I. INTRODUCTION
Nanometer-sized atomic clusters constitute a unique phase of matter that lies between the atoms and their bulk. Early studies on free atomic clusters have revealed that their structures and properties do not resemble those of their corresponding bulk materials. In addition, the properties of nanoclusters vary discontinuously and sensitively with the size, shape, and composition.1 These features have made the clusters promising and attractive building units for novel functional materials. A “superatom” concept has been developed based on the findings that clusters of certain elements, such as Na, Au, and Al, mimic the electronic structures of atoms in the Periodic Table.2–5 This concept explains the high stability of magic clusters by the closure of the electronic shells and has motivated researchers to synthesize a new class of materials using superatomic clusters as the building blocks, just as a variety of molecules are formed using atoms as the building blocks.
The key requirement for the nanocluster synthesis is atomic precision because the addition or removal of a single atom can drastically change the structure and properties of the nanocluster. The challenge of synthesis has been accomplished with gold clusters protected by monolayers of ligands, such as phosphines, thiolates (RS), alkynyls, and N-heterocyclic carbenes (Scheme 1).6–9 The superatomic concept can be also applied to the electronic structures of these monolayer-protected clusters (MPCs) of gold and silver.10–12 Another important requirement for establishing the structure–property correlation is the determination of the geometric structures of the metallic core and ligand layer (Scheme 1).13 Single crystal x-ray diffraction (SCXRD) analysis has been successfully applied to Au/Ag MPCs. Over the past two decades, the number of MPCs whose structures have been elucidated with atomic-level precision has been rapidly increasing.7,8 These studies revealed that the properties of MPCs can be controlled over a wide range by the size, composition, and atomic packing of the metal core as well as the nature of the ligands. The main purpose of this special issue is to provide a glimpse into the current status of development of novel nanoclusters and superatomic materials.
II. SUMMARY OF AREA COVERED
The following is a brief overview of the main topics covered in this special issue.
A. Geometric and electronic structures
Atomically precise synthesis and structure determination by SCXRD provide the basis for investigating the property–structure correlations of the MPCs. Table I summarizes the chemical compositions and core structures of MPCs experimentally determined in this special issue.14–20 It can be seen that icosahedral (Ih) M13 is the ubiquitous building unit of the ultrasmall metal clusters. In addition, the number and position of the dopant atoms are defined by the nature of the element and the specific interaction with the ligand. Ni8 adopts a unique folded sheet structure due to the ligand addition.20 Mullins et al. demonstrated, via first-principles density functional theory (DFT) calculations, that the 60-atom hollow shell of chiral-icosahedral symmetry (I-Au60) is robust and well retained in various bilayer structures, such as I-C60@I-Au60, I-Au32@I-Au602+, and Au60(MgCp)12.21 Nelli et al. revealed by molecular dynamics (MD) and metadynamics the unexpected diffusion pathways of Au or Ag impurity atoms within icosahedral Co or Cu clusters: the displacement of the impurity is coupled with the creation of vacancies in the central part of the cluster due to the presence of nonhomogeneous compressive stress.22
Formulaa . | Core structureb . | Reference . |
---|---|---|
[Au13L5Cl2]3+ (L = diarsine, diphosphine) | Ih Au13 | 14 |
[Ag24Pd1(SPhMe2)18]2− | Ih Pd@Ag12 | 15 |
[Au9Ag12(SAdm)4(dppm)6Cl6]3+ | Ih Au@Au6Ag6 | 16 |
Au11Ag6(dppm)4(SAdm)4(CN)4 | Ih Au@Au4Ag8 | 16 |
Au8Ag17(PPh3)10Cl10 | bIh (Au@Au3Ag9)2 | 17 |
[Ag23Pt2(PPh3)10Cl7]0 | bIh (Pt@Ag12)2–Ag | 18 |
[CuxAg20−x{S2P(OiPr)2}12] (x = 3, 4) | Ih Ag13 | 19 |
[Cu4Ag17{S2P(OiPr)2}12]+ | Ih Ag13 | 19 |
[Ni8(CNtBu)12]+ | Folded Ni8 sheet | 20 |
Formulaa . | Core structureb . | Reference . |
---|---|---|
[Au13L5Cl2]3+ (L = diarsine, diphosphine) | Ih Au13 | 14 |
[Ag24Pd1(SPhMe2)18]2− | Ih Pd@Ag12 | 15 |
[Au9Ag12(SAdm)4(dppm)6Cl6]3+ | Ih Au@Au6Ag6 | 16 |
Au11Ag6(dppm)4(SAdm)4(CN)4 | Ih Au@Au4Ag8 | 16 |
Au8Ag17(PPh3)10Cl10 | bIh (Au@Au3Ag9)2 | 17 |
[Ag23Pt2(PPh3)10Cl7]0 | bIh (Pt@Ag12)2–Ag | 18 |
[CuxAg20−x{S2P(OiPr)2}12] (x = 3, 4) | Ih Ag13 | 19 |
[Cu4Ag17{S2P(OiPr)2}12]+ | Ih Ag13 | 19 |
[Ni8(CNtBu)12]+ | Folded Ni8 sheet | 20 |
SAdm = adamantanethiolate and dppm = 1,1-bis(diphenylphosphino)methane.
Ih = icosahedron and bIh = biicosahedron shared by vertex Ag.
The structural motif and doping remarkably affect the electronic structures of MPCs as evidenced by optical and photophysical measurements. Shichibu et al. demonstrated that a subtle deformation of the Au13 core resulted in a notable difference in the optical absorption and photoluminescence properties.14 Miyamoto et al. theoretically demonstrated a substantial difference in the absorption onset energies between Au13 clusters with Ih and face-centered cubic (fcc) structures and ascribed it to the difference in the natural charges of the central Au atoms.23 Fagan et al. reproduced the high-resolution absorption spectra of [Au9(PPh3)8]3+ and [Au8(PPh3)7]2+ in the gas phase using time-dependent DFT calculations and particle-in-a-box analysis for an asymmetrical ellipsoidal superatomic core, respectively.24 Wang et al. theoretically proposed a Au2S network model to explain and predict the structural origin, evolution, and formation mechanism of Agn(SR)m containing quasi-fcc cores.25 Medves et al. proposed efficient algorithms, i.e., the complex polarizability polTDDFT approach and the hybrid-diagonal approximation, to predict the optical spectra of ligand-protected alloy clusters, such as Ag24Au(DMBT)18 (DMBT = 2,4-dimethylbenzenethiolate).26
B. Photophysical properties
One of the attractive properties of MPCs is photoluminescence (PL), which originates from the quantized, molecular-like electronic structures. Much effort has been made to tune the emission wavelengths and enhance the quantum yields (QYs) through the precise control of the structural parameters. A fundamental understanding of the relaxation dynamics has been deepened by joint studies on time-resolved spectroscopy and theoretical calculations.
Su et al. demonstrated that photocurrent-generating and PL properties are significantly affected by the difference in the core structures using two Ag60 clusters having different core–shell architectures (Ag12@Ag48 vs Ag14@Ag46).27 Petty et al. reported that DNA-bound Ag106+ exhibited prompt visible emission and metastable near-infrared PL and demonstrated that the metastable states could be tuned by synthetically modifying the DNA.28 Zhang et al. demonstrated that the fluorescent properties of Ag106+ are significantly affected by the host, a single-stranded DNA C4AC4TC3XT4 with guanosine and inosine.29 These results suggested that the brightness of the DNA-bound Ag cluster can be modulated by a single-site mutation of the DNA. Bootharaju et al. self-assembled Ag2Cl2(dppe)2 clusters [dppe = 1,2-bis(diphenylphosphino)ethane] through the ligand-exchange-induced transformation of [Pt2Ag23Cl7(PPh3)10] and observed intense PL with a QY of ∼18%, which is attributed to the metallophilic interactions and rigidification of the ligand shell.30 Havenridge and Aikens theoretically assigned dual PL from Au14Cd(SAdm)12 (SAdm = adamantanethiolate) to emissions from two minima on the first excited state.31 Yousefalizadeh and Stamplecoskie demonstrated diversity in terms of the excited state energy, relaxation dynamics, and PL properties using excitation emission matrix spectroscopy and pump–probe transient absorption spectroscopy on Au25(SR)18, Ag25(SR)18, and their alloys.32 Jeffries et al. resolved state-to-state energy relaxation dynamics in Au38(SC6H13)24 by employing femtosecond transient absorption and two-dimensional electronic spectroscopy.33
C. Redox and electronic properties
Redox and electronic properties are other interesting properties of MPCs arising from nonmetallic (molecular-like) electronic structures. Cowan et al. introduced novel structure–property relationships that predict electronic properties, such as the ionization potential and electron affinity, of thiolate-protected AgAu nanoclusters based on physically relevant descriptors.34 Chen et al. reported that [Ag16Au13L24]3− (L = alkynyl) with an icosahedral Au@Ag12 core underwent one-way oxidation by losing three electrons.35 Cesari et al. studied the redox properties of [Rh12E(CO)27]n− (n = 4 for E = Ge or Sn; n = 3 for E = Sb or Bi) with E@Ag12 icosahedral cores and showed that three clusters (all except E = Sn) act as electron reservoirs.36 Fiedler et al. demonstrated the emergence of electronic properties characteristic of a charge interaction between superatoms within the solid state material [Au9(PPh3)8](NO3)3−x(C60)x.37 López-Estrada et al. investigated the electronic structure and magnetically induced currents of doped nanoclusters [M@Au8(PPh3)8]q (M = Pd, Pt, Ag, Au, Cd, Hg, Ir, and Rh; and q = 0, +1, +2) having 6–10 delocalized valence electrons and found that the aromaticity of eight-electron cubic clusters increased in the order of M = Au < Pt < Pd.38
D. Catalytic properties
The catalytic application of metal nanoclusters could have a huge impact on the realization of a sustainable society by taking advantage of the high and selective reactivities. However, the development of cluster catalysts according to a rational design principle has been hampered by the challenges of the synthesis: size and compositions must be controlled with atomic precision. Typical approaches to this goal are as follows: (1) adsorption of MPCs with well-defined compositions on a solid support, (2) removal of ligands from MPCs with well-defined compositions adsorbed on the support, and (3) soft-landing of mass-selected cluster ions onto solid supports. Choi et al. demonstrated the production of syngas with controlled H2/CO ratios by CO2 electroreduction on specifically formulated mixtures of [Au25(SC6H13)18]− and [PtAu24(SC6H13)18]0 in aqueous media.39 Kawamura et al. found that Au144(pMBA)60 (pMBA = p-mercaptobenzoic acid) deposited on TiO2 enhanced the sonocatalytic activity for OH production by suppressing the electron–hole recombination.40 Yang et al. revealed that Au38S2(SAdm)20 with a body-centered cubic (bcc) Au14 core outperformed Au30(SAdm)18 and Au21(SAdm)15 with hexagonal close-packed (hcp) Au17 and fcc Au23 cores, respectively, in the CO2 cycloaddition with epoxides toward cyclic carbonates.41 Chen et al. successfully removed the PET (phenylethanethiolate) ligands from Au25(PET)18 supported on sulfur-doped graphene by thermal annealing and electrochemical biasing and showed that the resulting Au25 clusters catalyzed the electroreduction of CO2 to CO.42 Alotabi et al. demonstrated that the thermal-induced agglomeration of [Au9(PPh3)8]3+ on TiO2 could be suppressed by the photodeposition of the Cr2O3 layer.43 Valtera et al. found that Cu4 on titania prepared by soft-landing showed the highest activity for oxidative dehydrogenation of cyclohexene to benzene among Cun (n = 1–7).44
E. Ligand bonding
Ligand bonding is the key to the stability of MPCs. Matsuyama et al. revealed by thermogravimetry, differential thermal analysis, and mass spectrometric analysis that the ligand desorption temperature of [MAu8(PPh3)8]2+ (M = Pd, Pt) is higher than that of [Au9(PPh3)8]3+ because of the formation of a large bonding index of M–Au and a change in the Au–PPh3 bonding energy by the doping.45 Wang and Gao theoretically investigated the desorption energies of -SR-(Au-SR)n- and the resulting effects on the geometries and electronic structures of the remaining Au clusters.46 They found that the detachment of longer motifs normally required more energy and resulted in an upshift of the HOMO and a reduction in the HOMO–LUMO gap. Liu and Jiang predicted by DFT calculations that carboxylate groups could be promising ligands to support atomically precise metal clusters, especially for Cu and Ag.47
F. New type of nanoclusters
The conventional synthesis of MPCs is based on the chemical reduction of the relevant metal–organic compounds. New approaches have been developed to synthesize novel species that cannot be obtained by the conventional chemical reduction. Truttmann et al. reported the selective synthesis of Au16(PET)14 by the ligand exchange process of Au15(SG)13 (SG = glutathione).48 Emori et al. found that the irradiation of atmospheric pressure plasma to [Au9(PPh3)8]3+ in methanol yielded novel CN-containing clusters [Au9(PPh3)7(CN)1]2+ and [Au10(PPh3)7(CN)2]2+ as the main products via the sequential addition of AuCN unit(s) formed in situ.49 Dong et al. synthesized a highly stable gold hydride nanocluster, [Au22H3(dppee)7]3+ (dppee = bis(2-diphenylphosphino) ethyl ether), and proposed that the Au22 core is composed of two Au11 units bonded via two triangular faces, creating six uncoordinated Au sites that are bridged by three H atoms.50
G. Other systems
Daly et al. produced the novel argentate aggregates, [AgnPhn+1]−, by transmetalation reactions between phenyl lithium and silver cyanide.51 Mass-selected UV photodissociation spectroscopy and theoretical calculations suggested novel organometallic characteristics built from Ag2Ph subunits. Fang et al. theoretically predicted that the binding energy of the noble gas (Ng = Ne, Ar, Kr, and Xe) atoms becomes stronger with an increase in the cluster size and the electron affinity of the terminal ligands X in B12X11(Ng) (X = H, CN, and BO) and B12X10(Ng)2 (X = CN and BO).52 Nguyen et al. examined the structures and absorption spectra for CdS nanoplatelets (NPLs) with thicknesses of two and three monolayers (2 MLs and 3 MLs) and passivated by formate and acetate ligands.53 The spectral redshift for 3-ML CdS NPLs is attributed to the electron delocalization due to the expansion of the nanoplatelets in the lateral and vertical directions. Park et al. controlled the stoichiometry of both anions (P, As, S, and Se) and cations (In and Zn) at the interface of InP/ZnSe quantum dots (QDs) and correlated these changes with the resultant steady-state and time-resolved optical properties of the nanocrystals.54 The results show that the dynamics are not dramatically impacted at the ns timescale while the PL QYs are highly sensitive to the interfacial composition. Sikorska and Gaston theoretically predicted that the bandgap of a perovskite CsPbBr3 is narrowed by replacing the Cs+ cation with bimetallic superalkalis (LiMg, NaMg, LiCa, and NaCa).55
III. SUMMARY
This special issue highlights recent advances in the emerging field of atomically precise nanoclusters and cluster-assembled materials. We introduced the background and motivation of the research and provided a brief overview on how to control the structures and properties with a focus on the monolayer-protected metal clusters. We hope that the papers in this special issue will convince the readers of the bright future and opportunities in the materials science of nanoclusters and superatoms. We believe that new experimental and theoretical methods as well as interdisciplinary approaches will accelerate the exploration of new materials and deepen the understanding of the origin of their unique properties.
ACKNOWLEDGMENTS
We thank all the authors and reviewers who contributed to this issue, the journal editors Professor Lai-Sheng Wang and Professor Xiaoyang Zhu, and the editorial staff who assisted throughout its preparation.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
DATA AVAILABILITY
Data sharing is not applicable to this article as no new data were created or analyzed in this study.