You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/323475152

Systematic comparison of force fields for molecular dynamic simulation of


Au(111)/Ionic liquid interfaces

Article  in  Fluid Phase Equilibria · May 2018


DOI: 10.1016/j.fluid.2018.01.024

CITATIONS READS

16 273

4 authors, including:

Runxi Wang Sheng Bi


Huazhong University of Science and Technology Huazhong University of Science and Technology
20 PUBLICATIONS   297 CITATIONS    31 PUBLICATIONS   436 CITATIONS   

SEE PROFILE SEE PROFILE

Guang Feng
Huazhong University of Science and Technology
132 PUBLICATIONS   4,399 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

CO2 EOR View project

Electrical Double Layer View project

All content following this page was uploaded by Sheng Bi on 03 December 2018.

The user has requested enhancement of the downloaded file.


Fluid Phase Equilibria 463 (2018) 106e113

Contents lists available at ScienceDirect

Fluid Phase Equilibria


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / fl u i d

Systematic comparison of force fields for molecular dynamic


simulation of Au(111)/Ionic liquid interfaces
Runxi Wang a, b, Sheng Bi a, b, Volker Presser c, d, Guang Feng a, b, *
a
State Key Laboratory of Coal Combustion, School of Energy and Power Engineering, Huazhong University of Science and Technology (HUST), Wuhan
430074, China
b
Nano Interface Centre for Energy, School of Energy and Power Engineering, Huazhong University of Science and Technology, 430074, China
c
INM e Leibniz Institute for New Materials, 66123 Saarbrücken, Germany
d
Saarland University, 66123 Saarbrücken, Germany

a r t i c l e i n f o a b s t r a c t

Article history: Selecting the most suitable force field is a key to meaningful molecular dynamics (MD) simulation. To
Received 9 October 2017 select the appropriate gold force field to model the Au(111)/ionic liquid interface, a systematic com-
Received in revised form parison of four different widely used force fields of gold and a typical carbon force field has been studied
2 January 2018
by MD simulations with constant potential method. We calculated the ion adsorption behavior and
Accepted 22 January 2018
Available online 21 February 2018
differential capacitance of interfaces between the gold electrode and ionic liquid 1-butyl-1-
methylpyrrolidinium bis(trifluoromethylsulfonyl)imide ([PYR][TFSI]) in comparison with the experi-
mental results and showed the effects on the solid-liquid interfaces from the van der Waals interaction,
Keywords:
Au(111)/ionic liquid interface
image force effect and cumulative ions. Based on the comparison between the results of simulations and
Force fields experiments, we recommend two types of force fields to properly model the Au(111)/ionic liquid
Ion adsorption interfaces.
Differential capacitance © 2018 Published by Elsevier B.V.
Molecular dynamics

1. Introduction multilayered structure at electrified gold/RTILs interface and ion


arrangements change as a function of potential, which suggested
Room-temperature ionic liquids (RTILs) are promising electro- the RTILs interface can be shaped by the electrode potential and
lytes because of their remarkable advantages, such as low volatility, ions structure. Later, Mao et al. [11] obtained high quality AFM force
wide electrochemical windows, and designable properties [1e3]. curves to prove the existence of charged interior layers and neutral
Beyond that, they have been used in various applications including exterior layers. Gebbie et al. [12] employed a surface force appa-
energy storage, electrocatalysis, electrowetting, and electronic ratus (SFA) to investigate the electric screen of ionic liquids at
gating [4e7]. The interfacial structures of RTILs at solid-liquid in- charged gold and mica surfaces. Their results showed that both
terfaces determine the performance of such applications. There- Stern and diffuse layers are formed to screen the charged surfaces
fore, many studies have investigated fluid-solid interfaces with ILs. and the diffuse double layer consists of dissociated cations and
Among these, gold electrodes are commonly employed due to anions. In addition, neutron or X-ray reflectometry and surface
gold's high electrical conductivity, chemical stability, and compat- tunneling microscopy are further methods found in studies of solid/
ibility with conventional microfabrication procedures [8,9]. Rele- RTILs interfaces [13e17].
vant experimental studies have been carried out to investigate In addition to experimental research, there are many theoretical
gold/RTILs interfaces in order to reveal the mechanism of the works and simulations investigating gold/RTIL interfaces. Using
electric double layer (EDL) formation. For example, Atkin et al. [10] molecular dynamics (MD) simulations, Sha et al. [18] reported the
used atomic force microscopy (AFM) to reveal that there is a change of ion structure on gold electrode surface for different
applied voltages. They also calculated the interaction energy of
gold-ions and binding energy of cation-anion, which are difficult to
extract directly from experimental data. Ferreira et al. [19] used MD
* Corresponding author. State Key Laboratory of Coal Combustion, School of
simulations to investigate the orientation of ions on the solid sur-
Energy and Power Engineering, Huazhong University of Science and Technology
(HUST), Wuhan 430074, China. face and found the orientations of cations are more dependent on
E-mail address: gfeng@hust.edu.cn (G. Feng). the interface instead of the anions. Furthermore, Nikitina et al. [20]

https://doi.org/10.1016/j.fluid.2018.01.024
0378-3812/© 2018 Published by Elsevier B.V.
R. Wang et al. / Fluid Phase Equilibria 463 (2018) 106e113 107

combined MD simulations and quantum mechanical theory to electrode surfaces into account, for instance, a local redistribution
study the electron transfer reaction at the gold/RTIL interface. They of charges on the electrode surface due to interaction with RTILs.
found a higher energy barrier that impedes the charged reactants This approach has been verified by previous work [21,22], and it
from approaching the gold surface, which influences the electron was shown that this method is much more realistic comparing to
transfer. In these simulations, the accuracy of the force fields plays a assigning a uniform and constant partial on each electrode atom
crucial role in determining the amount of ions adsorption, molec- when simulating supercapacitors based on RTILs. The number of
ular orientation, energy barrier, and so on. However, so far most MD ion pairs was tuned to guarantee the electrolyte density in the
studies used different gold force fields to investigate the interface of center of channel was in accord with the bulk. Long-range Coulomb
gold/RTILs and the validation of the applicability of the force fields interactions were quantified by using the particle mesh Ewald
is usually ignored before modeling. Meanwhile, there are several (PME) method [23] with a correlation for slab-geometry [24]. An
kinds of force fields developed for gold, thus it is necessary to make FFT grid spacing of 0:1143_ nm and cubic interpolation for charge
systematic comparisons between them to reveal the suitable one distribution were used to compute the electrostatic interactions in
for modeling gold/RTILs interfaces. the reciprocal space. A cutoff length of 1.2 nm was used in the
In the present work, we used MD simulations to compare four calculation of electrostatic interactions in the real space. The non-
different types of often-used force fields of gold with a typical electrostatic interactions were computed by direct summation
carbon force field systematically, using the ionic liquid 1-butyl-1- with a cutoff length of 1.2 nm. LJ interactions between different
methylpyrrolidinium bis(trifluoromethylsulfonyl)imide ([PYR] atoms were calculated using Lorentz-Berthelot mixing rules. Sim-
[TFSI]) as the electrolyte. We investigated the electro-adsorption ulations were performed in the NVT ensemble using a customized
behavior and differential capacitance of Au(111)/RTILs interface MD code Gromacs [25], with a time step of 2 fs. The ensemble
and attributed the interfacial phenomena to van der Waals inter- temperature was maintained at 338 K using the Nose-Hoover
action, image force effects, and cumulative ions. thermostat. All configurations were first run by annealing from
800 K to 338 K for 20 ns, and then an 80 ns production run was
2. Simulation details performed for analysis.

2.1. Simulation methods 2.2. Force fields

The simulation unit cell is shown in Fig. 1. Ionic liquids were For the gold electrode surface, we choose four highly cited gold
enclosed between two electrodes with a size of 5.488  5.488 nm2. force fields that all used in modeling gold electrode. In this work,
Electrodes are modeled as Au (111) single-crystal surfaces and the form of the force fields adopts 12-6 Lennard-Jones potential,
graphene sheets. The distance between positive electrode and where rij is the distance and εij is the potential well describing the
negative electrode is 8.232 nm and the electrodes are fixed during interaction intensity between atom i and atom j.
the simulation, which should be wide enough to ensure that the 2 !12 !6 3
RTILs in the central portion of the system are bulk-like (see Fig. S1).  
4
sij sij 5
The expectation that [PYR][TFSI] in the central portion of the sys- VLJ rij ¼ 4εij 
rij rij
tem are bulk-like is corroborated by the observation that, beyond
positions ~3.0 nm from the electrode surface, the ionic space charge
density is practically zero (i.e., both cations and anions are
distributed homogeneously) and the electrical potential profile is
flat (see Fig. S2 and Fig. S3). Detailed number of the ionic liquids 2.2.1. Face-centered cubic metals (FCCM) force field
and electrode atoms are gathered in Table S1 for all studied sys- The face-centered cubic metals force field [26] was parameter-
tems. The equal potential was maintained on the electrode surfaces ized for simulating the surface and interface of face-centered cubic
and located at a plane across the center of surface atoms. This metals, such as Au, Ag, Al, and Cu. It was developed to simulate the
method was able to take the electronic polarizability of the interface between metals and organic, inorganic, and biological

Fig. 1. a) The schematic configuration of the simulation unit cell. b) All-atoms molecular structure of [PYR][TFSI].
108 R. Wang et al. / Fluid Phase Equilibria 463 (2018) 106e113

mixtures. The Lennard-Jones parameters were obtained by fitting Table 1


the density of metal, metal-vapor surface tension, metal-water Force fields parameters of electrodes.

interface tension, and elastic moduli to experimental results un- s (nm) ε (kJ/mol)
der ambient conditions. This force field is compatible with many FCCM 0.2629 22.1333
universal force fields (OPLS-AA, AMBER, CHARMM, COMPASS, CVFF, HPFF 0.2637 10.1800
and PCFF) to mimic the interface properties with water, bio- MaFF 0.3293 4.3932
polymers, organic molecules and inorganic components. Most of GolPa 0.3200 0.6500
Carbon 0.3400 0.3599
the existing work used this force field in metal/biopolymer inter-
a
face [27,28] to investigate molecular interactions, peptide adsorp- The LJ interaction is assigned to virtual sites in this force field. Because all
simulations are calculated by constant potential method, we do not consider the
tion and so on. Only a few studies have studied the EDL structure,
original polarization in this force field.
surface interaction, and capacitance when using gold as electrode
[18,19,29].
3. Results and discussion
2.2.2. Force field developed by Halicioǧlu and Pound (HPFF)
Halicioǧlu and Pound developed a simple method to calculate 3.1. Electric double-layer structure
12-6 Lennard-Jones potential parameters for metal, using crystal-
line state physical properties (such as cohesive energy, compress- To observe different ion electro-adsorption phenomena at
ibility, lattice constants, heat capacity, etc.) [30]. They found that Au(111)/[PYR][TFSI] interfaces when using four types of force fields,
the potential parameters are mainly dependent on DH (DH¼Hv-Hc, we calculated the density profiles of [PYR]þ and [TFSI]- ions as a
Hv, and Hc denote the total enthalpies of the vapor and the crys- function of distance from the electrode and the potential drop
talline states, respectively) and the volume of the crystal. Although across the EDL, shown in Figs. 2e3. The ion number density profiles
this force field is derived from experimental data of metal, several were calculated based on the location of the ion's center-of-mass as
previous works have used HPFF to investigate the diffusion of obtained from simulations. According to the electro-adsorption
adsorbed hydrocarbon species [31e33], and the dynamic behavior behavior, we divided the four force fields into two groups: strong
of atomic oxygen on metal surfaces [34]. So far, HPFF has been in- adsorption (FCCM and HPFF) and weak adsorption (MaFF and GolP).
tegrated into consistent Valence force field (CVFF) [35], which is a For strong adsorption, more cations can be adsorbed in the first
universal force field used by LAMPPS, Material Studio, etc. layer at neutral and negative potential electrodes. The peak value of
first layer is close to 80 #/nm3 at 3 V at 0.34 nm from the electrode
2.2.3. Force field developed by Maranas et al. (MaFF) surface, which is much higher than for the weak adsorption group.
The Force Field developed by Maranas et al. [36] utilizes the 12-6 At zero potential, the strong adsorption electrode also electro-
Lennard-Jones potential to describe the interaction of peptides and adsorbs more cations in the first layer. As potential becomes
water with gold surface. The size parameter s was set depending on increasingly positive when sweeping from 0 V, there is still a strong
the van der Waals radius of gold atom from the Universal Force adsorption of [PYR]þ. This observed strong adsorption of [PYR]þ at
Field [37]. Then they tuned the value of ε to make sure that water the positive potential surface is in agreement with the experi-
has a zero contact angle on gold surface. Finally, ε can be deter- mental phenomena deduced from neutron reflectometry [43]. In
mined by comparing the adsorption energy with ab initio result. line with observations for strong adsorption group, Lauw et al.
Similarly, this force field was mainly used for proteins, amino acids, experimentally found a cation [PYR]þ rich interface at a positively
and peptides, because most of such systems are water-based. charged Au(111) electrode, which indicates a strong adsorption of
[PYR]þ on the gold electrode. The similar strong adsorption of RTIL
2.2.4. GolP force field cations or anions on other metal surface has also been reported in
GolP is the abbreviation of Gold-Protein, which is designed for the literature [44,45]. However, the adsorption of [PYR]þ decreases
the interaction of proteins and the Au(111) surface, but not limited rapidly as the potential becomes more positive for the weak
to simulations of proteins [38]. Different from the above mentioned adsorption group.
three force fields, this force field introduced two Lennard-Jones For clearly comparing the different adsorption behavior, we also
virtual sites at the hollow site among gold atoms to drive the calculated the cation number density of [PYR][TFSI]/carbon surface,
adsorption of atoms to the correct site and added the image shown in Fig. 4. The number density of [PYR]þ in the first layer near
interaction. To confirm the reliability of the force field, quantum the carbon electrode shows the same pattern as can be seen for the
mechanical calculations of small molecule adsorption on gold weak adsorption group: the first layer has only one peak and the
surface were compared with experimental data. The GolP force peak value decreases significantly with the increase of positive
field has been proved that it can model the experimental phe- voltage. This finding implies that MaFF and GolP force fields may
nomenon in a physiological environment, but the interaction with not adequately reflect the strong adsorption of [PYR]þ at Au(111)
ionic liquids still needs further exploration. surface as revealed in experiment.
The force field of carbon was adapted from the work by Borodin Fig. 3 shows the number density of [TFSI]- as a function of the
et al. [39], which is widely used in modeling carbon/RTILs in- potential drop. The number of [TFSI]- in the first adsorbed layer
terfaces. The parameters of gold and carbon electrodes are listed in increases with the potential well (ε) of the force field. All four gold
Table 1. All studied force fields parameters are summarized in the force fields and carbon (Fig. 4b) present the same pattern, showing
FF files provided in Supporting Information. that anions adsorb at positive electrode and are rejected from the
A refined force field developed by Chaban et al. [40] from the negative electrode. This clearly shows that the interaction between
force field of Lopes et al. [41,42] was used for [PYR][TFSI], in which [TFSI]- and the electrode is weaker than for [PYR]þ. In addition,
they scaled partial charges on the IL so that the total charges on specific adsorption of [TFSI]- on the gold electrode is not observed
cation and anion are ± 0.8e, for the sake of refinement for more in our simulations (in contrast to [PYR]þ).
realistic reproduction of transport properties. This force field was
verified to show that it can accurately provide the thermodynamic 3.2. Van der Waals interaction
properties such as density and heat of vaporization (see Table S2
and Table S3). To understand the strong adsorption of [PYR]þ on the surface of
R. Wang et al. / Fluid Phase Equilibria 463 (2018) 106e113 109

Fig. 2. The number density of [PYR]þ as a function distance from the electrode surface for a series of potential across EDL (blue and red in floor color correspond to negative and
positive potential respectively). a) FCCM (s ¼ 0.2629 nm, ε ¼ 22.1333 kJ/mol), b) HPFF (s ¼ 0.2637 nm, ε ¼ 10.1800 kJ/mol), c) MaFF (s ¼ 0.3293 nm, ε ¼ 4.3932 kJ/mol), d) GolP
(s ¼ 0.3200 nm, ε ¼ 0.6500 kJ/mol). (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

Fig. 3. The number density of [TFSI]- as a function distance from the electrode surface for a series of potential across EDL (blue and red in floor color correspond to negative and
positive potential, respectively). a) FCCM (s ¼ 0.2629 nm, ε ¼ 22.1333 kJ/mol), b) HPFF (s ¼ 0.2637 nm, ε ¼ 10.1800 kJ/mol), c) MaFF (s ¼ 0.3293 nm, ε ¼ 4.3932 kJ/mol), d) GolP
(s ¼ 0.3200 nm, ε ¼ 0.6500 kJ/mol). (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

the gold electrode, we first investigated the van der Waals inter- the strong adsorption group could give stronger interactions with
action. As shown in Fig. 5a, the Lennard-Jones (LJ) potential be- cations; (2) the location of the potential well of the strong
tween the electrode and the cation [PYR]þ has been calculated for adsorption group is smaller than for the weak adsorption group;
all five force fields, including the one for carbon. There are two therefore, the first layer of cation is in closer proximity to the sur-
distinct observations: (1) the potential well of the strong adsorp- face. To eliminate the effect of the applied voltage (note that we do
tion group is lower than for weak adsorption group, which suggests consider the image force effect due to the electronic polarizability
110 R. Wang et al. / Fluid Phase Equilibria 463 (2018) 106e113

Fig. 4. The ions number density as a function distance from the carbon electrode surface for a series of potential across EDL (blue and red in floor color correspond to negative and
positive potential, respectively). a) [PYR]þ distribution, b) [TFSI]- distribution. (For interpretation of the references to color in this figure legend, the reader is referred to the Web
version of this article.)

Fig. 5. a) The LJ potential between electrode and carbon atom of [PYR]þ. b) The number density of [PYR]þ at applied potential of 0 V. To facilitate reading, the parameters of the force
fields are as follows. FCCM (s ¼ 0.2629 nm, ε ¼ 22.1333 kJ/mol), HPFF (s ¼ 0.2637 nm, ε ¼ 10.1800 kJ/mol), MaFF (s ¼ 0.3293 nm, ε ¼ 4.3932 kJ/mol), GolP (s ¼ 0.3200 nm,
ε ¼ 0.6500 kJ/mol), Carbon (s ¼ 0.3400 nm, ε ¼ 0.3599 kJ/mol).

of the electrode surfaces), we compared the cation number density in Fig. 6c and the surface charge distribution in Fig. 6d shows that
of the five force fields (including carbon) at 0 V in Fig. 5b. Consistent only few charges have been induced by the ionic liquids. This
with the previous data, more cations are electro-adsorbed at the change of the surface charge distribution is mainly due to the
electrode surface and the location of first peak is closer to the different adsorption of ionic liquids in the first layer. As mentioned
electrode surface. For the same reason, the strong adsorption group before, a strong adsorption electrode electro-adsorbs more ions
yields in the simulation for gold surfaces more electro-adsorbed and attract them closer to surface; thereby, stronger opposite
anions. Therefore, the LJ potential mainly determines the adsorp- charges are induced which results in lager image forces to electro-
tion behavior of ions, where the larger value of ε could drive more adsorb more ions in the first ionic layer.
ions to be adsorbed on the gold surface; the value of s relates the The specific adsorption of cations can be attributed to the
location of the first ion layer. combined effect of van der Waals interaction and the image force
effect. Van der Waals interaction plays a major role in the adsorp-
3.3. Image force effect tion behavior: a larger value of ε and smaller value of s cause that
more ions are electro-adsorbed closer to the electrode surface. The
Without applied voltage, the ion adsorption behavior is not only strong adsorption would induce more image charge at electrode
determined by the LJ potential but also affected by the image forces surface, which makes the interaction between ions and electrode
between ions and the electrode. To show the image force effect even stronger. Although this adsorption phenomenon was pre-
clearly, the 2D surface charge distribution has been illustrated in dicted in many theoretical studies, the image force is mostly
Fig. 6 (note that in MD modeling work with constant charge ignored in most previous simulation studies due to the heavy
method, there would be no results like such 2D surface charge computational cost. The deep effect of image forces still requires
distribution, since the constant charge method assume the charge further investigations in the future works. Interestingly, we found
on surface is zero everywhere at zero applied potential). As can be that there is no specific adsorption of anions, probably because of
seen, the charge distribution pattern changes when using different the geometry of anion different from that of cation: the cation
force fields. In Fig. 6aeb, most of the area is deep red or blue, which possesses a ring structure and a long alkyl chain, which would
suggests that the ions near the surface induced opposite charge. facilitate the cation closely to pack on the flat gold surface and then
Comparatively, the area and value of the induced charge are smaller generate a high attraction to the electrode.
R. Wang et al. / Fluid Phase Equilibria 463 (2018) 106e113 111

Fig. 6. 2D charge density distribution on gold surface at zero applied potential (unit: C/m2). a) FCCM (s ¼ 0.2629 nm, ε ¼ 22.1333 kJ/mol), b) HPFF (s ¼ 0.2637 nm, ε ¼ 10.1800 kJ/
mol), c) MaFF (s ¼ 0.3293 nm, ε ¼ 4.3932 kJ/mol), d) GolP (s ¼ 0.3200 nm, ε ¼ 0.6500 kJ/mol).

3.4. Differential capacitance liquids; therefore, we have calculated the cumulative number of
ions for all four different gold force fields in first EDLs layer as a
For better comparison, we also show the differential capacitance function of potential drop (Fig. 7b). Because of the strong adsorp-
of carbon in Fig. 7a. The differential capacitance (DC) curves in tion, the change of cumulative cations of FCCM force field is less
Fig. 7a were obtained from a local fit of surface charges and the EDL obvious compared to other force fields. The increase of net charge
potential drop (see Supporting Information for details). The dif- (i.e., the number of cations minuses that of anions) for FCCM from
ferential capacitance curves obtained by using FCCM and HPFF 0.5 V to 2.0 V is larger than that for weak adsorption group.
force fields have a characteristic camel shape and when using MaFF, Beyond 0.5 V, the variation changes slightly for FCCM. Thus, the
GolP, or carbon, we see a curve with bell shape. Lauw and Lockett differential capacitance curve of FCCM has larger amplitude than
et al. [43,46] measured the differential capacitance of [PYR][TFSI] at what we found for the weak adsorption group at negative potential
Au(111) surface by impedance spectroscopy and confirmed for this and become less defined at positive potential. For the HPFF force
system a camel-like shaped curve. In the latter work, the relatively field, the variation of cations and anions are steeper than weak
large ion sizes and nonhomogeneous ion polarization at the inter- adsorption group within the voltage range. Such a tendency in the
face were correlated with the camel shaped differential capacitance change of ions number leads to a higher capacitance of strong
curve. The camel-shaped differential capacitance has been theo- adsorption group compared with that of weak adsorption group.
retically predicted in the literature [47,48]. The results of strong adsorption group are qualitatively consistent
Overall, the differential capacitance curves of strong adsorption with experimental findings, which indicate that these force fields
electrodes have a larger amplitude than that of weak adsorption are more suitable for modeling the realistic interaction between
group and carbon; however, the peak value of weak adsorption ionic liquids and gold electrode surface. For a meaningful quanti-
group and carbon are similar. As reported in the literature [49,50], tative comparison, the simulations and experiments should be
the capacitance observed for gold electrodes is larger compared to done in a way where conditions should be identical, such as tem-
carbon surfaces when using the same ionic liquid. This aligns with perature, electrode structure, impurities and so on.
the issue that the weak adsorption group force field could not
precisely reproduce the differences between gold and carbon 4. Conclusion
electrodes. Previous studies [51] showed that the differential
capacitance is mainly determined by the electro-adsorption of ionic We systematically compared the accuracy of four different gold
112 R. Wang et al. / Fluid Phase Equilibria 463 (2018) 106e113

Fig. 7. a) Differential capacitance as a function of the EDL potential. b) The cumulative number of cations (solid lines) and anions (dashed lines). To facilitate reading, the parameters
of the force fields are as follows. FCCM (s ¼ 0.2629 nm, ε ¼ 22.1333 kJ/mol), HPFF (s ¼ 0.2637 nm, ε ¼ 10.1800 kJ/mol), MaFF (s ¼ 0.3293 nm, ε ¼ 4.3932 kJ/mol), GolP (s ¼ 0.3200 nm,
ε ¼ 0.6500 kJ/mol), Carbon (s ¼ 0.3400 nm, ε ¼ 0.3599 kJ/mol).

force fields when modeling Au(111)/RTILs interfaces at various were carried out at the National Supercomputing Center in Tianjin
electric potentials with MD simulations. The constant potential (Tianhe-1A) and Guangzhou (Tianhe II).
condition was imposed at the locations of all electrode atoms by
allowing the fluctuations of charges on the electrode surface. The Appendix A. Supplementary data
specific adsorption of cations was found at the strong adsorption
group of force fields even when the potential becomes positive, Supplementary data related to this article can be found at
which is in agreement with previous experimental work [43]. This https://doi.org/10.1016/j.fluid.2018.01.024.
specific adsorption results from stronger van der Waals interaction
and the image force effect. Moreover, MD simulations reveal that
References
FCCM and HPFF can result in a camel-shape of the calculated dif-
ferential capacitance pattern of Au/[PYR][TFSI] interfaces, while [1] M.V. Fedorov, A.A. Kornyshev, Ionic liquids at electrified interfaces, Chem. Rev.
MaFF and HPFF yield a bell shape. These curve shapes are mainly 114 (2014) 2978e3036.
due to the different adsorption behaviors of ions at electrode sur- [2] H. Niedermeyer, J.P. Hallett, I.J. Villar-Garcia, P.A. Hunt, T. Welton, Mixtures of
ionic liquids, Chem. Soc. Rev. 41 (2012) 7780e7802.
faces. Further observations show that the strong adsorption group [3] M. Galin  ski, A. Lewandowski, I. Ste˛ pniak, Ionic liquids as electrolytes, Elec-
of force fields is more suitable for modeling the realistic interaction trochim. Acta 51 (2006) 5567e5580.
between ionic liquids and gold electrode surface. However, the [4] F. Endres, Ionic liquids: solvents for the electrodeposition of metals and
semiconductors, ChemPhysChem: Eur. J. Chem. Phys. Phys. Chem. 3 (2002)
strongest adsorption of FCCM force field may complicate the
144e154.
simulation because more ions are crowed in the limited interface [5] S. Millefiorini, A.H. Tkaczyk, R. Sedev, J. Efthimiadis, J. Ralston, Electrowetting
space, which leads to declining diffusion of ions so that it requires of ionic liquids, J. Am. Chem. Soc. 128 (2006) 3098e3101.
[6] F. Chen, Q. Qing, J. Xia, J. Li, N. Tao, Electrochemical gate-controlled charge
much more time to reach the equilibrium and longer trajectories
transport in graphene in ionic liquid and aqueous solution, J. Am. Chem. Soc.
are needed for statistical analysis. Weak adsorption group of force 131 (2009) 9908e9909.
fields (MaFF and HPFF) are more suitable for simulating the inter- [7] A. Balducci, U. Bardi, S. Caporali, M. Mastragostino, F. Soavi, Ionic liquids for
face between biological molecules and gold in aqueous solution, hybrid supercapacitors, Electrochem. Commun. 6 (2004) 566e570.
[8] M.M. Collinson, Nanoporous gold electrodes and their applications in
since both force fields considered extra interaction with water analytical chemistry, ISRN Anal. Chem. 2013 (2013).
during the development process. MaFF used a modified TIP3P [9] R. Zhang, H. Olin, Porous gold filmsda short review on recent progress, Ma-
model [52] and GolP added extra van der Waals interaction be- terials 7 (2014) 3834e3854.
[10] R. Hayes, N. Borisenko, M.K. Tam, P.C. Howlett, F. Endres, R. Atkin, Double
tween gold and water molecules. Since water molecules occupy layer structure of ionic liquids at the Au(111) electrode interface: an atomic
most area of gold surface, weak adsorption group force fields are force microscopy investigation, J. Phys. Chem. C 115 (2011) 6855e6863.
preferred for biological simulation system; however, they may not [11] X. Zhang, Y.X. Zhong, J.W. Yan, Y.Z. Su, M. Zhang, B.W. Mao, Probing double
layer structures of Au (111)-BMIPF6 ionic liquid interfaces from potential-
be suitable for Au(111)/RTILs interfaces. This study provides dependent AFM force curves, Chem. Commun. (Camb) 48 (2012) 582e584.
important information for researchers to choose suitable force [12] M.A. Gebbie, M. Valtiner, X. Banquy, E.T. Fox, W.A. Henderson,
fields from existing gold force fields to study the gold/RTILs J.N. Israelachvili, Ionic liquids behave as dilute electrolyte solutions, Proc. Natl.
Acad. Sci. U.S.A. 110 (2013) 9674e9679.
interface.
[13] R. Steitz, V. Leiner, R. Siebrecht, R.v. Klitzing, Influence of the ionic strength on
the structure of polyelectrolyte films at the solid/liquid interface, Colloid.
Surface. Physicochem. Eng. Aspect. 163 (2000) 63e70.
Acknowledgments
[14] T. Albrecht, K. Moth-Poulsen, J.B. Christensen, J. Hjelm, T. Bjørnholm,
J. Ulstrup, Scanning tunneling spectroscopy in an ionic liquid, J. Am. Chem.
This work was supported by the National Natural Science Soc. 128 (2006) 6574e6575.
Foundation of China (51406060) and the Fundamental Research [15] E. Sloutskin, B.M. Ocko, L. Tamam, I. Kuzmenko, T. Gog, M. Deutsch, Surface
layering in ionic liquids: an X-ray reflectivity study, J. Am. Chem. Soc. 127
Funds for the Central Universities (2016YXZD006). V.P. thanks Prof. (2005) 7796e7804.
Eduard Arzt (INM) for his continued support. All the computations [16] U. Doman  ska, K. Paduszyn  ski, (Solidþ liquid) and (liquidþ liquid) phase
R. Wang et al. / Fluid Phase Equilibria 463 (2018) 106e113 113

equilibria measurements and correlation of the binary systems {tri-iso-butyl [34] S.M. Levine, S.H. Garofalini, Molecular dynamics simulation of atomic O on
(methyl) phosphonium tosylateþ alcohol, orþ hydrocarbon}, Fluid Phase Pt(111), Surf. Sci. 167 (2015) 198e206.
Equil. 278 (2009) 90e96. [35] P. Dauberosguthorpe, V.A. Roberts, D.J. Osguthorpe, J. Wolff, M. Genest,
[17] P. Paricaud, A. Galindo, G. Jackson, Recent advances in the use of the SAFT A.T. Hagler, Structure and energetics of ligand binding to proteins: Escherichia
approach in describing electrolytes, interfaces, liquid crystals and polymers, coli dihydrofolate reductase-trimethoprim, a drug-receptor system, Proteins
Fluid Phase Equil. 194 (2002) 87e96. Struct. Funct. Bioinf. 4 (1988) 31e47.
[18] M. Sha, Q. Dou, F. Luo, G. Zhu, G. Wu, Molecular insights into the electric [36] A.V. Verde, J.M. Acres, J.K. Maranas, Investigating the specificity of peptide
double layers of ionic liquids on Au (100) electrodes, ACS Appl. Mater. In- adsorption on gold using molecular dynamics simulations, Bio-
terfaces 6 (2014) 12556e12565. macromolecules 10 (2009) 2118e2128.
[19] E.S. Ferreira, C.M. Pereira, M.N.l.D. Cordeiro, D.J. dos Santos, Molecular dy- [37] A.K. Rappe , C.J. Casewit, K. Colwell, W. Goddard Iii, W. Skiff, UFF, a full periodic
namics study of the gold/ionic liquids interface, J. Phys. Chem. B 119 (2015) table force field for molecular mechanics and molecular dynamics simula-
9883e9892. tions, J. Am. Chem. Soc. 114 (1992) 10024e10035.
[20] V.A. Nikitina, S.A. Kislenko, R.R. Nazmutdinov, M.D. Bronshtein, G.A. Tsirlina, [38] F. Iori, R. Di Felice, E. Molinari, S. Corni, GolP: an atomistic force-field to
Ferrocene/ferrocenium redox couple at Au (111)/ionic liquid and Au (111)/ describe the interaction of proteins with Au (111) surfaces in water,
acetonitrile interfaces: a molecular-level view at the elementary act, J. Phys. J. Comput. Chem. 30 (2009) 1465e1476.
Chem. C 118 (2014) 6151e6164. [39] O. Borodin, Polarizable force field development and molecular dynamics
[21] C. Merlet, C. Pean, B. Rotenberg, P.A. Madden, P. Simon, M. Salanne, Simulating simulations of ionic liquids, J. Phys. Chem. B 113 (2009) 11463e11478.
supercapacitors: can we model electrodes as constant charge surfaces? [40] V.V. Chaban, I.V. Voroshylova, Systematic refinement of canongia
J. Phys. Chem. Lett. 4 (2013) 264e268. LopeseP adua force field for pyrrolidinium-based ionic liquids, J. Phys. Chem. B
[22] C. Pe an, C. Merlet, B. Rotenberg, P.A. Madden, P.-L. Taberna, B. Daffos, 119 (2015) 6242e6249.
M. Salanne, P. Simon, On the dynamics of charging in nanoporous carbon- [41] J.N. Canongia Lopes, J. Deschamps, A.A.H. P adua, Modeling ionic liquids using a
based supercapacitors, ACS Nano 8 (2014) 1576e1583. systematic all-atom force field, J. Phys. Chem. B 108 (2004), 11250e11250.
[23] U. Essmann, L. Perera, M.L. Berkowitz, T. Darden, H. Lee, L.G. Pedersen, [42] J.N. Canongia Lopes, A.A.H. P adua, Molecular force field for ionic liquids
A smooth particle mesh Ewald method, J. Chem. Phys. 103 (1995) 8577e8593. composed of triflate or bistriflylimide anions, J. Phys. Chem. B 108 (2004)
[24] I.-C. Yeh, M.L. Berkowitz, Ewald summation for systems with slab geometry, 16893e16898.
J. Chem. Phys. 111 (1999) 3155e3162. [43] Y. Lauw, M.D. Horne, T. Rodopoulos, V. Lockett, B. Akgun, W.A. Hamilton,
[25] B. Hess, C. Kutzner, D. Van Der Spoel, E. Lindahl, GROMACS 4: algorithms for A.R. Nelson, Structure of [C4mpyr][NTf2] room-temperature ionic liquid at
highly efficient, load-balanced, and scalable molecular simulation, J. Chem. charged gold interfaces, Langmuir: ACS J. surface. colloids 28 (2012)
Theor. Comput. 4 (2008) 435e447. 7374e7381.
[26] H. Heinz, R. Vaia, B. Farmer, R. Naik, Accurate simulation of surfaces and in- [44] C. Aliaga, S. Baldelli, Sum frequency generation spectroscopy and double-layer
terfaces of face-centered cubic metals using 12 6 and 9 6 Lennard-Jones capacitance studies of the 1-butyl-3-methylimidazolium Dicyanamide
potentials, J. Phys. Chem. C 112 (2008) 17281e17290. platinum interface, J. Phys. Chem. B 110 (2006) 18481e18491.
[27] H. Heinz, B.L. Farmer, R.B. Pandey, J.M. Slocik, S.S. Patnaik, R. Pachter, R.R. Naik, [45] M.T. Alam, M.M. Islam, T. Okajima, T. Ohsaka, Measurements of differential
Nature of molecular interactions of peptides with gold, palladium, and Pd Au capacitance at mercury/room-temperature ionic liquids interfaces, J. Phys.
bimetal surfaces in aqueous solution, J. Am. Chem. Soc. 131 (2009) Chem. C 111 (2007) 18326e18333.
9704e9714. [46] V. Lockett, M. Horne, R. Sedev, T. Rodopoulos, J. Ralston, Differential capaci-
[28] J. Feng, R.B. Pandey, R.J. Berry, B.L. Farmer, R.R. Naik, H. Heinz, Adsorption tance of the double layer at the electrode/ionic liquids interface, Phys. Chem.
mechanism of single amino acid and surfactant molecules to Au {111} sur- Chem. Phys. 12 (2010) 12499e12512.
faces in aqueous solution: design rules for metal-binding molecules, Soft [47] A.A. Kornyshev, Double-layer in ionic liquids: paradigm change? J. Phys.
Matter 7 (2011) 2113e2120. Chem. B 111 (2007) 5545e5557.
[29] K.C. Jha, H. Liu, M.R. Bockstaller, H. Heinz, Facet recognition and molecular [48] Y. Lauw, M. Horne, T. Rodopoulos, F. Leermakers, Room-temperature ionic
ordering of ionic liquids on metal surfaces, J. Phys. Chem. C 117 (2013) liquids: excluded volume and ion polarizability effects in the electrical
25969e25981. double-layer structure and capacitance, Phys. Rev. Lett. 103 (2009) 117801.
[30] T. Halicio, G.M. Pound, Calculation of potential energy parameters form [49] M.T. Alam, M.M. Islam, T. Okajima, T. Ohsaka, Capacitance measurements in a
crystalline state properties, Phys. Status Solidi 30 (1975) 619e623. series of room-temperature ionic liquids at glassy carbon and gold electrode
[31] H. Tamura, M. Yoshida, K. Kusakabe, C. Youngmo, R. Miura, M. Kubo, interfaces, J. Phys. Chem. C 112 (2008) 16600e16608.
K. Teraishi, A. Abhijit Chatterjee, A. Miyamoto, Molecular dynamics simulation [50] V. Lockett, M. Horne, R. Sedev, T. Rodopoulos, J. Ralston, Differential capaci-
of friction of hydrocarbon thin films, Langmuir: ACS J. surface. colloids 15 tance of the double layer at the electrode/ionic liquids interface, Phys. Chem.
(1999) 7816e7821. Chem. Phys.: PCCP 12 (2010) 12499e12512.
[32] D. Huang, Y. Chen, K.A. Fichthorn, A molecular-dynamics simulation study of [51] G. Feng, J. Zhang, R. Qiao, Microstructure and capacitance of the electrical
the adsorption and diffusion dynamics of short n-alkanes on Pt(111), Chem- double layers at the interface of ionic liquids and planar electrodes, J. Phys.
Inform 26 (1995) 11021e11030. Chem. C 113 (2009) 4549e4559.
[33] K.A. Fichthorn, P.G. Balan, Y. Chen, Simulation and analysis of the motion of n- [52] E. Neria, S. Fischer, M. Karplus, Simulation of activation free energies in mo-
butane on Pt(111), Surf. Sci. 317 (1994) 37e44. lecular systems, J. Chem. Phys. 105 (1996) 1902e1921.

View publication stats

You might also like