You are on page 1of 9

pubs.acs.

org/JPCC Article

A Molecular View of the Ionic Liquid Catalyst Interface of SCILLs:


Coverage-Dependent Adsorption Motifs of [C4C1Pyr][NTf2] on Pd
Single Crystals and Nanoparticles
Christian Schuschke, Lukas Fromm, Johannes Träg, Corinna Stumm, Chantal Hohner,
Roman Eschenbacher, Sebastian Grau, Dirk Zahn, Andreas Görling, Tanja Bauer,* and Jörg Libuda
Cite This: J. Phys. Chem. C 2021, 125, 13264−13272 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via INDIAN INST OF TECH KHARAGPUR on February 13, 2023 at 08:45:39 (UTC).

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Solid catalysts with ionic liquid layers (SCILLs) are


heterogeneous catalysts coated with a thin layer of an ionic liquid
(IL) to improve their selectivity. In this study, we investigated the
interplay of the room-temperature IL 1-butyl-1-methylpyrrolidinium
bis(trifluoromethylsulfonyl)imide [C4C1Pyr][NTf2] with transition
metal surfaces of both Pd(111) single crystals and Pd nanoparticles
(NPs) supported on highly oriented pyrolytic graphite (HOPG). To
this end, we combine theoretical insights obtained by density
functional theory (DFT) and molecular dynamics (MD) calcu-
lations with experimental data acquired via time-resolved infrared
reflection absorption spectroscopy (TR-IRAS) performed under
ultrahigh vacuum (UHV) conditions. IL monolayer films formed on
Pd(111) and Pd NPs are strongly bound to the surface via the SO2
moiety of the [NTf2]− anion. By combination of IRAS, DFT, and MD, we identify the most common adsorption motifs. Most
importantly, the binding motif of the anion changes as a function of the IL coverage. On supported Pd NPs, additional effects arise
from the particle morphology and from the diffusion of carbon from the support to the Pd NPs. Our results suggest that the effect of
the IL film on the catalytic activity should strongly depend on the local coverage and the morphology of the catalyst NPs.

■ INTRODUCTION
In many processes, catalytic modifiers are used to improve the
these interactions, it is possible to select specific ions,13−15
functionalize the ions,15,16 or modify their structure.17 This
selectivity and activity of noble metal based catalysts.1−4 One allows to tune the interaction strength with the surface, the
particular approach to improve the catalytic performance is the interaction with reactants, and also the wetting behavior on the
application of supported ionic liquid (IL) films in the form of a catalyst.
so-called supported ionic liquid phase (SILP) catalyst or a In spite of the commercial success of the SCILL concept, the
solid catalyst with ionic liquid layer (SCILL). In a SILP interaction of the IL with the active sites is poorly understood
catalyst, a molecular catalyst complex is dissolved in an IL film at the molecular level. Such molecular interactions at the IL/
coating a porous support. 5,6 In a SCILL material, a catalyst interface are addressed best with a surface science
conventional supported catalyst is coated with an IL film to approach.13,18,19 In several studies it has been demonstrated
improve the selectivity.7−9 Recently, such SCILL systems have that it is indeed possible to study the chemical interactions,
been developed for the use at industrial scale for selective orientation, and wetting of ILs on atomically defined
hydrogenation reactions.10 surfaces.20,21 To exploit the full potential of the surface science
Because ILs are highly versatile compounds in terms of their approach, such experiments are best performed under ultrahigh
chemical properties, IL-based catalytic modifiers can be vacuum (UHV) conditions, which is possible because of the
tailored to suit the specific demand of the process. Typical low vapor pressure of many ILs.18,19,22 Furthermore, several
ILs combine large organic cations with suitable anions, which
results in a vast number of ionic materials, many of which are
liquid even at room temperature and cover a wide range of Received: March 9, 2021
physicochemical properties.11,12 Of special interest for the Revised: May 31, 2021
catalytic properties is the molecular layer of the IL which is in Published: June 11, 2021
direct contact with the active site. The properties of these
adsorbed IL layers may be strikingly different from those of the
bulk IL due to the specific interactions at the interface. To tune
© 2021 The Authors. Published by
American Chemical Society https://doi.org/10.1021/acs.jpcc.1c02131
13264 J. Phys. Chem. C 2021, 125, 13264−13272
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 1. Investigated model systems: (a) molecular structures of the IL anion [NTf2]− and the cation [C4C1Pyr]+; (b) schematic representation of
the in situ IRAS setup to investigate the Pd(111) single crystal and Pd NPs on HOPG.

ILs are sufficiently stable to deposit them by physical vapor ization tools along with evaporation sources and a remote-
deposition (PVD).23,24 In this way, it is feasible to prepare controlled gas dosing system to prepare and investigate films in
ultraclean IL films of well-controlled thickness on atomically situ with a vacuum Fourier-transform infrared (FTIR)
controlled surfaces.25,26 Recently this approach was used to spectrometer (Bruker Vertex 80v).
study growth modes,27−29 the specific interactions of ions with Preparation of the Pd(111) Single Crystal. The
the surface,13,30 and the effect of preadsorbed reactants.8,31 Pd(111) single crystal (MaTecK) was cleaned by Ar +
Previously, a strong focus was on Pd and Pt catalysts, and it sputtering (1 keV, 300 K, 60 min, 8 × 10−5 mbar, Linde,
was shown that the site-specific interactions are different on purity 99.9999%). While in an O2 atmosphere (4 × 10−8 mbar,
the two noble metals.17,32 In the present work, we address a Linde, purity 99.999%), the sample was first annealed at 923 K
question that has been studied rarely in previous work. We for 10 min and then cooled below 300 K. Residual oxygen
investigate how the adsorption motifs of the IL change when atoms were removed by a subsequent flash to 923 K in UHV.
we change the IL loading on the catalyst surface. We believe The cleanliness of the sample was checked via IRAS,
that these adsorption motifs will be essential for the reactivity investigating the stepwise adsorption of CO at 300 K (1 ×
to be explored in future studies. To obtain detailed insights 10−6 mbar, Westfalen, 99.97%). The supplied CO was purified
into the adsorption motifs, we combine infrared reflection by both a home-built liquid nitrogen cold trap and a
absorption spectroscopy (IRAS) with density functional theory commercial adsorption filter (Pall Gaskleen II) to remove
(DFT) calculations and molecular dynamics (MD) simu- any traces of carbonyls. After the CO adsorption test, the
lations. As a model system we use films of the room- surface was briefly heated to 623 K to remove CO.
temperature IL 1-butyl-1-methylpyrrolidinium bis(trifluoro- Preparation of Pd NPs on HOPG. The HOPG crystal
methylsulfonyl)imide [C4C1Pyr][NTf2], prepared by PVD on (MikroMasch, ZYA, 0.4° mosaic spread) was cleaned via the
Pd(111) single crystal surfaces and Pd nanoparticles (NPs) “Scotch tape” method prior to transfer into the UHV chamber.
(see Figure 1a,b). Previously, we performed a similar study on The sample was annealed at 800 K for 3 min in UHV to
Pt single crystals and Pt NPs,33 which allows us to compare the remove residual gases. The HOPG sample was bombarded
two SCILL systems directly. For all model SCILLs, we observe with Ar+ (0.5 keV, 300 K, 60 min, 8 × 10−5 mbar) to create the
strongly adsorbing IL monolayer films covered by a bulklike nucleation sites necessary for the growth of Pd NPs.35,36 Pd
phase in which the orientation is successively lost. In spite of was deposited from a wire (Alfa Aesar, 99.9% metals basis, d =
the strong interaction with the metal, the local structure of the 1 mm) loaded into a commercial electron beam evaporator
adsorbed IL layer remains flexible. The molecular orientation (Focus EFM3). The deposition of Pd for 36 min (flux of 3 nA)
and adsorption motifs in these wetting layers change at a sample temperature of 300 K led to the formation of a Pd
dynamically as a function of the IL coverage and the surface film with an equivalent thickness of 15 Å, which was calibrated
morphology.


by a quartz crystal microbalance. The morphology of the Pd
NPs prepared was examined by scanning tunneling microscopy
METHODS (STM).
Experimental Setup. The IRAS experiments were STM Measurements and Characterization of the Pd
performed in an UHV system described in detail in an earlier NPs. The characterization of the Pd NPs was performed ex situ
publication.34 In brief, the system (base pressure of after the preparation and IRAS measurements. For this, we
1 × 10−10 mbar) features various preparation and character- used an AFM/STM system (Keysight Technologies, Series
13265 https://doi.org/10.1021/acs.jpcc.1c02131
J. Phys. Chem. C 2021, 125, 13264−13272
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 2. Characterization of the annealing behavior of Pd/HOPG (nominal thickness 15 Å) by ex situ STM images: (a) as prepared (300 K), (b)
after annealing to 500 K, (c) after annealing to 700 K, and by (d) IRAS and CO adsorption at 10−6 mbar CO (300 K, exponentially increasing
doses, ranging from 0.025 to 25.6 langmuirs, and ultimately spectra in a CO atmosphere). Acquisition parameters of the STM images: tunneling
current of 0.1 nA for all images and tunneling bias of 0.3 V for all images, except for the right panel of (a) (0.2 V).

5500 AFM/SPM), which is isolated from vibrations by a refer to the Supporting Information and Figure S1 therein. In
combination of active and passive noise damping. We brief, the CO adsorption on the Pd NPs on HOPG indicates
conducted the STM measurements using tips that were the presence of mostly (111) facets, with a strong contribution
manually cut from a PtIr (80:20) wire (MaTecK, 99.99%, of defect and edge sites. Particles annealed to 500 K in UHV
d = 0.25 mm). The STM images were obtained in constant show features that indicate the presence of carbon species
current mode by using a typical tunneling bias of 0.2−0.3 V originating from the HOPG support. These species occupy
applied to the sample and a tunneling current of 0.1 nA. The edge sites and strongly increase the signal of on-top CO. Above
STM images were postprocessed (row-aligning, data-leveling) 700 K, the particles sinter rapidly and the Pd surface decreases.
by using Gwyddion software (ver. 2.55).37 Deposition of [C4C1Pyr][NTf2]. The IL [C4C1Pyr][NTf2]
In Figure 2a−c, selected STM images are shown. They were (Merck KGA, ≥98%) was evaporated from a glass crucible
recorded ex situ as prepared at 300 K and after annealing to placed in a home-built Knudsen cell (see Figure 1a,b) to
500 and 700 K. We estimate a particle density of prepare the thin films (thickness ∼15 ML). Before the
4.6 × 1012 cm−2 for the Pd NPs on a sample investigated as experiments, the evaporator was separated from the main
prepared (Figure 2a), thus without the annealing step. From chamber by a gate valve, pumped by a separate high-vacuum
the Pd loading (average thickness of 15 Å, 5.7 × 1015 atoms line, and moderately baked for 24 h. Before the start of the
cm−2) and the unit cell parameters, we estimate the particles deposition, the evaporant was preheated to 400 K while the
contain 1300 atoms on average. Assuming a half-spherical sample was protected by a gate valve and a shutter to block the
shape, this corresponds to a particle size of 3.8 nm. Annealing beam path.
the sample to 500 K leads to a minor growth of the particle size Time-Resolved IRAS Experiments (TR-IRAS). IR spectra
(Figure 2b). From the particle density (2.3 × 1012 cm−2), we (rate of 1 spectrum/min, spectral resolution of 4 cm−1,
estimate one particle contains 2500 atoms on average, which p-polarized light only) were continuously acquired during the
corresponds to a calculated diameter of 4.7 nm. Further deposition of the IL. For the CO adsorption, spectra were
annealing to 700 K induces major particle growth (Figure 2c). recorded in UHV after the CO pulse. All data were referenced
The particles agglomerate to aggregates with an average to a background spectrum of the clean sample acquired prior
diameter of 16 nm. The particle density of these large to the experiment.
aggregates is 1.3 × 1011 cm−2, which corresponds to 45.000 Density Functional Theory (DFT) Calculations. DFT
atoms on average. calculations of adsorbed IL structures on the Pd(111) surface
The CO adsorption and thermal stability of the particles were performed with the Vienna Ab initio Simulation Package
were investigated by IRAS (Figure 2d). For a full analysis, we (VASP ver. 5.4.1).38−40 The projector-augmented wave
13266 https://doi.org/10.1021/acs.jpcc.1c02131
J. Phys. Chem. C 2021, 125, 13264−13272
The Journal of Physical Chemistry C


pubs.acs.org/JPCC Article

method (PAW)41 was used with an energy cutoff of 450 eV RESULTS AND DISCUSSION
together with the PBE exchange correlation functional.42 IR Spectrum of [C4C1Pyr][NTf2]. We first consider the IR
Additionally the DFT-D3 correction scheme by Grimme with spectra of [C4C1Pyr][NTf2] deposited by PVD on Pd(111)
the Becke−Johnson damping was employed.43,44 Energies and and Pd NPs/HOPG. The IR spectra of multilayer films (∼10
geometries were optimized until convergence of 10−7 eV and monolayers (ML), while 1 ML is defined as a closed double
forces smaller than 0.05 eV/Å were reached. The system was layer of chemisorbed species, i.e., ion pairs) are shown in
modeled by a supercell approach with a six- or four-layered Figure 3. For both substrates the spectra are nearly identical,
periodic surface unit cell separated by 20 Å of vacuum. The
bottom half of this metal surface was fixed to bulk geometry,
while the top half was allowed to relax. For the Pd(111)
surface, the experimental lattice constant of 3.890 Å was used.
Three different cells representing different coverages were
calculated. A cell with one isolated IL ion pair in a 6 × 6
orthogonal surface cell (four layers), allowing enough lateral
space to decouple the periodic images, was calculated by using
only the Γ-point. In a 4 × 4 orthogonal unit cell (six layers)
with a 3 × 3 × 1 Monkhorst−Pack k-point sampling45 one ion
pair increasingly interacts with the periodic images, modeling
an ideally “low-density” monolayer. The third cell was chosen
to mimic the monolayer density obtained from the MD
simulations explained subsequently (1.35 ion pairs nm−2).
Therefore, three ion pairs were calculated on a four-layered Figure 3. IRAS spectra of the [C4C1Pyr][NTf2] multilayer deposited
8 × 4 orthogonal surface unit cell which was sampled with a via PVD at 300 K, grown on Pd(111) and both pristine and annealed
Pd NPs.
1 × 2 × 1 Monkhorst−Pack k-point mesh. Adsorption energies
(Eads = Emol + Esurf − Esys) were estimated with one IL ion pair
in the gas phase (Emol) as reference. Vibrational spectra were
calculated within the harmonic approximation. The data were and the bands can be assigned according to previous
visualized by using the software Avogadro (ver. 1.1.1)46 and studies.29,32,33 In brief, the bands of the anion are found at
QVibeplot (ver. 1.6.0).47 1361 cm−1 (ν(SO2)as), 1231 cm−1 (ν(SO2)s + ν(CF3)s), and
Molecular Dynamics (MD). Fully atomistic MD simu- 1213 cm−1 (ν(NTf2−), comprising ν(SNS)as + ν(CS) +
lations were performed with the Large-scale Atomic/Molecular ν(CF3)s), 1145 cm−1 (ν(SO2)s + ν(CF3)as), and 1065 cm−1
Massively Parallel Simulator (LAMMPS, code ver. 2017)48 and (ν(SNS)as + ν(CF3)s + ν(SO2)s). Because these vibrations
the additional GPU package.49−51 The integration time step comprise contributions from various stretching modes, we only
was set to 1.0 fs. Electrostatic and van der Waals interactions list the strongest contribution in the following.
were described with the shifted cutoff potentials by using a Aliphatic deformations of the cation are visible at 1468 cm−1
distance delimiter of 12.0 Å. To impose constant temperature, (ν(CHx)). However, the cation bands are weak and will not be
the system was coupled to a Berendsen thermostat with a analyzed in detail. Comparison of the spectra with DFT and
relaxation constant of 0.5 ps. For the simulation of bulk IL on attenuated total reflection (ATR) data from previous experi-
the Pd surface, the simulation cell was coupled to an ments33 proves that the IL is deposited by PVD without
anisotropic Berendsen barostat (relaxation only in the decomposition.
z-direction) with a relaxation constant of 100.0 ps. The Growth of [C4C1Pyr][NTf2] on Pd(111). To identify
atoms of the Pd slab were kept at fixed positions during the specific adsorption motifs of the first IL layer, we performed
simulations to avoid the flying-ice-cube effect.52 The force field DFT calculations and MD simulations. The [NTf2]− anion can
parameters for [C4C1Pyr][NTf2] were taken from the GAFF2 interact strongly via both SO2 moieties with the surface, which
(ver. 2.1) parameter set (atom types sy and n2 for sulfur and allows for a variety of binding motifs displayed in Figure 4a.
nitrogen atoms of NTf2−).53 These parameters reproduce a We denote the motifs with two indices “i−j”, which indicate
the number of oxygen atoms of the respective SO2 moieties i
reasonably accurate potential energy landscape with respect to
and j interacting with the surface (for further details on the
a 2D scan of the dihedral angles C1−S1−N−S2 and S1−N−S2−
analysis, see the Supporting Information). The different
C2 (relative to a B3LYP/6-31G* reference), which are critical
adsorption motifs of the anion were subjected to quantum
for the overall structure of NTf2−. Atomic partial charges were chemical calculations to identify their vibrational/IR spectral
calculated with the restricted electrostatic potential (RESP) signature. The adsorption motifs of the cation are displayed in
method54 based on electrostatic potential calculations at the Figure 4b.
level of HF/6-31G* with the Gaussian 09 rev. D.01 code.55 In Figure 5, we compare (a) spectra simulated by DFT (for
The van der Waals parameters for the Pd atoms were adopted a full set and list of assignment, we refer to the Supporting
from Heinz et al.56 Local fluctuations of the image charges on Information) with the results of three TR-IRAS experiments
the metal surface induced by the IL are not considered, as (b−d). Specifically, we consider the deposition of the IL (b)
previous studies from our group indicated that the constant on Pd(111), (c) on Pd NPs, and (d) on Pd NPs that were
charge approach allows for a reasonable description of annealed to 500 K prior to deposition of the IL. The DFT
adsorption motifs in terms of structures.57 For an extensive spectra of [NTf2]− (see Figure 5a) depend strongly on the
description of the MD simulation procedure, please refer to the binding motif and the orientation of the anion (see Figure 4 for
Supporting Information. a sketch of the adsorption motifs).
13267 https://doi.org/10.1021/acs.jpcc.1c02131
J. Phys. Chem. C 2021, 125, 13264−13272
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 4. Adsorption motifs of the (a) anion [NTf2]− and (b) cation
[C4C1Pyr]+ on Pd(111) as suggested by MD.

Experimentally, we examined the film growth on a clean


Pd(111) surface (see Figure 5b) first. Time-resolved IR spectra
were acquired during the deposition of a multilayer film of the
IL at 300 K (left panel). The peak height was determined to
display the development of selected signals over time (right
panel). At low coverage, we observe two weak bands at 1233
and 1364 cm−1 and one stronger band at 1104 cm−1. With
increasing coverage, the peak at 1233 cm−1 shifts slightly to
lower wavenumbers and grows in intensity. We assign this
signal to contributions of ν(SO2)s. The two signals at 1364 and
1104 cm−1 behave differently. They are visible at low coverage
only, and they are weaker than the signals observed in the full
monolayer (described below). We assign them to the ν(SO2)as
and ν(SO2)s modes of anions specifically adsorbed to the Pd
surface. The experimental data strongly resemble the spectrum
of the “2−2” configuration suggested by DFT calculations for
low IL coverages (see Figure 4 for adsorption motifs). To
understand the spectral appearance, we have to consider the
metal-surface selection rule (MSSR), which only allows for
dynamic dipole moments to be observed with a component
along the surface normal.58 The ν(SO2)as mode is polarized
along the axis spanned by the two oxygen atoms. Therefore,
the SO2 group must be strongly tilted with respect to the
surface to be visible. The ν(SO2)s mode, on the other hand, is
polarized along the axis spanned by the center between the
oxygen atoms and the sulfur atom. Thus, the mode will be
visible if both oxygen atoms interact with the surface
simultaneously. The experimental spectra suggest that both
ν(SNS)as and ν(SO2)as modes are attenuated, and therefore,
they are polarized nearly parallel to the surface. This
observation is in line with the theoretical data which suggest
the adsorption of a majority of species as a cis-isomer via both
SO2 groups in a “2−2” fashion, while the CF3 groups are
directed toward the vacuum. Note that we expect a mixture of
Figure 5. Comparison of [C4C1Pyr][NTf2] film growth on different
prevalent adsorption motifs at this coverage but are only able substrates with theoretical data: (a) DFT calculations of isolated and
to describe to the most abundant species. 1 ML IL on Pd(111), TR-IRAS during the IL deposition via PVD at
In the next step, we investigated the growth of the IL layer 300 K onto (b) Pd(111), (c) pristine Pd NPs, (d) Pd NPs annealed
with MD simulations. We find that the IL molecules (or ion to 500 K. Left column: IRAS spectra; right column: area of all IL
pairs) of a submonolayer (0.5 ML) are not homogeneously bands in the range from range from 950 to 1400 cm−1 (labeled as
distributed over the Pd surface but tend to form IL islands. ν(NTf2−)) of selected IL and CO vibrations.

13268 https://doi.org/10.1021/acs.jpcc.1c02131
J. Phys. Chem. C 2021, 125, 13264−13272
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Hence, the system should rather be considered as a Pd surface monolayer coverage, most anions are located in the interior of
that is partially fully covered and partially empty. From a the islands. The tendency of the IL islands to contract will
technical point of view, this affects the MD simulations of cause individual anions to adopt an upright adsorption motif
submonolayer systems, as they suffer from an unfavorable (“2−0”), at least if the islands are sufficiently large. This
dependence on the simulation box size. With increasing box scenario explains why we start to see the spectral characteristics
size, we do not observe identical structures of homogeneously of the full monolayer already before full saturation is reached.
distributed IL islands of similar size, but rather structures of In a macroscopic picture, the behavior indicates the formation
one dominating island and one dominating void. Unfortu- of 3D structures, i.e., droplets, and partial dewetting. In
nately, the computationally accessible box sizes are not big summary, we suggest the formation of a monolayer, which is
enough to reach size consistency for such submonolayers. less ordered but still affected by the substrate in terms of
Accordingly, we cannot describe the most favorable adsorption orientation and conformation. The monolayer IL structure
motif or the distribution of prevalent motifs. comprises several adsorption motifs. Changes of the adsorption
The tendency to island growth is also indicated by the motif are driven by the growth of islands, which, in turn,
estimation of adsorption energy (the list of adsorption energies changes the prevailing binding mode. In specific, we observe a
is given in Table S5). The adsorption energy is lowest for transition from species mostly interacting with the surface in a
isolated species on the surface and increases significantly for “2−2” configuration to species which favor an upright
smaller cell sizes. These numbers indicate strong intermo- adsorption motif due to increasing intermolecular interactions
lecular interactions leading to the formation of islands instead (“2−0”, “2−1”).
of an evenly distributed film, which is in good agreement with As the layer thickness increases beyond the monolayer, we
the MD data. observe a linear increase of all bands, and the spectra more and
As the coverage increases, the signals of ν(SO2) at more resemble the ATR data published in previous work.33
1104 cm−1 disappear. The peak at 1233 cm−1 grows steadily This indicates the growth of a multilayer in random
and shifts to 1212 cm−1, while new signals are observed at orientation. The MD simulations show clearly, however, that
1359, 1139, and 1063 cm−1. As a result, we observe a spectrum the orientation is not lost for molecules interacting directly
that now rather resembles a combination of the configurations with the Pd surface. In the presence of a liquid IL phase on the
denoted as “2−1” and “2−0”, as suggested by DFT calculations Pd surface, anions at the IL/Pd interface preferably obtain the
for a monolayer coverage of the IL. Furthermore, the observed “2−2” adsorption motif previously observed for a submono-
spectrum is similar to the spectrum of the multilayer, but layer. In contrast to the monolayer, where we observed around
signals differ in shape and intensity. In particular, the signal at 40% of the anions in an upright adsorption motif (“2−0”), the
1139 cm−1 is far more intense and the one at 1359 cm−1 is interface layer between Pd and the IL liquid bulk phase shows
broader and weaker than in the multilayer. We assign the only around 5% of anions in this configuration (for detailed
signals at 1359, 1212, 1139, and 1063 cm−1 to ν(SO2)as, information about the distribution of the adsorption motifs, see
ν(NTf2−), ν(SO2)s, and ν(SNS)as, respectively. The strong the Supporting Information). Instead, the “flat” and spatially
symmetric signal of the SO2 group indicates that the oxygen more demanding motifs (“2−2”, “2−1”, “1−1” cis, and “1−1”
atoms remain mostly parallel with the surface, with a smaller trans) are more abundant in the presence of the bulk IL. These
contribution of tilted molecules giving rise to the signal of the results point to a packing effect: Within the monolayer,
ν(SO2)as mode. The appearance of ν(SNS)as indicates that individual anions leave their preferred “2−2” adsorption motif
species with uncoordinated SO2 groups are present. The loss of to allow further ions to adsorb directly on the Pd surface
species in “2−2” configuration is further supported by the instead of “floating” individually on the monolayer. The latter
absence of the characteristic signal of ν(SO2)s at 1104 cm−1. situation would be disfavored in terms of the Coulomb
To some extent the effect may, however, also be enhanced by interactions between ions.
intensity transfer.58,59 The situation is different if a liquid bulk IL phase is present.
Again, we rely on DFT spectra to support our assignments. Here, excess ions that detach from the monolayer are stabilized
We suggest the following structures upon comparison of by the bulk phase. Hence, the transition of the ML to the
theoretical and experimental data. Upon saturation of the interface layer between Pd and the IL bulk phase upon higher
monolayer, the molecules remain adsorbed via one SO2 group coverage is characterized by a density decrease (from 1.35 to
(“2−j”), but the other group is either uncoordinated (“2−0”) 1.16 anions nm−2). Noteworthy, we found a similar behavior
or adsorbed to the surface via only one oxygen in a for the cations, which tend to adsorb preferably in an
monodentate fashion (“2−1”). Such species are twisted and orientation that saves space on the surface. The majority of
strongly tilted with respect to the surface. To obtain additional cations point with their C4 tail away from the surface in the
information about the expected abundance of these species, we ML film, while this is not the case in the presence of the bulk
performed MD simulations. For the full IL monolayer, our IL phase (see Figure 4 for adsorption motifs).
simulations indicate the presence of around 40% of anions Lastly, we consider the evidence for the formation of a
bound in an upright adsorption motif (“2−0”), in good closed monolayer on the Pd(111) surface. During the film
agreement with the scenario described above. For a detailed growth, we first observe the growth of distinctive spectroscopic
discussion of the adsorption motifs and their abundance we signatures of a chemisorbed species, indicating interactions of
refer to the Supporting Information and Figure S7. the SO2 groups with the surface. This is followed by a sudden
The spectral changes around 0.5 ML might be associated transition toward the spectroscopic features assigned to the
with the growth of IL islands and a corresponding change of growth of a randomly oriented multilayer. This transition is
the ratio of anions being located at the edge and in the interior visualized in plots relating peak height to time (Figure 5, right
of IL islands, respectively. We assume that at the low coverage column).
the IL islands are isolated and comparably small so that most Related experiments employed the Stark effect60,61 on
molecules are located at the edges of these islands. Near adsorbed CO to prove the formation of an intact IL
13269 https://doi.org/10.1021/acs.jpcc.1c02131
J. Phys. Chem. C 2021, 125, 13264−13272
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

monolayer.31,33 Also, STM and X-ray photoelectron spectros- 5d). In the Supporting Information, we qualitatively discuss
copy (XPS) studied showed formation of a chemisorbed the adsorption of carbon species to defects, edges, and (100)
monolayer (wetting layer) on metal surfaces (see ref 62 and facets. We observe identical signals for both HOPG samples
references therein). Only after completion of the monolayer, during the growth of an IL submonolayer and monolayer
either 2D or 3D growth is observed. (assignment in the previous paragraph). However, the signals
Growth of [C4C1Pyr][NTf2] on Pd NPs on HOPG. Next, of the submonolayer are visible for a shorter time only, and the
we turn to the IL layers on Pd NPs. The Pd NPs feature mostly transition to the multilayer growth is less distinct. We attribute
(111) facets but are rich in defects as compared to a single both effects to the defect sites already partly occupied by
crystal surface. In Figure 5c, we show the IR data recorded carbon, making them less attractive for the IL. While we
during deposition of the IL on Pd NPs onto HOPG. A cannot describe these effects on an atomic level, we conclude
reference experiment, investigating the IL growth on clean that the presence of carbon species gives rise to less specific
HOPG, can be found in previous work.33 Because the signals binding and disorder in the monolayer. As for all other
of the multilayer spectrum on Pd NPs were discussed in the samples, the multilayer is not affected by the modification of
first section, we focus on the monolayer region in the the substrate and mainly shows molecules of random
following. orientation.
In the first few spectra, we observe bands at 1220, 1370, and Concludingly, similar adsorption motifs are observed on all
1116 cm−1 which are similar to those obtained from the three surfaces, but the prevailing adsorption motif is dependent
experiment on the Pd(111) surface. The peak at 1220 cm−1 on the IL coverage and morphology of the support. In general,
grows linearly and is assigned to coupled vibrations of [NTf2]−. the presence of smaller facets and of defects leads to greater
The latter two are assigned to ν(SO2)as and ν(SO2)s, disorder of the IL monolayer film.
respectively. They are visible for a short period only and
disappear at larger coverage. This behavior indicates the
presence of a temporary adsorption structure similar to that in
■ CONCLUSIONS
In conclusion, we investigated the assembly of a Pd-based
the submonolayer regime on Pd(111). Nevertheless, the model SCILL system. We combined theoretical insights
signals are less well-defined than on Pd(111), which points obtained by DFT and MD calculations with experimental
to differences in the adsorption geometry. We attribute this data acquired via time-resolved IRAS experiments performed
effect to the absence of large (111) facets which prevents the under UHV conditions. To this end, we deposited the room-
formation of large IL islands and adsorption in “2−2” temperature IL [C4C1Pyr][NTf2] onto Pd(111) and Pd NPs
configuration. Note that we also need to consider the MSSR, supported on HOPG. The main conclusions of the study are
which largely holds on HOPG (see ref 63 for details). Species summarized in the following:
adsorbed to tilted facets will show bands that were previously
forbidden. (1) Adsorption of isolated IL species: The growth of the IL
We thus combine the implications of the MSSR with the film on Pd(111) indicates that there are strong
coverage-dependent development of selected bands (ν(SO2)as intermolecular interactions with the IL submonolayer.
and ν(SO2)s), which show characteristic differences in the The [NTf2]− anion adopts cis configuration and adsorbs
DFT-derived spectra dependent on the adsorption motif. In specifically via both SO2 groups (“2−2” adsorption motif
particular, the bands of ν(SO2)as and ν(SO2)s are visible prevalent) to the Pd surface. The IL forms monolayer
simultaneously, which points to the presence of different types islands which rapidly grow with increasing coverage.
of SO2 groups, either in parallel or tilted with respect to the (2) Monolayer growth: Near monolayer coverage, the IL
surface. As a result, we tentatively assign the bands observed at adsorption motif within these islands changes. We
low coverage to adsorption at defect sites, which leads to identify mono- and bidentate binding motifs, in which
twisted, less space-consuming conformers (mostly “2−1”, less one SO2 group is lifted or partially lifted from the surface
“2−0”) along with a minor amount of flat-lying “2−2” anions. (“2−0”, “2−1” adsorption motifs prevalent). We
In particular, the “2−1” adsorption geometry is favored by attribute the change in the adsorption motif to the
smaller facets and abundant defect sites. formation of larger islands in which the IL layer is
With coverage increasing further, we observe signals of compressed. On Pd NPs, the absence of large facets
ν(SO2)as, ν(SO2)s, and ν(SNS)as at 1361, 1141, and hinders the formation of well-defined structure. Thus,
1064 cm−1, respectively. Several of the observed vibrations the adsorption motifs of the [NTf2]− are less specific
are polarized orthogonally in the molecule, which indicates a than on Pd(111).
loss of orientational order in the layer. For this reason, we
assign the signals to species binding to the surface in a less (3) Multilayer growth: Upon transition to the multilayer
defined orientation, involving various motifs associated with film, MD suggests a reorientation of the chemisorbed
monodentate and bidentate structures (mostly “2−0”, less “2− species on the Pd surface in the presence of the IL bulk.
1” and “2−2”). While the characteristic growth of islands is not We propose that this effect originates from the fact that
visible as clearly as on the Pd(111), there still is some ions expelled from the monolayer can be easily
accommodated in the multilayer film. For all substrates,
indication for a reorientation of the molecules as a function of
the preferential molecular orientation is lost for the
coverage (see the integrated peak area). After completion of
physisorbed species on top of the first monolayer.
the monolayer, the growth continues with the formation of a
multilayer consisting of molecules with nearly random Our results show that the adsorption motif and orientation
orientation. of the IL in the first monolayer depend strongly on the local IL
As a third sample, we investigated HOPG loaded with a coverage. Furthermore, the surface morphology also plays an
similar amount of Pd (15 Å) and annealed it to 500 K to important role. These findings will be of importance when the
induce the migration of carbon into/onto the Pd NPs (Figure IL film is used as a catalytic modifier. We expect that the effect
13270 https://doi.org/10.1021/acs.jpcc.1c02131
J. Phys. Chem. C 2021, 125, 13264−13272
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

of the IL on coadsorbed reactants and their reactivity will SPP 1708 “Material Synthesis near Room Temperature”
critically depend on the local IL coverage. (Project-ID 252578361).


*
ASSOCIATED CONTENT
sı Supporting Information
■ REFERENCES
(1) Welton, T. Ionic liquids in catalysis. Coord. Chem. Rev. 2004,
The Supporting Information is available free of charge at 248, 2459−2477.
https://pubs.acs.org/doi/10.1021/acs.jpcc.1c02131. (2) Wasserscheid, P.; Welton, T. Ionic Liquids in Synthesis, 2nd ed.;
Full set of DFT spectra; visualization of the binding Wiley: New York, 2007; Vol. 1.
motifs; list of adsorption energies; description of the (3) Steinrück, H.-P.; Wasserscheid, P. Ionic Liquids in Catalysis.
MD simulation procedure (PDF) Catal. Lett. 2015, 145, 380−397.


(4) Werner, S.; Haumann, M.; Wasserscheid, P. Ionic Liquids in
Chemical Engineering. Annu. Rev. Chem. Biomol. Eng. 2010, 1, 203−
AUTHOR INFORMATION 230.
Corresponding Author (5) Riisager, A.; Fehrmann, R.; Haumann, M.; Wasserscheid, P.
Tanja Bauer − Interface Research and Catalysis, ECRC, Supported Ionic Liquid Phase (SILP) catalysis: An Innovative
Friedrich-Alexander-Universität Erlangen-Nürnberg, 91058 Concept for Homogeneous Catalysis in Continuous Fixed-Bed
Reactors. Eur. J. Inorg. Chem. 2006, 2006, 695−706.
Erlangen, Germany; orcid.org/0000-0002-6399-2954;
(6) Mehnert, C. P. Supported Ionic Liquid Catalysis. Chem. - Eur. J.
Email: tanja.tb.bauer@fau.de 2005, 11, 50−56.
Authors (7) Kernchen, U.; Etzold, B. J. M.; Korth, W.; Jess, A. Solid Catalyst
with Ionic Liquid Layer (SCILL) − A New Concept to Improve
Christian Schuschke − Interface Research and Catalysis, Selectivity Illustrated by Hydrogenation of Cyclooctadiene. Chem.
ECRC, Friedrich-Alexander-Universität Erlangen-Nürnberg, Eng. Technol. 2007, 30, 985−994.
91058 Erlangen, Germany; orcid.org/0000-0002-5635- (8) Arras, J.; et al. How a Supported Metal is Influenced by an Ionic
1112 Liquid: In-depth Characterization of SCILL-Type Palladium Catalysts
Lukas Fromm − Lehrstuhl für Theoretische Chemie, Friedrich- and Their Hydrogen Adsorption. J. Phys. Chem. C 2010, 114, 10520−
Alexander-Universität Erlangen-Nürnberg, 91058 Erlangen, 10526.
Germany (9) Barth, T.; Korth, W.; Jess, A. Selectivity-Enhancing Effect of a
Johannes Träg − Lehrstuhl für Theoretische Chemie, CCC, SCILL Catalyst in Butadiene Hydrogenation. Chem. Eng. Technol.
Friedrich-Alexander-Universität Erlangen-Nürnberg, 91052 2017, 40, 395−404.
Erlangen, Germany (10) Szesni, N.; et al. Catalyst Composition for Selective
Corinna Stumm − Interface Research and Catalysis, ECRC, Hydrogenation with Improved Characteristics. U.S. Patent
WO2013057244A1, 2013.
Friedrich-Alexander-Universität Erlangen-Nürnberg, 91058 (11) Welton, T. Room-Temperature Ionic Liquids. Solvents for
Erlangen, Germany; orcid.org/0000-0003-4646-8234 Synthesis and Catalysis. Chem. Rev. 1999, 99, 2071.
Chantal Hohner − Interface Research and Catalysis, ECRC, (12) Wasserscheid, P.; Keim, W. Ionic LiquidsNew “Solutions”
Friedrich-Alexander-Universität Erlangen-Nürnberg, 91058 for Transition Metal Catalysis. Angew. Chem., Int. Ed. 2000, 39, 3772−
Erlangen, Germany; orcid.org/0000-0003-0803-147X 3789.
Roman Eschenbacher − Interface Research and Catalysis, (13) Maier, F.; et al. Insights into the Surface Composition and
ECRC, Friedrich-Alexander-Universität Erlangen-Nürnberg, Enrichment Effects of Ionic Liquids and Ionic Liquid Mixtures. Phys.
91058 Erlangen, Germany Chem. Chem. Phys. 2010, 12, 1905−1915.
Sebastian Grau − Interface Research and Catalysis, ECRC, (14) Lovelock, K. R. J.; et al. Influence of Different Anions on the
Friedrich-Alexander-Universität Erlangen-Nürnberg, 91058 Surface Composition of Ionic Liquids Studied Using ARXPS. J. Phys.
Chem. B 2009, 113, 2854−2864.
Erlangen, Germany
(15) Schernich, S.; et al. Interactions of Imidazolium-Based Ionic
Dirk Zahn − Lehrstuhl für Theoretische Chemie, CCC, Liquids with Oxide Surfaces Controlled by Alkyl Chain Functional-
Friedrich-Alexander-Universität Erlangen-Nürnberg, 91052 ization. ChemPhysChem 2013, 14, 3673−3677.
Erlangen, Germany (16) Xu, T.; et al. Gluing Ionic Liquids to Oxide Surfaces: Chemical
Andreas Görling − Lehrstuhl für Theoretische Chemie, Anchoring of Functionalized Ionic Liquids by Vapor Deposition onto
Friedrich-Alexander-Universität Erlangen-Nürnberg, 91058 Cobalt (II) Oxide. Angew. Chem. 2017, 129, 13310.
Erlangen, Germany; orcid.org/0000-0002-1831-3318 (17) Bauer, T.; et al. Palladium-Mediated Ethylation of the
Jörg Libuda − Interface Research and Catalysis, ECRC, Imidazolium Cation Monitored In Operando on a Solid Catalyst
Friedrich-Alexander-Universität Erlangen-Nürnberg, 91058 with Ionic Liquid Layer. ChemCatChem 2017, 9, 109−113.
Erlangen, Germany; orcid.org/0000-0003-4713-5941 (18) Steinrück, H.-P. Surface Science Goes Liquid ! Surf. Sci. 2010,
604, 481−484.
Complete contact information is available at: (19) Steinrück, H.-P.; et al. Surface Science and Model Catalysis
https://pubs.acs.org/10.1021/acs.jpcc.1c02131 with Ionic Liquid-Modified Materials. Adv. Mater. 2011, 23, 2571−
2587.
Notes (20) Cremer, T.; et al. Physical Vapor Deposition of [EMIM]-
The authors declare no competing financial interest. [Tf2N]: A New Approach to the Modification of Surface Properties


with Ultrathin Ionic Liquid Films. ChemPhysChem 2008, 9, 2185−
ACKNOWLEDGMENTS 2190.
(21) Xu, T.; et al. Interaction of Ester-Functionalized Ionic Liquids
This work was funded by the Deutsche Forschungsgemein- with Atomically-Defined Cobalt Oxides Surfaces: Adsorption,
schaft (DFG, German Research Foundation), Project-ID Reaction and Thermal Stability. ChemPhysChem 2017, 18, 3443−
431791331, SFB 1452. Furthermore, the authors acknowledge 3453.
support by the DFG via the projects 322419553 and (22) Wasserscheid, P. Volatile Times for Ionic Liquids. Nature 2006,
453560721 and additional support by the DFG within the 439, 797.

13271 https://doi.org/10.1021/acs.jpcc.1c02131
J. Phys. Chem. C 2021, 125, 13264−13272
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(23) Leal, J. P.; et al. The Nature of Ionic Liquids in the Gas Phase. (46) Hanwell, M. D.; et al. Avogadro: An Advanced Semantic
J. Phys. Chem. A 2007, 111, 6176−6182. Chemical Editor, Visualization, And Analysis Platform. J. Cheminf.
(24) Armstrong, J. P.; et al. Vapourisation of Ionic Liquids. Phys. 2012, 4, 17.
Chem. Chem. Phys. 2007, 9, 982−990. (47) Laurin, M. QVibeplot: A Program to Visualize Molecular
(25) Freund, H. J.; et al. Preparation and Characterization of Model Vibrations In Two Dimensions. J. Chem. Educ. 2013, 90, 944−946.
Catalysts: From Ultrahigh Vacuum to in Situ Conditions at the (48) Plimpton, S. Fast Parallel Algorithms for Short-Range
Atomic Dimension. J. Catal. 2003, 216, 223−235. Molecular Dynamics. J. Comput. Phys. 1995, 117, 1−19.
(26) Libuda, J. Reaction Kinetics on Model Catalysts: Molecular (49) Brown, W. M.; Wang, P.; Plimpton, S. J.; Tharrington, A. N.
Beam Methods and Time-Resolved Vibrational Spectroscopy. Surf. Implementing Molecular Dynamics on Hybrid High Performance
Sci. 2005, 587, 55−68. Computers − Short Range Forces. Comput. Phys. Commun. 2011, 182,
(27) Mehl, S.; et al. Ionic-Liquid-Modified Hybrid Materials 898−911.
Prepared by Physical Vapor Codeposition: Cobalt and Cobalt (50) Brown, W. M.; Kohlmeyer, A.; Plimpton, S. J.; Tharrington, A.
N. Implementing Molecular Dynamics on Hybrid High Performance
Oxide Nanoparticles in [C1C2Im][OTf] Monitored by in Situ IR
Computers − Particle−Particle Particle-Mesh. Comput. Phys.
Spectroscopy. Langmuir 2016, 32, 8613−8622.
Commun. 2012, 183, 449−459.
(28) Cremer, T.; et al. Interfaces of Ionic Liquids and Transition
(51) Brown, W. M.; Yamada, M. Implementing Molecular Dynamics
Metal SurfacesAdsorption, Growth, and Thermal Reactions of on Hybrid High Performance ComputersThree-Body Potentials.
Ultrathin [C1C1Im][Tf2N] Films on Metallic and Oxidised Ni(111) Comput. Phys. Commun. 2013, 184, 2785−2793.
Surfaces. Phys. Chem. Chem. Phys. 2012, 14, 5153−5163. (52) Harvey, S. C.; Tan, R. K.-Z.; Cheatham, T. E. The Flying Ice
(29) Sobota, M.; et al. Toward Ionic-Liquid-Based Model Catalysis: Cube: Velocity Rescaling in Molecular Dynamics Leads to Violation
Growth, Orientation, Conformation, and Interaction Mechanism of of Energy Equipartition. J. Comput. Chem. 1998, 19, 726−740.
the [Tf2N]− Anion in [BMIM][Tf2N] Thin Films on a Well-Ordered (53) Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D.
Alumina Surface. Langmuir 2010, 26, 7199−7207. A. Development and Testing of a General Amber Force Field. J.
(30) Schernich, S.; et al. Interactions Between the Room- Comput. Chem. 2004, 25, 1157−1174.
Temperature Ionic Liquid [C2C1Im][OTf] and Pd(111), Well- (54) Bayly, C. I.; Cieplak, P.; Cornell, W.; Kollman, P. A. A Well-
Ordered Al2O3, and Supported Pd Model Catalysts from IR Behaved Electrostatic Potential Based Method Using Charge
Spectroscopy. J. Phys. Chem. C 2014, 118, 3188−3193. Eestraints for Deriving Atomic Charges: The RESP Model. J. Phys.
(31) Bauer, T.; et al. Ligand Effects at Ionic Liquid-Modified Chem. 1993, 97, 10269−10280.
Interfaces: Coadsorption of [C2C1Im][OTf] and CO on Pd(111). J. (55) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Phys. Chem. C 2016, 120, 4453−4465. et al. Gaussian 09, rev. D.01; Gaussian, Inc.: Wallingford, CT, 2013.
(32) Sobota, M.; et al. Ligand Effects in SCILL Model Systems: Site- (56) Heinz, H.; Vaia, R. A.; Farmer, B. L.; Naik, R. R. Accurate
Specific Interactions with Pt and Pd Nanoparticles. Adv. Mater. 2011, Simulation of Surfaces and Interfaces of Face-Centered Cubic Metals
23, 2617−2621. Using 12−6 and 9−6 Lennard-Jones Potentials. J. Phys. Chem. C
(33) Schuschke, C.; et al. Dynamic CO Adsorption and Desorption 2008, 112, 17281−17290.
Through the Ionic Liquid Layer of a Pt Model Solid Catalyst with (57) Buraschi, M.; Sansotta, S.; Zahn, D. Polarization Effects in
Ionic Liquid Layers. J. Phys. Chem. C 2019, 123, 31057−31072. Dynamic Interfaces of Platinum Electrodes and Ionic Liquid Phases:
(34) Schuschke, C.; et al. Solar Energy Storage at an Atomically A Molecular Dynamics Study. J. Phys. Chem. C 2020, 124, 2002−
2007.
Defined Organic-Oxide Hybrid Interface. Nat. Commun. 2019, 10,
(58) Hoffmann, F. M. Infrared Reflection-Absorption Spectroscopy
2384.
of Adsorbed Molecules. Surf. Sci. Rep. 1983, 3, 107.
(35) Kettner, M.; Stumm, C.; Schwarz, M.; Schuschke, C.; Libuda, J.
(59) Hollins, P. The Influence of Surface Defects on the Infrared
Pd Model Catalysts on Clean and Modified HOPG: Growth, Spectra of Adsorbed Species. Surf. Sci. Rep. 1992, 16, 51−94.
Adsorption Properties, and Stability. Surf. Sci. 2019, 679, 64−73. (60) Boxer, S. G. Stark Realities. J. Phys. Chem. B 2009, 113, 2972−
(36) Kettner, M.; et al. Pd-Ga model SCALMS: Characterization 2983.
and Stability of Pd single Atom Sites. J. Catal. 2019, 369, 33−46. (61) Lambert, D. K. Vibrational Stark Effect of Adsorbates at
(37) Nečas, D.; Klapetek, P. Gwyddion: An Open-Source Software Electrochemical Interfaces. Electrochim. Acta 1996, 41, 623−630.
for SPM Data Analysis. Cent. Eur. J. Phys. 2012, 10. (62) Lexow, M.; Maier, F.; Steinrück, H.-P. Ultrathin Ionic Liquid
(38) Kresse, G.; Furthmüller, J. Efficiency Of Ab-Initio Total Energy Films on Metal Surfaces: Adsorption, Growth, Stability and Exchange
Calculations For Metals and Semiconductors Using A Plane-Wave Phenomena. Adv. Phys. X 2020, 5, 1761266.
Basis Set. Comput. Mater. Sci. 1996, 6, 15−50. (63) Heidberg, J.; Warskulat, M.; Folman, M. Fourier-Transform-
(39) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Infrared Spectroscopy of Carbon Monoxide Physisorbed on Highly
Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Oriented Graphite. J. Electron Spectrosc. Relat. Phenom. 1990, 54,
Rev. B: Condens. Matter Mater. Phys. 1996, 54, 11169−11186. 961−970.
(40) Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid
Metals. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 558−561.
(41) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the
Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter
Mater. Phys. 1999, 59, 1758−1775.
(42) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient
Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868.
(43) 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.
(44) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping
Function in Dispersion Corrected Density Functional Theory. J.
Comput. Chem. 2011, 32, 1456−1465.
(45) Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone
Integrations. Phys. Rev. B 1976, 13, 5188−5192.

13272 https://doi.org/10.1021/acs.jpcc.1c02131
J. Phys. Chem. C 2021, 125, 13264−13272

You might also like