Skip to main content
National Science Review logoLink to National Science Review
. 2023 Dec 22;11(4):nwad327. doi: 10.1093/nsr/nwad327

Synthesis and characterization of iron clusters with an icosahedral [Fe@Fe12]16+ Core

Gan Xu 1,c, Yun-Shu Cui 2,c, Xue-Lian Jiang 3, Cong-Qiao Xu 4, Jun Li 5,6,, Xu-Dong Chen 7,8,
PMCID: PMC10939364  PMID: 38487495

ABSTRACT

Iron-metal clusters are crucial in a variety of critical biological and material systems, including metalloenzymes, catalysts, and magnetic storage devices. However, a synthetic high-nuclear iron cluster has been absent due to the extreme difficulty in stabilizing species with direct iron−iron bonding. In this work, we have synthesized, crystallized, and characterized a (Tp*)4W4S12(Fe@Fe12) cluster (Tp* = tris(3,5-dimethyl-1-pyrazolyl)borate(1−)), which features a rare trideca-nuclear, icosahedral [Fe@Fe12] cluster core with direct multicenter iron−iron bonding between the interstitial iron (Fei) and peripheral irons (Fep), as well as Fep···Fep ferromagnetic coupling. Quantum chemistry studies reveal that the stability of the cluster arises from the 18-electron shell-closing of the [Fe@Fe12]16+ core, assisted by its bonding interactions with the peripheral tridentate [(Tp*)WS3]4− ligands which possess both S→Fe donation and spin-polarized Fe−W σ bonds. The ground-state electron spin is theoretically predicted to be S = 32/2 for the cluster. The existence of low oxidation-state (OS ∼ +1.23) iron in this compound may find interesting applications in magnetic storage, spintronics, redox chemistry, and cluster catalysis.

Keywords: Fe13 cluster, iron-sulfur cluster, low-valence iron, iron−iron bond, high spin complex


A low-valent iron cluster with icosahedral [Fe@Fe12]16+ core and multicenter iron-iron bonding has been synthetically achieved and computationally studied.

INTRODUCTION

Clusters with precise composition and geometric structure represent the intermediate state between atoms/molecules and condensed-phase matter. The in-depth understanding of cluster assembly, structure, and properties can provide key information for bulk materials. Cluster science is therefore of fundamental interest for the development of innovative materials [1–6]. As one of the most important elements on earth, iron has played critical roles in natural metalloenzymes such as nitrogenase and hydrogenase [7–9], as well as in key industrial processes such as the Haber-Bosch and Fischer-Tropsch processes [10,11]. Iron-metal clusters have been extensively studied due to their promising magnetic properties with such potential applications as magnetic storage [12–16]. They provide intrinsic information on how iron atoms are assembled through iron−iron interactions, which provides a fundamental understanding of how physical properties such as magnetism evolve from iron atoms to bulk materials [12–16]. However, most experimental studies of iron-metal clusters focus on gas-phase species using techniques such as molecular beam deflection or low-temperature matrix isolation [13,14,17–19]. While it is challenging to provide precise atomic structure of iron clusters in such experiments, synthetic iron clusters are essential as the precise cluster composition and crystal structure provide rich information on iron−iron bonding and inter-cluster interactions.

However, the synthesis of iron clusters relies exclusively on ligand bridging. Usually, peripheral ligands help to link the iron atoms and stabilize the whole cluster [20–24], and internal ligands are the key to bridging the iron centers to form high-nuclear clusters [25–27]. Direct iron−iron bonding between the interstitial and peripheral iron atoms not assisted by ligand-bridging is a prerequisite for assembling a high-nuclear pure iron-metal cluster core. To date, unsupported iron−iron bonding has only been observed in very limited cases of dimeric iron complexes [28–30], whereas a high-nuclearity synthetic iron cluster with pure [Fen] (n > 8) metal core is absent. The synthesis of high-nuclear iron-metal clusters has been challenging, hindering the understanding of the intrinsic nature of unsupported iron−iron bonding in iron materials. We report here the synthesis of an unprecedented icosahedral [Fe@Fe12] cluster compound featuring unsupported iron−iron bonding between the interstitial and peripheral iron atoms, in which the Fe oxidation state (OS) is +16/13 on average. This highly symmetric [Fe13] cluster, characterized by experimental techniques and quantum theoretical calculations, is the first crystallized iron cluster showing the precise atomic arrangement of the icosahedral [Fe13] core.

RESULTS AND DISCUSSION

Synthesis

In our experiment, the discrete iron ions were first assembled into [Fe3] moieties using a template, which were further mounted around an interstitial iron atom under reduced conditions to form the [Fe@Fe12]-containing cluster. Three equivalents of iron(II) were readily assembled into the [(Tp*)WFe3S33-Cl)Cl3]2− (Tp* = tris(3,5-dimethyl-1-pyrazolyl)borate(1−)) cluster containing a [Fe3] moiety [31], and then redox-free terminal ligand substitution with triethylphosphine ligands afforded [(Tp*)WFe3S3(μ3-Cl)(PEt3)3]1+ (1) as dark red crystals. Here the [(Tp*)WS3]4− moiety retains the low-valence W(III) and Fe(II) in 1 as in the precursor. Accordingly, the addition of sodium benzophenone ketyl solution (10 equiv.) into the premixed tetrahydrofuran (THF) solution of 1 (4 equiv.) and FeCl2(THF)1.5 (1 equiv.) followed by pentane diffusion, resulted in the formation of diamond-shaped brown-black crystals. This crystalline product was characterized by single-crystal X-ray diffraction to be a high-nuclearity neutral cluster [(Tp*)4W4S12Fe13] (2, 41% average yield based on eight experiments, highest yield at 52%), which features four [(Tp*)WS3]4− ligands around the [Fe13] cluster core. The synthetic process and structure of 2 is shown in Fig. 1. More experimental details are shown in Figs S1–S8 and Tables S1–S5 in the Supplementary Materials (SM).

Figure 1.

Figure 1.

Synthesis of [(Tp*)4W4S12(Fe@Fe12)] (2).

The significant stability of the icosahedral [Fe@Fe12] assembly in 2 is intriguing. It appears that the reduction of the iron centers in the ferrous [WFe3]9+ core of 1 and FeCl2(THF)1.5 to lower oxidation states triggers the direct iron−iron interaction in 2, leading to the aggregation of four [(Tp*WS3)Fe3] moieties around a central iron to form the icosahedral [Fe@Fe12] cluster core. According to the stoichiometric control of the synthetic design, the metal centers in starting materials 1 and FeCl2(THF)1.5 were reduced by 10 electrons from formal [W4Fe13]38+ (four [WFe3]9+ plus one Fe2+) to [W4Fe13]28+ during the reaction process. Indeed, both XPS (Fig. S2) and quantum-chemical theoretical studies (vide infra) of compound 2 suggest a +3 oxidation state of the W centers, thus rendering the [Fe@Fe12]16+ core an extremely low average oxidation state of OS = +16/13 (∼ +1.23) for these 13 Fe atoms.

Structure and bonding

Compound 2 crystallized in the space group RInline graphic, where the [Fe13]16+ core is surrounded by four [(Tp*)WS3]4− species in a tetrahedral arrangement. Previous time-of-flight mass spectrometry studies and first-principles calculations showed that [Fe13] cluster corresponding to a magic number was likely stable, consistent with its more frequent occurrence in gas phase distribution [15,19,32–34]. Indeed, our experimental observations have shown that compound 2 with [Fe13]16+ core is a thermodynamically stable product frequently observed in various reactions. Due to steric restraints generated from the ligands and crystal packing, the molecular symmetry of 2 in the solid state is slightly lowered from ideal Td to its subgroup with a chiral C3 symmetry, which has also slightly affected the Ih-symmetry [Fe13]16+ core to become a distorted icosahedron. In this [Fe13]16+ cluster core, 12 peripheral iron atoms (Fep) assemble around an interstitial iron atom (Fei) to form an Fe-centered icosahedral [Fe@Fe12]16+ core. The determined structure of the [Fe@Fe12]16+ cluster core is shown in Fig. 2.

Figure 2.

Figure 2.

Structure of the [Fe@Fe12]16+ cluster. Thermal ellipsoids are set at the 65% probability level. Symmetry codes: i: -y, x-y, z; ii: y-x, -x, z. The Fep−Fei bond lengths range from 2.483(1) to 2.530(1)  Å, with a mean value of 2.503 Å. The Fep···Fep distances are between 2.377(1) and 2.793(1)  Å, averaged at 2.647 Å.

A typical structural feature of the [Fe@Fe12]16+ core is that the peripheral-interstitial Fep−Fei bonds are unsupported, especially because they are completely ligand-free inside the Fe12 cage. Based on the quantum chemical calculations below, the Fep−Fei bonds in the [Fe@Fe12]16+ core of compound 2 belong to delocalized multi-center Fe−Fe bonds. Meanwhile, the peripheral iron atoms are shown to interact with neighboring ones via ferromagnetic coupling. The Fe−S coordination and W−Fe bonding are found to be the major interactions between the surrounding [(Tp*)WS3]4− species and the [Fe@Fe12]16+ core. Therefore, in compound 2, the unsupported Fep−Fei bonds and the Fep···Fep ferromagnetic coupling account for the key factors that assemble the [Fe@Fe12]16+ aggregate, which was further stabilized by the surrounding [(Tp*)WS3]4− species through Fe−S coordination and W−Fe bonding. In contrast, for the synthetic iron clusters with nuclearity <8, their cluster aggregation relies exclusively on Fe−Fe bonds bridged by external ligands and their stabilization requires extreme conditions such as −80°C in some cases [20–24].

Low-valence iron is potentially significant in crucial catalytic systems such as carbon fixation involving carbon monoxide dehydrogenase and nitrogen fixation utilizing nitrogenase [35,36]. From a synthetic point of view, such low-valence species have usually been stabilized by carefully designed low-coordinate phosphorus, nitrogen, and sulfur ligands [37–40]. The presence of low-valence iron in iron clusters is rather scarce [20,22,23]. The current case of compound 2 represents a unique example of a high nuclearity (>8) iron cluster containing multiple low-valence iron centers, which may correlate with versatile biomimetic and industrial processes.

Mössbauer studies

The zero-field 57Fe Mössbauer spectrum of compound 2 at 77 K is shown in Fig. 3, which has a slightly asymmetric doublet. Structurally, all 12 peripheral iron atoms of the icosahedron adopt the same coordination environment while the interstitial Fe atom interacts only with the 12 peripheral Fe atoms and has a completely different environment. Therefore, a two-component fitting was applied to the Mössbauer spectrum. The major component (blue fit, 92.3%) exhibits an isomer shift of 0.49 mm/s and the corresponding quadrupole splitting is 1.16 mm/s, which is assigned to the peripheral iron atoms. The other component (brown fit, 7.7%) has an isomer shift of 0.04 mm/s and no obvious quadrupole splitting belonging to the interstitial iron is observed.

Figure 3.

Figure 3.

Zero field Mössbauer spectrum of 2 obtained at 77 K. Simulation parameters: δ/|ΔEQ| (mm/s): blue fit (92.3%): 0.49(1)/1.16(1); brown fit (7.7%), 0.04(2)/0(2).

Quantum chemistry studies

Quantum chemistry studies were performed to probe the geometric and electronic structures of the [Fe@Fe12]16+ core and compound 2 (the computational details are given in the SM). A simplified model 2-H (see Fig. S9) with T symmetry was used as the model cluster, which is formed by replacing all methyl groups of 2 by hydrogen atoms to save computational time. The optimized structures of 2 and the simplified scheme 2-H show almost negligible differences in geometry, both of which agree well with the experimental crystal structure, as illustrated in Table S6. A variety of spin-multiplicity states have been evaluated (see Fig. S10) and the ground state is computationally predicted to have a spin quantum number of S = 32/2 for 2. While the particular spin multiplicity of the iron systems might change with different computational methods, the high spin feature hereby is remarkable.

As mentioned above, the [Fe@Fe12]16+ core of 2 has a pseudo-icosahedral structure, and the discussion will be based on the symmetrized Ih model for brevity. We first studied the interaction between Fei and 12 Fep by the fragment molecular orbital (FMO) approach. As shown in Fig. 4a, the twelve 4s orbitals of Fep span ag(S), t1u(P) bonding group orbitals (GOs) and hg(D), t2u(F) antibonding GOs. The interaction between GOs of Fep and atomic orbitals (AOs) of Fei can be classified into superatomic S-type, P-type and D-type. Such interaction gives rise to a 1S21P61D10 configuration of spherical jellium orbitals for [Fe@Fe12]16+ by involving the 4s-based bonding GOs (ag, t1u) of Fep, which results in occupied 1ag, 1t1u and 1hg orbitals that fulfills the ‘18-electron rule’. This 18-electron configuration provides considerable stability, as in the case of endohedral icosahedral M@Au12 (M = Mo, W) clusters [41,42]. The spin population of Fei is roughly 3, suggesting a 3d7 electronic configuration with an oxidation state of +1. This Fe(I) assignment is further supported by the calculated spin-polarized electron configuration of 3d7.81(4.95α+2.86β) (Table S7) for Fei by natural population analysis (NPA) of [Fe@Fe12]16+, where the lower occupation number of β−spin 3d orbitals (hg symmetry) than the α−spin can be attributed to their significant interaction with the 1hg orbitals derived from Fep−4s AOs.

Figure 4.

Figure 4.

Electronic structure. (a) Schematic Kohn-Sham MO diagram of Ih−[Fe@Fe12]16+ and contours of spherical jellium superatom orbitals. (b) Density of states (DOS) for 2-H. (c) Contours of α and β Pipek-Mezey LMOs of the Fe−W σ bond at the PBE level [50].

The bonding scheme (Fig. 4) can well explain why the low-valence iron cluster can be stable through Fe−Fe direct bonding, as in the case of MI8 (M = Mn, Zn) clusters with cubic aromaticity [43,44]. The [(Tp*)4W4S12(Fe@Fe12)] cluster can be considered as being composed of four [(Tp*)WIIIS3(FeII3)]2+ precursor fragments and a central Fe2+ ion to form the [(Tp*)4WIII4S12(FeII@FeII12)]10+ cluster with an icosahedral (Fe@Fe12)26+ cluster core. Upon formation of this tridecanuclear iron cluster, the Fe 4s-orbitals will form bonding orbitals of ag(S)+t1u(P) GOs, while the five 3d-orbitals of the central Fei will be stabilized via orbital interaction with Fe 4s-based hg GOs. The stabilization of these orbitals will cause the cluster to be reduced by 10 electrons via the formation of direct Fe−Fe bonds, which lead to the (Fe@Fe12)16+ cluster core with low OS = 16/13 (≈1.23).

On the other hand, besides the 18 electrons of the 1S21P61D10 superatomic configuration and the +16 charge for [Fe@Fe12]16+, 70 (Ne = 8 × 13 − 18 − 16) electrons are left for the 3d ‘band’ composed of 60 Fep 3d-based orbitals. Notably, the 3d orbitals are rather radially contracted (Fig. S12) due to quantum primogenic effects and can hardly form covalent bonds with each other at a distance of ∼2.5 Å because of insufficient orbital overlapping. Instead, the Fep−Fep interaction involves ferromagnetic coupling, leading to an energetically low-lying α-3d band stabilized by exchange interactions. The occupation pattern with 60 α electrons and 10 β electrons located at 60 3d-based Fep orbitals is favored for the pure [Fe@Fe12]16+ core, where the strong antibonding 3t1u(2P) orbitals possess the highest energy, as depicted in Fig. 4a. However, upon coordination by [(Tp*)WS3]4− ligands, the 2P superatom orbitals of [Fe@Fe12]16+ in spin α will be strongly destabilized by the P-type ligand group orbitals (LGOs) to yield three vacant antibonding orbitals with high energies, thus promoting spin flip of three α electrons on the 2P orbitals of [Fe@Fe12]16+ to β set, resulting in the cluster electronic state of [Fe@Fe12]16+ to be S = 44/2 (Fig. S13).

Consequently, the interaction between the [Fe@Fe12]16+ core and the ligands [(Tp*)WS3]4− was investigated by analyzing the density of states (DOS), as displayed in Fig. 4b. By comparing the DOS before and after ligation (Figs S14, S15), it can be found that the 3d band of Fe was broadened due to the interaction with the ligands and the increased distributions of β 3d states at the Fermi level, indicating the expected electron transfer from ligands. Additionally, principal interacting spin orbital (PISO) analysis [45] was carried out to extract the one-to-one interacting scenario between the [Fe@Fe12]16+ core and ligands. As illustrated in Table S8 and Figs S16 and S17, we identified two Fe−ligand interactions: the spin-polarized Fe−W interactions and the Fe−S interactions predominated by S-, P-, D-, F-type S→Fe donation. Furthermore, the bond distance of Fe−W (2.70 Å) is slightly longer than the sum of Pyykkö’s covalent radii for Fe and W atoms (2.53 Å) [46] and the calculated bond order is more than 0.50 (Table S1), indicating the existence of non-negligible Fe−W bonding. The DOS of W mainly distributed at ∼0.8 eV above the Fermi level for α states and ∼2.0 eV below the Fermi level for β states. Besides, the Fe−W σ bonds for spin α are Fe-dominated while remarkable increases of W contributions are found for spin β in the localized molecular orbitals (LMOs) (see Fig. 4c and Fig. S8), further demonstrating the spin-polarized character of the Fe−W bonds, which can also be captured in the spin density population in Fig. S19. Therefore, besides the intrinsic stability of the [Fe@Fe12]16+ core, the enhanced stability of 2 can be attributed to interactions from the ligands. First, electron donation from sulfur-based ligands stabilizes the Lewis-acidic sites of the Fe13 cluster core. Meanwhile, the paramagnetic W(III) centers are ‘magnetized’ by the high-spin Fe13 core, providing magnetic attachment between Fe and W, as shown in the spin-polarized Fe−W σ bonds. Both types of attractions enable the ligand-induced stabilization of the Fe13 core in 2.

To gain a quantitative and chemically intuitive understanding of the spin polarization of the Fe−W bonds, we also performed a spin natural orbital (SNO) analysis [47], where a larger absolute SNO eigenvalue |λ| corresponds to the bigger density difference between α and β densities and therefore stronger spin polarization. As shown in Fig. S20, 12 dominant SNO pairs with |λ| of 0.260–0.432 are assigned to 12 Fe−W bonds. Meanwhile, the exchange interaction between the α electron of the Fe−W bond and excess α electrons accumulated at the [Fe@Fe12]16+ core will stabilize the structure, which facilitates α spin density around Fe while inducing β spin density aggregation at the W atom. Thus, one can conclude that each W provides three β electrons and appears as W(III), while three adjacent Fe atoms offer three α electrons to form the spin-polarized Fe−W σ bonds, thus quenching 12 single electrons and leading to a ground state with S = 32/2 for 2-H.

CONCLUSIONS

In summary, an iron cluster with tridecanuclear [Fe@Fe12]16+ core featuring unsupported Fep−Fei bonding has been synthesized. The [Fe@Fe12]16+ cluster core with precise atomic structure is the first synthetic icosahedral [Fe13] cluster. This unique [Fe@Fe12]16+ cluster with low-valence iron centers is assembled by the multi-center Fep−Fei bonds and Fep···Fep ferromagnetic coupling, and is further stabilized by the peripheral [(Tp*)WS3]4− species through Fe−S coordination and W−Fe bonding. Theoretical studies indicate that the 18-electron structure of the icosahedral superatomic [Fe@Fe12]16+ core is responsible for the special intrinsic stability of the structure. The Fe−S interaction reveals S→Fe donation that polarized toward sulfur and the Fe−W interaction can be localized into 12 electron-sharing σ bonds with strong spin polarization. We also demonstrate that three β electrons of the W atom are involved in one Fe−W bond, leading to the oxidation state W(III) and the number of unpaired electrons for compound 2 to be 32. The special stability of this [Fe@Fe12] cluster bears resemblance in superatom characteristics to the recently found Rh@Rh12@Rh6 [48] and Co@Co12O8 clusters with cubic aromaticity [4]. The synthesis and crystallization of this Fe13 cluster underscores a feasible synthetic methodology toward reactive late transition-metal clusters with higher nuclearity.

NOTE ADDED IN THE PROOF

We noticed a latest paper on the same structure published by Scott et al. in Angew. Chem. Int. Ed. [49] after the manuscript had been submitted. Our synthesis approach differs significantly from theirs in both the starting materials and the strategy of cluster assembly, indicating that an Fe13 cluster can be synthesized through different approaches.

METHODS

Details of methods are included in the Supplementary data.

DATA AVAILABILITY

Crystallographic data can be obtained free of charge from the Cambridge Crystallographic Data Centre (www.ccdc.cam.ac.uk/data_request/cif) with CCDC numbers 2106794(2) and 2106795(1).

Supplementary Material

nwad327_Supplemental_File

ACKNOWLEDGEMENTS

We thank the staff from the BL17B1 beamline of the National Facility for Protein Science in Shanghai (NFPS) at Shanghai Synchrotron Radiation Facility, for assistance during data collection. We acknowledge the Center for Advanced Mössbauer Spectroscopy, Mössbauer Effect Data Center, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, for providing the Mössbauer measurement and analysis. We also thank Dr. Yang-Yang Sun for his help in compound characterizations. Computational resources are supported by the Center for Computational Science and Engineering at Southern University of Science and Technology.

Contributor Information

Gan Xu, Jiangsu Collaborative Innovation Center of Biomedical Functional Materials and Jiangsu Key Laboratory of New Power Batteries, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210023, China.

Yun-Shu Cui, Department of Chemistry and Guangdong Provincial Key Laboratory of Catalytic Chemistry, Southern University of Science and Technology, Shenzhen 518055, China.

Xue-Lian Jiang, Department of Chemistry and Guangdong Provincial Key Laboratory of Catalytic Chemistry, Southern University of Science and Technology, Shenzhen 518055, China.

Cong-Qiao Xu, Department of Chemistry and Guangdong Provincial Key Laboratory of Catalytic Chemistry, Southern University of Science and Technology, Shenzhen 518055, China.

Jun Li, Department of Chemistry and Guangdong Provincial Key Laboratory of Catalytic Chemistry, Southern University of Science and Technology, Shenzhen 518055, China; Department of Chemistry and Engineering Research Center of Advanced Rare-Earth Materials of Ministry of Education, Tsinghua University, Beijing 100084, China.

Xu-Dong Chen, Jiangsu Collaborative Innovation Center of Biomedical Functional Materials and Jiangsu Key Laboratory of New Power Batteries, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210023, China; State Key Laboratory of Coordination Chemistry, Nanjing University, Nanjing 210093, China.

FUNDING

This work was supported by the National Natural Science Foundation of China (92261107 and 22071110 to X.D.C., 22388102 and 22033005 to J.L., and 22103035 to C.Q.X.), the National Key Research and Development Program of China (2023YFA1506600, 2022YFA1503900 and 2022YFA1503000), Guangdong Basic and Applied Basic Research Foundation (2020A1515110282 to C.Q.X.), and Guangdong Provincial Key Laboratory of Catalysis (2020B121201002 to J.L.). X.D.C. also acknowledges financial support from the Priority Academic Program Development of Jiangsu Higher Educational Institutions, Jiangsu Collaborative Innovation Center of Biomedical Functional Materials, and State Key Laboratory of Coordination Chemistry in Nanjing University.

AUTHOR CONTRIBUTIONS

G.X. carried out the synthetic work and the analytical characterization, Y.S.C., X.L.J. and C.Q.X. performed the computational studies. J.L. directed the theoretical study, and X.D.C. directed the experimental study. All authors contributed to the discussion and co-writing of the manuscript.

Conflict of interest statement. None declared.

REFERENCES

  • 1. Kuang  T. A breakthrough of artificial photosynthesis. Natl Sci Rev  2016; 3: 2–3. 10.1093/nsr/nwv071 [DOI] [Google Scholar]
  • 2. Huang  B, Zhang  H, Gan  W  et al.  Nb12+—niobespherene: a full-metal hollow-cage cluster with superatomic stability and resistance to CO attack. Natl Sci Rev  2023; 10: nwac197. 10.1093/nsr/nwac197 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Zhang  J, Feng  P, Bu  X  et al.  Atomically precise metal chalcogenide supertetrahedral clusters: frameworks to molecules, and structure to function. Natl Sci Rev  2022; 9: nwab076. 10.1093/nsr/nwab076 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Geng  L, Weng  M, Xu  C-Q  et al.  Co13O8—metalloxocubes: a new class of perovskite-like neutral clusters with cubic aromaticity. Natl Sci Rev  2021; 8: nwaa201. 10.1093/nsr/nwaa201 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Yao  C, Guo  N, Xi  S  et al.  Atomically-precise dopant-controlled single cluster catalysis for electrochemical nitrogen reduction. Nat Commun  2020; 11: 4389. 10.1038/s41467-020-18080-w [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Li  X, Mitchell  S, Fang  Y  et al.  Advances in heterogeneous single-cluster catalysis. Nat Rev Chem  2023; 7: 754–67. 10.1038/s41570-023-00540-8 [DOI] [PubMed] [Google Scholar]
  • 7. Spatzal  T, Aksoyoglu  M, Zhang  L  et al.  Evidence for interstitial carbon in nitrogenase FeMo cofactor. Science  2011; 334: 940–40. 10.1126/science.1214025 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Lancaster  KM, Roemelt  M, Ettenhuber  P  et al.  X-ray emission spectroscopy evidences a central carbon in the nitrogenase iron-molybdenum cofactor. Science  2011; 334: 974–7. 10.1126/science.1206445 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Peters  JW, Lanzilotta  WN, Lemon  BJ  et al.  X-ray crystal structure of the Fe-only hydrogenase (CpI) from Clostridium pasteurianum to 1.8 angstrom resolution. Science  1998; 282: 1853–8. 10.1126/science.282.5395.1853 [DOI] [PubMed] [Google Scholar]
  • 10. Schlögl  R. Catalytic synthesis of ammonia—a ‘never-ending story’?  Angew Chem Int Ed  2003; 42: 2004–8. 10.1002/anie.200301553 [DOI] [PubMed] [Google Scholar]
  • 11. Khodakov  AY, Chu  W, Fongarland  P. Advances in the development of novel cobalt Fischer−Tropsch catalysts for synthesis of long-chain hydrocarbons and clean fuels. Chem Rev  2007; 107: 1692–744. 10.1021/cr050972v [DOI] [PubMed] [Google Scholar]
  • 12. Billas  IML, Châtelain  A, de Heer  WA. Magnetism from the atom to the bulk in iron, cobalt, and nickel clusters. Science  1994; 265: 1682–4. 10.1126/science.265.5179.1682 [DOI] [PubMed] [Google Scholar]
  • 13. Xu  X, Yin  S, Moro  R  et al.  Metastability of free cobalt and iron clusters: a possible precursor to bulk ferromagnetism. Phys Rev Lett  2011; 107: 057203. 10.1103/PhysRevLett.107.057203 [DOI] [PubMed] [Google Scholar]
  • 14. Niemeyer  M, Hirsch  K, Zamudio-Bayer  V  et al.  Spin coupling and orbital angular momentum quenching in free iron clusters. Phys Rev Lett  2012; 108: 057201. 10.1103/PhysRevLett.108.057201 [DOI] [PubMed] [Google Scholar]
  • 15. Bobadova-Parvanova  P, Jackson  KA, Srinivas  S  et al.  Density-functional investigations of the spin ordering in Fe13 clusters. Phys Rev B  2002; 66: 195402. 10.1103/PhysRevB.66.195402 [DOI] [Google Scholar]
  • 16. Wu  M, Kandalam  AK, Gutsev  GL  et al.  Origin of the anomalous magnetic behavior of the Fe13+ cluster. Phys Rev B  2012; 86: 174410. 10.1103/PhysRevB.86.174410 [DOI] [Google Scholar]
  • 17. Cox  DM, Trevor  DJ, Whetten  RL  et al.  Magnetic behavior of free-iron and iron oxide clusters. Phys Rev B  1985; 32: 7290–8. 10.1103/PhysRevB.32.7290 [DOI] [PubMed] [Google Scholar]
  • 18. Rohlfing  EA, Cox  DM, Kaldor  A. Photoionization measurements on isolated iron-atom clusters. Chem Phys Lett  1983; 99: 161–6. 10.1016/0009-2614(83)80551-X [DOI] [Google Scholar]
  • 19. Sakurai  M, Watanabe  K, Sumiyama  K  et al.  Magic numbers in transition metal (Fe, Ti, Zr, Nb, and Ta) clusters observed by time-of-flight mass spectrometry. J Chem Phys  1999; 111: 235–8. 10.1063/1.479268 [DOI] [Google Scholar]
  • 20. Lavallo  V, Grubbs  RH. Carbenes as catalysts for transformations of organometallic iron complexes. Science  2009; 326: 559–62. 10.1126/science.1178919 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Toniolo  D, Scopelliti  R, Zivkovic  I  et al.  Assembly of high-spin [Fe3] clusters by ligand-based multielectron reduction. J Am Chem Soc  2020; 142: 7301–5. 10.1021/jacs.0c01664 [DOI] [PubMed] [Google Scholar]
  • 22. Zhao  Q, Harris  TD, Betley  TA. [(HL)2Fe6(NCMe)m]n+ (m = 0, 2, 4, 6; n = −1, 0, 1, 2, 3, 4, 6): an electron-transfer series featuring octahedral Fe6 clusters supported by a hexaamide ligand platform. J Am Chem Soc  2011; 133: 8293–306. 10.1021/ja2015845 [DOI] [PubMed] [Google Scholar]
  • 23. Muñoz  SB, Daifuku  SL, Brennessel  WW  et al.  Isolation, characterization, and reactivity of Fe8Me12: Kochi's S = 1/2 species in iron-catalyzed cross-couplings with MeMgBr and ferric salts. J Am Chem Soc  2016; 138: 7492–5. 10.1021/jacs.6b03760 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Sánchez  RH, Willis  AM, Zheng  S-L  et al.  Synthesis of well-defined bicapped octahedral iron clusters [(trenL)2Fe8(PMe2Ph)2]n (n=0, −1). Angew Chem Int Ed  2015; 54: 12009–13. 10.1002/anie.201505671 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Bino  A, Ardon  M, Lee  D  et al.  Synthesis and structure of [Fe13O4F24(OMe)12]5−:  the first open-shell Keggin ion. J Am Chem Soc  2002; 124: 4578–9. 10.1021/ja025590a [DOI] [PubMed] [Google Scholar]
  • 26. Wen  W, Meng  Y-S, Jiao  C-Q  et al.  A mixed-valence {Fe13} cluster exhibiting metal-to-metal charge-transfer-switched spin crossover. Angew Chem Int Ed  2020; 59: 16393–7. 10.1002/anie.202005998 [DOI] [PubMed] [Google Scholar]
  • 27. Joris  VS, Rosa  P, Caneschi  A  et al.  Static and dynamic magnetic properties of an [Fe13] cluster. Phys Rev B  2006; 73: 014422. [Google Scholar]
  • 28. Hess  CR, Weyhermüller  T, Bill  E  et al. [{Fe(tim)}2]: an Fe–Fe dimer containing an unsupported metal–metal bond and redox-active N4 macrocyclic ligands. Angew Chem Int Ed  2009; 48: 3703–6. 10.1002/anie.200900725 [DOI] [PubMed] [Google Scholar]
  • 29. Lei  H, Guo  J-D, Fettinger  JC  et al.  Two-coordinate first row transition metal complexes with short unsupported metal−metal bonds. J Am Chem Soc  2010; 132: 17399–401. 10.1021/ja1089777 [DOI] [PubMed] [Google Scholar]
  • 30. Sunada  Y, Ishida  S, Hirakawa  F  et al.  Persistent four-coordinate iron-centered radical stabilized by π-donation. Chem Sci  2016; 7: 191–8. 10.1039/C5SC02601F [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Xu  G, Wang  Z, Ling  R  et al.  Ligand metathesis as rational strategy for the synthesis of cubane-type heteroleptic iron–sulfur clusters relevant to the FeMo cofactor. Proc Natl Acad Sci USA  2018; 115: 5089–92. 10.1073/pnas.1801025115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Roach  PJ, Woodward  WH, Castleman  AW  et al.  Complementary active sites cause size-selective reactivity of aluminum cluster anions with water. Science  2009; 323: 492–5. 10.1126/science.1165884 [DOI] [PubMed] [Google Scholar]
  • 33. Yuan  HK, Chen  H, Ahmed  AS  et al.  Density-functional study of Scn (n = 2–16) clusters: lowest-energy structures, electronic structure, and magnetism. Phys Rev B  2006; 74: 144434. 10.1103/PhysRevB.74.144434 [DOI] [Google Scholar]
  • 34. Dong  C, Gong  X. Magnetism enhanced layer-like structure of small cobalt clusters. Phys Rev B  2008; 78: 020409. 10.1103/PhysRevB.78.020409 [DOI] [Google Scholar]
  • 35. Rees  DC. Great metalloclusters in enzymology. Annu Rev Biochem  2002; 71: 221–46. 10.1146/annurev.biochem.71.110601.135406 [DOI] [PubMed] [Google Scholar]
  • 36. Hendrich  MP, Gunderson  W, Behan  RK  et al.  On the feasibility of N2 fixation via a single-site FeI/FeIV cycle: spectroscopic studies of FeI(N2)FeI, FeIV≡N, and related species. Proc Natl Acad Sci USA  2006; 103: 17107–12. 10.1073/pnas.0604402103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Hernández Sánchez  R, Bartholomew  AK, Powers  TM  et al.  Maximizing electron exchange in a [Fe3] cluster. J Am Chem Soc  2016; 138: 2235–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Nesbit  MA, Oyala  PH, Peters  JC. Characterization of the earliest intermediate of Fe-N2 protonation: CW and pulse EPR detection of an Fe-NNH species and its evolution to Fe-NNH2+. J Am Chem Soc  2019; 141: 8116–27. 10.1021/jacs.8b12082 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Creutz  SE, Peters  JC. Diiron bridged-thiolate complexes that bind N2 at the Fe(II)Fe(II), Fe(II)Fe(I), and Fe(I)Fe(I) redox states. J Am Chem Soc  2015; 137: 7310–3. 10.1021/jacs.5b04738 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. DeRosha  DE, Chilkuri  VG, Van Stappen  C  et al.  Planar three-coordinate iron sulfide in a synthetic [4Fe-3S] cluster with biomimetic reactivity. Nat Chem  2019; 11: 1019–25. 10.1038/s41557-019-0341-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Li  X, Kiran  B, Li  J  et al.  Experimental observation and confirmation of icosahedral W@Au12 and Mo@Au12 molecules. Angew Chem Int Ed  2002; 41: 4786–9. 10.1002/anie.200290048 [DOI] [PubMed] [Google Scholar]
  • 42. Pyykkö  P, Runeberg  N. Icosahedral WAu12: a predicted closed-shell species, stabilized by aurophilic attraction and relativity and in accord with the 18-electron rule. Angew Chem Int Ed  2002; 41: 2174–6. [DOI] [PubMed] [Google Scholar]
  • 43. Hu  H-C, Hu  H-S, Zhao  B  et al.  Metal–organic frameworks (MOFs) of a cubic metal cluster with multicentered MnI-MnI bonds. Angew Chem Int Ed  2015; 54: 11681–5. 10.1002/anie.201504758 [DOI] [PubMed] [Google Scholar]
  • 44. Cui  P, Hu  H-S, Zhao  B  et al.  A multicentre-bonded [ZnI]8 cluster with cubic aromaticity. Nat Commun  2015; 6: 6331. 10.1038/ncomms7331 [DOI] [PubMed] [Google Scholar]
  • 45. Sheong  FK, Zhang  J-X, Lin  Z. Principal interacting spin orbital: understanding the fragment interactions in open-shell systems. Phys Chem Chem Phys  2020; 22: 10076–86. 10.1039/D0CP00127A [DOI] [PubMed] [Google Scholar]
  • 46. Pyykkö  P, Atsumi  M. Molecular single-bond covalent radii for elements 1–118. Chem Eur J  2008; 15: 186–97. 10.1002/chem.200800987 [DOI] [PubMed] [Google Scholar]
  • 47. Sheong  F, Zhang  J-X, Lin  Z. Revitalizing spin natural orbital analysis: electronic structures of mixed-valence compounds, singlet biradicals, and antiferromagnetically coupled systems. J Comput Chem  2019; 40: 1172–84. 10.1002/jcc.25762 [DOI] [PubMed] [Google Scholar]
  • 48. Jia  Y, Xu  C-Q, Cui  C  et al.  Rh19: a high-spin super-octahedron cluster. Sci Adv  2023; 9: eadi0214. 10.1126/sciadv.adi0214 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Scott  A, Galico  D, Bogacz  I  et al.  High spin and reactive Fe13 cluster with exposed metal sites. Angew Chem Int Ed  2023; 62: e202313880. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Pipek  J, Mezey  PG. A fast intrinsic localization procedure applicable for ab initio and semiempirical linear combination of atomic orbital wave functions. J Chem Phys  1989; 90: 4916–26. 10.1063/1.456588 [DOI] [Google Scholar]

Associated Data

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

Supplementary Materials

nwad327_Supplemental_File

Data Availability Statement

Crystallographic data can be obtained free of charge from the Cambridge Crystallographic Data Centre (www.ccdc.cam.ac.uk/data_request/cif) with CCDC numbers 2106794(2) and 2106795(1).


Articles from National Science Review are provided here courtesy of Oxford University Press

RESOURCES