Skip to main content
Science Advances logoLink to Science Advances
. 2023 Aug 16;9(33):eadi0214. doi: 10.1126/sciadv.adi0214

Rh19: A high-spin super-octahedron cluster

Yuhan Jia 1,2,, Cong-Qiao Xu 3,, Chaonan Cui 1,, Lijun Geng 1, Hanyu Zhang 1, Yang-Yang Zhang 4, Shiquan Lin 1, Jiannian Yao 1,5, Zhixun Luo 1,2,*, Jun Li 3,4,*
PMCID: PMC10431703  PMID: 37585530

Abstract

Probing atomic clusters with magic numbers is of supreme importance but challenging in cluster science. Pronounced stability of a metal cluster often arises from coincident geometric and electronic shell closures. However, transition metal clusters do not simply abide by this constraint. Here, we report the finding of a magic-number cluster Rh19 with prominent inertness in the sufficient gas-collision reactions. Photoelectron spectroscopy experiments and global-minimum structure search have determined the geometry of Rh19 to be a regular Oh‑[Rh@Rh12@Rh6] with unusual high-spin electronic configuration. The distinct stability of such a strongly magnetic cluster Rh19 consisting of a nonmagnetic element is fully unveiled on the basis of its unique bonding nature and superatomic states. The 1-nanometer–sized Oh-Rh19 cluster corresponds to a fragment of the face-centered cubic lattice of bulk rhodium but with altered magnetism and electronic property. This cluster features exceptional electron-spin state isomers confirmed in photoelectron spectra and suggests potential applications in atomically precise manufacturing involving spintronics and quantum computing.


A magic-number metal cluster Rh19 is discovered with regular Oh-[Rh@Rh12@Rh6] and unusual high-spin electronic configuration.

INTRODUCTION

Developing functional materials with anticipated stability, regular structures, and well-defined components is one of the most important subjects in chemistry. Metal-atomic clusters hold promise in a variety of applications because of their versatile properties that can be tailored by cluster structure, size, and composition. Ongoing efforts in this field have been devoted to synthesizing monolayer-protected metal clusters and exploring pure metal clusters at reduced sizes with atomic precision (13). By solving the Schrödinger equation for a spherical nucleus in average central potentials, many stable clusters have been predicted with magic numbers of delocalized valence electrons (46). Meanwhile, metal clusters also find enhanced stability for those of favorable geometry within Mackay icosahedrons (7), i.e., N=1+i=1n(10i2+2), where n refers to the number of geometric shells, while N corresponds to magic numbers of atoms at 13, 55, 147···. Similar situations also allow metal clusters to exhibit magic numbers of atoms within Marks decahedra (8) and Leary tetrahedra (9, 10). However, it remains elusive to fully understand the diversity of atomic packing and stability of transition metal (TM) clusters.

Recent advances have led to the superatom concept (1116), which enables us to better understand itinerant electrons with delocalized orbitals in metal clusters akin to isolated atoms. For example, Al13 was found to exhibit an electronic configuration as a halogen atom (17, 18); in addition, the VNa8 cluster was demonstrated as a magnetic superatom showing a filled d subshell with a magnetic moment at 5.0 Bohr magneton (μB) (19). Because of the relative activity and less-contracted radial extension of the outermost d and f electrons, heavy TM clusters may undergo a shape relaxation without restriction to spherical symmetry (20), and the filling of superatomic orbitals may not simply mimic atoms in terms of Pauli exclusion principle and Hund’s rule (21), showing a diversity of structures and electronic/magnetic properties (10).

Rhodium, with an unfilled d electron shell configuration of 4d85s1, allows for diverse coordination numbers and valence states (e.g., +1, +3, and + 4). As the Rh─O bond dissociation energy is much larger than the Rh─Rh bond (~405 versus ~235 kJ mol−1), and Rhn clusters can readily coordinate with O2, it is challenging to prepare pure Rhn clusters, especially the anions (22). It is therefore difficult to identify magic-number rhodium clusters by the oxygen-etching reactions (18, 23, 24). Among the few studies devoted to understanding the reactivity of pure rhodium clusters (25), size effects in the reactions with N2O and alkanes were noted under single-collision conditions (26, 27), showing local minima of reaction rates at Rh5+, Rh19+, and Rh28+. Inasmuch as rhodium and its compounds are ubiquitous in chemical synthesis and petroleum engineering, radiology, and dentistry, as well as electrodes and electrical devices, understanding the structural stability and size-dependent properties is crucial for their applications.

We have recently developed an integrated instrument of time-of-flight mass spectrometry (TOF-MS) and photoelectron spectroscopy (PES) (figs. S1 and S2) that provides experimental techniques for preparing pure metal clusters and studying their reactivity and electronic structures (28, 29). Here, we report the formation and reactions of anionic Rhn (n = 3 to 33) clusters with a few typical gases (including O2, CO2, CH4, and CH3Br). A magic-number cluster Rh19 is identified, and its photoelectron spectra are measured. Through computational chemistry modeling, a high-spin super-octahedron structure of Oh Rh19 is determined, and the agreement between experiment and theory suggests the existence of a series of electron-spin state isomers (ESSIs) of the Rh19 cluster.

RESULTS

The finding of magic-number Rh19 cluster

The pure Rhn (n = 3 to 33) clusters were prepared via a laser evaporation (LaVa) source and reacted with O2, CO2, etc. in the downstream flow tube reactor combined with the TOF-MS instrument. It is found that the reactions of Rhn with O2 are too reactive to identify magic-number species (fig. S3), but the reactions of Rhn with CO2 exhibit interesting size dependence, as shown in Fig. 1A. Notably, most of these Rhn clusters react with CO2 to form Rhn(CO2)m adducts, whereas Rh8,9 and Rh19 are relatively inert. Especially noteworthy is the Rh19 cluster, which shows up progressively with an increased dose of the CO2 reactant and becomes dominant in the mass distribution (for details, see fig. S4, ESI). To verify the size-dependent reactivity and distinct stability of the Rh19, we carried out additional tests for the reactions with CH4 and CH3Br. It turns out that the reaction of Rhn with CH3Br (Fig. 1B) exhibits a similar phenomenon as Rhn reacting with CO2, although the reactivity with CH4 is too weak (fig. S5). To our surprise, Rh19 emerges again as the only dominant species in the mass spectra upon introducing a large dose of CH3Br gas reactant (fig. S6). In contrast, the neighboring counterparts Rh16–18 are rather reactive with CH3Br to form the products of RhnCHBr and RhnBr. While several previous studies have been devoted to understanding the reactivity of rhodium clusters (22), a magic number of rhodium clusters have not been identified. Our experimental observations are reminiscent of the previous findings of C60 and Al13 (17, 30) and reveal the prominent inertness of Rh19 in surviving sufficient gas collision reactions.

Fig. 1. MS observation.

Fig. 1.

Nascent mass distribution of rhodium cluster anions observed via TOF-MS, and after exposure to CO2 (A) and CH3Br (B) reactants in the downstream flow tube reactor, with different doses of the reactant gas (5% in helium) controlled by a 10 Hz pulsed valve, respectively. A 532-nm laser with a pulse energy of 15 to 35 mJ was used to generate the Rhn clusters. The Rhn are labeled with the number N.

Structure determination of octahedron Rh19

Figure 2A presents the PES of Rh19 taken at 355- and 266-nm laser excitation, respectively. The peaks at 2.72 and 3.45 eV are assigned to the vertical detachments of ground-state Rh19 to yield the ground state (X) and the first excited state (A) of Rh19, respectively. As shown below, the distinct spectral bands of the experimental PES spectra are well consistent with the simulated photoelectron detachment energies (fig. S10 and table S4). The profile with a broadband is essentially associated with the dense electronic energy levels of the neutral Rh19 cluster, especially the existence of ESSIs of Rh19 (infra vide).

Fig. 2. Determination of geometric structure.

Fig. 2.

(A) The experimental photoelectron spectra of the Rh19 clusters with a 355- and 266-nm laser, respectively. The peaks marked with X, A, B, and C are assigned to the electronic transition from the Oh Rh19 anion to the electron-detached Rh19 neutral, while those marked with x−1, x+1, and x+2 are associated with the ESSIs of Rh19 and Rh19 (table S2). (B) The lowest-energy structure of Rh19 determined by USPEX and TGMin, along with the van der Waals radius and (C) the Hirshfeld charge population. (D) Calculated spin density population of the Rh19 cluster. (E) Structural evolution sketch of Rh19 based on an Rh@Rh12 coordination, cubic Rh@Rh8 kernel coordination, octahedral Rh6 fusing, and tetrahedron Rh4 interpenetration.

To determine the structure and unveil the origin of distinct inertness of the Rh19 cluster, we conducted first-principles global-minimum (GM) search for the ground-state structure using the ab initio evolutionary algorithm of Universal Structure Predictor-Evolutionary Xtallography (USPEX) (31) embedded in Vienna Ab Initio Simulation Package (VASP) (32) and then optimized the structure of Rh19 with Gaussian 09 software (fig. S8). The global lowest energy structure of Rh19 corresponds to a regular octahedron, with a central Rh atom (Rhc), 12 edge atoms (Rhe), and 6 apex Rh atoms (Rha) anchored at the six square-surface centers of the Rh@Rh12 moiety. Meanwhile, an independent basin-hopping GM search by using TGMin code (33) confirmed the lowest energy structure of Rh19 within Oh symmetry and located a number of low-lying structural isomers. Notably, the density functional theory calculations indicate a ground state of 10 spin-unpaired electrons (figs. S9 and S10) with a calculated spin magnetic moments of 10.95 μB for the ground-state Rh19 (S = 10/2 with 0.58 μB per atom), which agrees well with the previously measured magnetic moments of 11.59 μB for the Rh19 neutral (i.e., S = 11/2 with 0.61 ± 0.08 μB per atom) (34). The geometric structure (Fig. 2B) of the Oh Rh19 shows a van der Waals radius of 11.2 Å, suggesting a 1 nm sized magnet of the Rh19 cluster.

From the energetics analysis (figs. S14 and S16), the highest occupied molecular orbital–lowest unoccupied molecular orbital (LUMO) gap and average Rh-Rh binding energy of the Oh Rh19 do not correspond to maximal values; however, the Oh Rh19 bears the third largest vertical electron detachment energy (VDE) and second largest adiabatic electron detachment energy among all the Rhn (n = 2 to 20) clusters. Furthermore, the Oh Rh19 shows the largest effective coordination number among these Rhn (n = 2 to 20) clusters (fig. S17); in addition, the Oh Rh19 shows the smallest energy gain for a CH3Br molecule adsorbed on the Rhn (n = 6, 17 to 20) clusters, which is consistent with the prominent inertness of Rh19 in the experimental observation. From the Hirshfeld population analysis (Fig. 2C and table S9), Rh19 shows positive charge distribution on the core Rhc atom, while negative fractional charges are evenly distributed on the 12 Rh12 shell atoms, and the 6 Rhe peripheral atoms are almost electroneutral. Further, the spin density analysis (Fig. 2D) shows that the unpaired spins mainly locate on the core-shell Rh@Rh12 instead of the six peripheral Rha atoms. The balanced electrostatic interactions and spin density between the Rhc atom and the shell Rhe atoms contribute to the stability of the [Rh@Rh12] frame, while the capped six Rha atoms resemble octahedral coordination ligands, thus rendering enhanced stability of such a high-spin octahedron cluster without being subjected to Jahn-Teller distortion. This is in contrast to the other Rhn clusters of lower symmetry (fig. S13).

From Born-Oppenheimer molecular dynamics simulations, the regular octahedral structure of Rh19 is dynamically stable up to 900 K (fig. S21). Whereas a few other 19-atom clusters were found to prefer a D5h double-icosahedron structure (fig. S18) (35), the D5h Rh19 is 1.88 eV higher in energy than the Oh structure (fig. S19). The Oh Rh19 has a smaller edge length (Oh 61.40 versus D5h 123.75 Å), also a smaller surface area (92.36 versus 98.25 Å2) and a smaller volume (63.26 versus 68.05 Å3) in geometry. The formation of a super-octahedron structure is consistent with the stacking principle based on generalized Wulff construction (36), which rationalizes the stability of such a magic-number cluster within Oh symmetry (37, 38).

Figure 2E shows the structure evolution of Rh19 based on different hypothetical growth modes by assuming that it begins with an Rh13, Rh9, Rh6, or Rh4 moiety and then grows up via vertex sharing, capping, or fusing. First, the octahedral Rh19 is viewed as an Rh13 motif capped by six Rh atoms (Rh13@Rh6), where the anchoring of six Rha atoms functions as protection to stabilize the cluster. The anchoring of six Rha atoms not only lowers the surface energy of square surfaces but also helps to balance the charge distribution thus passivating the Rh@Rh12 moiety (39, 40). Note that a metallic core of 13 atoms is also involved in many ligand-protected nanoclusters (41), showing diverse ligand accommodation and preferential growth on such a primitive cluster unit. In addition, Rh19 can be viewed as Rh9@Rh6@Rh4, for which a body-centered cubic Rh@Rh8 is formed first, and then the six faces are capped by six Rh atoms (two Rha and four Rhe atoms), followed by the addition of other four Rha atoms on the waist. This channel appears to be feasible by noting the relatively larger mass abundances of Rh8 and Rh9 in the MS experiments (Fig. 1). On the other hand, the Oh-Rh19 may also be viewed as the fusing of six Rh6 units, or tetrahedral Rh4 interpenetration, showing the likely multiple pathways in forming such an octahedral cluster.

Electronic configuration analysis

We further studied the electronic configuration of Rh19 to fully understand the origin of its distinctive stability. Figure 3A presents the projected densities of states (PDOS) for the α-orbitals (majority spin) of Rh19 (details in fig. S24), where the contributions by the 5s and 4d orbitals of the three types of atoms (Rhc, Rha, and Rhe) are displayed. It is seen that the 4d orbitals show a dominant contribution to the total DOS, while the Rhe-5s (red line) and Rha-5s (red dot line) contribute to the low-energy states and frontier states. Some of the frontier orbitals exhibit diffused superatomic orbital features (fig. S26), which is consistent with the results of the adaptive natural density partitioning (42) and electron localization function analyses (figs. S20 and S25). Among the delocalized superatomic orbitals in Rh19, the dominant contribution to the superatomic 1S orbital is made by the 5s electrons of the Rhc atom, while the 4d-based orbitals contribute to the superatomic 2S, 2P, 3S, and 3P orbitals (table S8). Especially interesting is that the superatomic 2S orbital is mainly contributed by the dz2 orbitals of the 12 Rhe atoms, while the superatomic 3S is dominated by the dz2 orbitals of the 6 Rha atoms in a local coordinate system (table S8). It is inferred that the electronic configuration within unique superatomic orbitals significantly reinforces the structure stability of the Rh19 cluster.

Fig. 3. Electronic structure.

Fig. 3.

PDOS and orbitals of Rh19, with different colors corresponding to the total density of states (TDOS) and contributions by the Rhc-5s, Rha-5s, Rhe-5s, Rhc-4d, Rha-4d, and Rhe-4d orbitals, respectively. Insets: Typical orbitals with superatomic characters corresponding to jellium 1S, 2S, 1P, 3S, and 2P shells for the Rh19 clusters. Iso value = 0.02 unless otherwise stated. HOMO, highest occupied molecular orbital.

In addition to the distinct superatomic feature (1113), the novelty of Rh19 also lies in its remarkable spin-unpaired electrons pertaining to a magnetic superatom of a cluster consisting of fully nonmagnetic Rh atoms (4345). Notably, bulk rhodium is also nonmagnetic. The origin of high-spin state of Rh19 is supposed to be a unity of opposites of superatomic itinerant electrons and localized d electrons. Such a high-spin magic-number cluster (S = 10/2) may have intriguing properties for spintronics and quantum information processing (e.g., photon or magnetic-field manipulated spin decoherency) (46). This cluster with 10 spin unpaired electrons could be a mimic of the predicted element Utn (Z = 130) (47).

Chemical bonding analysis

Figure 4 depicts Kohn-Sham molecular orbital energy–level correlations between [Rh@Rh12] and [Rh6] to unveil the bonding nature of Oh Rh19 (figs. S27 to S31). As the symmetry changes, the 5s-based orbitals with a1g + t1u + t2g + eg + t2u symmetries are split into two series. That is, the occupied bonding orbitals with a1g and t1u symmetries and vacant antibonding orbitals with t2g, a1g, eg, and t2u symmetries. Among them, the bonding 5s-based orbitals (6a1g + 11t1u) are dominated by the core-shell [Rh@Rh12] moiety, of which the stability can be reinforced by strong interactions with capped Rha atoms, pertaining to superatom-atom bonding nature analogous to that in Au20 (10, 48). Compared to the small energy-level splitting of a1g, t1u and eg in the shell [Rh6] unit as a result of weak orbital overlaps between the Rha atoms, the bonding and antibonding orbitals in the [Rh@Rh12@Rh6] cluster display a wide energy region with an obvious energy gap indicative of strong Rhe-Rhc and Rhe-Rhe interactions. All the 5s-based bonding orbitals are fully occupied within an eight-electron configuration of (a1g)2(t1u)6, whereas all the antibonding orbitals are vacant for Rh@Rh12, which obeys the 6n + 2 rule of cubic aromaticity (49, 50). In this regard, we have calculated the nucleus-independent chemical shift (NICS) values and found that the NICS(1)_zz is up to −38.7 parts per million (table S11), showing decent aromaticity of this cluster.

Fig. 4. Bonding analysis.

Fig. 4.

Kohn-Sham MO energy-level correlation diagram of the Oh [Rh@Rh12@Rh6] cluster. The insets show the contour plots (Iso value = 0.03) of the 5s-based orbitals (6t1u and 4a1g) in the kernel Oh-[Rh@Rh12] unit (left) in comparison with the related orbitals (11t1u and 6a1g) in the [Rh@Rh12@Rh6] cluster (right). An illustration in the upper right corner shows the superatom-atom interaction pattern of Rh13@Rh6. SOMO, singly occupied molecular orbital; ADE, adiabatic electron detachment energy.

Apart from the 5s manifold of orbitals, the 4d-based orbitals of [Rh@Rh12@Rh6] can be divided into three series (fig. S28): 77 fully occupied orbitals, 8 vacant orbitals, and 10 singly occupied molecular orbitals (SOMOs) that mainly derive from the [Rh@Rh12] moiety. The ground state has a stable half-filled open-shell configuration of (10t2g)3(15t1u)3(4eu)2(11eg)2 with S = 10/2. The energy levels of unoccupied orbitals provided by the outer atoms all increase during the linear superposition, accompanied by degressive occupied orbital energy levels, thereby reducing the total energy of the system. Compared to the vacant antibonding orbitals in [Rh@Rh12@Rh6], the SOMO also corresponds to an antibonding orbital but with an enlarged energy gap (0.92 eV between the α-SOMO and LUMO at the B3LYP level) compared with the kernel Oh-[Rh@Rh12] moiety. Notably, there is strong spin polarization in this system, leading to, for instance, the splitting of α- and β-MO of 7t1g as large as 0.86 eV.

DISCUSSION

It is worth noting that the PES of Rh19 cluster exhibits electronic hot bands before the first major electronic transition labeled as X (VDE1). As revealed by the Perdew–Burke–Ernzerhof (PBE) functional calculations, while the Rh19 cluster has a high-spin ground state S = 10/2, the other spin states due to different electron spin coupling (S = 0, 1, …, 10/2, 12/2…,16/2) lie in a small energy range, indicating the existence of numerous nearly degenerate ESSIs of the Rh19 cluster. Therefore, the broad PES hot bands can be understood via the electronic transitions from thermally populated ESSI levels of the Rh19 isomers to the corresponding neutral Rh19 following the PES selection rule. To illustrate the electronic hot band mechanism of the PES for such clusters, Fig. 5 sketches the electronic transitions from the Oh Rh19 ground state to the electron-detached Rh19 neutral. Because of the near degeneracy of the spin states of the anion the neutral, the ESSIs of the anion are populated approximately on the basis of Boltzmann distribution (with an ESSI population ratio nj/ni = e−(ɛj−ɛi)/kT). At room temperature (kT ~200 cm−1), there is likely a comparable time scale for vibrational relaxation, spin conversion, and vertical detachment to form the ESSIs of electron-detached neutral Rh19. The diversity of temperature-dependent population of the electronic and vibrational states of the high-spin clusters perplexes their PES measurements, reflected in the ESSIs of both cluster anion and the electron-detached neutral, as well as accelerative vibrations for a certain long-lived excited state. Notably, the super-octahedron Rh19 cluster has a spatial Oh symmetry but with different spin couplings at finite temperature. As the total wave function Ψ(R,S) = Ψ(R)Ψ(S) involves both electron-spin (S) isomers and likely low-lying spatial-structure (R) isomers, the PES feature of this cluster is rather complex, and concomitant electronic detachments from the populated ESSIs of the anion are inevitable for such high-spin clusters. It is therefore important to take ESSIs into account in the PES identification. In addition, the interactions between the spatial and spin coordinates are vital to understanding spin-orbit effects and scalar relativistic (SR) effects of TM clusters (51).

Fig. 5. PES mechanism involving ESSIs.

Fig. 5.

A sketch showing the electronic transition from the ESSIs of Oh Rh19 anion to the electron-detached Rh19 neutral, both electronic and vibrational states involved. The curves labeled as S0, S0′, and S1′ correspond to the electronic ground states and the first excitation state; the dot curves refer to ESSIs; and the horizon lines with sine wave mean vibrational states of the Rh19 anion and the Rh19 neutral. All the drop arrow lines depict the electronic transition between two states.

In summary, a magic-number TM cluster Rh19 is found with a super-octahedron structure that features high stability and multiple nearly-degenerate ESSIs. Such a highly magnetic Oh Rh19 provides superatomic building blocks of 1 nm sizes for developing previously unidentified magnetic micromaterials; the [Rh@Rh12@Rh6] structure bears six unsaturated apex Rh atoms indicating ideal active sites for catalysis. This high-spin Rh19 cluster shows a perfect example of the coexistence of multiple ESSIs while keeping the same spatial structure. This kind of stable metal-atomic clusters may find interest in atomically precise manufacturing involving quantum computing and spintronics.

MATERIALS AND METHODS

Experimental

The experiments were carried out by using a customized reflection time-of-flight mass spectrometer combined with the newly developed pulsed LaVa source, a compact flow tube reactor, and a 45° ion-extraction photoelectron velocity map imaging apparatus (29, 52, 53). A translatable and rotating rhodium disk (99.95%, ɸ = 16 mm) was used as the target of laser ablation to generate the Rhn clusters, with He (99.999%, 1 MPa) as a buffer gas. A 532-nm laser with a pulse energy of 15 to 35 mJ was used for the laser ablation. Downstream the cluster beam, different amounts of O2, CO2, CH4, and CH3Br (~5% in He, 0.1 MPa) were injected by a pulsed valve to react with rhodium clusters in the flow tube reactor.

Computational

Unbiased GM structure search to determine the ground-state structure of the Rh19 cluster was conducted by using both TGMin code (33) and ab initio evolutionary algorithm USPEX (31) embedded in the VASP software (32). The generalized gradient approximation with the PBE exchange-correlation functional was used for the energetics calculations, while orbital analysis and orbital energy–level correlation diagram are based on hybrid functional B3LYP in Gaussian and ADF software packages, respectively. Relativistic quantum chemical studies were performed, with SR effects considered by zero-order regular approximation to account for the mass velocity and Darwin effects.

Acknowledgments

We thank the supercomputers at Tsinghua National Laboratory for Information Science and Technology and the Center for Computational Science and Engineering (SUSTech).

Funding: This work was supported by the National Natural Science Foundation of China (grant nos. 92261113, 22033005, and 22038002), the National Key Research and Development Project (grant nos. 2020YFA0714602, 2022YFA1503900, and 2022YFA1503000), the Key Research Program of Frontier Sciences (CAS, grant no. QYZDBSSW-SLH024), and the Guangdong Provincial Key Laboratory of Catalysis (no. 2020B121201002).

Author contributions: Y.J., L.G., and H.Z. conducted the experiments. C.-Q.X. and C.C. conducted the calculations. Z.L. and J.L. contributed to the design of this project. All authors contributed to analyzing the data and writing the manuscript.

Competing interests: The authors declare that they have no competing interests.

Data and materials availability: All data needed to evaluate the conclusions in this paper are present in the paper and/or the Supplementary Materials.

Supplementary Materials

This PDF file includes:

Supplementary Text

Figs. S1 to S31

Tables S1 to S11

References

REFERENCES AND NOTES

  • 1.Tyo E. C., Vajda S., Catalysis by clusters with precise numbers of atoms. Nat. Nanotechnol. 10, 577–588 (2015). [DOI] [PubMed] [Google Scholar]
  • 2.Du Y., Sheng H., Astruc D., Zhu M., Atomically precise noble metal nanoclusters as efficient catalysts: A bridge between structure and properties. Chem. Rev. 120, 526–622 (2020). [DOI] [PubMed] [Google Scholar]
  • 3.Yoon B., Hakkinen H., Landman U., Worz A. S., Antonietti J. M., Abbet S., Judai K., Heiz U., Charging effects on bonding and catalyzed oxidation of CO on Au8 clusters on MgO. Science 307, 403–407 (2005). [DOI] [PubMed] [Google Scholar]
  • 4.Knight W. D., Clemenger K., de Heer W. A., Saunders W. A., Chou M. Y., Cohen M. L., Electronic shell structure and abundances of sodium clusters. Phys. Rev. Lett. 52, 2141–2143 (1984). [Google Scholar]
  • 5.Brack M., The physics of simple metal clusters: Self-consistent jellium model and semiclassical approaches. Rev. Mod. Phys. 65, 677–732 (1993). [Google Scholar]
  • 6.De Heer W. A., The physics of simple metal clusters: Experimental aspects and simple models. Rev. Mod. Phys. 65, 611–676 (1993). [Google Scholar]
  • 7.Kuo K. H., Mackay, anti-mackay, double-mackay, pseudo-mackay, and related icosahedral shell clusters. Struct. Chem. 13, 221–230 (2002). [Google Scholar]
  • 8.Marks L. D., Surface-structure and energetics of multiply twinned particles. Philos. Mag. A 49, 81–93 (1984). [Google Scholar]
  • 9.Leary R. H., Doye J. P. K., Tetrahedral global minimum for the 98-atom lennard-jones cluster. Phys. Rev. E 60, R6320–R6322 (1999). [DOI] [PubMed] [Google Scholar]
  • 10.Li J., Li X., Zhai H. J., Wang L. S., Au20: A tetrahedral cluster. Science 299, 864–867 (2003). [DOI] [PubMed] [Google Scholar]
  • 11.Jena P., Sun Q., Super atomic clusters: Design rules and potential for building blocks of materials. Chem. Rev. 118, 5755–5870 (2018). [DOI] [PubMed] [Google Scholar]
  • 12.Reber A. C., Khanna S. N., Superatoms: Electronic and geometric effects on reactivity. Acc. Chem. Res. 50, 255–263 (2017). [DOI] [PubMed] [Google Scholar]
  • 13.Luo Z., Castleman A. W., Special and general superatoms. Acc. Chem. Res. 47, 2931–2940 (2014). [DOI] [PubMed] [Google Scholar]
  • 14.Häkkinen H., Atomic and electronic structure of gold clusters: Understanding flakes, cages and superatoms from simple concepts. Chem. Soc. Rev. 37, 1847–1859 (2008). [DOI] [PubMed] [Google Scholar]
  • 15.Hirai H., Ito S., Takano S., Koyasu K., Tsukuda T., Ligand-protected gold/silver superatoms: Current status and emerging trends. Chem. Sci. 11, 12233–12248 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Zhao J., Du Q., Zhou S., Kumar V., Endohedrally doped cage clusters. Chem. Rev. 120, 9021–9163 (2020). [DOI] [PubMed] [Google Scholar]
  • 17.Bergeron D. E., Castleman A. W. Jr., Morisato T., Khanna S. N., Formation of Al13I: Evidence for the superhalogen character of Al13. Science 304, 84–87 (2004). [DOI] [PubMed] [Google Scholar]
  • 18.Bergeron D. E., Roach P. J., Castleman A. W. Jr., Jones N. O., Khanna S. N., Al cluster superatoms as halogens in polyhalides and as alkaline earths in iodide salts. Science 307, 231–235 (2005). [DOI] [PubMed] [Google Scholar]
  • 19.Zhang X., Wang Y., Wang H., Lim A., Gantefoer G., Bowen K. H., Reveles J. U., Khanna S. N., On the existence of designer magnetic superatoms. J. Am. Chem. Soc. 135, 4856–4861 (2013). [DOI] [PubMed] [Google Scholar]
  • 20.Reimann S. M., Koskinen M., Häkkinen H., Lindelof P. E., Manninen M., Magic triangular and tetrahedral clusters. Phys. Rev. B 56, 12147–12150 (1997). [Google Scholar]
  • 21.Medel V. M., Reveles J. U., Khanna S. N., Chauhan V., Sen P., Castleman A. W., Hund's rule in superatoms with transition metal impurities. Proc. Natl. Acad. Sci. U.S.A. 108, 10062–10066 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Zhao Y.-X., Zhao X.-G., Yang Y., Ruan M., He S.-G., Rhodium chemistry: A gas phase cluster study. J. Chem. Phys. 154, 180901 (2021). [DOI] [PubMed] [Google Scholar]
  • 23.Luo Z., Grover C. J., Reber A. C., Khanna S. N., Castleman A. W. Jr., Probing the magic numbers of aluminum–magnesium cluster anions and their reactivity toward oxygen. J. Am. Chem. Soc. 135, 4307–4313 (2013). [DOI] [PubMed] [Google Scholar]
  • 24.Yin B., Du Q., Geng L., Zhang H., Luo Z., Zhou S., Zhao J., Superatomic signature and reactivity of silver clusters with oxygen: Double magic Ag17 with geometric and electronic shell closure. CCS Chem. 3, 219–229 (2021). [Google Scholar]
  • 25.Ren Y., Yang Y., Zhao Y.-X., He S.-G., Size-dependent reactivity of rhodium cluster anions toward methane. J. Phys. Chem. C 123, 17035–17042 (2019). [Google Scholar]
  • 26.Adlhart C., Uggerud E., C–h activation of alkanes on Rhn+ (n=1–30) clusters: Size effects on dehydrogenation. J. Chem. Phys. 123, 214709 (2005). [DOI] [PubMed] [Google Scholar]
  • 27.Harding D., Ford M. S., Walsh T. R., Mackenzie S. R., Dramatic size effects and evidence of structural isomers in the reactions of rhodium clusters, Rhn+/−, with nitrous oxide. Phys. Chem. Chem. Phys. 9, 2130–2136 (2007). [DOI] [PubMed] [Google Scholar]
  • 28.Zhang H., Wu H., Jia Y., Geng L., Luo Z., Fu H., Yao J., An integrated instrument of DUV-IR photoionization mass spectrometry and spectroscopy for neutral clusters. Rev. Sci. Instrum. 90, 073101 (2019). [DOI] [PubMed] [Google Scholar]
  • 29.Zhang H., Geng L., Jia Y., Lei X., Luo Z., A 45-degree ion-extraction photoelectron velocity map imaging apparatus coupled to the DUV-IR mass spectrometry and spectroscopy instrument. Int. J. Mass Spectrom. 482, 116932 (2022). [Google Scholar]
  • 30.Kroto H. W., Heath J. R., Obrien S. C., Curl R. F., Smalley R. E., C60 – buckminsterfullerene. Nature 318, 162–163 (1985). [Google Scholar]
  • 31.Glass C. W., Oganov A. R., Hansen N., Uspex - evolutionary crystal structure prediction. Comput. Phys. Commun. 175, 713–720 (2006). [Google Scholar]
  • 32.Kresse G., Furthmüller J., Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comp. Mater. Sci. 6, 15–50 (1996). [DOI] [PubMed] [Google Scholar]
  • 33.Zhao Y., Chen X., Li J., Tgmin: A global-minimum structure search program based on a constrained basin-hopping algorithm. Nano Res. 10, 3407–3420 (2017). [Google Scholar]
  • 34.Cox A. J., Louderback J. G., Apsel S. E., Bloomfield L. A., Magnetism in 4d-transition metal clusters. Phys. Rev. B 49, 12295–12298 (1994). [DOI] [PubMed] [Google Scholar]
  • 35.Lloyd L. D., Johnston R. L., Theoretical analysis of 17–19-atom metal clusters using many-body potentials †. J. Chem. Soc. Dalton Trans., 307–316 (2000). [Google Scholar]
  • 36.Li S. F., Zhao X. J., Xu X. S., Gao Y. F., Zhang Z., Stacking principle and magic sizes of transition metal nanoclusters based on generalized wulff construction. Phys. Rev. Lett. 111, 115501 (2013). [DOI] [PubMed] [Google Scholar]
  • 37.Waldt E., Hehn A. S., Ahlrichs R., Kappes M. M., Schooss D., Structural evolution of small ruthenium cluster anions. J. Chem. Phys. 142, 024319 (2015). [DOI] [PubMed] [Google Scholar]
  • 38.Bumüller D., Hehn A.-S., Waldt E., Ahlrichs R., Kappes M. M., Schooss D., Ruthenium cluster structure change induced by hydrogen adsorption: Ru19. J. Phys. Chem. C 121, 10645–10652 (2016). [Google Scholar]
  • 39.Luo Z., Reber A. C., Jia M., Blades W. H., Khanna S. N., Castleman A. W. Jr., What determines if a ligand activates or passivates a superatom cluster? Chem. Sci. 7, 3067–3074 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Geng L., Weng M., Xu C.-Q., Zhang H., Cui C., Wu H., Chen X., Hu M., Lin H., Sun Z.-D., Wang X., Hu H.-S., Li J., Zheng J., Luo Z., Pan F., Yao J., Co13O8—metalloxocubes: A new class of perovskite-like neutral clusters with cubic aromaticity. Natl. Sci. Rev. 8, nwaa201 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Jia Y., Luo Z., Thirteen-atom metal clusters for genetic materials. Coord. Chem. Rev. 400, 213053 (2019). [Google Scholar]
  • 42.Zubarev D. Y., Boldyrev A. I., Developing paradigms of chemical bonding: Adaptive natural density partitioning. Phys. Chem. Chem. Phys. 10, 5207–5217 (2008). [DOI] [PubMed] [Google Scholar]
  • 43.Cox A. J., Louderback J. G., Bloomfield L. A., Experimental observation of magnetism in rhodium clusters. Phys. Rev. Lett. 71, 923–926 (1993). [DOI] [PubMed] [Google Scholar]
  • 44.Bae Y.-C., Osanai H., Kumar V., Kawazoe Y., Nonicosahedral growth and magnetic behavior of rhodium clusters. Phys. Rev. B 70, 195413 (2004). [Google Scholar]
  • 45.Ma L., Moro R., Bowlan J., de Heer W. A., Kirilyuk A., Multiferroic rhodium clusters. Phys. Rev. Lett. 113, 157203 (2014). [DOI] [PubMed] [Google Scholar]
  • 46.Graham M. J., Zadrozny J. M., Shiddiq M., Anderson J. S., Fataftah M. S., Hill S., Freedman D. E., Influence of electronic spin and spin-orbit coupling on decoherence in mononuclear transition metal complexes. J. Am. Chem. Soc. 136, 7623–7626 (2014). [DOI] [PubMed] [Google Scholar]
  • 47.Fricke B., Soff G., Dirac-fock-slater calculations for the elements Z = 100, fermium, to Z = 173. At. Data Nucl. Data Table 19, 83–95 (1977). [Google Scholar]
  • 48.Cheng L., Zhang X., Jin B., Yang J., Superatom-atom super-bonding in metallic clusters: A new look to the mystery of an Au20 pyramid. Nanoscale 6, 12440–12444 (2014). [DOI] [PubMed] [Google Scholar]
  • 49.Cui P., Hu H. S., Zhao B., Miller J. T., Cheng P., Li J., A multicentre-bonded [ZnI]8 cluster with cubic aromaticity. Nat. Commun. 6, 6331 (2015). [DOI] [PubMed] [Google Scholar]
  • 50.Hu H. C., Hu H. S., Zhao B., Cui P., Cheng P., Li J., Metal-organic frameworks (MOFs) of a cubic metal cluster with multicentered Mn(I)-Mn(I) bonds. Angew. Chem. Int. Ed. Engl. 54, 11681–11685 (2015). [DOI] [PubMed] [Google Scholar]
  • 51.Muñoz-Castro A., Arratia-Perez R., Spin–orbit effects on a gold-based superatom: A relativistic jellium model. Phys. Chem. Chem. Phys. 14, 1408–1411 (2012). [DOI] [PubMed] [Google Scholar]
  • 52.Hock C., Kim J. B., Weichman M. L., Yacovitch T. I., Neumark D. M., Slow photoelectron velocity-map imaging spectroscopy of cold negative ions. J. Chem. Phys. 137, 244201 (2012). [DOI] [PubMed] [Google Scholar]
  • 53.Leon I., Yang Z., Liu H. T., Wang L. S., The design and construction of a high-resolution velocity-map imaging apparatus for photoelectron spectroscopy studies of size-selected clusters. Rev. Sci. Instrum. 85, 083106 (2014). [DOI] [PubMed] [Google Scholar]
  • 54.Swart M., Van Duijnen P. T., Snijders J. G., A charge analysis derived from an atomic multipole expansion. J. Comput. Chem. 22, 79–88 (2001). [Google Scholar]
  • 55.Huang W., Sergeeva A. P., Zhai H. J., Averkiev B. B., Wang L. S., Boldyrev A. I., A concentric planar doubly π-aromatic B19− cluster. Nat. Chem. 2, 202–206 (2010). [DOI] [PubMed] [Google Scholar]
  • 56.Katsoulis D. E., A survey of applications of polyoxometalates. Chem. Rev. 98, 359–388 (1998). [DOI] [PubMed] [Google Scholar]
  • 57.Tolbert M. A., Mandich M. L., Halle L. F., Beauchamp J. L., Activation of alkanes by ruthenium, rhodium, and palladium ions in the gas phase: Striking differences in reactivity of first- and second-row metal ions. J. Am. Chem. Soc. 108, 5675–5683 (1986). [DOI] [PubMed] [Google Scholar]
  • 58.Wan X. K., Tang Q., Yuan S. F., Jiang D. E., Wang Q. M., Au19 nanocluster featuring a v-shaped alkynyl-gold motif. J. Am. Chem. Soc. 137, 652–655 (2015). [DOI] [PubMed] [Google Scholar]
  • 59.Velde G. T., Bickelhaupt F. M., Baerends E. J., Guerra C. F., Vangisbergen S. J. A., Snijders J. G., Ziegler T., Chemistry with ADF. J. Comput. Chem. 22, 931–967 (2001). [Google Scholar]
  • 60.Yu X., Oganov A. R., Popov I. A., Boldyrev A. I., d-AO spherical aromaticity in Ce6O8. J. Comput. Chem. 37, 103–109 (2016). [DOI] [PubMed] [Google Scholar]
  • 61.Wang J., Hao D., Ye J., Umezawa N., Determination of crystal structure of graphitic carbon nitride: Ab initio evolutionary search and experimental validation. Chem. Mater. 29, 2694–2707 (2017). [Google Scholar]
  • 62.Yu S., Huang B., Zeng Q., Oganov A. R., Zhang L., Frapper G., Emergence of novel polynitrogen molecule-like species, covalent chains, and layers in magnesium–nitrogen MgxNy phases under high pressure. J. Phys. Chem. C 121, 11037–11046 (2017). [Google Scholar]
  • 63.Kresse G., Joubert D., From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758–1775 (1999). [Google Scholar]
  • 64.Perdew J. P., Burke K., Ernzerhof M., Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). [DOI] [PubMed] [Google Scholar]
  • 65.Xing X., Hermann A., Kuang X., Ju M., Lu C., Jin Y., Xia X., Maroulis G., Insights into the geometries, electronic and magnetic properties of neutral and charged palladium clusters. Sci. Rep. 6, 19656 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Hill C. L., Introduction: Polyoxometalates-multicomponent molecular vehicles to probe fundamental issues and practical problems. Chem. Rev. 98, 1–2 (1998). [DOI] [PubMed] [Google Scholar]
  • 67.Duncan M. A., Invited review article: Laser vaporization cluster sources. Rev. Sci. Instrum. 83, 041101 (2012). [DOI] [PubMed] [Google Scholar]
  • 68.Echt O., Sattler K., Recknagel E., Magic numbers for sphere packings: Experimental verification in free xenon clusters. Phys. Rev. Lett. 47, 1121–1124 (1981). [Google Scholar]
  • 69.Miehle W., Kandler O., Leisner T., Echt O., Mass spectrometric evidence for icosahedral structure in large rare gas clusters: Ar, Kr, Xe. J. Chem. Phys. 91, 5940–5952 (1989). [Google Scholar]
  • 70.Lu T., Chen F., Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 33, 580–592 (2012). [DOI] [PubMed] [Google Scholar]
  • 71.Chen Z. F., Wannere C. S., Corminboeuf C., Puchta R., Schleyer P. v. R., Nucleus-independent chemical shifts (NICS) as an aromaticity criterion. Chem. Rev. 105, 3842–3888 (2005). [DOI] [PubMed] [Google Scholar]
  • 72.Schleyer P. v. R., Maerker C., Dransfeld A., Jiao H., Hommes N. J. R. v. E., Nucleus-independent chemical shifts: A simple and efficient aromaticity probe. J. Am. Chem. Soc. 118, 6317–6318 (1996). [DOI] [PubMed] [Google Scholar]
  • 73.Van Lenthe E., Baerends E. J., Optimized slater-type basis sets for the elements 1-118. J. Comput. Chem. 24, 1142–1156 (2003). [DOI] [PubMed] [Google Scholar]
  • 74.Mafune F., Tawaraya Y., Kudoh S., Reactivity control of rhodium cluster ions by alloying with tantalum atoms. J. Phys. Chem. A 120, 861–867 (2016). [DOI] [PubMed] [Google Scholar]
  • 75.van Lenthe E., Baerends E. J., Snijders J. G., Relativistic regular 2-component hamiltonians. J. Chem. Phys. 99, 4597–4610 (1993). [Google Scholar]
  • 76.Balteanu I., Balaj O. P., Fox-Beyer B. S., Rodrigues P., Barros M. T., Moutinho A. M. C., Costa M. L., Beyer M. K., Bondybey V. E., Size- and charge-state-dependent reactivity of azidoacetonitrile with anionic and cationic rhodium clusters Rhn. Organometallics 23, 1978–1985 (2004). [Google Scholar]
  • 77.Harris I. A., Kidwell R. S., Northby J. A., Structure of charged argon clusters formed in a free jet expansion. Phys. Rev. Lett. 53, 2390–2393 (1984). [Google Scholar]
  • 78.Waldt E., Ahlrichs R., Kappes M. M., Schooss D., Structures of medium-sized ruthenium clusters: The octahedral motif. ChemPhysChem 15, 862–865 (2014). [DOI] [PubMed] [Google Scholar]
  • 79.Wei G.-F., Liu Z.-P., Subnano Pt particles from a first-principles stochastic surface walking global search. J. Chem. Theory Comput. 12, 4698–4706 (2016). [DOI] [PubMed] [Google Scholar]
  • 80.Humphrey W., Dalke A., Schulten K., VMD: Visual molecular dynamics. J. Molec. Graph. 14, 33–38 (1996). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Text

Figs. S1 to S31

Tables S1 to S11

References


Articles from Science Advances are provided here courtesy of American Association for the Advancement of Science

RESOURCES