Alkyl Methyl Imidazolium Based Ionic Liquids at Au(111) Surface

Hybrid inorganic-organic systems (HIOS) are used for various applications such as catalysis, energy storage and conversion, photovoltaics, molecular s...
0 downloads 0 Views 2MB Size
Subscriber access provided by Bethel University

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Alkyl Methyl Imidazolium Based Ionic Liquids at Au(111) Surface: Anions and Alkyl Chain Cations Induced Interfacial Effects Shanmugasundaram Kamalakannan, Muthuramalingam Prakash, Muneerah Mogren Al-Mogren, Gilberte Chambaud, and Majdi Hochlaf J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.9b03242 • Publication Date (Web): 31 May 2019 Downloaded from http://pubs.acs.org on May 31, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Alkyl Methyl Imidazolium Based Ionic Liquids at Au(111) Surface: Anions and Alkyl Chain Cations Induced Interfacial Effects Shanmugasundaram Kamalakannan,† Muthuramalingam Prakash,*,† Muneerah Mogren Al-Mogren,$ Gilberte Chambaud,‡ Majdi Hochlaf*,‡ Department of Chemistry, SRM Institute of Science and Technology, Kattankulathur 603203, Tamil Nadu, India $ Chemistry Department, Faculty of Science, King Saud University, PO Box 2455, Riyadh 11451, Kingdom of Saudi Arabia ‡ Université Paris-Est, Laboratoire Modélisation et Simulation Multi Echelle, MSME UMR 8208 CNRS, †

77454 Marne la Vallée Cedex 2, France

*Corresponding Authors: M.P.: Department of Chemistry, SRM Institute of Science and Technology, Kattankulathur 603203, Tamil Nadu, India; Phone: +91 44 2741 7686. Email: M.P.: [email protected]; [email protected]; M.H.: [email protected].

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract The structure, stability and adsorption mechanism on Au(111) surface of hydrophilic/hydrophobic ionic liquids (ILs) with various length of alkyl chain group in cation were investigated using first principle approaches including electrostatic potential (ESP), electron density analysis (EDA) and dispersion corrected density functional theory (DFT-D3) methods. A suitable selection of ILs is considered. Indeed, we treat ILs of general formula [CnMIm]+[X]-, where X = Cl-, PF6- and TFSA-, MIm = methylimidazolium and where we varied the alkyl chain (Cn) length for n = 0, 2, 4, 6, 8 and 10. We found that the adsorption energies (Eads) of fluorinated ILs are lower (by ~30%) than those of non-fluorinated ILs. Computations show that the nature of the anion within the ILs and the alkyl chain length play a very important role to alter the interfacial interactions between ILs and gold surface. Indeed, the alkyl groups of the cation affect the cation – anion interaction strength within ILs due to specific modes of adsorption of alkyl chains on the Au(111) surface. Also, structural induced adsorption effects are observed during the variation of alkyl chain and anions. These are due to the modification of the local properties of the gold surface upon IL adsorption, where we identify, in some specific cases, an enhancement of the charge transfers between the ILs and the Au(111) surface through van der Waals and dispersive interactions.

2 ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1.

Introduction Hybrid inorganic-organic systems (HIOS) are used for various applications such as catalysis,

energy storage and conversion, photovoltaics, molecular sensors, switches, skeletal tissues and biominerals, drug delivery systems, catalysts, sensors, and polymer nanocomposites.1–11 We refer to the recent review by Heinz and Ramezani-Dakhel 7 for a detailed presentation of examples ranging from biomolecules on nanostructured metals up to building materials. Particularly, molecular adsorption with coinage metals are receiving a widespread attention. For instance, small molecules adsorbed on gold surface devices are important for the development of novel functional materials for optical, photovoltaic and electrode applications. To this end, the study of the adsorption mechanism and reactivity of molecules at different surfaces and solid-liquid interfaces are required. Several experimental and theoretical techniques have been devoted to understand the adsorption and reaction mechanisms of organic molecules interacting with a metal surface.1,12–16 For instance, Koch et al. investigated the adsorption of organic compounds (pentacene (PEN) and perfluoropentacene (PFP)) on Cu(111) surface using photoelectron spectroscopy, X-ray standing wave (XSW), and scanning tunneling microscopy. They also complemented their work by theoretical modelling using DFT (PBE) based method. They pointed out that PEN is closer to the Cu(111) surface than PFP due to weak intramolecular dipole interactions between PFP and the Cu(111) surface.1 Recent theoretical reports stated that N-based nucleophiles and aromatic π-systems form a Regium bond with coinage metals.17,18 These studies revealed that gold complexes present a stronger binding with N-based nucleophiles than the other metals. Let’s cite also the work of Tkatchenko and co-workers dealing with the electronic charge rearrangement at the metal/organic interfaces using PBE and PBE+vdW methods.19 These authors found that there are three plausible mechanisms for the charge rearrangement: (i) the Pauli pushback; (ii) charge transfer between the surface and the monolayer; and (iii) the intrinsic electronic dipole of the monolayer. The induced van der Waals (vdWs) interactions enhance the Pauli pushback effect and at the same time reduces the charger transfer at the interface. Metal surface - ionic liquids (ILs) based interface materials are used for energy storage applications.20–22 ILs are liquid at temperature below 100°C. ILs have an advantage over conventional organic solvents due to their wide electrochemical window, great tenability, high charge density, high thermal stability, and low volatility. The structural orientation and the nature of cations and anions of ILs define the specific capacitance of the supercapacitor and of various energy storage materials.23 All these properties make them well-recognized materials for electrolytes in batteries and supercapacitors and as catalytic conversion materials.24 Compared to conventional liquids, ILs have well structural order at the interface

9,25,

due to their

vdWs, H-bonding and coulombic interactions 26 either mutually or with the metal surface (e.g., Fe, Ag and 3 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Cu). For instance, Heinz and co-workers studied the molecular ordering of ILs at metal surfaces using DFT and molecular dynamics (MD) simulations. They investigated the structure and energetics of the selfassembly of 1-ethyl-3-methylimidazolium ethyl sulfate ([EMIM][ES]) on the crystallographic {111}, {100}, and {110} facets of gold. Single ion pairs to multilayers were considered. These authors showed that the adsorption mechanism is controlled by the interplay of soft epitaxy, ionic interactions, induced charges, and steric effects in connection with the geometry of the cation and the anion within the IL pair.14 Using surface force apparatus and atomic force microscopy (AFM), Lui et al. characterized electrical double layer (EDL) structures of ILs at metal-ILs interface. They found that increasing the size of the nonhalogen anion leads to decreasing the columbic, H-bond and interaction strength of ILs at the interface. The EDL formation is largely influenced by H-bonding, steric effects, and molecular interactions.27 More generally, these interfacial interactions decide on the performance of the electrochemical systems.28 In interfacial molecule-substrate systems, interactions occur through various kinds of covalent and noncovalent bonding.29,30 At the microscopic level, noncovalent interactions are very difficult to describe both experimentally and theoretically. Recently, we used first principle methodologies to treat the adsorption mechanism of various hydrophilic and hydrophobic ILs (1-butyl 3-methylimidazolium [BMIm]+[X]-, where X = Cl-, DCA, HCOO-, BF4-, PF6-, CH3SO3-, OTF- and TFSA-) interacting with Au(111) surface.31 A good agreement was found with experimental microscopic and macroscopic observations, hence validating our theoretical approach. Here, we perform a theoretical study of hydrophilic/hydrophobic fluorinated and non-fluorinated ILs interacting with Au(111). These ILs have the general formula [CnMIm]+[X]-, where X = Cl-, PF6- and TFSA-, MIm = methylimidazolium and where we varied the alkyl chain length (i.e. Cn) for n = 0, 2, 4, 6, 8, 10. Earlier experimental works showed that the alkyl chain length should affect the properties of this liquidsolid model interface, such as viscosity, surface tension, and conductivity.32–38 For instance, the work of Hu et al. dealing with various alkyl chain substituted anthraquinone adsorbed on the highly oriented pyrolytic graphite (HOPG)-interface revealed that the Eads increases by lengthening the alkyl chain due to strong vdWs interactions between the adsorbate-substrate and also from the H-bonds between the anthraquinone cores.34 In 2018, Chinwangso used X-ray photoelectron spectroscopy (XPS) to study a series of unsymmetrical spiroalkanedithiols adsorbed at a gold surface by increasing the methylene units. Similar effect was observed because of the enhancing of the alkyl chains dispersion interaction with the metal surface.35 Prior to that, Li and co-workers 36 established the influence of the alkyl chain length and anion species on the interfacial nanostructure of ILs at Au(111). They showed that the 1-ethyl-3methylimidazolium [EMIm]+ and 1-hexyl-3-methylimidazolium [HMIm]+ present enhanced interfacial structures than the 1-butyl-3-methylimidazolium [BMIm]+ since [EMIm]+ adsorbs strongly through 4 ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

cation···π stacking with a parallel carbon chain to the metallic surface and since [HMIm]+ has longer alkyl chain tails. Nevertheless, these adsorption induced effects by the alkyl chains of the cation of ILs on gold surface are still not understood. Besides, Pensado and Padua have studied the molecular interactions, solvation and ordering of ILs around ruthenium nanoparticles using MD simulations. The radial and spatial distribution function analyses reveal that nonpolar groups and side chains were preferentially directed away from surface. Also, they concluded that the nanoparticle stabilization does not depend on the electrostatic double layer and steric effect of the alkyl chain in the cations.39 Our work aims to understand the effects induced when an alkyl chain based ILs are close to a gold surface. To the best of our knowledge, this is the first theoretical study to interpret the role of alkyl chain lengths and of anions at the Au(111) surface. We have also studied the orientation and interactions between the cations and anions at this interface. Our study should be useful for designing suitable ILs-gold interfacial materials, and for improving the differential capacitance of the composite materials devoted for energy storage applications. 2.

Computational Details We performed computations of fluorinated and non-fluorinated ILs ([CnMIm]+[X]- (where n = 0,

2, 4, 6, 8, 10; X= Cl-, PF6- and TFSA-) interacting with an Au(111) surface. These electronic computations were done using Perdew−Burke−Ernzerhof (PBE) generalized gradient approximation (GGA) density functional theory (DFT) to define the exchange−correlation 40 with and without considering Grimme’s latest version of empirical correction term (DFT-D3) to study the effect of dispersion correction.40–42 This is mandatory to describe the interaction of the organic molecule at the metal surface and carbon based composite materials.43 Two types of computations are done: (i) “gas phase type computations” where isolated ILs are treated; and (ii) “liquid-solid periodic computations” to model the ILs-gold interface. The comparison of both sets of data should help for understanding the gold surface ILs adsorption induced energetic and structural effects. The structures of isolated ILs were optimized using the GAUSSIAN 16 and GaussView 6.0 software packages.44 All atoms in ILs were described with the 6-311++G** basis set.45 These geometry optimizations were done without constrains at the PBE(+D3) level, followed by frequency calculations to confirm that they correspond to a minimum on their potential energy surfaces. The binding energies (BEs) were deduced and corrected for the basis set superposition error (BSSE) employing the counterpoise (CP) procedure of Boys and Bernardi.46 BE([CnMIm]+[X]-)=(E[CnMIm]+[X]- - (E[X]-+E[CnMIm]+)

(1)

where E[CnMIm]+[X]-, E[X]-, and E[CnMIm]+ are the total energies of the [CnMIm]+[X]- IL complex, anion, and cation as computed in the full basis set of the IL complex. 5 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 32

The gas phase optimized geometries were used as starting point for the [CnMIm]+@Au(111) and [CnMIm]+[X]-@Au(111) bulk phase periodic computations. We used the CP2K/Quickstep code for bulk phase optimization.47 A hybrid basis set formalism known as a Gaussian and Plane Wave (GPW)

48

is

implemented in CP2K, where the Kohn−Sham orbitals are expanded in terms of contracted Gaussian type orbitals, whereas an auxiliary plane wave basis set is used to expand the electronic charge density. Indeed, valence electron density and pseudopotentials are expressed by a mixed GPW basis set scheme.49 All atoms, except Au, were described by the optimized MOLOPT-TZVP basis set 50, whereas we used the DZVPMOLOPT-SR-GTH (SR denotes a short range) basis set for Au atoms. A plane wave cutoff (400 Ry) was included. Orbital transformation was also performed.51 In this study, atomic structures of the Au(111) surface were taken from previously resolved global minimum structures.52 The slab consisted of three layers each has 48 Au atoms. Throughout the study, we fixed the bottom layer, and the upper two layers were optimized freely. Afterward, we optimized the structure of [CnMIm]+@Au(111) and [CnMIm]+[X]@Au(111) composite materials. The adsorption energies, Eads, of [CnMIm]+ and of [CnMIm]+[X]- on the Au(111) surface were evaluated using the following equation Eads(Y@Au(111)) = E[Complex] – (EAu(111)+EY)

(2)

where E[Complex], EAu(111) and EY are the total energies of Y@Au(111), of EAu(111) and of Y (= [CnMIm]+ or [CnMIm]+[X]- (where n = 0, 2, 4, 6, 8, 10)), respectively. For rationalization, we generated a cube density file from the optimized coordinates at the PBE+D3/TZVP level to discuss the charge distribution within the ILs and upon adsorption of [CnMIm]+ or [CnMIm]+[X]- species on the gold surface (cf. Supporting Information Figures S1-S8). We also performed a Löwdin charge transfer analysis of the complexes interacting with Au(111). Löwdin charge analysis shows the effect of the involvement of the anions within the complex and the possible enhancement of the charge transfer process in the local environment. The charge density difference is expressed by the following equation:

 (Y@Au(111))  (Au(111))  (Y)

(3)

where Y denotes [CnMIm]+ or [CnMIm]+[X]- (n = 0, 2, 4, 6, 8, 10)), (Y@Au(111)), Y and (Au(111)) are the total charge density of the system, charge densities of ILs, and charge density of gold surface, respectively. The corresponding data are given in Tables S1 and S2 of the Supporting Information. 3.

Results

3.1 Equilibrium Geometries of Ionic Liquids in Gas Phase The optimized structures of ILs [CnMIm]+[X]- (where n = 0, 2, 4, 6, 8, 10 and X = Cl- and PF6-, TFSA-) were computed at the PBE/6-311++G** and PBE+D3/6-311++G** levels of theory. They are shown in Figures 1 and 2 along with the distances between the cation and the anion (Å). In the following, the cationic part of the ILs is denoted as MIm for n = 0, EMIm for n = 2, BMIm for n=4, HMIm for n=6, 6 ACS Paragon Plus Environment

Page 7 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

OMIm for n=8 and DMIm for n=10 (cf. Figures 1 and 2). There are four different interionic interactions possible in various hydrophobic/hydrophilic ILs in gas phase. These are H-bonding, vdWs, dispersion and Coulomb forces.53–55 For instance, both hydrophilic and hydrophobic ILs can form multiple C-H donor sites within the ILs to stabilize these species.56,57 Since dispersion correction is important to describe these interactions, we will focus hereafter on discussing only PBE+D3 data. Calculations show that hydrophilic [Cl]- anions are directly H-bonded with the C2-H position of the cation (i.e. first carbon attached to MIm), due to the strong electrostatic and dispersive mode of interaction with anion (Figure 1). Whereas hydrophobic fluorinated anions are weakly interacting with the cation (Figure 2). The ion-pair distances vary from 1.5 to 2.7 Å. The shortest H-bonding distance (~1.493 Å, Figure 1) is for 1-methyl-imidazolium chloride [MIm]+[Cl]- IL. For fluorinated ILs, we have similar type of bonding where the F/O atoms of the corresponding anions form H-bonds with the C-Hs of the alkyl groups (Figure 2). The calculated BEs of all selected ILs [CnMIm]+[X]- (where n = 0, 2, 4, 6, 8, 10 and X = Cl- and PF6-, TFSA-) are given in Table 1. These BEs range from -83 to -160 kcal/mol. Due to the strong H-bonding and dipole interaction within [MIm]+[Cl]- IL, the absolute value of its BE is relatively high (~160 kcal/mol). Also, Table 1 shows that fluorinated ([CnMIm]+[PF6]- and [CnMIm]+[TFSA]-) ILs have close BEs. For instance, [MIm]+[PF6]- and [MIm]+[TFSA]- have BEs of -94 and -93 kcal/mol, respectively. When one compares non-fluorinated ([CnMIm]+[Cl]-) and fluorinated ([CnMIm]+[PF6]- and [CnMIm]+[TFSA]-) ILs, the later respective BEs significantly reduce (in absolute value by ~56%). As said above this is due to the strong electrostatic interactions between the Cl- with the C2-H of the alkyl groups, whereas the charge is delocalized over the fluorinated anion, which reduces the corresponding anion-cation interaction. The lowest BE was observed in the case of [DMIm]+[TFSA]- because of the hydrophobic perfluro moiety and the longer alkyl part.31 Our work shows that the alkyl groups affect the ion-pair interaction energy due to the electron donating nature of the longer alkyl chain. For n ≥ 2, a reduction of the BE is observed because of the weakening of the anion – cation bonding. The BEs follow this order: n = 0 > 2 > 4 > 6 > 8 > 10 for [CnMIm]+[Cl]-, with a slight variation for n ≥ 2. Indeed, ILs with n ≥ 2 exhibit close BEs, whatever the length of the alkyl. For explanation, we display in Figures S1-S4 the electrostatic potential maps of the ILs of interest in the present study. These pictures show the electron rich and poor sites that can be used to interact with Au(111) surface to stabilize the complex structure. In particular, we note that the negative electrostatic potentials (red regions) accumulate in the -Cl, -PF6, and -TFSA anions of the ILs, whereas higher positive charges are localized in the MIm+ (blue regions). When increasing the alkyl chain length, the positive potentials in the cationic part reduce. Due to the overall charge distribution on the ILs, the electrostatic potentials of the ILs significantly increase from -5.359 e-2 to -8.425 e-2 for the [CnMIm]+[Cl]7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

series. This tendency is completely reversed in the case of hydrophobic ILs such as [CnMIm]+ pairing with [PF6]- and [TFSA]-. 3.2 Structure and energetics of [CnMIm]+ (n = 0, 2, 4, 6, 8, 10) cations adsorbed on Au(111) surface We performed periodic calculations to study the role of alkyl chain length in the adsorption of alkyl methyl imidazolium ([CnMIm]+, n = 0, 2, 4, 6, 8, 10) on the Au(111) surface ([CnMIm]+@Au(111)) employing dispersion corrected PBE+D3 method. The optimized geometries and adsorption distances of all alkyl chain cations (in the absence of anion) on the gold surface are shown in Figure 3. The corresponding Eads are listed in Table 2. Figure 3 shows that all [CnMIm]+ cations lay parallel to the surface via cation···π stacking interactions established between the imidazolium aromatic ring and the Au(111) surface. This was already observed for other methylimidazole containing compounds interacting with (111) surfaces.14,58–61 For imidazole interacting with Au(111), we found however that the imidazole is orthogonal to the surface.62,63 Figure 3 shows also the further stabilization of the alkyl chains by anchor assisted H-bonds (AAHBs) on the gold surface. This leads to relatively large adsorption energies. For instance, we compute an Eads of -110.3 kcal/mol for [MIm]+ cation stacked on Au(111) surface. This is associated with a cation – surface distances of 2.6 – 3.6 Å. When the alkyl chain is longer, the calculated distance between [MIm]+ (i.e. center) to Au(111) surface gradually decreases. This is due to the involvement of –CH2 group(s) of alkyl chain which form(s) several C-H···Au H-bonded interaction with Au(111) surface. The calculated adsorption energies (Eads) vary from -110 to -183 kcal/mol with strong variation when the alkyl chain is lengthened. For instance, [HMIm]+ has stronger interfacial structure than [BMIm]+ and [EMIm]+ because of the enhancement of the adsorption strength between the alkyl tails and gold surface (via weak interactions). For n = 4 and 6, the respective alkyl end groups are protrude away from the surface. For n > 6, the alkyl groups are lying parallel to the surface and strongly adsorb on the gold surface. We found that the entire corresponding alkyl groups predominantly interact at top/bridge sites of the surface. Consequently, Eads of [DMIm]+ is significantly higher (-183.2 kcal/mol) than for the others. Anyway, the longer alkyl chains gain further stabilization (larger Eads in absolute value) due to the stronger adsorption of alkyl chains on the surface. 3.3 Structure and energetics of [CnMIm]+[X]-@Au(111) surface (where n =0, 2, 4, 6, 8, 10; X= Cl, PF6, TFSA) Using the PBE+D3/TZVP method, we performed periodic computations of [CnMIm]+[X]- attached to Au(111) surface. We have selected hydrophilic [Cl]- and hydrophobic [PF6]- and [TFSA]- anions. The optimized geometries of the [CnMIm]+[X]-@Au(111) (where n = 0, 2, 4, 6, 8, 10; X= Cl, PF6, TFSA) ILs are depicted in Figures 4 and 5. 8 ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 4 presents the optimized structures of [CnMIm]+[Cl]- adsorbed on Au(111) surface. Computations show that these ILs are strongly adsorbed on the surface where the [MIm]+ aromatic ring is parallel to the surface. These ILs have similar structures as those described above for gas phase respective species. We also observe alkyl chain –CH2 groups involved in AAHBs with the surface. As for [CnMIm]+@Au(111), this is mainly due to the stabilization of [MIm]+ ring on the gold surface through cation···π stacking interactions and the subsequent induced AAHBs with the metal surface. As said above, this is in line with previous SERS measurements and computations of methylimidazole containing compound interacting with (111) surfaces.14,58–61 It is observed also that the long chain alkyl groups are interacting with a parallel configuration to the surface (bridge/top sites) via AAHB with the gold surface. The [Cl]- anions are strongly chemisorbed to the gold surface due to columbic and electrostatic interactions between the anion and the gold surface. Interestingly, we found that the adsorption sites of the [Cl]- anions depend on the length of the alkyl chain. In the case of [MIm]+[Cl]-@Au(111) and of [EMIm]+[Cl]-@Au(111), Cl- anions strongly adsorb or chemisorb at the top gold surface sites. In [BMIm]+@Au(111) and [HMIm]+@Au(111), Cl- anions are chemisorbed at the bridge sites. For [OMIm]+[Cl]-@Au(111) and [DMIm]+[Cl]-@Au(111), Cl- anions are attached at the hcp (hexagonal close-packed) sites of the gold surface. These structural changes are due to local modifications of the electronic structure of the gold surface upon adsorption of the ILs as illustrated in Figure S6. Using molecular dynamics simulations, a similar adsorption mechanism was found for the adsorption of arginine and peptides in aqueous solution on a gold(111) surface by Heinz et al.8 These authors showed that the adsorbate molecules strongly adsorb to the metal surface on the epitaxial sites and elude from the top sites. They also strongly adsorb on the Au(111) facet rather than Au(100) and Au(110) surfaces. For [CnMIm]+[Cl]-@Au(111), the calculated stacking distances between [MIm]+ ring and Au(111) surface vary from ~3.5 to 3.7 Å, whereas the distances between the alkyl chains and gold surface decrease while increasing the number of methylene groups in the chain (Figure 4). The computed Eads values for [CnMIm]+[Cl]-@Au(111) surface are given in Table 2. They range from -107 to -50.0 kcal/mol. These Eads energies significantly reduce (up to 50%) when compared to those of [CnMIm]+@Au(111) (i.e. bare cation). Again, this is because of the local charge density modification induced by the Cl- anions attached to Au(111). Table 2 shows that the shorter alkyl chains have smaller Eads energies than the longer ones, which exhibit additional vdWs and AAHB with the gold surface. For instance, Eads of short alkyl chains (n = 0 and 2) are -50 and -72 kcal/mol, respectively. Whereas the largest in absolute value Eads (-107 kcal/mol) is for [DMIm]+[Cl]-. Figure 5 displays the optimized structures of fluorinated ILs interacting with Au(111). Both [CnMIm]+[PF6]- and [CnMIm]+[TFSA]- ILs are weakly adsorbed on the Au(111) surface. In the vicinity of 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Au(111), fluorinated ILs present different behaviors when compared to [CnMIm]+[Cl]- ILs. Indeed, the anions are not directly interacting with the surface. Instead, they interact with the cation, which is adsorbed on the metallic surface. This was already noticed for other hydrophobic ILs interacting with surfaces (e.g. [BMIm]+ with [X]- (BF4-, PF6-, CH3SO3-, OTF- and TFSA-)

31).

In addition, it is worth noting that the

equilibrium structures of [CnMIm]+[PF6]- ILs are different from those found in gas phase. In the present case, the anions are bound to the imidazolium core by -stacking rather than by the multiple H-bonds described above for isolated ILs. Close analysis of the [CnMIm]+[PF6]-@Au(111) structures, show that all alkyl chain end groups are not adsorbed on the surface whereas [CnMIm]+[TFSA]- ILs behave differently at these interfaces. Moreover, we found that the [MIm]+ aromatic ring cation is in parallel stacking with the surface for [CnMIm]+[PF6]ILs whereas the same aromatic ring cation is slightly tilted on the surface for [CnMIm]+[TFSA]-@Au(111) interface. This tilting effect is due to H-bond interactions taking place between the -C2H groups of the alkyls and the oxygen atom of the TFSA- anion. Consequently, the stacking distances between the center ring of [CnMIm]+[PF6]- and Au(111) vary from ~3.5 to 3.7 Å, whereas a significantly shorter distance (of ~ 2.4 3.0 Å) is computed between the [CnMIm]+[TFSA]- ILs and Au(111). These structural effects significantly affect the ILs Eads at the gold surface. The calculated Eads for [CnMIm]+[PF6]-@Au(111) and [CnMIm]+[TFSA]-@Au(111) vary from -25 to -47 and -24 to -54 kcal/mol, respectively (Table 2). Therefore, these hydrophobic anions decrease the Eads energies by 30% compared to those of the [Cl]- anion based ILs. This is due to the charge delocalized on [PF6]- anion which can induce strong electrostatic interactions in the C2-H and alkyl chain of [MIm]+, which are dominating the ion-pair – surface interaction. For [EMIm]+[ES]- adsorbed at Au(111), adsorption energies of ~-30 to -40 kcal/mol were computed by Heinz et al.14 These earlier data are close to those found here for the [EMIm]+[PF6]- and [EMIm]+[TFSA]ILs interacting with Au (111). They are however smaller (in absolute value) than those computed for Clcontaining ILs. 4.

Epitaxial Interactions at the Au(111)-ILs Interface Previous works showed the existence of soft epitaxial interactions of ILs on gold surfaces.8,14 This

is also the case for the ILs of interest in the present work. For illustration we present in Figure 6 top views of [BMIm]+ with Cl-, or PF6- or TFSA- adsorbed at Au (111), where we highlight the different adsorption sites and the epitaxial fcc lattice contacts. In the case of hydrophilic ILs, the nitrogen atoms of Im+ cation are strongly adsorbed on the top layer of the gold atoms. In the case of hydrophobic ILs, the nitrogen atoms of Im+ have contacts with the epitaxial sites of the gold atoms in the second layer. Furthermore, one can see that the hydrophilic Cl- anion is strongly coordinated with the epitaxial fcc lattice site, whereas the hydrophobic PF6- and TFSA- anions possess contacts with the epitaxial fcc and the top layer of the gold atoms. 10 ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

As discussed above, the alkyl chains are adsorbed on gold surface mainly via AAHBs. Figure 6 shows that the alkyl groups are interacting with the top and epitaxial sites of the gold atoms. The computed absolute values of the adsorption energies are larger for the long alkyl chains except for PF6- containing ILs, which is in line with the number of top and epitaxial contacts of these alkyl groups at the gold surface. A similar observation was found in the adsorption of amino acids in aqueous solution at the Au(111) surface using MD simulations.60 Among all amino acids, arginine has the strongest binding, which is attributed to the large molecular size of arginine allowing more contacts with epitaxial sites. Whereas, small molecular size amino acids such as alanine interact with lesser number of epitaxial sites of the gold atoms. 5.

Charge population analysis Distributed electronic charges of the individual atoms are quantitatively identified by using Löwdin

population analysis. The charges between ion-pair of ILs and those of the gold surface are given in Table S2 of the Supporting Information. Figure 7 shows the plot of electronic charge of Au surface after attachment of the [CnMIm]+[X]- (X=Cl-, PF6- and TFSA-) as derived from Löwdin charge transfer analysis. The obtained electronic charges indicate that there is a significant amount of charge transfer from the ILs to the Au(111) surface upon adsorption. In hydrophilic ILs (with Cl-) the amount of charge transfer gradually goes up with increasing alkyl chain length from n = 0 to 10. The partial atomic charges are -1.47 a.u. for n=0, -1.80 a.u. for n = 2, -2.11 a.u. for n = 4, -2.71 a.u. for n = 6, -2.48 a.u. for n = 8 and -2.76 a.u. for n = 10. It is interesting to note that there is a significant amount of charge transfer in the case of [DMIm]+[Cl]- (cf. Supporting Information Table S2). This is due to the chemisorption of Cl- anion at gold surface, which can create a local modification of the surface with a more electrostatic attractive surface favoring hence the adsorption of alkyl chain through AAHBs. For hydrophobic ILs@Au(111) surface (i.e. [PF6]- and [TFSA]-), the alkyl and anions induced charge modification of the surface are different. The maximum charge transfer ability of these two anions are -0.75 a.u. for [HMIm]+[PF6]-@Au(111) and -0.98 a.u. for [DMIm]+[TFSA]-@Au(111). The minimum and maximum charge distributions are observed for [CnMIm]+[PF6]-@Au(111) when n = 0 - 6, respectively. For n > 6, the charge density reduces due to the hydrophobic alkyl chain not fully adsorbing on the surface and the perturbations of the MIm ring cation···π stacking. In the case of [TFSA]- containing ILs, we observe a gradual increment of the charge vs. the alkyl chain variation. The various ions in the anionic parts (N, O, F, S) can adsorb on various sites of the Au(111) surface inducing a significant amount of charge on the gold surface. This will affect the Eads of ILs at gold surface. These findings are useful to develop composite materials for the energy storage and bio-sensor applications. The importance of induced charges was discussed in the recent work on biomolecular interactions at the metal interfaces.69 In this study we found similar effect, which is in good agreement with the charge transfer analysis (Figure 7 and 8) of the ILs@Au(111) complexes. For illustration, we show in Figure 9 the 11 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

layer by layer magnitude of charge potential distributions for [BMIm]+@Au(111), [BMIm]+[Cl]@Au(111), [BMIm]+[PF6]-@Au(111) and [BMIm]+[TFSA]-@Au(111) complexes. One can see that most pronounced effect is for [BMIm]+[Cl]-@Au(111) complexes. Indeed, The Cl- ion binds with different sites on the gold surface this is due to the high electronegativity of Cl atom. This high electronegativity of Clion can induce the charge transfer to the gold surface. Hence, the calculated binding strength is significantly higher (i.e. > 2 times) than the respective hydrophobic cases. 6.

Charge density difference analysis

The adsorption and desorption mechanism of ILs at gold surface can be identified using electron density difference (EDD) analysis. The electron density contour maps of ILs@Au(111) surface are given in Figures S5 – S8. The blue and red regions represent negative and positive charge density contours, respectively. Figure S5 (in Supporting Information) shows that the adsorption of [CnMIm]+@Au(111) surface, (where n=0, 2, 4, 6, 8, 10) induces negative (dark blue) charge density at the carbons of the MIm+ part. When the alkyl chain length of [CnMIm]+ interacting with Au(111) increases, the positive and neutral charges accumulate however at different positions of these chains, where a spread of negative and positive charges over the alkyl chain is observed. Similar findings are observed for the cationic parts of [CnMIm]+[Cl]@Au(111). Whereas there are significant differences for the anionic part because of the specific adsorption sites (i.e. top, bridge and hcp) of Cl- anion as described above. Again, this confirms the strong electrostatic interactions between surface Cl- ions and Au, which is tuned by the alkyl chain length induced effects. For [CnMIm]+[PF6]-@Au(111), the charge density is distributed only between imidazolium ring and Au surface. For [CnMIm]+[TFSA]-@Au(111), Figure S8 shows also that the oxygen atoms of TFSA play crucial roles in sustaining electrostatic behavior throughout the molecular structure, either in the alkyl chain or the imidazolium ring. These variations significantly affect the interfacial structure at the microscopic level and the interfacial interactions between ILs and surfaces. Similarly, previous studies of ILs at the graphite surface pointed out the occurrence of charge transfer phenomena at the carbon based interfacial structure.64 7.

Discussion

Our calculations reveal that the alkyl chain of ILs adsorption on gold surface depends on the nature of the anion and on the length of its alkyl chain. Significant interfacial interaction differences are observed. We found that there are two modes of adsorption of alkyl chain on the gold surface (i) AAHBs with the involvement of cation···π stacking (ii) anion induced anchor assisted H-bonds. Indeed, we found an anion dependent behavior on the gold surface where both cation···π and AAHBs are predominant in the case of hydrophilic (i.e. [CnMIm]+[Cl]-) anion, whereas hydrophobic ILs (i.e. [CnMIm]+[X]-, X=PF6 and TFSA) behave differently. The PF6- anion at gold surface induces only the cation···π stacking of MIm+ ring while 12 ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

TFSA- induces both interactions with a reduced magnitude of adsorption. The interplay between these two modes of interactions play vital role in the stability of the complexes and the electrostatic behavior of the surface. Recent report reveals that vdWs interactions enhance the Pauli pushback effect and simultaneously reduces the charger transfer at the interface. This is in very good agreement with our ILs@Au(111) models.19 Figure 8 shows a comparison between Eads of cations and of the corresponding ILs. This figure reveal that Eads gradually increases while lengthening the alkyl chain for the bare [CnMIm]+ cation. When anions are incorporated, similar features are observed except for [CnMIm]+[PF6]-. As stressed out above, this is due the stabilization away from the surface of [CnMIm]+[PF6]- ILs and only the central part of the [MIm]+ ring is stacked with the Au(111) surface. The analysis of the interfacial structures and energetics shows the competition between the strong electrostatic interactions within the ILs ion-pair, the dispersive and vdWs interactions, the AAHBs and charge transfer phenomena within the ILs and with the gold surface. These interactions are viewed to locally alter the gold surface and to depend strongly on the nature of ILs. Indeed, different structures are found for fluorinated hydrophobic ILs of interest in gas phase and interacting with gold surface; whereas close structures are found for [CnMIm]+[Cl]- ILs in both media. Such surface induced structural features were already noticed for polypeptides, oligopeptides proteins and polymers interacting with metal surface. For instance, surface-enhanced Raman spectroscopy (SERS) studies revealed that di- and tripeptides adsorb onto silver surface through the carboxyl terminus. This is due to the side chains of the amino acid residue closest to the carboxyl terminus. Whereas SERS spectra of the polypeptides and proteins are dominated by amide bands and aromatic side chain vibrations.65 For bisphenol A-polycarbonate (BPA-PC) attached to Ni(111) surface, Delle Site et al. found that alkylphenolic end groups repel from the surface but the benzene ring remains strongly adsorbed to the surface via the involvement of its π-cloud electron.66 In addition, they noticed an interplay between adsorption energies and conformational entropy of bisphenol A-polycarbonate (BPA-PC) at Ni(111) surface. Later on, this group showed morphological differences at this interface, which can vary the melts at the interface.67 Through the investigations of oligopeptides interacting with hydrophilic Pt(111) surface in the presence of water molecules, Ghiringhelli et al. showed that the nature of the functional group influences the metal surfaces and thus the surface-specific sequence design and devices.68 8.

Conclusions In this work, first principle computations are carried out to investigate the interfacial structure of

alkyl chain substituted MIm ILs adsorbed on the Au(111) surface. Calculations show that the anion plays an important role in the behavior of the alkyl group adsorption on the Au(111) surface. We found that there 13 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

are two modes of adsorption of alkyl chains on the gold surface either via AAHBs with the involvement of cation···π stacking or via anion induced AAHBs. We thus discussed the anion dependent behavior on the gold surface. We showed that cation···π and AAHBs are predominant in the case of hydrophilic ILs (i.e. [CnMIm]+[Cl]-), whereas hydrophobic ILs (i.e. [CnMIm]+[X]- X=PF6 and TFSA) behave differently. The PF6- anion at gold surface induces only cation···π stacking of MIm+ ring while TFSA- induces both interactions with a reduced magnitude of adsorption. The interplay between these two modes of interactions is vital for the stability of the complexes and for the electrostatic behavior of the surface. Furthermore, we showed that the alkyl chain length affect the interfacial interactions and thus change the Eads of these compounds on gold. This is due to the charge transfer between ion-pair and ILs@Au(111) surface. Particularly AAHB is involved in the stability of the complexes. It is interesting to note that the [CnMIm]+ with Cl- anion strongly chemisorbs on the Au(111) surface whereas [PF6]– and [TFSA]– are mostly attached to cationic part irrespectively the alkyl chain length. Therefore, the corresponding Eads are significantly depending on the ILs nature. The electron density and charge transfer analyses are also in good accordance with our Eads values. In sum, our work disentangles the complex nature of the interactions occurring at solid-liquid interfaces. These findings should help to understand, at the microscopic level, the adsorption mechanisms of ILs@Au(111) surface. In addition, our studies reveal that the electrostatic nature of the surface depends on the anion and the binding sites of the anion on this surface via the induced charge transfer processes between the surface and the liquid. This potentially helps to find suitable energy storage and electrochemical devices. In fine, our findings may be useful to develop interface materials for the various applications such as catalysis, energy storage, supercapacitor and battery technologies. This work can be extended by full consideration of the dynamics of the adsorption of ILs on this metallic surface. This should give insights on how the ILs reach the equilibrium structures at the gold surface and the adsorption sites identified by the present DFT electronic structure computations.14,69 Both approaches are complementary to understand at the microscopic level the properties of ILs close to metallic surfaces as stated in Ref [14]. The use of molecular dynamics should compensate on some limitations of DFT on accurately describing the long range parts of the interaction potentials, where the physically expected 1/r convergence of the image charge potential to zero at long range is not well reproduced. This is related to a strong electron delocalization error and an inaccurate description of the long range van der Waals interaction.69 ASSOCIATED CONTENT Supporting Information

14 ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Additional data of Lowdin charge transfer analysis of ILs@Gas and Au(111) surface. ESP map of ILs@Gas phase and Electron Density Maps of [CnMIm]+@Au(111) surface. AUTHOR INFORMATION Corresponding Authors *E-mail: [email protected] (M.P.). *E-mail: [email protected] (M.H.). ORCID: M. Prakash: 0000-0002-1886-7708 M. Hochlaf: 0000-0002-4737-7978 Acknowledgments The authors would like to extend their sincere appreciation to the Deanship of Scientific Research at King Saud University for funding the research through the Research Group Project No. RGP-333. S. K. thanks SRM Institute of Science and Technology (SRM-IST) Research Fellowship for his research work. M. P. thanks the Department of Science and Technology-Science and Engineering Research Board (DSTSERB) of India for the financial support (Grant number: ECR/2017/000891). The authors also thank SRMIST for providing the supercomputing facility and financial support.

REFERENCES (1)

(2) (3) (4) (5) (6) (7) (8)

Koch, N.; Gerlach, A.; Duhm, S.; Glowatzki, H.; Heimel, G.; Vollmer, A.; Sakamoto, Y.; Suzuki, T.; Zegenhagen, J.; Rabe, J. P.; Schreiber, F. Adsorption-Induced Intramolecular Dipole: Correlating Molecular Conformation and Interface Electronic Structure. J. Am. Chem. Soc. 2008, 130, 7300–7304. Flores, F.; Ortega, J.; Vázquez, H. Modelling Energy Level Alignment at Organic Interfaces and Density Functional Theory. Phys. Chem. Chem. Phys. 2009, 11, 8658–8675. Jhuang, J. Y.; Lee, S. H.; Chen, S. W.; Chen, Y. H.; Chen, Y. J.; Lin, J. L.; Wang, C. H.; Yang, Y. W. Adsorption and Reaction Pathways of 1H-Pyrazole on Cu(100) and O/Cu(100). J. Phys. Chem. C 2018, 122, 6195–6208. Tanaka, K. I. Surface Nano-Structuring by Adsorption and Chemical Reactions. Materials (Basel). 2010, 3, 4518–4549. Liu, W.; Tkatchenko, A. Modeling Adsorption and Reactions of Organic Molecules at Metal Surfaces. Acc. Chem. Res. 2014, 47, 3369–3377. Carrasco, J.; Liu, W.; Michaelides, A.; Tkatchenko, A. Insight into the Description of van Der Waals Forces for Benzene Adsorption on Transition Metal (111) Surfaces. J. Chem. Phys. 2014, 140, 084704. Sievers, C.; Scott, S. L.; Noda, Y.; Qi, L.; Albuquerque, E. M.; Rioux, R. M. Phenomena Affecting Catalytic Reactions at Solid−Liquid Interfaces. ACS Catal. 2016, 6, 8286–8307. Heinz, H.; Ramezani-Dakhel, H. Simulations of Inorganic-Bioorganic Interfaces to Discover New Materials: Insights, Comparisons to Experiment, Challenges, and Opportunities. Chem. Soc. Rev. 15 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(9) (10) (11) (12) (13) (14) (15) (16) (17) (18) (19) (20)

(21) (22) (23) (24) (25) (26) (27)

2016, 45, 412–448. Biedron, A. B.; Garfunkel, E. L.; Castner, E. W.; Rangan, S. Ionic Liquid Ultrathin Films at the Surface of Cu(100) and Au(111). J. Chem. Phys. 2017, 146, 054704–054710. Lu, Y.; Li, B.; Zheng, S.; Xu, Y.; Xue, H.; Pang, H. Syntheses and Energy Storage Applications of MxSy (M = Cu, Ag, Au) and Their Composites: Rechargeable Batteries and Supercapacitors. Adv. Funct. Mater. 2017, 27, 1703949. Wei, S.; Choudhury, S.; Tu, Z.; Zhang, K.; Archer, L. A. Electrochemical Interphases for HighEnergy Storage Using Reactive Metal Anodes. Acc. Chem. Res. 2018, 51, 80–88. Romaner, L.; Nabok, D.; Puschnig, P.; Zojer, E.; Draxl, C. A. Theoretical Study of PTCDA Adsorbed on the Coinage Metal Surfaces, Ag(111), Au(111) and Cu(111). New J. Phys. 2009, 11, 053010. Toyoda, K.; Hamada, I.; Lee, K.; Yanagisawa, S.; Morikawa, Y. Density Functional Theoretical Study of Pentacene/Noble Metal Interfaces with van Der Waals Corrections: Vacuum Level Shifts and Electronic Structures. J. Chem. Phys. 2010, 132, 134703. Jha, K. C.; Liu, H.; Bockstaller, M. R.; Heinz, H. Facet Recognition and Molecular Ordering of Ionic Liquids on Metal Surfaces. J. Phys. Chem. C 2013, 117, 25969–25981. Rosa, M.; Corni, S.; Felice, R. D. Enthalpy-Entropy Tuning in the Adsorption of Nucleobases at the Au(111) Surface. J. Chem. Theory Comput. 2014, 10, 1707–1716. Ting, E. C. M.; Popa, T.; Paci, I. Surface-Site Reactivity in Small-Molecule Adsorption: A Theoretical Study of Thiol Binding on Multi-Coordinated Gold Clusters. Beilstein. J Nanotechnol. 2016, 7, 53–61. Frontera, A.; Bauzá, A. Regium–π Bonds: An Unexplored Link between Noble Metal Nanoparticles and Aromatic Surfaces. Chem. - A Eur. J. 2018, 24, 7228–7234. Zierkiewicz, W.; Michalczyk, M.; Scheiner, S. Regium Bonds between Mn Clusters (M = Cu, Ag, Au and n = 2-6) and Nucleophiles NH3 and HCN. Phys. Chem. Chem. Phys. 2018, 20, 22498– 22509. Ferri, N.; Ambrosetti, A.; Tkatchenko, A. Electronic Charge Rearrangement at Metal/Organic Interfaces Induced by Weak van Der Waals Interactions. Phys. Rev. Mater. 2017, 1, 1–8. Cremer, T.; Wibmer, L.; Calderón, S. K.; Deyko, A.; Maier, F.; Steinrück, H. P. Interfaces of Ionic Liquids and Transition Metal Surfaces - Adsorption, Growth, and Thermal Reactions of Ultrathin [C 1C 1Im] [Tf 2N] Films on Metallic and Oxidised Ni(111) Surfaces. Phys. Chem. Chem. Phys. 2012, 14, 5153–5163. Mendonça, A. C. F.; Malfreyt, P.; Pádua, A. A. H. Interactions and Ordering of Ionic Liquids at a Metal Surface. J. Chem. Theory Comput. 2012, 8, 3348–3355. Maier, F.; Bhuin, R. G.; Lexow, M.; Talwar, T.; May, B.; Heller, B. S. J.; Steinrück, H. P. TimeDependent Changes in the Growth of Ultrathin Ionic Liquid Films on Ag(111). Phys. Chem. Chem. Phys. 2018, 20, 12929–12938. Buchner, F.; Tonigold, K. F.; Bozorgchenani, M.; Gross, A.; Behm, R. J. Interaction of a SelfAssembled Ionic Liquid Layer with Graphite(0001): A Combined Experimental and Theoretical Study. J. Phys. Chem. Lett. 2016, 7, 226–233. Watanabe, M.; Thomas, M. L.; Zhang, S.; Ueno, K.; Yasuda, T.; Dokko, K. Application of Ionic Liquids to Energy Storage and Conversion Materials and Devices. Chem. Rev. 2017, 117, 7190– 7239. Buchner, F.; Tonigold, K. F.; Uhl, B.; Alwast, D.; Wagner, N. Toward the Microscopic Identification of Anions and Cations at the Ionic Liquid | Ag ( 111 ) Interface : A Combined Experimental and Theoretical Investigation. ACS Nano 2013, 7, 7773–7784. Shi, B.; Wang, Z.; Wen, H. Research on the Strengths of Electrostatic and van Der Waals Interactions in Ionic Liquids. J. Mol. Liq. 2017, 241, 486–488. Liu, Z.; Zhang, P.; Wang, Y.; Kan, Y.; Chen, Y. Surface Force Apparatus Studies on the Surface Interaction of [Cnmim+][BF4-] and [Cnmim+J][PF6-] Ionic Liquids. 2017 IEEE Int. Conf. Manip. Manuf. Meas. Nanoscale 2018, 300–304. 16 ACS Paragon Plus Environment

Page 16 of 32

Page 17 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(28) (29) (30) (31) (32) (33) (34) (35) (36) (37) (38) (39) (40) (41) (42) (43) (44) (45) (46) (47)

Jo, S.; Park, S.; Noh, C.; Jung, Y. Computer Simulation Study of Differential Capacitance and Charging Mechanism in Graphene Supercapacitors : Effects of Cyano-Group in Ionic Liquids. Electrochim. Acta 2018, 284, 577–586. Barrena, E.; Lidón, P. E.; Munuera, C.; Torrelles, X.; Ferrer, S.; Jonas, U.; Salmeron, M.; Ocal, C. The Role of Intermolecular and Molecule-Substrate Interactions in the Stability of Alkanethiol Nonsaturated Phases on Au(111). J. Am. Chem. Soc. 2004, 126, 385–395. Roos, M.; Uhl, B.; Künzel, D.; Hoster, H. E.; Grob, A.; Behm, R. J. Intermolecular vs MoleculeSubstrate Interactions: A Combined STM and Theoretical Study of Supramolecular Phases on Graphene/Ru(0001). Beilstein J. Nanotechnol. 2011, 2, 365–373. Kamalakannan, S.; Prakash, M.; Chambaud, G.; Hochlaf, M. Adsorption of Hydrophobic and Hydrophilic Ionic Liquids at the Au ( 111 ) Surface. ACS Omega 2018, 3, 18039–18051. Heinz, H.; Vaia, R. A.; Farmed, B. L. Relation between Packing Density and Thermal Transitions of Alkyl Chains on Layered Silicate and Metal Surfaces. Langmuir 2008, 24, 3727–3733. Lopes, J. N. C.; Malfreyt, P.; Pensado, A. S.; Pádua, A. A. H.; Gomes, M. F. C. Effect of Alkyl Chain Length and Hydroxyl Group Functionalization on the Surface Properties of Imidazolium Ionic Liquids. Phys. Chem. Chem. Phys. 2011, 13, 13518. Hu, Y.; Miao, K.; Xu, L.; Zha, B.; Miao, X.; Deng, W. Effects of Alkyl Chain Number and Position on 2D Self-Assemblies. RSC Adv. 2017, 7, 32391–32398. Chinwangso, P.; Hill, L. R. S.; Marquez, M.; Lee, T. Unsymmetrical Spiroalkanedithiols Having Mixed Fluorinated and Alkyl Tailgroups of Varying Length: Film Structure and Interfacial Properties. Molecules 2018, 23, 2632. Li, H.; Endres, F.; Atkin, R. Effect of Alkyl Chain Length and Anion Species on the Interfacial Nanostructure of Ionic Liquids at the Au(111)-Ionic Liquid Interface as a Function of Potential. Phys. Chem. Chem. Phys. 2013, 15, 14624–14633. Chen, Z.; Ludwig, M.; Warr, G. G.; Atkin, R. Effect of Cation Alkyl Chain Length on Surface Forces and Physical Properties in Deep Eutectic Solvents. J. Colloid Interface Sci. 2017, 494, 373– 379. Liu, Z.; Cui, T.; Lu, T.; Ghazvini, M. S.; Endres, F. Anion Effects on the Solid/Ionic Liquid Interface and the Electrodeposition of Zinc. J. Phys. Chem. C 2016, 120, 20224–20231. Pensado, A. S.; Pádua, A. A. H. Solvation and Stabilization of Metallic Nanoparticles in Ionic Liquids. Angew. Chemie - Int. Ed. 2011, 50, 8683–8687. Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. Grimme, S. Semiempirical Hybrid Density Functional with Perturbative Second-Order Correlation Semiempirical Hybrid Density Functional with Perturbative Second-Order Correlation. J. Chem. Phys. 2006, 124, 034108–034116. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction ( DFT-D ) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104–15123. Prakash, M.; Chambaud, G.; Al-Mogren, M. M.; Majdi, H. Role of Size and Shape Selectivity in Interaction between Gold Nanoclusters and Imidazole : A Theoretical Study. J. Mol. Model 2014, 20, 2534. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; et al. Gaussian 16, Revision A.03; Gaussian, Inc.: Wallingford, CT, 2016. Hay, P. J.; Wadt, W. R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for K to Au Including the Outermost Core Orbitals Potentials for K to Au Including the Outermost Core Orbitals. J. Chem. Phys. 1985, 82, 299. Boys, S. F.; Bernardi, F. The Calculation of Small Molecular Interactions by the Differences of Separate Total Energies . Some Procedures with Reduced Errors. Mol. Phys. 2002, 19, 553–556. Vandevondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J. QUICKSTEP : 17 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(48) (49) (50) (51) (52) (53) (54) (55) (56) (57) (58) (59) (60) (61) (62) (63) (64) (65) (66)

Fast and Accurate Density Functional Calculations Using a Mixed Gaussian and Plane Waves Approach. Comput. Phys. Commun. 2005, 167, 103–128. Hutter, È.; Parrinello, M.; Lippert, G. Regular Article The Gaussian and Augmented-Plane-Wave Density Functional Method for Ab Initio Molecular Dynamics Simulations. Theor. Chem. Acc. 1999, 103, 124–140. Taylor, P.; Lippert, B. G.; Hutter, J.; Parrinello, M. A Hybrid Gaussian and Plane Wave Density Functional Scheme. Mol. Phys. 1997, 92, 477–488. Vandevondele, J.; Hutter, J. Gaussian Basis Sets for Accurate Calculations on Molecular Systems in Gas and Condensed Phases. J. Chem. Phys. 2012, 127, 114105–11409. Vandevondele, J.; Hutter, J. An Efficient Orbital Transformation Method for Electronic Structure Calculations An Efficient Orbital Transformation Method for Electronic Structure Calculations. J. Chem. Phys. 2003, 118, 4365–4369. Nguyen, M.; Pignedoli, C. A.; Treier, M.; Passerone, D. The Role of van Der Waals Interactions in Surface-Supported Supramolecular Networks. Phys. Chem. Chem. Phys. 2010, 12, 992–999. Fumino, K.; Reimanna, S.; Ludwig, R. Probing Molecular Interaction in Ionic Liquids by Low Frequency Spectroscopy: Coulomb Energy, Hydrogen Bonding and Dispersion Forces. Phys. Chem. Chem. Phys. 2014, 16, 21903–21929. Dong, V.; Zhang, S.; Wang, J. Understanding the Hydrogen Bonds in Ionic Liquids and Their Roles in Properties and Reactions. Chem. Commun. 2016, 52, 6744–6764. Hunt, P. A. Quantum Chemical Modeling of Hydrogen Bonding in Ionic Liquids. Top. Curr. Chem. 2017, 375, 225–246. Hunt, P. A.; Ashworth, C. R.; Matthews, R. P. Hydrogen Bonding in Ionic Liquids. Chem. Soc. Rev. 2015, 44, 1257–1288. Mendonça, A. C. F.; Pádua, A. A. H.; Malfreyt, P. Nonequilibrium Molecular Simulations of New Ionic Lubricants at Metallic Surfaces: Prediction of the Friction. J. Chem. Theory Comput. 2013, 9, 1600–1610. Heinz, H.; Farmer, B. L.; Pandey, R. B.; Slocik, J. M.; Patnaik, S. S.; Pachter, R.; Naik, R. R. Nature of Molecular Interactions of Peptides with Gold, Palladium, and Pd-Au Bimetal Surfaces in Aqueous Solution. J. Am. Chem. Soc. 2009, 131, 9704–9714. Coppage, R.; Slocik, J. M.; Briggs, B. D.; Frenkel, A. I.; Heinz, H.; Naik, R. R.; Knecht, M. R. Crystallographic Recognition Controls Peptide Binding For. J. Am. Chem. Soc. 2011, 133, 12346– 12349. Feng, J.; Pandey, R. B.; Berry, R. J.; Farmer, B. L.; Naik, R. R.; Heinz, H. Adsorption Mechanism of Single Amino Acid and Surfactant Molecules to Au {111} Surfaces in Aqueous Solution: Design Rules for Metal-Binding Molecules. Soft Matter 2011, 7, 2113–2120. Coppage, R.; Slocik, J. M.; Ramezani-Dakhel, H.; Bedford, N. M.; Heinz, H.; Naik, R. R.; Knecht, M. R. Exploiting Localized Surface Binding Effects to Enhance the Catalytic Reactivity of PeptideCapped Nanoparticles. J. Am. Chem. Soc. 2013, 135, 11048–11054. Benoit, D. M.; Izzaouihda, S.; Komiha, N.; Hassna, A. E. M.; Mahjoubi, K. Adsorption of Imidazole on Au(111) Surface: Dispersion Corrected Density Functional Study. Appl. Surf. Sci. 2016, 383, 233–239. Prakash, M.; Mathivon, K.; Benoit, D. M.; Chambaud, G.; Hochlaf, M. Carbon Dioxide Interaction with Isolated Imidazole or Attached on Gold Clusters and Surface : Competition between σ H-Bond and π Stacking. Phys. Chem. Chem. Phys. 2014, 16, 12503–12509. Gonazalez, G. V.; Fuente, A. G.; Vega, A.; Carrete, J.; Cabeza, O.; Gallego, L. J.; Varela, L. M. Density Functional Study of Charge Transfer at the Graphene/Ionic Liquid Interface. J. Phys. Chem. C 2018, 122, 15070–15077. Stewart, S.; Fredericks, P. M. Surface-Enhanced Raman Spectroscopy of Peptides and Proteins Adsorbed on an Electrochemically Prepared Silver Surface. Spectrochim. Acta - Part A Mol. Biomol. Spectrosc. 1999, 55, 1615–1640. Site, L. D.; Leon, S.; Kremer, K. BPA-PC on a Ni(111) Surface: The Interplay between Adsorption 18 ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(67) (68) (69)

Energy and Conformational Entropy for Different Chain-End Modifications. J. Am. Chem. Soc. 2004, 126, 2944–2955. Site, L. D.; Abrams, C. F.; Alavi, A.; Kremer, K. Polymers near Metal Surfaces: Selective Adsorption and Global Conformations. Phys. Rev. Lett. 2002, 89, 156103. Ghiringhelli, L. M.; Hess, B.; Vegt, N. F. A. V.; Site, L. D. Competing Adsorption between Hydrated Peptides and Water onto Metal Surfaces: From Electronic to Conformational Properties. J. Am. Chem. Soc. 2008, 130, 13460–13464. Geada, I. L.; Ramezani-Dakhel, H.; Jamil, T.; Sulpizi, M.; Heinz, H. Insight into Induced Charges at Metal Surfaces and Biointerfaces Using a Polarizable Lennard-Jones Potential. Nat. Commun. 2018, 9, 1–14.

Figure Captions: Figure 1. Optimized structures of [CnMIm]+[Cl]- complexes as computed at the PBE+D3/6-311++G** level along with the cation-anion distances in Å. Figure 2. Optimized structures of [CnMIm]+[PF6]- and of [CnMIm]+[TFSA]- complexes as computed at the PBE+D3/6-311++G** level along with the cation-anion distances in Å. 19 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 32

Figure 3. Optimized geometrical structures of [CnMIm]+@Au(111) surface was computed at the PBE+D3/TZVP level of theory and nearest interacting distances (in Å) between [CnMIm]+ and Au(111) surface. Figure 4. Optimized geometrical structures of Non-fluorinated [CnMIm]+[Cl]-@Au(111) surface was computed at the PBE+D3/TZVP level of theory and nearest interacting distances (in Å) between ILs and Au(111) surface. Figure

5.

Optimized

geometrical

structures

of

Fluorinated

[CnMIm]+[PF6]-@Au(111)

and

[TFSA]@Au(111) surface was computed at the PBE+D3/TZVP level of theory and nearest interacting distances (in Å) between ILs and Au(111) surface. Figure 6. Snapshots representation of epitaxial interations sites of various ILs@Au(111): (1) [BMIm]+, (2) [BMIm]+[Cl]-, (3) [BMIm]+[PF6]- and (4) [BMIm]+[TFSA]-. The red circles highlight the epitaxial soft contacts. Figure 7. Electronic charge of [CnMIm]+[X]-@Au(111) surface was computed by Löwdin Charge transfer analysis. Figure 8. Relation between calculated Eads with number of carbon atoms at [CnMIm]+[X]-@Au(111) surface by using (DFT+D3). Figure 9: Magnitude of charge potential distribution in ILs@Au(111) complexes with layer by layer analysis. (1) for [BMIm]+, (2) for [BMIm]+[Cl]-, (3) for [BMIm]+[PF6]- and (4) for [BMIm]+ [TFSA]complexes.

20 ACS Paragon Plus Environment

Page 21 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1. Optimized structures of [CnMIm]+[Cl]- complexes as computed at the PBE+D3/6-311++G** level along with the cation-anion distances in Å.

21 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. Optimized structures of [CnMIm]+[PF6]- and [CnMIm]+[TFSA]- complexes as computed at the PBE+D3/6-311++G** level along with the cation-anion distances in Å.

22 ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 3. Optimized geometrical structures of [CnMIm]+ adsorbed on Au(111) surface computed at the PBE+D3/TZVP level of theory. We give also the nearest distances (in Å) between [CnMIm]+ and the Au(111) surface.

23 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4. Optimized geometrical structures of Non-fluorinated [CnMIm]+[Cl]-@Au(111) surface was computed at the PBE+D3/TZVP level of theory and nearest interacting distances (in Å) between ILs and Au(111) surface.

24 ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure

5.

Optimized

geometrical

structures

of

fluorinated

[CnMIm]+[PF6]-@Au(111)

and

[CnMIm]+[TFSA]-@Au(111) as computed at the PBE+D3/TZVP level of theory and nearest interacting distances (in Å) between ILs and Au(111) surface.

25 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6. Snapshots representation of epitaxial interations sites of various ILs@Au(111): (1) [BMIm]+, (2) [BMIm]+[Cl]-, (3) [BMIm]+[PF6]- and (4) [BMIm]+[TFSA]-. The red circles highlight the epitaxial soft contacts.

26 ACS Paragon Plus Environment

Page 26 of 32

Page 27 of 32

0.0

-0.5

Electronic charge (e)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

-1.0

-1.5

-2.0 +

[CnMIm] [Cl]

-

+

-2.5

[CnMIm] [TFSA] +

[CnMIm] [PF6]

-

-

-3.0 0

2

4

6

8

10

No. of carbon atoms Figure 7. Electronic charge of Au surface at the [CnMIm]+[X]-@Au(111) interface as derived from Löwdin charge transfer analysis. See Table S2 for more details.

27 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 8. Relation between calculated PBE+D3/TZVP Eads with number of carbon atoms at [CnMIm]+[X]@Au(111) surface.

28 ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 9: Magnitude of charge potential distribution in ILs@Au(111) complexes with layer by layer analysis. (1) for [BMIm]+, (2) for [BMIm]+[Cl]-, (3) for [BMIm]+[PF6]- and (4) for [BMIm]+ [TFSA]complexes.

29 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 32

Table 1. Binding energies (BEs, kcal/mol) of optimized [CnMIm]+[X]- (X = Cl, PF6, TFSA) ILs in gas phase. Method/basis set

PBE/6-311++G(d,p)

PBE+D3/6-311++G(d,p)

X n=

BEs 0

2

4

6

8

10

Cl

-159.2 -95.9 -95.3 -94.9 -94.8

-94.8

PF6

-89.0

-80.3 -79.7 -79.4 -79.2

-78.9

TFSA

-87.0

-78.2 -77.2 -76.9 -76.7

-76.6

Cl

-159.8 -97.3 -96.9 -96.6 -96.5

-96.5

PF6

-91.4

-84.4 -84.3 -84.1 -84.0

-84.3

TFSA

-90.9

-84.0 -83.6 -83.5 -83.4

-83.3

30 ACS Paragon Plus Environment

Page 31 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table 2. Adsorption energies (Eads, kcal/mol) of [CnMIm]+ (alkyl chains) and of [CnMIm]+[X]- (X = Cl, PF6, TFSA) ILs on Au(111) surface. Methods

PBE/TZVP

ILs@Au(111)

Adsorption Energies (Eads) n= 0 – 10 0

2

4

6

8

10

[CnMIm]+

-82.6

-70.4

-81.1

-87.9

-76.2

-183.2

[CnMIm]+[Cl]-

-32.9

-25.8

-24.0

-31.9

-37.6

-38.2

[CnMIm]+[PF6]-

12.5

12.2

7.1

22.1

14.3

13.1

[CnMIm]+[TFSA]-

2.2

8.2

13.07

20.9

18.2

8.4

[CnMIm]+

-110.3

-109.0

-114.0

-125.9

-132.2

-183.2

PBE+D3

[CnMIm]+[Cl]-

-50.4

-72.1

-81.8

-94.3

-96.0

-107.3

/TZVP

MIm]+[PF

-32.9

-31.4

-46.6

-42.6

-24.8

-31.7

-24.2

-29.8

-28.6

-42.7

-50.0

-53.8

[Cn

]-

6

[CnMIm]+[TFSA]-

31 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

214x116mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 32 of 32