40
L. Tanzi et al. / Journal of Molecular Liquids 218 (2016) 39–49
patterns from measured X-ray and those from MD simulations can be
compared to provide a validation of the theoretical models here applied.
Alternatively to X-ray studies, vibrational spectroscopy can be
extremely useful for evaluating the nature of the cation–anion interac-
tions, identifying ion pair coupling and clarifying conformational iso-
mers [17-27]. For example, the presence and the force of hydrogen
bonding may be obtained from the vibrational frequencies of the OH
group of choline and the COO– group of carboxylate since directly in-
volved in such interaction. On the other hand non-empirical AIMD sim-
ulations have been recently and successfully applied to a wide number
of ionic liquids [28-32] to investigate their vibrational properties and as-
sign the observed absorptions.
(with nitrogen as sharpening atom) to enhance the resolution of the
curve at high q values and to decrease the truncation error in the calcu-
lation of the Fourier transform from the reciprocal space (q) to the direct
one (r). The expression for M(q) is
2
f ð0Þ
ꢀ
ꢁ
N
2
MðqÞ ¼
exp ꢀ0:01q
ð3Þ
2
f ðqÞ
N
Fourier transform of qI(q)M(q) led to radial distribution function
r.d.f.)
(
2r
π
qmax
0
2
DðrÞ ¼ 4πr ρ þ
∫
qIðqÞMðqÞ sinðrqÞdq
ð4Þ
0
2
. Experimental details
2
3
where ρ
0
(electrons /Å ) is the bulk number density. When the uniform
2
2
.1. Synthesis
distribution component is dropped (4πr ρ ), we obtain the differential
0
correlation function, Diff(r), which contains only the structural contri-
Reagents used in the synthesis were choline hydroxide solution
bution to the distribution function,
(
2
46 wt.% in H O, Aldrich) and salicylic acid(99%, Aldrich), both used
2
without further purification. Choline salicylate was synthesized by
dropwise addition of salicylic acid to the choline hydroxide solution in
ratio 1:1 (e.g. 0.12 mol:0.12 mol), stirring continuously at room temper-
ature and pressure for 12 h. Most of the water in the reaction mixture
was removed under reduced pressure, using a rotary evaporator at
Diff ðrÞ ¼ DðrÞ ꢀ 4πr ρ0
ð5Þ
2.3. Infrared and Raman spectra
7
0 °C for 4 h. The mixture was then dried in vacuo, heating at 70 °C
The Fourier transformed infrared (FTIR) spectrum was measured at
−
1
and stirring for 24 h. The purity of these BioILs was checked by 1H
NMR and 13C NMR spectroscopy, using a Bruker Avance III spectrome-
ter operating at 400 MHz and 100.6 MHz, respectively. The spectra have
confirmed the absence of any major impurities and the water final
content, evaluated by 1H NMR analysis, has been estimated below
room temperature from 4000 to 400 cm
after 100 scans using the
Perkin Elmer Spectrum two FT-IR spectrometer. Liquid samples were
placed as thin films between KBr plates. The Raman spectrum was
measured using a LABRAM confocal-microscope Raman spectrometer
by HORIBA Jobin Yvon using 5 mW at 632 nm excitation source and a
20× collection optics. The instrumental resolution is of the order of
0
.4 wt.%.
−
1
2
–3 cm . Background fluorescence has been fitted using a polynomial
2
.2. X-ray scattering
expression and subtracted from the data.
3. Computational details
The large angle X-ray scattering experiments were performed using
the non-commercial energy-scanning diffractometer built in the De-
partment of Chemistry at the University La Sapienza of Rome. (Patent
no. 01126484— 23 June 1993). Detailed description of instrument, tech-
nique and the experimental protocol (instrument geometry and scatter-
ing angles) of the data acquisition phase can be found in a series of
papers by our group [33-38]. The appropriate measuring time (i.e. num-
ber of counts) was chosen to obtain scattering variable (q) spectra with
high signal to noise ratio (500,000 counts on average).
3.1. Liquids by molecular dynamics
In order to describe the liquid structure we have followed two differ-
ent approaches. Local structural characterization with particular atten-
tion to the hydrogen bonding interaction occurring between ion pairs
was initially performed by ab initio methods.
Quantum-mechanical calculations on the choline–salicylate ion pair
were performed using the Gaussian 09 package [39]. Equilibrium
geometry and vibrational frequencies were obtained using Moller–
Plesset (MP2) and density functional theory (DFT) methods with the
B3LYP [40,41] exchange and correlation functional and employing the
The expression of q is:
q ¼ 4
π sinðθÞ
¼ 1:014E sinðλÞ
ð1Þ
λ
6
-311++G** basis set.
Dynamic effects were introduced at ab initio level by AIMD simula-
where E is expressed in keV and q in Å−1. The various angular data were
processed according to the procedure described in literature [33-38],
normalized to a stoichiometric unit of volume containing one ion pair
and combined to yield the total (static) “structure factor”, I(q),
tions performed on a model of 9 ion pairs using the Born–Oppenheimer
Molecular Dynamics (BOMD) method implemented in the CP2K code
[
42]. Potential energy calculations were carried out using BLYP
exchange-correlation functional [41,43] with the Grimme's dispersion
correction D3 [44] and the hybrid Gaussian and plane wave (GPW)
basis set [45]. Gaussian basis set was DZVP-MOLOPTSR-GTH and plane
waves expansion was developed in a periodic cubic system with unit
N
2
i
IðqÞ ¼ Ie:u:ðqÞ ꢀ ∑ xi f
ð2Þ
i¼1
3
where f
i
are the atomic scattering factors, x
i
are the number concentra-
cell edge of 14.53 Å and truncated at 320 Ry. Goedecker–Teter–Hutter
tions of the i-type atoms in the stoichiometric unit (i.e. the group of
(GTH) pseudopotentials [46,47] were employed to describe core elec-
trons. A pre-equilibration was performed employing classical molecular
dynamics within periodic boundary conditions using the two-body
Generalized Amber Force Field (GAFF) [48]. AIMD simulations started
from a snapshot of the classical MD simulation and the system was
equilibrated for 6.3 ps in the NVT ensemble at 300 K using the individual
thermostat for each degree of freedom for the first 3 ps and a global
Nosé–Hooverchain thermostat [49-51] for the remaining time.
Timestep was set to 0.4 fs. Trajectory was then collected for 20 ps in
particles used as reference for data normalization, the ion pair in this
2
case) and Ie.u. is the observed intensity in electron units (electrons ).
Such function depends on the scattering contributions of all the par-
ticles of the system, according to the pairwise distances between them
(
in the case of X-rays, the atoms scatter the radiation through the elec-
tron clouds surrounding them) and therefore, it gives a sort of mathe-
matical picture of the spatial disposition of the sample atoms. This
function is multiplied by q and q-dependent sharpening factor, M(q)