Water in Diesel

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Egyptian Journal of Petroleum (2013) 22, 517–530

Egyptian Petroleum Research Institute

Egyptian Journal of Petroleum


www.elsevier.com/locate/egyjp
www.sciencedirect.com

FULL LENGTH ARTICLE

Water-in-diesel fuel nanoemulsions: Preparation, stability


and physical properties
M.R. Noor El-Din a, Sabrnal H. El-Hamouly b, H.M. Mohamed a,
Marwa R. Mishrif a, Ahmad M. Ragab a,*

a
Egyptian Petroleum Research Institute (EPRI), 1 Ahmed El-Zomor St., Nasr City, Cairo 11727, Egypt
b
Chemistry Department, Faculty of Science, Menoufia University, Menoufia, Egypt

Received 12 March 2013; accepted 15 April 2013


Available online 5 December 2013

KEYWORDS Abstract In this work, water-in-diesel fuel nanoemulsions were prepared with mixed nonionic sur-
Emulsification; factants. Several mixtures of sorbitan monooleate and polyoxyethylene (20) sorbitan monooleate,
Ostwald ripening; with different Hydrophilic–Lipophilic Balance (HLB) values (9.6, 9.8, 10, 10.2 and 10.4) were pre-
Water-in-diesel fuel nano- pared to achieve the optimal HLB value. Three mixed surfactant concentrations were prepared at
emulsions; 6%, 8% and 10% to identify the optimum concentration. Five emulsions with different water con-
Interfacial tension; tents: 5%, 6%, 7%, 8% and 9% (wt./wt.) were prepared using high energy method at the optimum
Kinematic viscosity; conditions (HLB = 10 and mixed surfactant concentration = 10%). The effect of HLB value,
Density
mixed surfactant concentration and water content on the droplet size has been studied. The inter-
facial tension and thermodynamic properties of the individual and the blended emulsifiers were
investigated. Droplet size of the prepared nanoemulsions was determined by dynamic light scatter-
ing and the nanoemulsion stability was assessed by measuring the variation of the droplet size as a
function of time. From the obtained results, it was found that the mean droplet sizes were formed
between 49.55 and 104.4 nm depending on HLB value, surfactant concentration and water content
of the blended emulsifiers. The physical properties, kinematic viscosity and density, of the prepared
nanoemulsions and the effect of different temperatures on these properties were measured.
ª 2013 Production and hosting by Elsevier B.V. on behalf of Egyptian Petroleum Research Institute.
Open access under CC BY-NC-ND license.

1. Introduction

* Corresponding author. Tel.: +20 222745902, mobile: Nanoemulsions are a type of emulsions with uniform and
+201004995900; fax: +20 2227727433. extremely small droplet size (typically in the range
E-mail address: amragab.epri@yahoo.com (A.M. Ragab). 50–200 nm) or ‘‘milky’’ (up to 500 nm) [1,2]. Unlike micro-
Peer review under responsibility of Egyptian Petroleum Research emulsions, which are thermodynamically stable, nanoemul-
Institute. sions are only kinetically stable [3,4]. However, the long term
persistence of nanoemulsions (with no apparent flocculation
or coalescence) makes them unique, and they are sometimes
Production and hosting by Elsevier referred to as ‘‘approaching thermodynamic stability’’.

1110-0621 ª 2013 Production and hosting by Elsevier B.V. on behalf of Egyptian Petroleum Research Institute.
Open access under CC BY-NC-ND license. http://dx.doi.org/10.1016/j.ejpe.2013.11.006
518 M.R. Noor El-Din et al.

Nanoemulsions can be diluted without changing the droplet The main aim of the present work is to prepare water-in-
size distribution [5]. O/W nanoemulsions with droplet diame- diesel fuel nanoemulsion with different water contents via high
ters as low as 14 nm and high kinetic stability have been re- energy homogenizer. Two types of nonionic surfactants were
ported by Solans et al. [6,7]. However, w/o nanoemulsions adopted in these experiments. Each of them was blended with
have been receiving increased attention since they were first re- another to form five kinds of mixed surfactant with different
ported by Landfester et al. [8–10], who have described their HLB values used to stabilize water-in-diesel fuel nanoemul-
high energy emulsification methods and used them in a variety sion. Also, the effects of emulsifier concentrations and water
of organic and inorganic polymerization reactions [11]. contents on the stability of these nanoemulsions were studied
Depending on the preparation method, different droplet size at the optimum HLB. The interfacial tension is measured for
distributions are achieved, explaining why the route of prepa- the surfactants used, individually and for the blend between
ration can remarkably influence the emulsion stability [12]. them as an important factor affecting the stability of the for-
The formation of emulsions with droplet size in the 50– mation of the nanoemulsion. Also, measure the physical prop-
200 nm range [4], can be achieved either by high energy emul- erties, kinematic viscosity and density, of the prepared
sification methods (e.g. by high-shear stirring, high-pressure nanoemulsions and study the effect of different temperatures
homogenizers, or ultrasound generators) [13] or by low-energy on these properties.
emulsification methods (e.g. phase inversion temperature PIT)
[7,14]. Although high-energy emulsification methods allow a
great control of the droplet size and a large choice of compo- 2. Experimental
sition, low-energy emulsification methods are interesting be-
cause they take advantage of the energy stored in the system 2.1. Materials
to promote the formation of small droplets [13]. For any emul-
sion system, the choice of a right emulsifier is of very crucial In this study, technical-grade diesel fuel (CO-OP Petroleum
importance. Emulsions are usually formulated to obtain the Company, Cairo, Egypt) was used as the continuous emulsion
stable one. Generally, a single surfactant cannot produce the phase. The physical properties of this diesel fuel are summa-
desired stability [15]. Mixtures of different surfactant types of- rized in Table 1. The technical grade emulsifiers used through-
ten exhibit synergism in their effects on the properties of a sys- out the investigation namely: polyoxyethylene (20) sorbitan
tem for various applications [16–19]. This synergism can be monooleate (abbreviated T) (a hydrophilic surfactant with
attributed to nonideal mixing effects in the aggregates, and it HLB = 15) and sorbitan monooleate (abbreviated S) (a lipo-
results in critical micellization concentrations and interfacial philic surfactant with HLB = 4.3) were obtained from Sig-
tensions that are substantially lower than would be expected ma–Aldrich, Germany. Some physical properties and
on the basis of the properties of the unmixed surfactants alone. structures of T and S are listed in Table 2 and Fig. 1. The water
This situation has generated both theoretical and practical in all experiments was double distilled then deionized and fil-
interest in developing a quantitative understanding of the tered prior to use.
behavior of mixed surfactant systems, driven in part by the po-
tential applications of these systems in detergency [20,21], en-
2.2. Nanoemulsion formation
hanced oil recovery [22], and mineral flotation [23]. Sagitani
et al. [24] suggested that a proper HLB value of the surfactants
was a key factor for the formation of emulsion with minimal Nanoemulsions were prepared using a turbine homogenizer
droplets. Dai et al. [25] observed that the molecular structure (Ultraturrax pro 200, USA) as a high energy emulsification de-
of emulsifiers had a great effect on the droplet size of the final vice in two steps: Firstly, a pre-emulsion was prepared by addi-
emulsions. tion of water in different percentages (5, 6, 7, 8 and 9 wt.%) to
Transportation consumes the largest proportion of energy, a mixture of T and S, so-called TS, and diesel fuel to form five
and has seen the largest growth in demand in recent decades. emulsions. The rate of addition was kept approximately con-
This growth has largely come from new demand for per- stant (0.5 ml/min) with constant stirring at 600 rpm. In the sec-
sonal-use vehicles powered by internal combustion engines ond step, the prepared pre-emulsions were stirred at high speed
[26]. The vehicle emission causes increasing CO2 levels in the (30,000 rpm) for 5 min. All experiments were run at 30 C. Dif-
atmosphere, which may contribute to global warming [26]. ferent HLB values (9.6, 9.8, 10, 10.2 and 10.4) of the mixed
Much information is available in the literature on the increases surfactants (TSHLB) were derived by multiplying HLB values
in gasoline octane number that can be obtained by adding of T and S with their weight fractions.
water in the form of an emulsion or by injecting water into
the intake pipe [27]. A French Petroleum Company has devel- 2.3. Droplet size measurement
oped an emulsified diesel fuel called Aquazole. It contains
13 ± 3% water, 85 ± 3% diesel fuel, and 2–3% of surfactant
The mean droplet size of the prepared nanoemulsions was
additive package. The effects resulting from its use were re-
determined using dynamic light scattering equipment (Zetasiz-
ported in 2000 by Barnaud et al. [28]. Aquazole is in use today
er Nano-ZS, Malvern, UK) at 30 C.
in a large number of vehicles in France and other countries in
Southern Europe [28,29]. Water-in-diesel fuel emulsion is con-
sidered one of the possible alternative fuels for curtailing the 2.4. Interfacial tension measurements
emission pollution of combustion equipment such as diesel en-
gines and large power boilers [30,31]. Emulsion fuels typically The interfacial tension at a planar oil phase–aqueous phase
contain 5–20% of water, surfactant and the base fuel, such as boundary was measured using a tensiometer with a Du
kerosene or diesel. Nouy ring arrangement. Aqueous phases were prepared by
Water-in-diesel fuel nanoemulsions 519

Table 1 Physicochemical properties of diesel fuel.


Experiment Value Standard method
Color at 40 C 2 ASTM D1500-07
Specific gravity 60/60 F (gm/cm3) 0.8427 IP 160-87
Kinematics Viscosity at 40 C (mm2/S) 3.268 ASTM D445-12
Flash Point, C 67 ASTM D93-11
Pour Point, C 6 ASTM D97-11
Cloud Point, C 0 ASTM D2500-11
Distillate at 350 C, v/v% (min.) >85 Egyptian Standard ES16/2005
Water Content, wt.% Nil ASTM D4006-11

Table 2 Surfactants used throughout the investigation.


Material Appearance M. wt.% Density, g/ml at HLB Source
20 C
Span 80 ‘‘Sorbitan monooleate’’ Brown viscous liquid 428.61 0.99 4.3 Sigma–Aldrich Co., Germany
Tween 80 ‘‘Polyoxyethylene (20) Amber sticky liquid 1310 1.08 15 Sigma–Aldrich Co., Germany
sorbitan monooleate’’

2.6. Density measurement

Densities in kg/m3 of the diesel fuel and the prepared nano-


emulsions were determined using DE40 Density Meters,
Mettler–Toledo GmbH, Japan (with calibrated accuracy
0.5 kg/m3). The measurements were performed at a tempera-
Figure 1 Chemical structure of the used surfactants. ture range from 10 C to 70 C with 10 C intervals.

3. Results and discussion

dissolving each surfactant individually and mixed surfactants 3.1. Determination of the optimum HLB value
(TS) into water and stirring for at least 1 h before the solu-
tion was placed into the sample vessel. Oil phase (15 g) was
A proper surfactant HLB value is a key factor for the forma-
then carefully layered on top of the aqueous phase (15 g)
tion of stable emulsions [24,33]. Two types of nonionic surfac-
and left for 1 h at 30 C before each measurement. Ring
tant T and S were evaluated in our preliminary experiments.
buoyancy effects were taken into account for each oil phase
Mixtures of different surfactant types (T and S) often exhibit
by submerging the Du Nouy ring in the oil and adjusting the
synergism in their effects on the properties of a system
interfacial tension reading on the instrument to zero. Each
[16,21]. This synergism can be attributed to nonideal mixing
sample was analyzed three times and the data are reported
effects in the aggregates, and it results in critical micellization
as the average. The standard deviation of the interfacial
concentrations and interfacial tensions that are substantially
measurements was better than 0.3 mN m1. The cmc and
lower than would be expected on the basis of the properties
the interfacial tension at the cmc were determined from
of the unmixed surfactants alone [26]. However, only the com-
the breakpoint of interfacial tension and the logarithm of
bination of T and S had the dramatic effects on the emulsion’s
the concentration curve. From the slope of the linear
properties. Different ratios of T and S were used to turn the
decrease of c, the maximum surface excess concentration
HLB range from 9.6 to 10.4 of the emulsions.
(Cmax), the minimum surface area per surfactant (Amin),
The mixing ratios were adjusted to satisfy the proper HLB
the Gibbs free energy of adsorption (DGads) and the Gibbs
values for optimum emulsification conditions [33]. The mixed
free energy of micellization, (DGmic) were calculated by use
HLB values were calculated by the following equation:
of the Gibbs adsorption equations [32].
HLBTS ¼ HLBT % þ HLBS S% ð1Þ
2.5. Kinematic viscosity Where; HLBT, HLBS and HLBTS are the HLB values of T
(15.0), S (4.3) and the mixed surfactants (TS), and T% and S%
Kinematic viscosities in (cSt) of the diesel fuel and the pre- are the mass percentages of T and S in the mixed surfactants,
pared nanoemulsions were measured according to ASTM respectively. All HLB values used were obtained at 30 C. The
D445-12 using Cannon–Fenske routine viscometer (size no: HLB value (TSHLB) of the mixed surfactants was derived by
75). Kinematic viscosity measurements were performed at tem- multiplying HLB values of T and S with their weight fractions
peratures ranging from 10–70 C with 10 C intervals. A digital to obtain a (TSHLB) of 9.6, 9.8, 10, 10.2 and 10.4. Emulsions
oil bath was employed to prepare and control the required with 5 wt.% water content and 10 wt.% surfactant concentra-
measurement temperatures. tion were prepared at different (TSHLB) of 9.6, 9.8, 10, 10.2 and
520 M.R. Noor El-Din et al.

200 0.3
Water droplet size
180 PDI

0.25
160

140
0.2
120

r o , nm

PDI
100 0.15

80

0.1
60

40
0.05
20

0 0
9.6 9.8 10 10.2 10.4
TS HLB

Figure 2 Nanoemulsion r0 and PDI as a function of the TSHLB values for samples of water/TS/ diesel fuel at 5% water content, 10% TS,
different values of HLB (9.6, 9.8, 10, 10.2, 10.4 and 30 C).

imparting more rigidity and strength to the emulsifier film


Table 3 Effect of different TSHLB values on the initial radius through hydrogen bonding [36]. Also, the reduction of TSHLB
(r0) and poly-dispersity index (PDI) for water-in-diesel fuel may be attributed to the distribution of both emulsifiers (T and
nanoemulsion at 5 wt.% water content, 10 wt.% TS and 30 C. S) on the w/o interface used in our experiments. This means
that the ratio between (T) and (S) affect the adsorption of
Mixed HLB Tween 80, Span 80, r0, nm PDI
two molecules on the w/o interface and the formation of stable
values wt.% wt.%
nanoemulsions. As a result, regarding the structure and hydro-
9.6 49.5 50.5 145.4 0.247 phobicity of both surfactants (T and S), for the same type of
9.8 51.4 48.6 85.3 0.297
hydrophobe in the surfactant, the smaller the head group
10 53.3 46.7 49.55 0.113
(S), the more surfactant adsorbs at the interface and the lower
10.2 55.14 44.86 102 0.214
10.4 57 43 190.1 0.182 the surface tension [38]. Therefore, when the mixed surfactant
contains two different types of surfactants (T) and (S) that
have different solubility in water, the less water-soluble surfac-
tant (S) will be transported more rapidly between the water–oil
10.4. The relationship between the droplet size of the obtained interface than the more water-soluble surfactant (T) [39]. It is
emulsions (nm), poly-dispersity index (PDI) and the TSHLB is likely that the (S) surfactant orientated themselves at the
shown in Fig. 2. Results show that the mean droplet size water-in-oil interface (oil soluble surfactant), then the (T) sur-
decreased from 145.4 to 49.55 nm when the surfactant ratio factant by hydroxyl groups protruded into the aqueous phase,
of T:S wt.% was changed from 49.5:50.5 to 53.3:46.7, enabling them to form hydrogen bonds with the water mole-
respectively and gradually increased from 49.55 nm to cules and reducing the unfavorable contact between hydrocar-
190.1 nm when the surfactant ratio of (T) in the mixed surfac- bon chains and water molecules [36]. This means that the two
tant increased from 53.3:46.7 to 57:43, respectively. The mini- surfactants could arrange more compactly at the interface [36].
mum droplet size (49.55 nm) and the smallest Ostwald repining Accordingly, the combination of (T) and (S) could successfully
was obtained at TSHLB of 10 (the optimum TSHLB) as demon- enhance the physical stability of nanoemulsion [40]. Therefore,
strated in Table 3. The nanoemulsion with a transparent the emulsifier blend with TSHLB = 10 exhibited a synergetic
appearance appeared with the optimum TSHLB, while the activity on the properties of the system and was found to be
other nanoemulsions with different surfactant ratios have a optimum for the preparation of stable w/o emulsions in this
translucent appearance (Fig. 2). Several authors found that study.
surfactant mixtures could provide more stable emulsions with Similar trends were also observed in relation to the poly-
the minimum size than only one surfactant [34–36]. Further- dispersity index (PDI) versus different TSHLB values (see
more, the addition of a second surfactant to the dispersed Fig. 2). By inspection of the obtained data (Fig. 2), it was
phase decreased the rate of Ostwald ripening rate in w/o emul- found that the smallest PDI (0.113) was obtained at TSHLB va-
sions [37]. This phenomena may be regarded the mixed surfac- lue equal to 10. Other emulsions were considered to be poly-
tants were able to form a film around the dispersed droplets dispersed or wide-distributed because the poly-dispersity index
and to maintain the droplet stability by strengthening the was higher than 0.2. Accordingly, the formation of stable
interfacial film [6,11]. In such a system the hydrophilic and nanoemulsions was obtained with the optimum TSHLB and
lipophilic emulsifiers are thought to align alongside each other the smallest PDI [14].
Water-in-diesel fuel nanoemulsions 521

3.2. Nanoemulsion stability where r is the average droplet radius at time t, r0 is the initial
droplet radius and k is the frequency of rupture in the droplet
The w/o nanoemulsions prepared in this work exhibited a surface area [43]. Although the above equation has been devel-
good stability without phase separation for 2 weeks, with a oped for concentrated systems, we have used it in our
slight increase in the mean droplet size with time. The two nanoemulsion systems to show whether coalescence was the
common probable breakdown processes in these systems are driving force for the instability or not. As an illustration, Figs.
coalescence and/or Ostwald ripening [41]. If coalescence were 3 and 4 show plots of r2 vs. t for water/diesel fuel nanoemul-
the driving force for emulsion instability, changes in droplet sions at various TS concentrations and different water con-
size with time would be described by the following equation tents, respectively. In all cases, the plots did not follow the
[42]. predictions of Eq. (2), the Correlation coefficient (b) has a
low range (see Figs. 3 and 4), indicating that coalescence, as
r  2 ¼ r2
0  8p=3kt ð2Þ

4.5
6 % TS β = 0.9758
4 8 % TS β = 0.8406
10 % TS β = 0.8394

3.5

3
-2
x 10 , m

2.5
14

2
-2
r

1.5

0.5

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Time x105, Sec.

Figure 3 Nanoemulsion r2 as a function of time in system of water/TS/diesel fuel at 5% water content, HLB = 10, different
concentrations of TS and 30 C ((b): regression coefficient).

4.5

5% Water content β =0.975


4 6% Water content β =0.981
7% Water content β =0.986
3.5 8% Water content β =0.988
9% Water content β =0.984
3
r-2 x 1014, m-2

2.5

1.5

0.5

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4

Time x105, Sec.

Figure 4 Nanoemulsion r2 as a function of time in system of water/TS/diesel fuel at 10% TS, HLB = 10, different water contents and
30 C ((b): regression coefficient).
522 M.R. Noor El-Din et al.

30
5% Water content
6% Water content
7% Water content
8% Water content
25
9% Water content

20

3
r x10 , m
-22

15
3

10

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4

Time x10 5 , Sec.

Figure 5 Nanoemulsion r3 as a function of time in system of water/TS/diesel fuel at 10% TS, HLB = 10, different water contents and
30 C.

7
6 % TS
6.5 8 % TS
10 % TS
6
5.5
5
4.5
3
r x10 , m

4
-20

3.5
3
3

2.5
2
1.5
1
0.5
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Time x10 5 , Sec.

Figure 6 Nanoemulsion r3 as a function of time in system of water/TS/diesel fuel at 5% water content, HLB = 10, different
concentrations of TS and 30 C.

described by the model of [42], may not be the mechanism of dispersed phase, D is the diffusion coefficient of the dispersed
instability. phase in the continuous phase, d is the density of the internal
Another mechanism is Ostwald ripening which arises from phase, R is the gas constant and T is the absolute temperature.
the difference in internal pressure between droplets with differ- A stable nanoemulsion should have a small mean droplet
ent sizes. In this mechanism, larger droplets grow at the size and a large Ostwald ripening coefficient. Eq. (3) results
expense of smaller ones due to the molecular diffusion through in a linear relationship between the cube of the droplet radius
the continuous phase. The Lifshitz–lezov and Wagner (LSW) (r3) and time (t) as shown in Figs. 5 and 6. Ostwald ripening
theory [44,45] gives an expression for the Ostwald ripening rate rate can be calculated from the slope of each straight line. Ost-
(x) by the equation: wald ripening rate (x), initial radius (r0), poly-dispersity index
(PDI) and regression coefficients (b) of the prepared nano-
x ¼ dr3 =dt ¼ 8=9½ðC1 cVm =dRTÞ ð3Þ
emulsions at different water contents and different TS concen-
where r is the average radius of the droplet, t is the storage trations are shown in Tables 4 and 5 and illustrated in Figs. 5
time, C1 is the solubility of the dispersed phase in the contin- and 6 respectively. A good correlation of r3 and time was
uous phase, c is the interfacial tension between the dispersed obtained with regression coefficients (b)>0.99 for all the tested
phase and the medium, Vm is the molar volume of the emulsions (Tables 4 and 5 and Figs. 5 and 6), indicating that
Water-in-diesel fuel nanoemulsions 523

where, r is the radius of drop curvature, which has to be over-


Table 4 Ostwald ripening rate (x), initial radius (r0), poly-
come to break up a drop into smaller ones. Additional pro-
dispersity index (PDI) and regression coefficients (b) of water-
cesses occur during emulsification, such as adsorption of
in-diesel fuel nanoemulsion formed at different water contents,
surfactants and droplet collision, which may or may not lead
10% TS, HLB = 10 and 30 C.
to coalescence [1]. Concerning the Ostwald ripening rate (x),
Water content, r0, nm x, m3sec1 · 1027 PDI (b) it is clear from Fig. 6 and Table 5 that (x) decreases with
wt.% increasing surfactant concentration (from 4.9248 · 1020 to
5 49.55 2 0.113 0.9932 2 · 1027 m3 sec1) at 5% water content, HLB = 10 and
6 66.8 3 0.124 0.9948 30 C. The decrease in the Ostwald ripening rate (x) with an
7 78.88 5 0.145 0.9940 increase in the emulsifier concentration may be due to the
8 91.63 8 0.168 0.9945 following effects [48]:
9 104.4 10 0.180 0.9946
(1) The diffusion coefficient of micelles decreases as their
concentration increases, because crowding effects could
increase the viscosity.
Table 5 Ostwald ripening rate (x), initial radius (r0),
(2) The decrease in the average droplet size.
poly-dispersity index (PDI) and regression coefficients (b) of
(3) It is possible that association colloids other than micelles
water-in-diesel fuel nanoemulsion formed at different TS
are formed at relatively high emulsifier concentrations.
concentrations, 5% water content, HLB = 10 and 30 C.
(4) It is possible that the emulsifier molecules form a thick
Surfactant conc., r0, nm x, m3 sec1 PDI (b) multilayer around the emulsion droplets, which retards
wt.% the displacement of the water molecules from the drop-
6 135.80 4.9248 · 1020 0.262 0.9956 lets to the surrounding aqueous phase.
8 117.98 3.7792 · 1020 0.129 0.9915 (5) Preliminary aggregates formed at low emulsifier concen-
10 49.55 2 · 1027 0.113 0.9932 tration are loosely associated and likely to be slightly
interacting with water molecules.
(6) In our case and due to both surfactants (T and S) having
the same length of the hydrophobic part, an increase in
the Ostwald ripening was the driving force for the instability
emulsion stability with increasing emulsifier concentra-
[46]. Although the high stable emulsion was obtained with
tion may be caused by a closer packing of surfactant
10 wt.% TS, it is obvious that b for coalescence (b = 0.8394)
molecules at the oil–water interface, which results in a
exhibited a lower value than Ostwald ripening (b = 0.9932)
more rigid interfacial membrane. Overall, the droplet
at 10% TS concentration 5 wt.% water content and
size is reduced when the surfactant concentration rises,
HLB = 10. Another evidence indicates that coalescence was
causing an increase in the ratio of surfactant film thick-
not a driving force for emulsion stability (See Tables 4 and 5
ness to droplet radius. The relative thicker surfactant
and Figs. 3 and 4).
film can afford better steric stabilization against floccu-
A highly transparent nanoemulsion with an initial droplet
lation [49].
size of 49.55 nm was obtained with 10% TS concentration,
HLB = 10 and 5% water content. This droplet size value in-
Also, the evaluation of the particle size distribution for
creases slightly up to 70.96 nm after 360 h under the same
water-in-diesel fuel stabilized nanoemulsion containing 6, 8
conditions. This may be attributed to the instability of nano-
and 10 wt.% TS mixed surfactant concentrations is shown in
droplets, where the small water nanodroplets will disappear
Fig. 7. For 6 wt.% TS, (a) one large peak centrated around
by migration to large droplets via the continuous phase [43].
78.9 nm and (b) a small additional welded peak around
Instability of nanoemulsions may be due to the very small
241.6 nm (broad band dimension 210.5 d.nm and intensity per-
droplet size causing a large reduction in the gravity force
centage 61.7%) and (there is no clear band dimension due to
and the Brownian diffusion may prevent any increasing in
present overlap peak and intensity percentage 38.1%), respec-
droplet size under effect of Oswald ripening [47]. Therefore,
tively is obtained (see Fig. 7A), the mean size of water droplet
the kinetics of nanoemulsion decay may be attributed to Ost-
is 135.80 nm. When increasing the surfactant concentration to
wald ripening, as expected. Ostwald ripening rate decreases
8 wt.%, a truly bimodal distribution is observed with one peak
with the increase in emulsifier concentration from 6% to 10%.
denoted as c at 50.75 nm (slightly narrow band dimension
17.77 d.nm and intensity percentage 95.2%) and another
3.3. Factors affecting the nanoemulsion stability
broad band 295.3 nm denoted as e with band dimension
153.7 d.nm and intensity percentage 4.8%, respectively (see
3.3.1. Effect of emulsifier concentration Fig. 7B), the mean size of water droplet is 117.98 nm. Further
Another object of this paper was to evaluate the influence of increase in TS concentration to 10 wt.% results in one peak at
emulsifier concentration on the stability of w/o nanoemulsion. around 49.55 nm denoted as e (see Fig. 7C), the mean size of
Surfactant concentration is very important for a stable w/o water droplet is 49.55 nm. Overall, It can be concluded that
emulsion system [2]. Since, the emulsifier plays a major role an increase of surfactant concentration (6–10 wt.%), the drop-
in the formation of nanoemulsions. The high energy required let size distribution becomes narrow, and the initial water
for their formation can be understood from the consideration droplet radius declines from 135.80 to 49.55 nm, respectively
of the Laplace pressure (P) as shown in Eq. (4). (see Table 5). Sequentially, the bimodal distribution converts
to monomodal throughout the measurement period. This
P ¼ 2c=r ð4Þ may be attributed to the amount of surfactant as it determines
524 M.R. Noor El-Din et al.

Figure 7 Particle size distributions of water-in-diesel fuel nanoemulsions stabilized by 6, 8 and 10% TS, mixed surfactant concentration,
optimum TSHLB, 5% water content and 30 C. (A: 6%, B: 8% and 10% TS concentration).

the total interfacial area and hence the average size of the content (5, 6, 7, 8 and 9 wt.%), i.e. it was 49.55, 66.8, 78.88,
emulsion droplets [49]. 91.63, 104.4 nm, respectively after 360 h. The Ostwald ripening
Also, Fig. 7 shows that there was only a slight decrease in rate (x) increases with increasing water content (from 2 to
the mean droplet diameter with the increasing of the TSHLB 10 · 1027 m3 sec1) at 10 wt.% of mixed surfactant concen-
concentration from 6% to 10%, while the particle size tration and TSHLB = 10. When the dispersed water phase
distribution remained monomodal throughout the measure- increases in the emulsion, the interfacial film thickness possibly
ment period. decreases [47]. Therefore, the system tended to be instable due
to the limited amount of surfactants [47]. Accordingly, increas-
3.3.2. Effect of water content and oil weight fraction (F) ing of the water percentage in the emulsion could reduce the
The variations of water droplet size (r3) versus time (sec.) in stability of the produced nanoemulsion [50]. Also, with the in-
water/TS/diesel fuel nanoemulsion system with 10 wt.% TS crease of water loading, there occurs an increase in the size of
emulsifier concentration and HLB = 10 at 30 C are shown water droplets; this may be attributed to the constant energy
in Fig. 5. The initial droplet size and Ostwald ripening rates dissipated in each case with the increase of water loading,
of the prepared nanoemulsions are shown in Table 4 and illus- which affects the coarsening value of resultant water droplets
trated in Fig. 5. A good correlation of r3 and time is obtained to higher water droplet size.
with regression coefficients (b) > 0.99 for all emulsions tested Also, the influence of oil weight fraction (F) on the stability
in this study as shown in Table 4. Therefore, the emulsion of w/o nanoemulsions was studied. Oil weight fraction (F) is
breakdown could be attributed to Ostwald ripening mecha- defined as the weight fraction of oil in the total mixture of
nism. From Figs. 9 and 10 and Table 4, it was obvious that oil and water. Oil weight fraction (F) and surfactant concentra-
the droplet size (nm) increases with increasing of the water tion are important for a stable w/o emulsion system [2].
Water-in-diesel fuel nanoemulsions 525

110

105

100

95

90

r0, nm 85

80

75

70

65

60

55

50

45
0.9 0.91 0.92 0.93 0.94 0.95 0.96
F =O/(O+W)

Figure 8 Droplet size as a function of oil weight fraction F for the prepared emulsion at 10% TS, HLB = 10 and 30 C.

160
5% Water content
6% Water content
7% Water content
140 8% Water content
9% Water content

120
Droplet size, nm

100

80

60

40
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Time x105, Sec.

Figure 9 Water droplet size (nm) versus time (sec.) in a system of water/TS/diesel fuel at 10% TS, HLB = 10, different water contents
and 30 C.

Fig. 8 shows, the mean droplet size (nm) of the emulsion types often exhibit synergism in their effects on the properties
as a function of F and water contents. The nanoemulsion of a system [26]. Predominating, the interfacial tension de-
studied was composed of 10 wt.% TS concentration (with pends on the types of emulsifier used to stabilize the water/
optimum TSHLB value = 10) and had an oil weight fraction, oil nanoemulsion. For example, when the oil droplet contains
F, ranging from 0.91 to 0.95. Results show that the mean a sufficient concentration of low polarity emulsifier, the inter-
droplet size was increased from 49.5 to 104.4 nm with the facial tension at the water/oil interface is high. In contrast, the
decrease in F. interfacial tension decreases with high polarity emulsifier [50].
But, the presence of more than one surfactant molecule at the
3.3.3. Interfacial Tension (IFT) and thermodynamic properties interface leads to the decrease in the IFT comparing with the
of the used emulsifier individual emulsifier [50]. However, the HLB number concept
Emulsion droplets are normally stabilized by surfactants or only takes into account the surfactant molecule alone and not
amphiphilic polymers [43]. Mixtures of different surfactant the interaction of the surfactant with water and oil interface in
526 M.R. Noor El-Din et al.

(ccmc = 5.5 m Nm1) was adsorbed first and strongly on w/o


interface comparing with T (ccmc = 7.5 m Nm1), which took
more time to attain the w/o interface [39].
Accordingly, the lowering of ccmc (3.8 m Nm1) of TS was
due to the best synergistic effect between S and T which causes
a reduction in the droplet size, where the amount of surfactant
needed to produce the smallest droplet size depending on the
concentration of surfactant on the bulk which determines the
reduction in c, as given by the Gibbs adsorption equation, as
Figure 10 Photography of the prepared nanoemulsions with shows in Eq. (5).
different water contents [(1) Blank, (2) 5% water content, (3) 6% dc ¼ RTCd ln a ð5Þ
water content, (4) 7% water content, (5) 8% water content and (6)
9% water content]. where, R is the gas constant, T is the absolute temperature and
C is the surface excess (number of moles adsorbed per unit area
of the interface).
the overall emulsion system [51], the HLB temperature affected Also, Fig. 11 shows that a rapid initial decrease in interfa-
the lowering of interfacial tension [52]. cial tension with increasing the emulsifier concentration due to
The surface properties of the surfactants individually and the adsorption of emulsifier molecules to the boundary, which
the mixed surfactant, including the critical micelle concentra- was followed by a plateau region where the interfacial tension
tion (cmc), the values of interfacial tension at the cmc (ccmc), remained relatively constant due to the saturation of the
the maximum surface pressure (pcmc), the maximum surface boundary by emulsifier. It is likely that long-chain alcohol
excess concentration at surface saturation ‘‘effectiveness’’, molecules (T molecules) oriented themselves at the water–oil
(Ccmc) and the minimum surface area per surfactant molecule interface toward the aqueous phase, since, interaction here is
(Acmc) are listed in Table 6 and illustrated in Fig. 11. From presumably by H- bonding between the polar groups and the
these obtained data, it is obvious that the interfacial tension water droplet [32].
(ccmc) was decreased from 7.5 and 5.5 m Nm1 for T and S, Above a certain concentration (T = 5.5 · 104 mol/L and
respectively to 3.8 m Nm1 for TS. Regarding the structure S = 12.9 · 104 mol/L) the interface probably became
and hydrophobicity of S (HLB = 4.3) and T (HLB = 15), T saturated with either T or S molecules, therefore the interfacial
is more hydrophilic than S, and ccmc, it was clear that S tension did not change significantly as T surfactant (possess

Table 6 Interfacial tension and thermodynamic properties for emulsifiers at 30 C.


Emulsifier cmc, mol/L · 104 ccmc, mN/m pcmc, mN/m Cmax, mol/cm2 · 1010 Amin, nm2/molecule DGmic, kJ/mol DGads, kJ/mol
T 5.5 7.5 16.5 0.82 2.02 18.58 20.59
S 12.9 5.5 18.5 1.96 0.85 16.48 17.42
TS 14.3 3.8 20.2 0.94 1.77 16.23 18.39

23
S
21 T
TS
19
mN/m

17
T,

15
Interfacial tension

13

11

3
-11.5 -10.5 -9.5 -8.5 -7.5 -6.5 -5.5 -4.5 -3.5
ln C, mol/L

Figure 11 Interfacial tension (c) vs. ln C for water/emulsifiers/diesel fuel system at 30 C.
Water-in-diesel fuel nanoemulsions 527

Table 7 Kinematic viscosity (cSt) of diesel fuel and the prepared nanoemulsions of different water contents at different temperatures.
% of Water in nanoemulsions Kinematic Viscosity in (cSt) at; Regression Coefficients (b)
10 C 20 C 30 C 40 C 50 C 60 C 70 C
0.00 (Diesel fuel) 6.24 4.35 3.21 2.47 1.96 1.61 1.35 0.9922
5 9.70 7.27 5.64 4.51 3.71 3.09 2.61 0.9957
6 9.76 7.33 5.70 4.56 3.75 3.13 2.65 0.9960
7 9.82 7.39 5.76 4.61 3.79 3.17 2.69 0.9961
8 9.88 7.45 5.82 4.66 3.83 3.21 2.73 0.9963
9 9.94 7.51 5.88 4.71 3.87 3.25 2.77 0.9965

10.2
Diesel

9.2 5% W
6% W

8.2 7% W
8% W
Kinematic viscosity, cSt

7.2 9% W

6.2

5.2

4.2

3.2

2.2

1.2
5 15 25 35 45 55 65 75
Temp., °C

Figure 12 Kinematic viscosity of diesel fuel and the prepared nanoemulsions of water content 5–9 wt.% at temperatures 10–70 C.

Table 8 Density (kg/m3) of diesel fuel and the prepared nanoemulsions of different water contents at different temperatures.
% of Water in nanoemulsions Density in (kg/m3) at; Regression Coefficients (b)
10 C 20 C 30 C 40 C 50 C 60 C 70 C
0.00 (Diesel fuel) 0.8469 0.8396 0.8324 0.8251 0.8179 0.8106 0.8034 1
5 0.8671 0.8602 0.8534 0.8465 0.8397 0.8328 0.8260 1
6 0.8693 0.8624 0.8556 0.8487 0.8419 0.8350 0.8282 1
7 0.8703 0.8634 0.8566 0.8497 0.8429 0.8360 0.8292 1
8 0.8716 0.8647 0.8579 0.8510 0.8442 0.8373 0.8305 1
9 0.8728 0.8659 0.8591 0.8522 0.8454 0.8385 0.8317 1

more hydroxyl group) was added. Also, these results support 0.94 · 1010 mol/cm2. On the other hand, T exhibited the high-
our hypothesis that it is the strength of the interaction between est value of (Amin) (2.02 nm2/molecule) and S has the lowest
the emulsifier and the droplet surface that determines the sta- value of (Amin) (0.85 nm2/molecule), whereas TS has middle
bility of the droplets to coalescence. value between both (Amin) (1.77 nm2/molecule). This may be
The relatively small increase in the surface pressure at regarded to the structure and hydrophobicity of T and S emul-
saturation emulsifier concentrations suggests that T (p = sifiers. Because of the S surfactant, which have a low molecular
16.5 mN m1) and S (p = 18.5 mN m1) should be more weight and only one hydrophilic head (see Fig. 1 and Table 2),
strongly adsorbed to the surface of water emulsion droplets. large numbers of S molecules were needed to occupy the w/o
Table 6 indicates that S has the highest value of (Cmax), interface till saturation. Consequently, the minimum surface
(1.96 · 1010 mol/cm2) and T has the lowest value (0.82 · 10 area per surfactant molecule (Amin) was decreased. However,
10
mol/cm2), while the mixed surfactant TS has value of the low value of Amin suggests that the w/o interface was
528 M.R. Noor El-Din et al.

0.875
Diesel
5% W
6% W
0.86 7% W
8% W
9% W

0.845
Density, kg/m3

0.83

0.815

0.8
5 15 25 35 45 55 65 75
Temp., °C

Figure 13 Density of diesel fuel and the prepared nanoemulsions of water content 5–9 wt.% at temperatures 10–70 C.

close-packed; therefore, the orientation of the surfactant mol- to migrate from the bulk to the interface, therefore, a more
ecule at the interface is almost perpendicular to the interface negative value of DGads was obtained [32].
[53]. On the other hand, the smallest number of moles of (T) By comparing the data obtained from the interfacial ten-
occupied per cm2 to get the surface saturation was owing to sion properties and thermodynamic parameters of mixed sur-
the large molecular weight of T, which has a large hydrophobic factant TS (see Table 6) with the stability data of the
group (see Fig. 1). These hydrophilic groups (three terminal prepared nanoemulsions (Fig. 6), it is clear that there is a direct
hydroxyl groups per molecule) occupy a large surface area relationship between the interfacial tension reduction and the
per molecule, consequently the molecule is adsorbed and smallest droplet size obtained. This means that, the droplet
packed parallelly [53]. In case of TS, the values of (Cmax) radius decreases with the decreasing in interfacial tension.
and (Amin) are in the middle, the behavior is different. Where
the arrangement of S and T emulsifiers in the mixed TS on 3.4. Physical properties of the prepared nanoemulsions
the w/o interface may be influenced on these values, and as
such, the presence of large number of T with respect to S to Nanoemulsions have many interesting physical properties that
perform surface saturation, accordingly the value of Amin will are different from or are more extreme than those of larger
be in between as a result of increasing the area occupied by T. microscale emulsions [54]. In this section, we focus on a few
The results of the thermodynamic parameters of adsorption of the physical properties of nanoemulsions as an important
are shown in Table 6 for the investigated emulsifiers and give new class of soft materials. We do not intend to provide a com-
evidence on the relation between the surface active properties, prehensive review of all of the possible properties, but these
HLB of emulsifier’s used and the emulsification efficiency. particular properties serve as a few primary examples.
As can be seen, the Gibbs free energy of adsorption DGads, It is important to investigate the physical properties of
also varies strongly with the variation of HLB value of the water-in-diesel fuel nanoemulsions to assess the role of the
used emulsifiers. In general, the DGads increases with the in- water present within the diesel fuel nanoemulsions and the
crease of the emulsifier’s HLB. From the obtained data of interaction between the water droplet phase and the diesel fuel
DGmic, it can be concluded that the micellization process is continuous phase. These properties are, for example, density
spontaneous because DGmic < 0. Generally, DGads is lower and kinematic viscosity. Because the emulsion preparation,
than DGmic values. This indicates that, these surfactants favor storage, and transportation can be carried out at different tem-
adsorption more than micellization. Although, the TS has a peratures, it is important to study the effect of temperature on
lower interfacial tension (3.8 mN/m) compared with S the density and kinematic viscosity of the stable water-in-diesel
(5.5 mN/m) and T (7.5 mN/m), DGads of the surfactants in fuel nanoemulsions [55].
our work was ranked (according to the more negativity of
DGads) as follows; 17.42, 18.39 and 20.59 kJ/mol for S,
TS and T, respectively. This means that the surfactant (S) is 3.4.1. Kinematic viscosity of nanoemulsions
strongly adsorbed on the interface followed by surfactant High fuel viscosity can affect fuel atomization upon injection
(T). This may be attributed to the solubility of surfactant, into the cylinder and ultimately result in the formation of en-
where, the more hydrophobic emulsifier (S) needs low energy gine deposits [56]. To investigate the nanoemulsion viscosity,
Water-in-diesel fuel nanoemulsions 529

we used the Cannon–Fenske routine viscometer (size no: 75) References


for all test measurements. We studied the effects of tempera-
ture over the range of 10–70 C for pure diesel fuel alone [1] G.W.J. Lee, T.F. Tadros, Colloids Surf. 5 (1982) 105–115.
and all samples of stable nanoemulsions. Table 7 shows the [2] A. Forgiarini, J. Esquena, C. González, C. Solans, Langmuir 17
viscosity measurements as a function of temperature for pure (2001) 2076–2083.
diesel fuel and the prepared nanoemulsions. Fig. 12 shows [3] L. Wang, X. Li, G. Zhang, J. Dong, J. Eastoe, J. Colloid
the kinematic viscosity of diesel fuel and the prepared nano- Interface Sci. 314 (2007) 230–235.
[4] I. Sole, A. Maestro, C. González, C. Solans, J.M. Gutierrez,
emulsions at 10, 20, 30, 40, 50, 60 and 70 C for different stable
Langmuir 22 (2006) 8326–8332.
nanoemulsions of water content 5, 6, 7, 8 and 9 wt.%. [5] K. Shinoda, H. Saito, J. Colloid Interface Sci. 26 (1968) 70–74.
Fig. 12 shows a strong viscosity reduction nonlinearly [6] C. Solans, P. Izquierdo, J. Nolla, N. Azemar, M.J. Garcia-
with temperature for diesel fuel and all tested nanoemulsions. Celma, Curr. Opin. Colloid Interface Sci. 10 (2005) 102–110.
The accuracy evaluation of the measured viscosities of the [7] P. Fernandez, V. Andre, J. Rieger, A. Kuhnle, Colloids Surf. A:
diesel fuel and the prepared nanoemulsions was done by Physicochem. Eng. Aspects. 251 (2004) 53–58.
correlating them with temperature by means of linear, [8] K. Landfester, M. Willert, M. Antonietti, Macromolecules 33
exponential, power, and polynomial equations. The best cor- (2000) 2370–2376.
relation was obtained using a polynomial equation (b = 0.99) [9] H. Wu, C. Ramachandran, N.D. Weiner, B.J. Roessler, Int. J.
as shown in Table 7. Pharm. 220 (2001) 63–75.
[10] N. Uson, M.J. Garcia, C. Solans, Colloids Surf. A:
Physicochem. Eng. Aspects. 250 (2004) 415–421.
3.4.2. Density of nanoemulsions [11] T.F. Tadros, P. Izquierdo, J. Esquena, C. Solans, Adv. Colloid
Unlike viscosity, the density temperature dependence is Interface Sci. 108–109 (2004) 303–318.
almost linear. This result is in accordance with findings [12] A. Forgiarini, J. Esquena, C. González, C. Solans, Prog. Colloid
reported by Baroutian et al. [57]. Table 8 reports the density Polym. Sci. 118 (2001) 184–189.
for pure diesel fuel and the stable tested nanoemulsions at [13] S. Kentish, T.J. Wooster, M. Ashokkumar, S. Balachandran, R.
different temperatures 10, 20, 30, 40, 50, 60 and 70 C. The Mawson, L. Simons, Innov. Food Sci. Emerg. Technol. 9 (2008)
170–175.
effect of temperature on emulsion density is displayed in
[14] P. Izquierdo, J. Esquena, T.F. Tadros, J.C. Dederen, J. Feng,
Fig. 13 over the temperature range of 10–70 C. In Fig. 13, M.J. Garcia-Celma, N. Azemar, C. Solans, Langmuir 20 (2004)
over the water content range of 0–9 wt.%, the density of 6594–6598.
the nanoemulsion increases significantly with water addition. [15] S. Shahriar, Chem. Eng. Sci. 61 (2006) 3009–3017.
The results indicate that the diesel fuel and the prepared [16] B.M. Razavizadeh, M. Mousavi-Khoshdel, H. Gharibi, R.
nanoemulsions demonstrate temperature-dependent behavior Behjatmanesh-Ardakani, S. Javadian, B. Sohrabi, J. Colloid
and their densities decrease linearly with the increase in Interface Sci. 276 (2004) 197–207.
temperatures. The accuracy of the density data was further [17] X.Y. Hua, M.J. Rosen, J. Colloid Interface Sci. 90 (1982) 212–
evaluated by correlating them with temperature by means 219.
of a linear equation, and good correlations were obtained [18] L.-C. Peng, C.-H. Liu, C.-C. Kwan, K.-F. Huang, Colloids Surf.
A: Physicochem. Eng. Aspects. 370 (2010) 136–142.
(b = 1) as shown in Table 8.
[19] J. Araujo, E. Vega, C. Lopes, M.A. Egea, M.L. Garcia, E.B.
Souto, Colloids Surf. B: Biointerfaces. 72 (2009) 48–56.
4. Conclusion
[20] M.J. Rosen, Langmuir 7 (1991) 885–888.
[21] I. Reif, P. Somasundaran, Langmuir 15 (1999) 3411–3417.
A stable water-in-diesel fuel nanoemulsion with mean droplet [22] F. Scamehorn John, Phenomena in Mixed Surfactant Systems,
size from 49.55 nm to 190.1 nm has been obtained by high en- American Chemical Society, Washington, DC, 1986.
ergy method by adjusting the HLB values of the mixed surfac- [23] D.O. Shah, R.S. Schechter, Improved Oil Recovery by
tants. The main stability mechanism for the nanoemulsions has Surfactant and Polymer Flooding, Academic Press Inc., New
York, 1977.
been found to be Ostwald ripening, with reducing ripening rate
[24] H. Sagitani, J. Am. Oil Chem. Soc. 58 (1981) 738–743.
as the surfactant concentration rises. The stability of the stud-
[25] L. Dai, W. Li, X. Hou, Colloids Surf. A: Physicochem. Eng.
ied emulsions increases with increasing the total emulsifier con- Aspects. 125 (1997) 27–32.
centration from 6% to 10% at low water content. Referring to [26] Z. Fu, M. Liu, J. Xu, Q. Wang, Z. Fan, Fuel 89 (2010) 2838–
interfacial tension and thermodynamic properties of the used 2843.
emulsifiers, it was found that, the mixed surfactant (TS) has [27] V.S. Zavodov, V.A. Luksho, V.I. Neudakhin, G.G. Novitskii,
low interfacial tension compared with the individual emulsifi- E.V. Shatrov, Chem. Tech. Fuels Oils. 18 (1982) 25–29.
ers S and T. The kinematic viscosity of the diesel fuel and [28] F. Barnaud, P. Schmelzle, P. Schulz, SAE Technical Paper 2000-
the prepared nanoemulsions decreases nonlinearly with 01-1861, 2000.
increasing temperature from 10 C to 70 C. The accuracy [29] A. Lif, K. Holmberg, Adv. Colloid Interface Sci. 123–126 (2006)
231–239.
evaluation of the measured viscosities of the diesel fuel and
[30] G. Chen, D. Tao, Fuel Process. Technol. 86 (2005) 499–508.
the prepared nanoemulsions was done by correlation, and
[31] X. Liu, Y. Guan, Z. Ma, H. Liu, Langmuir 20 (2004) 10278–
the best correlation was obtained by using polynomial regres- 10282.
sion. The results indicate that the diesel fuel and the prepared [32] M.J. Rosen, Surfactants and Interfacial Phenomena, third ed.,
nanoemulsions demonstrate temperature-dependent behavior John Wiley & Sons Inc., Hoboken, New Jersey, 2004.
and their densities decrease linearly with the increase in tem- [33] F. Xie, B.W. Brooks, Colloids Surf. A: Physicochem. Eng.
peratures. The accuracy of the density data was further evalu- Aspects. 252 (2005) 27–32.
ated by correlation, and the best correlation was obtained [34] M. Porras, C. Solans, C. González, J.M. Gutiérrez, Colloids
using a linear equation. Surf. A: Physicochem. Eng. Aspects. 324 (2008) 181–188.
530 M.R. Noor El-Din et al.

[35] R.P. Gullapalli, B.B. Sheth, Eur. J. Pharm. Biopharm. 48 (1999) [48] B.P. Binks, J.H. Clint, P.D.I. Fletcher, S. Rippon, S.D.
233–238. Lubetkin, P.J. Mulqueen, Langmuir 15 (1999) 4495–4501.
[36] L. Wang, J. Dong, J. Chen, J. Eastoe, X. Li, J. Colloid Interface [49] W. Liu, D. Sun, C. Li, Q. Liu, J. Xu, J. Colloid Interface Sci. 303
Sci. 330 (2009) 443–448. (2006) 557–563.
[37] K. Welin-Berger, B. Bergenstahl, Int. J. Pharm. 200 (2000) 249– [50] M. Porras, C. Solans, C. Gonzalez, A. Martinez, A. Guinart,
260. J.M. Gutierrez, Colloids Surf. A: Physicochem. Eng. Aspects.
[38] J. Sjöblom, Emulsions and Emulsion Stability, second ed., 249 (2004) 115–118.
Taylor & Francis Group, USA, 2006. [51] M.J. Schick, Nonionic Surfactants: Physical Chemistry, Marcel
[39] M.R. Noor El-Din, in: Colloids Surf. A Physicochem. Eng. Dekker, New York, 1987.
Aspects 390 (2011) 189–198. [52] P. Izquierdo, J. Feng, J. Esquena, T.F. Tadros, J.C. Dederen,
[40] R. Chanamai, G. Horn, D.J. McClements, J. Colloid Interface M.J. Garcia, N. Azemar, C. Solans, J. Colloid Interface Sci. 285
Sci. 247 (2002) 167–176. (2005) 388–394.
[41] P. Taylor, Adv. Colloid Interface Sci. 75 (1998) 107–163. [53] F. Yang, G. Li, J. Qi, S.-M. Zhang, R. Liu, Appl. Surf. Sci. 257
[42] B. Deminiere, A. Colin, F.L. Calderon, J. Bibette, in: B.P. Binks (2010) 312–318.
(Ed.), Modern Aspects of Emulsion Science, The Royal Society [54] T.G. Mason, J.N. Wilking, K. Meleson, C.B. Chang, S.M.
of Chemistry, Cambridge, UK, 1998, pp. 261–291. Graves, J. Phys.: Condens. Matter. 18 (2006) R635.
[43] E.I. Taha, A.A. Zaghloul, A.M. Samy, S. Al-Saidan, A.A. [55] M.T. Ghannam, M.Y.E. Selim, Pet. Sci. Technol. 27 (2009) 396–
Kassem, M.A. Khan, Int. J. Pharm. 279 (2004) 3–7. 411.
[44] I.M. Lifshitz, V.V. Slyozov, J. Phys. Chem. Solids. 19 (1961) 35– [56] J. Čupera, P. Sedlák, P. Karafiát, Acta Universitatis
50. Agriculturae et Silviculturae Mendelianae Brunensis. 58 (2010)
[45] C.Z. Wagner, Z. Elektrochem, Elektrochem. 65 (1961) 581–591. 67–74.
[46] Y. Liu, F. Wei, Y. Wang, G. Zhu, Colloids Surf. A: [57] S. Baroutian, M.K. Aroua, A.A.A. Raman, N.M.N. Sulaiman,
Physicochem. Eng. Aspects. 389 (2011) 90–96. J. Chem. Eng. Data. 55 (2010) 504–507.
[47] I. Capek, Adv. Colloid Interface Sci. 107 (2004) 125–155.

You might also like