PEO copolymers on the micellization. It was realized that the
CMC values for Poloxamer copolymers with similar size PEO block and diverse PPO
remarkably decreased with increasing number of PO segments, proposing that copolymers
containing a larger hydrophobic (PPO) chain create micelles at lower concentrations. On the
other hand, CMC data for a group of copolymers with alike hydrophobic block but
23

hydrophilic block of different size reveal an increase in CMC with increasing number of EO
segments. In another report, it has been noted that the increase of initiator to PPO in dendritic
polyether surfactants with unchanged EO content elevated their CMC, indicating that the
efficiency of decreasing surface tension reduced with the increase of initiator to PPO [104].
However, the effect of PPO on the CMC is more prominent than that of PEO (when the
changes in CMC value are determined per PO or EO portion), implying that PPO is
predominantly liable for the micellization [105]. It was also observed that for a given
PPO/PEO ratio, Poloxamers with higher molecular weight form micelles at lower
concentrations (lower CMC). Besides, a decrease in CMC value was attained with an
increase in temperature [101].

Based on calculated thermodynamic properties of some selected demulsifiers, researchers,


however, found that the absolute value of Gibbs free energy of adsorption (ΔGads) is
somewhat larger than that of Gibbs free energy of micellization (ΔGmic) for all examined
demulsifiers, suggesting that demulsifier molecules favor adsorbing on the interface rather
than forming micelles [88].

2.6.1.3. Steric effects

Steric effects are among substantial factors affecting the demulsification efficacy of
surfactants because they are dependent on spatial configuration of surfactant molecules. It is
given that the solubility of a nonionic surfactant in water is associated with the hydration of
oxyethylene groups which brings about the binding of no less than two water molecules.
Generally, hydration numbers of 2-4 are often noted [106,107]. To illustrate the different
steric effects between sequential surfactants and reverse-sequential ones, Wu et al. [50]
argued that for Poloxamer and PoloxamerR copolymers with the same number of EO
monomers, the PEO block in sequential surfactant (Poloxamer) hydrates with a greater
number of water molecules, causing larger hydrophilicity which favors demulsification of
W/O emulsion. On the other hand, for PoloxamineR, in which all the PEO groups are placed
in the center of the molecule, the PEO blocks from the copolymer’s four arms are much close
to each other that create packed PEO moieties leading to far lesser hydration than surfactant
Poloxamine where the PEO blocks are located at the end of each arm. However, the steric
effect for smaller molecules can be disregarded.

In other words, steric effects are influenced by position isomerism in demulsifiers [50].
Adding sequential copolymer into a W/O emulsion, the hydrophilic PEO blocks at the ends
24

of copolymer molecule adjoin to water droplets, whereas the hydrophobic portion orients to
bulk oil phase, generating an interface with trivial steric repulsion. In contrast, the
hydrophobic tail in reverse-sequential block copolymer is ramified into the oil phase,
broadens the shear plane, and impedes the aggregation of water droplets. This argument may
explain the inability, or only partial capability, of reverse-sequential surfactants for breaking
the W/O emulsion.

2.6.1.4. Surface pressure-molecular area

Surface pressure-molecular area (π-A) compression isotherms of spread (Langmuir)


monolayers of PEOx-PPOy-PEOx triblock copolymers are commonly noted to represent their
structural transitions as well as physicochemical properties at the air-water interface [41,108].
Typically, surface pressure increases as area per molecule of copolymer decreases, but with a
steeper slope of isotherms in smaller areas of large copolymers (such as copolymers
Poloxamer 407 and Poloxamer F88), corresponding to surface pressure values above about
10 mN m-1 in which the isotherm slope change is due to a structural transition. Nonetheless,
other structural transition (or perhaps monolayer collapse) indicated by further slope change
(bend of isotherm) is apparent at even smaller areas for certain large copolymer molecule
(Poloxamer F88). However, the effect of molecular area variations on surface pressure in
small block copolymer (e.g., Poloxamer P85) reveals a somewhat different pattern from those
in larger molecules [108].

Zhang et al. [41] evidenced similar slopes of π-A curves for the PEO-PPO-PEO copolymers
with the same PEO mass fractions. In addition, based on the same PEO content, the minimum
area per molecule increased with the increase of copolymer molecular weight. On the other
hand, when the number of hydrophobic (PO) monomers is identical, an increase in the PEO
proportion led to the expansion of molecular structure, and thus the minimum area occupied
by the copolymer molecule augmented.

Noteworthy to say that, when demulsifier molecules adsorb at the oil-water interface, surface
pressure is elevated; hence, alterations in surface pressure with time may provide an easy way
to monitor the penetration of demulsifier molecules into the oil-water interface [109]. A
summary of some recent, synthesized chemical demulsifiers are presented in Table 1.
Table 1
Summary of some recent, synthesized chemical demulsifiers

Demulsifier Emulsion treated Emulsion Operational Demulsifier Advantages Reference


preparation conditions characteristics
Tweens Synthetic W/O Water in heavy crude DC b < 700 ppm High molecular weight; No corrosion effect due [20]
oil; Containing groups of to neutrality
WOR a = 1:9 (v/v) ketone and ester;
HLB ≥ 11.0
Hydrophobically Synthetic O/W Diesel oil in water DC = 0.5 g L-1; Synergetic effect of Superior phase [110]
modified chitosan stabilized by SDS c; pH = 4-9 cationic and hydrophobic separation obtained with
derivatives Oil conc. = 0.05% functionalities twice and four-fold less
(v/v); doses than unmodified
SDS conc. = 0.0001 chitosan and cationic
M polyacrylamide
flocculants,
respectively;
Phase separation
independent of pH
values in the studied
range
Alkyl phenol Synthetic O/W Heavy crude oil in DC = 60 ppm; - Complete water [92]
formaldehyde distilled water T = 50 °C separation of emulsion
ethoxylated-propoxylated emulsified by
DBSNa d surfactant;
Oil content = 90
vol.%;
Surfactant conc. = 0.5
vol.%
Dendritic polyether Synthetic W/O Water in crude oil; DC = 150 mg L-1 Low PEO amount (20 High demulsification [104]
PAE82 WOR = 7:13 (m/m) wt.%) ratio (91.92%) achieved
in 15 min
Dendrimer copolymer Oilfield W/O (from - T = 60 °C PPO:PEO = 3:1 (wt/wt); Excellent [111]
SD31 Shengli, China) Highest number of demulsification
polymeric branches (most performance due to
complex molecular increased penetrability
structure) to the oil-water interface
REB e Oilfield O/W (from - DC = 10 g L-1; P-DcE:CaCl2:CPAM = High removal rates for [112]
26

Liaohe, China) T = 70 °C; 20:600:1.2 (m/m); mineral oil (98.04%)


pH = 7.76 Prominent effect of CaCl2 and COD (94.48%);
on promoting synergistic Inexpensive;
demulsification generated Environmentally
by REB’s constituents (P- friendly;
DcE, CaCl2, and CPAM) Low BOD f to COD g
ratio of REB effluent
offered good
performance in
anaerobic digestion
Bisphenol A Oilfield W/O - DC = 300 ppm 34 ethylene oxide units 100% demulsification [88]
after 58 min
(PE 6100 + RPE 3110) h Oilfield W/O (Aksaz - T = 80 °C PE 6100: RPE 3110 = Demulsifier dissolved in [113]
field, Kazakhstan) 1:1; kerosene was more
Both block copolymers efficient for dewatering
have similar PEO and than that in benzene;
PPO block contents; Dewatering degree
The demulsifier mixture equaled to 51.95%
dissolved in kerosene or
benzene (with conc. of
1%)
a
WOR: water-to-oil ratio
b
DC: demulsifier concentration
c
SDS: sodium dodecyl sulfate
d
DBSNa: dodecyl benzene sulfonic acid sodium salt
e
REB: calcium chloride + poly (dimethylamine-co-epichlorohydrin) (P-DcE) + cationic polyacrylamide (CPAM)
f
BOD: biological oxygen demand
g
COD: chemical oxygen demand
h
(PE 6100 + RPE 3110): mixture of commercial oil-soluble nonionic surfactants based on alkylene oxide block copolymers
2.6.2. Ionic surfactants

Cationic and amphoteric surfactants have barely been applied for the treatment of O/W
emulsions generated from SP and ASP flooding processes, though some of which have been
recognized promising for such applications [55,77], as listed down in Table 2.
Table 2
List of some efficient ionic surfactants proposed for demulsification of SP and/or ASP emulsions

Demulsifier Type of Emulsion treated Name/type of Emulsion Operational Outcome(s) Reference


demulsifier emulsifying preparation conditions
surfactant(s)
a
C8TAB Cationic surfactant Synthetic SP O/W Alcohol Crude oil in brine DC c = 200 ppm; The water quality of [55]
propoxylate emulsified by SP T = 25 ºC the treated emulsion
sulfate + alkyl components; was remarkably better
sulfonate Oil content = 30%; than that of untreated
(mixture of Alcohol emulsion upon 3 h
anionic propoxylate sulfate separation time
surfactants) conc. = 0.15%;
Alkyl sulfonate
conc. = 0.05%;
HPAM b conc. =
0.12%
d
C12TAB Cationic surfactant Synthetic ASP Crude oil in brine DC = 200 ppm; The water quality of
O/W emulsified by ASP T = 25 ºC the treated emulsion
components; was remarkably better
Oil content = 30%; than that of untreated
Na2CO3 conc. = emulsion upon 3 h
1%; separation time
NaHCO3 conc. =
0.043% ;
Alcohol
29

propoxylate sulfate
conc. = 0.038%;
Alkyl sulfonate
conc. = 0.113%;
HPAM conc. =
0.15%
Cocobetaine Amphoteric Synthetic SP O/W Petrostep S-1, Crude oil in brine DC = 75 ppm; Substantial clearing of [77]
i.e., sodium salt
surfactant stabilized by SP T = 25 ºC the aqueous phase was
of a propoxylated
branched C16-C17 components; obtained after 21 h
alcohol sulfate, +
Oil cut = 30%; separation time
Petrostep S-2,
i.e., sodium salt Petrostep S-1 conc. (remaining oil conc.:
of an internal
= 0.15 wt.%; 218 ppm);
olefin sulfonate
(mixture of Petrostep S-2 conc. Negligible amounts of
anionic
= 0.05 wt.%; water in the oil phase
surfactants)
HPAM conc. = 0.12
wt.%
a
C8TAB: octyltrimethylammonium bromide
b
HPAM: partially hydrolyzed polyacrylamide polymer
c
DC: demulsifier concentration
d
C12TAB: dodecyltrimethylammonium bromide
2.6.3. Ionic liquids

An ionic liquid (IL) is a salt in the liquid form. ILs are made from large organic cations that
can be symmetric or asymmetric, and anions. The cation is invariably organic (heterocyclic
or acyclic), and the anion can be a halogen, inorganic (such as [BF4]-, [PF6]-, [SbF6]-, [AlCl4]-
, [FeCl4]-, [AuCl4]-, [InCl4]-, [NO3]-, [SCN]-, etc.), or organic (namely [AcO] -, [N(OTf2)]-,
[CF3CO2]-, [CF3SO3]-, [PhCOO] -, [C(CN)2]-, [RSO4]-, [OTs]-, etc.). The properties and
composition of the ILs depend on their combinations of cation and anion.
Hydrophilic/hydrophobic features of ILs rely significantly on the structure of the cations and
anions. Non-volatility, non-flammability, catalytic attributes, environmental friendliness, and
thermal and electrochemical stabilities are among the most important properties of the ILs
that have famed them as attractive compounds for a variety of applications, particularly at the
petroleum industry [114,115]. The surfactant-like behavior of peculiar ILs has identified
them as potential chemical components for EOR processes, emulsion stabilizers, and/or
demulsifying agents [116-118]. To date, the demulsifying ILs have mostly been used in
conjunction with the other physical treatment means, namely microwave irradiation, in order
to boost up the separation efficiency of emulsions [119,120]. The limited number of
publications on the use of ILs as chemical demulsifiers [118-120], however, highlights the
necessity of further research to screen and identify more capable species for the targeted
treatment.

2.7. Oil-water interfacial properties

The oil-water interfacial properties including IFT (which is defined as the Gibbs energy
required to increase the unit interfacial area between two immiscible liquids at constant
temperature and pressure; its unit is mN m-1 or dyne cm-1), interfacial viscosity, life time, and
thinning rate of oil film in a W/O emulsion system are affected in the presence of a
demulsifier. Most of the researches conducted have dealt with the effect of demulsification on
IFT and interfacial viscosity at the oil-water interface [15,121]. However, the demulsification
performance is also connected to life time and thickness of oil film as well as dewatering rate
of emulsion. The more the reduction of oil film life time and the faster the thinning of oil film
are, the greater the dehydration rate of emulsion becomes. Kang et al. [122] discovered that
addition of demulsifiers to a synthetic W/O emulsion (no matter what type of demulsifier was
used) decreased the interfacial elasticity by replacing the emulsifiers at the interfacial film,
thereby reducing the strength and life time of oil film together with film thickness. The film
31

was broken once the thinning of film thickness reached a critical value. Hence, the lower the
oil-water interfacial elasticity is, the higher the demulsifying effectiveness will be.

Similarly, both the IFT and the interfacial viscosity of the W/O emulsion can be decreased by
using either water-soluble or oil-soluble demulsifiers; however, they are decreased to very
lower point when water-soluble demulsifiers are applied [122]. This effect may be attributed
to the greater surface adsorption rate when demulsifier exists in water than when it is present
in the oil phase [109]. In addition, for the systems with analogous type of demulsifiers, it was
discovered that increase of demulsifier concentration lowered the IFT and interfacial
viscosity, enhanced the dewatering rate, and improved the demuslfication productivity [122].
Further, it was observed that an O/W emulsion system with lowest equilibrium IFT attained a
good separation, good water quality, and low basic sediment and water (BS&W) value [55].
Optimal demulsification capability, or the least emulsion stability, of both W/O and O/W
systems were accomplished when noticeably low IFTs were reached [55,123]. On the other
hand, Zaki [92] declared that increasing anionic surfactant concentration in aqueous phase of
their examined crude O/W emulsion reduced the oil-water IFT value and consequently
enhanced the emulsion stability. These results are in accordance with the Gibbs adsorption
isotherm where a decrease in the IFT of an emulsion implies an increase in concentration of
surfactant molecules at the oil-water interface. Conversely, it was reported that IFT increased
after addition of demulsifier into the emulsion [124]. Jones and colleagues [125] also
discerned hardly different IFT values for various emulsions having widely different stabilities
at 25 ˚C. Equally, lowered emulsion IFT after addition of a demulsifier did not necessarily
dictate an improved resolution capacity [15]. It can, therefore, be apperceived that IFT is not
an imperative factor in evaluation of emulsion stability, and such other parameters as zeta
potential as well as interfacial rheology may also be considered.

IFT is a function of time, temperature, and emulsifier(s)/demulsifier(s) type and


concentration [25,109]. The variation in crude oil-water IFT with time is described by
consideration of demulsifier adsorption process happening in two steps [109]: primarily, the
demulsifier molecules transfer from the surface next to the interface, hence the IFT declines
rapidly. Once the surface layer is drained, demulsifier molecules have to diffuse from the
bulk phase, which takes place slowly, and thus the IFT decreases gradually. Eventually, the
interface gets saturated and the equilibrium IFT is obtained. It is believed that the diffusion
rate of demulsifier molecules to the interface rises with emulsion temperature as well as
demulsifier concentration; the IFT therefore falls faster at higher temperatures and
32

concentrations. Development of ultra-low IFT in ASP flooding technology depends on


several factors including alkali, surfactant, and crude oil properties. Surfactants have the most
prominent effect on reducing IFT, though alkalis lower IFT as well, and polymers have
negligible influence on IFT reduction [124,126]. Alkali reacts with acidic molecules in crude
oil and forms some new soap-like surfactants which accumulate at the crude oil-water
interface and synergistically decrease the oil-water IFT; the stronger the alkali, the larger the
IFT declines [124,127].

The IFT values of oil-brine systems comprising 2% of anionic surfactant (petroleum


sulfonate) along with 1% of a short-chain alcohol used as “co-surfactant” or “co-solvent” at
low salinities (≤ 1.4%), where the O/W emulsion coexists with nearly pure oil phase (Winsor
type I), as well as at high salinities (≥ 1.7%), where the W/O emulsion coexists with excess
brine (Winsor type II), usually exceed 0.01 mN m-1 [74]. However, in the region of
intermediate salinities (between 1.4% and 1.7%) in which the alleged middle/surfactant phase
is in equilibrium with virtually pure oil and pure brine (Winsor type III), ultralow IFTs
(between 0.0001 and 0.001 mN m-1) are obtained with brine and oil at its low- (1.4%) and
high-salinity (1.7%) parts of existence, respectively [74].

The complex interfacial dilatational modulus is a primary factor in controlling emulsion


stability, defined as the IFT increment per unit partial interfacial area change, which
encompasses two components (i.e., the dilatational elastic modulus and the dilatational
viscous modulus) as a function of frequency [15,126]. The characteristic relaxation time
induced by interfacial film drainage upon addition of a demulsifier is measured from peak
frequency of the plot of dilatational viscous modulus against low frequency range [15]. Both
elastic and viscous moduli of crude oil-water interface pass through a minimum value at
certain concentration of an effective demulsifier [96], signifying the optimum demulsifier
concentration as enucleated in Section 2.5.7. Viscoelastic properties of crude oil-water
interface are drastically improved once an ultra-high-molecular-weight polymer (such as
partially hydrolyzed polyacrylamide or its modified products) is added to the aqueous phase
in chemical EOR [128].

Sun et al. [129] compared the effect of two commercial demulsifiers with different structures
(branch chain and straight chain) on dilatational viscoelasticity of oil-water interfacial film
consisting of surface-active fractions from crude oil. They observed that demulsifier with
branch-chain structure had larger aptitude to replace film-forming substances than straight-
33

chain demulsifier, owing to the smaller size of free spaces between branch-chain demulsifier
molecules to which surface-active agents of either small or large molecular size were unable
to enter. Consequently, after adding a branch-chain demulsifier, interfacial film of large and
small size surface-active fraction molecules represented viscoelastic properties much alike to
those of demulsifier alone. On the other hand, when a straight-chain demulsifier was
employed, surface-active fractions of small molecular weight could simply penetrate to the
vacancies between demulsifier molecules and generate a mix-adsorption film with
demulsifier; thus, the resulting interfacial film revealed diverse viscoelastic characteristics
from those of demulsifier and surface-active fractions alone. In other words, the reciprocity
of demulsifier and surface-active substances at the oil-water interface is significantly
dependent upon the configuration of demulsifier as well as the molecular size of surface-
active agents.

Kim and coworkers [130] achieved a good relationship between the demulsifier efficiency
and dynamic interfacial film elasticity, suggesting that the demulsifier performance is
inversely proportional to the resultant interfacial elasticity. Indeed, an effective demulsifier
for a rapid coalescence is the one representing fast film stress-relaxation kinetics
corresponding to low interfacial elasticity which is linked to the demulsifier’s diffusivity and
dynamic interfacial activity [15,130]. In other words, faster diffusivity and larger dynamic
interfacial activity of a demulsifier lead to greater demulsification performance. The term
“diffusivity” is important since the process of demulsifier adsorption at the oil-water interface
is diffusion-controlled, which therefore demands an exploration into the kinetics of
demulsifier adsorption. However, while dynamic interfacial activity of a demulsifier may
determine its demulsifying effectiveness, static (equilibrium) interfacial activity or tension of
a demulsifier does not demonstrate an apparent correlation with demulsifying efficiency
[130].

The elastic modulus of the copolymer PEO-PDMS-PEO at crude oil-water interface was
determined employing liquid-liquid Langmuir balance method [131]; its value converged to
zero once surface pressure was increased above 15 mN m-1 pointing to the interfacial film
solubility in one or both phases due to the interactions of native surface-active substances
with copolymer comprising hydrophilic PEO and hydrophobic PDMS moieties that resulted
in the interfacial film destabilization. It is presumed that the stability of emulsions is
associated with their elastic behavior at the oil-water interface [132].
34

2.8. Zeta potential

Zeta potential is a key measure of the surface charge of dispersed drops in an emulsion; its
value pertains to the structure of electrical double layer formed around the droplets. The
increased thickness of electrical double layer increases the absolute value of zeta potential,
the charge repulsion, and consequently the emulsion mechanical stability. The charges of
internal phase droplets are founded through three ways: adsorption, ionization, and friction
[24]. Acoustic and electroacoustic spectroscopies are the preferred methods for determination
of zeta potential than the traditional approaches (light scattering, electrophoresis,
sedimentation, etc.), due to the omission of emulsion dilution step needed for most
conventional techniques [133]. In emulsion systems generated from ASP flooding process,
the effects of alkali, surfactant, and polymer on zeta potential are related to the properties of
these chemicals [124]; in general, the adsorption of these chemicals including an anionic
surfactant at the oil-water interface leads to a decrease in zeta potential (i.e., results in more
negative zeta potential), owing to the enhancement of negative electric charge density on the
surface of oil globules. However, the result has been recognized to be reverse for high alkali
concentrations at which zeta potential represents an ascending trend, which may be ascribed
to the presence of sodium ion (in NaOH) decreasing the thickness of electrical double layer
on the oil droplets surface [134].

Addition of a suitable demulsifier to an anionic surfactant-stabilized O/W emulsion promotes


the zeta potential (reduces the anionic charge on the surface of oil globs), reflecting that
demulsifier molecules are adsorbed at the oil-water interface and replace the emulsifying
surfactant, thereby lowering the electrostatic repulsion between the dispersed droplets and
subsequently facilitating their closeness and coalescence leading to emulsion resolution
[55,124]. Less negative zeta potentials can be obtained by increasing demulsifier
concentration [55].

2.9. Demulsifier partitioning

Most commercial demulsifiers employed to rupture the W/O emulsions are soluble in oil. The
interfacial activity of the oil-soluble demulsifiers is dominated by the diffusion rate from the
bulk phase to the interface plus the adsorption obstacle at the oil-water interface. In specific
cases, some oil-soluble demulsifier components can be dissolved (partitioned) into the
dispersed phase (water droplets in the case of W/O emulsion); this process is referred to as
“partitioning,” which is distinguished as a key factor in demulsification mechanism. The
35

partitioning/distribution coefficient, defined as the ratio of demulsifier concentration in the


aqueous phase to that in the oil phase, is markedly affected by temperature, hydrocarbon-
phase polarity, and electrolyte content in water [29,135,136]. The partitioned demulsifier
components in the droplet phase can vigorously influence the oil-water dynamic interfacial
characteristics, such as the IFT gradient or the Marangoni-Gibbs stabilizing effect [136]. In
order to accomplish a quality performance for destabilizing a generic emulsion, a demulsifier
primarily dissolved in the continuous phase should hold the following properties
[109,135,136]:

1) The demulsifier has to be able to partition into the dispersed phase in order that its
partitioning coefficient between the water and oil phases approximates unity.
2) The demulsifier concentration in the droplet must be adequate to guarantee sufficient
diffusion flux to the interface.
3) The interfacial activity of the demulsifier must be sufficiently large to suppress the IFT
gradient as well as the Marangoni-Gibbs stabilizing effect, thereby expediting the film
drainage and thinning rate, and enhancing coalescence.

A partitioned demulsifier increases the static/dynamic interfacial activity and submits a high
adsorption rate to the interface while efficiently reducing the film dilatational modulus
together with dynamic film/interfacial tensions, yielding a supreme demulsification
performance. Hence, without demulsifier partitioning, the demulsifier is unable to effectively
diminish the film dilatational modulus owing to the lack of adsorption on the film interface.

2.10. Aging of W/O emulsions

Stability of crude oil emulsions has been observed to enhance with aging [125]. However, the
aging process takes long times for crude oil-brine interface, which proves the time-dependent
stability of crude oil emulsions. This implication is mainly due to the slow adsorption of
natural surfactants present in the crude oil at the interface, which accordingly prolongs the
formation of a saturated interface [109].

The effect of W/O emulsions aging for a 60-day period on their droplet size distribution,
water content, dynamic viscosity, and stability as well as on their rheological properties were
studied by Maia Filho et al. [132,137] using one sample of petroleum emulsion and four
samples of model (synthetic) systems containing various oil phase characteristics, and the
following conclusions were drawn:
36

 Asphaltenes in crude oil play a decisive role in aging of W/O emulsion.


 The most noticeable impact of aging is by changing the interfacial film of emulsions, as
specific properties of an emulsion including droplet size distribution, emulsified water
content, and dynamic viscosity experience trifling alterations.
 Aging can make severe impairment to the demulsification process of petroleum emulsion
by drastically lowering the amount of water separation in aged emulsion (below 30 days)
compared with that in fresh emulsion system.
 Real petroleum emulsion was more influenced by aging than the examined synthetic
emulsions with similar asphaltenes content, implying the more stability of the latter. This
finding is possibly due to faster migration of asphaltenes to the interface in the model
emulsions, resulting in increased elastic modulus of interfacial film; and as the higher
elastic modulus corresponds to superior emulsion stability, the model emulsions became
more stable. As a result, the model emulsions did not display water separation at any
aging days even after addition of demulsifier, on the contrary to the petroleum emulsion
representing a decreasing water separation by aging less than 30 days.
 The aptitude of emulsions to aging is governed by solubility of asphaltene molecules in
the dispersant medium (oil phase). The higher the solubility of asphaltenes in the oil
phase is, or the more asphaltenes-solvent interactions are, the lesser their migration to the
oil-water interface will be, and thus emulsion with lower stability is formed. Aging of
petroleum emulsion caused more asphaltene aggregates, primarily solubilized in medium
because of solvency effect of the oil phase, to gradually move toward the surfaces of
droplets and fill in the cavities present at the interface. However, after 30 days of aging,
the interfacial film was cohesive and resistant enough to avoid the effectual interaction of
demulsifier with the dispersed water phase. Besides, among the emulsion systems with
the same asphaltenes content, the model emulsion with great aromaticity of the oil phase,
leading to robust interaction of medium with asphaltenes and consequently slower
migration of asphaltenes to surface, demonstrated a more pronounced effect of aging.

2.11. Demulsification mechanisms

To date, numerous chemical demulsification mechanisms have been introduced in the


literature, insinuating the complexity of destabilization processes involved. This
sophistication principally originates from the multifaceted demulsification knowledge
incorporating colloid chemistry, surfactants science, and interfacial phenomena altogether. In
37

this part, we intend to briefly go through the mechanisms of emulsion breaking caused by
various demulsifying surfactants (i.e., nonionic and ionic surfactants), and at last reach a
simplified conclusion applicable to the systems with relevant surfactant class.

2.11.1. Nonionic demulsifiers

Daniel-David et al. [138] employed several techniques to identify the destabilization


mechanism of W/O emulsions using polysiloxane copolymers. Eventually, they proposed a
two-step mechanism as follows:

1. Adsorption of the copolymer into packing defect spots at the crude oil-water interface
because of its great susceptibility to the oil-water interface;
2. Disruption of the network of asphaltene aggregates by adsorbed copolymer molecules,
provoking the emulsion breakup and coalescence of water droplets.

Le Follotec et al. [99] related the demulsifying efficiency of a copolymer to its spontaneous
curvature and described the demulsification process, as shown in Fig. 2. In W/O emulsion
systems, the oil layer between the two water droplets has to break to allow the coalescence of
water drops, i.e. small but greatly curved hole must nucleate into the oil layer. This process is
mostly followed by an energy hindrance concerned with the new interface formation. The
effective copolymers with moderate hydrophobic segment may contribute to the hole
nucleation, leading to the separation of W/O emulsions, while in O/W emulsion systems they
adversely function by obstructing the nucleation of a gap in the water layer between two oil
droplets, and thus favor the emulsion stabilization. In contrast, the bulky central hydrophobic
portion of an inefficient triblock copolymer unfavorably hinders the hole formation, prevents
coalescence, and consequently improves the stability of W/O emulsions.
38

Fig. 2. (a) Coalescence of two water droplets during emulsion destabilization by efficient
molecules, and (b) connection without coalescence with inefficient molecules [99].

In an earlier report, a model based on the “revisited oriented wedge theory” was established
to examine the relationship between the equilibrium phase behavior of oil-water-surfactant
mixtures and macroemulsion type and stability [139]. It was argued that both the equilibrium
phase behavior and emulsion stability are dependent on the bending elasticity of surfactant
monolayer at the oil-water interface. As such, from the developed model, the following points
were elicited:

i. At positive spontaneous curvatures, O/W emulsions are very stable, and at negative
values W/O emulsions are stable, whereas in the balanced state of the surfactant film (i.e.,
zero interfacial curvature), an emulsion normally breaks for a broad range of systems.
39

ii. The monolayer at the edge of a nucleation hole in the emulsion film is deeply curved,
which causes the coalescence barrier to be acutely dependent on the sign and the absolute
value of the monolayer spontaneous curvature.
iii. The thickness of the emulsion film is allowed to change to lessen the free energy of the
nucleation hole.
iv. Double-tailed surfactants are prone to stabilize W/O emulsions, while single-tailed
surfactants stabilize O/W emulsions.

According to the above theory, it was inferred that demulsification phenomena can be
generally induced by:

- Lowering the absolute value of the monolayer spontaneous curvature by the addition of
opposite-packing-type surfactants.
- Decreasing the bending elasticity of monolayer by adding short-chain alcohols.

Wang et al. [104] studied the demulsification mechanism of an O/W emulsion demulsified by
dendritic polyether surfactants applying single droplet protocol proposed by Cockbain and
McRoberts [140]. In this regard, the breakup of the emulsion is apportioned into two
consecutive stages [140]: firstly, the demulsified water present between the bulk oil phase
and the oil droplet covered with natural emulsifier is drained into the aqueous phase. At this
stage, the droplet to separate has not yet ruptured. This step is called drainage process, and
the time at which the droplets are “about to break” is referred to as drainage time. Secondly,
the bulk oil phase is completely emerged as the oil drops coalesce with the oil bulk, and a
fully flat oil-water interface is formed. Exploiting this mechanism, there were three terms
examined, namely drainage time, half life time (half the time it lasts for the droplets to
disappear), and rupture rate constant. The droplets stability was measured by their half life
time, and the grouping of all these three terms was used to determine the interfacial film
strength.

It was concluded that all the above-mentioned terms were reliant on the demulsifier
concentration and the EO content. Both the drainage time and half life time decreased with
either an increase in demulsifier concentration or a decrease of EO content, and the rupture
rate constant increased with an enhancement in demulsification performance. However, the
drainage stage was found to be more significantly affected by the variables such as EO
content. As a result, the drainage stage was proposed as the rate-determining step for
demulsification utilizing dendritic polyethers.
40

Indeed, the demulsification mechanism of W/O emulsions may be interpreted by illustration


of film drainage process in the presence of demulsifier along with the effect of concentration
of natural crude oil surfactants and demulsifier on crude oil-water IFT [109,135], as depicted
in Fig. 3. When two water drops approach each other as a consequence of gravitational
forces, agitation, thermal convection, or so forth, the thickness of the intermediate oil phase
film shrinks. The shear stress pertaining to film drainage is likely to accumulate the innate
surfactant molecules outside the film, and hence their concentration inside the film is
reduced. Subsequently, an IFT gradient is developed with low IFT outside the film and high
tension inside the film. As shown in the Fig. 3, demulsifier molecules are adsorbed in the
voids left by innate surfactant molecules in the film. As apparent in the changes of IFT with
time, the adsorption rate of demulsifier at the crude oil-water interface occurs far more
rapidly than that of crude oil natural surfactants. Adsorption of demulsifier annuls the
produced IFT gradient and improves film drainage. Lastly, the film gets very thin, and
because of the propinquity of the dispersed phase droplets, the van der Waals forces of
attraction overcome and droplets coalesce.

It should also be noted that the adsorption of a competent hydrophilic demulsifier with an
appropriate concentration at the interface beside asphaltene molecules produces an
amphiphilic mixture which unveils the same affinity towards both phases; this condition
corresponds to achievement of an optimum formulation in which the dispersed droplets fuse
quickly upon contact with one another as if there was no emulsifier at the interface [45].
41

Fig. 3. (a) Film drainage in the presence of demulsifier; (b) Influence of concentration of
innate crude oil surfactants and demulsifier on crude oil-water interfacial tension [109].

2.11.2. Ionic demulsifiers

Until now, the mechanism of ionic demulsifying substances has not been elaborately
illuminated in the related published works. However, the effectiveness of cationic
demulsifiers in resolution of O/W emulsions stabilized by anionic surfactants has been
justified by ion pair formation of these two ionic species, resulting in neutralization of their
charged heads [55,77,141,142]. In this manner, analogous to the mechanism of ion pair
adsorption at solution-air interfaces [143], cationic-anionic (also known as catanionic) pairs
are primarily formed in the brine and then adsorbed at the oil-water interface, displacing the
anionic emulsifier. Another supposition may be the adsorption of cationic surfactant into the
voids left between the anionic surfactant followed by ion pair formation at the interface itself.
In either plausible mechanism, the emulsion destabilization process is fulfilled, and both of
these possibilities might even happen concomitantly. As a result, the negative charge on the
surfaces of oil drops triggered by the previously adsorbed emulsifying surfactant is alleviated,
and thus the electrostatic repulsion force between the dispersed globs is weakened, leading
them to get nearer and fuse in order that ultimately phase separation takes place.
42

2.11.3. Conclusion

The concepts of terminologies R ratio 1, HLB 10, PIT, HLD 0, and zero spontaneous
interfacial curvature in demulsification phenomenology all equivalently hint the balanced
state of the mixture of demulsifier and emulsifier(s) or of the demulsifier alone at the
interface (i.e., neither too hydrophilic, nor too hydrophobic) of W/O or O/W type of
emulsion, whereby the least emulsion stability is accomplished. Accordingly, the following
mechanisms are inducted:

 W/O emulsions: an efficient hydrophilic nonionic demulsifier with a proper


concentration diffuses to the voids left between emulsifier(s) (e.g., natural surfactants in
crude oil emulsions, or nonionic surfactants) at the surface of water droplets so that the
combination of demulsifier and emulsifier(s) attains an optimum formulation in which the
mixture reveals the same tendency to the oil phase as to the water phase. However, even
if it is argued that a highly surface-active demulsifier completely displaces the
emulsifier(s) at the interface, the demulsifier has to satisfy the defined optimum
formulation in order for the water drops to coalesce promptly. In the latter case where
only the demulsifier is present at the interface, its partitioning coefficient approaches
unity (i.e., the demulsifier equally partitions into both phases). A schematic illustration of
interfacial film thinning process between two droplets in water-in-crude oil emulsions
upon the addition of a capable demulsifier is displayed in Fig. 4. The interfacial film is
weakened, drained, and gets thinner before it is ruptured and droplets coalesce as a result
of the predominance of van der Waals forces of attraction between the dispersed drops.

Fig. 4. Schematic image of interfacial film thinning process between two droplets in
water-in-crude oil emulsions after the addition of an efficient chemical demulsifier.
43

 O/W emulsions: for the resolution of O/W emulsions stabilized by nonionic surfactants,
the same demulsification mechanism as discussed for W/O emulsions is expected with the
exception that adequate amount of an effective hydrophobic nonionic demulsifier has to
be added to the relevant emulsion to offset the hydrophilicity of emulsifying agent(s) at
the interface. Yet, the destabilization mechanism of anionic surfactant-stabilized O/W
emulsions using a competent cationic demulsifier is presumed to be ‘catanionic pair
adsorption/formation’ at the interface, both possibly occurring simultaneously. The
interfacial adsorption of the cationic-anionic pairs is basically followed by the
displacement of the anionic emulsifier.
44

3. Biological demulsification

Research regarding biological demulsification has been commenced since early 1980s
[144,145]. Microbial or bio- demulsifiers are proposed to be potential alternatives to
chemically synthesized demulsifiers commonly applied in petroleum industries. Microbial
demulsifiers have three main advantages over the chemical demulsifiers [146]:

1- Low toxicity of biological compounds,


2- Facile biodegradability of the compounds,
3- Exclusive properties of natural products that synthesized chemicals do not possess.

While physicochemical demulsification routes demand high capital costs, and the emulsions
produced at the wellhead are required to be delivered to processing facilities, an efficient
biological demulsification may be exploited immediately at the wellhead, thus reducing the
transportation expenses as well as high equipment expenditures [147]. The main benefit of
microbial demulsification is that, provided hydrocarbons present in the emulsion can sustain
the microbial growth, both microorganism growth and demulsification may take place at the
same time [148,149]. Meanwhile, biodemulsifiers can be recycled and reused without
noticeable loss of activity [150], and might be highly effective under extreme conditions
[151].

Most researchers have mainly focused on maximization of both biodemulsifier production


yield and biodemulsification rate, which are extremely medium-dependent, by varying the
operating variables, such as carbon source, nitrogen source, pH, salinity, temperature, and so
forth. The most competent biodemulsifier is the one that can sustain its stability when
subjected to a wide range of pH, temperature and salinity. In the following sections, it is
attempted to introduce some of the factors that may potently influence on either
biodemulsifier production, demulsification performance towards W/O and/or O/W systems,
or both.

3.1. Cell concentration

Increase of cell concentration has different effects on demulsification of W/O and O/W
emulsions. This difference can be realized by comparing the corresponding half-life values of
such emulsions. Half-life of emulsion may be used as a parameter to describe the emulsion
stability, and it is defined as the time at which the emulsion volume has decayed to half its
primary volume [6]. Clearly, the lesser the half-life value is, the faster the decay of emulsion
45

would be. Das [152] found that an elevation in the concentration of kerosene-washed cells of
an isolated Micrococcus species had a direct effect on demulsification of O/W emulsion with
a decrease in its half-life value, while the same increase in the cell content had an opposite
impact on demulsification of W/O system by raising its associated half-life time. This
delayed rupture of W/O emulsion with increased quantities of microbial cells beyond a
certain value may be owing to creation of a defensive layer around the dispersed drops rather
than assisting them to coalesce [148].

3.2. Chemical treatment

Washing of microbial cells, which contain lipid components, with lipid-solvating solvents
definitely alters the demulsifying characteristics of respective cells. Generally, n-alkane-
grown microorganisms are inclined to gather remarkable amounts of lipid compounds in their
cell wall [153]. These lipid materials may be removed from the cell wall through post-harvest
washing with lipid-solvating solvents, which consequently influences on the cell wall
hydrophobic affinity to a varied degree depending on the type of solvents [152]. In this
respect, some of the proposed solvents include n-pentane for elimination of neutral lipids
[148,154], n-hexane for deletion of certain glycolipids [155,156], chloroform-methanol-water
(CMW) for removal of total extractable lipids [148,157], and kerosene, being a combination
of n-alkanes, for exclusion of lipid materials [158].

According to the calculated half-lives of a model O/W emulsion, Das [152] observed faster
decays of emulsion with pentane- and kerosene-washed Micrococcus sp. cells, but more
sluggish decays with hexane- and CMW-treated cells than with unwashed cell. The longest
half-life value, or the most delayed demulsification, of the examined O/W emulsion was
evidenced for CMW-washed cells, probably because of removal of a notable proportion of
cell wall-bound lipids [148,157]. On the other hand, washing of the cells with none of the
aforementioned lipid-solubilizing solvents proved beneficial for demulsification of a W/O
system [152]. Similarly, in another report, washed pelleted cells of a mixed bacterial culture
demonstrated a declined demulsification performance for W/O emulsion in comparison with
pelleted cells, connoting that some demulsification activity was destroyed by washing as a
result of either some cell soluble component concerned or the loss of cells [159]. Likewise,
solvents with different properties used for treatment of bacterial cells leave distinctive effects
on their demulsification activity. Huang et al. [151] noticed elevations in demulsification
ratio of the cells of strain Alcaligenes sp. S-XJ-1 after separate chemical treatments with
46

compounds dichloromethane and acetone, but dramatic decline in demulsification ratio of the
cells after treatment with methanol or an aqueous alkali as compared with the untreated cells
used for demulsifying a model W/O system.

3.3. Substrate

The type of bacterial growth substrate has a significant impact on demulsifying efficiency of
microbial cells. This may be ascribed to their ability to alter the cell surface chemical
composition [160]. However, their high cost creates the greatest barrier for large-scale
production of biodemulsifiers [161]. It was found that microbial cell culture grown on
petroleum hydrocarbon substrates (such as crude oil, 10W30 motor oil, and diesel) revealed
high growth and demulsification capabilities for petroleum oil emulsion, whereas culture
grown on non-petroleum (carbohydrates) substrates (namely canola oil, starch, sucrose, and
glucose) did not attain high demulsification, even though all of the carbohydrates splendidly
supported the bacterial growth [159]. This outcome reflects that petroleum hydrocarbon
substrates may stimulate larger assembly of cell-related components or biosurfactants
accountable for demulsification ability. Alternatively, the carbon source has a decisive effect
on demulsification efficacy.

Biosurfactants or biological surface-active agents are produced by microorganisms typically


grown on hydrocarbons as water-insoluble substrates. Nonetheless, some biosurfactants have
been reported which are grown on water-soluble substrates (such as sugars, glycerol, or
ethanol) [162]. Indeed, most cell wall-associated biodemulsifiers have been produced on
hydrophobic carbon sources, such as crude oil [163], kerosene [164], tetradecane [152], and
hexadecane [144,145]. This is supposedly due to the existence of amphiphilic compounds,
like a biodemulsifier, on microbial cell wall, which may accelerate the consumption of
hydrophobic substrates, and therefore aid bacteria to endure in a hydrocarbon-contaminated
environment. Also, greater cell surface hydrophobic capacity may be achieved in presence of
cell wall-associated biodemulsifiers [165]. Numerous substrates can function as hydrophobic
carbon sources, including oils, fatty acids, and their respective esters, alkanes, etc. [166]. In
contrast, reports on production of biodemulsifiers from hydrophilic carbon sources have
rarely been presented, except for the production of the extracellular biodemulsifier, acetoin,
by Bacillus Subtilis grown on glucose, applied for the separation of a model O/W emulsion
system [167]. However, combined hydrophilic and hydrophobic carbon sources have also
been introduced, such as utilization of glucose along with liquid paraffin for the synthesis of
47

extracellular biodemulsifier from demulsifying strain Bacillus mojavensis XH1, where


glucose took part in the demulsifier biosynthesis, and liquid paraffin enhanced the
lipophilicity as well as sprinkle of biosurfactant [168]. Liu et al. [165] experimentally
concluded the following assertions by comparing several hydrophobic carbon sources with
various hydrophilic substrates for the production of a biodemulsifier from Alcaligenes sp. S-
XJ-1, used to beak a simulated crude oil emulsion:

 Cell wall-related biodemulsifiers were more favorably produced by hydrophobic


carbon sources than hydrophilic ones.
 Hydrophobic substrates resulted in higher demulsification ratios and surface activities.
 Shorter emulsion half-life values were achieved on the biodemulsifier generated from
hydrophobic carbon sources.
 When hydrophobic substrates were employed, “microbial adhesion to hydrocarbon”
(denoting the cell surface hydrophobicity) was at least 3 times that when hydrophilic
substrates were used.

Exploitation of industrial wastes and byproducts as substrates for the bulk production of
biosurfactants is recommended to reduce the overall cost of their production processes, and in
the meantime, battle the pollution effect of the wastes [150]. Liu et al. [169] proposed that the
waste frying oil obtained from frying flour was an efficient, cheap carbon source for the
synthesis of biodemulsifier by Dietzia sp. S-JS-1, and the resultant biodemulsifier exhibited
excellent demulsification capability and higher reaction rate than its counterpart produced on
paraffin as carbon source, given the equal amount of substrates and period of cultivation
applied. Higher biomass production as well as a further varied chemical structure of
biodemulsifier generated from the associated waste frying oil contributed to such distinctions.

3.4. Nitrogen source

In addition to carbon source, nitrogen source is a crucial element of culture media that
profoundly affects the biosurfactant (including biodemulsifier) production [165]. However,
according to literature, there is no unique nitrogen source conducive to the synthesis and
demulsifying performance of all biodemulsifiers. A synergistic effect between carbon and
nitrogen sources is usually encountered. For example, in the production of biodemulsifier
from Alcaligenes sp. S-XJ-1, high demulsifying ability was achieved when ammonium nitrate
as the nitrogen source was used together with paraffin as the hydrophobic carbon source,
48

whereas a poor demulsification activity was observed when ammonium nitrate was utilized
with sodium citrate as the hydrophilic substrate [165]. Nevertheless, apart from
soluble/insoluble carbon and organic/inorganic nitrogen sources, this synergism can be
generalized to additional medium nutrients, such as phosphorous and metal ions, for
enhancement of both biosynthesis and demulsifying productivity of biodemulsifiers through
identification of optimum medium composition [169].

The biosurfactant-producing microorganisms are seemingly selective, or exhibit a preference,


toward the nature of nitrogen source. For instance, Brevibacilis brevis HOB1 revealed a
larger reduction in surface tension of medium with sodium nitrate than ammonium nitrate
[170], whereas Alcaligenes sp. S-XJ-1 demonstrated a greater reduction in surface tension
with ammonium nitrate than sodium nitrate [165]. It was, therefore, realized that ammonium
ions hindered the consumption of nitrate ions for the production of biosurfactant from
Brevibacilis brevis HOB1 strain due to the less decrease in surface tension of medium when
ammonium nitrate was utilized as the nitrogen source [170].

3.5. pH

Impacts of variation in medium pH on both microbial cell growth and surface tension
reduction of medium are of crucial importance. Neutral (pH = 7.0) or near-neutral
environment has often been reported for the production of various biodemulsifiers
[168,169,171-174]. Amirabadi et al. [173] recognized that maximum production of
biodemulsifier from Paenibacillus alvei ARN63 was obtained at medium pH = 7.0, and pH
values larger than 7.0 were unable to provide an appropriate condition for biodemulsifier
production. On the other hand, Liu and colleagues [165,175] suggested that an alkaline
environment of the primary culture medium supported the growth of Alcaligenes sp. S-XJ-1,
and advanced its demulsifying ability. Additionally, optimization of demulsifying ability may
be implemented by adjusting the pH value of targeted emulsion. It was found that the
increase of pH (to 10.0) of a waste crude oil emulsion remarkably promoted the dewatering
capability of biodemulsifier rhamnolipid, and attained separated phases with higher clarity
since rhamnolipid could be more hydrophilic in an alkaline environment, favoring the
destabilization of the W/O type emulsion, while reducing pH inhibited the water separation
[176]. Commensurately, it may be expected that certain biodemulsifiers express further
hydrophobicity in acidic surroundings, helpful for the breakup of O/W emulsion systems; as
in an earlier study concerning food products, a specific microbial strain could demulsify all
49

respective milk proteins-stabilized O/W emulsions at pH = 3.0 among tested pHs in the range
of 3-7 [177].

3.6. Thermal treatment

Although elevations of emulsion temperature bring about some useful effects leading to a
faster coalescence of dispersed phase droplets (as discussed in Section 2.5.2), there is often
an optimal incubation temperature at which maximum biodemulsifier production and highest
demulsification rate are gained. Some investigators have found optimum incubation
temperature of their demulsifying microbes in the range of 35-40 °C [165,173,178]. It was
observed that a further increase in temperature of the cell-free broth (supernatant) of
demulsifying strain P. alvei beyond its optimum temperature (37 °C) up to 80 °C did not
reveal a noticeable fall in its demulsification performance and surface activity for breaking a
heavy crude oil emulsion, indicative of thermal stability of the related biodemulsifier [173].
On the other hand, Li et al. [168] viewed that the extracellular biodemulsifier of Bacillus
mojavensis XH1 maintained a high demulsifying ratio when exposed to temperatures below
75 °C, but its demulsifying ratio at higher temperatures (from 75 °C to 120 °C) was
drastically declined to a value less than half that obtained before further increase of
temperature. Moreover, biodemulsifiers usually respond in a different manner from one
another when undergo a harsh thermal treatment, depending on their tolerance or stability
against variations of temperature. For example, the biodemulsifying capabilities of a mixed
bacterial culture [159] and the culture of Pseudomonas aeruginosa MSJ [178] were
deteriorated and completely destroyed, respectively, by autoclaving process (conducted at
121 °C), yet this treatment left no effect on the demulsification activity of bacterium
Rhodococcus sp. PR-1 [179].

3.7. Culture age

In the literature, it has not been reported a consistent trend for the relationship between
demulsification activity and culture age. For instance, an increase in demulsification activity
by a mixed microbial culture in a model petroleum emulsion was observed with increasing
the age of culture [159], while demulsification activity from the whole culture of Nocardia
amarae fluctuated with culture age [144]. Furthermore, Nadarajah et al. [159] suggested that
the demulsification activity of a mixed microbial culture was more dependent on microbial
counts (representing a positive correlation) rather than inoculum age. On the other hand, the
age-dependent demulsification performance of whole cultures of Nocardia amarae did not
50

always parallel the age-dependent variations in biomass [144]. It is generally believed that
cell surface hydrophobicity improves with culture age [180]. The larger the degree of
hydrophobicity, the higher the rate of demulsification of O/W emulsions, arising from the
stronger interaction between the cells and oil particles. On the contrary, younger hydrophilic
cells are most beneficial for demulsification of W/O emulsions [180]. Despite this argument,
microbial demulsifiers produced from Alcaligenes sp. S-XJ-1 [165] and from Rhodococcus
sp. PR-1 [179], possessing hydrophobic cell surfaces, were claimed to be useful in breaking
W/O emulsions. These conflicting outcomes are presumably originated from differing growth
conditions of the reported bacteria.

Apart from the above-mentioned parameters, agitation speed of demulsifier in emulsion,


residence time for phase separation and settling, phase volume ratio (sample
volume:emulsion volume), and reaction mixture volume are among the major factors
affecting the emulsion breaking efficiency [147,181].

3.8. Mechanism

Demulsification mechanism caused by biological agents has not yet been fully elucidated.
Explicitly, biodemulsifiers are adsorbed at the oil-water interface, and displace the
emulsifiers so as to destabilize emulsions. However, what active substances basically
contribute to improved biodemulsifying efficiency and how the composition of those
substances is affected are the matters demanding thorough, persuasive investigations. In
general, microorganisms are deemed to modify emulsion properties by employing their
hydrophobic cell surfaces or the ambivalent hydrophobic/hydrophilic feature of
biosurfactants to replace the emulsifying agents present at the oil-water interface [159]. The
prevalent lipophilic moiety of biosurfactants is the hydrocarbon chain of a fatty acid; the
hydrophilic moiety stems from the ester or alcohol functional groups of neutral lipids, the
carboxylate group of fatty acids or amino acids, the phosphate-containing segments of
phospholipids, and the carbohydrates of glycolipids [155].

In terms of functionality, biodemulsifiers are categorized into cellular (cell-bound) and


extracellular metabolic biodemulsifiers. Demulsification ability in cellular biodemulsifying
agents, such as those produced by Nocardia [144], Corynebacterium [145,182], Micrococcus
[152], Rhodococcus [179], Ochrobactrum anthropi [181], Alcaligenes [171], and hybrid
bacteria [159,163], resides in the bacterial cell surface, whereas demulsifying capability in
extracellular biodemulsifiers, including acetoin [167], fengycin [173], protein and lipopeptide
51

[168], and fatty acids and carbohydrates [178], is present in their culture supernatant. Cellular
demulsifiers have been found more promising than extracellular ones for demulsification of
emulsions, due to possession of more intricate demulsification mechanism [160].
Nevertheless, from an industrial prospect, extracellular metabolites released into the culture
medium are appealing owing to the simplicity of their recovery process [183].

The cell wall-bound biodemulsifiers identified so far include proteins [175], mycolic acid
[179], and fatty acid components [161], or composite substances such as protein-lipid-
carbohydrate [151]. The demulsifying activity of bacterial cells is a joint action between
profuse cell wall-related compounds and functional groups, structure of demulsifying cells,
and cell surface physicochemical features such as hydrophobicity and charge of cell surface
[151,175]. As mentioned above, cell surface chemical composition of demulsifying strains is
dependent on the nature of carbon source on which they have been grown. The
demulsification mechanism is, thus, tightly linked to the sort of substrate selected. In this
respect, Huang and coworkers [160] employed particular analytical tools to characterize the
cell surface chemical compositions of the demulsifying microbe Alcaligenes sp. S-XJ-1
cultivated with different carbon sources, and they witnessed the superior demulsification ratio
of a model W/O emulsion on cell wall-associated proteins or lipids, present on the cells
grown on alkane or fatty acid esters, respectively, than that on polysaccharide existing on the
cells cultivated with glucose. However, some researchers discovered that, in W/O emulsions,
the demulsifying effectiveness of biomass grown on alkane was better than that cultivated
with fatty acid ester [160,169]. On the other hand, another study argued that use of vegetable
oils, as cheap carbon sources, with a higher saturation degree and larger contents of C 16 and
C18 fatty acids was advantageous for the synthesis of greatly efficient biodemulsifiers whose
cell wall was, as a result, bound with fatty acids that could partially form cell surface
hydrophobicity and consequently facilitate the oil dewatering process [161].

Besides, optimization of microbial demulsification of complex W/O petroleum emulsions


might be accomplished by considering such parameters as cell surface hydrophobicity as well
as cell surface charge [145]. Both hydrophobicity and electrokinetic potential of cell surface
governing cell adhesion are synchronically impressed by cultivation conditions [184]. In line
with this statement, it was realized that greater cell surface hydrophobicity and a lower
negative charge of an alkaliphilic demulsifying bacterium, cultured under optimal conditions,
promoted its microbial adhesion to hydrocarbon as well as its adsorption at the oil-water
interface, leading to enhanced W/O emulsion breaking ability [175].
52

Incidentally, further probes on various bacteria with variant growth conditions are essential to
illuminate the effects of their cell surface hydrophobicity and surface charge on biological
destabilization of O/W emulsions as well since such studies concerning this type of
emulsions have rarely been reported.

Some of the current biodemulsifier-producing microorganisms are summarized in Table 3.


Table 3
Summary of some current biodemulsifier-producing microorganisms applied for demulsification
Microorganism Isolated Substrate Nitrogen Emulsion(s) Emulsion Operational Microorganism Outcomes Reference
from source treated preparation conditions characteristics
Alcaligenes sp. Petroleum- Liquid NH4NO3 Model W/O Kerosene:water CC a = 500 mg High Demulsification [185]
S-XJ-1 polluted soil paraffin = 2:3; L-1; hydrophobicity ratio b of 81.3%
Distilled water T = 35 °C of cell surface; within 24 h RRC c :
(containing Presence of [151,160,161,16
Span 80) mixed amphiphilic 5,171,175,186]
with aviation compounds in
kerosene the cell walls
(containing
Tween 80)
Mixed bacterial Petroleum- Light sour - Model W/O Kerosene:water CC = 10% (v/v); 24 old Demulsification [159]
culture d contaminate mix crude = 70:30; T = 50 °C inoculum; activity e of 96%
d soil oil Water Demulsifying in 24 h, offering RRC: [163]
stabilized with activity economically
hydrophilic principally feasible
surfactant related to cells; demulsification
(Tween 80), Superior compared with
and kerosene demulsification pure bacterial
stabilized with capability of cultures
lipophilic floating cells
surfactant (due to their
(Span 80) hydrophobicity)
Various - CC = 5 mL than that of Demulsification
oilfield W/O biodemulsifier pelleted cells activity of 99%,
added to 45 g of and water and
each field solid contents of
emulsion emulsions
sample; decreased to 1%
T = 50 °C or lower in 72 h
Micrococcus sp. - n- - Model O/W Stabilized with CC = 0.2-4.0 mg 140 h old; An increase of [152]
tetradecane Tween 60-Span mL-1; Washed with cell
60 surfactants T = 28-30 °C kerosene concentration
from 2 to 4 mg
mL-1 reduced the
54

emulsion’s half-
life by 78%
Model W/O Stabilized with 140 h old; An increase in
Poloxamer L92 unwashed cells cell
surfactant applied concentration
above 2.0 mg
mL-1 could not
decrease the
emulsion’s half-
life to a lower
value;
Unwashed cells
found more
effective for
demulsification
Rhodococcus sp. Sewage of - - Model W/O - T = 55 °C Hydrophobic Complete [179]
PR-1 Dagang surfaces of demulsification f
oilfield, bacterial cells achieved in 8 h; RRC: [187]
China characterized by Freezing/thawing
mycolic acid and autoclaving
with carbon had no effect on
chain length of demulsification
27-54 process, but
ultrasonic
disposal and
contact with
organic solvent
inhibited its
activity
Ochrobactrum Oil-polluted Ethanol NH4NO3; Model Crude CC = 3 mL Whole cells Demulsification [181]
anthropi RIPI5-1 sandy bank NH3+; W/O/W oil:produced sample added to presented high activity for whole
of Siri NO32- water = 1:1 6 mL emulsion; demulsifying culture and
Island, Iran (v/v); T = 50 °C ability during whole cells was
Crude oil late-exponential 63% and 72%,
stabilized with growth and respectively,
Span 80, and stationary after 8.5 h
produced water phases, while
with Tween 80 whole culture
had high
55

demulsifying
activity during
early-
exponential
growth phase;
Demulsifying
activity of whole
cells clearly
correlated with
cell surface
hydrophobicity
Pseudomonas Petroleum- Diesel oil NH4NO3; Model W/O Kerosene:water CC = 40 mg wet The extracellular Demulsifying [178]
aeruginosa MSJ contaminate Yeast = 2:1 (v/v); cells / 300 μL biodemulsifier activity of cells
d soil extract Emulsion culture composed of and supernatants
stabilized with supernatant fatty acids and was 90.3% and
Tween 80 added to 2 ml carbohydrates; 95.3%,
emulsion; A positive respectively,
T = 35 °C correlation after 72 h
Model O/W Kerosene:water observed Demulsifying
= 1:1 (v/v); between activity of cells
Emulsion demulsification and supernatants
stabilized with performance and was 96.7% and
Span 80-Tween cell surface 99.7%,
80 hydrophobicity respectively,
of the strain after 72 h
Streptomyces sp. Antarctica Soluble (NH4)2SO4 Model O/W Kerosene:water CC = 108 spores Demulsification Spores from solid [174]
AA8321 and coastal starch = 2:1 (v/v), per mL associated with culture showed
regions of emulsified by emulsion; surface demulsifying
Korea Triton X-100 T = Room hydrophobicity ability, but
temperature of aerial spores submerged
culture did not;
aerial spores
displayed 95%
demulsification
after 1 min.
56

Bacillus Petroleum- Glucose NH4Cl; Model O/W Optimal ratio CC = 2 mL The extracellular Absence of yeast [168]
mojavensis XH1 contaminate and liquid Yeast of biodemulsifier biodemulsifier extract,
d soil paraffin extract kerosene:water solution added contained two ammonium RRC: [188]
= 65:35 (v/v); to 5 ml components, a chloride or
Span 60 emulsion; protein and a phosphate
dissolved in 1 T = 35 °C lipopeptide inhibited
L kerosene, and biodemulsifier’s
Tween 60 activity;
dissolved in 1 Under optimal
L water production
medium
conditions,
demulsifying
ratio was 35.5%,
and
biodemulsifier
yield 2.07 g L-1
a
CC: cell concentration
b

c
RRC: relevant research conducted
d
Mixed bacterial culture is comprising of isolates Acinetobacter calcoaceticus, A. calcoaceticus BVALC, Acinetobacter radioresistans, Alcaligenes latus, Pseudomonas
aeruginosa, Pseudomonas carboxydohydrogena, Sphingobacterium thalophilum, Kingella denitrificans and Rhodococcus globerulus.

f
Complete demulsification is related to the complete separation of oil and water phases in the emulsion, or 100% demulsification activity
4. Membrane demulsification

Exploitation of membrane separation process for oily wastewater treatment has been
triggered since 1973. The uniformity of the permeate (treated water) quality irrespective of
influent variations, absence of extraneous chemicals, high separation efficiency, useful for
separation of micron and submicron droplets, compact setting, and lower energy cost than
thermal treatments are perceived as the main benefits of membrane treatment. Nonetheless,
high capital costs for very large effluent volumes, fairly low permeate flux, fouling and
degradation of polymeric membranes during function, and their sensitivity to polar and
chlorinated solutions are the constraints to the wide-ranging membrane applications [189-
191]. In membrane demulsification, the core emphasis has been given to optimization of
permeate flux, separation efficiency, and minimization of membrane fouling.

Membrane filtration processes involve rejection of some components in a solution by the


membrane. Prior to reaching the steady state, the convective flow of these components
oriented toward the membrane surface is larger than that caused by diffusion backflow
directed to the bulk solution. As a result, build-up of the rejected components occurs at the
membrane surface. This recognized phenomenon is regarded as concentration polarization
bringing about several negative impacts on membrane operations, such as increase in osmotic
pressure, decline in transmembrane flux, and variations in rejection properties [192,193].
Low transmembrane pressure (TMP) plus slow flow rates of the emulsion passing across the
membrane are the grounds for the formation of concentration polarization [194,195].

As such, it is necessary to increase the backward transport of solute to the bulk solution by
some ways so as to reduce this build-up at the membrane surface. The most common means
of tackling this difficulty is to configure the membrane module in order that the flow of feed
solution is tangential to the surface of membrane; the flow can be either turbulent or high-
shear-stress laminar [192].

Ultrafiltration (UF) and reverse osmosis (RO) have been found effective for treatment of
soluble oil emulsions. In addition to high oil rejection capacity, these types of membranes
have good performance in rejection of total organic carbon (TOC), and reduction of chemical
oxygen demand (COD) and total surface charge (TSC) from the O/W emulsions [1,196,197].
Outwardly, microfiltration (MF) membranes have more frequently been used for resolving
O/W emulsions than W/O emulsions [198-200], while porous glass membranes have more
58

often been exploited to treat W/O emulsions than wastewater [201-203]. Demulsification by
the use of ion-exchange resin membrane was also proposed several decades ago [204].

4.1. Separation efficiency

The emulsified drop rejection features of a membrane rely principally on the membrane
nature (van der Waals forces, Donnan exclusion effect, etc.), membrane material, pore size
distribution, operational conditions (cross-flow velocity, TMP, temperature, etc.), and the
capillary pressure of oil droplets in its pores [198,205]. The oil rejection coefficient, Ro, (or
separation efficiency) in percentage for the membrane separation of O/W emulsions is
calculated using Eq. (5):

(5)

The oil-water separation efficiency is absolutely linked to the properties of membrane


materials, which hold different surface characteristics including surface tension, surface
charge, surface energy, and ion sorption in aqueous environment. For instance, it was
reported that the surface of alumina membranes had a strong tendency to adsorb
hydrocarbons or oil, whereas zirconia membranes expressed a weak adsorption, which is
attributed to rather high surface polarity of the latter weakening its adsorption capability
[206].

4.2. Permeate flux

Effect of TMP on permeate flux (given by Eq. (6)) is explained as follows: initially, with
increasing the pressure, the transmembrane flux also increases until a maximum flux is
reached, due to an increase in the driving force across the membrane; further increase of
pressure then causes the flux to decrease until the flux becomes independent of pressure. The
collapse of flux with increasing pressure is because of the enhanced resistance of the
thickening gel layer (as a result of concentration polarization) formed by further
accumulation of solute on the membrane surface, and also the penetration of larger number of
droplets into the membrane pores evincing pore plugging at a faster rate; however, once the
solute’s convective mass transfer to the membrane surface is balanced by its diffusive flow
back to the bulk solution, the permeate flux has attained steady state condition at which it is
unfettered by pressure, and is solely controlled by the rate of backward diffusive transport
from its surface to the solution [192,205,207-209].
59

(6)

Where,

J: permeate (or transmembrane) flux


Vp: permeate volume
A: effective membrane area
t: experimental time required for permeation

Superior flux stability may be achieved by modification of membrane operating conditions,


such as cross-flow velocity, TMP, feed temperature, and so on [198,210]. Larger permeate
flux has been observed for membranes with higher surface porosity (more number of pores),
larger pore size, lower feed oil concentration, greater hydrophilicity, intermittent operation,
higher temperature, and/or enhanced cross-flow rate (CFR) [190,194,198,208,209]. Yet,
increasing CFR at constant TMP diminishes the percentage of oil rejection because the oil
droplets are fragmented at higher CFR, ending in oil diffusion into the pores together with
permeate. The overall outcome of operating at greater CFR is the lessening of concentration
polarization effect [209].

Clearly, the membranes with larger pore size acquire higher permeate flux, essentially as a
consequence of higher permeability and greater dissidence with both creation of
concentration polarization layer and obstruction of smaller pores [194]. In contrast, better
demulsification efficiency was obtained through membranes with smaller pore size, which
subsequently required higher TMP [200,201].

4.3. Membrane fouling

Fouling is a complicated event arising from adsorption of feed solution components on the
membrane surface and in its pores (pore plugging) [193]. The following points are some
indications of membrane fouling or loss of permeability [199,205]:

 Fouled membranes illustrate critical surface tensions far lower than virgin membranes,
owing to the adhesion of the macromolecules, such as oil and surfactants (especially oil),
to the membrane surface.
 Fouled membranes are less hydrophilic (more oil wet) than the unused ones. In other
words, fouling modifies the wettability of membrane.
60

 At some pores, the operating (transmembrane) pressure surpasses the capillary pressure
so that the oil drops can be distorted and enter to the membrane pores. This usually
happens at high operating pressures where concentration polarization enhances the
probability of emulsified oil globules contacting the pores, thereby reducing the flux
below the value found for lower operating pressures.

The total resistance towards permeation is the amalgamation of resistances of membrane


hydraulic, gel/cake layer, osmotic pressure, and reversible and irreversible fouling [193,211].
Reversible fouling may be removed by backflushing/backwashing the membrane after a
typical trial, whereas irreversible fouling due to pore adsorption may not be readily resolved
by cleaning procedures.

Surface coating of membrane applied to tune its hydrophilicity, and change of the operation
mode (e.g., from parallel-flow to cross-flow) are the means reported to minimize/eliminate
the fouling effect [210,212]. However, high hydrophilicity of a membrane for breakup of an
O/W emulsion, for instance, may not always guarantee its lack of fouling by oil droplets
[208].

4.4. Membrane cleaning

Cleaning or regeneration of membrane can be obtained through backwash with hot water and
hot alkali solution, alkaline wash followed by acid wash, or the micellar solution of sodium
dodecyl sulphate-n-pentanol-water system to maintain its consistent separation performance
[195,198,205]. However, the porosity, pore size, and pore surface area of certain membrane
substances may change if subjected to harsh cleaning solutions. It was found that a typical
hydrophobic membrane was more alkali-resistant and its pore structure was less affected than
an examined hydrophilic membrane when exposed to a cleaning alkaline solution due to the
non-wetting surface of the former [203].

4.5. Mechanism

Membrane demulsification mechanism is explained by either sieving effect or coalescence.


The sieving effect is based on the difference between the droplet size of emulsion and pore
size of membrane, whereas coalescence occurs on the surface of membrane due to droplet
interactions with membrane material [200,201]. Once the droplets coalesce on the membrane
surface, they may penetrate into the pores after deformation, or simply join the gel layer
formed on the surface, and then migrate through the pores under pressure. Due to the increase
61

in droplets size and surfactant removal, emulsion on the other side of membrane is not
stabilized and two separated liquid phases are produced, and subsequently the water phase is
settled down quickly. Even though the pores are initially filled with the coalesced droplets,
they can be completely pushed out by the advanced continuous phase. Where the membrane
is the coalescing medium, hydrophilic or hydrophobic degree of the membrane surface for
W/O or O/W emulsion, respectively, governs their separation effectiveness.

Some of the existing membranes utilized for demulsification are reviewed in Table 4.
Table 4
Summary of some present membranes used for demulsification

Membrane Type of Fabrication method Pore Emulsion(s) Emulsion Feed oil conc. Operational Outcomes Reference
material membrane size treated preparation conditions
Hydrophilic UF a Immersion - Synthetic Machine oil-in- 800 ppm TMP b = 0.1 MPa; Ro d > 99%; [213]
cellulose hollow precipitation O/W deionised water CFR c = 54 L h-1; Final oil conc. = 3.3
fiber (no surfactant T = 20 ºC ppm;
added) Fouling-resistant;
Tolerant to wide pH
range (1-14)
ZrO2 membrane MF e Particle sintering 0.2 µm Industrial - 5 g L-1 TMP = 0.1-0.15 J g = 1500 L m-2 h-1 [206]
on asymmetric O/W MPa; bar-1;
Al2O3 support (steelworks) CFV f = 3-5 m s-1 Ro = 99.8%;
High, stable flux
due to slight fouling
Polysulfone UF Immersion 3.62 nm Synthetic Crude oil-in- 100 ppm TMP = 103.4 kPa; J = 126.1 L m-2 h-1; [209]
precipitation O/W water (no CFR = 3.1 E-6 m3 s- Ro = 94.4%;
1
surfactant ; Oil conc. in
added) pH = 6.12; permeate < 10 ppm
T = 25 ºC
Industrial - 366 ppm pH = 8.73; J = 128 L m-2 h-1;
O/W T = 25 ºC Ro = 77.5%;
(produced Oil conc. in
water) permeate = 50 ppm
Ceramic- UF Immersion 5 µm Synthetic SDS i-stabilized 100 ppm TMP = 0.4 MPa; J = 190 L m-2 h-1; [214]
supported precipitation and O/W oil-in-water Ro > 98.5%;
polymer interfacial Oil conc. in
composite h polymerization permeate < 1.6
ppm;
Fouling-resistant;
Steady flux
Modified MF Nano-TiO2 coating 0.2 µm Synthetic Hydraulic oil < 4 g L-1 CFV = 5 m s-1; Weakened fouling; [210]
commercial O/W 32-in-water TMP = 0.16 MPa; Filtrate oil conc. <
ceramic with stabilized by T = 40 ºC 10 ppm
hydrophilic nano- Tween 80 and
TiO2 coating Span 80
Hydrophilic NaA MF In-situ hydrothermal 1.2 µm Synthetic Tween 80- 100 ppm TMP = 50 kPa Ro > 99%; [195]
zeolite on porous synthesis O/W stabilized Oil conc. in
ceramic tube lubricant oil-in- permeate < 1 ppm;
water J = 85 L m-2 h-1;
63

Stable against oil


fouling and cleaning
solution
Shirasu porous Porous glass Leaching 0.86 µm Synthetic Brine-in- - TMP = 392 kPa; Maximum [202]
glass W/O kerosene T = 293 K demulsification
containing a efficiency of 91%
lipophilic
surfactant
(tetraglycerin
condensed
ricinoleic acid
ester)
Janus Hybrid Self-initiated 10 mm Synthetic Various oil-in- - TMP = 0.09 MPa High separation [215]
polymer/CNT j photografting and O/W water emulsions efficiency and
photopolymerization stabilized by promising flux for
Tween 80 (oil the examined O/W
phase: toluene, and W/O emulsions
chloroform, or due to anisotropic
hexane) wettability of the
Synthetic Various water- membrane (e.g.,
W/O in-oil emulsions filtrate oil purity for
stabilized by water-in-toluene
Span 80 (oil emulsion >
phase: toluene, 99.98%);
chloroform, or Selective removal
hexane) of organic solvents
from water with
high sorption
capacity and good
recyclability
a
UF: ultrafiltration
b
TMP: transmembrane pressure
c
CFR: cross-flow ratio
d
Ro: oil rejection coefficient
e
MF: microfiltration
f
CFV: cross-flow velocity
g
J: permeate flux
h
porous ceramic membrane substrate + polyvinylidene fluoride (PVDF) ultrafiltration sub-layer (polymer) + polyamide/polyvinyl alcohol (PVA) composite thin top-layer
i
SDS: sodium dodecyl sulfate
j
CNT: carbon nanotube
5. Electrical demulsification

Electrical demulsification is generally considered advantageous from the perspective of


energy consumption as compared with other demulsification approaches, such as heating or
centrifugation, in addition to ecological validity [71,216]. However, the drawback of this
process is formation of fine secondary droplets during the coalescence, making it more
burdensome to separate the produced smaller water drops. This effect is turned out either due
to imposing too high electric field on the W/O emulsion, or when the dispersed aqueous
drops have been enlarged too much [8]. There are two competing processes to determine
whether these secondary droplets are formed; one is necking as a result of droplet
deformation due to electric field, and the other is pumping of the content of the water drops
into the oil phase, which is driven by the IFT of the droplet liquid. The ratio of electrical
stress energy required for the former process to the energy needed for the latter one,
introduced by the product of Weber and Ohnesorge dimensionless numbers, has been
reported to be a useful tool for optimization of the design of electro-coalescers with regard to
liquids characteristics, exerted electric potential, and electrode structure and geometry [8].

The technologies and mechanisms associated with electrostatic enhancement of coalescence


of W/O emulsions have already been reviewed by researchers [217,218].

5.1. Influential features

The destabilization efficiency of W/O emulsions subjected to alternating current (AC)


electric field is dependent upon various factors including temperature, shear rate, water
fraction, oil characteristics, emulsion residence time, and the features of electric field applied
(such as field strength, duration, waveform, and frequency). It has been observed that
elevating temperature, strength (voltage), duration, or frequency of the imposed electric field
led to noticeably enhanced dehydration performance of model W/O emulsion [19,216,219-
221]. However, demulsification productivity of such emulsion substantially changed with
different waveforms in declining order of square, sinusoidal and triangular [219].

The optimum frequency is typically influenced by the dielectrics, rheological and electrical
characteristics of the dispersed phase, as well as the size of droplets in emulsion [222,223].
Primary size of droplets in an emulsion significantly affects the optimum frequency of an
electric field exerted for emulsion separation. In this respect, non-homogenized emulsions
containing bigger droplet size require much lower optimal frequency to be demulsified in
65

comparison with homogenized emulsions having small, narrowly distributed dispersed


droplets [224]. The frequency magnitude may prominently depend on the electric field
strength; the choice of optimum frequency is particularly essential at low voltages [19].
However, with regard to the growth of droplets diameter during the dehydration process of
crude oil emulsions and variation of optimum frequency of an electric field with droplets
size, Zhang et al. [224] discovered that in the case of varying frequency, an appropriate
selection and combination of frequencies resulted in improved dehydration performance of
the pulsed direct current (DC) electric field in comparison with the most effective (optimal)
frequency when it was kept constant.

In contrast to large voltages, large frequencies do not usually influence on increment in


droplets distortion and emulsion resolution [19]. The droplets may even not have adequate
time to respond to rapid variations in alternating current at too high frequencies, giving rise to
formation of shorter chains of drops [225]. Crude oil emulsions are far more rapidly (and
perhaps irreversibly) flocculated when subjected to certain electric field voltage owing to
greater number of droplets and hence lower distance between them, leading to reduction in
emulsion viscosity [19,226]. That is why low concentrated emulsions typically undergo
laminar or turbulent flow by internal mixing to initially aid in approaching of droplets and
forming chains [19]. The crude oil viscosity, principally owing to its paraffinic content and
asphaltene, is a dominant, representative factor in evaluation of crude oil electrostatic
dehydration and desalting efficiencies as it reflects both chemical (paraffinic content,
asphaltene, and metal, such as Fe and Ni, fractions) and electrical (crude oil conductivity)
effects of the crude oil samples in all bench, pilot-plant and industrial scales [222,227-229].
The oil viscosity of W/O emulsions under electric field affects the coalescence rest time of
water droplets. The higher the oil viscosity, the longer the coalescence rest time; this is
because the impact velocity of the falling water drops in high-viscosity oil is lower than that
in low-viscosity oil [220].

Contrary to W/O emulsions, electrical demulsification has seldom been exploited to


destabilize O/W emulsions. Utilization of high electric field to O/W emulsions undeniably
initiates the electrolysis of both the aqueous phase and the electrodes, resulting in the
contamination of chemicals existing in the emulsion [230]. Voltages higher than optimum
voltage for breaking O/W emulsions may also trigger their re-emulsificaion [231].
Conversely, high electric field has commonly been applied to demulsify W/O emulsions and
low electric field was proved ineffective in separation of W/O systems [217,230].
66

Efficiency and effect of electrostatic coalescence of O/W emulsions are dependent on type
and concentration of their chemical constituents (such as oil, electrolyte, and surfactant
applied as emulsifier) and water content. Ichikawa et al. [230] employed a low external
electric field (1-10 V/cm) to demulsify some model O/W emulsions and discovered that:

 Increasing the electrolyte concentration or decreasing the ionic surfactant concentration


enhanced the electrical demulsification;
 O/W emulsions stabilized by a nonionic polymer surfactant were not electrically
demulsified since they are stabilized by steric effects caused by the adsorption of polymer
molecules on the surface of droplets, but not by the electrostatic interactions developed
from the surface charges of emulsion particles;
 Dense O/W emulsions (low water content) were quickly demulsified, whereas dilute
(high water content) emulsions demulsified solely near an electrode after being
compacted by electrophoresis of the oil droplets, indicating that the reciprocal contact of
the drops across the thin intervening water phase is crucial for an instant demulsification.

5.2. Mechanism

Varied mechanisms have been described for the electrostatic separation of W/O emulsions.
An external electric field can promote the coalescence not only between the dispersed
droplets, but also between droplets and interface. In general, a three-step coalescence
mechanism is introduced: approaching of dispersed droplets, interfacial film thinning, and
ultimately film breakup leading to coalescence of droplets [18,217]. The overall governing
step is often the film thinning process which is appreciably improved applying an effectual
electric field [8]. Other mechanisms proposed are chain formation of droplets, formation of
intermolecular bonds, dipole-dipole coalescence, electrophoresis, dielectrophoresis, random
collisions, and electrorefining [217]. Besides optical microscopy and coalescence
measurement techniques, electrical conductivity is considered as a useful tool to distinguish
the water droplets coalescence from their chain formation when subjected to high voltage
gradients, as related to the crude oil-water interfacial rheological features; incompressible oil-
water interface hampers the droplet-droplet coalescence and establishes chains of water
droplets (high conductivity), yet compressible interface contributes to an efficient
coalescence followed by marginal conduction [232]. Rayat and Feyzi [233] have recently
proposed a modified thermodynamic model introducing dipole-induced dipole attraction
force in electrical demulsification of W/O emulsions, taking into account both water droplet
67

size and the adsorbed layer thickness of asphaltenic film around the water globules, and
found it to be a more realistic model which could predict the electrical field strength, required
for emulsion breakup, remarkably lower than that of the unmodified model which considered
the dipole-dipole interactions irrespective of mutual induction, as evaluated in an earlier
report [234]. Both AC and pulsed DC modes of electric field entail induced dipoles on the
water drops, resulting in attraction and coalescence between contiguous droplets [224].

The most underlying force for stabilization of charged oil droplets in water is the double layer
force produced by the overlap of diffuse electric double layers of the proximate droplets
[12,235]. The strength of this force is measured by the concentration of ions present in water
plus the surface charge densities of adjoining droplets. The mechanism of electrical
demulsification of O/W emulsions is schematically depicted in Fig. 5 [235]: migration of
surface ions takes place under an external electric field in order that the electrostatic potential
gradient brought about by electric field is mitigated. As a result, the surface charge density at
the right side of the biggest droplet, which is vertical to the applied field direction, becomes
the lowest. As the decrease in surface charge density lessens the height of potential energy
barrier, a small droplet close to the surface with the least surface charge density is firstly
fused with the largest droplet. Fusion of the small drops further enlarges the volume of the
biggest droplet. The charge density on the right hand of the biggest droplet is thus more
reduced. Decrement of the surface charge density diminishes the potential barrier, and speeds
up the coalescence process. The imposed external field consequently demulsifies the
emulsion.

Fig. 5. Schematic illustration of the mechanism of electrical demulsification [235].


68

6. Microwave irradiation

Demulsification by microwave irradiation was seemingly pioneered by Wolf in 1986 [236].


There is ample evidence indicating the superiority of microwave heating over the
conventional heating schemes for breaking emulsions due to faster separation rate and less
energy consumption of the former, which themselves are indebted to high penetration power
of microwaves enabling the “volumetric” and “selective” heating of the dispersed phase
[9,16,65,119,237-241]. The advantages of microwave heating technology help it to be applied
as a standalone pathway for demulsification of emulsions in petroleum industry, by
eliminating or minimizing the consumption of chemical demulsifiers [242,243]. Microwaves
generate electromagnetic waves with wavelength range of 1 mm to 100 cm and frequencies
between 300 GHz and 300 MHz, enabling them to deliver far higher reaction productivities
than traditional heating methods at a specific temperature or even at lower temperatures in a
variety of applications, including inorganic or organic synthesis, food processing,
environmental chemistry, analytical chemistry, and polymerization [238,244]. Likewise,
while applied for demulsification of W/O emulsions, microwave heating may be employed as
a tool to migrate (separate) specific acidic compounds (like naphtenic acids which are
severely corrosive for refinery facilities) from crude oil toward the aqueous phase, provided
the existing operational variables, particularly temperature, are optimized [239,245].

6.1. Effectual factors

Optimization of microwave-induced demulsification of emulsions is typically based on


microwave power and time [5,72,238,246,247]. Ordinarily, the higher the microwave
irradiation power, the more rapidly the emulsion can break and also the greater the phase
separation occurs [65,236]. Nonetheless, this improvement in demulsification efficiency as a
result of increasing the microwave power may result in overheating of O/W systems [65,72].
The power degree for continuous treatment of emulsion is, however, usually from about 1 W
to about 500 W per gallon of emulsion [236]. The time during which emulsions are in contact
with radiation is routinely short, spanning even few seconds. The cases of long radiation (≥
30 min) may bring about excessive energy inputs, thereby prompting turbulence and mixing
of water and oil [248], which are clearly against the emulsion breakup.

Under microwave irradiation treatment, an increase in surfactant concentration, oil-to-water


volumetric ratio, or aging of W/O emulsions, and a reduction in their water content decrease
their separation efficiency and demulsification rate [246,249,250], while an upturn in
69

surfactant concentration or solution pH of O/W emulsions detracts from their demulsification


competence [72,247]. Furthermore, in the mode of combined demulsification utilizing
microwave radiation concurrently with the addition of chemical demulsifier, the longer time
interval between these two treatments would eventuate in decreased separation productivity
of crude oil emulsions [249].

In contrast, Kuo and Lee [72] recognized that microwave-induced separation of an examined
cutting O/W emulsion (50 ml) was enhanced by escalation of salt (NaCl) dosage, settling
time, and initial oil concentration up to their optimal values of 14 g L-1, 1 h, and 10 g L-1,
respectively, at the pH = 9.5, microwave irradiation power of 280 W, and irradiation time of
2 min. In a continuous system, the settling time depends on the applied microwave power,
energy input, and particularly upon the flow rate, since the first two parameters themselves
are linked to the flow rate (turbulence). Binner and co-workers [240] established a lab-scale
continuous microwave system preceding a settling container for the treatment of a model
crude oil emulsion, and discerned that at high flow rates (9-12 L min-1), where turbulence is
dominant, the settling times were independent of energy input, whereas at low flow rates (1-6
L min-1) they decreased upon increasing the power density.

Salt-assisted microwave demulsification of both W/O and O/W emulsions have frequently
been announced in the literature [65,72,247,251,252], mechanism of which is described in
Section 6.3. In this regard, Xia et al. [251] studied the microwave separation of a W/O
emulsion (microwave power: 850 W) in presence and absence of small quantity (0.02 mol L -
1
) of varied inorganic salts, and drew the following conclusions:

 All the Cl- salts (NaCl, KCl, MgCl2, CaCl2), except for the AlCl3, promoted the
demulsification rate and efficiency, resulting in complete separation in 120-150 s;
 The demulsification capabilities of Cl- salts were moderately higher than those of Na+
salts (NaBr, NaI, NaNO3, Na2CO3, Na2SO4);
 The temperature of emulsions in the presence of inorganic salts was greater than that in
the absence of inorganic salts;
 The sample temperature rose faster in the presence of inorganic salts.

In the case of varying electrolyte concentrations, the optimal salinity below and above which
the formation of O/W and W/O emulsions, respectively, is likely has to be taken into
consideration, as discussed in Section 2.5.1 and Section 2.7.
70

One should, however, always take note of the interplay of the operating variables (i.e.,
positive and negative synergisms) involved in the treatment of a typical demulsification
technique including microwave heating. The concept of “synergy” is qualified as a
quantitative/qualitative measure of the combined effect of two (or more) variables on a
system in which the resultant effect may be either below, above, or equal to the additive of
the individual effects of the influential variables under investigation; if the produced effect of
the system is above the related additive measure, this is considered a positive synergy; if it is
less, then it is a negative synergy. This is the definition of synergy from a quantitative point
of view; though, from a qualitative prospect, the positive synergy will be related to the
appearance of new phenomena that are not among those of the interacting systems. On the
other hand, if some of the phenomena of the interacting systems disappear due to the
interaction, then this is called negative synergy [253].

For example, Fortuny et al. [68] pointed out that increasing the pH (from 7 to 12) of various
water-in-crude oil emulsions at high water content (45%), T = 130 ºC and with no salt
content under microwave treatment (t = 15 min) led to highly effective demulsification, yet
no free water was obtained when the high pH value of 12 and the salt content of 30,000 ppm
(other variables remained unchanged) were simultaneously involved, suggesting that the
increment of pH was efficient only if no salt was introduced, disclosing a negative synergy
between the pH and salinity impacts. It is well-established that increase of pH and decrease of
salinity contribute to destabilization of W/O emulsion systems, as elaborated in Section 2.5.1
and Section 2.5.3.

However, the effect of frequency changes from the electromagnetic source on the
demulsification performance of emulsions has not been reported. Deeper studies are,
therefore, still demanded to optimize the operational variables engaged, to shed light on the
detailed mechanism of this burgeoning separation course, and finally to provide oil qualities
that can meet the requirements of oil transportation and refinery thereafter as well as the
water qualities that can comply with discharge rules or re-injection targets.

6.2. Lateral and radial irradiations

Samanta and Basak [254] carried out a theoretical analysis to compare the heating effect of
lateral and radial irradiations on both W/O and O/W emulsion samples confined within 2D
microwave cylinders. They concluded that the selection of lateral/radial irradiation mainly
hinges on the emulsion dielectric properties and the sample size; for example, the emulsion
71

samples (both W/O and O/W) containing smaller size of dispersed phase drops (with fraction
of 0.2) were able to spatially absorb greater average power when exposed to radial irradiation
than those having larger droplet size (with fraction of 0.5), which was attributed to larger
effective dielectric loss of emulsions with lower fraction of dispersed phase (0.2). On the
other hand, lateral irradiation was suited as the optimum heating strategy for the emulsions
with larger sample diameters [244].

6.3. Mechanism

Exposure of W/O emulsions to the microwave electromagnetic field gives rise to molecular
rotation and ionic conduction of dispersed brine phase (but causes no change in the molecular
configuration), reduction of the viscosity of the oil phase (due to increase in temperature),
and improvement of interfacial film thinning progression as a result of the expansion of water
drops arising from their augmented internal pressure during radiation, facilitating the contact
between droplets and subsequently accelerating the demulsification phenomenon and
ameliorating separation productivity. The resultant molecular rotation is accountable for
internal heating, neutralization of the zeta potential of the internal phase, and breaking the
hydrogen bonds between surfactant and water molecules, that all eventuate in the fall of
emulsion stability. However, the conductive migration of dissolved ions (i.e., ionic
conduction) in the emulsion in contact with microwave field also contributes to its internal
heating due to the resistance against the ion flow. As a result, the principal mechanism by
which microwave heating enhances the demulsification capability of emulsions is assigned to
its thermal effects rather than non-thermal. In accordance with this mechanism, saline
emulsions under microwave are separated more quickly since the salts alter the dielectric
properties of the water phase (apart from changing the IFT of emulsion), leading to improved
heating of the dispersed brine compared with pure water, thereby developing a greater
thermal gradient in the continuous phase, which promotes coalescence probability
[68,246,252]. Similar mechanism has also been proposed for O/W emulsions, and the
expedient impact of salts dissolved in the aqueous phase on their resolution is attributed to
the proliferation in solvent conductivity, enabling the acceleration of heating rate as well as
the reduction of electrical double layer thickness at the oil-water interface [65,247,255].
Nonetheless, beyond optimum electrolyte concentration, the ionic conduction and the dipole
rotation of water molecules are incompatibly affected, triggering a decrease in temperature
and consequently declined separation efficacy [247].
72

7. Summary

This paper has strived to have an overview of certain widely applied, on-going individual
demulsification techniques including chemical, biological, and some physical means by
noticing the marked influential factors to either stability or instability of emulsions.
Nevertheless, the impressive factors expatiated for each method are not exclusively confined
to those debated here, and a greater number may be expected in practice. It is worth
reiterating that when optimizing the various operational variables (such as temperature, pH,
salinity, etc.) engaged in a typical demulsification process of an emulsion, their interplay with
each other has to be considered to identify the positive and/or negative synergism(s) incurred
in the relevant procedure (as clarified in Section 6.1). In all the demulsification techniques
employed in petroleum industry, both separation efficiency and treatment rate are taken note
of, and the primary goal of all these methods is to accelerate the interfacial film thinning
process by weakening/draining the stabilizing film encapsulating the dispersed drops and, in
ideal cases, at the same time hampering the adsorption of emulsifier(s) at the oil-water
interface, resulting in the predominance of van der Waals forces of attraction between the
scattered droplets, which eventually prompts the film rupture and droplets coalescence. Each
of these destabilizing approaches has clearly its own advantages and drawbacks as summed
up in Table 5. Although they were not included within the scopes of this review note, the
combinatorial modes of separation are plentifully accessible in the literature, all targeting
excelled demulsification performance relative to each of their associated singly applied
treatment pathways. It is, however, obvious enough that a promising emulsion-breaking route
not only has to mind its robustness and capability for a variety of emulsions, but also has to
take into account the environmental impacts by meeting the legislated ecological rules for the
case of exposure or disposal, and, last but not least, has to minimize the economic burden
imposed on the oil industry.
Table 5
Advantages and disadvantages of the demulsification methods overviewed

Demulsification method Advantages Disadvantages

Chemical Tunable structure of demulsifier for the targeted emulsion; Chemical contamination of emulsions;
Fast separation Hazardous to human and/or environment if disposed;
Impedes subsequent emulsification of emulsions;
Chemicals might be expensive

Biological Environmental compatibility; Might be sensitive/vulnerable to operational/treatment


Biodegradability; conditions
Unique properties that chemicals may not have
Reusable without considerable loss of activity

Membrane Suitable for the separation of micron or submicron droplets; Concentration polarization;
No chemical contamination of emulsions; Fouling;
High oil removal efficiency; Requires periodic cleaning process;
Lower energy cost than thermal treatments Might be vulnerable to acidic/alkaline cleaning solutions;
Relatively low permeate flux;
High capital cost for large effluent volumes

Electrical No chemical contamination of emulsions; Confined to low water concentration of W/O emulsions;
Environmental credibility; Formation of secondary droplets in emulsions
Lower energy cost than conventional heating & centrifugation

Microwave irradiation Fast heating of emulsion in contact; Might be necessary to add chemical demulsifier to reach the
Volumetric heating of dispersed phase; desired separation efficiency
Selective heating of dispersed phase;
Decreases the need for high heating energy, unlike conventional heating;
No chemical contamination of emulsions;
Does not need high electric voltage
Acknowledgements

The research financial supports (SGRA and IGRF) awarded by the Universiti Putra Malaysia
are gratefully acknowledged.

References

[1] H. Zhang, S. Fang, C. Ye, M. Wang, H. Cheng, H. Wen, X. Meng, Treatment of waste
filature oil/water emulsion by combined demulsification and reverse osmosis, Sep. Purif.
Technol. 63 (2008) 264-268.

[2] J. Zhang, J. Li, R.W. Thring, X. Hu, X. Song, Oil recovery from refinery oily sludge via
ultrasound and freeze/thaw, J. Hazard. Mater. 203-204 (2012) 195-203.

[3] W. Zhang, P. Xiao, D. Wang, Central treatment of different emulsion wastewaters by an


integrated process of physicochemically enhanced ultrafiltration and anaerobic-aerobic
biofilm reactor, Bioresour. Technol. 159 (2014) 150-156.

[4] V. Rajaković, D. Skala, Separation of water-in-oil emulsions by freeze/thaw method and


microwave radiation, Sep. Purif. Technol. 49 (2006) 192-196.

[5] S. Wang, X. Xu, J. Yang, J. Gao, Effect of the carboxymethyl chitosan on removal of
nickel and vanadium from crude oil in the presence of microwave irradiation, Fuel Process.
Technol. 92 (2011) 486-492.

[6] J.J. Sheng, Modern Chemical Enhanced Oil Recovery: Theory and Practice, Gulf
Professional Publishing, New York, 2010.

[7] http://energy.gov/fe/science-innovation/oil-gas-research/enhanced-oil-recovery, access


date: May 29, 2016 (time: 7:05 PM).

[8] M. Mousavichoubeh, M. Shariaty-Niassar, M. Ghadiri, The effect of interfacial tension on


secondary drop formation in electro-coalescence of water droplets in oil, Chem. Eng. Sci. 66
(2011) 5330-5337.

[9] I.N. Evdokimov, A.P. Losev, Microwave treatment of crude oil emulsions: Effects of
water content, J. Petrol. Sci. Eng. 115 (2014) 24-30.

[10] http://www.indexmundi.com/commodities/glossary/light-vs-heavy-crude-oil, access


date: April 15, 2016 (time: 2:10 PM).

[11] D.G. Thompson, A.S. Taylor, D.E. Graham, Emulsification and demulsification related
to crude oil production, Colloids Surf. 15 (1985) 175-189.

[12] S.S. Dukhin, J. Sjӧblom, Ø. Sӕther, An experimental and theoretical approach to the
dynamic behavior of emulsions, in: J. Sjӧblom (Ed.), Emulsions and Emulsion Stability:
Surfactant Science Series/61, CRC Press, Boca Raton, 2005, pp. 1-101.
75

[13] Y. Kawashima, T. Hino, H. Takeuchi, T. Niwa, K. Horibe, Shear-induced phase


inversion and size control of water/oil/water emulsion droplets with porous membrane, J.
Colloid Interface Sci. 145 (1991) 512-523.

[14] J.-L. Salager, Formulation concepts for the emulsion maker, in: F. Nielloud, G. Marti-
Mestres (Eds.), Pharmaceutical Emulsions and Suspensions, Marcel Dekker, New York,
2000, pp. 19-72.

[15] S. Mukherjee, A.P. Kushnick, Effect of demulsifiers on interfacial properties governing


crude oil demulsification, in: J.K. Borchardt, T.F. Yen (Eds.), Oilfield Chemistry Enhanced
Recovery and Production Stimulation, Am. Chem. Soc., Washington, DC, 1989, pp. 364-374.

[16] L. Xia, S. Lu, G. Cao, Stability and demulsification of emulsions stabilized by


asphaltenes or resins, J. Colloid Interface Sci. 271 (2004) 504-506.

[17] G. Hu, J. Li, G. Zeng, Recent development in the treatment of oily sludge from
petroleum industry: A review, J. Hazard. Mater. 261 (2013) 470-490.

[18] M. Elektorowicz, S. Habibi, R. Chifrina, Effect of electrical potential on the electro-


demulsification of oily sludge, J. Colloid Interface Sci. 295 (2006) 535-541.

[19] G. Rodionova, S. Keleşoğlu, J. Sjöblom, AC field induced destabilization of water-in-oil


emulsions based on North Sea acidic crude oil, Colloids Surf. A 448 (2014) 60-66.

[20] N.H. Roodbari, A. Badiei, E. Soleimani, Y. Khaniani, Tweens demulsification effects on


heavy crude oil/water emulsion, Arab. J. Chem. (2011).
http://dx.doi.org/10.1016/j.arabjc.2011.08.009.

[21] J.D. McLean, P.K. Kilpatrick, Effects of asphaltene solvency on stability of water-in-
crude-oil emulsions, J. Colloid Interface Sci. 189 (1997) 242-253.

[22] J.J. Oren, G.D.M. Mackay, Electrolyte and pH effect on emulsion stability of water-in-
petroleum oils, Fuel 56 (1977) 382-384.

[23] M.K. Poindexter, S.C. Marsh, Inorganic solid content governs water-in-crude oil
emulsion stability predictions, Energy Fuels 23 (2009) 1258-1268.

[24] X. Yang, W. Tan, Y. Bu, Demulsification of asphaltenes and resins stabilized emulsions
via the freeze/thaw method, Energy Fuels 23 (2009) 481-486.

[25] R.A. Mohammed, A.I. Bailey, P.F. Luckham, S.E. Taylor, Dewatering of crude oil
emulsions 2. Interfacial properties of the asphaltic constituents of crude oil, Colloids Surf. A
80 (1993) 237-242.

[26] P.M. Spiecker, K.L. Gawrys, C.B. Trail, P.K. Kilpatrick, Effects of petroleum resins on
asphaltene aggregation and water-in-oil emulsion formation, Colloids Surf. A 220 (2003) 9-
27.

[27] B.F. Graham, E.F. May, R.D. Trengove, Emulsion inhibiting components in crude oils,
Energy Fuels 22 (2008) 1093-1099.
76

[28] P.-C. Schorling, D.G. Kessel, I. Rahimian, Influence of the crude oil resin/asphaltene
ratio on the stability of oil/water emulsions, Colloids Surf. A 152 (1999) 95-102.

[29] B.E. Chistyakov, Theory and practical application aspects of surfactants, in: D. Möbius,
R. Miller, V.B. Fainerman (Eds.), Surfactants: Chemistry, Interfacial Properties,
Applications, Elsevier, Amsterdam, 2001, pp. 511-618.

[30] T.C. Kowalski, R.W. Pike, Mixture of gelatin, weighting agent and chelate compound,
US Patent No. 6,326,335 B1 (2001).

[31] N.H. Zaid, Dispersed oil soluble corrosion inhibitor and water soluble phosphonate scale
inhibitor composition, US Patent No. 4,713,184 A (1987).

[32] A. Fakhru’l-Razi, A. Pendashteh, L.C. Abdullah, D.R.A. Biak, S.S. Madaeni, Z.Z.
Abidin, Review of technologies for oil and gas produced water treatment, J. Hazard. Mater.
170 (2009) 530-551.

[33] Z. Youyi, H. Qingfeng, J. Guoqing, M. Desheng, W. Zhe, Current development and


application of chemical combination flooding technique, Petrol. Explor. Develop. 40 (2013)
96-103.

[34] M. Dong, S. Ma, Q. Liu, Enhanced heavy oil recovery through interfacial instability: A
study of chemical flooding for Brintnell heavy oil, Fuel 88 (2009) 1049-1056.

[35] N. Yan, J.H. Masliyah, Demulsification of solids-stabilized oil-in-water emulsions,


Colloids Surf. A 117 (1996) 15-25.

[36] N. Yan, C. Kurbis, J.H. Masliyah, Continuous demulsification of solids-stabilized oil-in-


water emulsions by the addition of fresh oil, Ind. Eng. Chem. Res. 36 (1997) 2634-2640.

[37] A.P. Sullivan, P.K. Kilpatrick, The effects of inorganic solid particles on water and
crude oil emulsion stability, Ind. Eng. Chem. Res. 41 (2002) 3389-3404.

[38] B.P. Binks, S.O. Lumsdon, Influence of particle wettability on the type and stability of
surfactant-free emulsions, Langmuir 16 (2000) 8622-8631.

[39] N.D. Denkov, I.B. Ivanov, P.A. Kralchevsky, D.T. Wasan, A possible mechanism of
stabilization of emulsions by solid particles, J. Colloid Interface Sci. 150 (1992) 589-593.

[40] N.S. Ahmed, A.M. Nassar, N.N. Zaki, H.K. Gharieb, Stability and rheology of heavy
crude oil-in-water emulsion stabilized by an anionic-nonionic surfactant mixture, Petrol Sci.
Technol. 17 (1999) 553-576.

[41] Z. Zhang, G.Y. Xu, F. Wang, S.L. Dong, Y.M. Li, Characterization and demulsification
of poly(ethylene oxide)–block–poly(propylene oxide)–block–poly(ethylene oxide)
copolymers, J. Colloid Interface Sci. 277 (2004) 464-470.

[42] W.C. Griffin, Classification of surface-active agents by HLB, J. Soc. Cosmet. Chem. 1
(1949) 311-326.
77

[43] W.D. Bancroft, The theory of emulsification, V, J. Phys. Chem. 17 (1913) 501-519.

[44] H.T. Davis, Factors determining emulsion type: hydrophile-lipophile balance and
beyond, Colloids Surf. A 91 (1994) 9-24.

[45] M. Rondón, P. Bouriat, J. Lachaise, J.-L. Salager, Breaking of water-in-crude oil


emulsions. 1. Physicochemical phenomenology of demulsifier action, Energy Fuels 20 (2006)
1600-1604.

[46] W.C. Griffin, Calculation of HLB values of non-ionic surfactants, J. Soc. Cosmet.
Chem. 5 (1954) 249-256.

[47] J. Wu, Y. Xu, T. Dabros, H. Hamza, Effect of demulsifier properties on destabilization


of water-in-oil emulsion, Energy Fuels 17 (2003) 1554-1559.

[48] I. Kailey, X. Feng, Influence of structural variations of demulsifiers on their


performance, Ind. Eng. Chem. Res. 52 (2013) 785-793.

[49] Y. Xu, J. Wu, T. Dabros, H. Hamza, Optimizing the polyethylene oxide and
polypropylene oxide contents in diethylenetriamine-based surfactants for destabilization of a
water-in-oil emulsion, Energy Fuels 19 (2005) 916-921.

[50] J. Wu, Y. Xu, T. Dabros, H. Hamza, Effect of EO and PO positions in nonionic


surfactants on surfactant properties and demulsification performance, Colloids Surf. A 252
(2005) 79-85.

[51] P.A. Winsor, Solvent Properties of Amphiphilic Compounds, Butterworths Scientific,


London, 1954.

[52] A. Goldszal, M. Bourrel, Demulsification of crude oil emulsions: Correlation to


microemulsion phase behavior, Ind. Eng. Chem. Res. 39 (2000) 2746-2751.

[53] A.A. Pena, G. Hirasaki, C. Miller, Chemically induced destabilization of water-in-crude-


oil, Ind. Eng. Chem. Res. 44 (2005) 1139-1149.

[54] K. Shinoda, H. Arai, The correlation between phase inversion temperature in emulsion
and cloud point in solution of nonionic emulsifier, J. Phys. Chem. 68 (1964) 3485-3490.

[55] D. Nguyen, N. Sadeghi, C. Houston, Chemical interactions and demulsifier


characteristics for enhanced oil recovery applications, Energy Fuels 26 (2012) 2742-2750.

[56] J. Boyd, C. Parkinson, P. Sherman, Factors affecting emulsion stability, and the HLB
concept, J. Colloid Interface Sci. 41 (1972) 359-370.

[57] J.-L. Salager, Microemulsions, G. Broze (Ed.), Handbook of Detergents: Properties,


Marcel Dekker, New York, 1999, pp. 253-302.

[58] J.-L. Salager, Emulsion phase inversion phenomena, in: J. Sjӧblom (Ed.), Emulsions and
Emulsion Stability, CRC Press, New York, 2006, pp. 185-226.
78

[59] J.-L. Salager, N. Marquez, A. Graciaa, J. Lachaise, Partitioning of ethoxylated


octylphenol surfactants in microemulsion−oil−water systems: Influence of temperature and
relation between partitioning coefficient and physicochemical formulation, Langmuir 16
(2000) 5534-5539.

[60] M. Bourrel, J.-L. Salager, R.S. Schechter, W.H. Wade, A correlation for phase behavior
of nonionic surfactants, J. Colloid Interface Sci. 75 (1980) 451-461.

[61] J.L. Salager, J.C. Morgan, R.S. Schechter, W.H. Wade, E. Vasquez, Optimum
formulation of surfactant/water/oil systems for minimum interfacial tension or phase
behavior, Soc. Pet. Eng. J. 19 (1979) 107-115.

[62] R.E. Anton, N. Garces, A. Yajure, A correlation for three-phase behavior of cationic
surfactant-oil-water systems, J. Dispersion Sci. Technol. 18 (1997) 539-555.

[63] J.-L. Salager, The fundamental basis for the action of a chemical dehydrant: Influence of
the physical and chemical formulation on the stability of an emulsion, Int. Chem. Eng. 30
(1990) 103-116.

[64] J.L. Salager, M. Mińana-Perez, M. Pérez-Sánchez, M. Ramirez-Gouveia, C.I. Rojas,


Surfactant-oil-water systems near the affinity inversion Part III: The two kinds of emulsion
inversion, J. Dispersion Sci. Technol. 4 (1983) 313-329.

[65] R. Martínez-Palou, R. Cerón-Camacho, B. Chavéz, A.A. Vallejo, D. Villanueva-


Negrete, J. Castellanos, J. Karamath, J. Reyes, J. Aburto, Demulsification of heavy crude oil-
in-water emulsions: A comparative study between microwave and thermal heating, Fuel 113
(2013) 407-414.

[66] A. Bera, A. Mandal, B.B. Guha, Synergistic effect of surfactant and salt mixture on
interfacial tension reduction between crude oil and water in enhanced oil recovery, Chem.
Eng. Data 59 (2014) 89-96.

[67] A.G. Gaonkar, Effects of salt, temperature, and surfactants on the interfacial tension
behavior of a vegetable oil/water system, J. Colloid Interface Sci. 149 (1992) 256-260.

[68] M. Fortuny, C.B.Z. Oliveira, R.L.F.V. Melo, M. Nele, R.C.C. Coutinho, A.F. Santos,
Effect of salinity, temperature, water content, and pH on the microwave demulsification of
crude oil emulsions, Energy Fuels 21 (2007) 1358-1364.

[69] M. Moradi, V. Alvarado, S. Huzurbazar, Effect of salinity on water-in-crude oil


emulsion: Evaluation through drop-size distribution proxy, Energy Fuels 25 (2011) 260-268.

[70] X. Wang, A. Brandvik, V. Alvarado, Probing interfacial water-in-crude oil emulsion


stability controls using electrorheology, Energy Fuels 24 (2010) 6359-6365.

[71] C.E. Perles, P.L.O. Volpe, A.J.F. Bombard, Study of the cation and salinity effect on
electrocoalescence of water/crude oil emulsions, Energy Fuels 26 (2012) 6914-6924.
79

[72] C.-H. Kuo, C.-L. Lee, Treatment of a cutting oil emulsion by microwave irradiation,
Sep. Sci. Technol. 44 (2009) 1799-1815.

[73] M.K. Sharma, D.O. Shah, Macro- and microemulsions in enhanced oil recovery, in:
D.O. Shah (Ed.), Macro- and Microemulsions Theory and Applications, Am. Chem. Soc.,
Washington, D.C., 1985, pp. 149-172.

[74] C.A. Miller, P. Neogi, Interfacial Phenomena: Equilibrium and Dynamic Effects, second
ed., CRC Press, New York, 2008.

[75] R.A. Mohammed, A.I. Bailey, P.F. Luckhum, S.E. Taylor, Dewatering of crude oil
emulsions 3. Emulsion resolution by chemical means, Colloids Surf. A 83 (1994) 261-271.

[76] N.H. Abdurahman, Y.M. Rosli, N.H. Azhari, B.A. Hayder, Pipeline transportation of
viscous crudes as concentrated oil-in-water emulsions, J. Pet. Sci. Eng. 90-91 (2012) 139-
144.

[77] G.J. Hirasaki, C.A. Miller, O.G. Raney, M.K. Poindexter, D.T. Nguyen, J. Hera,
Separation of produced emulsions from surfactant enhanced oil recovery processes, Energy
Fuels 25 (2011) 555-561.

[78] J.E. Strassner, Effect of pH on interfacial films and stability of crude oil-water
emulsions, J. Pet. Technol. 20 (1968) 303-312.

[79] C. Dalmazzone, C. Noïk, J.-F. Argillier, Impact of chemical enhanced oil recovery on
the separation of diluted heavy oil emulsions, Energy Fuels 26 (2012) 3462-3469.

[80] C.-M. Chen, C.-H. Lu, C.-H. Chang, Y.-M. Yang, J.-R. Maa, Influence of pH on the
stability of oil-in-water emulsions stabilized by a splittable surfactant, Colloids Surf. A 170
(2000) 173-179.

[81] http://petrowiki.org/Stability_of_oil_emulsions, access date: January 5, 2015 (time:


11:15 AM).

[82] B.P. Binks, R. Murakami, S.P. Armes, S. Fujii, Effect of pH and salt concentration on
oil-in-water emulsions stabilized solely by nanocomposite microgel particles, Langmuir 22
(2006) 2050-2057.

[83] I. Silva, B. Borges, R. Blanco, M. Rondón, J.-L. Salager, J.C. Pereira, Breaking of
water-in-crude oil emulsions. 5. Effect of acid-alkaline additives on the performance of
chemical demulsifiers, Energy Fuels 28 (2014) 3587-3593.

[84] N.N. Zaki, M.E. Abdel-Raouf, A.-A.A. Abdel-Azim, Propylene oxide-ethylene oxide
block copolymers as demulsifiers for water-in-oil emulsions, II. Effects of temperature,
salinity, pH-value, and solvents on the demulsification efficiency, Monatshefte für Chemie
127 (1996) 1239-1245.

[85] A.-A.A. Abdel-Azim, M.E.-S. Abdel-Raouf, A.-R.M. Abdul-Raheim, N.E.-S. Maysor,


Sugar-based ethoxylated amine surfactants as demulsifiers for crude oil emulsions: 2-
Demulsification of different types of crudes, Brazil. J. Pet. Gas 4 (2010) 155-165.
80

[86] M.Y. Pletnev, Chemistry of surfactants, in: V.B. Fainerman, D. Möbius, R. Miller
(Eds.), Surfactans: Chemistry, Interfacial Properties, Applications, Elsevier Science,
Amsterdam, 2001, pp. 1-97.

[87] A.M. Al-Sabagh, M.R.N. El-Din, S. Abo-El Fotouh, N.M. Nasser, Investigation of the
demulsification efficiency of some ethoxylated polyalkylphenol formaldehydes based on
locally obtained materials to resolve water-in-oil emulsions, J. Disp. Sci. Technol. 30 (2009)
267-276.

[88] A.M. Al-Sabagh, N.G. Kandile, R.A. El-Ghazawy, M.R.N. El-Din, Synthesis and
evaluation of some new demulsifiers based on bisphenols for treating water-in-crude oil
emulsions, Egypt. J. Pet. 20 (2011) 67-77.

[89] B. Borges, M. Rondón, O. Sereno, J. Asuaje, Breaking of water-in-crude-oil emulsions.


3. Influence of salinity and water-oil ratio on demulsifier action, Energy Fuels 23 (2009)
1568-1574.

[90] O. Mouraille, T. Skodvin, J. Sjöblom, J-L. Peytavy, Stability of water-crude oil


emulsion: role played by the stage of solvation of asphaltenes and by waxes, J. Disp. Sci.
Technol. 19 (1998) 339-367.

[91] V.B. Menon, D.T. Wasan, Demulsification, in: P. Becher (Ed.), Encyclopedia of
Emulsion Technology, Vol. 2: Applications, Marcel Dekker, New York, 1985, pp. 1-76.

[92] N.N. Zaki, Surfactant stabilized oil-in-water emulsions for pipeline transportation of
viscous crude oils, Colloids Surf. A 125 (1997) 19-25.

[93] A.M. Al-Sabagh, N.G. Kandile, M.R.N. El-Din, Functions of demulsifiers in the
petroleum industry, Sep. Sci. Technol. 46 (2011) 1144-1163.

[94] http://petrowiki.org/Oil_demulsification, access date: January 17, 2015 (time: 4:25 PM).

[95] Y.B. Zhou, L. Chen, X.M. Hou, J. Lu, Modified resin coalescer for oil-in-water
emulsion treatment: Effect of operating conditions on oil removal performance, Ind. Eng.
Chem. Res. 48 (2009) 1660-1664.

[96] N.T. Opedal, I. Kralova, C. Lesaint, J. Sjӧblom, Enhanced sedimentation and


coalescence by chemicals on real crude oil systems, Energy Fuels 25 (2011) 5718-5728.

[97] http://petrowiki.org/Oil_emulsions, access date: March 19, 2015 (time: 7:40 PM).

[98] R. Pal, Effect of droplet size on the rheology of emulsions, AIChE J. 42 (1996) 3181-
3190.

[99] A. Le Follotec, I. Pezron, C. Noik, C. Dalmazzone, L. Metlas-Komunjer, Triblock


copolymers as destabilizers of water-in-crude oil emulsions, Colloids Surf. A 365 (2010)
162-170.

[100] R.J. Hunter, Foundations of Colloid Science, Oxford University Press, Oxford, 1987.
81

[101] P. Alexandridis, V. Athanassiou, S. Fukuda, T.A. Hatton, Surface activity of


poly(ethy1ene oxide)-block-poly(propy1ene oxide)-block-poly(ethy1ene oxide) copolymers,
Langmuir 10 (1994) 2604-2612.

[102] Z. Zhou, B. Chu, Light-scattering study on the association behavior of triblock


polymers of ethylene oxide and propylene oxide in aqueous solution, J. Colloid Interface Sci.
126 (1988) 171-180.

[103] P. Linse, M. Malmsten, Temperature-dependent micellization in aqueous block


copolymer solutions, Macromolecules 25 (1992) 5434-5439.

[104] J. Wang, F.L. Hu, C.Q. Li, J. Li, Y. Yang, Synthesis of dendritic polyether surfactants
for demulsification, Sep. Purif. Technol. 73 (2010) 349–354.

[105] P. Alexandridis, J.F. Holzwarth, T.A. Hatton, Micellization of poly(ethylene oxide)-


poly(propylene oxide)-poly(ethylene oxide) triblock copolymers in aqueous solutions:
Thermodynamics of copolymer association, Macromolecules 27 (1994) 2414-2425.

[106] H. Saito, K. Shinoda, The solubilization of hydrocarbons in aqueous solutions of


nonionic surfactants, J. Colloid Interface Sci. 24 (1967) 10-15.

[107] K. Shinoda, T. Ogawa, The solubilization of water in nonaqueous solutions of nonionic


surfactants, J. Colloid Interface Sci. 24 (1967) 56-60.

[108] B.R. Blomqvist, T. Wärnheim. P.M. Claesson, Surface rheology of PEO-PPO-PEO


triblock copolymers at the air-water interface: comparison of spread and adsorbed layers,
Langmuir 21 (2005) 6373-6384.

[109] A. Bhardwaj, S. Hartland, Dynamics of emulsification and demulsification of water in


crude oil emulsions, Ind. Eng. Chem. Res. 33 (1994) 1271-1279.

[110] S. Bratskaya, V. Avramenko, S. Schwarz, I. Philippova, Enhanced flocculation of oil-


in-water emulsions by hydrophobically modified chitosan derivatives, Colloids Surf. A 275
(2006) 168-176.

[111] Z. Zhang, G. Xu, F. Wang, S. Dong, Y. Chen, Demulsification by amphiphilic


dendrimer copolymers, J. Colloid Interface Sci. 282 (2005) 1-4.

[112] K. Tong, Y. Zhang, P.K. Chu, Evaluation of calcium chloride for synergistic
demulsification of super heavy oil wastewater, Colloids Surf. A 419 (2013) 46-52.

[113] A.O. Adilbekova, K.I. Omarova, A. Karakulova, K.B. Musabekov, Nonionic


surfactants based on polyoxyalkylated copolymers used as demulsifying agents, Colloids
Surf. A (2014). http://dx.doi.org/10.1016/j.colsurfa.2014.11.004.

[114] R. Martίnez-Palou, P.F. Sánchez, Perspectives of ionic liquids applications for clean
oilfield technologies, in: A. Kokorin (Ed.), Ionic Liquids: Theory, Properties, New
Approaches, InTech, Rijeka, 2011, pp. 567-630.
82

[115] Y. Sun, L. Shi, Basic ionic liquids with imidazole anion: New reagents to remove
naphthenic acids from crude oil with high total acid number, Fuel 99 (2012) 83-87.

[116] A.Z. Hezave, S. Dorostkar, S. Ayatollahi, M. Nabipour, B. Hemmateenejad,


Investigating the effect of ionic liquid (1-dodecyl-3-methylimidazolium chloride ([C12mim]
[Cl])) on the water/oil interfacial tension as a novel surfactant, Colloids Surf. A 421 (2013)
63-71.

[117] R.K. Goyal, N.S. Jayakumar, M.A. Hashim, Emulsion stabilization using ionic liquid
[BMIM]+[NTf2]− and performance evaluation on the extraction of chromium, J. Hazard.
Mater. 195 (2011) 55-61.

[118] C.A. Flores, E.A. Flores, E. Hernández, L.V. Castro, A. García, F. Alvarez, F.S.
Vázquez, Anion and cation effects of ionic liquids and ammonium salts evaluated as
dehydrating agents for super-heavy crude oil: Experimental and theoretical points of view, J.
Mol. Liq. 196 (2014) 249-257.

[119] D. Guzmán-Lucero, P. Flores, T. Rojo, R. Martínez-Palou, Ionic liquids as demulsifiers


of water-in-crude oil emulsions: Study of the microwave effect, Energy Fuels 24 (2010)
3610-3615.

[120] R.C.B. Lemos, E.B. da Silva, A. dos Santos, R.C.L. Guimarães, B.M.S. Ferreira, R.A.
Guarnieri, C. Dariva, E. Franceschi, A.F. Santos, M. Fortuny, Demulsification of water-in-
crude oil emulsions using ionic liquids and microwave irradiation, Energy Fuels 24 (2010)
4439-4444.

[121] R.A. Mohammed, A.I. Bailey, P.F. Luckham, S.E. Taylor, The effect of demulsifiers
on the interfacial rheology and emulsion stability of water-in-crude oil emulsions, Colloids
Surf. A 91 (1994) 129-139.

[122] W. Kang, G. Jing, H. Zhang, M. Li, Z. Wu, Influence of demulsifier on interfacial film
between oil and water, Colloids Surf. A 272 (2006) 27-31.

[123] M. Razi, M.R. Rahimpour, A. Jahanmiri, F. Azad, Effect of a different formulation of


demulsifiers on the efficiency of chemical demulsification of heavy crude oil, J. Chem. Eng.
Data 56 (2011) 2936-2945.

[124] R. Zhang, C. Liang, D. Wu, S. Deng, Characterization and demulsification of produced


liquid from weak base ASP flooding, Colloids Surf. A 290 (2006) 164-171.

[125] T.J. Jones, E.L. Neustadter, K.P. Whittingham, Water-in-crude oil emulsion stability
and emulsion destabilization by chemical demulsifiers, J. Can. Pet. Technol. (1978).
http://dx.doi.org/10.2118/78-02-08.

[126] B. Wang, T. Wu, Y. Li, D. Sun, M. Yang, Y. Gao, F. Lu, X. Li, The effects of oil
displacement agents on the stability of water produced from ASP
(alkaline/surfactant/polymer) flooding, Colloids Surf. A 379 (2011) 121-126.

[127] M. Li, M. Lin, Z. Wu, A.A. Christy, The influence of NaOH on the stability of
paraffinic crude oil emulsion, Fuel 84 (2004) 183-187.
83

[128] X. Cao, Y. Li, S. Jiang, H. Sun, A. Cagna, L. Dou, A study of dilational rheological
properties of polymers at interfaces, J. Colloid Interface Sci. 270 (2004) 295-298.

[129] T. Sun, L. Zhang, Y. Wang, S. Zhao, B. Peng, M. Li, J. Yu. Influence of demulsifiers
of different structures on interfacial dilational properties of an oil–water interface containing
surface-active fractions from crude oil. J. Colloid Interface Sci. 255 (2002) 241-247.

[130] Y.H. Kim, D.T. Wasan, P.J. Breen, A study of dynamic interfacial mechanisms for
demulsification of water-in-oil emulsions, Colloids Surf. A 95 (1995) 235-247.

[131] D. Daniel-David, I. Pezron, C. Dalmazzone, C. Noїk, D. Clausse, L. Komunjer, Elastic


properties of crude oil/water interface in presence of polymeric emulsion breakers. Colloids
Surf. A 270-271 (2005) 257-262.

[132] D.C. Maia Filho, J.B.V.S. Ramalho, G.M.S. Lucas, E.F. Lucas, Aging of water-in-
crude oil emulsions: Effect on rheological parameters, Colloids Surf. A 405 (2012) 73-78.

[133] A.S. Dukhin, P.J. Goetz, Acoustic and electroacoustic spectroscopy for characterizing
concentrated dispersions and emulsions, Adv. Colloid Interface Sci. 92 (2001) 73-132.

[134] S. Deng, R. Bai, J.P. Chen, G. Yu, Z. Jiang, F. Zhou, Effects of


alkaline/surfactant/polymer on stability of oil droplets in produced water from ASP flooding,
Colloids Surf. A 211 (2002) 275-284.

[135] M.A. Krawczyk, D.T. Wasan, C.S. Shetty, Chemical demulsification of petroleum
emulsions using oil-soluble demulsifiers, Ind. Eng. Chem. Res. 30 (1991) 367-375.

[136] Y.H. Kim, D.T. Wasan, Effect of demulsifier partitioning on the destabilization of
water-in-oil emulsions, Ind. Eng. Chem. Res. 35 (1996) 1141-1149.

[137] D.C. Maia Filho, J.B.V.S. Ramalho, L.S. Spinelli, E.F. Lucas, Aging of water-in-crude
oil emulsions: Effect on water content, droplet size distribution, dynamic viscosity and
stability, Colloids Surf. A 396 (2012) 208-212.

[138] D. Daniel-David, A. Le Follotec, I. Pezron, C. Dalmazzone, C. Noїk, L. Barré, L.


Komunjer, Destabilisation of Water-in-Crude Oil Emulsions by Silicone Copolymer
Demulsifiers, Oil Gas Sci. Technol. 63 (2008) 165-173.

[139] A. Kabalnov, H. Wennerström, Macroemulsion stability: The oriented wedge theory


revisited, Langmuir 12 (1996) 276-292.

[140] E.G. Cockbain, T.S. McRoberts, The stability of elementary emulsion drops and
emulsions, J. Colloid Sci. 8 (1953) 440-451.

[141] A. Upadhyaya, E.J. Acosta, J.F. Scamehorn, D.A. Sabatini, Microemulsion phase
behavior of anionic-cationic mixtures: Effect of tail branching, J. Surfactants Deterg. 9
(2006) 169-179.
84

[142] E.W. Kaler, K.L. Herrington, A.K. Murthy, J.A.N. Zasadzinski, Phase behavior and
structures of mixtures of anionic and cationic surfactants, J. Phys. Chem. 96 (1992) 6698-
6707.

[143] M.J. Rosen, Surfactants and Interfacial Phenomena, third ed., John Wiley & Sons, New
Jersey, 2004.

[144] W.L. Cairns, D.G. Cooper, J.E. Zajic, J.M. Wood, N. Kosaric, Characterization of
Nocardia amarae as a potent biological coalescing agent of water-in-oil emulsions, Appl.
Environ. Microbiol. 43 (1982) 362-366.

[145] A.L. Stewart, N.C.C. Gray, W.L. Cairns, N. Kosaric, Bacteria-induced de-
emulsification of water-in-oil petroleum emulsions, Biotechnol. Lett. 5 (1983) 725-730.

[146] H. Bach, D.L. Gutnick, Potential applications of bioemulsifiers in the oil industry, in:
R. Vazquez-Duhalt, R. Quintero-Ramirez (Eds.), Petroleum Biotechnology, Developments
and Perspectives, Elsevier, Amsterdam, 2004, pp. 233-282.

[147] A. Singh, J.D.V. Hamme, O.P. Ward, Surfactants in microbiology and biotechnology:
Part 2. Application aspects, Biotechnol. Adv. 25 (2007) 99-121.

[148] N. Kosaric, W.L. Cairns, N.C.C. Gray, Microbial de-emulsifiers, in: N. Kosaric, W.L.
Cairns, N.C.C. Gray (Eds.), Biosurfactants and Biotechnology, Marcel Dekker, New York,
1987, pp. 247-321.

[149] D.F. Gerson, The biophysics of microbial surfactants: growth on insoluble substrates,
in: N. Kosaric (Ed.), Biosurfactants: Production Properties Applications, Marcel Dekker,
New York, 1993, pp. 269-286.

[150] N. Kosaric, Biosurfactants in industry, Pure Appl. Chem. 64 (1992) 1731-1737.

[151] X. Huang, K. Peng, Y. Feng, J. Liu, L. Lu, Separation and characterization of effective
demulsifying substances from surface of Alcaligenes sp. S-XJ-1 and its application in water-
in-kerosene emulsion, Bioresour. Technol. 139 (2013) 257-264.

[152] M. Das, Characterization of de-emulsification capabilities of a Micrococcus species,


Bioresour. Technol. 79 (2001) 15-22.

[153] C.A. Boulton, C. Ratledge, Biosynthesis of lipid precursors to surfactant production, in:
N. Kosaric, W.L. Cairns, N.C.C. Gray (Eds.), Biosurfactants and Biotechnology, Marcel
Dekker, New York, 1987, pp. 47-87.

[154] C.R. Macdonald, D.G. Cooper, J.E. Zajic, Surface active lipids from Nocardia
erythropolis grown on hydrocarbons, Applied Environmental Microbiology 41 (1981) 117-
123.

[155] S. Lang, F. Wagner, Structure and properties of biosurfactants, in: N. Kosaric, W.L.
Cairns, N.C.C. Gray (Eds.), Biosurfactants and Biotechnology, Marcel Dekker, New York,
1987, pp. 21-45.
85

[156] C. Syldatk, F. Wagner, Production of biosurfactants, in: N. Kosaric, W.L. Cairns,


N.C.C. Gray (Eds.), Biosurfactants and Biotechnology, Marcel Dekker, New York, 1987, pp.
89-120.

[157] M. Kates, Techniques of Lipidology: Isolation, Analysis, and Identification of Lipids,


Elsevier, New York, 1972.

[158] D.M. Considine (Ed.), Chemical and Process Technology Encyclopedia, Mc Graw-Hill,
New York, 1974.

[159] N. Nadarajah, A. Singh, O.P. Ward, De-emulsification of petroleum oil emulsion by a


mixed bacterial culture, Process Biochem. 37 (2002) 1135-1141.

[160] X. Huang, K. Peng, L. Lu, R. Wang, J. Liu, Carbon source dependence of cell surface
composition and demulsifying capability of Alcaligenes sp. S-XJ-1, Environ. Sci. Technol.
(2014). http://dx.doi.org/10.1021/es404636j.

[161] X.F. Huang, M.X. Li, L.J. Lu, S. Yang, J. Liu, Relationship of cell-wall bound fatty
acids and the demulsification efficiency of demulsifying bacteria Alcaligenes sp. S-XJ-1
cultured with vegetable oils, Bioresour. Technol. 104 (2012) 530-536.

[162] I.M. Banat, Biosurfactants production and possible uses in microbial enhanced oil
recovery and oil pollution remediation: A review, Bioresour. Technol. 51 (1995) 1-12.

[163] N. Nadarajah, A. Singh, O.P. Ward, Evaluation of a mixed bacterial culture for de-
emulsification of water-in-petroleum oil emulsions, World J. Microbiol. Biotechnol. 18
(2002) 435-440.

[164] J.C. Lee, K.Y. Lee, Emulsification using environmental compatible emulsifiers and de-
emulsification using DC field and immobilized Nocardia amarae, Biotechnol. Lett. 22 (2000)
1157-1163.

[165] J. Liu, X.F. Huang, L.J. Lu, J.C. Xu, Y. Wen, D.H. Yang, Q. Zhou, Optimization of
biodemulsifier production from Alcaligenes sp. S-XJ-1 and its application in breaking crude
oil emulsion, J. Hazard. Mater. 183 (2010) 466-473.

[166] I.N.A. Van Bogaert, K. Saerens, C. De Muynck, D. Develter, W. Soetaert, E.J.


Vandamme, Microbial production and application of sophorolipids, Appl. Microbiol.
Biotechnol. 76 (2007) 23-34.

[167] K.L. Janiyani, H.J. Purohit, R. Shanker, P. Khanna, De-emulsification of oil-in-water


emulsions by Bacillus subtilis, World J. Microbiol. Biotechnol. 10 (1994) 452-456.

[168] X. Li, A. Li, C. Liu, J. Yang, F. Ma, N. Hou, Y. Xu, N. Ren, Characterization of the
extracellular biodemulsifier of Bacillus mojavensis XH1 and the enhancement of
demulsifying efficiency by optimization of the production medium composition, Process
Biochem. 47 (2012) 626-634.
86

[169] J. Liu, X.F. Huang, L.J. Lu, J.C. Xu, Y. Wen, D.H. Yang, Q. Zhou, Comparison
between waste frying oil and paraffin as carbon source in the production of biodemulsifier by
Dietzia sp. S-JS-1, Bioresour. Technol. 100 (2009) 6481-6487.

[170] N.I.A. Haddad, J. Wang, B. Mu, Isolation and characterization of a biosurfactant


producing strain, Brevibacilis brevis HOB1, J. Ind. Microbiol. Biotechnol. 35 (2008) 1597-
1604.

[171] X.F. Huang, J. Liu, L.J. Lu, Y. Wen, J.C. Xu, D.H. Yang, Q. Zhou, Evaluation of
screening methods for demulsifying bacteria and characterization of lipopeptide bio-
demulsifier produced by Alcaligenes sp., Bioresour. Technol. 100 (2009) 1358-1365.

[172] X.F. Huang, W. Guan, J. Liu, L.J. Lu, J.C. Xu, Q. Zhou, Characterization and
phylogenetic analysis of biodemulsifier-producing bacteria, Bioresour. Technol. 101 (2010)
317-323.

[173] S.Sh. Amirabadi, A. Jahanmiri, M.R. Rahimpour, B. Rafie nia, P. Darvishi, A. Niazi,
Investigation of Paenibacillus alvei ARN63 ability for biodemulsifier production: Medium
optimization to break heavy crude oil emulsion, Colloids Surf. B 109 (2013) 244-252.

[174] S.H. Park, J.H. Lee, S.H. Ko, D.S. Lee, H.K. Lee, Demulsification of oil-in-water
emulsions by aerial spores of a Streptomyces sp., Biotechnol. Lett. 22 (2000) 1389-1395.

[175] J. Liu, L.J. Lu, X.F. Huang, J.J. Shang, M.X. Li, J.C. Xu, H.P. Deng, Relationship
between surface physicochemical properties and its demulsifying ability of an alkaliphilic
strain of Alcaligenes sp. S-XJ-1, Process Biochem. 46 (2011) 1456-1461.

[176] X. Long, G. Zhang, C. Shen, G. Sun, R. Wang, L. Yin, Q. Meng, Application of


rhamnolipid as a novel biodemulsifier for destabilizing waste crude oil, Bioresour. Technol.
131 (2013) 1-5.

[177] M.H. Ly, M. Aguedo, S. Goudot, M.L. Le, P. Cayot, J.A. Teixeira, T.M. Le, J.M.
Belin, Y. Waché, Interactions between bacterial surfaces and milk proteins, impact on food
emulsions stability, Food Hydrocolloids 22 (2008) 742-751.

[178] J.O.P.A. Coutinho, M.P.S. Silva, P.M. Moraes, A.S. Monteiro, J.C.C. Barcelos, E.P.
Siqueira, V.L. Santos, Demulsifying properties of extracellular products and cells of
Pseudomonas aeruginosa MSJ isolated from petroleum-contaminated soil, Bioresour.
Technol. 128 (2013) 646-654.

[179] T. Ma, F.L. Liang, Y.W. Xi, R.L. Liu, Performance of demulsification by Rhodococcus
sp. PR-1, Environ. Sci. 27 (2006) 1191-1196.

[180] N.C.C. Gray, A.L. Stewart, W.L. Cairns, N. Kosaric, Bacteria induced de-
emulsification of oil-in-water petroleum emulsions, Biotechnol. Lett. 6 (1984) 419-424.

[181] G. Mohebali, A. Kaytash, N. Etemadi, Efficient breaking of water/oil emulsions by a


newly isolated de-emulsifying bacterium, Ochrobactrum anthropi strain RIPI5-1, Colloids
Surf. B 98 (2012) 120-128.
87

[182] Z. Duvnjak, N. Kosaric, Deemulsification of petroleum w/o emulsions by selected


bacterial and yeast cells, Biotechnol. Lett. 9 (1987) 39-42.

[183] S.B. Batista, A.H. Mounteer, F.R. Amorim, M.R. Tótola, Isolation and characterization
of biosurfactant/bioemulsifier-producing bacteria from petroleum contaminated sites,
Bioresour. Technol. 97 (2006) 868-875.

[184] M.C.M. van Loosdrecht, J. Lyklema, W. Norde, G. Schraa, A.J.B. Zehnder,


Electrophoretic mobility and hydrophobicity as a measure to predict the initial steps of
bacterial adhesion, Appl. Environ. Microbiol. 53 (1987) 1898-1901.

[185] Y. Wen, H. Cheng, L.J. Lu, J. Liu, Y. Feng, W. Guan, Q. Zhou, X.F. Huang, Analysis
of biological demulsification process of water-in-oil emulsion by Alcaligenes sp. S-XJ-1,
Bioresour. Technol. 101 (2010) 8315-8322.

[186] J. Liu, K. Peng, X. Huang, L. Lu, H. Cheng, D. Yang, Q. Zhou, H. Deng, Application
of waste frying oils in the biosynthesis of biodemulsifier by a demulsifying strain Alcaligenes
sp. S-XJ-1, J. Environ. Sci. 23 (2011) 1020-1026.

[187] M.A. Wilkinson, D.G. Cooper, Testing of microbial demulsifiers with heavy crude
emulsions, Biotechnol. Lett. 7 (1985) 406-408.

[188] N. Hou, D. Li, F. Ma, J. Zhang, Y. Xu, J. Wang, C. Li, Effective biodemulsifier
components secreted by Bacillus mojavensis XH-1 and analysis of the demulsification
process, Biodegradation (2013). http://dx.doi.org/10.1007/s10532-013-9679-5.

[189] M. Cheryan, N. Rajagopalan, Membrane processing of oily streams. Wastewater


treatment and waste reduction, J. Membr. Sci. 151 (1998) 13-28.

[190] A.C. Hong, A.G. Fane, R.P. Burford, The effects of intermittent permeate flow and
crossflow on membrane coalescence of oil-in-water emulsions, Desalination 144 (2002) 185-
191.

[191] J. Zhong, X. Sun, C. Wang, Treatment of oily wastewater produced from refinery
processes using flocculation and ceramic membrane filtration, Sep. Purif. Technol. 32 (2003)
93-98.

[192] E. Matthiasson, B. Sivik, Concentration polarization and fouling, Desalination 35


(1980) 59-103.

[193] M.K. Ko, J.J. Pellegrino, Determination of osmotic pressure and fouling resistance and
their effects of performance of ultrafiltration membranes, J. Membr. Sci. 74 (1992) 141-157.

[194] R.D. Colle, C.A. Fortulan, S.R. Fontes, Manufacture of ceramic membranes for
application in demulsification process for cross-flow microfiltration, Desalination 245 (2009)
527-532.

[195] J. Cui, X. Zhang, H. Liu, S. Liu, K.L. Yeung, Preparation and application of
zeolite/ceramic microfiltration membranes for treatment of oil contaminated water, J. Membr.
Sci. 325 (2008) 420-426.
88

[196] P. Lipp, C.H. Lee, A.G. Fane, C.J.D. Fell, A fundamental study of the ultrafiltration of
oil-water emulsions, J. Membr. Sci. 36 (1988) 161-177.

[197] M. Karhu, T. Kuokkanen, J. Rämö, M. Mikola, J. Tanskanen, Performance of a


commercial industrial-scale UF-based process for treatment of oily wastewaters, J. Environ.
Manage. 128 (2013) 413-420.

[198] M. Hlavacek, Break-up of oil-in-water emulsions induced by permeation through a


microfiltration membrane, J. Membr. Sci. 102 (1995) 1-7.

[199] F.F. Nazzal, M.R. Wiesner, Microfiltration of oil-in-water emulsions, Water Environ.
Res. 68 (1996) 1187-1191.

[200] N.M. Kocherginsky, C.L. Tan, W.F. Lu, Demulsification of water-in-oil emulsions via
filtration through a hydrophilic polymer membrane, J. Membr. Sci. 220 (2003) 117-128.

[201] D. Sun, X. Duan, W. Li, D. Zhou, Demulsification of water-in-oil emulsion by using


porous glass membrane, J. Membr. Sci. 146 (1998) 65-72.

[202] M. Kukizaki, M. Goto, Demulsification of water-in-oil emulsions by permeation


through Shirasu-porous-glass (SPG) membranes, J. Membr. Sci. 322 (2008) 196-203.

[203] C. Liu, T.-M. Xiao, J. Zhang, L. Zhang, J.-L. Yang, M. Zhang, Effect of membrane
wettability on membrane fouling and chemical durability of SPG membranes used in a
microbubble-aerated biofilm reactor, Sep. Purif. Technol. (2014).
http://dx.doi.org/10.1016/j.seppur.2014.03.001.

[204] K. Shinoda, Demulsification by the use of ion-exchange resin membrane, J. Colloid


Sci. 16 (1961) 85-87.

[205] S. Lee, Y. Aurelle, H. Roques, Concentration polarization, membrane fouling and


cleaning in ultrafiltration of soluble oil, J. Membr. Sci. 19 (1984) 23-38.

[206] C. Yang, G. Zhang, N. Xu, J. Shi, Preparation and application in oil-water separation of
ZrO2/α-Al2O3 MF membrane, J. Membr. Sci. 142 (1998) 235-243.

[207] T. Mohammadi, M. Kazemimoghadam, M. Saadabadi, Modeling of membrane fouling


and flux decline in reverse osmosis during separation of oil in water emulsions, Desalination
157 (2003) 369-375.

[208] A. Rezvanpour, R. Roostaazad, M. Hesampour, M. Nystrӧm, C. Ghotbi, Effective


factors in the treatment of kerosene-water emulsion by using UF membranes, J. Hazard.
Mater. 161 (2009) 1216-1224.

[209] B. Chakrabarty, A.K. Ghoshal, M.K. Purkait, Cross-flow ultrafiltration of stable oil-in-
water emulsion using polysulfone membranes, Chem. Eng. J. 165 (2010) 447-456.
89

[210] Q. Chang, J. Zhou, Y. Wang, J. Liang, X. Zhang, S. Cerneaux, X. Wang, Z. Zhu, Y.


Dong, Application of ceramic microfiltration membrane modified by nano-TiO2 coating in
separation of a stable oil-in-water emulsion, J. Membr. Sci. 456 (2014) 128-133.

[211] H. Falahati, A.Y. Tremblay, Flux dependent oil permeation in the ultrafiltration of
highly concentrated and unstable oil-in-water emulsions, J. Membr. Sci. 371 (2011) 239-247.

[212] L. Yang, A. Thongsukmak, K.K. Sirkar, K.B. Gross, G. Mordukhovich, Bio-inspired


onboard membrane separation of water from engine oil, J. Membr. Sci. 378 (2011) 138-148.

[213] H.-J. Li, Y.-M. Cao, J.-J. Qin, X.-M. Jie, T.-H. Wang, J.-H. Liu, Q. Yuan,
Development and characterization of anti-fouling cellulose hollow fiber UF membranes for
oil-water separation, J. Membr. Sci. 279 (2006) 328-335.

[214] W. Shu, C. Liangyin, C. Wenmei, Fouling-resistant composite membranes for


separation of oil-in-water microemulsions, Chinese J. Chem. Eng. 14 (2006) 37-45.

[215] J. Gu, P. Xiao, J. Chen, J. Zhang, Y. Huang, T. Chen, Janus polymer/carbon nanotube
hybrid membranes for oil/water separation, ACS Appl. Mater. Interfaces 6 (2014) 16204-
16209.

[216] C.-M. Lee, G.W. Sams, J.P. Wagner, Power consumption measurements for ac and
pulsed dc for electrostatic coalescence of water-in-oil emulsions, J. Electrostat. 53 (2001) 1-
24.

[217] J.S. Eow, M. Ghadiri, A.O. Sharif, T.J. Williams, Electrostatic enhancement of
coalescence of water droplets in oil: a review of the current understanding, Chem. Eng. J. 84
(2001) 173-192.

[218] J.S. Eow, M. Ghadiri, Electrostatic enhancement of coalescence of water droplets in


oil: a review of the technology, Chem. Eng. J. 85 (2002) 357-368.

[219] C. Lesaint, W.R. Glomm, L.E. Lundgaard, J. Sjӧblom, Dehydration efficiency of AC


electrical fields on water-in-model-oil emulsions, Colloids Surf. A 352 (2009) 63-69.

[220] M. Chiesa, S. Ingebrigtsen, J.A. Melheim, P.V. Hemmingsen, E.B. Hansen, Ø. Hestad,
Investigation of the role of viscosity on electrocoalescence of water droplets in oil, Sep. Purif.
Technol. 50 (2006) 267-277.

[221] P. Suemar, E.F. Fonseca, R.C. Coutinho, F. Machado, R. Fontes, L.C. Ferreira, E.L.
Lima, P.A. Melo, J.C. Pinto, M. Nele, Quantitative evaluation of the efficiency of water-in-
crude-oil emulsion dehydration by electrocoalescence in pilot-plant and full-scale units, Ind.
Eng. Chem. Res. 51 (2012) 13423-13437.

[222] J.S. Eow, M. Ghadiri, Drop-drop coalescence in an electric field: the effects of applied
electric field and electrode geometry, Colloids Surf. A 219 (2003) 253-279.

[223] S.B. Sample, B. Raghupathy, C.D. Hendricks, Quiescent distortion and resonant
oscillations of a liquid drop in an electric field, Int. J. Eng. Sci. 8 (1970) 97-109.
90

[224] Y. Zhang, Y. Liu, R. Ji, F. Wang, B. Cai, H. Li, Application of variable frequency
technique on electrical dehydration of water-in-oil emulsion, Colloids Surf. A 386 (2011)
185-190.

[225] T.Y. Chen, R.A. Mohammed, A.I. Bailey, P.F. Luckham, S.E. Taylor, Dewatering of
crude oil emulsions 4. Emulsion resolution by the application of an electric field, Colloids
Surf. A 83 (1994) 273-284.

[226] A.E. Bresciani, C.F.X. de Mendonça, R.M.B. Alves, C.A.O. Nascimento, Modeling of
kinetics of water droplets coalescence in crude oil emulsion subjected to an electrical field,
10th International Symposium on Process Systems Engineering - PSE2009, 1947-1952.

[227] R.C.C. Coutinho, J.C. Pinto, M. Nele, A. Hannisdal, J. Sjӧblom, Evaluation of water-
in-crude-oil emulsion stability using critical electric field: Effect of emulsion preparation
procedure and crude oil properties, J. Disp. Sci. Technol. 32 (2011) 923-934.

[228] G. Liu, X. Xu, J. Gao, Study on the compatibility of asphaltic crude oil with the electric
desalting demulsifiers, Energy Fuels 17 (2003) 543-548.

[229] G. Liu, X. Xu, J. Gao, Study on the compatibility of high-paraffin crude oil with
electric desalting demulsifiers, Energy Fuels 17 (2003) 625-630.

[230] T. Ichikawa, K. Itoh, S. Yamamoto, M. Sumita, Rapid demulsification of dense oil-in-


water emulsion by low external electric field I. Experimental evidence, Colloids Surf. A 242
(2004) 21-26.

[231] M. Hosseini, M.H. Shahavi, Electrostatic Enhancement of coalescence of oil droplets


(in nanometer scale) in water emulsion, Chin. J. Chem. Eng. 20 (2012) 654-658.

[232] S.E. Taylor, Investigations into the electrical and coalescence behavior of water-in-
crude oil emulsions in high voltage gradients, Colloids Surf. 29 (1988) 29-51.

[233] K. Rayat, F. Feyzi, Influence of external electric field on the polarity of water droplets
in water-in-oil emulsion phase transition, Colloids Surf. A 375 (2011) 61-67.

[234] F.L.M.C. Silva, F.W. Tavares, M.J.E.M. Cardoso, Phase transition of water–in–oil
emulsions over influence of an external electric field, Colloids Surf. A 326 (2008) 10-17.

[235] T. Ichikawa, Electrical demulsification of oil-in-water emulsion, Colloids Surf. A 302


(2007) 581-586.

[236] N.O. Wolf, Use of microwave radiation in separating emulsions and dispersions of
hydrocarbons and water, US Patent No. 4,582,629 (1986).

[237] C.S. Fang, B.K.L. Chang, P.M.C. Lai, W.J. Klaila, Microwave demulsification, Chem.
Eng. Commun. 73 (1988) 227-239.

[238] S. Wang, J. Yang, X. Xu, Effect of the cationic starch on removal of Ni and V from
crude oils under microwave irradiation, Fuel 90 (2011) 987-991.
91

[239] E.B. da Silva, D. Santos, M.P. de Brito, R.C.L. Guimarães, B.M.S. Ferreira, L.S.
Freitas, M.C.V. de Campos, E. Franceschi, C. Dariva, A.F. Santos, M. Fortuny, Microwave
demulsification of heavy crude oil emulsions: Analysis of acid species recovered in the
aqueous phase, Fuel 128 (2014) 141-147.

[240] E.R. Binner, J.P. Robinson, S.W. Kingman, E.H. Lester, B.J. Azzopardi, G. Dimitrakis,
J. Briggs, Separation of oil/water emulsions in continuous flow using microwave heating,
Energy Fuels 27 (2013) 3173-3178.

[241] S.-H. Kim, S.Y. Li, G.-R. Yi, D.J. Pine, S.-M. Yang, Microwave-assisted self-
organization of colloidal particles in confining aqueous droplets, J. Am. Chem. Soc. 128
(2006) 10897-10904.

[242] P. Parvasi, A.K. Hesamedini, A. Jahanmiri, M.R. Rahimpour, A novel modeling and
experimental study of crude oil desalting using microwave, Sep. Sci. Technol. 49 (2014)
1029-1044.

[243] A.H. Nour, R.M. Yunus, A.H. Nour, Demulsification of water-in-oil emulsions by
microwave heating technology, World Academy Sci. Eng. Technol. 4 (2010) 144-149.

[244] S.K. Samanta, T. Basak, Efficient processing of oil–water emulsions confined within
2D cylinders with various microwave irradiations: Role of metallic annulus, Food Res. Int. 43
(2010) 148-166.

[245] S. Mutyala, C. Fairbridge, J.R.J. Paré, J.M.R. Bélanger, S. Ng, R. Hawkins, Microwave
applications to oil sands and petroleum: A review, Fuel Process. Technol. 91 (2010) 127-135.

[246] C.-C. Chan, Y.-C. Chen, Demulsification of w/o emulsions by microwave radiation,
Sep. Sci. Technol. 37 (2002) 3407-3420.

[247] C.-H. Kuo, C.-L. Lee, Treatment of oil/water emulsions using seawater-assisted
microwave irradiation, Sep. Purif. Technol. 74 (2010) 288-293.

[248] C.S. Fang, P.M.C. Lai, Microwave heating and separation of water-in-oil emulsions, J.
Microwave Power E.E. 30 (1995) 46-57.

[249] B.M.S. Ferreira, J.B.V.S. Ramalho, E.F. Lucas, Demulsification of water-in-crude oil
emulsions by microwave radiation: Effect of aging, demulsifier addition, and selective
heating, Energy Fuels 27 (2013) 615-621.

[250] W. Tan, X.-G. Yang, X.-F. Tan, Study on demulsification of crude oil emulsions by
microwave chemical method, Sep. Sci. Technol. 42 (2007) 1367-1377.

[251] L.-X. Xia, S.-W. Lu, G.-Y. Cao, Salt-assisted microwave demulsification, Chem. Eng.
Commun. 191 (2004) 1053-1063.

[252] E.R. Binner, J.P. Robinson, S.A. Silvester, S.W. Kingman, E.H. Lester, Investigation
into the mechanisms by which microwave heating enhances separation of water-in-oil
emulsions, Fuel 116 (2014) 516-521.
92

[253] S.S.E.H. Elnashaie, S.S. Elshishini, Dynamic Modelling, Bifurcation, and Chaotic
Behaviour of Gas-Solid Catalytic Reactors, Gordon and Breach, Amsterdam, 1996.

[254] S.K. Samanta, T. Basak, Efficient microwave processing of oil–water emulsion


cylinders with lateral and radial irradiations, Food Res. Int. 42 (2009) 1337-1350.

[255] C.S. Fang, D.M.C. Lai, B.K.L. Chang, W.J. Klaila, Oil recovery and waste reduction
by microwave radiation, Environ. Prog. 8 (1989) 235-238.
93

Highlights

 The most common demulsification approaches of W/O and O/W emulsions are
overviewed.
 The significant parameters affecting the demulsification efficiency are discussed.
 Surfactant is the major constituent determining the stability of emulsions.
 All demulsification techniques enhance the rate of interfacial film thinning.

You might also like