How The Structure of Pyrrolidinium Ionic Liquids Is Susceptible To High Pressure

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Article

pubs.acs.org/JPCB

How the Structure of Pyrrolidinium Ionic Liquids Is Susceptible to


High Pressure
Shobha Sharma, Aditya Gupta, and Hemant K. Kashyap*
Department of Chemistry, Indian Institute of Technology Delhi, Hauz Khas, New Delhi 110016, India

ABSTRACT: The structural landscape of room-temperature ionic liquids


(RTILs) with longer cationic alkyl tail(s) exhibits polarity ordering (PO)
along with charge ordering (CO). In polarity ordering, which is also
Downloaded via TEVA PHARMACEUTICAL INDUSTRIES LTD on January 3, 2021 at 10:53:14 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

referred to as intermediate-range ordering, polar groups are separated by


segregated domains of apolar groups and vice versa. Charge ordering
resembles alternation of positive−negative charge groups. In this work,
how these two characteristic orderings respond to applied external
pressure has been investigated via molecular dynamics simulations. The
present study complements the recent experimental studies of Yoshimura
et al. (J. Phys. Chem. B 2015, 119, 8146−8153) and computational studies
of Russina et al. (Phys. Chem. Chem. Phys. 2015, 17, 29496−29500)
wherein the authors described in detail the effects of pressure on the structural and conformational changes in imidazolium based
ionic liquids. Our simulations predict that for 1-alkyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)amide, Pyrr1,n+/NTf2−
with n = 8 and 10, the PO and CO fade when the external pressure increases from ambient pressure to 10000 bar. We observe
that the apolar tail group as well as the polar group correlations are susceptible to the applied pressure. The decrease of polar−
polar and apolar−apolar correlations at higher pressure is accompanied by the enhancement in the polar−apolar correlations and
increased stability/probability of gauche conformations along the cationic tails.

1. INTRODUCTION The pioneering experimental studies of Yoshimura and co-


workers63,67 on the series of imidazolium ionic liquids have
Stimuli-responsive materials that are capable of making
revealed that the ionic liquids could go through structural as
conformational and/or physiochemical changes on exertion of
well as conformational changes when the pressure is increased
external stimuli have gained a wide range of applications in
up to GPa level. The authors also showed that the ionic liquids
industry, including biosensing, drug delivery, and separation.1 could possess multiple phases or render structural transitions at
In the recent past, room-temperature ionic liquids (RTILs) around 7 GPa.
have found a great deal of attention because of their unique Recently, Russina et al.70 have reported that, upon increasing
properties and diverse applications.2,3 RTIL properties can be hydrostatic pressure on 1-octyl-3-methylimidazolium tetrafluor-
fine-tuned by making minor changes in the chemical oborate (Omim+/BF4−), the correlations corresponding to
composition of their constituent ions. The RTIL response polar moieties bearing strong electrostatic interactions are
functions can also be altered by changing the chemical nature of unaffected by the pressure, but the conformations of the alkyl
the ions. In order to conceive RTILs as smart materials, it is tails are influenced by pressure. Very much similar to the work
vital to understand how this new class of liquids respond to of Margulis73 and Pádua64 on the RTILs with diether
external stimuli, such as electric field, temperature, and stress or containing cationic tails, the authors claimed that the
pressure. Structural properties of RTILs under ambient enhancement of the gauche conformations along the tail leads
conditions and at elevated temperatures have been thoroughly to curling of the cation tail toward the imidazolium ring, and
examined by experiments4−33 and computer simulation34−49 the overall effect is suppression of PO in the RTIL studied.
studies, and these works are nicely summarized and discussed According to recent studies, pyrrolidinium cation based
in the recent review articles.6,50−54 Many RTILs show RTILs show more influence of tail modification on their
anomalous properties when they are subject to electric field55 microscopic properties than the other classes of ionic liquids.74
and temperature.56 However, very little has been addressed Cytotoxic studies of ionic liquids have also revealed that these
about the pressure-responsiveness of RTILs.52,57−72 Shah and ionic liquids are less toxic than well-known piperidinium and
Maginn studied the structure and dynamics of imidazolium imidazolium ionic liquids,74 which makes them more environ-
ionic liquids containing significantly long alkyl chains and ment friendly. In the present study, the choice of the 1-alkyl-1-
concluded that the thermodynamics and dynamics of these methylpyrrolidinium bis(trifluoromethylsulfonyl)amide
RTILs could be influenced by pressure.57 Molecular dynamics
(MD) studies done by Zhao et al. on 1-butyl-3-methylimida- Received: February 2, 2016
zolium hexafluorophosphate RTIL indicated the deformation of Revised: March 7, 2016
alkyl chains at higher pressures.61 Published: March 9, 2016

© 2016 American Chemical Society 3206 DOI: 10.1021/acs.jpcb.6b01133


J. Phys. Chem. B 2016, 120, 3206−3214
The Journal of Physical Chemistry B Article

(Pyrr1,n+/NTf2−) ionic liquids with longer tail length stems Table 1. Simulated Equilibrium Densities (in g/cm3) of the
from the fact that these liquids show pronounced charge and RTILs at 295 K for Different Pressures
polarity orderings under ambient conditions. In these liquids,
pressure (in bar) Pyrr1,8+/NTf2− Pyrr1,10+/NTf2−
the constituent species (cation head, cation tail, and anion) are
structurally complex and can be conformationally excited upon 100 1.343 1.301
changing pressure. Several studies on the structure of this class 1000 1.377 1.335
of ionic liquids have already been pursued.10,38,47,48,73,75,76 5000 1.467 1.424
However, to the best of our knowledge, only limited studies 10000 1.534 1.488
have been reported about their response toward a range of
applied pressure. The Margulis and Castner groups have done
comprehensive studies on the structure of RTILs comprising 1- computed properties at 5000 bar have not been presented here
(cyclohexylmethyl)-1-methylpyrrolidinium, 1-(2-ethylhexyl)-1- for clarity. The X-ray scattering static structure function, S(q),
methylpyrrolidinium, and 1-alkyl-1-methylpyrrolidinium cati- and its ionic and subionic components were computed using
ons paired with bis(trifluoromethylsulfonyl)amide anion via X- the methodology proposed in the literature.44,48,75,76 The exact
ray scattering experiments and molecular dynamics simula- definitions of cation head and tail groups used in this study are
tions.47,48,73,75,76 The MD results were in excellent agreement
with the X-ray scattering experiments under ambient
conditions, which allowed the authors to unambiguously assign
the characteristic orderings that to the peaks in total S(q).
Recently, a nice work by Smith and co-workers77 on the
structure of dialkylpyrrolidinium-based RTILs showed that the
RTILs with longer chains could exhibit a structural transition
induced by two-dimensional thin film confinement. The
present work is focused on understanding the structural
responsiveness of 1-alkyl-1-methylpyrrolidinium bis-
(trifluoromethylsulfonyl)amide ionic liquids with increasing
pressure that is isotropically applied to the liquid. Attention is
given on the total and partial X-ray scattering structure
functions along with representative radial distribution functions Figure 1. Definitions of the cation (head), cation (tail), and CTS
(RDFs). The aim of the present work is to know the group in the cation tails. In the CTS group, only carbon atoms are
susceptibility of charge and polarity orderings to the pressure included.
increase. The present study also focuses on understanding the
possibility of interdigitation and/or curling of cationic tails
along with conformational changes in the anion at very high provided in Figure 1. In brief, the X-ray scattering static
pressure. structure function S(q) was computed using eq 1.
S(q) =
2. COMPUTATIONAL DETAILS
n n L /2 sin qr
All simulations were carried out in the isothermal−isobaric ρo ∑i = 1 ∑ j = 1 xixj fi (q)f j (q) ∫ 4πr 2[gij(r ) − 1] qr
W (r ) dr
0
ensemble using the GROMACS-4.6.5 package.78,79 A total of n n
[∑i = 1 xi fi (q)][∑ j = 1 xj f j (q)]
3000 ion pairs in a cubic box were used for all the systems
studied. Periodic boundary conditions and minimum image (1)
convention were applied. Nonpolarizable atomistic force-field In eq 1, gij(r) is the radial distribution function (RDF) for the
parameters from all-atom optimized potentials for liquid atoms of type i and j. The RDFs include both intra- and
simulations (OPLS-AA)80−83 in conjunction with those of intermolecular pairs. xi is the mole fraction of atoms of type i,
Lopes and Pádua84,85 (CL&P) were used to model the Pyrr1,n+/ and f i(q) is the X-ray atomic form factor for ith type atoms.89 ρo
NTf2− RTILs.48,75 The leapfrog algorithm with 1 fs time step = Natom/⟨V⟩ is the total number density, and L is the box
was used for integration of equations of motion. The simulation length. W(r) is a Lorch window function defined as W(r) =
boxes for each RTIL were thoroughly equilibrated by raising sin(2πr/L)/(2πr/L).90,91 The Lorch window function mini-
and lowering the temperatures, and scaling down the partial mizes the effects of finite truncation of r. For a moderately large
charges. Each system with full atomic charges was allowed to system size, the W(r) function has negligible effects on the
equilibrate for at least 15 ns at 295 K temperature and at physical meaning of the peaks in S(q). To appreciate the
different pressures (100, 1000, 5000, and 10000 bar) using a characteristic ordering at different length scales, we used ionic,
velocity-rescale thermostat and a Parrinello−Rahman86 baro- subionic, and polarity based partitioning of the total structure
stat, respectively. The last 3 ns of the trajectories was saved at function S(q). Other details of the methodology are described
every 100 fs for the computation of the RTIL’s properties. To in the literature.44,75,76 In brief, the total S(q) can be split into
get better statistical averages, the properties of the Pyrr1,8+ its cationic, Scat−cat(q), anionic, San−an(q), and cationic−anionic,
system were averaged over two independent trajectories. The Scat−an(q) + San−cat(q), subcomponents. The cationic sub-
cutoff radius for the short-range interactions was set to 1.2 nm. component can be further decomposed into its head−head,
Electrostatic interactions were evaluated using the Particle tail−tail, and head−tail subcomponents, S cat−cat (q) =
Mesh Ewald (PME)87,88 summation technique with an Scathead−cathead(q) + Scathead−cattail(q) + Scattail−cathead(q) + Scattail−cattail(q).
interpolation order of 6 and a Fourier grid spacing of 0.08 Similarly, the cation−anion term can be split into cation head−
nm. The equilibrium values of the bulk densities at 295 K for anion and cation tail−anion, [Scat−an(q) + San−cat(q)] =
different pressures are provided in Table 1. A few of the Scathead−an(q) + San−cathead(q) + Scattail−an(q) + San−cattail(q). We can
3207 DOI: 10.1021/acs.jpcb.6b01133
J. Phys. Chem. B 2016, 120, 3206−3214
The Journal of Physical Chemistry B Article

group the different terms in the above functions into polar, 3.2. Charge Ordering. In Figure 3, we have shown cation−
apolar, and cross polar−apolar contributions76 as SPolar−Polar(q) cation, anion−anion, and cation−anion partial X-ray SSFs for
= San−an(q) + Scathead−cathead(q) + Scathead−an(q) + San−cathead(q),
S Apolar−Apolar (q) = S cat tail −cat tail (q), and [S Polar−Apolar (q) +
SApolar−Polar(q)] = Scathead−cattail(q) + Scattail−cathead(q) + Scattail−an(q) +
San−cattail(q).75,76

3. RESULTS AND DISCUSSION


3.1. X-ray Scattering Structure Function, S(q). As can
be seen in Figure 2, the polarity ordering or polar−apolar

Figure 2. Pressure dependence of the simulated total structure


function, S(q), for Pyrr1,8+/NTf2− (upper panel) and Pyrr1,10+/NTf2−
(lower panel) RTILs at 295 K. Figure 3. Cation−cation (black line), anion−anion (red line), and
cation−anion (blue line) partial S(q)s of (a) Pyrr1,8+/NTf2− and (b)
Pyrr1,10+/NTf2− RTILs at different pressures.

alternation peaks that occur below q = 0.5 Å−1 and charge the Pyrr1,8+ and Pyrr1,10+ systems at different pressures. Peaks
ordering peaks that occur at around q = 0.7 Å−1 in the X-ray and antipeaks observed at around q = 0.7 Å−1 in these functions
scattering structure function (SSF) of the Pyrr1,8+ and Pyrr1,10+ mean that, at the corresponding real-space characteristic
systems gradually dim upon increasing pressure. The distances, DCO = 2π/qCO, the appreciation or depreciation in
appearance of the polarity peaks below q < 0.5 Å−1 and charge the probability of finding cations (or anions) with respect to
ordering peaks at 0.5 Å−1 < q < 1 Å−1 implies that polarity another cation (or anion) are in the same phase44,48,75,76 and
ordering occurs at relatively longer characteristic distances than with respect to anion (or cation) are off-phase.44,48,75,76 For
the charge ordering which occur at shorter length scales but both of the RTILs, the amplitudes of the cation−cation, anion−
persists for longer range. Note that these two peaks are also
anion, and cation−anion partial components gradually decrease
observed for 1 bar pressure and 295 K temperature (ambient
as pressure increases. This means that the charge ordering in
conditions) and have been reported previously by the Margulis
these systems does get diminished or weaker as the pressure
and Castner groups.48,75,76 Under ambient conditions, the
heights of both PO and CO peaks are more than that for the increases. Therefore, the decreased intensity of this peak in the
lowest pressure studied in this work. Therefore, considering the X-ray SSF at higher pressure is not due to mere cancellations of
previous works48,75,76 and the present study, one can infer that the peaks (constructive interference) and antipeaks (destructive
the two orderings fade when the external pressure increases interference). Also, corresponding real-space characteristic
from ambient pressure to 10000 bar. While the polarity distances, DCO = 2π/qCO, for higher pressure are larger than
ordering peak shows a slight shift toward higher q-values with those for the lower pressures. This clearly indicates that the
increasing pressure, the charge ordering peak shifts toward average cation (head)−cation (head), cation (head)−anion,
lower q-values upon increasing pressure. Also, the polar−apolar and anion−anion separations comprising CO are larger at the
alternation peak almost disappears at the highest pressure, higher pressures but the corresponding correlations are weaker.
whereas the charge ordering peak still exists even at very high The weakening of CO at higher pressures is also supported by
pressure. This clearly suggests that the polar−apolar domain the radial distribution functions (RDFs) for the cationic and
segregation in these liquids has enhanced sensitivity toward the anionic nitrogen atoms (Figure 4), namely, Ncat−Nan, Ncat−Ncat,
pressure than the charge ordering. To know the exact origin of and Nan−Nan. Clearly, the amplitudes of the peaks correspond-
this behavior of these two characteristic peaks (or orderings), ing to the nearest solvation shells in all of the RDFs decrease
we computed the ionic and subionic components of the X-ray when the pressure is increased from 100 to 10000 bar, with the
SSF proposed in the literature.44,48,75,76 Ncat−Nan RDF being the most sensitive among all three RDFs.
3208 DOI: 10.1021/acs.jpcb.6b01133
J. Phys. Chem. B 2016, 120, 3206−3214
The Journal of Physical Chemistry B Article

Figure 4. Ncat−Nan (black line), Ncat−Ncat (blue line), and Nan−Nan


Figure 5. Cation−cation (black line), cation (head)−cation (head)
(red line) radial distribution functions for (a) Pyrr1,8+/NTf2− and (b)
(red line), cation (tail)−cation (tail) (blue line), and cation (head)−
Pyrr1,10+/NTf2− RTILs at different pressures.
cation (tail) (green line) partial S(q)s of (a) Pyrr1,8+/NTf2− and (b)
Pyrr1,10+/NTf2− RTILs at different pressures.
The susceptibility of subcationic correlations toward pressure
can also be observed for the cation (head)−cation (head), to tail−tail correlations. This observation is complemented by
cation (tail)−cation (tail), and cation (head)−cation (tail) the RDF for the group of three consecutive terminal carbon
S(q)s (Figures 5). As already explained by others,48,75,76 cation atoms, defined as CTS (Figure 1), in the cationic tail at
heads that are separated by anions contribute to cation different pressures, as shown in Figure 8a. The decrease in the
(head)−cation (head) for CO ordering, whereas cation heads heights of the closest contact peaks, occurring below r = 0.6
that are separated by the segregated domains of apolar groups nm, in the CTS−CTS RDF as the pressure increases clearly
contribute to the cation (head)−cation (head) peak corre- shows weakening of apolar−apolar correlations. The opposite
sponding to PO ordering. It also turns out that the cation trend is observed in the case of cation nitrogen (Ncat) and CTS
(tail)−cation (tail) correlations have significant contribution to correlations with pressure change (see Figure 8b). While the
the PO peak along with cation (head)−cation (head), and both nearest neighbor Ncat−CTS correlations are enhanced with
correlations get weaker upon increasing pressure. To get more increasing pressure, the second solvation shell correlations are
accurate details of charge and polarity orderings, partial X-ray decreased as pressure increases. Moreover, upon increasing
SSFs for subcationic components (head and tail) with anion pressure, the distance of closest approach between the tails as
were also computed. From Figure 6, it is clearly visible that, well as Ncat and the CTS group decrease. The adjacency
upon increasing pressure, the amplitudes of both cation correlation (AC) peak45,48,75,76 at around q = 1.3 Å−1 in S(q),
(head)−anion and cation (tail)−anion correlations correspond- which is primarily rendered by close-contact distances between
ing to PO and CO peaks are suppressed. both cationic and anionic atoms, slightly shifts toward larger q-
3.3. Polarity Ordering. With regard to the polarity values along with decreased intensity with increasing pressure
ordering behavior toward pressure, in Figure 7, we have for both of the RTILs.
shown the polar−polar, polar−apolar, and apolar−apolar partial 3.4. Dihedral Distributions, Cation Tail Curling, and
X-ray SSFs for both RTILs. In all of the partial components, the Anion Conformers. To elucidate on the conformational
PO peaks at around q = 0.4 Å−1 slightly shift toward larger q- aspect of the cation alkyl tails, we computed the probability
values (smaller distances) as the pressure is increased. Here the distribution function for the number of trans dihedrals along
polar component is comprised of the cation head and anion the alkyl chain of the cations (Figure 9). In agreement with the
and the apolar groups belong to the cation tail (see Figure previous study,70 the overall probability of finding a greater
1).44,48,75,76 The subcomponents of polar−polar, apolar−apolar, number of gauche kinks/defects along the cationic tail is
and polar−apolar X-ray SSFs are provided in Figures 5 and 6. enhanced as pressure increases. This increase in the number of
From these figures, it is obvious that the polar−polar peaks and gauche kinks in the cationic tails facilitates bending or curling of
polar−apolar antipeaks are also susceptible to the pressure the cation tails toward their head groups, which can be clearly
change along with the apolar−apolar peaks which correspond seen in the intramolecular distribution functions for cationic
3209 DOI: 10.1021/acs.jpcb.6b01133
J. Phys. Chem. B 2016, 120, 3206−3214
The Journal of Physical Chemistry B Article

Figure 6. Cation−anion (black line), cation (head)−anion (red line), Figure 7. Polar−polar (black line), polar−apolar (red line), and
and cation (tail)−anion (blue line) partial S(q)s of (a) Pyrr1,8+/NTf2− apolar−apolar (blue line) partial S(q)s of (a) Pyrr1,8+/NTf2− and (b)
and (b) Pyrr1,10+/NTf2− RTILs at different pressures. Pyrr1,10+/NTf2− RTILs at different pressures.

nitrogen and the terminal carbon (CTerminal) of the alkyl tail


(Figure 10). We also notice that the probability of finding two
or three trans conformations along the Pyrr1,8+ cation and three
or four along the Pyrr1,10+ cation increases considerably when
the pressure is changed from 1000 to 10000 bar. We also
observe significant broadening of the nearest neighbor peak in
the Ncat−CTerminal intramolecular distribution function at the
highest pressure.
It has been demonstrated previously42,44,63,92 that in liquids
the NTf2− anions exist in equilibrium between two conforma-
tional states about the C−S−S−C pseudodihedral angles, viz.,
gauche or C1 and trans or C2 conformers. The distribution of
intramolecular C−C (carbon−carbon) distance in the anion
also displays two peaks for both of the ionic liquids; the shorter
distance peak resembles the C1 conformer, and the longer C−C
separation peak corresponds to the C2 conformer. Moreover, at
ambient pressure and temperature, the C 2 conformer
dominates. In Figure 11, we show that as the pressure increases
from 100 to 10000 bar the probability of finding C2 conformers
of the anion in both liquids decreases and that of C1 increases.

4. CONCLUSION
In this study, we have used molecular dynamics simulations to
appreciate the structures of Pyrr1,8+/NTf2− and Pyrr1,10+/NTf2− Figure 8. (a) Intermolecular RDFs for the CTS group of the cationic
RTILs at different pressures. The simulated X-ray scattering tail for Pyrr1,8+/NTf2− (left panel) and Pyrr1,10+/NTf2− (right panel)
structure functions show that the characteristic ordering peaks, RTILs at different pressures. (b) Intermolecular RDFs for cation
especially charge-ordering and polarity-ordering, in both RTILs nitrogen and the CTS group of the cationic tail for Pyrr1,8+/NTf2− (left
are susceptible to external pressure; the strength of charge- panel) and Pyrr1,10+/NTf2− (right panel) RTILs at different pressures.
ordering and polarity-ordering decreases as the applied pressure
increases. Upon increasing pressure, a significant decrease in
3210 DOI: 10.1021/acs.jpcb.6b01133
J. Phys. Chem. B 2016, 120, 3206−3214
The Journal of Physical Chemistry B Article

Figure 11. (a) Probability distribution function of C−S−S−C dihedral


angles for NTf2−. (b) Intramolecular distance distribution function for
Figure 9. Probability distribution for the number of trans dihedrals the carbon atoms of NTf2−.
along the cationic tail. The atoms considered are from the ring
nitrogen through the terminal carbon. In this study, a dihedral angle is
defined as trans when the dihedral angle is within the range 180 ± 20°.
pressures; this leads to increased bending or curling of the tails
at higher pressure. The probability of finding gauche or C1
conformers of the anion is enhanced at higher pressure, and
that of the trans or C2 conformers is diminished.

■ AUTHOR INFORMATION
Corresponding Author
*E-mail: hkashyap@chemistry.iitd.ac.in. Phone: +91-(0)11-
26591518. Fax: +91-(0)11-26581102.
Notes
The authors declare no competing financial interest.

Figure 10. Intramolecular distribution function for the distances


■ ACKNOWLEDGMENTS
We sincerely thank Professor Ranjit Biswas, Professor Claudio
between cationic nitrogen and terminal carbon (CTerminal) of the J. Margulis, and Professor Edward W. Castner, Jr., for support
longest tail for Pyrr1,8+/NTf2− (left panel) and Pyrr1,10+/NTf2− (right
and encouragement. S.S. and A.G. thank CSIR-UGC, India, for
panel) RTILs at different pressures.
a fellowship. We also thank IIT Delhi for providing the
supercomputing facility. This work is supported by the
the nearest neighbor pair correlations for both polar−polar and Department of Science and Technology (DST), India, through
a grant awarded to H.K.K. (Grant No. SB/FT/CS-124/2014).


apolar−apolar groups of the RTILs is also observed. On the
other hand, the polar−apolar nearest neighbor correlations are
stronger at higher pressure. This finding is different from those REFERENCES
observed by Russina et al.70 for Omim+/BF4−, wherein only (1) Stuart, M. A. C.; Huck, W. T. S.; Genzer, J.; Muller, M.; Ober, C.;
apolar moieties are influenced by pressure change. In the Stamm, M.; Sukhorukov, G. B.; Szleifer, I.; Tsukruk, V. V.; Urban, M.;
et al. Emerging Applications of Stimuli-responsive Polymer Materials.
present case, the susceptibility of the polar group correlations Nat. Mater. 2010, 9, 101−113.
toward the applied pressure could be because of the difference (2) Welton, T. Room-Temperature Ionic Liquids. Solvents for
in the anion and the cation ring; the imidazolium ring and BF4− Synthesis and Catalysis. Chem. Rev. 1999, 99, 2071−2084.
have a lower conformational flexibility than the pyrrolidinium (3) Armand, M.; Endres, F.; MacFarlane, D. R.; Ohno, H.; Scrosati,
ring and NTf2−. The alkyl tails of cations possess an increased B. Ionic-Liquid Materials for the Electrochemical Challenges of the
number of gauche defects at higher pressures than that in lower Future. Nat. Mater. 2009, 8, 621−629.

3211 DOI: 10.1021/acs.jpcb.6b01133


J. Phys. Chem. B 2016, 120, 3206−3214
The Journal of Physical Chemistry B Article

(4) Gomes, M.; Lopes, J. N. C.; Pádua, A. A. H. Thermodynamics (22) Gontrani, L.; Russina, O.; Celso, F. L.; Caminiti, R.; Annat, G.;
and Micro Heterogeneity of Ionic Liquids. Top. Curr. Chem. 2009, Triolo, A. Liquid Structure of Trihexyltetradecylphosphonium
290, 161−183. Chloride at Ambient Temperature: An X-Ray Scattering and
(5) Wishart, J. F.; Castner, E. W., Jr. The Physical Chemistry of Ionic Simulation Study. J. Phys. Chem. B 2009, 113, 9235−9240.
Liquids. J. Phys. Chem. B 2007, 111, 4639−4640. (23) Pott, T.; Méléard, P. New Insight into the Nanostructure of
(6) Castner, E. W., Jr.; Wishart, J. F. Spotlight on Ionic Liquids. J. Ionic Liquids: A Small Angle X-Ray Scattering (SAXS) Study on
Chem. Phys. 2010, 132, 120901. Liquid Tri-Alkyl-Methyl-Ammonium Bis(trifluoromethanesulfonyl)-
(7) Castner, E. W., Jr.; Margulis, C. J.; Maroncelli, M.; Wishart, J. F. amides and Their Mixtures. Phys. Chem. Chem. Phys. 2009, 11,
Ionic Liquids: Structure and Photochemical Reactions. Annu. Rev. Phys. 5469−5475.
Chem. 2011, 62, 85−105. (24) Fujii, K.; Mitsugi, T.; Takamuku, T.; Yanaguchi, T.;
(8) Russina, O.; Triolo, A. New Experimental Evidence Supporting Umebayashi, Y.; Ishiguro, S.-i. Effect of Methylation at the C2
the Mesoscopic Segregation Model in Room Temperature Ionic Position of Imidazolium on the Structure of Ionic Liquids Revealed by
Liquids. Faraday Discuss. 2012, 154, 97−109. Large Angle X-Ray Scattering Experiments and MD Simulations.
(9) Russina, O.; Triolo, A.; Gontrani, L.; Caminiti, R. Mesoscopic Chem. Lett. 2009, 38, 340−341.
Structural Heterogeneities in Room-Temperature Ionic Liquids. J. (25) Zheng, W.; Mohammed, A.; Hines, L. G.; Xiao, D.; Martinez, O.
Phys. Chem. Lett. 2012, 3, 27−33. J.; Bartsch, R. A.; Simon, S. L.; Russina, O.; Triolo, A.; Quitevis, E. L.
(10) Castiglione, F.; Moreno, M.; Raos, G.; Famulari, A.; Mele, A.; Effect of Cation Symmetry on the Morphology and Physicochemical
Appetecchi, G. B.; Passerini, S. Structural Organization and Transport Properties of Imidazolium Ionic Liquids. J. Phys. Chem. B 2011, 115,
Properties of Novel Pyrrolidinium-Based Ionic Liquids with 6572−6584.
Perfluoroalkyl Sulfonylimide Anions. J. Phys. Chem. B 2009, 113, (26) Santos, C. S.; Murthy, N. S.; Baker, G. A.; Castner, E. W., Jr.
10750−10759. Communication: X-Ray Scattering from Ionic Liquids with Pyrrolidi-
(11) Lee, H. Y.; Shirota, H.; Castner, E. W. Differences in Ion nium Cations. J. Chem. Phys. 2011, 134, 121101.
Interactions for Isoelectronic Ionic Liquid Homologs. J. Phys. Chem. (27) Aoun, B.; Goldbach, A.; Gonzalez, M. A.; Kohara, S.; Price, D.
Lett. 2013, 4, 1477−1483. L.; Saboungi, M.-L. Nanoscale Heterogeneity in Alkyl-Methylimida-
(12) Bradley, A. E.; Hardacre, C.; Holbrey, J. D.; Johnston, S.; zolium Bromide Ionic Liquids. J. Chem. Phys. 2011, 134, 104509.
McMath, S. E. J.; Nieuwenhuyzen, M. Small-Angle X-Ray Scattering (28) Fujii, K.; Kanzaki, R.; Takamuku, T.; Kameda, Y.; Kohara, S.;
Studies of Liquid Crystalline 1-Alkyl-3-Methylimidazolium Salts. Kanakubo, M.; Shibayama, M.; Ishiguro, S.; Umebayashi, Y.
Chem. Mater. 2002, 14, 629−635. Experimental Evidences for Molecular Origin of Low-Q Peak in
(13) Triolo, A.; Mandanici, A.; Russina, O.; Rodriguez-Mora, V.;
Neutron/X-Ray Scattering of 1-Alkyl-3-Methylimidazolium Bis-
Cutroni, M.; Hardacre, C.; Nieuwenhuyzen, M.; Bleif, H.-J.; Keller, L.;
(trifluoromethanesulfonyl)amide Ionic Liquids. J. Chem. Phys. 2011,
Ramos, M. A. Thermodynamics, Structure, and Dynamics in Room
135, 244502.
Temperature Ionic Liquids: The Case of 1-Butyl-3-Methyl Imidazo-
(29) Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. Pronounced
lium Hexafluorophosphate ([bmim][PF6]). J. Phys. Chem. B 2006,
Sponge-like Nanostructure in Propylammonium Nitrate. Phys. Chem.
110, 21357−21364.
(14) Triolo, A.; Russina, O.; Bleif, H.-J.; Di Cola, E. Nanoscale Chem. Phys. 2011, 13, 13544−13551.
(30) Umebayashi, Y.; Hamano, H.; Seki, S.; Minofar, B.; Fujii, K.;
Segregation in Room Temperature Ionic Liquids? J. Phys. Chem. B
2007, 111, 4641−4644. Hayamizu, K.; Tsuzuki, S.; Kameda, Y.; Kohara, S.; Watanabe, M.
(15) Triolo, A.; Russina, O.; Fazio, B.; Triolo, R.; DiCola, E. Liquid Structure of and Li+ Ion Solvation in Bis-
Morphology of 1-Alkyl-3-Methylimidazolium Hexafluorophosphate (trifluoromethanesulfonyl)amide Based Ionic Liquids Composed of
Room Temperature Ionic Liquids. Chem. Phys. Lett. 2008, 457, 1-Ethyl-3-Methylimidazolium and N-Methyl-N-Propylpyrrolidinium
362−365. Cations. J. Phys. Chem. B 2011, 115, 12179−12191.
(16) Atkin, R.; Warr, G. G. The Smallest Amphiphiles: Nanostruc- (31) Triolo, A.; Russina, O.; Arrighi, V.; Juranyi, F.; Janssen, S.;
ture in Protic Room-Temperature Ionic Liquids with Short Alkyl Gordon, C. Quasielastic Neutron Scattering Characterization of the
Groups. J. Phys. Chem. B 2008, 112, 4164−4166. Relaxation Processes in a Room Temperature Ionic Liquid. J. Chem.
(17) Fukuda, S.; Takeuchi, M.; Fujii, K.; Kanzaki, R.; Takamuku, T.; Phys. 2003, 119, 8549−8557.
Chiba, K.; Yamamoto, H.; Umebayashi, Y.; Ishiguro, S.-i. Liquid (32) Hardacre, C.; Holbrey, J. D.; Nieuwenhuyzen, M.; Youngs, T. G.
Structure of N-Butyl-N-Methylpyrrolidinium Bis- A. Structure and Solvation in Ionic Liquids. Acc. Chem. Res. 2007, 40,
(trifluoromethanesulfonyl) Amide Ionic Liquid Studied by Large 1146−1155.
Angle X-Ray Scattering and Molecular Dynamics Simulations. J. Mol. (33) Hardacre, C.; Holbrey, J. D.; Mullan, C. L.; Youngs, T. G. A.;
Liq. 2008, 143, 2−7. Bowron, D. T. Small Angle Neutron Scattering from 1-Alkyl-3-
(18) Fujii, K.; Seki, S.; Fukuda, S.; Takamuku, T.; Kohara, S.; Methylimidazolium Hexafluorophosphate Ionic Liquids ([Cnmim]-
Kameda, Y.; Umebayashi, Y.; Ishiguro, S.-i. Liquid Structure and [PF6], n = 4, 6, and 8). J. Chem. Phys. 2010, 133, 74510−74517.
Conformation of a Low-Viscosity Ionic Liquid, N-Methyl-N-Propyl- (34) Urahata, S. M.; Ribeiro, M. C. Structure of Ionic Liquids of 1-
Pyrrolidinium Bis(fluorosulfonyl) Imide Studied by High-Energy X- Alkyl-3-Methylimidazolium Cations: A Systematic Computer Simu-
Ray Scattering. J. Mol. Liq. 2008, 143, 64−69. lation Study. J. Chem. Phys. 2004, 120, 1855−1863.
(19) Xiao, D.; Hines, L. G., Jr.; Li, S.; Bartsch, R. A.; Quitevis, E. L.; (35) Del Popolo, M. G.; Voth, G. A. On the Structure and Dynamics
Russina, O.; Triolo, A. Effect of Cation Symmetry and Alkyl Chain of Ionic Liquids. J. Phys. Chem. B 2004, 108, 1744−1752.
Length on the Structure and Intermolecular Dynamics of 1,3- (36) Wang, Y.; Voth, G. A. Unique Spatial Heterogeneity in Ionic
Dialkylimidazolium Bis(trifluoromethanesulfonyl)amide Ionic Liquids. Liquids. J. Am. Chem. Soc. 2005, 127, 12192−12193.
J. Phys. Chem. B 2009, 113, 6426−6433. (37) Wang, Y.; Voth, G. A. Tail Aggregation and Domain Diffusion
(20) Triolo, A.; Russina, O.; Fazio, B.; Appetecchi, G. B.; Carewska, in Ionic Liquids. J. Phys. Chem. B 2006, 110, 18601−18608.
M.; Passerini, S. Nanoscale Organization In Piperidinium-Based Room (38) Borodin, O.; Smith, G. D. Structure and Dynamics of N-Methyl-
Temperature Ionic Liquids. J. Chem. Phys. 2009, 130, 164521. N-Propylpyrrolidinium Bis(trifluoromethanesulfonyl)imide Ionic
(21) Russina, O.; Triolo, A.; Gontrani, L.; Caminiti, R.; Xiao, D.; Liquid from Molecular Dynamics Simulations. J. Phys. Chem. B
Hines, L. G., Jr.; Bartsch, R. A.; Quitevis, E. L.; Pleckhova, N.; Seddon, 2006, 110, 11481−11490.
K. R. Morphology and Intermolecular Dynamics of 1-Alkyl-3- (39) Bhargava, B. L.; Devane, R.; Klein, M. L.; Balasubramanian, S.
Methylimidazolium Bis(trifluoromethanesulfonyl)amide Ionic Liquids: Nanoscale Organization in Room Temperature Ionic Liquids: A
Structural and Dynamic Evidence of Nanoscale Segregation. J. Phys.: Coarse Grained Molecular Dynamics Simulation Study. Soft Matter
Condens. Matter 2009, 21, 424121. 2007, 3, 1395−1400.

3212 DOI: 10.1021/acs.jpcb.6b01133


J. Phys. Chem. B 2016, 120, 3206−3214
The Journal of Physical Chemistry B Article

(40) Bhargava, B.; Klein, M.; Balasubramanian, S. Structural Hexafluorophosphate [bmim][PF6]. High Pressure Res. 2011, 31, 35−
Correlations and Charge Ordering in a Room-Temperature Ionic 38.
Liquid. ChemPhysChem 2008, 9, 67−70. (59) Imai, Y.; Takekiyo, T.; Abe, H.; Yoshimura, Y. Pressure- and
(41) Canongia Lopes, J. N.; Pádua, A. A. H. Nanostructural Temperature-induced Raman Spectral Changes of 1-Butyl-3-Methyl-
Organization in Ionic Liquids. J. Phys. Chem. B 2006, 110, 3330−3335. imidazolium Tetrafluoroborate. High Pressure Res. 2011, 31, 53−57.
(42) Canongia Lopes, J. N.; Shimizu, K.; Pádua, A. A. H.; (60) Su, L.; Li, M.; Zhu, X.; Wang, Z.; Chen, Z.; Li, F.; Zhou, Q.;
Umebayashi, Y.; Fukuda, S.; Fujii, K.; Ishiguro, S.-i. A Tale of Two Hong, S. In Situ Crystallization of Low-Melting Ionic Liquid
Ions: The Conformational Landscapes of Bis- [BMIM][PF6] under High Pressure up to 2 GPa. J. Phys. Chem. B
(trifluoromethanesulfonyl)amide and N,N-Dialkylpyrrolidinium. J. 2010, 114, 5061−5065.
Phys. Chem. B 2008, 112, 1465−1472. (61) Zhao, Y.; Liu, X.; Lu, X.; Zhang, S.; Wang, J.; Wang, H.; Gurau,
(43) Annapureddy, H. V. R.; Kashyap, H. K.; De Biase, P. M.; G.; Rogers, R. D.; Su, L.; Li, H. The Behavior of Ionic Liquids under
Margulis, C. J. What Is the Origin of the Prepeak in the X-Ray High Pressure: A Molecular Dynamics Simulation. J. Phys. Chem. B
Scattering of Imidazolium-Based Room-Temperature Ionic Liquids? J. 2012, 116, 10876−10884.
Phys. Chem. B 2010, 114, 16838−16846. (62) Russina, O.; Fazio, B.; Schmidt, C.; Triolo, A. Structural
(44) Santos, C. S.; Annapureddy, H. V. R.; Murthy, N. S.; Kashyap, Organization and Phase Behaviour of 1-Butyl-3-Methylimidazolium
H. K.; Castner, E. W., Jr.; Margulis, C. J. Temperature-Dependent Hexafluorophosphate: an High Pressure Raman Spectroscopy Study.
Structure of Methyltributylammonium Bis(trifluoromethylsulfonyl)- Phys. Chem. Chem. Phys. 2011, 13, 12067−12074.
amide: X Ray Scattering and Simulations. J. Chem. Phys. 2011, 134, (63) Yoshimura, Y.; Takekiyo, T.; Imai, Y.; Abe, H. Pressure-Induced
064501. Spectral Changes of Room-Temperature Ionic Liquid, N,N-Diethyl-N-
(45) Kashyap, H. K.; Santos, C. S.; Annapureddy, H. V. R.; Murthy, Methyl-N-(2-Methoxyethyl)ammonium Bis(trifluoromethylsulfonyl)-
N. S.; Margulis, C. J.; Castner, E. W., Jr. Temperature-Dependent imide, [DEME][TFSI]. J. Phys. Chem. C 2012, 116, 2097−2101.
Structure of Ionic Liquids: X-Ray Scattering and Simulations. Faraday (64) Shimizu, K.; Bernardes, C. E. S.; Triolo, A.; Canongia Lopes, J.
Discuss. 2012, 154, 133−143. N. Nano-segregation in Ionic Liquids: Scorpions and Vanishing
(46) Siqueira, L. J. A.; Ribeiro, M. C. C. Charge Ordering and Chains. Phys. Chem. Chem. Phys. 2013, 15, 16256−16262.
Intermediate Range Order in Ammonium Ionic Liquids. J. Chem. Phys. (65) Faria, L. F. O.; Penna, T. C.; Ribeiro, M. C. C. Raman
2011, 135, 204506. Spectroscopic Study of Temperature and Pressure Effects on the Ionic
(47) Li, S.; Banuelos, J. L.; Guo, J.; Anovitz, L.; Rother, G.; Shaw, R. Liquid Propylammonium Nitrate. J. Phys. Chem. B 2013, 117, 10905−
W.; Hillesheim, P. C.; Dai, S.; Baker, G. A.; Cummings, P. T. Alkyl 10912.
Chain Length and Temperature Effects on Structural Properties of (66) Abe, H.; Takekiyo, T.; Hatano, N.; Shigemi, M.; Hamaya, N.;
Pyrrolidinium-Based Ionic Liquids: A Combined Atomistic Simulation Yoshimura, Y. Pressure-Induced Frustration-Frustration Process in 1-
and Small-Angle X-Ray Scattering Study. J. Phys. Chem. Lett. 2012, 3, Butyl-3-Methylimidazolium Hexafluorophosphate, a Room-Temper-
125−130. ature Ionic Liquid. J. Phys. Chem. B 2014, 118, 1138−1145.
(48) Kashyap, H. K.; Santos, C. S.; Murthy, N. S.; Hettige, J. J.; Kerr, (67) Yoshimura, Y.; Shigemi, M.; Takaku, M.; Yamamura, M.;
K.; Ramati, S.; Gwon, J.; Gohdo, M.; Lall-Ramnarine, S. I.; Wishart, J. Takekiyo, T.; Abe, H.; Hamaya, N.; Wakabayashi, D.; Nishida, K.;
F.; et al. Structure of 1-Alkyl-1-Methylpyrrolidinium Bis- Funamori, N.; et al. Stability of the Liquid State of Imidazolium-Based
(trifluoromethylsulfonyl)amide Ionic Liquids with Linear, Branched, Ionic Liquids under High Pressure at Room Temperature. J. Phys.
and Cyclic Alkyl Groups. J. Phys. Chem. B 2013, 117, 15328−15337. Chem. B 2015, 119, 8146−8153.
(49) Shimizu, K.; Bernardes, C. E. S.; Canongia Lopes, J. N. Structure (68) Faria, L. F. O.; Ribeiro, M. C. C. Phase Transitions of Triflate-
and Aggregation in the 1-Alkyl-3-Methylimidazolium Bis- Based Ionic Liquids under High Pressure. J. Phys. Chem. B 2015, 119,
(trifluoromethylsulfonyl)imide Ionic Liquid Homologous Series. J. 14315−14322.
Phys. Chem. B 2014, 118, 567−576. (69) Li, H.; Wang, Z.; Chen, L.; Wu, J.; Huang, H.; Yang, K.; Wang,
(50) Maginn, E. J. Molecular Simulation of Ionic Liquids: Current Y.; Su, L.; Yang, G. Kinetic Effect on Pressure-Induced Phase
Status and Future Opportunities. J. Phys.: Condens. Matter 2009, 21, Transitions of Room Temperature Ionic Liquid, 1-Ethyl-3-Methyl-
373101. imidazolium Trifluoromethanesulfonate. J. Phys. Chem. B 2015, 119,
(51) Russina, O.; Fazio, B.; Di Marco, G.; Caminiti, R.; Triolo, A. In 14245−14251.
The Structure of Ionic Liquids; Caminiti, R., Gontrani, L., Eds.; Soft and (70) Russina, O.; Celso, F. L.; Triolo, A. Pressure-responsive
Biological Matter; Springer International Publishing: Switzerland, Mesoscopic Structures in Room Temperature Ionic Liquids. Phys.
2014; pp 39−61. Chem. Chem. Phys. 2015, 17, 29496−29500.
(52) Hayes, R.; Warr, G. G.; Atkin, R. Structure and Nanostructure in (71) Suarez, S. N.; Rúa, A.; Cuffari, D.; Pilar, K.; Hatcher, J. L.;
Ionic Liquids. Chem. Rev. 2015, 115, 6357−6426. Ramati, S.; Wishart, J. F. Do TFSA Anions Slither? Pressure Exposes
(53) Murphy, T.; Atkin, R.; Warr, G. G. Scattering from Ionic the Role of TFSA Conformational Exchange in Self-Diffusion. J. Phys.
Liquids. Curr. Opin. Colloid Interface Sci. 2015, 20, 282−292. Chem. B 2015, 119, 14756−14765.
(54) Araque, J. C.; Hettige, J. J.; Margulis, C. J. Modern Room (72) Mariani, A.; Caminiti, R.; Campetella, M.; Gontrani, L. Pressure-
Temperature Ionic Liquids, a Simple Guide to Understanding Their Induced Mesoscopic Disorder in Protic Ionic Liquids: First Computa-
Structure and How It May Relate to Dynamics. J. Phys. Chem. B 2015, tional Study. Phys. Chem. Chem. Phys. 2016, 18, 2297−2302.
119, 12727−12740. (73) Kashyap, H. K.; Santos, C. S.; Daly, R. P.; Hettige, J. J.; Murthy,
(55) Lauw, Y.; Horne, M. D.; Rodopoulos, T.; Lockett, V.; Akgun, B.; N. S.; Shirota, H.; Castner, E. W.; Margulis, C. J. How Does the Ionic
Hamilton, W. A.; Nelson, A. R. J. Structure of [C4mpyr][NTf2] Room- Liquid Organizational Landscape Change When Nonpolar Cationic
Temperature Ionic Liquid at Charged Gold Interfaces. Langmuir 2012, Alkyl Groups are Replaced by Polar Isoelectronic Diethers? J. Phys.
28, 7374−7381. Chem. B 2013, 117, 1130−1135.
(56) Hettige, J. J.; Kashyap, H. K.; Margulis, C. J. Communication: (74) Salminen, J.; Papaiconomou, N.; Kumar, R. A.; Lee, J.-M.; Kerr,
Anomalous Temperature Dependence of the Intermediate Range J.; Newman, J.; Prausnitz, J. M. Physicochemical Properties and
Order in Phosphonium Ionic Liquids. J. Chem. Phys. 2014, 140, Toxicities of Hydrophobic Piperidinium and Pyrrolidinium Ionic
111102. Liquids. Fluid Phase Equilib. 2007, 261, 421−426.
(57) Shah, J. K.; Maginn, E. J. Molecular Dynamics Investigation of (75) Kashyap, H. K.; Hettige, J. J.; Annapureddy, H. V. R.; Margulis,
Biomimetic Ionic Liquids. Fluid Phase Equilib. 2010, 294, 197−205. C. J. SAXS Anti-Peaks Reveal the Length-Scales of Dual Positive-
(58) Takekiyo, T.; Hatano, N.; Imai, Y.; Abe, H.; Yoshimura, Y. Negative and Polar-Apolar Ordering in Room-Temperature Ionic
Pressure-induced Phase Transition of 1-Butyl-3-Methylimidazolium Liquids. Chem. Commun. 2012, 48, 5103−5105.

3213 DOI: 10.1021/acs.jpcb.6b01133


J. Phys. Chem. B 2016, 120, 3206−3214
The Journal of Physical Chemistry B Article

(76) Kashyap, H. K.; Margulis, C. J. (Keynote) Theoretical


Deconstruction of the X-Ray Structure Function Exposes Polarity
Alternations in Room Temperature Ionic Liquids. ECS Trans. 2013,
50, 301−307.
(77) Smith, A. M.; Lovelock, K. R. J.; Gosvami, N. N.; Licence, P.;
Dolan, A.; Welton, T.; Perkin, S. Monolayer to Bilayer Structural
Transition in Confined Pyrrolidinium-Based Ionic Liquids. J. Phys.
Chem. Lett. 2013, 4, 378−382.
(78) 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, 435−447.
(79) 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, 1701−1718.
(80) Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. Development
and Testing of the OPLS All-Atom Force Field on Conformational
Energetics and Properties of Organic Liquids. J. Am. Chem. Soc. 1996,
118, 11225−11236.
(81) Kaminski, G.; Jorgensen, W. L. Performance of the AMBER94,
MMFF94, and OPLS-AA Force Fields for Modeling Organic Liquids.
J. Phys. Chem. 1996, 100, 18010−18013.
(82) Rizzo, R. C.; Jorgensen, W. L. OPLS All-Atom Model for
Amines: Resolution of the Amine Hydration Problem. J. Am. Chem.
Soc. 1999, 121, 4827−4836.
(83) Price, M. L. P.; Ostrovsky, D.; Jorgensen, W. L. Gas-Phase and
Liquid-State Properties of Esters, Nitriles, and Nitro Compounds with
the OPLS-AA Force Field. J. Comput. Chem. 2001, 22, 1340−1352.
(84) Canongia Lopes, J. N.; Pádua, A. A. H. Molecular Force Field
for Ionic Liquids Composed of Triflate or Bistriflylimide Anions. J.
Phys. Chem. B 2004, 108, 16893−16898.
(85) Canongia Lopes, J. N.; Pádua, 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, 19586−19592.
(86) Parrinello, M.; Rahman, A. Polymorphic Transitions in Single
Crystals: A New Molecular Dynamics Method. J. Appl. Phys. 1981, 52,
7182−7190.
(87) Darden, T.; York, D.; Pedersen, L. Particle Mesh Ewald: An
N.log(N) Method for Ewald Sums in Large Systems. J. Chem. Phys.
1993, 98, 10089−10092.
(88) Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.;
Pedersen, L. G. A Smooth Particle Mesh Ewald Method. J. Chem. Phys.
1995, 103, 8577−8593.
(89) Hahn, T.; Shmueli, U.; Wilson, A. J. C.; Prince, E., Ed.;
International Tables for Crystallography; International Union of
Crystallography: Chester, England, 2006; Vol. C.
(90) Lorch, E. Neutron Diffraction by Germania, Silica and
Radiation-Damaged Silica Glasses. J. Phys. C: Solid State Phys. 1969,
2, 229−237.
(91) Du, J.; Benmore, C. J.; Corrales, R.; Hart, R. T.; Weber, J. K. R.
A Molecular Dynamics Simulation Interpretation of Neutron and X-
Ray Diffraction Measurements on Single Phase Y2O3-Al2O3 Glasses. J.
Phys.: Condens. Matter 2009, 21, 205102.
(92) Fujii, K.; Fujimori, T.; Takamuku, T.; Kanzaki, R.; Umebayashi,
Y . ; I s h i g u r o , S . C o n f o r m a t i o n a l E q u i l i b r i u m o f Bi s -
(trifluoromethanesulfonyl) Imide Anion of a Room-Temperature
Ionic Liquid: Raman Spectroscopic Study and DFT Calculations. J.
Phys. Chem. B 2006, 110, 8179−8183.

3214 DOI: 10.1021/acs.jpcb.6b01133


J. Phys. Chem. B 2016, 120, 3206−3214

You might also like