63
doi.org/10.26434/chemrxiv.13643579.v1 Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters: Similar Absorption Spectra Disguise Distinct Geometries and Electronic Structures Yavuz S. Ceylan, Rebecca Gieseking Submitted date: 26/01/2021 Posted date: 27/01/2021 Licence: CC BY-NC 4.0 Citation information: Ceylan, Yavuz S.; Gieseking, Rebecca (2021): Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters: Similar Absorption Spectra Disguise Distinct Geometries and Electronic Structures. ChemRxiv. Preprint. https://doi.org/10.26434/chemrxiv.13643579.v1 Ligands can dramatically affect the electronic structure of gold nanoclusters (NCs) and provide a useful handle to tune the properties required for nanomaterials that have high performance for important functions like catalysis. Recently, questions have arisen about the nature of the interactions of hydride and halide ligands with Au NCs: hydride and halide ligands have similar effects on the absorption spectra of Au NCs, which suggested that the interactions of the two classes of ligands with the Au core may be similar. Here, we elucidate the interactions of halide and hydride ligands on phosphine-protected gold clusters via theoretical investigations. The computed absorption spectra using time-dependent density functional theory are in reasonable agreement with the experimental spectra, confirming that the computational methods are capturing the ligand-metal interactions accurately. Despite the similarities in the absorption spectra, the hydride and halide ligands have distinct geometric and electronic effects. The hydride ligand behaves as a metal dopant and contributes its two electrons to the number of superatomic electrons, while the halides act as electron-withdrawing ligands and do not change the number of superatomic electrons. Clarifying the binding modes of these ligands will aid in future efforts to use ligand derivatization as a powerful tool to rationally design Au NCs for use in functional materials. File list (2) download file view on ChemRxiv Au9_MS_2021-01-21_RMG.pdf (1.59 MiB) download file view on ChemRxiv Au9_SI_2021-01-21_RMG.pdf (2.16 MiB)

Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

doi.org/10.26434/chemrxiv.13643579.v1

Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters: SimilarAbsorption Spectra Disguise Distinct Geometries and ElectronicStructuresYavuz S. Ceylan, Rebecca Gieseking

Submitted date: 26/01/2021 • Posted date: 27/01/2021Licence: CC BY-NC 4.0Citation information: Ceylan, Yavuz S.; Gieseking, Rebecca (2021): Hydride- and Halide-SubstitutedAu9(PH3)83+ Nanoclusters: Similar Absorption Spectra Disguise Distinct Geometries and ElectronicStructures. ChemRxiv. Preprint. https://doi.org/10.26434/chemrxiv.13643579.v1

Ligands can dramatically affect the electronic structure of gold nanoclusters (NCs) and provide a usefulhandle to tune the properties required for nanomaterials that have high performance for important functionslike catalysis. Recently, questions have arisen about the nature of the interactions of hydride and halideligands with Au NCs: hydride and halide ligands have similar effects on the absorption spectra of Au NCs,which suggested that the interactions of the two classes of ligands with the Au core may be similar. Here, weelucidate the interactions of halide and hydride ligands on phosphine-protected gold clusters via theoreticalinvestigations. The computed absorption spectra using time-dependent density functional theory are inreasonable agreement with the experimental spectra, confirming that the computational methods arecapturing the ligand-metal interactions accurately. Despite the similarities in the absorption spectra, thehydride and halide ligands have distinct geometric and electronic effects. The hydride ligand behaves as ametal dopant and contributes its two electrons to the number of superatomic electrons, while the halides actas electron-withdrawing ligands and do not change the number of superatomic electrons. Clarifying thebinding modes of these ligands will aid in future efforts to use ligand derivatization as a powerful tool torationally design Au NCs for use in functional materials.

File list (2)

download fileview on ChemRxivAu9_MS_2021-01-21_RMG.pdf (1.59 MiB)

download fileview on ChemRxivAu9_SI_2021-01-21_RMG.pdf (2.16 MiB)

Page 2: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

1

Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters: Similar

Absorption Spectra Disguise Distinct Geometries and Electronic Structures

Yavuz S. Ceylan, Rebecca L. M. Gieseking*

Department of Chemistry, Brandeis University

415 South Street, Waltham, Massachusetts 02453

* Corresponding author: [email protected]; (781)-736-2511

Page 3: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

2

Abstract

Ligands can dramatically affect the electronic structure of gold nanoclusters (NCs) and provide a

useful handle to tune the properties required for nanomaterials that have high performance for

important functions like catalysis. Recently, questions have arisen about the nature of the

interactions of hydride and halide ligands with Au NCs: hydride and halide ligands have similar

effects on the absorption spectra of Au NCs, which suggested that the interactions of the two

classes of ligands with the Au core may be similar. Here, we elucidate the interactions of halide

and hydride ligands on phosphine-protected gold clusters via theoretical investigations. The

computed absorption spectra using time-dependent density functional theory are in reasonable

agreement with the experimental spectra, confirming that the computational methods are capturing

the ligand-metal interactions accurately. Despite the similarities in the absorption spectra, the

hydride and halide ligands have distinct geometric and electronic effects. The hydride ligand

behaves as a metal dopant and contributes its two electrons to the number of superatomic electrons,

while the halides act as electron-withdrawing ligands and do not change the number of superatomic

electrons. Clarifying the binding modes of these ligands will aid in future efforts to use ligand

derivatization as a powerful tool to rationally design Au NCs for use in functional materials.

Page 4: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

3

1. Introduction

Atomically precise noble metal nanoclusters (NCs) composed of a metal core surrounded by

ligands have been studied widely in recent years for their catalytic activity for industrially

important reactions,1–8 particularly as photocatalysts.9–13 Although gold nanoparticles and

nanostructures have plasmonic properties that reflect a continuum of electronic states, gold NCs

with diameters less than a few nm have discrete orbital energies and exciton-like excited-state

properties.14,15 For this reason, gold NCs often have higher photocatalytic activity than larger

nanoparticles.16,17 Tuning the electronic structures of gold NCs is a critical way to control their

catalytic activity.8,18 The precise structural control of noble metal NCs has enabled detailed study

of catalytic reaction mechanisms19,20 and elucidation of the active sites,21–23 as well as investigation

of the effects of structural changes as small as single atoms.22,24–26

Ligands play a critical role in determining the electronic structure and catalytic activity of Au NCs.

The ligands can dramatically affect many properties, including which NC sizes are stable, the

geometric structure of the Au core, and the NC electronic structure.27,28 Ligand substitution can

dramatically change the activity and selectivity of catalytic reactions,29–31 as well as the reaction

mechanism.32 Deliberate choice of ligand shapes33 and ligand-metal bond strengths34 can promote

the formation of undercoordinated metal atoms that act as active sites for catalysis. Thus,

understanding the connection between the choice of ligand, the NC size and geometry, and the

chemical properties of the resulting NCs is a necessary step toward designing NCs with properties

that are tuned to enhance their catalytic or photocatalytic activity.30,35–38 A detailed understanding

Page 5: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

4

of the geometric and electronic properties of the ligands is key to developing structure/function

relationships in NC-based nanomaterials.

Several classes of ligands have been shown to have different binding modes and electronic effects

on Au NCs. Thiolate ligands typically form “staple” motifs (-SR-Au-SR- or similar) on the NC

surface; in contrast, alkene, phosphine, and halide ligands typically bind directly to the Au core.39

The electronic effects of the ligands are also quite distinct. The electronic structure of Au NCs can

be understood in terms of a superatomic model,40–42 which has been widely used to understand and

predict NC properties.43–50 In this model, the Au valence orbitals (6s and 6p) create superatomic

molecular orbitals (MOs) that are delocalized across the metal core with shapes that resemble those

of atomic orbitals (1S, 1P, 1D, 2S, etc.); ‘magic number’ occupations that correspond to filled

superatomic shells have high stability. A bare, neutral Au NC has one superatomic electron per

Au atom. Ligands such as thiolates and halides typically act as electron-withdrawing ligands, and

each ligand reduces the number of electrons in the superatomic Au core by one; in contrast, weaker

ligands like phosphines do not modify the number of superatomic electrons in the Au core.40

Hydride ligands have been of particular interest in recent years because of their unique interactions

with the Au NC. Binding of hydrogen to NCs is a critical step in catalysis of reactions such as

hydrogen evolution,51 and some Au NCs bind hydrogen more strongly than well-known Pt

catalysts.52 Early photoelectron experiments suggested that hydrogen atoms behaved like

monovalent metal atoms when adsorbed on bare Au NCs;53 these results were supported by DFT

calculations of the electronic structures of similar NCs.54 Based on these results, the hydride ligand

has been described as a metal dopant rather than as a traditional ligand. In recent years, this

Page 6: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

5

description of hydrogen as a metal dopant has been extended to ligand-protected Au NCs based

on both computational55–57 and experimental55,56,58 evidence. Unlike typical electron-withdrawing

and weak ligands, the two electrons of the hydride are added to the superatomic electron count,

increasing the total number of electrons in superatomic orbitals.

The Au9(PPh3)83+ NC is an accessible platform to understand the behavior of halide and hydride

ligands. Au9(PPh3)83+ was one of first Au NCs discovered and is accessible in gram-scale

quantities,59 and the NC has been widely studied using both experimental and theoretical

tools.46,56,58,60,61 Two distinct geometries of the NC have been observed in crystal structures,

depending on the choice of anion and solvent: the metal core may have either approximate D4d

symmetry62 or approximate D2h symmetry.63,64 In both geometries, the central Au atom is bonded

only to eight outer Au atoms, making it a low-coordination site that is accessible as a binding site

for an additional hydride ligand.55 In the gas phase, this NC can undergo collision-induced

interconversion between forms with different effective sizes.65

Recently, Johnson et al. showed that the Au9(PPh3)8H2+, Au9(PPh3)8Cl2+, and Au9(PPh3)8Br2+ NCs

have surprisingly similar gas-phase absorption spectra using a technique that couples mass

spectroscopy to electronic absorption spectra (Figure 1).66 In contrast with the unsubstituted

Au9(PPh3)83+ NC, all three ligand-substituted NCs have similar absorption spectra between 2.25

and 3.40 eV, with plateaus in their absorption spectra across a 0.5 eV range. The similarities are

unexpected because the hydride and halide ligands are typically thought to have different binding

modes with the Au core. Johnson et al. proposed that the similarities in the absorption spectra are

due to similarities in the electronic structures of the ligand-substituted NCs: either hydride,

Page 7: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

6

chloride, and bromide all behave as metal dopants, or all three ligands are all electron-withdrawing.

Either of these interpretations would substantially modify the current model of ligand behavior in

Au NCs.

Figure 1. Experimental absorption spectra of (a) Au9(PPh3)83+, (b) Au9(PPh3)8H2+, (c) Au9(PPh3)8Cl2+, and (d) Au9(PPh3)8Br2+. Data is from ref.66

Here, we examine the unsubstituted Au9(PPh3)83+ NC and the hydride- and halide-substituted NCs

using density functional theory (DFT) to understand the effect of the ligands on the NC geometric

and electronic structures and absorption spectra. We show that the superficially similar absorption

spectra in these NCs occur despite substantial differences in the geometric and electronic

structures. Instead, the hydride-substituted NC has properties that suggest that the hydride ligand

Page 8: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

7

is a metal dopant that binds to the central Au atom, consistent with previous theoretical results. In

contrast, the halide ligands behave like electron-withdrawing ligands and unexpectedly bind in a

bridging position to two outer Au atoms. These results highlight the importance of detailed

theoretical study to elucidate the properties of Au NCs.

2. Computational Methods

The geometries of the Au NCs were optimized using a two-step process. Our initial geometries

were based on symmetrized versions of the D4d63 and D2h

64 crystal structures of Au9(PH3)83+, with

the experimental PPh3 ligands replaced with PH3 to reduce computational cost. Starting from each

geometry, we constructed eleven different structures of each ligand-protected Au NC with

different initial positions of the hydride or halide ligand, including structures with the ligand near

the central Au atom and structures with the ligand near the outer Au atoms. The initial

optimizations were performed using the BP86 functional67,68 and a double-ζ (DZ) Slater-type basis

set with a frozen-core approximation for the Au 1s-5p orbitals; the BP86 function has been widely

used for Au NCs69–72 and is a low computational cost GGA-type functional. In a second step, for

each ligand-substituted NC, all distinct local minima with energies less than 10.0 kcal/mol higher

than the most stable structure were re-optimized using the revTPSS73 functional with a triple-ζ

polarized (TZP) Slater-type all-electron basis set. In previous benchmarking studies, the revTPSS

functional has performed particularly well for the relative energies and geometries of Au NCs.74

For all NCs within 11.9 kcal/mol of the most stable structure in the revTPSS optimizations, the

lowest 200 excited states in dipole-allowed symmetry groups were computed using a time-

dependent DFT (TD-DFT) approach with the SAOP functional75 and TZP all-electron basis set.

Page 9: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

8

Relativistic effects were incorporated using the zeroth-order regular approximation (ZORA)76 in

both ground and excited state calculations. All calculations were performed using the Amsterdam

Density Functional (ADF) 2018 software package.77,78

The absorption spectra were simulated by applying a Lorentzian broadening to the TD-DFT stick

spectrum with a full width at half maximum (FWHM) of 0.0087. To compute the superatomic

contributions to the absorption spectra, we visually identified the occupied molecular orbitals

(MOs) with superatomic character. The superatomic character of each excited state was computed

as the weighted percentage of the excitations contributing to the state that involve superatomic

occupied MOs, consistent with our previous definitions.79 The superatomic contributions to the

absorption spectra were computed by scaling the oscillator strength of each excited state by its

superatomic character.

3. Results

We first examine the unsubstituted Au9(PPh3)83+, which has been widely studied experimentally

and computationally;46,56,58,60,61 our results are largely consistent with previous results but lay

important groundwork to understand the effects of ligand substitution. We then turn to the hydride-

substituted and halide-substituted clusters, comparing their geometries and electronic structures

and absorption spectra.

3.1. Unsubstituted Au9(PH3)83+ nanocluster

Page 10: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

9

As mentioned in the Introduction, crystal structures of Au9(PR3)83+ with both D4d and D2h

symmetry have been observed;62–64 previous theoretical work has focused on the structure with

D2h symmetry.46,49 We have computed the properties of the NC structures with both symmetries,

with the PPh3 ligands replaced with PH3 to reduce computational cost. The optimized Au9(PH3)83+

structures with these two symmetries are shown in Figure 2a,d. In both geometries, one Au atom

is located in the center of the NC and is bound to eight peripheral Au atoms; this atom is thus . At

the revTPSS/TZP level with relativistic corrections, the structure with D4d symmetry is 1.2

kcal/mol more stable than the structure with D2h symmetry.

Figure 2. (a, d) Top and side views of the optimized geometries, (b, e) HOMO and HOMO-l superatomic orbitals, and (c, f) computed absorption spectra of the (a, b, c) D4d and (d, e, f)

D2h structures of Au9(PH3)83+.

Page 11: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

10

In Au NCs, we can identify superatomic molecular orbitals (MOs), which are composed primarily

of the Au 6s and 6p atomic orbitals and are delocalized across the Au core with shapes that

resemble those of atomic orbitals.40 Au9(PR3)83+ has been shown previously to have six electrons

in superatomic orbitals,46 with occupations of 1S21P4. Our results are consistent, and the HOMO

and HOMO-1 for both the D4d and D2h symmetries correspond to the 1Py and 1Px superatomic

orbitals, respectively (Figure 2b,e). Because of the NC symmetry, the HOMO and HOMO-1

orbitals are degenerate for the D4d structure but not for D2h. For both structures, the 1Pz superatomic

orbital aligned along the short axis of the NC is higher in energy that 1Py and 1Px superatomic

orbitals because of the overall ellipsoidal NC shape, and is thus unoccupied. Mixing of the

superatomic orbitals with the Au 5d orbitals and ligand-based orbitals makes distinguishing the 1S

superatomic orbital difficult; thus, we focus here only on the 1P superatomic orbitals.

The computed absorption spectra for the two structures of the Au9(PH3)83+ NC (Figure 2c,f)

include several low-energy absorption peaks. The absorption spectra of the D2h and D4d structures

are quite distinct, and both structures have significantly fewer low-energy absorption peaks than

are visible in the experimental spectrum (Figure 1a). These differences are expected because it

has been previously shown that the phenyl groups introduce many more absorption peaks below 3

eV.46 Because the D2h structure has fewer degenerate MOs, it exhibits more distinct absorption

peaks than the D4d structure. Our computed absorption spectrum for the D2h structure is largely

consistent with the previously computed spectrum,46 with expected small shifts in the absorption

energies because of differences in the functional used.

Page 12: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

11

To understand the absorption spectra, we have also computed the contribution of the superatomic

MOs to each excited state.79 The low-lying unoccupied MOs of most Au NCs have significant

superatomic character, so the analysis is based solely on the occupied MOs in each excited state.

In the absorption spectrum of the D2h structure, the peaks at 2.63 eV and 3.87 eV have primarily

superatomic character (80% and 70%, respectively) involving excitations from the 1P superatomic

orbitals. In contrast, the peaks at 3.17 eV and 3.57 eV have < 15% superatomic character. The

peak at 3.17 eV primarily involves HOMO-2, and the peak at 3.57 eV is largely an excitation from

HOMO-5 and HOMO-6; all of these MOs involve a combination of the Au 5d and ligand atomic

orbitals. For the D4d structure, the peak at 2.87 eV is 96% superatomic, and the somewhat weaker

absorption peak at 3.57 eV has 61% superatomic character. The higher-energy absorption peaks

have low superatomic character, indicating that they arise primarily from the Au 5d and ligand

orbitals.

3.2. Hydride-substituted Au9(PH3)8H2+ nanocluster

To determine the most stable geometry of the hydride-substituted Au9(PH3)8H2+ NC, we used a

two-step process. For initial screening, we constructed and optimized eleven different structures

based on both the D2h and D4d structures of the unsubstituted NC, with the hydride ligand placed

near either the central Au atom or near one or two of the outer Au atoms; these optimizations were

performed at the BP86/DZ level. After removing duplicates in the optimized geometries, we

selected all structures within 11.93 kcal/mol of the most stable structure for reoptimization at the

revTPSS/TZP level; this functional has previously yielded accurate geometries and relative

energies for Au NCs.74 We focus here solely on the most stable geometry from this second

Page 13: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

12

optimization; this structure is 6.8 kcal/mol stable than any other structure, an order of magnitude

larger than thermal energy at room temperature. More details on the higher-energy structures are

available in the SI.

The most stable structure of Au9(PH3)8H2+ has the hydride ligand bonded directly to the low-

coordination central atom of the Au9 NC and has approximately C2v symmetry (Figure 3a). Four

of the outer Au atoms form a rectangle nearly co-planar with the central Au, and four are below

that plane. This geometry is consistent with the previously published structure of this NC.40 In this

geometry, the distance between the central Au atom and the hydride ligand is 0.19 Å shorter than

the sum of the covalent radii81 of the two atoms, and substantially shorter than the sum of the van

der Waals radii (Table 1).82 This suggests that the Au-H interaction involves strong orbital mixing,

rather than the electron-withdrawing character typical of many anionic ligands. To enable more

direct comparison to experimental results, we have also optimized Au9(PPh3)8H2+ with the full

PPh3 ligands used experimentally (Figure 3d). The geometry of Au9(PPh3)8H2+ is quite similar to

that of Au9(PH3)8H2+ and shows a 0.03 Å elongation of the Au-H bond, suggesting that the change

in the ligands has minimal effects on the nature of the metal-ligand interaction.

Page 14: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

13

Figure 3. (a, d) Optimized structures, (b, e) superatomic orbitals, and (c, f) absorption spectra of (a, b, c) Au9(PH3)8H2+ and (d, e, f) Au9(PPh3)8H2+. Ligands are excluded from

panel d for clarity.

Table 1. Au-ligand (Au-L) interatomic distances (dAu-L) and comparison to the sums of the van der Waals (𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺) and covalent (𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺) radii of the two atoms. For the halide

ligands, Aua and Aub indicate the two Au atoms closest to the ligand.

Au-L Bond dAu-L dAu-L – 𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺 dAu-L – 𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺𝚺 Au-H (PH3 ligands) 1.62 -1.24 -0.19 Au-H (PPh3 ligands) 1.65 -1.21 -0.16

Aua-F 2.24 -0.89 0.09 Aub-F 2.57 -0.56 0.42 Aua-Cl 2.66 -0.75 0.23 Aub-Cl 2.67 -0.74 0.24 Aua-Br 2.75 -0.76 0.17 Aub-Br 2.76 -0.75 0.18

Page 15: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

14

The lack of electron-withdrawing character in the Au-H bond can be confirmed by examining the

Hirshfeld atomic charges,83 which have been commonly used to understand the electronic

properties of Au NCs.48,49,83,84 The Hirshfeld charges on both the hydride and attached central Au

atom are close to zero (Table 2), which is unsurprising given that Au is slightly more

electronegative than H (electronegativities of 2.54 and 2.20, respectively). These results are

consistent with previous charge analyses of hydride-substituted Au NCs.57 Replacement of the PH3

ligands with PPh3 ligands has minimal effects on the charge distribution, which is consistent with

the small changes in the geometric structure.

Table 2. Hirshfeld charges of the hydride and halide ligands and the Au atoms bound to those ligands. For the halide ligands, Aua and Aub indicate the two Au atoms closest to the

ligand.

H (PH3 ligands) H (PPh3 ligands) F Cl Br Ligand -0.082 -0.096 -0.477 -0.314 -0.260 Aua -0.003 -0.004 0.043 -0.007 -0.013 Aub ----- ----- 0.002 -0.007 -0.015

The hydride ligand substantially affects the electronic structure of Au9(PH3)8H2+ and produces a

more isotropic structure. Unlike Au9(PH3)83+, which has two occupied 1P superatomic orbitals, in

Au9(PH3)8H2+ all three 1P superatomic orbitals are fully occupied and are relatively similar in

energy. The orbital energy changes induced by the hydride ligand are large enough to cause a

reordering of the superatomic orbital energies: in Au9(PH3)8H2+, the 1Px and 1Py superatomic

orbitals (HOMO and HOMO-1) are higher in energy than the 1Pz superatomic orbital (HOMO-2)

as shown in Figure 3b, which is a reversal the ordering seen in the unsubstituted NC. Thus,

Page 16: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

15

Au9(PH3)8H2+ can be viewed as a nearly spherical superatomic NC with a closed shell of eight

superatomic electrons (1S not shown for the same reason given for the unsubstituted NC),

consistent with previous computational results.85 The Au−H bond has little effect on the 1Px and

1Py superatomic orbitals because of the NC symmetry, and both orbitals are composed primarily

of the Au 6s and 6p atomic orbitals. The newly occupied 1Pz superatomic orbital is created mainly

by a bonding interaction between the 6s orbital of the central Au atom and the 1s orbital of the

hydride ligand and thus includes significant contributions from both the Au 6s and 6p orbitals and

the H 1s orbital. As it is case for Au9(PH3)8H2+, Au9(PPh3)8H2+ also has three occupied 1P

superatomic orbitals (Figure 3e), and the 1Pz superatomic orbital is the lowest in energy of the

three.

We have also computed the absorption spectra of both Au9(PH3)8H2+ (Figure 3c) and

Au9(PPh3)8H2+ (Figure 3f) to determine the effect of the ligands on the absorption spectrum.

Au9(PH3)8H2+ exhibits distinct absorption peaks near 2.8 and 3.4 eV, in contrast with the plateau

in the experimental absorption spectrum between 2.5 and 3.3 eV (Figure 1b). The excited states

below 3.5 eV are almost completely superatomic. The absorbing states at 2.3 eV and in the range

of the 2.8 – 2.9 eV involve the 1Px and 1Py superatomic orbitals, whereas the absorbing states in

the 3.3 – 3.4 eV range primarily involve the 1Pz superatomic orbital. The excited states above 3.5

eV have a mix of superatomic and ligand character. The computed absorption spectrum of

Au9(PPh3)8H2+ (Figure 3d) is much more similar to the experimental absorption spectrum, with a

plateau from 2.2 eV to 3.7 eV. For reasons of computational cost, excited states above 3.8 eV

could not be computed. The inclusion of the full phenyl groups also introduces higher-energy

occupied ligand MOs that reduce the energy at which the excited states switch from superatomic

Page 17: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

16

to ligand-based. The excited states below 3.1 eV have dominantly superatomic character; the

excited states above 3.1 eV involve a mix of superatomic and ligand character. Overall, these

results show that the experimental absorption spectrum of Au9(PPh3)8H2+ is consistent with the

hydride ligand behaving as a metal dopant, and there is no evidence that the hydride ligand has

electron-withdrawing character.

3.3. Halide-substituted Au9(PH3)8X2+ nanoclusters

We now turn to the halide substituted NCs Au9(PPh3)8X2+, where X is F, Cl, or Br. The chloride-

and bromide-substituted structures have been studied experimentally;66 we also include the

fluoride-substituted NC to extend the series. Similar to the previous sections, we have replaced the

experimental PPh3 ligands with PH3 to reduce computational cost. Our approach to determining

the most stable structures of the halide-substituted NCs is similar to what we previously described

for the hydride-substituted NCs. We constructed eleven different structures for each halide with

the halide placed either near the center Au atom or near one or two of the outer Au atoms; structures

were based on both the D2h and D4d structures of Au9(PH3)83+. Several low-energy structures for

each halide from our initial screening at the BP86/DZ level were re-optimized at the revTPSS/TZP

level.

Surprisingly, the most stable structures of the halide-substituted NCs are quite different from the

geometry of the hydride-substituted NC, despite the strong similarities in the experimental

absorption spectra. We have not found any stable geometries where the halide interacts directly

with the central Au atom. In the most stable structures (Figure 4), the halide bridges two adjacent

Page 18: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

17

outer Au atoms. For the fluoride, chloride, and bromide ligands, these most stable structures are

8.1, 3.8, and 5.9 kcal/mol more stable than the second most stable structures, respectively. In all

of these structures, the central Au atom is bonded only to the outer eight Au atoms and retains its

low coordination. More details on the higher-energy structures are provided in the SI.

Figure 4. (a, d, g) Optimized structures, (b, e, h) superatomic orbitals, and (c, f, i) absorption spectra of (a, b, c) Au9(PH3)8F2+, (d, e, f) Au9(PPh3)8Cl2+, and (g, h, i)

Au9(PPh3)8Br2+.

To understand the interactions of the halide ligands with the Au NC, we have analyzed the bond

lengths and atomic charges, analogous to our exploration of the hydride ligand. The chloride and

bromide ligands adsorb almost symmetrically, with two nearly equal Au-X bond lengths (Table

1). In contrast, the fluoride ligand adsorbs asymmetrically with one short and one long Au-F

interaction. All Au-X interactions are significantly shorter than the sum of the van der Waals radii

of the two atoms, but are longer than the sum of the covalent radii. This is in contrast with the Au-

Page 19: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

18

H bond, which is shorter than the sum of the covalent radii. These longer-distance interactions

suggest that the Au-X interactions have significantly less covalent character than the Au-H

interaction. The Hirshfeld atomic charges (Table 2) show substantial negative charge on all three

halides, which is likewise in contrast with the extremely small negative charge on the hydride

ligand. This suggests that the halide ligands all act as electron-withdrawing ligands, rather than

contributing their electrons to the Au NC as metal dopants. Since all three halides are significantly

more electronegative than Au, it is unsurprising for these elements to have electron-withdrawing

behavior. The electronic structure of the Au9(PH3)8X2+ NCs is also consistent with our assignment

of the halide ligands as electron-withdrawing ligands. These NCs have two occupied 1P

superatomic orbitals (HOMO, HOMO -1 in Figure 4) that are nearly degenerate; the halides do

not introduce an additional occupied superatomic orbital, in contrast with the hydride ligand. The

superatomic orbitals have primarily Au 6s and 6p character but also include minor contributions

from the halides, which increase with increasing halide size.

Even though the electronic structures and geometries of the halide-substituted NCs are quite

distinct from those of the hydride-substituted NC, the computed absorption spectra of the halide-

substituted NCs (Figure 4) have peaks in reasonable agreement with the experimental spectra

(Figure 1c,d), especially given our computational simplifications to the phosphine ligands.

Because the halide-substituted NCs have one fewer occupied superatomic MO than the hydride-

substituted NC, the excited states switch from mostly superatomic to mostly non-superatomic at

lower energies. The excited states in the three halide-substituted NCs are superatomic only up to

2.5-2.7 eV, as compared to 3.5 eV for the hydride-substituted NC. The higher-energy excited states

are mostly ligand-based. Among the halide-substituted NCs, the excited states of the fluoride-

Page 20: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

19

substituted NC have slightly more superatomic character than the chloride- and bromide-

substituted NCs, which may be due to the smaller contributions of the fluoride atomic orbitals to

the NC frontier MOs. Overall, these calculations show that the similar experimental absorption

spectra of the hydride- and halide-substituted Au9 NCs are coincidental and disguise major

differences in the geometries and electronic structures of these NCs. The halide ligands clearly act

as electron-withdrawing ligands, in strong contrast to the metal dopant character of the hydride

ligands.

4. Conclusions

Based on similarities in the experimental absorption spectra, it was recently proposed that hydride

and halide substituents have similar binding modes with the Au9(PH3)83+ NC.66 This proposal was

in contrast with previous expectations that the hydride ligand would act as a metal dopant, whereas

the halide ligands would have electron-withdrawing character. By computing the geometries and

absorption spectra of the unsubstituted Au9(PH3)83+ NC and its hydride- and halide-substituted

derivatives using DFT-based methods, we have shown that the hydride and halide ligands have

distinctly different binding geometries and electronic effects. The similarities in the experimental

absorption spectra of these NCs disguise these substantial differences.

Consistent with previous DFT-based studies, we have showed that the hydride ligand in

Au9(PH3)8H2+ binds to the central Au atom. The hydride ligand contributes its two electrons to the

superatomic electron count, giving the NC one more occupied superatomic MO and a 1S21P6

superatomic occupation. This is consistent with a description of the hydride ligand acting as a

Page 21: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

20

metal dopant. In contrast, the halide ligands bind in a position bridging two of the outer Au atoms;

these ligands are electron-withdrawing and do not change the number of occupied superatomic

orbitals, resulting in a 1S21P4 superatomic occupation. Whereas the hydride ligand occupies a

coordination site at the low-coordinated central Au atom, the halide ligands do not change the

coordination of the central Au atom.

These results highlight the power of detailed theoretical analysis of the geometries and electronic

structures of noble metal NCs to understand the role of ligands and interpret experimental data. In

particular, they underline differences in the ability of ligands to modify the coordination

environment of metal atoms, which is a feature that often correlates strongly with the catalytic

activity of metal NCs. Understanding the roles of ligands in determining the properties of noble

metal nanoclusters will support future efforts to use ligand derivatization as a powerful tool for the

rational design of metal nanoclusters as efficient catalysts.

Acknowledgments

The authors thank Prof. Christopher Johnson for helpful discussion. This work was supported by

start-up funding at Brandeis University. Computations were performed using Brandeis

University's High Performance Computing Cluster, which is partially funded by DMR-MRSEC

1420382.

Supporting Information

Relative energies and absorption frames of studied clusters. Excited-state properties and Cartesian

coordinates of the most stable bare and ligand-protected Au clusters.

Page 22: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

21

References

(1) Schwerdtfeger, P. Gold Goes Nano - From Small Clusters to Low-Dimensional

Assemblies. Angew. Chemie - Int. Ed. 2003, 42 (17), 1892–1895.

https://doi.org/10.1002/anie.200201610.

(2) Shang, L.; Dong, S.; Nienhaus, G. U. Ultra-Small Fluorescent Metal Nanoclusters:

Synthesis and Biological Applications. Nano Today 2011, 6 (4), 401–418.

https://doi.org/10.1016/j.nantod.2011.06.004.

(3) Li, G.; Jin, R. Atomically Precise Gold Nanoclusters as New Model Catalysts. Acc. Chem.

Res. 2013, 46 (8), 1749–1758. https://doi.org/10.1021/ar300213z.

(4) Fang, J.; Zhang, B.; Yao, Q.; Yang, Y.; Xie, J.; Yan, N. Recent Advances in the Synthesis

and Catalytic Applications of Ligand-Protected, Atomically Precise Metal Nanoclusters.

Coord. Chem. Rev. 2016, 322, 1–29. https://doi.org/10.1016/j.ccr.2016.05.003.

(5) Zhao, J.; Ge, L.; Yuan, H.; Liu, Y.; Gui, Y.; Zhang, B.; Zhou, L.; Fang, S. Heterogeneous

Gold Catalysts for Selective Hydrogenation: From Nanoparticles to Atomically Precise

Nanoclusters. Nanoscale 2019, 11 (24), 11429–11436.

https://doi.org/10.1039/C9NR03182K.

(6) Du, Y.; Sheng, H.; Astruc, D.; Zhu, M. Atomically Precise Noble Metal Nanoclusters as

Efficient Catalysts: A Bridge between Structure and Properties. Chem. Rev. 2020, 120 (2),

526–622. https://doi.org/10.1021/acs.chemrev.8b00726.

(7) Tian, S.; Cao, Y.; Chen, T.; Zang, S.; Xie, J. Ligand-Protected Atomically Precise Gold

Nanoclusters as Model Catalysts for Oxidation Reactions. Chem. Commun. 2020, 56 (8),

1163–1174. https://doi.org/10.1039/c9cc08215h.

Page 23: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

22

(8) Liu, C.; Yan, C.; Lin, J.; Yu, C.; Huang, J.; Li, G. One-Pot Synthesis of Au 144 (SCH 2

Ph) 60 Nanoclusters and Their Catalytic Application. J. Mater. Chem. A 2015, 3 (40),

20167–20173. https://doi.org/10.1039/C5TA05747G.

(9) Chen, H.; Liu, C.; Wang, M.; Zhang, C.; Luo, N.; Wang, Y.; Abroshan, H.; Li, G.; Wang,

F. Visible Light Gold Nanocluster Photocatalyst: Selective Aerobic Oxidation of Amines

to Imines. ACS Catal. 2017, 7, 3632–3638. https://doi.org/10.1021/acscatal.6b03509.

(10) Higaki, T.; Li, Q.; Zhou, M.; Zhao, S.; Li, Y.; Li, S.; Jin, R. Toward the Tailoring

Chemistry of Metal Nanoclusters for Enhancing Functionalities. Acc. Chem. Res. 2018, 51

(11), 2764–2773. https://doi.org/10.1021/acs.accounts.8b00383.

(11) Liu, C.; Ren, X.; Lin, F.; Fu, X.; Lin, X.; Li, T.; Sun, K.; Huang, J. Structure of the Au

23− x Ag x (S‐Adm) 15 Nanocluster and Its Application for Photocatalytic Degradation of

Organic Pollutants. Angew. Chemie Int. Ed. 2019, 58 (33), 11335–11339.

https://doi.org/10.1002/anie.201904612.

(12) Chai, O. J. H.; Liu, Z.; Chen, T.; Xie, J. Engineering Ultrasmall Metal Nanoclusters for

Photocatalytic and Electrocatalytic Applications. Nanoscale 2019, 11 (43), 20437–20448.

https://doi.org/10.1039/C9NR07272A.

(13) Qin, Z.; Zhao, D.; Zhao, L.; Xiao, Q.; Wu, T.; Zhang, J.; Wan, C.; Li, G. Tailoring the

Stability, Photocatalysis and Photoluminescence Properties of Au11 Nanoclusters: Via

Doping Engineering. Nanoscale Adv. 2019, 1 (7), 2529–2536.

https://doi.org/10.1039/c9na00234k.

(14) Jin, R. Quantum Sized, Thiolate-Protected Gold Nanoclusters. Nanoscale 2010, 2 (3),

343–362. https://doi.org/10.1039/b9nr00160c.

(15) Qian, H.; Zhu, M.; Wu, Z.; Jin, R. Quantum Sized Gold Nanoclusters with Atomic

Page 24: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

23

Precision. Acc. Chem. Res. 2012, 45 (9), 1470–1479. https://doi.org/10.1021/ar200331z.

(16) Negishi, Y.; Matsuura, Y.; Tomizawa, R.; Kurashige, W.; Niihori, Y.; Takayama, T.;

Iwase, A.; Kudo, A. Controlled Loading of Small Aun Clusters (n = 10-39) onto

BaLa4Ti4O15 Photocatalysts: Toward an Understanding of Size Effect of Cocatalyst on

Water-Splitting Photocatalytic Activity. J. Phys. Chem. C 2015, 119 (20), 11224–11232.

https://doi.org/10.1021/jp5122432.

(17) Wang, Y.; Liu, X.; Wang, Q.; Quick, M.; Kovalenko, S.; Chen, Q.; Koch, N.; Pinna, N.

Insights into Charge Transfer at the Atomically Precise Nanocluster/Semiconductor

Interface for in-Depth Understanding the Role of Nanocluster in Photocatalytic System.

Angew. Chemie Int. Ed. 2020. https://doi.org/10.1002/anie.201915074.

(18) Abbas, M. A.; Kamat, P. V.; Bang, J. H. Thiolated Gold Nanoclusters for Light Energy

Conversion. ACS Energy Lett. 2018, 3 (4), 840–854.

https://doi.org/10.1021/acsenergylett.8b00070.

(19) Li, G.; Jin, R. Gold Nanocluster-Catalyzed Semihydrogenation: A Unique Activation

Pathway for Terminal Alkynes. J. Am. Chem. Soc. 2014, 136 (32), 11347–11354.

https://doi.org/10.1021/ja503724j.

(20) Koklioti, M. A.; Skaltsas, T.; Sato, Y.; Suenaga, K.; Stergiou, A.; Tagmatarchis, N.

Mechanistic Insights into the Photocatalytic Properties of Metal Nanocluster/Graphene

Ensembles. Examining the Role of Visible Light in the Reduction of 4-Nitrophenol.

Nanoscale 2017, 9 (27), 9685–9692. https://doi.org/10.1039/C7NR02944F.

(21) Liu, C.; Abroshan, H.; Yan, C.; Li, G.; Haruta, M. One-Pot Synthesis of Au 11 (PPh 2 Py)

7 Br 3 for the Highly Chemoselective Hydrogenation of Nitrobenzaldehyde. ACS Catal.

2016, 6 (1), 92–99. https://doi.org/10.1021/acscatal.5b02116.

Page 25: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

24

(22) Xu, J.; Xu, S.; Chen, M.; Zhu, Y. Unlocking the Catalytic Activity of an Eight-Atom Gold

Cluster with a Pd Atom. Nanoscale 2020, 12 (10), 6020–6028.

https://doi.org/10.1039/C9NR10198E.

(23) Jin, R.; Li, G.; Sharma, S.; Li, Y.; Du, X. Toward Active-Site Tailoring in Heterogeneous

Catalysis by Atomically Precise Metal Nanoclusters with Crystallographic Structures.

Chem. Rev. 2020, acs.chemrev.0c00495. https://doi.org/10.1021/acs.chemrev.0c00495.

(24) Cai, X.; Saranya, G.; Shen, K.; Chen, M.; Si, R.; Ding, W.; Zhu, Y. Reversible Switching

of Catalytic Activity by Shuttling an Atom into and out of Gold Nanoclusters. Angew.

Chemie Int. Ed. 2019, 58 (29), 9964–9968. https://doi.org/10.1002/anie.201903853.

(25) Liu, Y.; Chai, X.; Cai, X.; Chen, M.; Jin, R.; Ding, W.; Zhu, Y. Central Doping of a

Foreign Atom into the Silver Cluster for Catalytic Conversion of CO 2 toward C−C Bond

Formation. Angew. Chemie Int. Ed. 2018, 57 (31), 9775–9779.

https://doi.org/10.1002/anie.201805319.

(26) Choi, W.; Hu, G.; Kwak, K.; Kim, M.; Jiang, D.; Choi, J.-P.; Lee, D. Effects of Metal-

Doping on Hydrogen Evolution Reaction Catalyzed by MAu 24 and M 2 Au 36

Nanoclusters (M = Pt, Pd). ACS Appl. Mater. Interfaces 2018, 10 (51), 44645–44653.

https://doi.org/10.1021/acsami.8b16178.

(27) Jin, R.; Zeng, C.; Zhou, M.; Chen, Y. Atomically Precise Colloidal Metal Nanoclusters

and Nanoparticles: Fundamentals and Opportunities. Chem. Rev. 2016, 116 (18), 10346–

10413. https://doi.org/10.1021/acs.chemrev.5b00703.

(28) Liu, P.; Qin, R.; Fu, G.; Zheng, N. Surface Coordination Chemistry of Metal

Nanomaterials. J. Am. Chem. Soc. 2017, 139 (6), 2122–2131.

https://doi.org/10.1021/jacs.6b10978.

Page 26: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

25

(29) Li, G.; Abroshan, H.; Liu, C.; Zhuo, S.; Li, Z.; Xie, Y.; Kim, H. J.; Rosi, N. L.; Jin, R.

Tailoring the Electronic and Catalytic Properties of Au 25 Nanoclusters via Ligand

Engineering. ACS Nano 2016, 10 (8), 7998–8005.

https://doi.org/10.1021/acsnano.6b03964.

(30) Wan, X.; Wang, J.; Nan, Z.; Wang, Q. Ligand Effects in Catalysis by Atomically Precise

Gold Nanoclusters. Sci. Adv. 2017, 3, e1701823. https://doi.org/10.1126/sciadv.1701823.

(31) Higaki, T.; Li, Y.; Zhao, S.; Li, Q.; Li, S.; Du, X.; Yang, S.; Chai, J.; Jin, R. Atomically

Tailored Gold Nanoclusters for Catalytic Application. Angew. Chemie Int. Ed. 2019, 58

(25), 8291–8302. https://doi.org/10.1002/anie.201814156.

(32) Nasaruddin, R. R.; Chen, T.; Li, J.; Goswami, N.; Zhang, J.; Yan, N.; Xie, J. Ligands

Modulate Reaction Pathway in the Hydrogenation of 4-Nitrophenol Catalyzed by Gold

Nanoclusters. ChemCatChem 2018, 10 (2), 395–402.

https://doi.org/10.1002/cctc.201701472.

(33) Yuan, S.-F.; Lei, Z.; Guan, Z.-J.; Wang, Q.-M. Atomically Precise Preorganization of

Open Metal Sites on Gold Nanocluster with High Catalytic Performance. Angew. Chemie

Int. Ed. 2020, anie.202012499. https://doi.org/10.1002/anie.202012499.

(34) Zhao, S.; Austin, N.; Li, M.; Song, Y.; House, S. D.; Bernhard, S.; Yang, J. C.;

Mpourmpakis, G.; Jin, R. Influence of Atomic-Level Morphology on Catalysis: The Case

of Sphere and Rod-Like Gold Nanoclusters for CO 2 Electroreduction. ACS Catal. 2018, 8

(6), 4996–5001. https://doi.org/10.1021/acscatal.8b00365.

(35) Goel, S.; Velizhanin, K. A.; Piryatinski, A.; Ivanov, S. A.; Tretiak, S. Ligand Effects on

Optical Properties of Small Gold Clusters: A TDDFT Study. J. Phys. Chem. C 2012, 116

(5), 3242–3249. https://doi.org/10.1021/jp208732k.

Page 27: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

26

(36) Goel, S.; Velizhanin, K. A.; Piryatinski, A.; Tretiak, S.; Ivanov, S. A. DFT Study of

Ligand Binding to Small Gold Clusters. J. Phys. Chem. Lett. 2010, 1 (6), 927–931.

https://doi.org/10.1021/jz1000193.

(37) Shabaninezhad, M.; Abuhagr, A.; Sakthivel, N. A.; Kumara, C.; Dass, A.; Kwak, K.; Pyo,

K.; Lee, D.; Ramakrishna, G. Ultrafast Electron Dynamics in Thiolate-Protected

Plasmonic Gold Clusters: Size and Ligand Effect. J. Phys. Chem. C 2019, 123 (21),

13344–13353. https://doi.org/10.1021/acs.jpcc.9b01739.

(38) Johnson, G. E.; Laskin, J. Understanding Ligand Effects in Gold Clusters Using Mass

Spectrometry. Analyst 2016, 141 (12), 3573–3589. https://doi.org/10.1039/c6an00263c.

(39) Günter Schmid, Benedetto Corain, N. T. Metal Nanoclusters in Catalysis and Materials

Science: The Issue of Size Control; Elsevier Ltd, 2011.

(40) Walter, M.; Akola, J.; Lopez-Acevedo, O.; Jadzinsky, P. D.; Calero, G.; Ackerson, C. J.;

Whetten, R. L.; Gronbeck, H.; Hakkinen, H. A Unified View of Ligand-Protected Gold

Clusters as Superatom Complexes. Proc. Natl. Acad. Sci. USA 2008, 105, 9157–9162.

https://doi.org/10.1073/pnas.0801001105.

(41) Häkkinen, H. Electronic Shell Structures in Bare and Protected Metal Nanoclusters. Adv.

Phys. X 2016, 1 (3), 467–491. https://doi.org/10.1080/23746149.2016.1219234.

(42) Lopez-Acevedo, O.; Akola, J.; Whetten, R. L.; Grönbeck, H.; Häkkinen, H. Structure and

Bonding in the Ubiquitous Icosahedral Metallic Gold Cluster Au 144(SR) 60. J. Phys.

Chem. C 2009, 113 (13), 5035–5038. https://doi.org/10.1021/jp8115098.

(43) Mingos, D. M. P. Structural and Bonding Patterns in Gold Clusters. Dalt. Trans. 2015, 44

(15), 6680–6695. https://doi.org/10.1039/c5dt00253b.

(44) Zhu, M.; Aikens, C. M.; Hendrich, M. P.; Gupta, R.; Qian, H.; Schatz, G. C.; Jin, R.

Page 28: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

27

Reversible Switching of Magnetism in Thiolate-Protected Au 25 Superatoms. J. Am.

Chem. Soc. 2009, 131 (7), 2490–2492. https://doi.org/10.1021/ja809157f.

(45) Lugo, G.; Schwanen, V.; Fresch, B.; Remacle, F. Charge Redistribution Effects on the

UV-Vis Spectra of Small Ligated Gold Clusters: A Computational Study. J. Phys. Chem.

C 2015, 119 (20), 10969–10980. https://doi.org/10.1021/jp511120j.

(46) Karimova, N. V.; Aikens, C. M. Time Dependent Density Functional Theory Study of

Magnetic Circular Dichroism Spectra of Gold Clusters Au 9 (PH 3 ) 8 3+ and Au 9 (PPh 3

) 8 3+ . J. Phys. Chem. A 2016, 120 (48), 9625–9635.

https://doi.org/10.1021/acs.jpca.6b10063.

(47) Karimova, N. V.; Aikens, C. M. Optical Properties of Small Gold Clusters Au8L82+ (L =

PH3, PPh3): Magnetic Circular Dichroism Spectra. J. Phys. Chem. C 2017, 121 (35),

19478–19489. https://doi.org/10.1021/acs.jpcc.7b05630.

(48) Weerawardene, K. L. D. M.; Aikens, C. M. Theoretical Insights into the Origin of

Photoluminescence of Au25(SR)18- Nanoparticles. J. Am. Chem. Soc. 2016, 138 (35),

11202–11210. https://doi.org/10.1021/jacs.6b05293.

(49) Muniz-Miranda, F.; Menziani, M. C.; Pedone, A. Assessment of Exchange-Correlation

Functionals in Reproducing the Structure and Optical Gap of Organic-Protected Gold

Nanoclusters. J. Phys. Chem. C 2014, 118 (14), 7532–7544.

https://doi.org/10.1021/jp411483x.

(50) Muñoz-Castro, A. On the Ligand-Core Interaction in Ligand-Protected Gold Superatoms.

Insights from Au25(XR)18 (X = S, Se, Te): Via Relativistic DFT Calculations. Phys.

Chem. Chem. Phys. 2019, 21 (24), 13022–13029. https://doi.org/10.1039/c9cp02077b.

(51) Kwak, K.; Choi, W.; Tang, Q.; Jiang, D.; Lee, D. Rationally Designed Metal Nanocluster

Page 29: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

28

for Electrocatalytic Hydrogen Production from Water. J. Mater. Chem. A 2018, 6 (40),

19495–19501. https://doi.org/10.1039/C8TA06306K.

(52) Hu, G.; Wu, Z.; Jiang, D. Stronger-than-Pt Hydrogen Adsorption in a Au 22 Nanocluster

for the Hydrogen Evolution Reaction. J. Mater. Chem. A 2018, 6 (17), 7532–7537.

https://doi.org/10.1039/C8TA00461G.

(53) Buckart, S.; Ganteför, G.; Kim, Y. D.; Jena, P. Anomalous Behavior of Atomic Hydrogen

Interacting with Gold Clusters. J. Am. Chem. Soc. 2003, 125 (46), 14205–14209.

https://doi.org/10.1021/ja036544t.

(54) Mondal, K.; Agrawal, S.; Manna, D.; Banerjee, A.; Ghanty, T. K. Effect of Hydrogen

Atom Doping on the Structure and Electronic Properties of 20-Atom Gold Cluster. J.

Phys. Chem. C 2016, 120 (33), 18588–18594. https://doi.org/10.1021/acs.jpcc.6b04584.

(55) Takano, S.; Hirai, H.; Muramatsu, S.; Tsukuda, T. Hydride-Doped Gold Superatom (Au 9

H) 2+ : Synthesis, Structure, and Transformation. J. Am. Chem. Soc. 2018, 140 (27),

8380–8383. https://doi.org/10.1021/jacs.8b03880.

(56) Takano, S.; Hasegawa, S.; Suyama, M.; Tsukuda, T. Hydride Doping of Chemically

Modified Gold-Based Superatoms. Acc. Chem. Res. 2018, 51 (12), 3074–3083.

https://doi.org/10.1021/acs.accounts.8b00399.

(57) Hu, G.; Tang, Q.; Lee, D.; Wu, Z.; Jiang, D. E. Metallic Hydrogen in Atomically Precise

Gold Nanoclusters. Chem. Mater. 2017, 29 (11), 4840–4847.

https://doi.org/10.1021/acs.chemmater.7b00776.

(58) Hirai, H.; Takano, S.; Tsukuda, T. Synthesis of Trimetallic (HPd@M 2 Au 8 ) 3+

Superatoms (M = Ag, Cu) via Hydride-Mediated Regioselective Doping to (Pd@Au 8 )

2+. ACS Omega 2019, 4 (4), 7070–7075. https://doi.org/10.1021/acsomega.9b00575.

Page 30: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

29

(59) Cariati, F.; Naldini, L. Preparation and Properties of Gold Atom Cluster Compounds :

Octakis- (Triary1phosphine)Enneagold Trianion. J.C.S. Dalt. 1972, 20, 2286–2287.

(60) Takano, S.; Yamazoe, S.; Tsukuda, T. A Gold Superatom with 10 Electrons in

Au13(PPh3)8(p-SC6H4CO2H)3. APL Mater. 2017, 5 (5), 3–8.

https://doi.org/10.1063/1.4976018.

(61) Das, A.; Liu, C.; Byun, H. Y.; Nobusada, K.; Zhao, S.; Rosi, N.; Jin, R. Structure

Determination of [Au18(SR)14]. Angew. Chemie - Int. Ed. 2015, 54 (10), 3140–3144.

https://doi.org/10.1002/anie.201410161.

(62) Hall, K. P.; Theoblad, B. R. C.; Gilmour, D. I.; Mingos, D. M. P.; Welch, A. J. Synthesis

and Structural Characterization of [Au9{P(p-C6H4OMe)3}8](BF4)3; a Cluster with a

Centred Crown of Gold Atoms. J. Chem. Soc. Chem. Commun. 1982, 3 (10), 528.

https://doi.org/10.1039/c39820000528.

(63) Schulz-Dobrick, M.; Jansen, M. Supramolecular Intercluster Compounds Consisting of

Gold Clusters and Keggin Anions. Eur. J. Inorg. Chem. 2006, 2006 (22), 4498–4502.

https://doi.org/10.1002/ejic.200600790.

(64) Wen, F.; Englert, U.; Gutrath, B.; Simon, U. Crystal Structure, Electrochemical and

Optical Properties of [Au 9(PPh3)8](NO3)3. Eur. J. Inorg. Chem. 2008, 9 (1), 106–111.

https://doi.org/10.1002/ejic.200700534.

(65) Hirata, K.; Chakraborty, P.; Nag, A.; Takano, S.; Koyasu, K.; Pradeep, T.; Tsukuda, T.

Interconversions of Structural Isomers of [PdAu 8 (PPh 3 ) 8 ] 2+ and [Au 9 (PPh 3 ) 8 ]

3+ Revealed by Ion Mobility Mass Spectrometry. J. Phys. Chem. C 2018, 122 (40),

23123–23128. https://doi.org/10.1021/acs.jpcc.8b04722.

(66) Cirri, A.; Hernández, H. M.; Johnson, C. J. Hydride, Chloride, and Bromide Show Similar

Page 31: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

30

Electronic Effects in the Au9(PPh3)83+ Nanocluster. Chem. Commun. 2020, 56 (8),

1283–1285. https://doi.org/10.1039/c9cc08009k.

(67) Becke, A. D. Density-Functional Exchange-Energy Approximation with Correct

Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098–3100.

https://doi.org/10.1103/PhysRevA.38.3098.

(68) Perdew, J. P. Density-Functional Approximation for the Correlation Energy of the

Inhomogeneous Electron Gas. Phys. Rev. B 1986, 33, 8822–8824.

https://doi.org/10.1103/PhysRevB.33.8822.

(69) Aikens, C. M.; Schatz, G. C. TDDFT Studies of Absorption and SERS Spectra of Pyridine

Interacting with Au20. J. Phys. Chem. A 2006, 110, 13317–13324.

(70) Fihey, A.; Kloss, B.; Perrier, A.; Maurel, F. Density Functional Theory Study of the

Conformation and Optical Properties of Hybrid Aun-Dithienylethene Systems (n = 3, 19,

25). J. Phys. Chem. A 2014, 118 (26), 4695–4706. https://doi.org/10.1021/jp501542m.

(71) Bae, G.; Aikens, C. M. Time-Dependent Density Functional Theory Studies of Optical

Properties of Au Nanoparticles: Octahedra, Truncated Octahedra, and Icosahedra. J. Phys.

Chem. C 2015, 119, 23127−23137. https://doi.org/10.1021/jp300789x.

(72) Chen, L.; Gao, Y.; Cheng, Y.; Li, H.; Wang, Z.; Li, Z.; Zhang, R.-Q. Nonresonant

Chemical Mechanism in Surface-Enhanced Raman Scattering of Pyridine on M@Au 12

Clusters. Nanoscale 2016, 8, 4086–4093. https://doi.org/10.1039/C5NR07246H.

(73) Perdew, J. P.; Ruzsinszky, A.; Csonka, G. I.; Constantin, L. A.; Sun, J. Workhorse

Semilocal Density Functional for Condensed Matter Physics and Quantum Chemistry.

Phys. Rev. Lett. 2009, 103 (2), 10–13. https://doi.org/10.1103/PhysRevLett.103.026403.

(74) Baek, H.; Moon, J.; Kim, J. Benchmark Study of Density Functional Theory for Neutral

Page 32: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

31

Gold Clusters, Aun (n = 2-8). J. Phys. Chem. A 2017, 121 (12), 2410–2419.

https://doi.org/10.1021/acs.jpca.6b11868.

(75) Gritsenko, O. V; Schipper, P. R. T.; Baerends, E. J. Approximation of the Exchange-

Correlation Kohn–Sham Potential with a Statistical Average of Different Orbital Model

Potentials. Chem. Phys. Lett. 1999, 302, 199–207.

(76) van Lenthe, E.; Snijders, J. G.; Baerends, E. J. The Zero-Order Regular Approximation for

Relativistic Effects: The Effect of Spin–Orbit Coupling in Closed Shell Molecules. J.

Chem. Phys. 1996, 105, 6505–6516.

(77) te Velde, G.; Bickelhaupt, F. M.; Baerends, E. J.; Fonseca Guerra, C.; van Gisbergen, S. J.

A.; Snijders, J. G.; Ziegler, T. Chemistry with ADF. J. Comput. Chem. 2001, 22 (9), 931–

967. https://doi.org/10.1002/jcc.1056.

(78) Amsterdam Density Functional 2018; SCM, Theoretical Chemistry: Vrije Universiteit,

Amsterdam, The Netherlands.

(79) Gieseking, R. L. M.; Ashwell, A. P.; Ratner, M. A.; Schatz, G. C. Analytical Approaches

To Identify Plasmon-like Excited States in Bare and Ligand-Protected Metal

Nanoclusters. J. Phys. Chem. C 2020, 124 (5), 3260–3269.

https://doi.org/10.1021/acs.jpcc.9b10569.

(80) Segala, M.; Schneider, F. S. S.; Caramori, G. F.; Parreira, R. L. T. Evaluation of Electron

Donation as a Mechanism for the Stabilisation of Chalcogenate-Protected Gold

Nanoclusters. ChemPhysChem 2016, 17 (19), 3102–3111.

https://doi.org/10.1002/cphc.201600552.

(81) Sanderson, R. T. Electronegativity and Bond Energy. J. Am. Chem. Soc. 1983, 105 (8),

2259–2261. https://doi.org/10.1021/ja00346a026.

Page 33: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

32

(82) Bondi, A. Van Der Waals Volumes and Radii. J. Phys. Chem. 1964, 68, 441–451.

(83) Hirshfeld, F. L. Bonded-Atom Fragments for Describing Molecular Charge Densities.

Theor. Chim. Acta 1977, 44, 129–138. https://doi.org/10.1007/BF00549096.

(84) Muniz-Miranda, F.; Presti, D.; Menziani, M. C.; Pedone, A. Electronic and Optical

Properties of the Au22[1,8-Bis(Diphenylphosphino) Octane]6 Nanoclusters Disclosed by

DFT and TD-DFT Calculations. Theor. Chem. Acc. 2016, 135 (1), 1–9.

https://doi.org/10.1007/s00214-015-1764-x.

(85) Omoda, T.; Takano, S.; Tsukuda, T. Toward Controlling the Electronic Structures of

Chemically Modified Superatoms of Gold and Silver. Small 2020, 2001439, 1–18.

https://doi.org/10.1002/smll.202001439.

Page 35: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Supporting Information:

Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters: Similar

Absorption Spectra Disguise Distinct Geometries and Electronic Structures

Yavuz S. Ceylan, Rebecca L. M. Gieseking*

Department of Chemistry, Brandeis University

415 South Street, Waltham, Massachusetts 02453

* Corresponding author: [email protected]; (781)-736-2511

Page 36: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Au9(PH3)83+ (D2h and D4d symmetries)

Table S1. Excited states of Au9(PH3)83+ (D2h symmetry). All excited states between 2.0 and 3.9

eV with oscillator strengths larger than 0.02 are listed.

Energy (eV) f Superatomic character (%)

Weight (%) Transition

2.52 0.0210 99.19 51.20 HOMO LUMO

47.48 HOMO LUMO+2

2.63 0.1097 80.22 40.12 HOMO-1 LUMO

39.65 HOMO LUMO+2

18.64 HOMO-2 LUMO

3.17 0.2754 15.15 73.32 HOMO-2 LUMO

3.51 0.0207 18.13 15.64 HOMO LUMO+5

3.57 0.0810 1.15 89.50 HOMO-5 LUMO+1

3.62 0.0250 0.84 89.33 HOMO-6 LUMO

3.87 0.7635 69.8 48.39 HOMO LUMO+2

17.32 HOMO-1 LUMO+1

16.66 HOMO-2 LUMO+2

Table S2. Excited states of Au9(PH3)83+ (D4d symmetry). All excited states between 2.0 and3.9 eV

with oscillator strengths larger than 0.02 are listed.

Energy (eV) f Superatomic character (%)

Weight (%) Transition

2.87 0.176 95.75 89.03 HOMO LUMO+1

3.57 0.0911 60.86 57.84 HOMO LUMO+3

19.87 HOMO-2 LUMO+1

16.85 HOMO-2 LUMO

3.82 0.1751 11.72 73.75 HOMO-2 LUMO+1

9.99 HOMO-9 LUMO

3.87 0.0218 50.1 49.16 HOMO LUMO+4

24.98 HOMO-7 LUMO

Page 37: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Figure S1. Kohn–Sham orbital energy diagram for the (top) D2h and (bottom) D4d symmetries of Au9(PH3)8

3+. The horizontal colored bar for each Kohn-Sham orbital indicates the percent contributions of the atomic orbitals of Au (sp) in yellow, Au (d) in blue, and P (sp) in violet. The y axis shows the orbital energy.

Page 38: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Au9(PH3)8H2+

Table S3. Relative energies at the revTPSS/TZP level of Au9(PH3)8H2+. d2h # and d4d # represent different optimized geometries derived from different starting positions of the hydride on the unsubstituted D2h and D4d symmetries of Au9(PH3)8

3+. The structure indicated in bold (d2h 2) is the most stable structure, described in the manuscript.

(kcal/mol) d2h 1 9.18 d2h 2 0.00 d2h 3 11.93 d2h 4 8.61 d4d 1 6.75 d4d 2 7.81

Page 39: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Front view Top view d2h 1

d2h 2

d2h 3

d2h 4

d4d 1

d4d 2

Figure S2. Optimized structures of Au9(PH3)8H2+ at revTPSS/TZP level of theory

Page 40: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Figure S3. Absorption spectra of the optimized structures of Au9(PH3)8H2+.

0

12

3

4

56

7

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 1 / Au9(PH3)8H2+

0

1

2

3

4

5

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 2 / Au9(PH3)8H2+

0

1

2

3

4

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 3 / Au9(PH3)8H2+

0123456

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 4 / Au9(PH3)8H2+

0

1

2

3

4

5

6

7

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d4d 1 / Au9(PH3)8H2+

0

1

2

3

4

5

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d4d 1 / Au9(PH3)8H2+

Page 41: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Figure S4. Kohn–Sham orbital energy diagram for the most stable structure of Au9(PPh3)8H2+. The horizontal colored bar for each Kohn-Sham orbital indicates the percent contributions of the atomic orbitals of Au (sp) in yellow, H (s) in green, Au (d) in blue, and P (sp) in violet. The y axis shows the orbital energy.

Page 42: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Table S4. Excited states of Au9(PH3)8H2+ (D2h symmetry (d2h 2)). All excited states between 2.0 and 3.9 eV with oscillator strengths larger than 0.01 are listed.

Energy (eV) f Superatomic character (%)

Weight (%) Transition

2.30 0.035 99.61 83.57 HOMO LUMO+1

2.72 0.0129 99.81 57.84 HOMO LUMO+3

19.87 HOMO LUMO+4

2.82 0.1196 97.83 65.50 HOMO-1 LUMO

12.07 HOMO-1 LUMO+1

2.86 0.1202 98.24 42.76 HOMO-1 LUMO+1

31.42 HOMO LUMO+2

2.88 0.0251 99.76 42.91 HOMO-2 LUMO+1

33.04 HOMO-2 LUMO

3.33 0.0749 97.17 86.44 HOMO-2 LUMO+3

3.40 0.095 97.11 42.19 HOMO-2 LUMO+2

27.15 HOMO-2 LUMO+4

17.14 HOMO-1 LUMO+4

3.60 0.0126 94.83 93.70 HOMO LUMO+6

3.75 0.0315 7.44 82.74 HOMO-3 LUMO

3.82 0.0245 90.96 87.99 HOMO-1 LUMO+5

Page 43: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Au9(PH3)8F2+

Table S5. Relative energies at the revTPSS/TZP level of Au9(PH3)8F2+. d2h # and d4d # represent different optimized geometries derived from different starting positions of the fluoride on the unsubstituted D2h and D4d symmetries of Au9(PH3)8

3+. The structure indicated in bold (d4d 2) is the most stable structure, described in the manuscript.

(kcal/mol) d2h 1 10.60 d2h 2 10.70 d2h 3 8.11 d4d 1 10.23 d4d 2 0.00

Page 44: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Front view Top view

d2h 1

d2h 2

d2h 3

d4d 1

d4d 2

Figure S5. Optimized structures of Au9(PH3)8F2+ at revTPSS/TZP level of theory

Page 45: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Figure S6. Absorption spectra of fluoride ligated structures.

012345678

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 1 / Au9(PH3)8F2+

0

1

2

3

4

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 2 / Au9(PH3)8F2+

0

1

2

3

4

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 3 / Au9(PH3)8F2+

0

1

2

3

4

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d4d 1 / Au9(PH3)8F2+

0

1

2

3

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d4d 2 / Au9(PH3)8F2+

Page 46: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Table S6. Excited states of Au9(PH3)8F2+ (D4d symmetry (d4d 2)). All excited states between 2.0 and 3.9 eV with oscillator strengths larger than 0.01 are listed.

Energy (eV) f Superatomic character (%)

Weight (%) Transition

2.24 0.0161 99.12 65.79 HOMO LUMO

2.43 0.0356 95.84 52.67 HOMO -1 LUMO

19.76 HOMO LUMO +1

15.72 HOMO LUMO +2

2.51 0.0432 97.18 79.35 HOMO -1 LUMO +1

2.61 0.0509 94.28 79.58 HOMO LUMO +2

2.77 0.0554 93.32 83.53 HOMO -1 LUMO +2

2.87 0.0283 4.43 72.94 HOMO -2 LUMO

18.00 HOMO -2 LUMO +1

2.98 0.0457 9.43 74.60 HOMO -2 LUMO +1

11.21 HOMO -2 LUMO

3.25 0.0151 70.94 29.31 HOMO LUMO +4

29.07 HOMO -1 LUMO +3

23.94 HOMO -2 LUMO +2

3.33 0.0165 1.27 76.16 HOMO -3 LUMO +2

19.20 HOMO -4 LUMO

3.46 0.1022 17.72 54.79 HOMO -4 LUMO

14.48 HOMO LUMO +5

11.52 HOMO -3 LUMO +2

3.67 0.1323 46.21 25.07 HOMO -1 LUMO +5

17.62 HOMO -6 LUMO

10.93 HOMO -4 LUMO +2

9.21 HOMO -4 LUMO +1

3.80 0.0672 34.79 26.16 HOMO -1 LUMO +6

23.80 HOMO -3 LUMO +3

18.76 HOMO -4 LUMO +2

3.85 0.0571 6.4 41.57 HOMO -3 LUMO +3

33.78 HOMO -2 LUMO +4

Page 47: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Figure S7. Kohn–Sham orbital energy level diagram for the Au9(PH3)8F2+ structure. Each KS orbital is drawn to indicate the relative contributions (line length with color labels) of the atomic orbitals of Au (sp) in gold, F (s) in green, Au (d) in blue, and P (sp) in purple. The y axis shows the M.O. and energy level. X axis shows the percentage of the specific atom contributions.

Au9(PH3)8Cl2+

Table S7. Relative energies at the revTPSS/TZP level of Au9(PH3)8Cl2+. d2h # and d4d # represent different optimized geometries derived from different starting positions of the chloride on the unsubstituted D2h and D4d symmetries of Au9(PH3)8

3+. The structure indicated in bold (d4d 2) is the most stable structure, described in the manuscript.

Name (kcal/mol) d2h 1 3.77 d2h 2 8.93 d2h 3 6.16 d4d 1 9.44 d4d 2 0.00

Page 48: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Front view Top view

d2h 1

d2h 2

d2h 3

d4d 1

d4d 2

Figure S8. Optimized structures of Au9(PH3)8Cl2+ at revTPSS/TZP level of theory

Page 49: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Figure S9. Absorption spectra of chloride ligated structures.

0

1

2

3

4

5

6

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 1 / Au9(PH3)8Cl2+

0

1

2

3

4

5

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 2 / Au9(PH3)8Cl2+

0

1

2

3

4

5

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 3 / Au9(PH3)8Cl2+

0

1

2

3

4

5

6

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d4d 1 / Au9(PH3)8Cl2+

0

1

2

3

4

5

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d4d 2 / Au9(PH3)8Cl2+

Page 50: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Table S8. Excited states of Au9(PH3)8Cl2+ (D4d symmetry (d4d 2)). All excited states between 2.0 and 3.9 eV with oscillator strengths larger than 0.01 are listed.

Energy (eV) f Superatomic character (%)

Weight (%) Transition

2.39 0.033 85.23 37.49 HOMO LUMO +1

23.47 HOMO -1 LUMO

19.22 HOMO -1 LUMO +1

10.00 HOMO -2 LUMO

2.64 0.0187 76.55 37.30 HOMO LUMO +2

36.16 HOMO -1 LUMO +2

9.56 HOMO -3 LUMO +1

9.24 HOMO -2 LUMO

2.70 0.0401 82.05 40.20 HOMO -1 LUMO +2

35.56 HOMO LUMO +2

11.40 HOMO -3 LUMO

2.80 0.0535 11.54 46.67 HOMO -3 LUMO

36.00 HOMO -4 LUMO

2.86 0.0731 10.86 72.91 HOMO -3 LUMO +1

3.12 0.0489 3.41 89.18 HOMO -3 LUMO +2

3.40 0.0631 39.74 29.97 HOMO -5 LUMO

26.92 HOMO LUMO +4

25.12 HOMO -5 LUMO +1

3.46 0.0409 57.9 35.04 HOMO -5 LUMO +1

17.84 HOMO -1 LUMO +4

15.60 HOMO LUMO +5

12.77 HOMO LUMO +4

10.49 HOMO -1 LUMO +5

3.52 0.0381 90.04 51.76 HOMO -1 LUMO +5

36.16 HOMO LUMO +5

3.58 0.0241 19.96 62.10 HOMO -3 LUMO +3

9.56 HOMO -5 LUMO +2

9.47 HOMO -1 LUMO +5

3.60 0.033 27.5 33.41 HOMO -3 LUMO +3

31.24 HOMO -5 LUMO +2

Page 51: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

10.57 HOMO LUMO +5

9.54 HOMO -1 LUMO +5

3.82 0.2062 23.35 25.36 HOMO -8 LUMO +1

19.24 HOMO -5 LUMO +2

9.82 HOMO LUMO +5

8.96 HOMO -7 LUMO +1

Figure S10. Kohn–Sham orbital energy level diagram for the Au9(PH3)8Cl2+ structure. Each KS orbital is drawn to indicate the relative contributions (line length with color labels) of the atomic orbitals of Au (sp) in gold, Cl(s) in green, Au (d) in blue, and P (sp) in purple. The y axis shows the M.O. and energy level. X axis shows the percentage of the specific atom contributions.

Page 52: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Au9(PH3)8Br2+

Table S9. Relative energies at the revTPSS/TZP level of Au9(PH3)8Br2+. d2h # and d4d # represent different optimized geometries derived from different starting positions of the bromide on the unsubstituted D2h and D4d symmetries of Au9(PH3)8

3+. The structure indicated in bold (d4d 2) is the most stable structure, described in the manuscript.

Name (kcal/mol) d2h 1 6.74 d2h 2 10.98 d2h 3 5.88 d4d 1 9.42 d4d 2 0.00

Page 53: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Front view Top view

d2h 1

d2h 2

d2h 3

d4d 1

d4d 2

Figure S11. Optimized structures of Au9(PH3)8Br2+ at revTPSS/TZP level of theory

Page 54: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Figure S12. Absorption spectra of bromide ligated structures.

0

1

2

3

4

5

6

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 1 / Au9(PH3)8Br2+

0

1

2

3

4

5

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 2 / Au9(PH3)8Br2+

0

1

2

3

4

5

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d2h 3 / Au9(PH3)8Br2+

0

1

2

3

4

5

6

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d4d 1 / Au9(PH3)8Br2+

0

1

2

3

4

2 3 4 5

Abso

rptio

n In

tens

ity

Energy (eV)

d4d 2 / Au9(PH3)8Br2+

Page 55: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Table S10. Excited states of Au9(PH3)8Br2+ (D4d symmetry (d4d 2)). All excited states between 2.0 and 3.9 eV with oscillator strengths larger than 0.01 are listed.

Energy (eV) f Superatomic character (%)

Weight (%) Transition

2.34 0.0339 84.39 62.73 HOMO LUMO +1

17.63 HOMO -1 LUMO

11.06 HOMO -2 LUMO +1

2.69 0.0357 45.68 42.88 HOMO LUMO +2

20.36 HOMO -3 LUMO

15.00 HOMO -4 LUMO +1

12.61 HOMO -2 LUMO +2

2.79 0.1116 15.63 46.94 HOMO -4 LUMO

25.97 HOMO -2 LUMO +1

3.08 0.0618 3.65 91.20 HOMO -4 LUMO +2

3.41 0.0541 36.97 51.81 HOMO -5 LUMO

22.73 HOMO LUMO +4

12.13 HOMO LUMO +5

3.43 0.0655 14.17 71.71 HOMO -5 LUMO +1

8.89 HOMO -2 LUMO +3

3.54 0.0044 5.32 90.80 HOMO -1 LUMO +5

3.58 0.0362 48.84 47.42 HOMO LUMO +5

25.76 HOMO -5 LUMO +2

14.99 HOMO -2 LUMO +4

3.78 0.159 22.96 37.43 HOMO -5 LUMO +2

19.04 HOMO LUMO +5

9.21 HOMO -8 LUMO +1

Page 56: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Figure S13. Kohn–Sham orbital energy level diagram for the Au9(PH3)8Br2+ structure. Each KS orbital is drawn to indicate the relative contributions (line length with color labels) of the atomic orbitals of Au (sp) in gold, Br (s) in green, Au (d) in blue, and P (sp) in purple. The y axis shows the M.O. and energy level. X axis shows the percentage of the specific atom contributions.

Page 57: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Cartesian Coordinates of the Most Stable Structures

Au9(PH3)83+ (D2h symmetry)

Au 0.000000000 0.000000000 0.000000000 Au -2.390600000 0.000000000 1.389600000 Au 2.390600000 0.000000000 1.389600000 Au 2.390600000 0.000000000 -1.389600000 Au -2.390600000 0.000000000 -1.389600000 Au -1.420800000 2.381800000 0.000000000 Au -1.420800000 -2.381800000 0.000000000 Au 1.420800000 -2.381800000 0.000000000 Au 1.420800000 2.381800000 0.000000000 P -3.661800000 0.000000000 3.435500000 P 3.661800000 0.000000000 3.435500000 P 3.661800000 0.000000000 -3.435500000 P -3.661800000 0.000000000 -3.435500000 P -2.667500000 4.505800000 0.000000000 P -2.667500000 -4.505800000 0.000000000 P 2.667500000 -4.505800000 0.000000000 P 2.667500000 4.505800000 0.000000000 H -1.918100000 -5.712600000 0.000000000 H -3.542800000 -4.754600000 1.091000000 H -3.542800000 -4.754600000 -1.091000000 H 3.542800000 4.754600000 1.091000000 H 3.542800000 4.754600000 -1.091000000 H 1.918100000 5.712600000 0.000000000 H 3.542800000 -4.754600000 1.091000000 H 1.918100000 -5.712600000 0.000000000 H 3.542800000 -4.754600000 -1.091000000 H -3.542800000 4.754600000 -1.091000000 H -3.542800000 4.754600000 1.091000000 H -1.918100000 5.712600000 0.000000000 H 5.070700000 0.000000000 3.275400000 H 3.477600000 1.097600000 4.314000000 H 3.477600000 -1.097600000 4.314000000 H -3.477600000 1.097600000 -4.314000000 H -3.477600000 -1.097600000 -4.314000000 H -5.070700000 0.000000000 -3.275400000 H 3.477600000 1.097600000 -4.314000000 H 5.070700000 0.000000000 -3.275400000 H 3.477600000 -1.097600000 -4.314000000 H -3.477600000 -1.097600000 4.314000000 H -3.477600000 1.097600000 4.314000000 H -5.070700000 0.000000000 3.275400000

Page 58: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Au9(PH3)83+ (D4d symmetry)

Au 0.000000000 0.000000000 0.000000000 Au 2.330600000 0.965400000 -1.035400000 Au -2.330600000 -0.965400000 -1.035400000 Au -0.965400000 2.330600000 -1.035400000 Au 0.965400000 -2.330600000 -1.035400000 Au -2.330600000 0.965400000 1.035400000 Au 2.330600000 -0.965400000 1.035400000 Au 0.965400000 2.330600000 1.035400000 Au -0.965400000 -2.330600000 1.035400000 P 3.975800000 1.646800000 -2.637100000 P -3.975800000 -1.646800000 -2.637100000 P -1.646800000 3.975800000 -2.637100000 P 1.646800000 -3.975800000 -2.637100000 P -3.975800000 1.646800000 2.637100000 P 3.975800000 -1.646800000 2.637100000 P 1.646800000 3.975800000 2.637100000 P -1.646800000 -3.975800000 2.637100000 H -4.384000000 -2.985800000 -2.594900000 H -5.211200000 -0.988600000 -2.594900000 H -3.645500000 -1.510000000 -3.991100000 H 3.645500000 1.510000000 -3.991100000 H 4.384000000 2.985800000 -2.594900000 H 5.211200000 0.988600000 -2.594900000 H 2.985800000 -4.384000000 -2.594900000 H 0.988600000 -5.211200000 -2.594900000 H 1.510000000 -3.645500000 -3.991100000 H -1.510000000 3.645500000 -3.991100000 H -2.985800000 4.384000000 -2.594900000 H -0.988600000 5.211200000 -2.594900000 H 4.384000000 -2.985800000 2.594900000 H 5.211200000 -0.988600000 2.594900000 H 3.645500000 -1.510000000 3.991100000 H -3.645500000 1.510000000 3.991100000 H -4.384000000 2.985800000 2.594900000 H -5.211200000 0.988600000 2.594900000 H -2.985800000 -4.384000000 2.594900000 H -0.988600000 -5.211200000 2.594900000 H -1.510000000 -3.645500000 3.991100000 H 1.510000000 3.645500000 3.991100000 H 2.985800000 4.384000000 2.594900000 H 0.988600000 5.211200000 2.594900000

Page 59: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Au9(PH3)8H2+ (D2h symmetry (d2h 2))

Au -0.020900000 0.052600000 -1.036800000 Au -1.343400000 0.716900000 1.226400000 Au 1.336100000 -0.636200000 1.195400000 Au 2.633300000 0.349700000 -1.269600000 Au -2.677500000 -0.266300000 -1.206500000 Au -1.338700000 2.373300000 -1.271700000 Au -1.088200000 -2.104600000 0.390100000 Au 1.273800000 -2.274600000 -1.317000000 Au 1.065200000 2.186300000 0.392300000 P -2.642300000 1.226900000 3.119100000 P 2.606800000 -1.213600000 3.095300000 P 4.832600000 0.702300000 -1.964300000 P -4.852000000 -0.735800000 -1.925900000 P -2.316800000 4.345300000 -2.070500000 P -2.015400000 -4.104100000 1.309500000 P 2.241500000 -4.236300000 -2.159600000 P 2.030600000 4.203500000 1.211400000 H -1.131000000 -5.158700000 1.610200000 H -2.727200000 -4.054000000 2.524400000 H -2.950100000 -4.812800000 0.529500000 H 2.678800000 4.218300000 2.462100000 H 3.035300000 4.804100000 0.427700000 H 1.186200000 5.318600000 1.380600000 H 3.433600000 -4.709300000 -1.578100000 H 1.481000000 -5.420500000 -2.144600000 H 2.625000000 -4.214000000 -3.513800000 H -2.745500000 4.309300000 -3.410700000 H -3.480400000 4.848500000 -1.458000000 H -1.536200000 5.515900000 -2.102700000 H 3.998000000 -1.005700000 3.027600000 H 2.337300000 -0.556400000 4.311400000 H 2.596700000 -2.547100000 3.547700000 H -5.218000000 -0.265300000 -3.201100000 H -5.193900000 -2.093800000 -2.070300000 H -5.957500000 -0.288700000 -1.177400000 H 5.450300000 1.920100000 -1.621500000 H 5.835400000 -0.196000000 -1.552700000 H 5.063300000 0.702900000 -3.352800000 H -2.434000000 0.499000000 4.306100000 H -2.616500000 2.534500000 3.642200000 H -4.034500000 1.061500000 2.988600000 H -0.040800000 0.049600000 -2.659100000

Page 60: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Au9(PH3)8F2+ (D4d symmetry (d4d 2))

Au 0.196700000 -0.356700000 -0.220600000 Au 1.936700000 1.378600000 -1.565400000 Au -2.425900000 -0.257100000 -1.389500000 Au -0.794600000 2.137100000 -0.985100000 Au 1.197800000 -2.847700000 0.171600000 Au -1.977500000 0.630200000 1.122200000 Au 2.835300000 -0.533400000 0.366400000 Au 1.203300000 1.828900000 1.045800000 Au -1.562500000 -2.229600000 0.585900000 P 2.892500000 1.847400000 -3.661100000 P -3.461600000 -0.709400000 -3.450100000 P -1.516000000 4.269200000 -1.729300000 P 2.275100000 -4.969000000 0.322300000 P -3.837600000 1.218700000 2.518600000 P 5.087700000 -0.515400000 1.027900000 P 1.450500000 2.920800000 3.102700000 P -3.086400000 -3.687800000 1.625300000 H -4.165500000 -1.920200000 -3.580000000 H -4.445400000 0.193400000 -3.892800000 H -2.651600000 -0.771800000 -4.598600000 H 2.036500000 1.878800000 -4.776900000 H 3.551000000 3.080200000 -3.820300000 H 3.885800000 0.973800000 -4.139800000 H 3.384900000 -5.052500000 1.183400000 H 1.541200000 -6.080200000 0.779200000 H 2.837100000 -5.517500000 -0.845700000 H -1.534600000 4.544500000 -3.109300000 H -2.813400000 4.690500000 -1.384200000 H -0.773500000 5.373600000 -1.273100000 H 5.863300000 -1.678400000 0.863300000 H 5.937300000 0.425700000 0.418200000 H 5.348900000 -0.239700000 2.382100000 H -3.589600000 0.989100000 3.883500000 H -4.222400000 2.571400000 2.570700000 H -5.108300000 0.624600000 2.394800000 H -4.413900000 -3.730200000 1.161200000 H -2.795700000 -5.064300000 1.658800000 H -3.308200000 -3.451100000 2.993700000 H 1.557000000 2.040200000 4.187700000 H 2.540700000 3.780000000 3.350000000 H 0.381000000 3.734800000 3.498400000 F -0.615000000 1.281000000 2.774900000

Page 61: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Au9(PH3)8Cl2+ (D4d symmetry (d4d 2))

Au 0.135000000 -0.395500000 0.017100000 Au 1.862900000 1.431400000 -1.326000000 Au -2.433500000 -0.281700000 -1.274700000 Au -0.893100000 2.089500000 -0.641400000 Au 1.180700000 -2.891500000 0.214500000 Au -2.127900000 0.392100000 1.356200000 Au 2.791000000 -0.577400000 0.514700000 Au 1.211100000 1.787100000 1.310800000 Au -1.609900000 -2.401500000 0.497500000 P 2.900400000 1.801800000 -3.451400000 P -3.481500000 -0.632200000 -3.406900000 P -1.776100000 4.225400000 -1.336600000 P 2.255300000 -5.067600000 0.213400000 P -4.056800000 1.004800000 2.686100000 P 5.060600000 -0.610900000 1.291400000 P 2.169400000 3.615400000 2.576500000 P -3.203600000 -4.029000000 1.258700000 H -4.210800000 -1.813500000 -3.590000000 H -4.436100000 0.299500000 -3.836600000 H -2.665400000 -0.675600000 -4.544100000 H 2.104600000 1.787800000 -4.604300000 H 3.573600000 3.014700000 -3.644900000 H 3.904100000 0.908100000 -3.845500000 H 3.081400000 -5.385400000 1.299900000 H 1.462700000 -6.223600000 0.188400000 H 3.128800000 -5.369700000 -0.839900000 H -1.633500000 4.608000000 -2.676700000 H -3.145700000 4.474100000 -1.172300000 H -1.265400000 5.373900000 -0.716200000 H 5.672200000 -1.850000000 1.523400000 H 6.049600000 -0.005600000 0.505200000 H 5.319800000 0.009400000 2.520200000 H -3.960400000 0.733600000 4.056500000 H -4.395400000 2.362900000 2.743500000 H -5.326000000 0.468400000 2.428600000 H -4.406400000 -4.172600000 0.554300000 H -2.814100000 -5.372600000 1.331300000 H -3.708300000 -3.861300000 2.554900000 H 2.763200000 3.275600000 3.799000000 H 3.173100000 4.442400000 2.054200000 H 1.274300000 4.594400000 3.025700000 Cl -0.326200000 0.852800000 3.272600000

Page 62: Hydride- and Halide-Substituted Au9(PH3)83+ Nanoclusters

Au9(PH3)8Br2+ (D4d symmetry (d4d 2))

Au 0.141500000 -0.286200000 0.073400000 Au 1.941400000 1.428700000 -1.294000000 Au -2.348200000 -0.338200000 -1.287400000 Au -0.870800000 2.140800000 -0.754300000 Au 1.173000000 -2.784300000 0.160500000 Au -2.177200000 0.569300000 1.296500000 Au 2.752100000 -0.474500000 0.736200000 Au 1.182100000 1.964800000 1.279500000 Au -1.567200000 -2.276200000 0.714300000 P 3.020300000 1.766900000 -3.405400000 P -3.374300000 -0.906900000 -3.375400000 P -1.746900000 4.212500000 -1.633600000 P 2.054100000 -5.020900000 -0.101000000 P -4.232400000 1.195800000 2.414800000 P 4.928600000 -0.695100000 1.718200000 P 2.195000000 3.902300000 2.331500000 P -3.079600000 -3.897400000 1.627400000 H -3.202400000 -2.203900000 -3.875800000 H -4.768900000 -0.801300000 -3.461200000 H -3.019500000 -0.168100000 -4.512400000 H 2.227900000 1.742900000 -4.560600000 H 3.697300000 2.975500000 -3.618600000 H 4.021900000 0.865800000 -3.787500000 H 3.431100000 -5.229400000 0.061000000 H 1.567600000 -6.045800000 0.721600000 H 1.884100000 -5.630600000 -1.352000000 H -2.014000000 4.305200000 -3.006300000 H -2.974400000 4.667700000 -1.132200000 H -0.994000000 5.384200000 -1.472300000 H 5.563800000 -1.942000000 1.643200000 H 5.975000000 0.122300000 1.267600000 H 5.017300000 -0.462400000 3.096200000 H -4.569600000 0.449500000 3.551900000 H -4.274700000 2.481900000 2.968800000 H -5.471500000 1.185400000 1.758900000 H -4.338900000 -4.059700000 1.033200000 H -2.668900000 -5.236500000 1.666700000 H -3.463600000 -3.725100000 2.963100000 H 2.755400000 3.702500000 3.600400000 H 3.246600000 4.605600000 1.726400000 H 1.344800000 4.976300000 2.626400000 Br -0.399800000 1.071400000 3.341500000