ArticlePDF Available

Dielectric relaxation of 1-butyl-3-methylimidazolium tetrafluoroborate/TX-100/toluene microemulsions: Structure transition, percolation mechanism, interfacial polarization and electrical properties of microdroplets

Authors:

Abstract and Figures

The dielectric relaxation spectra of the non-aqueous ionic liquid (IL) microemulsions composed of 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF4])/p-(1,1,3,3-tetramethylbutyl)phenoxypolyoxyethyleneglycol (TX-100)/toluene were obtained, the measured frequency is between 1 MHz to 3 GHz. A unique dielectric relaxation located at 10-100 MHz was observed. The direct current (dc) conductivity data were obtained from the total dielectric loss. Near the percolation transition, weight fraction dependence of dc conductivity and static dielectric constant, frequency dependence of permittivity and loss angle all suggest that static percolation occurs in this hydrophilic IL microemulsion. Phase boundaries of IL-in-toluene (IL/O), bicontinuous (B.C.) and toluene-in-IL (O/IL) subregions were determined by the inflection points in the dependent curve of relaxation parameters varied with weight fraction. The mechanism of this dielectric relaxation is attributed to the interfacial polarization of droplets by analyzing relaxation time with Maxwell-Wagner theory. The phase parameters, which reflect the interior properties of the spherical dispersed particle, were calculated by Hanai theory. The dependence of phase parameters on the variation of sample composition was properly explained. For O/IL subregion, the continuous phase should be an IL/TX-100 binary solvent rather than the pure IL. For IL/O subregions, a certain amount of toluene was dissolved in the droplets.
Content may be subject to copyright.
Journal Pre-proof
Dielectric relaxation of 1-butyl-3-
methylimidazolium tetrafluoroborate/TX-
100/toluene microemulsions: structure transition,
percolation mechanism, interfacial polarization and
electrical properties of microdroplets
Zhen Li, Zhefeng Fan, Zhen Chen, Yiwei Lian
PII: S0927-7757(21)01696-4
DOI: https://doi.org/10.1016/j.colsurfa.2021.127827
Reference: COLSUA127827
To appear in: Colloids and Surfaces A: Physicochemical and Engineering
Aspects
Received
date:
21 August 2021
Revised date: 14 October 2021
Accepted
date:
26 October 2021
Please cite this article as: Zhen Li, Zhefeng Fan, Zhen Chen and Yiwei Lian,
Dielectric relaxation of 1-butyl-3-methylimidazolium tetrafluoroborate/TX-
100/toluene microemulsions: structure transition, percolation mechanism,
interfacial polarization and electrical properties of microdroplets, Colloids and
Surfaces A: Physicochemical and Engineering Aspects, (2021)
doi:https://doi.org/10.1016/j.colsurfa.2021.127827
This is a PDF file of an article that has undergone enhancements after acceptance,
such as the addition of a cover page and metadata, and formatting for readability,
but it is not yet the definitive version of record. This version will undergo
additional copyediting, typesetting and review before it is published in its final
form, but we are providing this version to give early visibility of the article.
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.
© 2021 Published by Elsevier.
Dielectric relaxation of 1-butyl-3-methylimidazolium tetrafluoroborate/TX-
100/toluene microemulsions: structure transition, percolation mechanism,
interfacial polarization and electrical properties of microdroplets
Zhen Li,a,b Zhefeng Fan,a* Zhen Chen,c and Yiwei Lianb
a School of Chemistry and Material Science, Shanxi Normal University, Linfen, Shanxi, 041001, China
b Department of Applied Chemistry, School of Material Science and Engineering, Hebei University of Engineering,
Handan, Hebei, 056038, China
c Department of Applied Chemistry, School of Science, Anhui Agricultural University, Hefei, Anhui, 230036, China
*Corresponding author. E-mail addresses: zhefengfan@126.com
Abstract
The dielectric relaxation spectra of the non-aqueous ionic liquid (IL) microemulsions composed of 1-butyl-3-
methylimidazolium tetrafluoroborate ([bmim][BF4])/p-(1,1,3,3-tetramethylbutyl)phenoxypolyoxyethyleneglycol (TX-
100)/toluene were obtained, the measured frequency is between 1 MHz to 3 GHz. A unique dielectric relaxation located at
10-100 MHz was observed. The direct current (dc) conductivity data were obtained from the total dielectric loss. Near the
percolation transition, weight fraction dependence of dc conductivity and static dielectric constant, frequency dependence of
permittivity and loss angle all suggest that static percolation occurs in this hydrophilic IL microemulsion. Phase boundaries
of IL-in-toluene (IL/O), bicontinuous (B.C.) and toluene-in-IL (O/IL) subregions were determined by the inflection points in
the dependent curve of relaxation parameters varied with weight fraction. The mechanism of this dielectric relaxation is
attributed to the interfacial polarization of droplets by analyzing relaxation time with Maxwell-Wagner theory. The phase
parameters, which reflect the interior properties of the spherical dispersed particle, were calculated by Hanai theory. The
dependence of phase parameters on the variation of sample composition was properly explained. For O/IL subregion, the
continuous phase should be an IL/TX-100 binary solvent rather than the pure IL. For IL/O subregions, a certain amount of
toluene was dissolved in the droplets.
Keywords Waterless microemulsions, bicontinuous structure, loss angle, phase parameters, static dielectric constant
1. Introduction
Ionic liquids (ILs) are electrolytes or salts that are liquid below 100 °C [
1
]. Due to their special physical and chemical
properties, such as high thermal stability, nonflammability, negligible vapor pressure, environmental friendliness, wide liquid
range and can be easily tuned through careful selection of cation-anion combinations, ILs can serve as all kinds of (polar or
nonpolar) solvent and surfactant in colloid dispersion systems [
2
-
6
]. Over the past few years, IL based microemulsions, in
which either the aqueous phase, the oil phase, the surfactant or even two of these components are replaced by ILs, become an
attractive topic [
7
-
38
]. This is because IL microemulsions have both the advantages of ILs and microemulsions, which can
not only overcome solubility limitations of ILs to dissolve a number of chemicals including some hydrophilic and
hydrophobic substances but also provide hydrophobic or hydrophilic nano domains, and thus broaden the utility of ILs.
Journal Pre-proof
2
The most intensely investigated IL microemulsions are waterless microemulsions in which water has been replaced by an
IL. The advantages of these systems lie in: they are especially useful when water should be avoided; the microemulsion
structure can be maintained over a wider temperature range compared to water [7] and can circumvent the solubility
limitations of ILs in nonpolar solvents. Han et al. reported the first non-aqueous microemulsion with IL as a polar phase in
2004 [8], they characterized 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF4])/TX-100/cyclohexane by using
phase diagram, conductivity, dynamic light scattering and freeze-fracture transmission electron microscopy measurements.
Eastoe et al. performed small angle neutron scattering experiment on the same microemulsion, which demonstrated the
formation of surfactant-stabilized dispersed nanodroplets with IL cores [9]. Chakrabarty et al. also studied this microemulsion
and discussed the effect of confinement on rotational and solvation relaxation of probe molecule [10]. No change of IL
([bmim][BF4]) and surfactant (TX-100), by replacing the cyclohexane with toluene, benzene, p-xylene, respectively, the
variety of waterless IL microemulsions is expanded [11-15]. Ghosh et al. [16] also extended IL microemulsions without
water by changing the type of IL, namely, they prepared different microemulsions composed of TX-100, cyclohexane and IL
with the variation of alkyl chain length of the [CnSO4]- anion (n=4,6,8), and they proposed that the long octayl chain of octayl
sulfate allows the anion to align itself along with the TX-100 surfactant which increase the rigidity of the system. Mandal et
al. [17] provided an IL microemulsion that all components were biologically acceptable ([C2mim][C4SO4]/Tween-80/Span-
20/IPM). A wide variety of other ILs are suitable for the preparation of IL-based nonaqueous microemulsion, such as
[bmim][PF6] [18-20], [N3111][NTf2] [21], [C2mim][Tf2N] [22] and [P13][Tf2N] [23]. Researchers have successfully
characterized these systems by using phase diagram, UVvisible spectroscopy, DLS, FTIR, electrical conductivity, small
angle X-ray scattering and diffusion-ordered spectroscopy NMR measurements. In addition, there are numerous other (non-
ionic, cationic as well as anionic) surfactants can form non-aqueous IL microemulsion [24-29]. Even some nonaqueous IL
microemulsions without surfactant have been reported [30,31]. Nonaqueous IL microemulsions have been applied as reaction,
separation, and extraction media, and applied in biocatalysis, polymerization, (nano-)material synthesis, drug delivery etc.
[5].
The understanding of physical and chemical properties of nonaqueous IL microemulsion are supplemented by the
analysis of dielectric behavior recently. Over a wide frequency range, multiple dielectric relaxations were observed in IL
microemulsions such as [bmim][BF4]/TX-100/cyclohexane [32-34], [bmim][PF6]/TX-100/ethyleneglycol [35,36],
[bmim][BF4]/TX-100/benzene [37], [bmim][BF4]/TX-100/p-xylene [38] and [bmim][BF4]/TX-100/triethylamine [
39
]. In
these reports, a theoretical and quantitative analysis of the direct current (dc) conductivity data (obtained from the dielectric
measurement) were carried out, which indicated the static percolation process took place in these systems [32,35,37-39]. In
addition, that conclusion was supported by the quantitative analysis for the frequency dependence of the dielectric constant
[32,39]. However, previous studies only used the conductivity method to discussed the percolation process qualitatively
[8,11-13,19-22,30]. The observed dielectric relaxation around 10-100 MHz was ascribed to an interfacial polarization, which
is a strong evidence that interfaces exist in such systems [33,35,36,39]. The phase parameters represent the individual
electrical properties of the constituent phase were calculated according to Hanai theory [33,35,36,39]. The obvious problem
with the dielectric study of nonaqueous IL microemulsions is that only five such systems have been studied. Compared with
conventional conductivity measurements, dielectric spectroscopy has a major advantage for investigating percolation of
microemulsions, namely, this technique allows the determination of the percolation mechanism by using four different
methods at the same time: concentration dependence of (1) dc conductivity and (2) static dielectric constant, frequency
Journal Pre-proof
3
dependence of (3) permittivity and (4) loss angle. Method (1) was widely used in many nonaqueous IL microemulsions [32-
39]; in [bmim][BF4]/TX-100/cyclohexane methods (1) and (3) were both used [33]; finally, all four methods have only been
used simultaneously in [bmim][BF4]/TX-100/triethylamine [39]. This shows that the dielectric spectrum method has an
important potential application in judging the percolation mechanism of anhydrous IL microemulsion.
Based on the above considerations, we measured the weight fraction dependent dielectric relaxation spectra of
[bmim][BF4]/TX-100/toluene at a constant temperature. The dielectric behavior (including loss angle and static dielectric
constant) of this system was analyzed in light of the percolation theories and Hanai theory, the structure transition
information, percolation mechanism, dynamics properties (interfacial polarization) and interior electrical properties (the
phase parameters) of IL/O and O/IL droplets were studied. This study expanded the research scope for the dielectric analysis
of nonaqueous IL-based microemulsion, and thus is a supplement to the understanding of the physical and chemical
properties of these systems.
2. Experiment
2.1. Preparation of IL-based microemulsion
Fig. 1. Chemical structure of TX-100, [bmim][BF4] and toluene.
The IL [bmim][BF4] (purity > 99%) was obtained from Shanghai Cheng Jie Chemical Co. Ltd., China. The water was
less than 10000ppm and the residual chlorides was less than 800ppm in the IL sample. TX-100 (p-(1,1,3,3-
tetramethylbutyl)phenoxypolyoxyethyleneglycol) (laboratory grade) was purchased from Sigma-Aldrich LLC. America.
Toluene (analytical grade) was produced by Beijing Chemical Reagent Factory, China. The structural formula of TX-100,
[bmim][BF4] and toluene are shown in Fig. 1. According to the literature [11], by mixing appropriate weight fractions of
[bmim][BF4], TX-100 and toluene, a series of microemulsions of different weight fractions were prepared, the experimental
path is shown in Fig. 2. In these samples, the weight ratio of toluene/TX-100 is fixed at 1:1, and the weight fraction of
[bmim][BF4] ranged from 3.39 wt% to 60.07 wt%, the interval is approximately 3.00 wt%. We prepared other two samples
which are marked as (I) and (II) in Fig. 2; the weight fractions of [bmim][BF4], TX-100 and toluene are 16, 68 and 16 wt%
for (I) and 9, 83 and 8 wt% for (II). For comparison, we prepared two [bmim][BF4]/TX-100/triethylamine samples which
noted as (III) and (IV); the weight fractions of [bmim][BF4], TX-100 and triethylamine are 16, 67 and 17 wt% for (III) and 9,
82 and 9 wt% for (IV). The molar fractions of IL, TX-100 and oil of samples (I) and (III) are both 20, 30 and 50 %, and the
ratio for samples (II) and (IV) are both 15, 50 and 35 %. All mixtures were transparent and were allowed to stand for over
seven days prior to dielectric measurement to equilibrate these systems.
Journal Pre-proof
4
Fig. 2. Experimental path diagram for [bmim][BF4]/TX-100/toluene systems, which denotes the change of IL content when the weight ratio
of toluene/TX-100 is fixed at 1.
2.2. Dielectric measurement
The dielectric measurements were performed on an RF impedance analyzer (Keysight E4991B, USA) that allows for a
continuous frequency measurement from 1 MHz to 3 GHz. This impedance analyzer was tested and calibrated in accordance
with the procedure recommended by the manufacturer. The complex permittivity were obtained with a coaxial sample cell
[
40
] located at the end of a coaxial line, which is composed of an outer conductor, with an inner diameter of 3.5 mm and a
length of 25 mm, and an inner conductor, with an outer diameter of 2 mm and a length of 10 mm. The temperature of weight
fraction dependent dielectric measurements was 20 ± 0.1 °C. The permittivity and total dielectric loss at every measured
frequency were directly obtained.
3. Method and models
3.1. Dielectric spectra analysis
When a microemulsion with high conductivity is placed in an applied electric field of angular frequency
(
2f

,
f
is the frequency), the dielectric properties of this sample can be generally characterized in terms of the complex
permittivity
*
.
* ' "total
j

(1a)
(1b)
where
'
is the permittivity,
21j
,
"total
is the total dielectric loss,
"
is the dielectric loss of the sample,
0
is the
permittivity of vacuum,
l
is the dc conductivity (which can be obtained by fitting the part of
"total

that has an index
of -1 [
41
,
42
,
43
]).
Journal Pre-proof
5
If the measured sample presents a dielectric relaxation, the dielectric spectrum can be further characterized by a set of
dielectric parameters. These parameters can be obtained by fitting the experiment data with the empirical function containing
a Cole-Cole term as below [
44
,
45
]
*1
hj


(2)
where
h
is the high-frequency limit of permittivity,
is the dielectric increment,
is the relaxation time,
(
01

) is the Cole-Cole parameter (indicating the dispersion of the relaxation time). The static dielectric constant is
defined as
sh
. By using these dielectric parameters, we can not only analyze the relaxation mechanism but also
calculate the phase parameters.
3.2. Percolation models
Fig. 3. Diagrammatic drawing of (a) [bmim][BF4]-in-toluene (IL/O) subregion, (b) bicontinuous (B.C.) subregion, (c) toluene-in-
[bmim][BF4] (O/IL) subregion. Both (a) and (c) can be modeled as spherical droplets dispersed in a continuous phase.
As illustrated in Fig. 3, when the weight fraction of IL increases, the microstructure of the IL microemulsions will
change from a IL-in-oil (IL/O for short; see Fig. 3a for the schematic diagram) droplets to an oil-in-IL (O/IL for short; see
Fig. 3c for the schematic diagram) type, this is the so-called percolation phenomenon. There are two theories to describe the
transition mechanism between IL/O and O/IL. The dynamic percolation model [
46
,
47
] considered that percolation clusters
are formed during the percolation process. The static percolation attributes the percolation to the appearance of a
bicontinuous (B.C.) structure [
48
] as shown in Fig. 3b. Conductivity and dielectric relaxation spectroscopy are the frequently
used techniques to investigate microstructure and structural change on the basis of the percolation theories. Both the dynamic
and static percolation models hold that the percolation transition takes place at a certain threshold
p
c
[46,47,48]. Below and
above
p
c
, the relations between the dc conductivity and the weight fraction of water (or hydrophilic) phase are as below,
respectively:
s
lp
cc

for
p
cc
(3)
Journal Pre-proof
6
lp
cc

for
p
cc
(4)
Similar with Eq. (3), there is another scaling relation between static dielectric constant and weight fraction of hydrophilic
phase below
p
c
[41,
49
,
50
].
s
sp
cc

for
p
cc
(5)
Moreover, s and μ are related to a universal exponent
u
through Eq. (6):
us
(6)
The value of
u
can be obtained directly by
'

and

curves near
p
c
by using Eqs. (7) and (8), respectively
[41,49,50].
1
'u
(7)
11-
2
pu

(8)
Where

is the loss angle of sample and is defined as
"/ '
. The plateau (or maximum) of

provides the asymptotic value
p
of the loss angle, which gives
u
according to Eq. (8). In these equations, s ≈ 1.2 and u
0.61 in dynamic theory [46,47], s 0.7 and u 0.73 in static theory [48], and μ 1.9 both in dynamic and static theory
[46,47,48].
3.3. Calculation of phase parameters (Hanai theory)
Journal Pre-proof
7
Fig. 4. Diagrammatic drawing of Hanai model: the continuous phase has the permittivity of
m
and the conductivity of
m
, in which the
dispersed phase has the permittivity of
p
and the conductivity of
p
, the volume fraction of the dispersed phase is
.
Relaxation parameters represent the collective properties of microemulsions, while phase parameters represent the
individual electrical properties of constituent phase. As illustrated in Fig. 4, phase parameters include the permittivity of the
dispersed phase
p
, the permittivity of the continuous phase
m
, the conductivity of the dispersed phase
p
, the
conductivity of the continuous phase
m
and the volume fraction occupied by the dispersed phase
. For a system of
spherical droplets as depicted in Fig. 3a and 3c, with high volume fractions dispersed in a continuous phase, according to the
Hanai theory, can be modeled as
1/3
**
**
*=1
*
pm
mp

 


(9)
On the basis of this equation, the formulae related to the limiting values of low- and high-frequencies are given by
1/3
=1
hp
m
m p h

 


(10)
31
3m p p m
s
l p l m p l p m
(11)
31
3m l l m
h
h p h m p h p m
(12)
1/3
=1
lpm
m p l

 


(13)
where
10hl

is the high-frequency limit of conductivity. After cumbersome and mathematical treatments, the
phase parameters are calculated from the dielectric parameters. By using the phase parameters, the relaxation time of the
interfacial polarization, according to Maxwell-Wagner theory, can be calculated as below:
0
2
2
m p m p
MW
m p m p

(14)
4. Results and discussion
Journal Pre-proof
8
Fig. 5. Three-dimensional representations for the frequency dependences of the (a)
'
(permittivity), (b)
"total
(total dielectric loss) and
(c)
"
(dielectric loss of sample) for [bmim][BF4]/TX-100/toluene microemulsions with different IL weight fractions.
Fig. 5a and b show the directly measured frequency dependence of complex dielectric permittivity of microemulsions
[Bmim][BF4]/TX-100/toluene with different IL weight fractions when the weight ratio of toluene/TX-100 is fixed as 1:1. Fig.
5a shows the variation of permittivity as a function of frequency. Fig 5b shows the spectra of total dielectric loss. As an
example, Fig. 6 shows the total dielectric loss
"total
directly measured (open circles) when the IL weight fraction is 11.94
wt%.
"total
depends on frequency with a power of -1 in the low frequency range, which indicates that
"total
includes the
contribution of dc conductivity
0
/
l
. Therefore, according to Eq. (1b) we fit the data at low frequency, the fitting result
is represented by the straight line in Fig. 6, thus the dc conductivity
l
of the sample was obtained. Further subtracting
0
/
l
from
"total
, the dielectric loss of sample
"
(represented by the solid circles in Fig. 6) can be obtained.
"
and
l
of each measured sample, obtained by this method, are respectively shown in Fig. 5c and Table 1.
Fig. 6. Example of obtaining dc conductivity and dielectric loss of sample from the total dielectric loss at an IL weight fraction of 11.94
wt%.
A remarkable dielectric relaxation was observed in Fig. 5a, which is confirmed from the corresponding peak position of
dielectric loss of sample as shown in Fig. 5c. In addition, as shown by the arrows in Fig. 5a and c, these spectra and the
relaxation they exhibited have three distinct different weight fraction dependence stages. This phenomenon should relate to
that the experimental path spans three different subregions (IL/O, B.C. and O/IL) of the microemulsions.
Journal Pre-proof
9
Table 1
DC conductivity and dielectric parameters of the relaxations for [bmim][BF4]/TX-100/toluene with different IL weight fractions.
IL
c
(wt%)
Micro-
structure
l
(mS/cm)
h
(ns)
s
3.39
IL/O
0.194±0.001
3.82±0.06
5.10±0.08
2.53±0.06
0.58±0.01
8.92
6.41
IL/O
0.726±0.001
3.85±0.09
7.35±0.04
2.33±0.07
0.63±0.01
11.20
9.22
IL/O
1.364±0.002
3.86±0.08
8.05±0.08
2.24±0.07
0.67±0.01
11.91
11.94
IL/O
2.202±0.002
3.86±0.07
8.81±0.06
1.87±0.08
0.67±0.01
12.67
15.08
IL/O
3.513±0.003
3.87±0.08
9.12±0.06
1.42±0.08
0.66±0.01
12.99
18.25
IL/O
4.964±0.004
3.88±0.08
9.35±0.07
1.29±0.07
0.68±0.01
13.23
20.94
B.C.
6.167±0.005
3.90±0.07
9.38±0.07
1.25±0.08
0.64±0.01
13.28
23.93
B.C.
8.258±0.007
3.91±0.07
9.34±0.08
1.08±0.08
0.64±0.01
13.25
27.15
B.C.
10.43±0.01
3.95±0.07
9.20±0.08
0.978±0.003
0.64±0.01
13.15
29.99
B.C.
15.27±0.02
4.02±0.08
9.25±0.07
0.903±0.003
0.62±0.01
13.27
33.18
B.C.
18.45±0.02
4.05±0.08
9.15±0.06
0.790±0.003
0.64±0.01
13.20
36.00
O/IL
20.58±0.03
4.09±0.09
9.38±0.07
0.767±0.004
0.63±0.01
13.47
39.11
O/IL
21.93±0.02
4.09±0.09
9.5±0.08
0.788±0.004
0.60±0.01
13.59
42.16
O/IL
23.74±0.03
4.10±0.07
9.76±0.08
0.730±0.004
0.60±0.01
13.86
45.14
O/IL
25.31±0.03
4.10±0.07
9.98±0.07
0.739±0.004
0.58±0.01
14.08
47.99
O/IL
26.44±0.05
4.10±0.08
10.00±0.07
0.702±0.005
0.57±0.01
14.10
50.97
O/IL
30.17±0.05
4.11±0.08
10.11±0.08
0.695±0.005
0.58±0.01
14.22
53.80
O/IL
32.60±0.04
4.12±0.08
10.36±0.06
0.679±0.005
0.57±0.01
14.48
56.88
O/IL
34.85±0.05
4.13±0.07
10.51±0.07
0.644±0.005
0.57±0.01
14.64
60.07
O/IL
37.37±0.05
4.14±0.09
10.55±0.08
0.625±0.005
0.59±0.01
14.69
4.1. The percolation in IL-based microemulsion
Fig. 7. Dc conductivity of [bmim][BF4]/TX-100/toluene as a function of IL weight fraction. The arrow points out the percolation threshold
which is obtained from Fig. 9.
As shown in table 1, the dc conductivity of the lowest IL weight fraction is 0.194 mS/cm, and the value is 37.37 mS/cm for
the highest IL weight fraction. That is to say with increasing of IL weight fraction, the dc conductivity of [bmim][BF4]/TX-
100/toluene changed by two orders of magnitude, which is known as an electric percolation [
51
-
53
]. This phenomenon means
that the microstructure of this IL microemulsion changed from [bmim][BF4]-in-toluene droplets (as shown in Fig. 3a) to
toluene-in-[bmim][BF4] droplets (as shown in Fig. 3c) through a transition structure. As introduced in section 3.2, there are
Journal Pre-proof
10
two models, namely, percolation clusters and bicontinuous structure to describe this transition structure. Theoretical analysis
of dc conductivity and complex permittivity is needed to determine what the structure is. Fig. 7 shows the dependence of dc
conductivity listed in table 1 on IL weight fraction. Neither the inflection points indicating the boundary of different
microregions in this microemulsion nor the percolation threshold can be detected in this curve, due to the high conductivity
of ILs.
Fig. 8. Permittivity of [bmim][BF4]/TX-100/toluene microemulsions vs IL weight fraction for different frequencies.
The dependences of
'
for seven different frequencies on IL weight fraction are shown in Fig. 8. Clearly, there exists an
inflexion which provides the percolation threshold at approximately 27.15-36.00 wt% in this curves. In the case of not only
traditional microemulsion but also IL microemulsions [35,38,39,41,49], the maximum of the dependence of
/
l IL
d dc
on
IL
c
corresponds to the value of the threshold
p
IL
c
. Here, the percolation threshold 29.98 wt% is obtained from the same way
as shown Figure 9 and is pointed by the arrows both in Fig. 7 an Fig. 9. This value falls within the range of the percolation
threshold (27.15-36.00 wt%) shown in Fig. 8, which demonstrates the effectiveness of these two methods.
Fig. 9. The dependence of
/
l IL
d dc
on
IL
c
, the percolation threshold
p
IL
c
(=29.98 wt%) is obtained from the maximum point.
Journal Pre-proof
11
Fig. 10. Scaling behavior of the dc conductivity below and above the percolation threshold. The line is the scaling fit result of data below
percolation threshold according to Eq. (3).
Figure 10 depicts the variation of dc conductivity as a function of percolation distance divided by the percolation
threshold. Note that the dc conductivity is plotted in logarithmic coordinates. In this figure, the data below percolation
threshold (open circles) can be divided into two linear dependencies. The previous linear relationship was fitted by using Eq.
(3) with the slope of -0.7. The fitting solid line can describe properly the scaling behavior of the dc conductivity data which
having located the percolation process when below the percolation threshold. Thus, s=0.7 i.e. the static percolation process
takes place in the microemulsion.
Fig. 11. Dependence of static dielectric constant on
/
pp
IL IL IL
c c c
, which is plotted in a logarithmic scale. The line is the fitting
result of data below percolation threshold according to Eq. (5).
The static dielectric constant of every measured sample is obtained according to
sh
, which will be
introduced in section 4.2 as following. The dependence of
log s
on
log /
pp
IL IL IL
c c c
is shown in Fig. 11.
According to Eq. (5) with the slope -0.7, we get the fitting line, which can properly describe the scaling behavior of static
dielectric constant in the data range of 49.70 to 78.62 wt%. This phenomenon also indicates that the static percolation process
takes place in this microemulsion, which is consistent with the conclusion obtained from the analysis of dc conductivity data.
The analysis of physical significance for other conductivity and static dielectric constant data are failed, which is possibly
caused by: (1) the higher electrical conductivity of IL microemulsion than that of traditional microemulsion; (2) the
superposition of interfacial polarization and percolation process (this conclusion will be introduced in the second paragraph
Journal Pre-proof
12
of section 4.3). In above discussions of
l
and
s
, the data which can conform to the theoretical predictions are very limited,
the other exponents μ and u cannot be obtained. These two limitations thus make it is necessary to further decide theoretically
the percolation mechanism by the discussion of frequency-dependent dielectric behavior.
Fig.12. Scaling dependence of permittivity on frequency for the IL weight fraction of 29.98 wt%. The slope of the straight line gives u =
0.73.
Compared with conductivity measurements, dielectric spectroscopy has a major advantage for investigating percolation
of microemulsions, because the universal parameter u can be directly determined by the measured dielectric spectra, namely
via Eq. (7) and (8). The evolution of
log '
with
log f
is shown in Fig. 12. By using the slope of linear part of this curve -
0.27, u is calculated as 0.73, which manifests again the system undergoes a static percolation process. The exponent u is also
accessible from the loss angle following Eq. (8). The variations of
vs frequency for the percolation threshold is shown in
Fig. 13. The horizontal line corresponds to
p
=0.39 and then u=73, which is consistent with the value obtained from Fig. 12.
Fig. 13. Variation of the loss angle
as a function of frequency for IL weight fraction of 29.98 wt%.
In summary, four data curves provide evidence that a bicontinuous structure is formed in this [bmim][BF4]/TX-
100/toluene IL microemulsion during the percolation process. Fig. 3b presents a simplified portrayal of bicontinuous
structure, namely, IL and oil tubes are entangled and crossed with each other, with IL and hydrocarbon regions stretching
over large distances. In addition, the rapid migration of ions in the IL tunnels results in the percolation phenomenon observed
in Fig. 7, which makes the conductivity increase by two orders of magnitude.
Journal Pre-proof
13
4.2 Structural transition of IL microemulsion
Fig. 14. Variations of
'
() and
"
() vs
f
, the solid lines are the theoretical curve of Eq. (2),
IL
c
=15.08 wt%.
As an example, Fig. S in the Supplementary Material shows the complex plane plot of [bmim][BF4]/TX-100/toluene at an
IL weight fraction of 20.94 wt%. The situation is much closer to the Cole-Cole Plot (the dotted line is the Debye function), so
we finally used the ColeCole function, namely Eq. (2), to fit all of the complex dielectric spectra shown in Fig. 5a and c.
During this process, a satisfactory estimation of the dielectric parameters is obtained by using the nonlinear least-squares
method to minimize the sum of the residuals for the complex permittivity. To manifest the goodness of the fitting process,
Fig. 14 shows the fitting curves (solid lines) are in agreement with the experimental data (open symbols) when the IL weight
fraction in the microemulsion is 15.08 wt%, as an example. The values of dielectric parameters
h
,
,
, and
were
listed in Table 1 and the weight fraction dependence of them are shown in Fig. 15. In Fig. 15, all the data curves have two
inflection points at the same positions, namely, 19.30 and 33.84 wt%. Since section 4.1 has already proved the microstructure
of this IL microemulsion changed from IL/O to O/IL through the bicontinuous state (B.C.), by using these two weight
fractions in Fig. 15, the experimental path is divided into IL/O, B.C. and O/IL subregions. In addition, the positions of these
two inflection points are consistent with the phase boundaries which are reported in the literature [11], which indicates that
the dielectric parameters obtained by our fitting process are reasonable. The dependence of dc conductivity on IL weight
fraction, as shown in Fig. 7, does not have inflection points; meanwhile, platforms cannot be detected in Fig. 7 either, on the
contrary, we can detect two platforms (or slow growth) in both IL/O and O/IL subregions in the curve of
h
vs IL weight
fraction as shown in Fig. 15d. Those all demonstrates that the dielectric parameters are more sensitive to structure of the
system than conductivity data. In summary, two phase transitions are occurring between the data-range we have selected in
Fig. 2.
Journal Pre-proof
14
Fig. 15. Plot of (a)
(hollow symbols), (b)
, (c)
, (d)
h
against IL weight fraction. In (a), solid symbols represent the relaxation
time of interfacial polarization
MW
of samples.
4.3 Relaxation mechanism
The most direct way to determine the relaxation mechanism for the observed dielectric relaxation is to compare the
experimental and theoretical values of the relaxation time. As introduced in section 3.2, both IL/O and O/IL droplets can be
modeled as spherical droplets dispersed in a continuous phase. In addition, because the dielectric constants of [bmim][BF4],
TX-100 and toluene are significantly different, in the IL microemulsion made up by using these three materials, the electrical
parameter of dispersed and continuous phase should also be different, and the interfacial polarization should be observed in
our measured frequency range, as summarized in section 3.3. On the basis of this idea, we calculated the phase (electrical)
parameters including the relaxation time of interfacial polarization according to Eq. (9-14) by using the dielectric parameters,
the results are listed in Table 2.
For convenient comparison, the dependence of
MW
on IL weight fraction is also shown in Fig. 15a. In this figure, we can
see that not only the theoretical values
MW
(solid circles) for every sample in both IL/O and O/IL subregions are very close
to the experimental values
(hollow circles), but also they have the same weight fraction dependence trend. Thus, we
believe that the dielectric relaxation observed in the [bmim][BF4]/TX-100/toluene microemulsions should be attributable to
the interfacial polarization, which was also observed in other IL-based [33,35,36,39] and traditional [
54
,
55
] microemulsions.
The existence of interfacial polarization indicates that: (1) the previous spherical IL/O and O/IL models is reasonable, which
strongly proves the existence of microdroplet structure coated by surfactant interfacial membrane; (2) the dielectric properties
of continuous phase and dispersed phase are significantly different; (3) in Fig. 15 the two inflection points are taken as the
boundaries between these three subregions (IL/O, B.C. and O/IL) is reasonable. The superposition of these two polarizations
results in only part of the data for
s
in Fig. 10 can match the theoretical prediction. In addition, Maxwell-Wagner
polarization for suspensions is more specifically ascribed to the fluctuation of space charge (arising in solution neighboring
the particle surface) within the Debye length (a measure of the thickness of the double layer characterizing the size of the
Journal Pre-proof
15
diffuse ion cloud), as reported by O’Brein et al. [
56
,
57
] and Dukhin et al. [
58
,
59
]. This double layer and then the Debye
length may exist not only on the surface of O/IL and IL/O droplets but also on the surface wall of oil and IL channels in the
B.C. subregion. That is why the dicontinuous phase also has a dielectric relaxation and the relaxation time in this sub-region
transitions continuously from IL/O to O/IL microregions, as shown in Fig. 15a. The relaxation time distribution index
is
in the range of 0.65 to 0.70, which deviates from 1, indicating that there are at least two types of relaxation mechanisms in
this observed dielectric relaxation: one is the interfacial polarization which is just identified; the other one is related to the
percolation process introduced in section 4.1 because percolation can be regarded as a polarization process [39,41,50].
As shown in Table 1, the value of dielectric increment falls within the range of a typical interfacial polarization relaxation.
The sum of interface in the system is proportional to the dielectric increment [35,39], thus the weight fraction dependence of
interface can be seen in Fig. 15c through the
IL
c
curve: (1) from 3.39 to 18.25wt%, namely, in IL/O subregion, more
and more IL molecules keep inserting into the micelle cores, which induce an increase of interface and is why
increases
from 5.10 to 9.35; (2) from 20.94wt% to 33.18wt%,
is a constant in the B.C. subregion, and the average value of it is
9.26; this phenomenon infers that the interface sum is not change during the formation of IL and oil channels (namely, the
IL/O droplets break at the same time the O/IL droplets form) in this subregion [
60
-
62
]; (3) in O/IL subregion, the volume
fraction of oil droplets increase (as discussed in the fifth paragraph of section 4.4) makes the sum of interface increases, that
is why above 36.00wt%,
still increases.
4.4 Analysis of the phase parameters
Table 2
Phase parameters of [bmim][BF4]/TX-100/toluene with different IL weight fractions.
IL
c
(wt%)
Micro-structure
m
m
(mS/cm)
p
p
(mS/cm)
MW
(ns)
3.39
IL/O
2.40
0.0490
7.50
6.56
0.382
2.20
6.41
IL/O
2.40
0.134
6.32
11.1
0.461
1.51
9.22
IL/O
2.40
0.210
5.76
11.6
0.518
1.28
11.94
IL/O
2.40
0.294
5.44
14.5
0.558
1.06
15.08
IL/O
2.40
0.423
5.25
18.9
0.589
8.37
18.25
IL/O
2.40
0.534
5.06
21.8
0.623
7.54
36.00
O/IL
10.7
4.71
2.40
34.1
0.718
1.17
39.11
O/IL
10.8
4.97
2.40
36.1
0.723
1.13
42.16
O/IL
11.0
5.30
2.40
38.9
0.726
1.08
45.14
O/IL
11.2
5.57
2.40
41.4
0.729
1.04
47.99
O/IL
11.3
5.81
2.40
42.9
0.732
1.02
50.97
O/IL
11.4
6.61
2.40
48.8
0.734
0.909
53.80
O/IL
11.6
7.05
2.40
52.8
0.735
0.859
56.88
O/IL
11.7
7.50
2.40
56.3
0.737
0.818
60.07
O/IL
11.8
8.04
2.40
60.1
0.738
0.774
The IL/O and O/IL microemulsions are both spherical droplets (with a high volume fraction
) dispersed in a continuous
phase as shown in Fig. 4, whose phase parameters can be all calculated according to the Hanai theory introduced in section
3.3. When
m
in the IL/O microregion and
p
in the O/IL microregion are respectively taken as the dielectric constant of
Journal Pre-proof
16
pure toluene (2.40), other phase parameters of each sample was calculated according to Eqs. (9-13), the results were listed in
Table 2. To better show the weight fraction dependence, the phase parameters of these two subregions listed in table 2 were
plotted in Fig. 16.
Compared with traditional microemulsions, ILs often require more surfactants to form stable microemulsions. In addition,
the range of IL/O and O/IL microregions is narrow in the selected experimental path in this paper, the structural
transformation is rapid. Both of these two factors mean that the measured sample should be a dense system. This prediction
can be illustrated by our calculation, namely, the minimum of the volume fraction is 38.2% for the dispersed phase. This
calculation result also shows that the using of Hanai theory is reasonable.
Fig. 16. Evolution of (a)
, (b)
p
, (c)
m
and (d)
p
as a function of IL weight fraction for IL/O subregions. Plot of (e)
, (f)
m
and (g)
p
against toluene weight fraction for O/IL subregions.
In the IL/O subregion, as shown in Fig. 16a, with increasing IL weight fraction,
also increases because the IL
constantly solubilizes into the micropool of droplets. As shown in Fig. 16b,
p
decreases as the IL weight fraction increases,
and their values are lower than the values of both pure IL (13.06 [33]) and pure TX-100 (7.59 [33]) but higher than the value
of pure toluene 2.40. This means that the dispersed phase should be a mixture of IL, TX-100 and toluene under our
calculation setting. In this subregion, the continuous phase is formed by oil with low conductivity, and the dispersion phase is
formed by ILs with high conductivity, which indicates that, as shown in Figs. 16c and 16d, the calculated result of Hanai
theory, namely, the conductivity of the continuous phase
m
is lower than the conductivity of the dispersed phase
p
, are
reasonable. With the increasing of IL weight fraction, the IL dissolved in the continuous phase (although very little) increases,
and the number of ions in the disperse phase also increases, which is why the calculated values of
m
and
p
both increase.
Journal Pre-proof
17
In the O/IL subregion, as listed in Table 2, for every sample, only when
m
is taken the value calculated according to the
formula below, can other phase parameters have certain physical significance.
m TX TX IL IL
(15)
where
TX
and
IL
are the volume fractions of TX-100 and IL, which are both calculated according to the weight ratio of
IL/TX-100 introduced in Section 2.1,
TX
and
IL
are the dielectric constant of TX-100 (7.59) and IL (13.06) [33]. This
shows that the continuous phase consists of the IL mixed with a certain amount of TX-100. There are two possible reasons
for this phenomenon: (1) the stable IL microemulsion needs a large number of surfactants; (2) there is no significant
difference between
TX
and
IL
, thus the contribution of TX-100 to the permittivity of the continuous phase cannot be
ignored. The similar situations have been found in [bmim][BF4]/TX-100/cyclohexane [33] and [bmim][BF4]/TX-
100/triethylamine [39].
In the O/IL microemulsion, with the increase of toluene weight fraction from 20.07 to 31.95 wt%, the volume fraction of
dispersed phase should increase because the oil molecules constantly solubilize into micropool of oil. However, the
calculated volume fraction of dispersed phase
decreased slightly by 2% instead as shown in Fig. 16e, compared to an
increase of nearly 12% in the oil weight fraction. This phenomenon can be explained as: as illustrated in Fig. 3a to Fig. 3c,
when the droplet changing from IL/O to O/IL with increasing IL weight fraction, the orientation of the TX-100 hydrophilic
chain changed from facing inward towards the IL micropool to facing outward towards the IL continuous phase, which
makes more space for the palisade layer to fill with more ILs, as illustrated by Fig. 17. This volume increase effect overcomes
the volume decrease caused by the decrease of oil weight fraction, as a result, the volume fraction of O/IL droplets increases
slightly. The value of
increases until close to the maximum stacking density of spheres 74%, which means that the
spherical microemulsion structure will soon be broken down as the weight fraction of the oil decrease (TX-100 content
decrease at the same time) because the single phase IL microemulsion will break down and become a multiphase system.
Similarly, in the IL microemulsion [bmim][BF4]/TX-100/triethylamine [39], although only the oil phase was different,
also decreases with the oil content increases in the O/IL microemulsion because the same surfactant (TX-100 with a long
hydrophilic chain) constitutes the palisade layer at the interface of oil droplets. In contrast, for IL microemulsion
[bmim][BF4]/ethanol/toluene in our preparation article, although IL and oil are the same,
in the O/IL microemulsion
increases as the oil weight fraction increases because the palisade layer at the interface of oil droplets is ethanol which does
not have a long hydrophilic chain.
Journal Pre-proof
18
Fig. 17. The volume fraction of dispersed phase
increases slightly, although the oil weight fraction decreases.
In the O/IL droplets, the hydrophilic chain of TX-100 directing towards the outside continuous phase, thus more space
for the palisade layer to fill with ILs, which is also the reason why the electrical conductivity of the dispersed phase
p
further increases with the decrease of weight fraction of toluene (the weight fraction of IL increases) as shown in Fig. 16g. In
the O/IL subregion,
m
(Fig. 16f) is less than
p
(Fig. 16g), possibly because the binary solvent, which contains a large
amount of nonionic surfactant, is used as the continuous phase. The conductivity of the continuous phase of the IL/O
microzone (Fig. 16c) is less than that of the O/IL microzone (Fig. 16f), which is also reasonable, because the continuous
phase of the former is just oil, and the continuous phase of the latter is an IL/TX-100 binary solvent that contains ILs. In
conclusion, according to our calculation results, no matter the O/IL or IL/O subregions, the dispersed phase in the IL based
microemulsion has a higher conductivity than traditional microemulsions. This conclusion is supported by a qualitative report
in the literatures [8,9].
In summary, Hanai theory successfully models IL/O and O/IL microregion as spherical particles with high volume fraction,
and the calculation results is reasonable, thus indicates that: (1) again confirms the phase transition between these three
microregions in our measured experimental path; (2) the concentration ranges chosen for these two spherical microregions
are reasonable; (3) there indeed exists interfacial polarization in these spherical dispersed particle; (4) the large volume of
droplets leads to a small vibration range of them; the large volume fraction of droplets makes them easier to contact each
other and connect together to form a double channel structure; moreover, a large number of surfactants exist, the surfactants
have long hydrophilic chains, the hydrophilic chains interact strongly with the IL, which together make the ions in the
droplets difficult to break through the thick surfactant layer coated in the surface. As a result, the ion is hard to exchange
between the droplets, thus the structure of percolation clusters cannot form; these may be the reasons why the static
percolation occurs in this IL microemulsion.
4.5 Comparison of dielectric spectra of the two kinds of IL microemulsion
Journal Pre-proof
19
Fig. 18. Comparison of complex dielectric spectra for [bmim][BF4]/TX-100/toluene (○ permittivity □ dielectric loss) and
[bmim][BF4]/TX-100/triethylamine (● permittivity, ■ dielectric loss) with the same molar fraction of IL/surfactant/oil (a) 20, 30 and 50 %,
(b) 15, 50 and 35 %.
The dielectric spectra of sample (I) and sample (III) described in Section 2.1 are placed in Fig. 18a, the hollow circle and
hollow square are the permittivity and dielectric loss data of sample (I), the solid circle and solid square are the permittivity
and dielectric loss data of sample (III). Similarly, dielectric spectra of samples (II) and (IV) are shown in Figure 18b, the
hollow symbols are the data of sample (II), the solid symbols are the data of sample (IV), the circular symbols represent the
permittivity, and the square symbols represent the dielectric loss. As can be seen from Figure 18a, although the oil phase that
compose the microemulsion are different, both the permittivity and dielectric loss spectra are very similar, and there is no
essential difference. Figure 18b even shows the spectra almost overlapping with each other, no matter permittivity and
dielectric loss data. This phenomenon can be explained as: the dielectric constants of the two oil phases, namely toluene (2.40)
and triethylamine (2.44), are very close to each other; and the substances that constitute the interface film (TX-100) and the
polar phase ([bmim][BF4]) are the same. In conclusion, the difference between the dielectric properties of the continuous
phase and the dispersed phase is the main factor for the dielectric relaxation phenomenon in this frequency band, while the
molecular details of the oil phase are the secondary factor.
5. Conclusion
We studied the dielectric properties of three subregions for IL microemulsion [bmim][BF4]/TX-100/toluene. The three
microstructures of this IL-based microemulsion, namely, IL-in-oil, B.C. and oil-in-IL were divided by the weight fraction
dependence of dielectric parameters. Both the conductivity behavior and dielectric behavior indicate that the static
percolation occurs when the phase change from IL/O droplet to O/IL droplet, which infers that a network of IL tubes in an oil
matrix or a network of oil tubes in a IL matrix exist in this system. The only dielectric relaxation observed may be caused by
the interfacial polarization, which indicates that there indeed exists interface in IL/O and O/IL microdroplets caused by the
dielectric difference between the dispersed phase and continuous phase. The conductivity and permittivity of the droplet and
the continuous phase were calculated quantitatively. The electrical conductivity of droplets in the IL microemulsion is much
higher than that of droplets in the traditional microemulsion. The dielectric relaxation is more affected by the static dielectric
constant of each component than by the molecular structural details of oil.
Journal Pre-proof
20
Acknowledgement
This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.
References
[
1
] A. Aljuhani, W.S. El-sayed, P.K. Sahu, N. Rezki, M.R. Aouad, R. Salghi, M. Messali, Microwave-
assisted synthesis of novel imidazolium, pyridinium and pyridazinium-based ionic liquids and / or
salts and prediction of physicochemical properties for their toxicity and antibacterial activity, J.
Mol. Liq. 249 (2018) 747753.
[
2
] B. K. Paul and S. P. Moulik, Ionic liquid-based surfactant science: Formulation,
characterization, and applications. Weinheim: Wiley-VCH, 2015.
[
3
] Sivapragasam M., Wilfred C.D. (2020) Biological Applications of Ionic Liquids-Based Surfactants:
A Review of the Current Scenario. In: Inamuddin, Asiri A. (eds) Nanotechnology-Based Industrial
Applications of Ionic Liquids. Nanotechnology in the Life Sciences. Springer, Cham.
[
4
] W. Qian, T. John and Y. Feng, Frontiers in poly(ionic liquid)s: syntheses and applications, Chem.
Soc. Rev., 2017, 46, 1124-1159.
[
5
] M. Hejazifar, O. Lanaridi, K. Bica-Schröer, Ionic liquid based microemulsions: A review, Journal
of Molecular Liquids 303 (2020) 112264.
[
6
] L. G. Chen, S. H. Strassburg, H. Bermudez, Micelle co-assembly in surfactant/ionic liquid
mixtures, J. Colloid Interface Sci. 477 (2016) 40-45.
[
7
] O. Zech, S. Thomaier, A. Kolodziejski, D. Touraud, I. Grillo, W. Kunz, Ethylammonium nitrate in
high temperature stable microemulsions, J. Colloid Interface Sci. 347 (2) (2010) 227232.
[
8
] H. Gao, J. Li, B. Han, W. Chen, J. Zhang, R. Zhang, D. Yan, Microemulsions with ionic liquid
polar domains, Phys. Chem. Chem. Phys. 6 (11) (2004) 29142916.
[
9
] J. Eastoe, S. Gold, S. E. Rogers, A. Paul, T. Welton, R. K. Heenan, I. Grillo, Ionic liquid-in-
oil microemulsions. J Am Chem Soc. 127 (2005) 73027303.
Journal Pre-proof
21
[
10
] D. Chakrabarty, D. Seth, A. Chakraborty, N. Sarkar, Dynamics of solvation and rotational
relaxation of coumarin 153 in ionic liquid confined nanometer-sized microemulsions. J. Phys. Chem. B
109 (2005) 57535758.
[
11
] Y. Gao, S. Wang, L. Zheng, S. Han, X. Zhang, D. Lu, L. Yu, Y. Ji, G. Zhang, Microregion
detection of ionic liquid microemulsions. J Colloid Interface Sci. 301 (2006) 612616.
[
12
] Y. Gao, J. Zhang, H. Xu, X. Zhao, L. Zheng, X. Li, L. Yu, Structural studies of 1-butyl-3-
methylimidazolium tetrafluoroborate/TX-100/p-xylene ionic liquid microemulsions. ChemPhysChem 7
(2006) 15541561.
[
13
] Y. Gao, N. Li, L. Zheng, X. Bai, L. Yu, X. Zhao, J. Zhang, M. Zhao, Z. Li, Role of solubilized
water in the reverse ionic liquid microemulsion of 1-butyl-3-methylimidazolium tetrafluoroborate/TX-
100/benzene. J. Phys. Chem. B;111 (2007) 25062513.
[
14
] N. Li, Y Gao., L. Zheng, J. Zhang, L. Yu, X. Li, Studies on themicropolarities of bmimBF4/TX-
100/toluene ionic liquid microemulsions and their behaviors characterized by UVVisible spectroscopy.
Langmuir 23 (2007) 10911097.
[
15
] Y. Gao, N. Li, S. Zhang, L. Zheng, X. Li, B. Dong, L. Yu, Organic solvents induce the formation
of oil-in-ionic liquid microemulsion aggregations. J. Phys. Chem. B 113 (2009) 13891395.
[
16
] S. Ghosh, C. Banerjee, S. Mandal, V. G. Rao, N. Sarkar, Effect of alkyl chain of room
temperature ionic liquid (RTILs) on the phase behavior of [C2mim][CnSO4]/TX-100/cyclohexane
microemulsions: solvent and rotational relaxation study. J. Phys. Chem. B 117 (2013) 58865897.
[
17
] S. Mandal, S. Ghosh, C. Banerjee, J. Kuchlyan, D. Banik, N. Sarkar, A novel ionic liquid in-oil
microemulsion composed of biologically acceptable components: an excitation wavelength dependent
fluorescence resonance energy transfer study. J. Phys. Chem. B 117 (2013) 32213231.
[
18
] J. Li, J. Zhang, H. Gao, B. Han, L. Gao, Nonaqueous microemulsion-conta iningionic liquid
[bmim][PF6] as polar microenvironment, Colloid. Polym. Sci. 283 (12) (2005) 13711375.
Journal Pre-proof
22
[
19
] Y. Zheng, W. Eli, Study on the polarity of [bmim][PF6]/tween80/toluene microemulsion
characterized by UVvisible spectroscopy, J. Dispers. Sci. Technol. 30 (5) (2009) 698703.
[
20
] S. Cheng, X. Fu, J. Liu, J. Zhang, Z. Zhang, Y. Wei, B. Han, Study of ethylene glycol/TX-
100/ionic liquid microemulsions, Colloids Surf. A Physicochem. Eng. Asp. 302 (1-3) (2007) 211215.
[
21
] R. Pramanik, S. Sarkar, C. Ghatak, V.G. Rao, P. Setua, N. Sarkar, Microemulsions with
surfactant TX-100, cyclohexane, and an ionic liquid investigated by conductance, DLS, FTIR
measurements, and study of solvent and rotational relaxation within this microemulsion, J. Phys.
Chem. B 114 (22) (2010) 75797586.
[
22
] S. Sarkar, R. Pramanik, C. Ghatak, V.G. Rao, N. Sarkar, Characterization of 1-ethyl-3-
methylimidazolium bis(trifluoromethylsulfonyl) imide ([Emim][Tf2N])/TX-100/cyclohexane ternary
microemulsion: investigation of photo induced electron transfer in this RTIL containing
microemulsion, J. Chem. Phys. 134 (7) (2011) 074507.
[
23
] R. Pramanik, S. Sarkar, Ghatak C., V. G. Rao, N. Sarkar, Ionic liquid containing microemulsions:
probe by conductance, dynamic light scattering, diffusion-ordered spectroscopy NMR measurements, and
study of solvent relaxation dynamics. J. Phys. Chem. B 115 (2011) 23222330.
[
24
] D. Blach, J. J. Silber, N. M. Correa, R. D. Falcone, Electron donor ionic liquids entrapped in
anionic and cationic reverse micelles. Effects of the interface on the ionic liquidsurfactant
interactions. Phys. Chem. Chem. Phys. 15 (2013) 1674616757.
[
25
] J. C. Thater, V. Gérard, C. Stubenrauch, Microemulsions with the ionic liquid ethylammonium
nitrate: phase behavior, composition, andmicrostructure. Langmuir 30 (2014) 82838289.
[
26
] R. Atkin, G. G. Warr, Phase behavior and microstructure of microemulsions with a room-
temperature ionic liquid as the polar phase. J. Phys. Chem. B 111 (2007) 93099316.
[
27
] O. C. Thater, T. Sottmann, C. Stubenrauch, Alcohol as tuning parameter in an IL containing
microemulsion: the quaternary system EAN-n-octane-C12E3-1-octanol. Colloids Surf., A 494 (2016) 139
146.
Journal Pre-proof
23
[
28
] F. Wang, Z. Zhang, D. Li, J. Yang, C. Chu, L. Xu, Dilution method study on the interfacial
composition, thermodynamic properties, and structural parameters of the [bmim][BF4] + Brij-35 + 1-
butanol + toluene microemulsion. J. Chem. Eng. Data 56 (2011) 33283335.
[
29
] R. Pramanik, S. Sarkar, C. Ghatak, P. Setua, N. Sarkar, Effect of polymer, poly(ethylene
glycol)(PEG-400), on solvent and rotational relaxation of coumarin-480 in an ionic liquid containing
microemulsions. Phys. Chem. Chem. Phys. 12 (2010) 38783886.
[
30
] J. Xu, L. Zhang, A. Yin, W. Hou, Y. Yang, Nonaqueous ionic liquid microemulsions of 1-butyl-3-
methylimidazolium tetrafluoroborate, toluene and ethanol, Soft Matter 9 (2013) 64976504.
[
31
] F. Xu, L. Chen, A. Wang, Z. Yan, Influence of surfactant-free ionic liquid microemulsions
pretreatment on the composition, structure and enzymatic hydrolysis of water hyacinth, Bioresour.
Technol. 208 (2016) 1923.
[
32
] Y. Lian, K. Zhao, Dielectric analysis of micelles and microemulsions formed in a hydrophilic
ionic liquid. I. Interaction and percolation, J. Phys. Chem. B 115 (2011) 1136811374.
[
33
] Y. Lian, K. Zhao, Broadband dielectric spectroscopy of micelles and microemulsions formed in a
hydrophilic ionic liquid: the relaxation mechanism and interior parameters, New J. Chem. 42 (2018)
26052615.
[
34
] J. Wang , W. Li , K. Zhao, Effects of ionic liquids on microstructure and thermal stability of
microemulsions by broadband dielectric spectroscopy, Colloids Surf., A 610 (2021) 125739.
[
35
] K. Chen, K. Zhao, Dielectric analysis of the [Bmim][PF6]/TX-100/ethyleneglycol nonaqueous
microemulsions: microstructures and percolation, Colloids Surf., A 461 (2014) 5056.
[
36
] K. Chen, K. Zhao, Dielectric analysis on the phase behavior of ionic liquid-containing
nonaqueous microemulsions, Colloid Polym. Sci. 293 (2015) 833840.
[
37
] C. Zhang, Z. Zhen, L. Ma, K. Zhao, Dielectric relaxation of nonaqueous ionic liquid
microemulsions: polarization, microstructure, and phase transition, RSC Adv. 7 (2017) 1373313741.
Journal Pre-proof
24
[
38
] C. Zhang, H. Rao, K. Zhao, Dielectric insights into the microcosmic behavior of ionic liquid-
based self-assembly—microemulsions/micelles, J. Phys. Chem. B 122 (2018) 7170−7177.
[
39
] Z. Li, Z. Fan, Y. Lian, Z. Chen, Dielectric analysis of percolation, interface polarization,
and phase behavior of 1-butyl-3-methylimidazolium tetrafluoroborate/TX-100/triethylamine
microemulsions, New J. Chem. 45 (2021) 1716317175.
[
40
] Z. Chen, R. Nozaki, Time-dependent phase behavior of 5CB-DDAB-water microemulsions monitored by
means of dielectric spectroscopy. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 2011, 84 (1 Pt 1),
011401.
[
41
] A. Ponton, T.K. Bose, Dielectric study of percolation in an oil-continuous microemulsion, J.
Chem. Phys. 94 (1991) 6879−6886.
[
42
] C. Daguenet, P. J. Dyson, I. Krossing, A. Oleinikova, J. Slattery, C. Wakai, H. Weingärtner,
Dielectric response of imidazolium-based room-temperature ionic liquids, J. Phys. Chem. B, 110 (2006)
12682−12688.
[
43
] F. Bordi, C. Cametti, S.Sennato, D. Viscomi, Radiofrequency dielectric loss relaxation in
polyion-induced liposome aggregates, J. Colloid Interface Sci., 309 (2007) 366-372.
[
44
] K. S. Cole, R. H. Cole, Dispersion and absorption in dielectrics I. Alternating current
characteristics, J. Chem. Phys. 9 (1941) 341-351.
[
45
] K. Asami, Characterization of heterogeneous systems by dielectric spectroscopy, Prog. Polym.
Sci. 27 (2002) 16171659.
[
46
] M. Laguës, Electrical conductivity of microemulsions : a case of stirred percolation. J. Phys.,
Lett. 40 (1979) 331−333.
[
47
] G. S. Grest, I. Webman, S.A. Safran, A.L.R. Bug, Dynamic percolation in microemulsions. Phys.
Rev. A: At., Mol., Opt. Phys. 33 (1986) 2842−2845.
[
48
] Y. Talmon, S. Prager, Statistical thermodynamics of phase equilibria in microemulsions. J. Chem.
Phys. 69 (1978) 2984−2991.
Journal Pre-proof
25
[
49
] C. Mathew, Z. Saidi, J. Peyrelasse, C. Boned, Viscosity, conductivity, and dielectric
relaxation of waterless glycerol-sodium bis(2-ethylhexyl)sulfosuccinate-isooctane microemulsions:
The percolation effect, Phys. Rev. A 43(2) (1991), 873−882.
[
50
] S. Schrödle, R. Buchner, W. Kunz, Percolating microemulsions of nonionic surfactants probed by
dielectric spectroscopy, ChemPhysChem 6 (2005) 1051 1055.
[
51
] A. Maitra, C. Mathew and M. Varshney, Closed and open structure aggregates in microemulsions
and mechanism of percolative conduction. J. Phys. Chem., 1990, 94, 5290-5292.
[
52
] S. Hait, A. Sanyal and S. Moulik, Physicochemical studies on microemulsions. 8. The effects of
aromatic methoxy hydrotropes on droplet clustering and understanding of the dynamics of conductance
percolation in water/oil microemulsion systems. J. Phys. Chem. B, 2002, 106, 12642-12650.
[
53
] X. Zhang, Y. Chen, J. Liu, C. Zhao and H. Zhang, Investigation on the structure of
water/AOT/IPM/alcohols reverse micelles by conductivity, dynamic light scattering, and small angle
X-ray scattering. J. Phys. Chem. B 116 (2012) 3723-3734.
[
54
] K. He, K. Zhao, J. Chai, G. Li, Dielectric analysis of the APG/n-butanol/cyclohexane/water
nonionic microemulsions, J. Colloid Interface Sci. 313 (2007) 630637.
[
55
] K. Chen and K. Zhao, Dielectric relaxation behavior of ternary systems of water/toluene/Triton
X-100: the effects of water and oil contents on microemulsion structure, Colloid Polym. Sci. 292
(2014) 557566.
[
56
] R.W. O’Brein, The response of a colloidal suspension to an alternating electric field, Adv.
Colloid Interface Sci. 16(1) (1982) 281320.
[
57
] R.W. O’Brien, The high-frequency dielectric dispersion of a colloid, J. Colloid Interface Sci.
113(1) (1986) 8193.
[
58
] S.S. Dukin and V.N. Shilov, Dielectric phenomena and the double layer in disperse system and
polyelectrolytes, Wiley, New York, 1974.
Journal Pre-proof
26
[
59
] S.S. Dukhin, Electrochemical characterization of the surface of a small particle and
nonequilibrium electric surface phenomena, Adv. Colloid Interface Sci. 61 (1995) 1749.
[
60
] D. Langevin, Micelles and microemulsions, Annu. Rev. Phys. Chem. 43(1) (1992) 341369.
[
61
] J. Sjöblom, R. Lindberg, S. E. Friberg., Microemulsions—phase equilibria characterization,
structures, applications and chemical reactions, Adv. Colloid Interface Sci. 65(96) (1996) 125-287.
[
62
] U .Olsson, K. Shinoda, B. Lindman, Change of the structure of microemulsions with the
hydrophile-lipophile balance of nonionic surfactant as revealed by NMR self-diffusion studies, J.
Phys. Chem. 90(17) (1986) 4083-4088.
CRediT authorship contribution statement
Zhen Li: Conceptualization, Methodology, Formal analysis, Investigation, Resources, Data Curation,
Writing - Original Draft. Zhefeng Fan: Investigation, Resources, Writing - Review & Editing,
Supervision. Zhen Chen: Resources. Yiwei Lian: Resources.
Declaration of interests
The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.
The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:
Journal Pre-proof
27
Graphical Abstract
The dielectric properties of three subregions, namely, IL-in-toluene, bicontinuous and toluene-in-IL for
non-aqueous ionic liquid microemulsion [bmim][BF4]/TX-100/toluene.
Journal Pre-proof
... 24−35 At present, this method has been used to study the microemulsion formed by IL replacing only water 25,29−35 or only oil. 24,[26][27][28]32 In view of the distinct advantages that dielectric spectra can be used to judge the microemulsion percolation mechanism by multiple physical quantities simultaneously, in these studies, the percolation mechanisms of IL microemulsions such as H 2 33,34 frequency dependence of permittivity 25,33−35 and loss angle. 33−35 The multipolarization of IL microemulsions has been studied by DRS: the interfacial polarization is observed, 27,28,31,33−35 which is strong evidence that interfaces exist in such systems; other dielectric relaxation mechanisms are directional movement of the IL anions along the PEO chains of TX-100 29,30 or the result of superimposed dipole rotation polarization of cations on the long axis direction and the polarization of the EO segment in TX-100. ...
... Percolation Theories. With temperature being constant, as the weight fraction of the dispersed phase increases and passes a critical composition (c p ), microemulsions undergo a percolation process, 25,29,[33][34][35]39 namely, the structure of the microemulsion changes from a water-in-oil (W/O) droplets to oil-in-water (O/W) droplets. The schematic diagrams of these two subregions are illustrated in Figure 2; because the PIL replaces the polar phase water in this paper, Figure 2a,c represents the subregions of PIL-in-oil (PIL/O) and oilin-PIL (O/PIL), respectively. ...
... This section will first discuss the calculated relaxation time to obtain the relaxation mechanism and then discuss the other phase parameters in the following section. The high-frequency relaxation time of the PIL/O subregion samples in this paper, as shown in Table 1 34 The dielectric relaxations of IL microemulsions in the literature 33−35 have been proved to be caused by interfacial polarization, which indicates that the high-frequency dielectric relaxation in this paper may also be caused by interfacial polarization. A more reliable way to judge the percolation mechanism is to compare the experimental and theoretical values of relaxation time, Figure 6e not only includes the experimental value τ h of high-frequency dielectric relaxation but also shows the theoretical values τ MW of the interface polarization time, from which we can find that for seven samples τ MW values are nearly equal to τ h values, and their values for other samples are very close or at least in the same order of magnitude. ...
... Dielectric relaxation spectroscopy (DRS) is a powerful method for studying traditional [14][15][16][17][18][19][20], nonaqueous IL [21][22][23][24][25][26][27][28][29] and surfactant-free microemulsions [30,31]. DRS allows the determination of J o u r n a l P r e -p r o o f the percolation mechanism by using several different ways at the same time, such as the concentration dependence of static dielectric constant [14,15,28,29], relaxation frequency [14,15] and dc conductivity [15,16,[21][22][23][24][25][27][28][29]; frequency dependence of permittivity [15,16,21,28,29] and loss angle [15,16,28,29]. ...
... Dielectric relaxation spectroscopy (DRS) is a powerful method for studying traditional [14][15][16][17][18][19][20], nonaqueous IL [21][22][23][24][25][26][27][28][29] and surfactant-free microemulsions [30,31]. DRS allows the determination of J o u r n a l P r e -p r o o f the percolation mechanism by using several different ways at the same time, such as the concentration dependence of static dielectric constant [14,15,28,29], relaxation frequency [14,15] and dc conductivity [15,16,[21][22][23][24][25][27][28][29]; frequency dependence of permittivity [15,16,21,28,29] and loss angle [15,16,28,29]. DRS is an effective tool to study the multipolarization in these systems [17][18][19][20][21][22][23][24][25][26][27][28][29]. ...
... Dielectric relaxation spectroscopy (DRS) is a powerful method for studying traditional [14][15][16][17][18][19][20], nonaqueous IL [21][22][23][24][25][26][27][28][29] and surfactant-free microemulsions [30,31]. DRS allows the determination of J o u r n a l P r e -p r o o f the percolation mechanism by using several different ways at the same time, such as the concentration dependence of static dielectric constant [14,15,28,29], relaxation frequency [14,15] and dc conductivity [15,16,[21][22][23][24][25][27][28][29]; frequency dependence of permittivity [15,16,21,28,29] and loss angle [15,16,28,29]. DRS is an effective tool to study the multipolarization in these systems [17][18][19][20][21][22][23][24][25][26][27][28][29]. ...
Article
Full-text available
A dielectric relaxation was observed at approximately 700 MHz in the surfactant-free microemulsion which is composed of 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF4]), ethanol and toluene. The measured frequency was between 1 MHz and 3 GHz. Two inflection points (22.39 and 31.09 wt%) on the curve of the dielectric parameters vs weight fraction of IL give the phase boundaries between IL/O and B.C., B.C. and O/IL sub-regions. The permittivity of the frequency at the beginning of the relaxation phenomenon and the maximum of the derivative of dc conductivity both give 31.09 wt% as the percolation threshold. Both frequency dependences of permittivity and loss angle provide the critical exponent u = 0.72, which suggest that a static percolation occurs in the microemulsion near the critical threshold. This conclusion is also supported by the scaling dependence of dc conductivity on IL weight fraction. The results of phase parameters obtained by Hanai theory show that: the dielectric relaxation of this system is proved to be caused by the interfacial polarization, which indicates that dispersed particles indeed exist in this surfactant-free microemulsion; in the IL/O sub-region, disperse particle not only contains IL and TX-100, but also includes oil molecule; and for the O/IL sub-region the continuous phase should be the IL/ethanol binary solvent rather than the pure IL; the main reasons for the occurring of the static percolation in this system may be the large volume fraction of dispersed particle and the existent of IL molecule. The substitution of ethanol for surfactant causes change in interface properties but no change in percolation mechanism, indicating that ethanol indeed distributes at the interface and molecular details at the interface do not determine the mechanism of percolation
Article
Full-text available
Two nonaqueous ionic liquid (IL) microemulsions (toluene/TX-100/[bmim][PF6] and [bmim][BF4]/TX-100/benzene) were studied by dielectric spectroscopy covering a wide frequency range (40 Hz to 110 MHz). A unique relaxation was observed in the radio frequency (RF) range. By methodically analyzing the dependence of relaxation parameters on the ILs content, the microstructures of the microemulsions were identified. Additionally, based on the interfacial polarization theory and Einstein equation, the mechanism of the relaxation caused by the fluctuation of IL anions along the TX-100 PEO chain was confirmed, what's more, according to the dependence of the dc conductivity of the microemulsions on IL concentration, it was concluded that the hydrophilicity of the IL in the nonaqueous IL microemulsions may play a crucial role in the electrical conduction mechanism: our analysis results suggest that a dynamic percolation process occurs in the toluene/TX-100/[bmim][PF6] system in which IL is hydrophobic, while a static percolation happens in benzene/TX-100/[bmim][BF4] where IL is hydrophilic. The otherness of relaxation time provides evidence that there is a possible coupling effect between IL and TX-100. Moreover, there are hints that all of the disparities, such as relaxation time, percolation type, ion migration rate and the size of different micro zones, may just stem from the different hydrophobicity of the two kinds of IL.
Article
The dielectric properties of the nonaqueous microemulsions 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF4])/p-(1,1,3,3-tetramethylbutyl)phenoxypolyoxyethyleneglycol (TX-100)/triethylamine, in which ionic liquid (IL) serves as the polar phase, were measured as a function of IL weight fraction...
Article
Phase behaviors of TX-100/[bmim][PF6]/water and TX-100/[bmim][BF4]/cyclohexane systems were investigated by dielectric relaxation spectroscopy. The dielectric analysis of the two systems showed that there were two relaxations in both systems at 1 ns and 50 ps, besides, all of them were originated from superimposed dipole polarization although induced by different groups. The relaxation parameters, which were obtained by fitting Cole-Cole equations to the dielectric data, were used to discuss the phase behaviors and percolation dynamics of microemulsions. It was found that the TX-100/[Bmim][PF6]/ water system presented a temperature-dependent phase transition, while the TX-100/[Bmim][BF4]/ cyclohexane system was not sensitive to temperature in the measured range, which proved that the hydrophilic/hydrophobic ionic liquid has a great influence on the thermal stability of microemulsion. For the TX-100/[bmim][PF6]/water system, the critical percolation temperature Tp was obtained by analysing the conductivity properties and verified by analyzing dielectric increment and relaxation time. Besides, according to percolation theory, there is the dynamic percolation process occured and caused the phase transion of TX-100/[bmim][PF6]/water microemulsion, however, that doesn't occurred in TX-100/[Bmim][BF4]/ cyclohexane system, which caused the different situation between two microemulsions in the rising the temperature. This work provides a new insight into the phase transition of sub-regions of IL-based microemulsion.
Chapter
The paramount features of Ionic Liquids-Based Surfactant, (ILBS) stems from their remarkable properties such as low critical micelle concentration (CMC), high solubilization ability, and better wetting and foaming properties compared to commercially available surfactants. Currently, the types of ILBS synthesized includes those of imidazolium, pyrrolidinium, pyridinium, piperidinium, and amino acid based. The combination of surfactants and ionic liquids (ILs) are particularly relevant to applications such as heterogeneous catalysis and emulsion technologies that rely on the control over their interfacial properties. The ability of ILBS to depict features of ionic liquids (ILs) while aggregating in dilute aqueous solution, and the possibility of switching between these two states protrude applications led to its plentiful applications in the biological scope. This chapter provides a current overview on the biological applications of ILBS which included those of the drug delivery, biomolecule extraction, biocompound separation, enzyme catalysis, and others.
Article
Over the past few years an increasing number of studies dealt with microemulsions, in which either the aqueous phase, the oil phase, the surfactant or even two of these components constitute of an ionic liquid. Many examples demonstrated that water and oil are not necessarily the required polar and non-polar components that contribute to the formation of microemulsions but can be replaced by targeted ionic liquids. Moreover, in many cases ionic liquids are used as amphiphiles that assist in the formation of microemulsions in aqueous or non-aqueous media due to their unique surface activity. In the present review, we highlight the properties of three types of ionic liquid based microemulsions, namely non-aqueous ionic liquid microemulsions, aqueous ionic liquid microemulsions and ionic liquid/oil/water microemulsions. Various characterization techniques applied for the confirmation of their formation and the detailed elucidation of their properties are discussed herein. Additionally, we present a comprehensive summary of the recent applications of ionic liquid based microemulsions in different fields, such as synthesis, (bio-)catalysis, polymerization, (nano-)material preparation, drug delivery and separations.
Article
Dielectric relaxation spectroscopy of ([Bmim][BF4]/TX-100/P-xylene) microemulsions and ([Bmim][BF4]/TX-100) micelles ionic liquids (ILs) formed was measured. A specific dielectric relaxation dependent on the concentration of ILs was observed in the range of 106~108Hz. Combined dielectric parameters with the Einstein displacement equation and Bruggeman’s effective-medium approximation, the interaction between [Bmim][BF4] and TX-100 in microemulsions/micelles was presented: Due to electrostatic interaction and van der Waals' force, [Bmim][BF4] is bound around the PEO chains of TX-100, and once added electric field, ions of [Bmim][BF4] will move along the PEO chain. The dependence of dielectric parameters such as relaxation time, permittivity on the mass fraction of ILs presents an evidence for our proposals about the transition of both systems with the increase of ILs content. In addition, it was confirmed that percolation is a unique phenomenon in microemulsions and the percolation of [Bmim][BF4]/TX-100/P-xylene belongs to static percolation. The transition process of micelles with the change of ILs content is presented from the dielectric view.
Article
Three striking dielectric relaxations located at 100MHz, 1GHz and 10GHz respectively were observed both in binary mixtures of p-(1,1,3,3-tetramethylbutyl)phenoxypolyoxyethyleneglycol (TX-100) and 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF4]), and [bmim][BF4]/TX-100/cyclohexane microemulsions in a wide frequency range. The mechanisms of the dielectric relaxations were proposed and the detailed characterizations of low-frequency relaxation(around 100 MHz), which is identify as interface polarization, were analyzed theoretically. The dielectric parameters (relaxation time dielectric increment) were obtained by fitting the dielectric spectra with Debye plus Cole-Cole formula. The low-frequency relaxation mechanisms of both binary and ternary systems are contributed to interfacial polarization, supported by the good agreement between relaxation times calculated by Maxwell-Wagner theory and the values obtained from experiment. Meanwhile, by analyzing the linear ionic liquid content dependence of dielectric increment of low-frequency relaxation, we find that this relaxation involves the dynamics of ionic liquids which located at the interface. Besides, the phase parameters, which reflect the interior properties of the binary and ternary systems, were calculated by Hanai theory, and the dependence of them on the variation of sample composition are properly explained; the distribution of cyclohexane in dispersed phase is obtained quantitatively, which is of important in the study field of microstructure of ionic liquid microemulsions.
Article
An efficient, eco-friendly, simple and facile synthesis of a novel class of imidazolium, pyridinium and pyridazinium-based ionic liquids and/or salts is described under both conventional procedure and microwave irradiation. The newly synthesized ILs were well-characterized by FT-IR, ¹H NMR, ¹³C NMR, mass spectrometry and elemental analysis. Their in vitro antibacterial against a panel of Gram-positive and Gram-negative bacteria was measured by determination of the inhibition zone (IZ), minimum inhibitory concentration (MIC) and minimum bactericidal concentration (MBC) and the results revealed that ILs containing imidazolium cation are very effective antibacterial agents, especially, 1,2-dimethyl-3-(3-phenoxypropyl)-1H-imidazol-3-ium bromide 8. A correlation of structure and activities relationship of these ILs with respect to Lipinski rule of Five, drug likeness, toxicity profiles and other physico-chemical properties of drugs are described and verified experimentally.
Article
We review recent works on the synthesis and application of poly(ionic liquid)s (PILs). Novel chemical structures, different synthetic strategies and controllable morphologies are introduced as a supplement to PIL systems already reported. The primary properties determining applications, such as ionic conductivity, aqueous solubility, thermodynamic stability and electrochemical/chemical durability, are discussed. Furthermore, the near-term applications of PILs in multiple fields, such as their use in electrochemical energy materials, stimuli-responsive materials, carbon materials, and antimicrobial materials, in catalysis, in sensors, in absorption and in separation materials, as well as several special-interest applications, are described in detail. We also discuss the limitations of PIL applications, efforts to improve PIL physics, and likely future developments.
Article
Hypothesis: The phase behavior of amphiphiles is known to depend on their solvent environment. The organic character of ionic liquids suggested the possibility to tune surfactant aggregation, even in the absence of water, by selection of appropriate ionic liquid chemistry. To that end the behavior of the surfactant sodium dodecylsulfate in a chemically similar imidazolium ionic liquid, 1-ethyl-3-methyl imidazolium ethylsulfate, was explored. Experiments: The solubility of sodium dodecylsulfate in 1-ethyl-3-methyl imidazolium ethylsulfate was determined, establishing the Krafft temperature. Tensiometry was performed to obtain interfacial properties such as the surface excess and area per molecule. Pulsed-field gradient spin-echo NMR was used to determine the diffusion coefficients of all the major species, including micelles, as a function of surfactant concentration. Importantly, all three methods provided consistent values for the critical micelle concentration. Findings: Analysis of tensiometry data suggests, and is confirmed by NMR results, that the ionic liquid ions are incorporated along with surfactants into micelles, revealing a complex micellization behavior. In light of these findings past studies with ternary mixtures of surfactants, ionic liquids, and water may merit additional scrutiny. Given the large number of ionic liquids, this work suggests opportunities to further control micelle formation and properties.