You are on page 1of 10

Article

pubs.acs.org/JPCC

Proton Microenvironment and Interfacial Structure of Sulfonic-AcidFunctionalized Ionic Liquids


Weida Shan, Qiwei Yang, Baogen Su, Zongbi Bao, Qilong Ren, and Huabin Xing*
Key Laboratory of Biomass Chemical Engineering of Ministry of Education, College of Chemical and Biological Engineering, Zhejiang
University, Hangzhou 310027, China

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

S Supporting Information
*

ABSTRACT: Sulfonic-acid-functionalized ionic liquids


(SFILs) have shown promising performance in catalytic
reactions and development of new proton conductive
materials. In this work, molecular dynamic simulation based
on the ion pair charge approach has been performed to
investigate the structural characteristic of several typical SFILs
in both bulk liquids and the vacuumliquid interfacial region.
The results showed that anions play a critical role in
determining the microenvironment of the terminal sulfonic
acid proton in SFILs, and less basic anions lead to a much
weaker association of sulfonic acid protons with anions. A
signicant aggregation of sulfonic acid side chains existed in SFILs because of the strong interaction between dierent sulfonic
acid groups. A sharp change in the ordering preference of alkyl tails and pyridine rings at vacuumliquid interfaces was observed
after the introduction of a sulfonic acid group to the side chain of the cation of ionic liquids, and the properties of the anions have
a remarkable inuence on the preferential location of anions in the interfacial region. These results can aid in understanding of
the physiochemical properties of SFILs and thus facilitate the development of new SFILs and proton conductive materials.

1. INTRODUCTION
Over the past decade, Brnsted acidic materials have attracted
enormous attention for various applications, such as catalysis,
fuel cell electrolytes, and proton exchange membrane.19
Among the many types of Brnsted acidic materials, sulfonicacid-functionalized Brnsted acidic ionic liquids (SFILs) have
been considered one of the most promising species. Cole et al.
designed the rst SFILs, which featured an alkanesulfonic acid
group covalently tethered to the imidazolium or phosphonium
cation, and then, a variety of SFILs including ammonium-based
SFILs, pyridinium-based SFILs,10 benzimidazole-based SFILs,11
and even modied silica-based SFILs12 have been synthesized
in the past decade. These SFILs have been widely used as novel
acid catalyst or dual solventcatalyst for alkylation reaction,
Fischer esterication,10,13,14 polymerization,15,16 biodiesel production,17,18 and hydrolysis of cellulose,12,19 due to their
distinctive properties, such as high catalytic activities, tunable
acidities, superior reactionseparation coupling behavior and
recyclability. In addition, acidic ionic materials bearing an active
proton or an available proton site have recently been the
subject of extensive study for their possible use in constructing
water-free proton transport channels5,6,2025 by suitably
designing their molecular shape and nanostructured aggregation, which can be useful for the development of new
electrolytes for the next generation of fuel cells. The
microenvironment of the proton, i.e., the microstructure and
interaction around the acidic proton, is the key factor for these
Brnsted acidic materials in determining the proton con 2015 American Chemical Society

ductivity as well as the acid catalytic activity. However, the


current understanding of the proton microenvironment and
heterogeneous nanoaggregation in SFILs remains very limited,
which signicantly hinders the design and application of SFILs
and their derivative materials.
An accurate molecular description of the gasliquid
interfacial structure is quite important for the use of ionic
liquids (ILs)26 in electrochemical application,2729 gas-separation process,3033 and heterogeneous catalytic reaction.34 Even
though a large number of experimental and simulation studies
have been conducted to illustrate the interfacial structure of
some functionalized ILs such as amine-functionalized ILs,
alkoxysilane-functionalized ILs, and thioether-functionalized
ILs,35,36 the interfacial structure of SFILs remains unclear to
many researchers and hinders the application of these materials
in catalysis reaction and fuel cell design.
Therefore, in this study, we focused on investigating the
structural characteristics of some typical SFILs (Figure 1) in
both bulk liquid and gasliquid interfacial regions through
molecular dynamic (MD) simulation, including N-propanesulfonic acid pyridinium triuoromethylsufate ([PSPy][Tfo]),
N-propane-sulfonic acid pyridinium hydrogen sulfate([PSPy][HSO4]), N-propane-sulfonic acid pyridinium dihydrogen
phosphate([PSPy][H2PO4]), and N-hexane-sulfonic acid pyrReceived: March 24, 2015
Revised: July 5, 2015
Published: August 14, 2015
20379

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

Article

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

The Journal of Physical Chemistry C

Figure 1. Schematic structure and atom labels of the studied ILs.

at +1 e and 1 e, respectively.28,32,36,4145 However, this result


is in fact an ideal case, which requires the assumption that every
ion was isolated from each other with negligible charge-transfer
interionic interactions. The cationanion interaction in
functionalized ILs, such as SFILs, is much stronger than that
in common ILs; thus, the common unit charge will deviate for
these ILs and may lead to inaccuracy in the simulation results.
For example, it has been conrmed by experiment that unit
charge could lead to an overestimated electrostatic interaction
and thus slow down the simulated dynamic properties even if
for the nonfunctionalized IL.46 In addition, recently, a new
charge derivation approach47 based on united IL pairs has been
proposed and showed a better prediction performance for
experimental results than isolated ion-based charge derivation.
Therefore, for all ILs in this work, the molecular geometries
of ion pairs were optimized at the B3LYP/6-311++g(d,p) level,
and then we adopted a similar approach based on the optimized
ion pair structure to derive the atomic charges using the
restrained electrostatic potential (RESP)48 method instead of
the traditional isolated-ion approach. The derived total charge
of cation and anion was about +0.8 e and 0.8 e, respectively,
signicantly deviated from the assumption used in the isolatedion approach. The ion-pair charge approach can compensate
the strong charge-transfer and polarization phenomenon that
occurs in the microscopic environment of SFILs, and is crucial
for good simulation of the bulky and interfacial structures of
SFILs. In fact, when we tried to use the isolated-ion approach
to derive atomic charges, the simulation always failed due to the
program errors.
2.3. Simulation Details. MD simulations were performed
with the GROMACS 4.5.449,50 software package. Each
simulation system was composed of 512 ion pairs. The leapfrog
algorithm with a time step of 1.0 fs was adopted to integrate
Newtons equation of motion. The Lennard-Jones interaction
and electrostatic forces were truncated at a radius of 1.5 nm.
The Coulombic interaction was treated with the particle-mesh
Ewald (PME) summation. Given that SFILs generally have

idinium triuoromethylsufate ([HSPy][Tfo]). Moreover, Npropyl-pyridinium triuoromethylsufate ([PPy][Tfo]) with no


sulfonic acid functionalization was also investigated as a
benchmark for comparison (Figure 1). The resulting
information sheds light on the proton microenvironment and
interfacial structure of SFILs, which are highly instructive for
the further design of SFILs and new proton conductive
materials.

2. MODELS AND SIMULATION DETAILS


2.1. Force Field Model. A standard all-atom force eld was
used to model molecular interaction among dierent species,
where the total potential energy consists of four kinds of
potential energy, as shown in eq 1:
E=

K r(r r0)2

Bonds

Dihedrals
N

i=1

K ( 0)2

Angles

[1 + cos(n )]

12 6

qq

4ij ij ij + i j
r
rij
rij
j=i+1
ij

(1)

where N is the number of atoms, ij and ij are the LennardJones parameters for atom pair, and qi is the partial charge on
atom i. The Lennard-Jones terms for cross interactions were
derived using LorentzBerthelot combining rules. The forceeld parameters including the atom types, Lennard-Jones
parameters, and molecular terms were mostly taken from the
Generalized Amber force eld (GAFF), which was found to be
capable of simulating imidazolium-based ILs and other
functional ILs with good prediction performance by many
researchers.32,3638,43 The force-eld parameters developed by
Lopes et al.3840 was adopted for several functional groups in
this work.
2.2. Charge Parameterization. For most of the reported
simulations of ILs, the charges of the cation and anion were set
20380

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

Article

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

The Journal of Physical Chemistry C

and [PSPy][Tfo], as shown in Figure 2. In both ILs, the rst


maximum value of RDFs between the negatively charged
oxygen atom of anion (O23) and the carbon atoms of pyridine
ring of cation (C4, C5, C6) are more than 1.5, indicating the
preferential location of anions around the pyridine ring of the
cations. In addition, the intensity of the rst peak decreases in
the order of RDF O23C4 > RDF O23C5 > RDF O23C6,
which is similar to those observed in a common IL 1butylpyridinium tetrauoroborate and indicates that the anions
prefer to locate near the side chain.45 However, the RDFs for
the terminal carbon of alkyl chain with anions (O23C9 curves
in Figure 2a and Figure 2b) show a sharp contrast in both ILs,
where the intensity of the rst peak of RDF O23C9 in
[PPy][Tfo] (Figure 2a) is only 1.1, but that in [PSPy][Tfo]
(Figure 2b) is up to 1.9. Moreover, the maximum peak of RDFs
between O23 and the oxygen atoms of the sulfonic acid group
(O13) (pink curves in Figure 2b) shows a similar behavior to
the RDFs of O23C4, which illustrates the fact that, after
introducing a sulfonic acid group into the terminal position of
the alkyl chain of [PPy][Tfo], a new strong interaction site
appears between the sulfonic acid group of cation and the
anion.
We also calculated the spatial distribution functions (SDFs)
of anions around the cations in [PPy][Tfo] and [PSPy][Tfo].
As shown in Figure 3a, a concentrated distribution of red
isosurfaces of [Tfo] anion can be found around the N atom of
the pyridine ring in [PPy][Tfo], and it is away from the end of
the alkyl chain. However, in Figure 3b for [PSPy][Tfo], the
protuberant isosurface around the sulfonic acid group
(especially around the sulfonic acid proton) indicates a high
probability of the anions [Tfo] distributed around the sulfonic
acid group in the end of alkyl chain. This phenomenon also
supports the strong interaction between the sulfonic acid group
and anions. The results of SDFs were consistent with the RDFs
analysis.
The unique properties of SFILs, such as their catalytic
activities and proton conductivities, depend signicantly on the
microenvironment of the proton in the sulfonic acid
group.5,52,53 Therefore, the RDFs between the sulfonic acid
proton (H13) of [PSPy]+ and the main oxygen atom of the
dierent anions (i.e., the oxygen atom, which exhibits a stronger
interaction with H13 than the other oxygen atoms) have been
computed to investigate the microenvironment of the sulfonic
acid proton (Figure 4). The rst maximum peak is located at a
distance of 2.29 for [PSPy][H2PO4], but 2.40 for
[PSPy][HSO4] and 2.48 for [PSPy][Tfo], and the peak
intensities for these three SFILs are 5.45, 3.50 and 3.35,

high viscosity and are applied at relatively high temperatures,


the system was initially simulated at 298 K to derive the bulk
density data in the canonical (NPT) ensemble, and then
simulated at 373 K for a duration of 1.5 ns with the trajectory of
the last 0.5 ns collected to analyze the bulk structure. After
NPT equilibrium was reached, the periodic boundary cubic box
of ILs was elongated along the Z-axis, creating a liquid slab
perpendicular to Z-axis in the middle of the box with a vacuum
phase on either side. Following this step, the system was
equilibrated for another 7 ns at 423 and 473 K in the canonical
(NVT) ensemble, and the trajectory of the last 3 ns was derived
for subsequent analysis to decrease the sampling noise. Similar
approaches of increasing simulation temperature have been
adopted by many researchers in MD simulations to improve the
sampling behavior32,33,42,44,51 due to the slow dynamic
properties of highly viscous ILs, and the simulation temperatures in this work was determined by considering both the
common catalytic reaction temperature and the electrolyte
operation temperature for SFILs. In all cases, the trajectory data
for structure analysis was collected every 0.5 ps, and the
Berendsen thermostat/barostat algorithm was utilized for
environment control. As no force-eld parameter improvement
was performed to t the experimental values, the predicted
densities for the investigated SFILs from MD simulation are in
good agreement with the experimental values with deviations
less than 4% (Table 1).
Table 1. Liquid Densities of Some Sulfonic-AcidFunctionalized ILs at 298 K from Simulation and
Experiment
density (g/cm3)
ILs

[PSPy][HSO4]
[PSPy][H2PO4]
[PSPy][Tfo]

simulation

experiment

1.45
1.46
1.42

1.50
1.48
1.44

The SFILs was synthesized by mixing euqimolar PSPy with H2SO4,


H3PO4, or HTfo, and the reaction mixtures were stirred at 353 K for
72 h. The high purity of SFILs was evaluated by NMR and MS. The
detailed experimental procedure is similar to our previous paper.10

3. RESULTS AND DISCUSSION


3.1. Bulk Liquid Structure and the Microenvironment
of Proton. Various sitesite radial distribution functions
(RDFs) were calculated to give us insight into the [PPy][Tfo]

Figure 2. Sitesite RDFs of [PPy][Tfo] and [PSPy][Tfo] between the oxygen atoms on [Tfo] and the heavy atoms on the cation.
20381

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

Article

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

The Journal of Physical Chemistry C

studied SFILs, could facilitate the proton transfer and thus


increase the proton conductivity in the electrochemical
applications.
In general, the above results show that the type of anion
plays a critical role in determining the interaction strength
between the sulfonic acid protons and anions in SFILs, which
provides us insight into tuning the proton microenvironment
and the performance of SFILs as novel proton conductive
materials and catalyst.
3.2. Nanostructured Aggregation Features. Nanostructured aggregation is one of the most appealing features
of ILs,42,54 which generally leads to the nanosegregation of
polar and nonpolar domains and accounts for the good
solubility of a variety of polar or nonpolar molecules in ILs as
well as the IL-mediated construction of self-assembled
nanochannels.20,22,55,56 To investigate the nanostructured
aggregation in SFILs, we calculated the intermolecular atom
atom RDFs between the equivalent terminal atoms along the
alkyl chain, as shown in Figure 5.

Figure 3. Colored SDFs for the anions around the cations with the
isovalue 3.5 (a) [PPy][Tfo] (b) [PSPy][Tfo], where subscripts v and
h stand for the vertical and horizontal views on the pyridine ring plane.
(Red isosurfaces represent the anions [Tfo], and small cyan, white,
blue, yellow, and red spheres represent carbon, hydrogen, nitrogen,
sulfur, and oxygen atoms, respectively.)

Figure 5. Intermolecular RDFs of equivalent terminal atoms of side


chain of SFILs and [PPy][Tfo].

The rst peak located at 3.8 is present for the terminal


alkyl atoms (C9) in nonfunctionalized [PPy][Tfo], indicating
the aggregation of the alkyl tails. There are also three peaks
located at about 5 for the sulfur atoms (S1) of the sulfonic
acid group in [PSPy][Tfo], [PSPy][HSO4], and [PSPy][H2PO4], which illustrates that the sulfonic acid group attached
to the terminal position of the alkyl chain could also form selfassembled nanostructures, where the dissociated protons could
hop among the adjacent cations and anions.20,24,25,57 The S1
S1 peak values follow the order [PSPy][Tfo] > [PSPy][HSO4]
> [PSPy][H2PO4], showing that the sulfonic acid group of
[PSPy][Tfo] possesses slightly stronger nanoaggregation than
the other two SFILs, and this result might be mainly attributed
to the weaker electrostatic interaction between sulfonic acid
group and anion, which could improve the mobility of the
terminal sulfonic acid group and thus strengthen its selfassembly behavior. This characteristic of SFILs has been
utilized in constructing regular molecular proton and water
pathways via the appropriate arrangement of the hydrophobic
and hydrophilic part of the molecular structure.20,22 In addition,
some intermolecular atomatom RDFs (as shown in Figure
S1) between atoms of anions have been computed, showing us
that the hydrogen-bonding network could also exist between
anions, which might facilitate the formation of anionic
nanoaggregation.

Figure 4. Sitesite RDFs of [PSPy][Tfo], [PSPy][HSO4] and


[PSPy][H2PO4] between the sulfonic acid proton and the oxygen
atom on the anions.

respectively, indicating that the main oxygen atom of the anions


could interact with the sulfonic acid proton (H13) of the
cations through a hydrogen-bonding interaction, following the
order [PSPy][Tfo] < [PSPy][HSO4] < [PSPy][H2PO4]. This
order is considered a result of the dierent acidities of the
conjugate acids of each anion. Among HTfo, H2SO4 and
H3PO4, the super-acidic HTfo (pKa = 14.0, 298.15 K) has a
much stronger acidity than H2SO4 (pKa = 3.0, 298.15 K) and
H3PO4 (pKa = 2.1, 298.15 K); thus, the [Tfo] anion has a
much lower basicity than [HSO4] and [H2PO4], and the
attraction of the anion to the sulfonic acid proton in
[PSPy][Tfo] is the weakest among these three SFILs.
An ecient proton transfer in an IL matrix contributes to the
eective proton conductivity of IL-based electrolytes.53 Therefore, the most unconstrained microenvironment around the
sulfonic acid proton in [PSPy][Tfo], resulting from the weakest
interaction between [Tfo] and the proton among the three
20382

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

Article

The Journal of Physical Chemistry C

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

Figure 6. Mass density proles of cations and anions of ILs at 423 K.

Figure 7. Coulombic charge density proles along the z-direction. (a) Entire liquid slab for [PPy][Tfo]. (b) Zoomed-in on the top interface for
[PPy][Tfo]. (c) Entire liquid slab for [PSPy][Tfo]. (d) Zoomed-in on the top interface for [PSPy][Tfo].

3.3. Mass Density Proles for VacuumIL Interface.


The simulation of the gasliquid interfacial structure of SFILs
was approximately performed by simulating the vacuumliquid
interfacial structure. The mass density proles (MDPs) of
cations and anions were calculated on the basis of the location
of ionic centers of mass with respect to their average bulk
density along the z-direction.36 In Figure 6, the vacuum side is
on the right of the Z-axis, where the mass densities of the
cations and anions are zero, and the bulk liquid is on the left
with the interfacial region between it and the vacuum. As Figure
6 shows, clear density oscillations can be found for both cations
and anions in the interfacial region and propagate into the bulk
interior, indicating the existence of nanostructured organization
in ILs as reported in previous reports.54,58 For three [Tfo]based ILs, the [Tfo] anions rather than the cations appear in
the outer position of the interface and closer to the vacuum
side. In addition, the MDPs of [PSPy][Tfo] in Figure 6 show
enhanced densities for both cation and anion in the interfacial
region compared to that of [PPy][Tfo], indicating the

introduction of sulfonic acid group has a signicant inuence


on the interfacial density oscillations. By comparing the mass
density prole of three SFILs bearing the same cation
[PSPy]+at 423 K ([PSPy][HSO4] and [PSPy][H2PO4] are
shown in Figure S3), a notable eect of the variety of anion on
the interfacial structure of SFILs was observed, whereas the
irregular density oscillation of the anions in [PSPy][HSO4] and
[PSPy][H2PO4] may result from the presence of a hydrogenbonding network among the anions, as we discussed before.
Overall, the mass density distribution of the ions presents a
general depiction of the ordering preferences and density
enrichment of cations and anions in the interfacial region, and
the following number density distribution analysis in section 3.5
provides detailed interfacial structure of SFILs at atomic level.
3.4. Coulombic Charge Density Proles. Figure 7 shows
the Coulombic charge density proles of [PPy][Tfo] and
[PSPy][Tfo] along the z direction. The Coulombic charge
density proles of [PSPy][HSO4], [PSPy][H2PO4], and
[HSPy][Tfo] are listed in the Supporting Information. The
20383

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

Article

The Journal of Physical Chemistry C

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

Figure 8. Number density proles of the backbone atoms of [PPy][Tfo], [PSPy][Tfo], and [HSPy][Tfo] at 423 K (dashed lines correspond to the
atoms of anions at the interface).

Figure 9. Number density proles of the backbone atoms of [PSPy][HSO4] and [PSPy][H2PO4] at 423 K (dashed lines correspond to the atoms of
anions at the interface).

NDPs of three ILs ([PPy][Tfo], [PSPy][Tfo] and [HSPy][Tfo]) are presented in Figure 8. As observed in Figure 8a, the
terminal carbon atoms (C9) of the propyl groups of
[PPy][Tfo] seem to locate in the outmost position of the
interfacial region, successively followed by C8 atoms and C7
atoms, which is in good agreement with the simulation and
experimental results in the previous report that the alkyl chain
was inclined to be perpendicular to the surface and pointing
into the vacuum.59,60 As for the distribution of the anion
[Tfo], represented by the C3 and S2 atoms of the anion, tends
to be aligned with the surface normal and the anionic CF3
groups are exposed at the surface. However, as shown in Figure
8b, when adding sulfonic acid groups into the terminal methyl
group of the propyl group, the ordering preference of the alkyl
tails of the [PSPy]+ cation is almost the opposite of that
observed in [PPy]+, with the order of C7atoms, C8 atoms, C9
atoms, and S1 atoms from the vacuum side into the bulk liquid.
Moreover, it can be inferred, from the position of the terminal
carbon atom (C6) of pyridine ring in Figure 8a and 8b, that the
pyridine ring of the [PSPy]+ cation prefers to locate in the
outer region of the surface, whereas the pyridine ring in [PPy]+
tends to stay in the inner part. The sharp contrast in ordering
preference of the alkyl tails and pyridine rings illustrates the
reverse ordering of the cations of these two ILs at the interface,
presumably due to the addition of the sulfonic acid group that
could enhance the polarity of the side chain of cation, indicating
that the tail-tuning strategy has a signicant impact on the
microstructure of the interface. Moreover, the interfacial
structure of [HSPy][Tfo] shown in Figure 8c is analogous to

charge density proles were calculated by counting the


electrostatic partial charge on each atom averaged over each
z-slab perpendicular to the charged interface. As shown in
Figure 7a,c, at the interface, a maximum value of charge density
can be found for both the cation and anion in these two ILs,
which is in agreement with the mass density peak observed in
Figure 5. The total charge density prole of the ion pair
indicates that it is essentially neutral along the entire liquid slab.
As shown in Figure 7b, the charge density proles of
[PPy][Tfo] (green line) shows a small negative charge peak
on the extreme outer edge of the interface, followed by a net
positive charge domain between 8.25 and 8.60 nm, which is
similar to the charge density distribution of the common IL
[bmim][BF4].36 However, in [PSPy][Tfo], the charge variation
on the top outermost layer of interface is slight, followed by an
obvious negative charge peak (green line in Figure 7d), which
suggests that the addition of sulfonic acid functional group into
ILs change the charge distribution of interface on some level.
3.5. Number Density Proles. The number density
proles (NDPs) for selected atoms with respect to the z-axis
were constructed by dividing the z-space into 600 bins, with
each bin entry weighted by the atomic masses and then divided
by the product of the atomic mass and the stoichiometric
coecient of the atom or atoms in the bin, so the number
density values can be obtained normalized on a per
stoichiometric coecient basis.51 The distribution of the
sulfonic acid group was represented using the NDPs of the
sulfur atoms (S1) in the sulfonic acid group, and the bulk liquid
is on the left side and the vacuum is on the right side. The
20384

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

Article

The Journal of Physical Chemistry C

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

Figure 10. Snapshots of the simulation boxes containing 512 ions, using a color code (the anions and pyridinium rings were colored in red and the
side chain in green) to investigate the interfacial structure (a) [PPy][Tfo] and (b) [PSPy][Tfo].

Figure 11. Orientational ordering parameters of ILs (a) [PPy][Tfo], (b) [PSPy][Tfo], and (c) [HSPy][Tfo] along the Z-axis.

anions and pyridinium rings in red and the side chain in green.
The comparison between the bulk phase and the interface
phase in each picture shows that the side chain of [PPy][Tfo]
shown in Figure 10a exhibits a trend pointing toward the
vacuum side, whereas the side chain of [PSPy][Tfo] shown in
Figure 10b was covered by the anions and pyridinium rings.
This observation is consistent with that drawn from analyzing
the number density ordering of the atoms shown in Figure 8a,b.
3.6. Orientational Ordering Parameters. Figure 11 and
12 illustrate the distribution of the orientational ordering of the

that of [PSPy][Tfo]. Nevertheless, the carbon atoms in the


alkyl chain of the [HSPy]+ cation are found to stay in the same
region of the surface with high probability and thus exhibits a
more compact interfacial structure than that of [PSPy][Tfo],
due to the exibility of the long aliphatic chain in [HSPy]+.
Comparing Figure 9 with Figure 8b, the eect of the dierent
anions on the SFIL interfacial structure can be partially
illustrated. The pyridinium rings of the cations of these three
SFILs ([PSPy][Tfo], [PSPy][HSO4], [PSPy][H2PO4]) are all
apt to locate in the outer position of the surface, whereas their
side chains, including the alkyl chains and sulfonic acid groups,
tend to stay in the inner part of the interface. Nevertheless, the
side-chain ordering (C7, C8, C9, S1) of the cations in these
three SFILs shows no uniform preference at the interface and
depends on the types of anions, most likely due to the
dierences in electrostatic and dispersion interactions between
the side chain and the anions. Moreover, the location of the
anions in these three SFILs, compared with the same carbon
atoms (C6) of the pyridinium rings, shows signicant contrast,
suggesting some information where the hydrophilic anion
([HSO4], [H2PO4]) locates in the inner layer of the
interface, while the slightly hydrophobic anion ([Tfo]) locates
in the outer layer of the interface with the hydrophobic part
(CF3) pointing toward the vacuum side, as we observed
before. The weaker interaction between the less basic anion
[Tfo] and the cations sulfonic acid group in [PSPy][Tfo]
than that in [PSPy][HSO4] and [PSPy][H2PO4] may also
account for the interfacial dierence, which implies that the
enhancement in the polarity and basicity of the anion could
lead to increased probability for the anions to occupy the inner
part of the interface.
To gain a visual description of the interfacial structure of the
studied ILs, snapshots of the simulation box for [PPy][Tfo]
and [PSPy][Tfo] from the MD trajectory data are shown in
Figure 10. A color code has been adopted to represent the

Figure 12. Orientational ordering parameters of ILs (a) [PSPy][HSO4], and (b) [PSPy][H2PO4] along the Z-axis.

side chain and pyridine rings, which can be studied using the
following correlation function:
1
P2( ) = < [3cos2( ) 1]>
2
where is the angle between a specic vector in the moleculexed structure and the z-axis parallel to the surface normal. The
z-direction space was divided into 4050 bins, and each vector
event was sorted into the correct bin. The angle between the
vector and z-aixs was calculated for each molecule in the
system, then giving second Legendre polynomial P2() for each
20385

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

The Journal of Physical Chemistry C

specic vector. The average of second Legendre polynomial


P2() can be obtained by averaging the summation of P2()
values for each bin, which is useful to attain a quantitative
description of molecular orientation within the liquid interfacial
region. It ranges from 0.5 to 1, with a value of 0.5
representing two specic perpendicular vectors and a value of 1
implying that they are parallel. As shown in Figure 11a and
Figure 11b, the peak values of the C7C9 vector and the C7
S1 vector in [PPy][Tfo] and [PSPy][Tfo] at the surface are
both in the range of 0.2 to 0.30. This result suggests that the
side chains of these two cations both prefer to align with the
surface normal, which is in agreement with the number density
ordering shown in Figure 8a and 8b. From the negative value of
the C7S1 vector in Figure 11c, a distinct orientational
distribution with the long side chain of [HSPy]+ parallel to the
interface plane can be observed, due to its long aliphatic chain.
Moreover, it can be inferred from the negative value of the
N3C6 vector in Figure 11b and 11c that the pyridine ring
plane of [PSPy][Tfo] and [HSPy][Tfo] tends to lie at at the
surface, similar to the previous simulation results for
imidazolium-based ionic liquids.61 When tethering [HSO4]
and [H2PO4] to the same cation [PSPy]+, the side chains in
both ILs prefer to adopt an orientation relatively parallel to the
surface plane, deduced from the negative value of the C7S1
vector as shown in Figure 12. Moreover, the second Legendre
Polynomial vanishes in the bulk region for all ILs as expected
due to the isotropic orientation in that region.

Article

ASSOCIATED CONTENT

S Supporting Information
*

The Supporting Information is available free of charge on the


ACS Publications website at DOI: 10.1021/acs.jpcc.5b02814.
Atom charges for all ionic liquids and density proles of
ILs (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: xinghb@zju.edu.com.
Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
The research was supported by the National Natural Science
Foundation of China (21222601, 91434115, and 21436010),
the Zhejiang Provincial Natural Science Foundation of China
(LR13B060001), the Fundamental Research Funds for the
Central Universities of China (2014XZZX003-17), and the
Program for New Century Excellent Talents in University of
China (NCET-13-0524).

REFERENCES

(1) Susan, M. A. B. H.; Noda, A.; Mitsushima, S.; Watanabe, M.


Brnsted Acid-Base Ionic Liquids and Their Use as New Materials for
Anhydrous Proton Conductors. Chem. Commun. 2003, 8, 938939.
(2) Greaves, T. L.; Drummond, C. J. Protic Ionic Liquids: Properties
and Applications. Chem. Rev. 2008, 108, 206237.
(3) Mishra, A. K.; Kuila, T.; Kim, D.; Kim, N. H.; Lee, J. H. Protic
Ionic Liquid-Functionalized Mesoporous Silica-Based Hybrid Membranes For Proton Exchange Membrane Fuel Cells. J. Mater. Chem.
2012, 22, 2436624372.
(4) Cole, A. C.; Jensen, J. L.; Ntai, I.; Tran, K. L. T.; Weaver, K. J.;
Forbes, D. C.; Davis, J. H. Novel Brnsted Acidic Ionic Liquids and
Their Use as Dual Solvent-Catalysts. J. Am. Chem. Soc. 2002, 124,
59625963.
(5) Kim, S. Y.; Kim, S.; Park, M. J. Enhanced Proton Transport in
Nanostructured Polymer Electrolyte/Ionic Liquid Membranes under
Water-Free Conditions. Nat. Commun. 2010, 1, 17.
(6) Huang, J.; Luo, H.; Liang, C.; Sun, I. W.; Baker, G. A.; Dai, S.
Hydrophobic Brnsted Acid-Base Ionic Liquids Based on PAMAM
Dendrimers with High Proton Conductivity and Blue Photoluminescence. J. Am. Chem. Soc. 2005, 127, 1278412785.
(7) Xing, H.; Liao, C.; Yang, Q.; Veith, G. M.; Guo, B.; Sun, X.; Ren,
Q.; Hu, Y.; Dai, S. Ambient Lithium-SO2 Batteries with Ionic Liquids
as Electrolytes. Angew. Chem., Int. Ed. 2014, 53, 20992103.
(8) Chen, L.; Mullen, G. E.; Le Roch, M.; Cassity, C. G.; Gouault, N.;
Fadamiro, H. Y.; Barletta, R. E.; OBrien, R. A.; Sykora, R. E.; Stenson,
A. C.; et al. On the Formation of a Protic Ionic Liquid in Nature.
Angew. Chem. 2014, 126, 1195611959.
(9) Wang, C.; Luo, H.; Jiang, D. E.; Li, H.; Dai, S. Carbon Dioxide
Capture by Superbase-Derived Protic Ionic Liquids. Angew. Chem.
2010, 122, 61146117.
(10) Xing, H.; Wang, T.; Zhou, Z.; Dai, Y. Novel Brnsted-Acidic
Ionic Liquids for Esterifications. Ind. Eng. Chem. Res. 2005, 44, 4147
4150.
(11) Kore, R.; Srivastava, R. Synthesis and Applications of Novel
Imidazole and Benzimidazole Based Sulfonic Acid Group Functionalized Bronsted Acidic Ionic Liquid Catalysts. J. Mol. Catal. A: Chem.
2011, 345, 117126.
(12) Amarasekara, A. S.; Owereh, O. S. Synthesis of a Sulfonic Acid
Functionalized Acidic Ionic Liquid Modified Silica Catalyst and
Applications in the Hydrolysis of Cellulose. Catal. Commun. 2010, 11,
10721075.

4. CONCLUSION
In this work, we have investigated the microscopic structural
features of some representative SFILs by performing MD
simulation using the ion pair charge approach.
When the sulfonic acid group is tethered to the terminal
position of the alkyl chain, a strong interaction between the
sulfonic acid group and the anions was found. The type of
anions plays a critical role in determining the intensity of the
interaction between the sulfonic acid proton and the anions in
the order [PSPy][Tfo] < [PSPy][HSO4] < [PSPy][H2PO4],
which has a pronounced eect on the proton microenvironments of the SFILs. Less basic anions can lead to a much
weaker association of sulfonic acid protons with anions due to
the reduced protonanion interactions. In addition, the
terminal sulfonic acid group (tail functional group) could
lead to a nanostructured aggregation in the bulk phase similar
to the nonfunctionalized alkyl chain does.
The analysis of various density proles reveals that, for
dierent SFILs, the side chains of the cations all tend to locate
in the inner part of the interface, with the pyridinium rings of
the cations exposed in the outer region at the surface, which is
opposite to the case of nonfunctionalized IL. The enhancement
in the polarity and basicity of the anion could lead to increased
probability for the anions to locate in the inner part of the
interface. The alkyl chain of the [HSPy]+ cation exhibits a more
compact interfacial structure than [PSPy]+ due to the exibility
of its long aliphatic chain.
The present results show some structural characteristics of
SFILs in both the bulk liquid and interfacial regions, and the
ndings oer some insight into the proton microenvironments
of SFILs, which is highly instructive for the development of
new SFILs and proton conductive materials.
20386

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

Article

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

The Journal of Physical Chemistry C

Neuendorf, S.; GA Ke, V. Ionic Liquids for Propene-Propane


Separation. Chem. Eng. Technol. 2010, 33, 6373.
(32) Xu, H.; Han, Z.; Zhang, D.; Zhan, J. Interface Behaviors of
Acetylene and Ethylene Molecules with 1-Butyl-3-methylimidazolium
Acetate Ionic Liquid: A Combined Quantum Chemistry Calculation
and Molecular Dynamics Simulation Study. ACS Appl. Mater. Interfaces
2012, 4, 66466653.
(33) Lynden-Bell, R. M.; Del Popolo, M. G.; Youngs, T. G.;
Kohanoff, J.; Hanke, C. G.; Harper, J. B.; Pinilla, C. C. Simulations of
Ionic Liquids, Solutions, and Surfaces. Acc. Chem. Res. 2007, 40, 1138
1145.
(34) Hallett, J. P.; Welton, T. Room-Temperature Ionic Liquids:
Solvents for Synthesis and Catalysis. 2. Chem. Rev. 2011, 111, 3508
3576.
(35) Kolbeck, C.; Niedermaier, I.; Deyko, A.; Lovelock, K. R. J.;
Taccardi, N.; Wei, W.; Wasserscheid, P.; Maier, F.; Steinruck, H.
Influence of Substituents and Functional Groups on the Surface
Composition of Ionic Liquids. Chem. - Eur. J. 2014, 20, 39543965.
(36) Xing, H.; Yan, Y.; Yang, Q.; Bao, Z.; Su, B.; Yang, Y.; Ren, Q.
Effect of Tethering Strategies on the Surface Structure of AmineFunctionalized Ionic Liquids: Inspiration on the CO2 Capture. J. Phys.
Chem. C 2013, 117, 1601216021.
(37) Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D.
A. Development and Testing of A General Amber Force Field. J.
Comput. Chem. 2004, 25, 11571174.
(38) Canongia Lopes, J. N.; Padua, A. A. H.; Shimizu, K. Molecular
Force Field for Ionic Liquids IV: Trialkylimidazolium and Alkoxycarbonyl-Imidazolium Cations; Alkylsulfonate and Alkylsulfate
Anions. J. Phys. Chem. B 2008, 112, 50395046.
(39) Canongia Lopes, J. N.; Padua, A. A. H. Molecular Force Field
for Ionic Liquids III: Imidazolium, Pyridinium, and Phosphonium
Cations; Chloride, Bromide, and Dicyanamide Anions. J. Phys. Chem. B
2006, 110, 1958619592.
(40) Canongia Lopes, J. N.; Padua, A. A. H. Molecular Force Field
for Ionic Liquids Composed of Triflate or Bistriflylimide Anions. J.
Phys. Chem. B 2004, 108, 1689316898.
(41) Morrow, T. I.; Maginn, E. J. Molecular Dynamics Study of the
Ionic Liquid 1-n-Butyl-3-methylimidazolium Hexafluorophosphate. J.
Phys. Chem. B 2002, 106, 1280712813.
(42) Cadena, C.; Zhao, Q.; Snurr, R. Q.; Maginn, E. J. Molecular
Modeling and Experimental Studies of the Thermodynamic and
Transport Properties of Pyridinium-Based Ionic Liquids. J. Phys. Chem.
B 2006, 110, 28212832.
(43) Verevkin, S. P.; Zaitsau, D. H.; Emel'yanenko, V. N.;
Yermalayeu, A. V.; Schick, C.; Liu, H.; Maginn, E. J.; Bulut, S.;
Krossing, I.; Kalb, R. Making Sense of Enthalpy of Vaporization
Trends for Ionic Liquids: New Experimental and Simulation Data
Show a Simple Linear Relationship and Help Reconcile Previous Data.
J. Phys. Chem. B 2013, 117, 64736486.
(44) Paredes, X.; Fernandez, J.; Padua, A. A. H.; Malfreyt, P.;
Malberg, F.; Kirchner, B.; Pensado, A. S. Using Molecular Simulation
to Understand the Structure of [C2C1im]+-Alkylsulfate Ionic Liquids:
Bulk and Liquid-Vapor Interfaces. J. Phys. Chem. B 2012, 116, 14159
14170.
(45) Sun, H.; Qiao, B.; Zhang, D.; Liu, C. Structure of 1Butylpyridinium Tetrafluoroborate Ionic Liquid: Quantum Chemistry
and Molecular Dynamic Simulation Studies. J. Phys. Chem. A 2010,
114, 39903996.
(46) Koddermann, T.; Paschek, D.; Ludwig, R. Molecular Dynamic
Simulations of Ionic Liquids: A Reliable Description of Structure,
Thermodynamics and Dynamics. ChemPhysChem 2007, 8, 2464
2470.
(47) Zhang, Y.; Maginn, E. J. A Simple AIMD Approach to Derive
Atomic Charges for Condensed Phase Simulation of Ionic Liquids. J.
Phys. Chem. B 2012, 116, 1003610048.
(48) Bayly, C. I.; Cieplak, P.; Cornell, W.; Kollman, P. A. A WellBehaved Electrostatic Potential Based Method using Charge Restraints
for Deriving Atomic Charges: the RESP Model. J. Phys. Chem. 1993,
97, 1026910280.

(13) Li, X.; Eli, W. A Green Approach for the Synthesis of Long
Chain Aliphatic Acid Esters at Room Temperature. J. Mol. Catal. A:
Chem. 2008, 279, 159164.
(14) Fang, D.; Zhou, X.; Ye, Z.; Liu, Z. Brnsted Acidic Ionic Liquids
and Their Use as Dual Solvent-Catalysts for Fischer Esterifications.
Ind. Eng. Chem. Res. 2006, 45, 79827984.
(15) Ren, H.; Ying, H.; Sun, Y.; Wu, D.; Ma, Y.; Wei, X. Synthesis of
Poly(lactic acid)-Poly(ethylene glycol) Copolymers using MultiSO3H-Functionalized Ionic Liquid as the Efficient and Reusable
Catalyst. Polym. Bull. 2014, 71, 11731195.
(16) Zhang, S.; Lefebvre, H.; Tessier, M.; Fradet, A. Influence of
Brnsted Acid Ionic Liquid Structure on Hydroxyacid Polyesterification. Green Chem. 2011, 13, 27862793.
(17) Zhang, L.; Cui, Y.; Zhang, C.; Wang, L.; Wan, H.; Guan, G.
Biodiesel Production by Esterification of Oleic Acid over Brnsted
Acidic Ionic Liquid Supported onto Fe-Incorporated SBA-15. Ind. Eng.
Chem. Res. 2012, 51, 1659016596.
(18) Yanfei, H.; Xiaoxiang, H.; Qing, C.; Lingxiao, Z. Transesterification of Soybean Oil to Biodiesel by Brnsted-Type Ionic
Liquid Acid Catalysts. Chem. Eng. Technol. 2013, 36, 15591567.
(19) Zhang, C.; Fu, Z.; Dai, B.; Zen, S.; Liu, Y.; Xu, Q.; Kirk, S. R.;
Yin, D. Biochar Sulfonic Acid Immobilized Chlorozincate Ionic Liquid:
An Efficiently Biomimetic and Reusable Catalyst for Hydrolysis of
Cellulose and Bamboo under Microwave Irradiation. Cellulose 2014,
21, 12271237.
(20) Soberats, B.; Yoshio, M.; Ichikawa, T.; Taguchi, S.; Ohno, H.;
Kato, T. 3D Anhydrous Proton-Transporting Nanochannels Formed
by Self-Assembly of Liquid Crystals Composed of a Sulfobetaine and a
Sulfonic Acid. J. Am. Chem. Soc. 2013, 135, 1528615289.
(21) Rondla, R.; Lin, J. C. Y.; Yang, C. T.; Lin, I. J. B. Strong
Tendency of Homeotropic Alignment and Anisotropic Lithium Ion
Conductivity of Sulfonate Functionalized Zwitterionic Imidazolium
Ionic Liquid Crystals. Langmuir 2013, 29, 1177911785.
(22) Ichikawa, T.; Kato, T.; Ohno, H. 3D Continuous Water
Nanosheet as a Gyroid Minimal Surface Formed by Bicontinuous
Cubic Liquid-Crystalline Zwitterions. J. Am. Chem. Soc. 2012, 134,
1135411357.
(23) Noda, A.; Susan, M. A. B. H.; Kudo, K.; Mitsushima, S.;
Hayamizu, K.; Watanabe, M. Brnsted Acid-Base Ionic Liquids as
Proton-Conducting Nonaqueous Electrolytes. J. Phys. Chem. B 2003,
107, 40244033.
(24) Perttu, E. K.; Szoka, F. C. Zwitterionic Sulfobetaine Lipids that
Form Vesicles with Salt-Dependent Thermotropic Properties. Chem.
Commun. 2011, 47, 1261312615.
(25) Yoshizawa, M.; Ohno, H. Anhydrous Proton Transport System
Based on Zwitterionic Liquid and HTFSI. Chem. Commun. 2004,
18281829.
(26) Wu, J.; Jiang, T.; Jiang, D.; Jin, Z.; Henderson, D. A Classical
Density Functional Theory for Interfacial Layering of Ionic Liquids.
Soft Matter 2011, 7, 1122211231.
(27) Bayley, P. M.; Best, A. S.; MacFarlane, D. R.; Forsyth, M.
Transport Properties and Phase Behaviour in Binary and Ternary Ionic
Liquid Electrolyte Systems of Interest in Lithium Batteries.
ChemPhysChem 2011, 12, 823827.
(28) Zhang, X.; Huo, F.; Liu, Z.; Wang, W.; Shi, W.; Maginn, E. J.
Absorption of CO2 in the Ionic Liquid 1-n-Hexyl-3-methylimidazolium Tris(pentafluoroethyl)trifluorophosphate ([hmim][FEP]): A
Molecular View by Computer Simulations. J. Phys. Chem. B 2009,
113, 75917598.
(29) Fedorov, M. V.; Lynden-Bell, R. M. Probing the Neutral
Grapheme-Ionic Liquid Interface: Insights from Molecular Dynamics
Simulations. Phys. Chem. Chem. Phys. 2012, 14, 25522556.
(30) Li, R.; Xing, H.; Yang, Q.; Zhao, X.; Su, B.; Bao, Z.; Yang, Y.;
Ren, Q. Selective Extraction of 1-Hexene Against n-Hexane in Ionic
Liquids with or without Silver Salt. Ind. Eng. Chem. Res. 2012, 51,
85888597.
(31) Mokrushin, V.; Assenbaum, D.; Paape, N.; Gerhard, D.;
Mokrushina, L.; Wasserscheid, P.; Arlt, W.; Kistenmacher, H.;
20387

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

Article

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): August 20, 2015 | doi: 10.1021/acs.jpcc.5b02814

The Journal of Physical Chemistry C


(49) Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A.
E.; Berendsen, H. J. C. GROMACS: Fast, Flexible, and Free. J.
Comput. Chem. 2005, 26, 17011718.
(50) Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. GROMACS
4: Algorithms for Highly Efficient, Load-Balanced, and Scalable
Molecular Simulation. J. Chem. Theory Comput. 2008, 4, 435447.
(51) Perez-Blanco, M. E.; Maginn, E. J. Molecular Dynamics
Simulations of CO2 at an Ionic Liquid interface: Adsorption,
Ordering, and Interfacial Crossing. J. Phys. Chem. B 2010, 114,
1182737.
(52) Xing, H.; Wang, T.; Zhou, Z.; Dai, Y. The Sulfonic AcidFunctionalized Ionic Liquids with Pyridinium Cations: Acidities and
Their Acidity-Catalytic Activity Relationships. J. Mol. Catal. A: Chem.
2007, 264, 5359.
(53) Kumar, M.; Venkatnathan, A. Mechanism of Proton Transport
in Ionic-Liquid-Doped Perfluorosulfonic Acid Membranes. J. Phys.
Chem. B 2013, 117, 1444914456.
(54) Canongia Lopes, J. N. A.; Padua, A. A. H. Nanostructural
Organization in Ionic Liquids. J. Phys. Chem. B 2006, 110, 33303335.
(55) Gao, X.; Lu, F.; Shi, L.; Jia, H.; Gao, H.; Zheng, L.
Nanostructured Aqueous Lithium-Ion Conductors Formed by the
Self-Assembly of Imidazolium-Type Zwitterions. ACS Appl. Mater.
Interfaces 2013, 5, 1331213317.
(56) Kim, O.; Jo, G.; Park, Y. J.; Kim, S.; Park, M. J. Ion Transport
Properties of Self-Assembled Polymer Electrolytes: The Role of
Confinement and Interface. J. Phys. Chem. Lett. 2013, 4, 21112117.
(57) Ichikawa, T.; Kato, T.; Ohno, H. 3D Continuous Water
Nanosheet as a Gyroid Minimal Surface Formed by Bicontinuous
Cubic Liquid-Crystalline Zwitterions. J. Am. Chem. Soc. 2012, 134,
1135411357.
(58) Wang, Y.; Voth, G. A. Unique Spatial Heterogeneity in Ionic
Liquids. J. Am. Chem. Soc. 2005, 127, 1219212193.
(59) Baldelli, S. Influence of Water on the Orientation of Cations at
the Surface of a Room-Temperature Ionic Liquid: A Sum Frequency
Generation Vibrational Spectroscopic Study. J. Phys. Chem. B 2003,
107, 61486152.
(60) Aliaga, C.; Baker, G. A.; Baldelli, S. Sum Frequency Generation
Studies of Ammonium and Pyrrolidinium Ionic Liquids Based on the
Bis-trifluoromethanesulfonimide Anion. J. Phys. Chem. B 2008, 112,
16761684.
(61) Bhargava, B. L.; Balasubramanian, S. Layering at an Ionic
Liquid-Vapor Interface: A Molecular Dynamics Simulation Study of
[bmim][PF6]. J. Am. Chem. Soc. 2006, 128, 1007310078.

20388

DOI: 10.1021/acs.jpcc.5b02814
J. Phys. Chem. C 2015, 119, 2037920388

You might also like