Next Article in Journal
Fate of Dissolved Organic Matter and Cooperation Behavior of Coagulation: Fenton Combined with MBR Treatment for Pharmaceutical Tail Water
Previous Article in Journal
Chemical Properties, Preparation, and Pharmaceutical Effects of Cyclic Peptides from Pseudostellaria heterophylla
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring Single-Molecular Magnets for Quantum Technologies

1
UCL Department of Physics and Astronomy, University College London, Gower Street, London WC1E 6BT, UK
2
Yunnan Key Laboratory of Metal-Organic Molecular Materials and Devices, Kunming University, Kunming 650214, China
3
Key Laboratory of Yunnan Provincial Higher Education Institutions for Organic Optoelectronic Materials and Devices, Kunming University, Kunming 650214, China
4
School of Physical Science and Technology, Kunming University, Kunming 650214, China
5
School of Physics, Peking University, Chengfu Road 209, Beijing 100871, China
6
Department of Chemical Engineering, Pohang University of Science and Technology, 77 Cheongam-Ro, Nam-Gu, Pohang 37673, Republic of Korea
7
School of Physics and Astronomy, Center for Optoelectronics Engineering Research, Yunnan University, Kunming 650091, China
*
Author to whom correspondence should be addressed.
Current address: Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China.
Molecules 2025, 30(12), 2522; https://doi.org/10.3390/molecules30122522
Submission received: 12 December 2024 / Revised: 31 May 2025 / Accepted: 5 June 2025 / Published: 9 June 2025
(This article belongs to the Section Physical Chemistry)

Abstract

:
A single-molecule magnet (SMM) is a molecule that functions as a magnet. SMMs can be explored not only for emerging technology but also the fundamental science of their quantum nature, nanometer sizes, and their ease of engineering. This review encompasses the state-of-the-art experiments and theories developed so far for SMMs. We briefly explore their experimental synthesis and characterization. In the experimental synthesis, we cover ‘Click Chemistry’ and supramolecular chemistry. The main experimental characterizations comprise superconducting quantum interference devices, electron paramagnetic resonance, neutron scattering, and X-ray magnetic circular dichroism. The theoretical and computational works based on the density functional theory, the post-Hartree–Fock methods, and the theory of open quantum systems are discussed. Moreover, we exemplify the numerous promising research areas for SMMs by discussing quantum technologies. We envision a brilliant future for the fundamental research and emerging applications of SMMs.

1. Introduction

As the name suggests, a single-molecule magnet (SMM) is a molecule that acts like a tiny magnet. SMMs have been developed since the first reported synthesis of [ Mn 12 O 12 ( O 2 CR)16( H 2 O)4] ( Mn 12 ). Mn 12 is a high-spin molecule displaying slow magnetic relaxation at low temperatures [1]. This type of molecule-based magnet was first named a single-molecule magnet after the fabrication of [ Mn IV Mn 3 III O 3 X] [2]. In Figure 1, we have shown some typical examples of SMMs—Mn 12 (the SMM synthesized in the early 1990s) [1], Dy 4 polynuclear lanthanide SMM [3], 3 d 4 f Co 2 -Dy 2 SMM [4], and the Dy(III) single-ion magnet (SIM) that contains only one magnetic metal ion [5].
SMMs have attracted unprecedented research interest since their birth, owing to their promise for many research areas, such as magnetic storage, spintronics, quantum technologies, and structural biology [1,2,6,7,8,9,10,11,12,13,14,15,16] (Figure 2). Moreover, SMMs could play an important role as a platform for studying fundamental quantum science through a large variety of spin models and the flexibility of ligand engineering. The material demands of emerging technologies [10,12] provide fantastic opportunities to physicists and chemists to model, design, and synthesize new SMM species.
SMMs are also suitable for the study of quantum tunneling between different magnetic states (bi-stable states), raising intensive interest in molecular quantum science [17,18]. SMMs can be used as the molecular quantum bit (qubit) that has a two-dimensional Hilbert space. If the SMM has a stable spin state higher than 1 2 , it can even work as a qudit with d dimensions (normally d > 2 ). This could offer a series of benefits for quantum technology, such as more robust quantum information storage [12,19]. Molecular spintronics, exploring spin-polarized electrical currents for information processing using molecules, could also be based on SMMs, taking advantage of the great tunability offered by supramolecular chemistry [6,10].
There are already plenty of great dedicated review articles focusing on the state-of-the-art synthesis, structural, and magnetic characterizations of SMMs [9,14]. Complementary to these topics, this review will endeavor to cover (i) the experimental synthesis and characterization in brief, (ii) the theoretical and computational perspectives of SMMs, and (iii) the promising potential of SMMs for emerging technologies including high-density information storage, opto-spintronics, quantum technology, and single-molecule toroics.
The following discussion will fall into five sections. In Section 2, state-of-the-art experimental developments will be briefly covered. In Section 2.2, we will discuss the important magnetic properties of SMMs. In Section 3, we will cover theories and simulations. In Section 4, we will present the applications of SMMs to QT. At the end, in Section 5, we will draw some general conclusions.

2. The Experimental Developments of SMMs

2.1. Chemical Synthesis and Molecular Ligand Engineering

An extensive range of SMMs have been synthesized, which indicates the great power of molecular engineering associated with Click Clack reactions in supramolecular chemistry [20].
The key principle for fabricating SMMs is that they need to have not only excellent magnetic properties but also stability under normal conditions. For example, SMMs will not decompose in the presence of atmospheric oxygen and natural humidity. This criterion is important for simplifying the production of SMM-based technologies. Supramolecular chemistry is of secondary nature, since it is most important to first determine the principles of constructing the necessary coordination units using specially created ligands. Then, using non-covalent (i.e., supramolecular) interactions will ensure the correct arrangement of SMMs in the crystalline space, effectively suppressing the intermolecular magnetic exchange interactions undesirable for this type of magnet. Also, supramolecular chemistry approaches can be useful for the monolayer application of the most promising SMM on the surface.
To fabricate SMMs containing 3 d and/or 4 f magnetic chemical elements effectively, there are two main approaches: (a) using ligands carefully designed to obtain appropriate coordination functional groups for 3 d , 4 f , or both metal ions, and (b) exploring co-ligands to assist the self-assembly processes to integrate lanthanide ions into existing ligands [21]. The two approaches can be called the designed assembly approach (DAA) and assisted self-assembly approach (ASA), respectively. Moreover, site-targeted reactions are another promising route to synthesize SMMs. Establishing the correlations between the molecular structure and the properties is important for the iterative predictable tuning of SMM properties by changing the ligand structure. For the 3 d electrons, the orbital overlap is dominant, while the bonding between lanthanides carrying 4 f electrons and the ligand is ionic.
In the DAA, one can design organic ligands such as Schiff-based ligands. Schiff-based ligands could provide an ideal route to fabricate SMMs as they can have distinct coordinate pockets, thus binding many kinds of metal ions. In addition, oximes and pyridonate ligands can also be used for the DAA. In addition, amino acids, alkylol amines, flexible hexa-dentate ligands, Calix[4]arenes (H4L54, bowl-shaped ligands), and hexaimine macrocyclic ligands (H6L55R, R = Pr, Ox, Ph, Bu) can be useful for the DAA. For the ASA, acetate and nitrate can bridge co-ligands, while Hfac , N 3 , SCN , and C 2 O 4 can work as terminal ligands. A more detailed review of the synthesis of SMMs can be found in Refs. [21,22]. Although the DAA and ASA can be used to synthesize stable structures, most coordinated structures could be detrimental to the expected SMM properties. However, highly active reagents can be used to remove unnecessary coordinated anions or solvents. The design of organic ligands compatible with the electronic structure of magnetic elements is therefore crucial.

2.2. Experimental Characterizations

Up to now, huge experimental efforts have been made to study the magnetic properties of SMMs, which are governed by quantum mechanics in principle. The quantum mechanical spin Hamiltonian below governs the magnetic behavior of SMMs under an external magnetic field.
H ^ = i , j J i j s ^ i · s ^ j + i D i s ^ i z 2 + E i ( s ^ i x 2 s ^ i y 2 ) + i μ B B · g e · s ^ i + i , j A i j s ^ i · I ^ j + j μ N B · g N · I ^ j
Here, s ^ i , j ( I ^ i , j ) denotes spins due to the electron (nuclear) in the SMM, J i j refers to the exchange interaction between spins s ^ i and s ^ j , A i j is the hyperfine interaction between s ^ i and I ^ j , D i and E i are the uniaxial and biaxial magnetic anisotropy for spin i, respectively, B is the magnetic field, g e ( g N ) is the electron (nuclear) g-tensor, and μ B ( μ N ) is the Bohr magneton (nuclear magneton).
The blocking temperature ( T B ) and effective tunneling energy barrier ( U eff ) are the crucial parameters for the magnetic properties of SMMs. Previously, T B was defined as the temperature at which the magnetic relaxation time is 100 s [14]. This definition provides a standard method to compare the magnetic properties of SMMs. A more flexible definition for T B is the temperature at which the magnetic relaxation is slower with reference to the time scale of the characterization instrument.
In addition, T B can be defined from alternative perspectives including thermal energy, hysteresis, magnetic susceptibility, zero-field-cooling (ZFC), field-cooling (FC) divergence, and the magnetic tunneling effect [7,23,24,25,26]. In terms of the competition between the thermal energy and the magnetic anisotropy, T B can be defined as the temperature below which the thermal energy can not overcome the magnetic anisotropy. This is consistent with the definition based on spin relaxation because spin relaxation is driven by thermal fluctuations. T B can also be defined as the highest temperature at which magnetic hysteresis can be observed. From the point of view of cooling, T B can be defined as the temperature where the ZFC and FC curves diverge. The peak of the out-of-phase component χ of the temperature-dependent AC magnetic susceptibility can also refer to the blocking temperature. The temperature at which the quantum tunneling effect is negligible can also be defined as T B .
U eff is defined as the spin inversion energy barrier from spin-up to spin-down states, or vice versa, which is related to quantum tunneling of magnetization (QTM). Suppression of QTM is important for the further development of SMMs [27,28,29]. U eff defines how SMMs relax at a particular temperature T as described in the formula τ 1 = τ 0 1 e U eff / k B T . Here, τ is the thermal relaxation time, τ 0 is the attempt time (typically between 10 10 and 10 9 s), and k B is the Boltzmann constant. The 3 d -block SMMs show the following relationship between U eff and the magnetic anisotropy: U eff = | D | S 2 and U eff = | D | ( S 2 1 4 ) for the integer and the half-integer spin quantum number S, respectively. However, the correlation between the U eff , the magnetic anisotropy, and the spin quantum number is much more complicated [9] for f-block SMMs. SMMs, carrying a spin angular moment, are a natural candidate for quantum information storage units. SMMs can inherently work as a qudit, where d refers to the dimension of the quantum information storage unit represented by the magnetic moment. For example, spin- 1 2 is a qubit, with two dimensions spanned by spin-up and spin-down. To assess SMMs’ properties for quantum computing, the knowledge of the spin–lattice relaxation time ( T 1 ) and the spin–spin relaxation time T 2 is necessary. The spin–lattice relaxation time is the lifetime for the spin to survive as a classical magnet, whereas the spin–spin relaxation time is a measure of the time limit of quantum coherence.
U e f f and T B are important parameters for slow magnetic relaxation. More deeply speaking, the magnetic anisotropy, quantum tunneling, and the crystal field environment, especially the symmetry, play a crucial role in slow magnetic relaxation. A summary of U eff and T B is illustrated in Figure 3, where we have shown polynuclear oxo-bridged transition metal SMMs (in black) [1,30,31,32], cyano-bridged SMMs (in blue) [33,34], lanthanide SIM (in pink) [35], 3 d 4 f SMM (in green) [36,37,38], polynuclear lanthanide SMMs (in purple) [39,40], actinide-based SMMs (in orange) [41,42], transition metal SIMs (in brown) [43,44,45], organometallic lanthanide SIMs (in red) [5,46,47,48,49,50,51,52,53,54], and a di-lanthanide SMM (in cyan) [55].
Progress has been made through continuous efforts to engineer ligands and metallic ions. The best SMM so far, the dysprosium-based lanthanide SIM, has reached a T B up to 80 K and a U eff exceeding 1500 cm−1. Recently, there has been substantial development on combining radicals into SMM molecular structures to enhance exchange interactions and the coercive field (up to 3 Tesla) by using Ln4 metallocene complexes ( T B = 6.4 K, U eff = 91 cm 1 ) [56]. Dysprosium (III) has also been bound with dicarbollide ligands to improve the magnetic relaxation ( T B = 6.8 K, U eff = 559 cm 1 ) [57]. Research on SMMs also points to combining SMMs into other material structures such as metallic–organic frameworks and interesting surfaces to stabilize the SMM molecular structures, which could facilitate the design of quantum technology devices based on SMMs [58,59,60]. A high blocking temperature of 17 K ( U eff = 427 cm 1 ) has been reported for SMMs on a graphene/Ir surface. A monolayer of SMMs on a graphene surface has impressive performance with T B = 28 K, U eff = 556 cm 1 [60]. A blocking temperature of 28 K on a surface is indeed encouraging for using SMMs as spintronic or quantum devices. More recently, di-lanthanide SMMs have been fabricated, suggesting a comparable performance to an organometallic lanthanide SIM, which points to a promising route to enhance magnetic properties by forming metal–metal bonds—an idea extended from transition metals [55]. However, the highest blocking temperature achieved so far is still far below room temperature, which needs further theoretical and experimental efforts.
The experimental characterization methods have been summarized in Figure 4, and mainly include superconducting quantum interference devices (SQUIDs), electron paramagnetic resonance (EPR), neutron scattering (NS), X-ray magnetic circular dichroism (XMCD), and spin-polarized scanning tunneling microscopy (SP-STM) [12].
Observation of magnetic properties can be performed by using SQUID, which explores superconducting loops that are sensitive to a tiny change in magnetic flux. Two types of SQUID measurements are generally used, including AC and direct current (DC). AC SQUID can be used to observe the relaxation of SMMs while DC SQUID can be used for static magnetization measurements. Using DC SQUID, the magnetic exchange interaction and magnetic anisotropy can be extracted by fitting the experimental data to a spin Hamiltonian, as shown in Equation (1).
Neutron scattering (NS), using a neutron beam to probe a material’s physical properties, is a very powerful technique for studying nanoscale materials’ crystal structure, magnetic structure, magnetic excitation, spin waves, etc. [61]. Elastic NS can be used to probe static crystal and magnetic structures. Inelastic neutron scattering (INS) is another important experimental technique to capture the magnetic properties of SMMs. In INS experiments, spins exchange energy and momentum with neutrons, which allows us to probe magnetic ordering, spin waves, and magnetic excitation; these properties are closely related to exchange interactions and magnetic anisotropy. Usually, we can obtain scattering cross-sections as a function of the wave vector ( Q ) and energy transfer (E), thus providing four-dimensional INS. Recently, inelastic neutron scattering experiments have been considered as a crucial tool to study SMMs [12,62].
Electron paramagnetic resonance (EPR) is a powerful tool to study the dynamics due to the interactions in SMMs, including the Zeeman energy, exchange interaction, zero-field splitting, and hyperfine interaction. The principle of EPR is to use a continuous-wave or pulsed microwave field to excite the system and record the dynamical magnetic response from it. In this way, the parameters in the spin Hamiltonian in Equation (1) can be fitted by using experimental spectra and the system is characterized. EPR is also a nice method to measure the shift and anisotropy of g-factors [63]. Nowadays, time-resolved EPR (TREPR) is applied widely to observe the spin dynamics in real time, especially under optical excitation [64,65,66]. TREPR could be a very powerful technique for spin-based quantum computing, where observing real-time spin dynamics is important for quantum operations including initialization, gate operations, and readout.
X-ray magnetic circular dichroism (XMCD) is another powerful experimental tool to observe the magnetic properties of SMMs, and it explores the difference between spectra using left- and right-circular light. XMCD is especially suitable for studying the magnetic properties of SMMs on surfaces [67,68,69,70]. The downside is that it is difficult for XMCD to address individual molecules on surfaces because it is a bulk measurement. It might be possible to develop confocal XMCD that can focus the light on a small spot such that individual molecules could be addressed optically. For this purpose, recently, nanoscale imaging for magnetic domains has been achieved experimentally [71].
Scanning tunneling microscopy (STM) was developed to use a metallic tip to probe the substances on surfaces through the tunneling electric current between the tip and surface. Spin-polarized STM (SP-STM) adds spin polarization on top of STM through a magnetically coated STM tip, which can sense magnetic substances on the surface. SP-STM would be a useful tool to study SMMs on surfaces and has the potential to effectively address individual molecules [72]. Recently, DyPc 2 double-decker molecules have been investigated using SP-STM, and readout of spin states in different magnetic fields has been achieved [73].
Apart from the methods we discussed in detail above, there are also others used to characterize SMMs [74]. Single-crystal X-ray diffraction (XRD) and X-ray absorption spectroscopy, such as extended X-ray absorption fine structure (EXAFS) and X-ray absorption near edge structure (XANES), are effective for establishing magnetic–structural correlations [75,76]. By correlations, we mean the relationship between the influence of ligand modifications on the structure of coordination units, i.e., the parameters of magnetic anisotropy of the metal ion, and accordingly, the key physical and chemical properties of SMMs (slow relaxation of magnetization, blocking temperature, effective tunneling energy barrier, the shape of the hysteresis, and relaxation mechanisms). Reliable primary data on molecular and crystal structures from XRD and EXAFS/XANES are also very important for harnessing the full power of quantum chemical calculation methods. Electronic absorption spectroscopy (EAS) can also be used to study the electronic structure of the ground state of SMMs [77]. Moreover, luminescence spectroscopy (LS) can be employed to reveal the dependence of the energy of magnetic sub-levels on the applied magnetic field [78]. Recently, nuclear magnetic resonance (NMR) spectroscopy has also gained popularity for SMM studies [79,80]. Since NMR spectrometers are very widespread and their availability is higher than that of magnetometers, they can be an easily accessible alternative for measurements of molecular magnetic susceptibility by the Evans method and for the determination of the g-tensors, zero-field splitting energy, and ligand field parameters for SMMs in solids and solutions [79].

3. The Theoretical Developments of SMMs

The computation of the electronic structures of SMMs can be used to gain important insights into their magnetic properties, including exchange interactions and magnetic anisotropy, both of which are crucial to understand the magnetic properties of SMMs. To calculate magnetic anisotropy, we only need to take into account the spin–orbit interaction and spin–spin dipolar interactions for magnetic anisotropy. By contrast, we need to perform a set of calculations with distinct spin configurations to obtain the exchange interaction. Extensive computational and theoretical works have been carried out for SMMs, including studies on the electronic structure [81,82,83,84,85,86,87,88], exchange interactions [87,89,90,91,92,93,94,95,96,97,98], spin–phonon coupling [99], the effect of spin–orbit interactions [100], and relaxation rates [101,102,103,104,105,106]. A more general and comprehensive review about molecular spectroscopy is given by Ref. [107], which not only discusses the magnetic properties but also vibrational and rotational spectroscopies.
The first-principles computational methods widely used include density functional theory (DFT) [108] and quantum chemistry post-Hartree–Fock (HF) [109]. DFT is a rigorous theory that relies on charge density to predict the magnetic properties of a many-body system, whereas post-HF methods, such as configuration interaction (CI), are based on HF calculations to construct multi-determinant wave functions to obtain a better description of ground and excited states. The DFT methods include local-density approximation, generalized gradient approximation (GGA), meta-GGA, DFT + U and hybrid-exchange DFT [110,111]. As the localized d- and f-electrons are important for the magnetic properties of SMMs, we will focus our discussion on DFT+U and hybrid-exchange DFT. DFT+U, based on the conventional DFT framework, introduces a Hubbard-U parameter in the Hamiltonian to account for the electron correlation and localized orbitals due to d or f block magnetic atoms [112]. Hybrid-exchange DFT tries to mix the conventional DFT exchange–correlation functional (normally GGA) with an optimal portion of the HF exact exchange between Kohn–Sham orbitals, which is widely used to take into account electron correlations, especially in strongly correlated systems. However, the effectiveness of hybrid-exchange DFT is still controversial, which leads to the recent dedication to improve the hybrid methods or explore alternative routes [113].
The calculations of exchange interactions for SMMs, which is a challenging yet beneficial task, have been carried out predominantly by using the framework of hybrid-exchange DFT such as in Ref. [89], which is relatively efficient but on the other hand controversial in terms of calculation accuracy. Post-HF methods widely used for molecular magnets include CI, complete active space self-consistent field (CASSCF) [114], etc. CASSCF is similar to CI, but a complete active space (a set of molecular orbitals) can be defined normally near the gap between the highest-occupied molecular orbital and the lowest-unoccupied molecular orbital. An advantage of CASSCF over CI is that the molecular orbital coefficients and the CI coefficients can be determined self-consistently. The post-HF methods can provide highly accurate results but are much more expensive than DFT. In particular, the difference-dedicated configuration interaction (DDCI) can be explored to compute the exchange interaction quantitatively and analyze the mechanism for exchange interactions [115]. It is therefore very desirable to develop a theoretical framework in between, i.e., a more accurate computational method than DFT but less expensive than DDCI. Most recent work points to the decomposition method to compute exchange interactions part by part [116]. Moreover, CASSCF can also be used to compute the electronic structure, g-factors, and magnetic anisotropy [102,117,118,119]. CASSCF/CASPT2 (CAS-second-order perturbation theory)/RASSI-SO (restricted active space state interaction–spin orbit)/atomic mean-field approximation (AMFI) has been widely used for the computation of the parameters in the spin Hamiltonian, shown in Equation (1) [117]. The calculated results can be compared with the experiments carried out using the techniques aforementioned [117].
DFT, DDCI, and CASSCF use the same principles to compute the exchange interaction, i.e., comparing the difference in the total interaction energies with a variety of spin configurations. In fact, much of the computational time has been spent twice for the same matrix elements of either one-electron or two-electron potentials. From this point of view, Green’s function perturbation theory (GFPT) might be useful, as this theoretical frame can directly find the spin-flipping route according to the interaction V and propagator G 0 , which is essentially similar to Feynman’s path integral formalism. However, even with GFPT methods, we still need to be careful about the accuracy of the matrix elements of the Fock matrix from either DFT or quantum chemistry calculations based on HF, which are needed for GFPT calculations.
The DFT simulations of SMMs on surfaces are particularly interesting as they are associated with a proper isolation of MMs for many useful purposes such as spintronics. Recently, simulations of this type have been carried out for coordinated lanthanide–transition-metal clusters on gold surfaces [120,121].
Another interesting research direction is to predict the optically driven exchange interaction between spins in SMMs, which can shed light on the control of molecular spins and magnetic interactions. In addition, fundamentally, this is a dynamical process involving not only the electronic state but also vibrational and rotational degrees of freedom [107].
Moreover, spin dynamics simulations within a theoretical framework of open quantum systems are becoming increasingly important [19] for a comprehensive description of how SMMs interact with the environment. The Markovian approximation for quantum dynamics, which can describe the weak coupling limit, could be useful in studying quantum systems while taking into account the perturbation from the environment [122]. This type of simulation could be useful for spin systems under external stimuli such as optical excitation. The Orbach barrier is related to the Orbach spin relaxation process involving intermediate phonon states. The Orbach barrier is then the energy needed for the Orbach process involving two phonons. Recently, a theoretical work based on the Redfield equation and non-equilibrium Green’s function method has been carried out to interpret the unexpected low Orbach barrier that is normally set by the magnetic anisotropy [86].
The software for computing the electronic structure and magnetic properties of SMMs includes Gaussian, SIESTA (Spanish Initiative for Electronic Simulations with Thousands of Atoms), ORCA, MOLCAS, CP2K, WIEN2K, Turbomole, and Amsterdam Modelling Suites (AMS) [88,123,124,125,126,127,128,129,130]. Most of these software packages use a Gaussian-type basis set, while AMS uses the Slater-type orbital [88].

4. Quantum Science and Technologies Based on SMMs

Quantum technologies in general include quantum sensing, quantum communication, and quantum computing. Quantum sensing, which has been performed already, is another promising direction for SMMs [131]. The quantum sensor (SMM) can interact with the external field or systems; the resulting change in spin properties can be considered the information acquired during sensing. Otherwise, the SMM quantum sensor can be embedded within the system or field under probing to obtain information about this system. Magnetic molecules can work as quantum thermometers, local temperature sensors, and local probes. In particular, molecular spin ensembles could be used to detect dark matter particles by using resonance with the effective field of axions (“cold” dark matter particles). Quantum communication relying on the excellent portability and tunability of molecular networks has also been proposed and assessed in detail [132]. Recent experimental and theoretical works have pointed to promising optically addressable spin-bearing molecules for quantum circuits and computing [10,12,19,133]. First-principles computational methods using the theory of open quantum systems have been reviewed in Ref. [133].

5. Conclusions and Outlook

In summary, we have reviewed the development of the research on SMMs so far and discussed their huge potential for emerging science and technology. We believe that only a tiny fraction of the literature about SMMs has been covered here, with many important works outside the scope of this paper. For chemical synthesis, we briefly cover click chemistry and supramolecular chemistry, followed by some details about the ASA and DAA methods. The experimental developments for the characterization of SMMs have been discussed briefly, including SQUID, EPR, NS, XMCD, XRD, XAFS/XANES, EAS, LS, and NMR. Here, TREPR on SMMs is particularly promising for developing quantum technology. XMCD has the potential to address SMMs on surfaces. The characterization of T B , U eff , T 1 , and T 2 parameters has also been discussed. T B is still much lower than room temperature, thus hindering the practical applications of SMMs. We expect more exciting research on the enhancement of T B in the near future. The measurements of U eff , T 1 , and T 2 so far show that SMMs are very promising for emerging technologies such as quantum computing and spintronics. In addition, new theoretical techniques need to be developed to achieve more efficient computation of the exchange interaction, which is a crucial parameter for SMMs. Open quantum system simulations need to be strengthened to be able to match up with the experiments that are normally performed along with the noise from the environment. A few fascinating emerging technologies, namely high-density information storage, opto-spintronics, quantum technology, and single-molecule toroics, have been discussed. SMMs could be a good platform to realize opto-spintronics by combining the spin of the metal and the optical response of the ligand. Quantum technology is another area where SMMs could thrive as quantum information storage units for molecular networks. In summary, we anticipate an exciting era for research on SMMs.

Author Contributions

W.W. designed, planned and wrote the paper. T.H. checked literatures. J.Z., T.Z. and H.W. proofread the paper. All authors have read and agreed to the published version of the manuscript.

Funding

W. W. thanks the funding support from UK Research Councils Basic Technology Program (EP/F041349/1), UKRI Atomically Deterministic Doping and Readout For Semiconductor Solotronics Project (ADDRFSS, EP/M009564/1), the EU Horizon 2020 Project Marketplace (No. 760173), and the UK science and technology facilities council for funding. H. W. thanks the National Natural Science Foundation of China (No. 62164006), Yunnan Fundamental Research Projects (No.202201AS070008). This work was also supported by the National Natural Science Foundation of China (62164006), Yunnan Provincial Department of Science and Technology Project (202201AS070008, 202105AG070025), Scientific Research Fund of Yunnan Education Department (2023Y0886).

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The authors gratefully acknowledge the colleagues at Kunming University, UCL and Imperial College London for helpful discussions.

Conflicts of Interest

There are no conflicts to declare.

Nomenclature

COT 2 1,3,5,7-cyclooctatetraene
Cp iPr 5 penta-iso-propylcyclopentadienyl
Cp Me 4 H tetramethylcyclopentadienyl
Cp ttt C 5 H 2 t Bu3-1,2,4
Cp * pentamethylcyclopentadienyl
Dmp2,6-dimesitylphenyl
mntmaleonitriledithiolate
O t Butert-butoxide
Pcphthalocyanine
Piv2,2-dimethylpropanoate
Pypyridine
TTTT Br 2, 2 , 2 -(((nitrilotris(ethane-2,1 -diyl))tris(azanediyl))tris(methylene))tris-(4-bromophenol)

References

  1. Sessoli, R.; Tsai, H.L.; Schake, A.R.; Wang, S.; Vincent, J.B.; Folting, K.; Gatteschi, D.; Christou, G.; Hendrickson, D.N. High-spin molecules: [Mn12O12 (O2CR)16 (H2O)4]. J. Am. Chem. Soc. 1993, 115, 1804–1816. [Google Scholar] [CrossRef]
  2. Aubin, S.M.J.; Wemple, M.W.; Adams, D.M.; Tsai, H.L.; Christou, G.; Hendrickson, D.N. Distorted MnIVMnIII3 Cubane Complexes as Single-Molecule Magnets. J. Am. Chem. Soc. 1996, 118, 7746–7754. [Google Scholar] [CrossRef]
  3. Blagg, R.J.; Ungur, L.; Tuna, F.; Speak, J.; Comar, P.; Collison, D.; Wernsdorfer, W.; McInnes, E.J.; Chibotaru, L.F.; Winpenny, R.E. Magnetic relaxation pathways in lanthanide single-molecule magnets. Nat. Chem. 2013, 5, 673–678. [Google Scholar] [CrossRef]
  4. Liu, J.L.; Wu, J.Y.; Huang, G.Z.; Chen, Y.C.; Jia, J.H.; Ungur, L.; Chibotaru, L.F.; Chen, X.M.; Tong, M.L. Desolvation-driven 100-fold slow-down of tunneling relaxation rate in Co (II)-Dy (III) single-molecule magnets through a single-crystal-to-single-crystal Process. Sci. Rep. 2015, 5, 16621. [Google Scholar] [CrossRef]
  5. Guo, F.S.; Day, B.M.; Chen, Y.C.; Tong, M.L.; Mansikkamäki, A.; Layfield, R.A. Magnetic hysteresis up to 80 kelvin in a dysprosium metallocene single-molecule magnet. Science 2018, 362, 1400–1403. [Google Scholar] [CrossRef]
  6. Bogani, L.; Wernsdorfer, W. Molecular spintronics using single-molecule magnets. Nat. Mater. 2008, 7, 179–186. [Google Scholar] [CrossRef] [PubMed]
  7. Gao, S.; Affronte, M. Molecular Nanomagnets and Related Phenomena; Springer: Berlin/Heidelberg, Germany, 2015. [Google Scholar]
  8. Ungur, L.; Chibotaru, L.F. Strategies toward high-temperature lanthanide-based single-molecule magnets. Inorg. Chem. 2016, 55, 10043–10056. [Google Scholar] [CrossRef]
  9. Shao, D.; Wang, X.Y. Development of Single-Molecule Magnets. Chin. J. Chem. 2020, 38, 1005–1018. [Google Scholar] [CrossRef]
  10. Coronado, E. Molecular magnetism: From chemical design to spin control in molecules, materials and devices. Nat. Rev. Mater. 2020, 5, 87–104. [Google Scholar] [CrossRef]
  11. Spree, L.; Popov, A.A. Recent advances in single molecule magnetism of dysprosium-metallofullerenes. Dalton Trans. 2019, 48, 2861–2871. [Google Scholar] [CrossRef]
  12. Moreno-Pineda, E.; Wernsdorfer, W. Measuring molecular magnets for quantum technologies. Nat. Rev. Phys. 2021, 3, 645–659. [Google Scholar] [CrossRef]
  13. Wang, J.H.; Li, Z.Y.; Yamashita, M.; Bu, X.H. Recent progress on cyano-bridged transition-metal-based single-molecule magnets and single-chain magnets. Coord. Chem. Rev. 2021, 428, 213617. [Google Scholar] [CrossRef]
  14. Martynov, A.G.; Horii, Y.; Katoh, K.; Bian, Y.; Jiang, J.; Yamashita, M.; Gorbunova, Y.G. Rare-earth based tetrapyrrolic sandwiches: Chemistry, materials and applications. Chem. Soc. Rev. 2022, 51, 9262–9339. [Google Scholar] [CrossRef]
  15. Ungur, L.; Lin, S.Y.; Tang, J.; Chibotaru, L.F. Single-molecule toroics in Ising-type lanthanide molecular clusters. Chem. Soc. Rev. 2014, 43, 6894–6905. [Google Scholar] [CrossRef]
  16. Zabala-Lekuona, A.; Seco, J.M.; Colacio, E. Single-Molecule Magnets: From Mn12-ac to dysprosium metallocenes, a travel in time. Coord. Chem. Rev. 2021, 441, 213984. [Google Scholar] [CrossRef]
  17. Yamashita, M. Next generation multifunctional nano-science of advanced metal complexes with quantum effect and nonlinearity. Bull. Chem. Soc. Jpn. 2021, 94, 209–264. [Google Scholar] [CrossRef]
  18. Shahed, S.M.F.; Ara, F.; Hossain, M.I.; Katoh, K.; Yamashita, M.; Komeda, T. Observation of Yu–Shiba–Rusinov states and inelastic tunneling spectroscopy for intramolecule magnetic exchange interaction energy of terbium phthalocyanine (TbPc) species adsorbed on superconductor NbSe2. Acs Nano 2022, 16, 7651–7661. [Google Scholar] [CrossRef]
  19. Gaita-Ariño, A.; Luis, F.; Hill, S.; Coronado, E. Molecular spins for quantum computation. Nat. Chem. 2019, 11, 301–309. [Google Scholar] [CrossRef] [PubMed]
  20. Glaser, T. Rational design of single-molecule magnets: A supramolecular approach. Chem. Commun. 2011, 47, 116–130. [Google Scholar] [CrossRef]
  21. Liu, K.; Shi, W.; Cheng, P. Toward heterometallic single-molecule magnets: Synthetic strategy, structures and properties of 3d–4f discrete complexes. Coord. Chem. Rev. 2015, 289-290, 74–122. [Google Scholar] [CrossRef]
  22. Maniaki, D.; Pilichos, E.; Perlepes, S.P. Coordination Clusters of 3d-Metals That Behave as Single-Molecule Magnets (SMMs): Synthetic Routes and Strategies. Front. Chem. 2018, 6, 461. [Google Scholar] [CrossRef]
  23. Gatteschi, D.; Sessoli, R.; Villain, J. Molecular Nanomagnets; Oxford University Press: New York, NY, USA, 2006; Volume 5. [Google Scholar]
  24. Friedman, J.R.; Sarachik, M.; Tejada, J.; Ziolo, R. Macroscopic measurement of resonant magnetization tunneling in high-spin molecules. Phys. Rev. Lett. 1996, 76, 3830. [Google Scholar] [CrossRef] [PubMed]
  25. Sessoli, R.; Gatteschi, D.; Caneschi, A.; Novak, M. Magnetic bistability in a metal-ion cluster. Nature 1993, 365, 141–143. [Google Scholar] [CrossRef]
  26. Thomas, L.; Lionti, F.; Ballou, R.; Gatteschi, D.; Sessoli, R.; Barbara, B. Macroscopic quantum tunnelling of magnetization in a single crystal of nanomagnets. Nature 1996, 383, 145–147. [Google Scholar] [CrossRef]
  27. Liu, H.; Li, J.F.; Yin, B. The coexistence of long τ QTM and high Ueff as a concise criterion for a good single-molecule magnet: A theoretical case study of square antiprism dysprosium single-ion magnets. Phys. Chem. Chem. Phys. 2022, 24, 11729–11742. [Google Scholar] [CrossRef]
  28. Yin, B.; Li, C.C. A method to predict both the relaxation time of quantum tunneling of magnetization and the effective barrier of magnetic reversal for a Kramers single-ion magnet. Phys. Chem. Chem. Phys. 2020, 22, 9923–9933. [Google Scholar] [CrossRef]
  29. Yin, B.; Luo, L. The anisotropy of the internal magnetic field on the central ion is capable of imposing great impact on the quantum tunneling of magnetization of Kramers single-ion magnets. Phys. Chem. Chem. Phys. 2021, 23, 3093–3105. [Google Scholar] [CrossRef]
  30. Tasiopoulos, A.J.; Vinslava, A.; Wernsdorfer, W.; Abboud, K.A.; Christou, G. Giant single-molecule magnets: A Mn84 torus and its supramolecular nanotubes. Angew. Chem. 2004, 116, 2169–2173. [Google Scholar] [CrossRef]
  31. Sangregorio, C.; Ohm, T.; Paulsen, C.; Sessoli, R.; Gatteschi, D. Quantum tunneling of the magnetization in an iron cluster nanomagnet. Phys. Rev. Lett. 1997, 78, 4645. [Google Scholar] [CrossRef]
  32. Milios, C.J.; Vinslava, A.; Wernsdorfer, W.; Moggach, S.; Parsons, S.; Perlepes, S.P.; Christou, G.; Brechin, E.K. A record anisotropy barrier for a single-molecule magnet. J. Am. Chem. Soc. 2007, 129, 2754–2755. [Google Scholar] [CrossRef]
  33. Sokol, J.J.; Hee, A.G.; Long, J.R. A cyano-bridged single-molecule magnet: Slow magnetic relaxation in a trigonal prismatic MnMo6(CN)18 cluster. J. Am. Chem. Soc. 2002, 124, 7656–7657. [Google Scholar] [CrossRef] [PubMed]
  34. Qian, K.; Huang, X.C.; Zhou, C.; You, X.Z.; Wang, X.Y.; Dunbar, K.R. A single-molecule magnet based on heptacyanomolybdate with the highest energy barrier for a cyanide compound. J. Am. Chem. Soc. 2013, 135, 13302–13305. [Google Scholar] [CrossRef] [PubMed]
  35. Ishikawa, N.; Sugita, M.; Ishikawa, T.; Koshihara, S.y.; Kaizu, Y. Lanthanide double-decker complexes functioning as magnets at the single-molecular level. J. Am. Chem. Soc. 2003, 125, 8694–8695. [Google Scholar] [CrossRef] [PubMed]
  36. Osa, S.; Kido, T.; Matsumoto, N.; Re, N.; Pochaba, A.; Mrozinski, J. A tetranuclear 3d–4f single molecule magnet: [CuIILTbIII (hfac) 2]2. J. Am. Chem. Soc. 2004, 126, 420–421. [Google Scholar] [CrossRef]
  37. Langley, S.K.; Chilton, N.F.; Ungur, L.; Moubaraki, B.; Chibotaru, L.F.; Murray, K.S. Heterometallic tetranuclear [LnIII2CoIII2] complexes including suppression of quantum tunneling of magnetization in the [DyIII2CoIII2] single molecule magnet. Inorg. Chem. 2012, 51, 11873–11881. [Google Scholar] [CrossRef]
  38. Huang, X.C.; Zhou, C.; Wei, H.Y.; Wang, X.Y. End-on azido-bridged 3d–4f complexes showing single-molecule-magnet property. Inorg. Chem. 2013, 52, 7314–7316. [Google Scholar] [CrossRef]
  39. Tang, J.; Hewitt, I.; Madhu, N.; Chastanet, G.; Wernsdorfer, W.; Anson, C.E.; Benelli, C.; Sessoli, R.; Powell, A.K. Dysprosium triangles showing single-molecule magnet behavior of thermally excited spin states. Angew. Chem. 2006, 118, 1761–1765. [Google Scholar] [CrossRef]
  40. Rinehart, J.D.; Fang, M.; Evans, W.J.; Long, J.R. Strong exchange and magnetic blocking in N23–radical-bridged lanthanide complexes. Nat. Chem. 2011, 3, 538–542. [Google Scholar] [CrossRef]
  41. Mills, D.P.; Moro, F.; McMaster, J.; Van Slageren, J.; Lewis, W.; Blake, A.J.; Liddle, S.T. A delocalized arene-bridged diuranium single-molecule magnet. Nat. Chem. 2011, 3, 454–460. [Google Scholar] [CrossRef]
  42. Mougel, V.; Chatelain, L.; Pécaut, J.; Caciuffo, R.; Colineau, E.; Griveau, J.C.; Mazzanti, M. Uranium and manganese assembled in a wheel-shaped nanoscale single-molecule magnet with high spin-reversal barrier. Nat. Chem. 2012, 4, 1011–1017. [Google Scholar] [CrossRef]
  43. Aromí, G.; Brechin, E.K. Synthesis of 3D Metallic Single-Molecule Magnets. In Single-Molecule Magnets and Related Phenomena; Springer: Berlin/Heidelberg, Germany, 2006; pp. 1–67. [Google Scholar]
  44. Zadrozny, J.M.; Xiao, D.J.; Atanasov, M.; Long, G.J.; Grandjean, F.; Neese, F.; Long, J.R. Magnetic blocking in a linear iron (I) complex. Nat. Chem. 2013, 5, 577–581. [Google Scholar] [CrossRef] [PubMed]
  45. Bunting, P.C.; Atanasov, M.; Damgaard-Møller, E.; Perfetti, M.; Crassee, I.; Orlita, M.; Overgaard, J.; van Slageren, J.; Neese, F.; Long, J.R. A linear cobalt (II) complex with maximal orbital angular momentum from a non-Aufbau ground state. Science 2018, 362, eaat7319. [Google Scholar] [CrossRef]
  46. Jiang, S.D.; Wang, B.W.; Su, G.; Wang, Z.M.; Gao, S. A Mononuclear Dysprosium Complex Featuring Single-Molecule-Magnet Behavior. Angew. Chem. 2010, 122, 7610–7613. [Google Scholar] [CrossRef]
  47. Goodwin, C.A.; Ortu, F.; Reta, D.; Chilton, N.F.; Mills, D.P. Molecular magnetic hysteresis at 60 kelvin in dysprosocenium. Nature 2017, 548, 439–442. [Google Scholar] [CrossRef]
  48. Meihaus, K.R.; Long, J.R. Magnetic blocking at 10 K and a dipolar-mediated avalanche in salts of the bis (η8-cyclooctatetraenide) complex [Er (COT)2]. J. Am. Chem. Soc. 2013, 135, 17952–17957. [Google Scholar] [CrossRef] [PubMed]
  49. Qin, L.; Singleton, J.; Chen, W.P.; Nojiri, H.; Engelhardt, L.; Winpenny, R.E.; Zheng, Y.Z. Quantum Monte Carlo Simulations and High-Field Magnetization Studies of Antiferromagnetic Interactions in a Giant Hetero-Spin Ring. Angew. Chem. Int. Ed. 2017, 56, 16571–16574. [Google Scholar] [CrossRef] [PubMed]
  50. Chen, Y.C.; Liu, J.L.; Ungur, L.; Liu, J.; Li, Q.W.; Wang, L.F.; Ni, Z.P.; Chibotaru, L.F.; Chen, X.M.; Tong, M.L. Symmetry-supported magnetic blocking at 20 K in pentagonal bipyramidal Dy (III) single-ion magnets. J. Am. Chem. Soc. 2016, 138, 2829–2837. [Google Scholar] [CrossRef]
  51. Liu, J.; Chen, Y.C.; Liu, J.L.; Vieru, V.; Ungur, L.; Jia, J.H.; Chibotaru, L.F.; Lan, Y.; Wernsdorfer, W.; Gao, S.; et al. A stable pentagonal bipyramidal Dy (III) single-ion magnet with a record magnetization reversal barrier over 1000 K. J. Am. Chem. Soc. 2016, 138, 5441–5450. [Google Scholar] [CrossRef]
  52. Gupta, S.K.; Rajeshkumar, T.; Rajaraman, G.; Murugavel, R. An air-stable Dy (III) single-ion magnet with high anisotropy barrier and blocking temperature. Chem. Sci. 2016, 7, 5181–5191. [Google Scholar] [CrossRef]
  53. Ding, Y.S.; Chilton, N.F.; Winpenny, R.E.; Zheng, Y.Z. On approaching the limit of molecular magnetic anisotropy: A near-perfect pentagonal bipyramidal dysprosium (III) single-molecule magnet. Angew. Chem. Int. Ed. 2016, 55, 16071–16074. [Google Scholar] [CrossRef]
  54. Li, Z.H.; Zhai, Y.Q.; Chen, W.P.; Ding, Y.S.; Zheng, Y.Z. Air-Stable Hexagonal Bipyramidal Dysprosium (III) Single-Ion Magnets with Nearly Perfect D6h Local Symmetry. Chem. Eur. J. 2019, 25, 16219–16224. [Google Scholar] [CrossRef] [PubMed]
  55. Gould, C.A.; McClain, K.R.; Reta, D.; Kragskow, J.G.; Marchiori, D.A.; Lachman, E.; Choi, E.S.; Analytis, J.G.; Britt, R.D.; Chilton, N.F.; et al. Ultrahard magnetism from mixed-valence dilanthanide complexes with metal-metal bonding. Science 2022, 375, 198–202. [Google Scholar] [CrossRef] [PubMed]
  56. Mavragani, N.; Errulat, D.; Gálico, D.A.; Kitos, A.A.; Mansikkamäki, A.; Murugesu, M. Radical-Bridged Ln4 Metallocene Complexes with Strong Magnetic Coupling and a Large Coercive Field. Angew. Chem. 2021, 133, 24408–24415. [Google Scholar] [CrossRef]
  57. Jin, P.B.; Zhai, Y.Q.; Yu, K.X.; Winpenny, R.E.; Zheng, Y.Z. Dysprosiacarboranes as Organometallic Single-Molecule Magnets. Angew. Chem. Int. Ed. 2020, 59, 9350–9354. [Google Scholar] [CrossRef]
  58. Aulakh, D.; Pyser, J.B.; Zhang, X.; Yakovenko, A.A.; Dunbar, K.R.; Wriedt, M. Metal–organic frameworks as platforms for the controlled nanostructuring of single-molecule magnets. J. Am. Chem. Soc. 2015, 137, 9254–9257. [Google Scholar] [CrossRef]
  59. Paschke, F.; Birk, T.; Enenkel, V.; Liu, F.; Romankov, V.; Dreiser, J.; Popov, A.A.; Fonin, M. Exceptionally High Blocking Temperature of 17 K in a Surface-Supported Molecular Magnet. Adv. Mater. 2021, 33, 2102844. [Google Scholar] [CrossRef] [PubMed]
  60. Spree, L.; Liu, F.; Neu, V.; Rosenkranz, M.; Velkos, G.; Wang, Y.; Schiemenz, S.; Dreiser, J.; Gargiani, P.; Valvidares, M.; et al. Robust single molecule magnet monolayers on graphene and graphite with magnetic hysteresis up to 28 K. Adv. Funct. Mater. 2021, 31, 2105516. [Google Scholar] [CrossRef]
  61. Boothroyd, A.T. Principles of Neutron Scattering from Condensed Matter; Oxford University Press: New York, NY, USA, 2020. [Google Scholar]
  62. Dunstan, M.A.; Mole, R.A.; Boskovic, C. Inelastic Neutron Scattering of Lanthanoid Complexes and Single-Molecule Magnets. Eur. J. Inorg. Chem. 2019, 2019, 1090–1105. [Google Scholar] [CrossRef]
  63. Weil, J.A.J.A. Electron Paramagnetic Resonance Elementary Theory and Practical Applications, 2nd ed.; Weil, J.A., Bolton, J.R., Eds.; Wiley-Interscience: Hoboken, NJ, USA, 2007. [Google Scholar]
  64. Atzori, M.; Tesi, L.; Morra, E.; Chiesa, M.; Sorace, L.; Sessoli, R. Room-temperature quantum coherence and rabi oscillations in vanadyl phthalocyanine: Toward multifunctional molecular spin qubits. J. Am. Chem. Soc. 2016, 138, 2154–2157. [Google Scholar] [CrossRef]
  65. Nolden, O.; Fleck, N.; Lorenzo, E.R.; Wasielewski, M.R.; Schiemann, O.; Gilch, P.; Richert, S. Excitation energy transfer and exchange-mediated quartet state formation in porphyrin-trityl systems. Chem.-Eur. J. 2021, 27, 2683–2691. [Google Scholar] [CrossRef]
  66. Bayliss, S.; Laorenza, D.; Mintun, P.; Kovos, B.; Freedman, D.; Awschalom, D. Optically addressable molecular spins for quantum information processing. Science 2020, 370, 1309–1312. [Google Scholar] [CrossRef] [PubMed]
  67. Cornia, A.; Mannini, M.; Sainctavit, P.; Sessoli, R. Chemical strategies and characterization tools for the organization of single molecule magnets on surfaces. Chem. Soc. Rev. 2011, 40, 3076–3091. [Google Scholar] [CrossRef] [PubMed]
  68. Mannini, M.; Pineider, F.; Sainctavit, P.; Joly, L.; Fraile-Rodríguez, A.; Arrio, M.A.; Cartier dit Moulin, C.; Wernsdorfer, W.; Cornia, A.; Gatteschi, D.; et al. X-ray magnetic circular dichroism picks out single-molecule magnets suitable for nanodevices. Adv. Mater. 2009, 21, 167–171. [Google Scholar] [CrossRef]
  69. Liu, C.M.; Sun, R.; Wang, B.W.; Wu, F.; Hao, X.; Shen, Z. Homochiral ferromagnetic coupling Dy2 single-molecule magnets with strong magneto-optical faraday effects at room temperature. Inorg. Chem. 2021, 60, 12039–12048. [Google Scholar] [CrossRef]
  70. Greber, T.; Seitsonen, A.P.; Hemmi, A.; Dreiser, J.; Stania, R.; Matsui, F.; Muntwiler, M.; Popov, A.; Westerström, R. Circular dichroism and angular deviation in x-ray absorption spectra of Dy2ScN@C80 single-molecule magnets on h-BN/Rh (111). Phys. Rev. Mater. 2019, 3, 014409. [Google Scholar] [CrossRef]
  71. Kfir, O.; Zayko, S.; Nolte, C.; Sivis, M.; Möller, M.; Hebler, B.; Arekapudi, S.S.P.K.; Steil, D.; Schäfer, S.; Albrecht, M.; et al. Nanoscale magnetic imaging using circularly polarized high-harmonic radiation. Sci. Adv. 2017, 3, eaao4641. [Google Scholar] [CrossRef]
  72. Rogez, G.; Donnio, B.; Terazzi, E.; Gallani, J.L.; Kappler, J.P.; Bucher, J.P.; Drillon, M. The quest for nanoscale magnets: The example of [Mn12] single molecule magnets. Adv. Mater. 2009, 21, 4323–4333. [Google Scholar] [CrossRef]
  73. Frauhammer, T.; Chen, H.; Balashov, T.; Derenbach, G.; Klyatskaya, S.; Moreno-Pineda, E.; Ruben, M.; Wulfhekel, W. Indirect spin-readout of rare-earth-based single-molecule magnet with scanning tunneling microscopy. Phys. Rev. Lett. 2021, 127, 123201. [Google Scholar] [CrossRef]
  74. Novikov, V.V.; Nelyubina, Y.V. Modern physical methods for the molecular design of single-molecule magnets. Russ. Chem. Rev. 2021, 90, 1330. [Google Scholar] [CrossRef]
  75. Zhu, Z.; Zhao, C.; Feng, T.; Liu, X.; Ying, X.; Li, X.L.; Zhang, Y.Q.; Tang, J. Air-stable chiral single-molecule magnets with record anisotropy barrier exceeding 1800 K. J. Am. Chem. Soc. 2021, 143, 10077–10082. [Google Scholar] [CrossRef]
  76. Rubín, J.; Arauzo, A.; Bartolomé, E.; Sedona, F.; Rancan, M.; Armelao, L.; Luzón, J.; Guidi, T.; Garlatti, E.; Wilhelm, F.; et al. Origin of the Unusual Ground-State Spin S=9 in a Cr10 Single-Molecule Magnet. J. Am. Chem. Soc. 2022, 144, 12520–12535. [Google Scholar] [CrossRef] [PubMed]
  77. Mossin, S.; Tran, B.L.; Adhikari, D.; Pink, M.; Heinemann, F.W.; Sutter, J.; Szilagyi, R.K.; Meyer, K.; Mindiola, D.J. A mononuclear Fe (III) single molecule magnet with a 3/2 ↔ 5/2 spin crossover. J. Am. Chem. Soc. 2012, 134, 13651–13661. [Google Scholar] [CrossRef] [PubMed]
  78. Marin, R.; Brunet, G.; Murugesu, M. Shining new light on multifunctional lanthanide single-molecule magnets. Angew. Chem. Int. Ed. 2021, 60, 1728–1746. [Google Scholar] [CrossRef]
  79. Evans, D. 400. The determination of the paramagnetic susceptibility of substances in solution by nuclear magnetic resonance. J. Chem. Soc. (Resumed) 1959, 2003–2005. [Google Scholar] [CrossRef]
  80. Carretta, P.; Lascialfari, A. NMR-MRI, μSR and Mössbauer Spectroscopies in Molecular Magnets; Springer Science & Business Media: Milan, Italy, 2007. [Google Scholar]
  81. Foguet-Albiol, D.; O’Brien, T.A.; Wernsdorfer, W.; Moulton, B.; Zaworotko, M.J.; Abboud, K.A.; Christou, G. DFT Computational Rationalization of an Unusual Spin Ground State in an Mn12 Single-Molecule Magnet with a Low-Symmetry Loop Structure. Angew. Chem. Int. Ed. 2005, 44, 897–901. [Google Scholar] [CrossRef]
  82. Ge, N.; Zhai, Y.Q.; Deng, Y.F.; Ding, Y.S.; Wu, T.; Wang, Z.X.; Ouyang, Z.; Nojiri, H.; Zheng, Y.Z. Rationalization of single-molecule magnet behavior in a three-coordinate Fe (III) complex with a high-spin state (S=5/2). Inorg. Chem. Front. 2018, 5, 2486–2492. [Google Scholar] [CrossRef]
  83. Chilton, N.F. Molecular Magnetism. Annu. Rev. Mater. Res. 2022, 52, 79–101. [Google Scholar] [CrossRef]
  84. Atanasov, M.; Ganyushin, D.; Sivalingam, K.; Neese, F. A modern first-principles view on ligand field theory through the eyes of correlated multireference wavefunctions. In Molecular Electronic Structures of Transition Metal Complexes II; Springer: Berlin/Heidelberg, Germany, 2012; pp. 149–220. [Google Scholar]
  85. Atanasov, M.; Aravena, D.; Suturina, E.; Bill, E.; Maganas, D.; Neese, F. First principles approach to the electronic structure, magnetic anisotropy and spin relaxation in mononuclear 3d-transition metal single molecule magnets. Coord. Chem. Rev. 2015, 289, 177–214. [Google Scholar] [CrossRef]
  86. Gu, L.; Wu, R. Origins of slow magnetic relaxation in single-molecule magnets. Phys. Rev. Lett. 2020, 125, 117203. [Google Scholar] [CrossRef]
  87. Ruiz, E.; Cirera, J.; Cano, J.; Alvarez, S.; Loose, C.; Kortus, J. Can large magnetic anisotropy and high spin really coexist? Chem. Commun. 2008, 1, 52–54. [Google Scholar] [CrossRef]
  88. Amsterdam Modeling Suite, SCM. Available online: http://www.scm.com (accessed on 4 June 2025).
  89. Rajaraman, G.; Ruiz, E.; Cano, J.; Alvarez, S. Theoretical determination of the exchange coupling constants of a single-molecule magnet Fe10 complex. Chem. Phys. Lett. 2005, 415, 6–9. [Google Scholar] [CrossRef]
  90. Cauchy, T.; Ruiz, E.; Alvarez, S. Exchange interactions in a Fe5 complex: A theoretical study using density functional theory. Inorganica Chim. Acta 2008, 361, 3832–3835. [Google Scholar] [CrossRef]
  91. Cauchy, T.; Ruiz, E.; Alvarez, S. Exchange coupling interactions in a Fe6 complex: A theoretical study using density functional theory. Phys. Condens. Matter 2006, 384, 116–119. [Google Scholar] [CrossRef]
  92. Ruiz, E.; Cano, J.; Alvarez, S.; Gouzerh, P.; Verdaguer, M. Theoretical study of the exchange coupling interactions in a polyoxometalate Fe9W12 complex. Polyhedron 2007, 26, 2161–2164. [Google Scholar] [CrossRef]
  93. Vignesh, K.R.; Martin, R.B.; Miller, G.; Rajaraman, G.; Murray, K.S.; Langley, S.K. {MnIII2LnIII2}(Ln= Gd, La or Y) butterfly complexes: Ferromagnetic exchange observed between bis-μ-alkoxo bridged manganese (III) ions. Polyhedron 2019, 170, 508–514. [Google Scholar] [CrossRef]
  94. Bellini, V.; Olivieri, A.; Manghi, F. Density-functional study of the Cr 8 antiferromagnetic ring. Phys. Rev. 2006, 73, 184431. [Google Scholar] [CrossRef]
  95. Vignesh, K.R.; Langley, S.K.; Murray, K.S.; Rajaraman, G. Quenching the Quantum Tunneling of Magnetization in Heterometallic Octanuclear {TMIII4DyIII4}(TM=Co and Cr) Single-Molecule Magnets by Modification of the Bridging Ligands and Enhancing the Magnetic Exchange Coupling. Chem.-Eur. J. 2017, 23, 1654–1666. [Google Scholar] [CrossRef]
  96. Singh, S.K.; Beg, M.F.; Rajaraman, G. Role of Magnetic Exchange Interactions in the Magnetization Relaxation of {3d–4f} Single-Molecule Magnets: A Theoretical Perspective. Chem. Eur. J. 2016, 22, 672–680. [Google Scholar] [CrossRef]
  97. Cano, J.; Costa, R.; Alvarez, S.; Ruiz, E. Theoretical study of the magnetic properties of an Mn12 single-molecule magnet with a loop structure: The role of the next-nearest neighbor interactions. J. Chem. Theory Comput. 2007, 3, 782–788. [Google Scholar] [CrossRef]
  98. Gomez-Coca, S.; Ruiz, E.; Kortus, J. Single-molecule magnet Fe 9 supramolecular dimers: A theoretical approach to intramolecular and intermolecular exchange interactions. Chem. Commun. 2009, 29, 4363–4365. [Google Scholar] [CrossRef]
  99. Moseley, D.H.; Stavretis, S.E.; Thirunavukkuarasu, K.; Ozerov, M.; Cheng, Y.; Daemen, L.L.; Ludwig, J.; Lu, Z.; Smirnov, D.; Brown, C.M.; et al. Spin–phonon couplings in transition metal complexes with slow magnetic relaxation. Nat. Commun. 2018, 9, 2572. [Google Scholar] [CrossRef] [PubMed]
  100. Srnec, M.; Chalupský, J.; Fojta, M.; Zendlová, L.; Havran, L.; Hocek, M.; Kyvala, M.; Rulíšek, L. Effect of Spin-Orbit Coupling on Reduction Potentials of Octahedral Ruthenium (II/III) and Osmium (II/III) Complexes. J. Am. Chem. Soc. 2008, 130, 10947–10954. [Google Scholar] [CrossRef] [PubMed]
  101. Atanasov, M.; Neese, F. Computational studies on vibronic coupling in single molecule magnets: Impact on the mechanism of magnetic relaxation. Proc. J. Physics. Conf. Ser. 2018, 1148, 012006. [Google Scholar] [CrossRef]
  102. Gómez-Coca, S.; Urtizberea, A.; Cremades, E.; Alonso, P.J.; Camón, A.; Ruiz, E.; Luis, F. Origin of slow magnetic relaxation in Kramers ions with non-uniaxial anisotropy. Nat. Commun. 2014, 5, 4300. [Google Scholar] [CrossRef]
  103. Castro-Alvarez, A.; Gil, Y.; Llanos, L.; Aravena, D. High performance single-molecule magnets, Orbach or Raman relaxation suppression? Inorg. Chem. Front. 2020, 7, 2478–2486. [Google Scholar] [CrossRef]
  104. Briganti, M.; Santanni, F.; Tesi, L.; Totti, F.; Sessoli, R.; Lunghi, A. A complete ab initio view of Orbach and Raman spin–lattice relaxation in a dysprosium coordination compound. J. Am. Chem. Soc. 2021, 143, 13633–13645. [Google Scholar] [CrossRef]
  105. Reta, D.; Kragskow, J.G.; Chilton, N.F. Ab initio prediction of high-temperature magnetic relaxation rates in single-molecule magnets. J. Am. Chem. Soc. 2021, 143, 5943–5950. [Google Scholar] [CrossRef]
  106. Lunghi, A.; Totti, F.; Sessoli, R.; Sanvito, S. The role of anharmonic phonons in under-barrier spin relaxation of single molecule magnets. Nat. Commun. 2017, 8, 14620. [Google Scholar] [CrossRef]
  107. Barone, V.; Alessandrini, S.; Biczysko, M.; Cheeseman, J.R.; Clary, D.C.; McCoy, A.B.; DiRisio, R.J.; Neese, F.; Melosso, M.; Puzzarini, C. Computational molecular spectroscopy. Nat. Rev. Methods Prim. 2021, 1, 38. [Google Scholar] [CrossRef]
  108. Parr, R.G.; Yang, W. Density-Functional Theory of Atoms and Molecules (International Series of Monographs on Chemistry); Oxford University Press: New York, NY, USA, 1994. [Google Scholar]
  109. Szabo, A.; Ostlund, N.S. Modern Quantum Chemistry: Introduction to Advanced Electronic Structure Theory, 1st ed.; Dover Publications, Inc.: Mineola, NY, USA, 1996. [Google Scholar]
  110. Cohen, A.J.; Mori-Sánchez, P.; Yang, W. Challenges for density functional theory. Chem. Rev. 2012, 112, 289–320. [Google Scholar] [CrossRef]
  111. Verma, P.; Truhlar, D.G. Status and challenges of density functional theory. Trends Chem. 2020, 2, 302–318. [Google Scholar] [CrossRef]
  112. Himmetoglu, B.; Floris, A.; De Gironcoli, S.; Cococcioni, M. Hubbard-corrected DFT energy functionals: The LDA+ U description of correlated systems. Int. J. Quantum Chem. 2014, 114, 14–49. [Google Scholar] [CrossRef]
  113. Janesko, B.G. Replacing hybrid density functional theory: Motivation and recent advances. Chem. Soc. Rev. 2021, 50, 8470–8495. [Google Scholar] [CrossRef] [PubMed]
  114. Olsen, J. The CASSCF method: A perspective and commentary. Int. J. Quantum Chem. 2011, 111, 3267–3272. [Google Scholar] [CrossRef]
  115. García, V.; Castell, O.; Caballol, R.; Malrieu, J. An iterative difference-dedicated configuration interaction. Proposal and test studies. Chem. Phys. Lett. 1995, 238, 222–229. [Google Scholar] [CrossRef]
  116. David, G.; Ferré, N.; Le Guennic, B. Consistent Evaluation of Magnetic Exchange Couplings in Multicenter Compounds in KS-DFT: The Recomposition Method. J. Chem. Theory Comput. 2022, 19, 157–173. [Google Scholar] [CrossRef]
  117. Gupta, T.; Rajaraman, G. Modelling spin Hamiltonian parameters of molecular nanomagnets. Chem. Commun. 2016, 52, 8972–9008. [Google Scholar] [CrossRef]
  118. Swain, A.; Sarkar, A.; Rajaraman, G. Role of Ab Initio Calculations in the Design and Development of Organometallic Lanthanide-Based Single-Molecule Magnets. Chem.-Asian J. 2019, 14, 4056–4073. [Google Scholar] [CrossRef]
  119. Wu, X.; Li, J.F.; Yin, B. The interpretation and prediction of lanthanide single-ion magnets from ab initio electronic structure calculation: The capability and limit. Dalton Trans. 2022, 51, 14793–14816. [Google Scholar] [CrossRef]
  120. Cornia, A.; Mannini, M. Single-molecule magnets on surfaces. In Molecular Nanomagnets and Related Phenomena; Springer: Berlin/Heidelberg, Germany, 2015; pp. 293–330. [Google Scholar]
  121. Liu, J.; Li, J.; Xu, Z.; Zhou, X.; Xue, Q.; Wu, T.; Zhong, M.; Li, R.; Sun, R.; Shen, Z.; et al. On-surface preparation of coordinated lanthanide-transition-metal clusters. Nat. Commun. 2021, 12, 1619. [Google Scholar] [CrossRef]
  122. Breuer, H.P.; Petruccione, F. The Theory of Open Quantum Systems; Oxford University Press: New York, NY, USA, 2007. [Google Scholar]
  123. Frisch, M. gaussian09. 2009. Available online: http://www.gaussian.com (accessed on 4 June 2025).
  124. Artacho, E.; Anglada, E.; Diéguez, O.; Gale, J.D.; García, A.; Junquera, J.; Martin, R.M.; Ordejón, P.; Pruneda, J.M.; Sánchez-Portal, D.; et al. The SIESTA method; developments and applicability. J. Phys. Condens. Matter 2008, 20, 064208. [Google Scholar] [CrossRef] [PubMed]
  125. Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The ORCA quantum chemistry program package. J. Chem. Phys. 2020, 152, 224108. [Google Scholar] [CrossRef]
  126. Neese, F. Software update: The ORCA program system—Version 5.0. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2022, 12, e1606. [Google Scholar] [CrossRef]
  127. Aquilante, F.; De Vico, L.; Ferré, N.; Ghigo, G.; Malmqvist, P.å.; Neogrády, P.; Pedersen, T.B.; Pitoňák, M.; Reiher, M.; Roos, B.O.; et al. MOLCAS 7: The next generation. J. Comput. Chem. 2010, 31, 224–247. [Google Scholar] [CrossRef]
  128. Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. cp2k: Atomistic simulations of condensed matter systems. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2014, 4, 15–25. [Google Scholar] [CrossRef]
  129. Blaha, P.; Schwarz, K.; Madsen, G.K.H.; Kvasnicka, D.; Luitz, J.; Laskowski, R.; Tran, F.; Marks, L.D. WIEN2k: An Augmented Plane Wave plus Local Orbitals Program for Calculating Crystal Properties; Vienna University of Technology: Vienna, Austria, 2018. [Google Scholar]
  130. Balasubramani, S.G.; Chen, G.P.; Coriani, S.; Diedenhofen, M.; Frank, M.S.; Franzke, Y.J.; Furche, F.; Grotjahn, R.; Harding, M.E.; Hättig, C.; et al. TURBOMOLE: Modular program suite for ab initio quantum-chemical and condensed-matter simulations. J. Chem. Phys. 2020, 152, 184107. [Google Scholar] [CrossRef] [PubMed]
  131. Troiani, F.; Ghirri, A.; Paris, M.; Bonizzoni, C.; Affronte, M. Towards quantum sensing with molecular spins. J. Magn. Magn. Mater. 2019, 491, 165534. [Google Scholar] [CrossRef]
  132. Laorenza, D.W.; Freedman, D.E. Could the Quantum Internet Be Comprised of Molecular Spins with Tunable Optical Interfaces? J. Am. Chem. Soc. 2022, 144, 21810–21825. [Google Scholar] [CrossRef]
  133. Lunghi, A.; Sanvito, S. Computational design of magnetic molecules and their environment using quantum chemistry, machine learning and multiscale simulations. Nat. Rev. Chem. 2022, 6, 761–781. [Google Scholar] [CrossRef]
Figure 1. Typical examples of SMMs. (a) Mn 12 [1], (b) polynuclear lanthanide SMM Dy 4 [3], (c) 3 d 4 f SMM [4], and (d) lanthanide SIM [5]. Here, carbon is in gray, nitrogen in blue, oxygen in red, manganese in purple, bromine in dark red, dysprosium in green, and cobalt in dark pink.
Figure 1. Typical examples of SMMs. (a) Mn 12 [1], (b) polynuclear lanthanide SMM Dy 4 [3], (c) 3 d 4 f SMM [4], and (d) lanthanide SIM [5]. Here, carbon is in gray, nitrogen in blue, oxygen in red, manganese in purple, bromine in dark red, dysprosium in green, and cobalt in dark pink.
Molecules 30 02522 g001
Figure 2. A summary of important scientific and application potentials for SMMs.
Figure 2. A summary of important scientific and application potentials for SMMs.
Molecules 30 02522 g002
Figure 3. A summary (log–log plot) for the effective spin conversion barrier ( U eff ) and blocking temperature ( T B ) for typical SMMs up to now. We have simplified the compound names to save space in the figure. Here, 1 = Mn 12 , 2 = Mn 84 , 3 = Mn 6 , 4 = Fe 8 , 5 = Mn II Mo 6 III , 6 = Mn 2 II , 7 = TbPc 2 , 8 = { Cu 2 Tb 2 } * , 9 = { Cr 2 Dy 2 , 10 = { Cu 2 Tb 2 } * * , 11 = [ Dy 4 K 2 O ( OtBu ) 12 , 12 = [ Dy ( μ - OH ) ( DBP ) 2 ( THF ) 2 , 13 = { [ ( Me 3 Si ) 2 N ] 2 ( THF ) } 2 ( μ - η 2 : η 2 N 2 , 14 = [ ( Cp Me 4 H 2 Ln ) 2 ( μ - N 2 . ] , 15 = ( Cp * ) Ln ( COT ) ,   16 = [ Er ( COT ) 2 ] , 17 = [ Dy ( Cpttt ) 2 , 18 = [ Dy ( Cp iPr 5 ) ( Cp * ) ] + , 19 = [ DyZn 2 ( TTTT Br ) 2 / n ( MeOH ) ] + , 20 = [ Dy ( bbpen ) Br ] , 21 = [ Dy ( OPCy 3 ) 2 ( H 2 O ) 4 ] 3 + , 22 = [ Dy ( OP t Bu ( NHiPr 2 ) 2 ( H 2 O ) 5 ] 3 + , 23 = [ Dy ( O t Bu ) 2 ( py ) 5 ] + , 24 = [ Dy ( L E ) ( 4 - MeO - PhO ) 2 , 25 = [ UO 2 ( salen ) ] 2 Mn ( Py ) 3 } 6 , 26 = [ Fe ( C ( SiMe 3 ) 3 ) 2 ] , 27 = [ ( sIPr ) CoNDmp ] , 28 = Co ( C ( SiMe 2 ONaph ) 3 ) 2 , 29 = ( Cp iPr 5 )2 Dy 2 I 3 , and 30 = ( Cp iPr 5 )2 Tb 2 I 3 . Notice that [ Dy 4 K 2 O ( OtBu ) 12 (11) (in purple) and Co ( C ( SiMe 2 ONaph ) 3 ) 2 (28) (in brown) have similar parameters, so they overlap each other. We have also added a linearly fitted line to show the trend for the optimization of the SMMs in the previous work. This trending line points to the further optimization expected in the future, going towards a higher blocking temperature and a larger effective energy barrier.
Figure 3. A summary (log–log plot) for the effective spin conversion barrier ( U eff ) and blocking temperature ( T B ) for typical SMMs up to now. We have simplified the compound names to save space in the figure. Here, 1 = Mn 12 , 2 = Mn 84 , 3 = Mn 6 , 4 = Fe 8 , 5 = Mn II Mo 6 III , 6 = Mn 2 II , 7 = TbPc 2 , 8 = { Cu 2 Tb 2 } * , 9 = { Cr 2 Dy 2 , 10 = { Cu 2 Tb 2 } * * , 11 = [ Dy 4 K 2 O ( OtBu ) 12 , 12 = [ Dy ( μ - OH ) ( DBP ) 2 ( THF ) 2 , 13 = { [ ( Me 3 Si ) 2 N ] 2 ( THF ) } 2 ( μ - η 2 : η 2 N 2 , 14 = [ ( Cp Me 4 H 2 Ln ) 2 ( μ - N 2 . ] , 15 = ( Cp * ) Ln ( COT ) ,   16 = [ Er ( COT ) 2 ] , 17 = [ Dy ( Cpttt ) 2 , 18 = [ Dy ( Cp iPr 5 ) ( Cp * ) ] + , 19 = [ DyZn 2 ( TTTT Br ) 2 / n ( MeOH ) ] + , 20 = [ Dy ( bbpen ) Br ] , 21 = [ Dy ( OPCy 3 ) 2 ( H 2 O ) 4 ] 3 + , 22 = [ Dy ( OP t Bu ( NHiPr 2 ) 2 ( H 2 O ) 5 ] 3 + , 23 = [ Dy ( O t Bu ) 2 ( py ) 5 ] + , 24 = [ Dy ( L E ) ( 4 - MeO - PhO ) 2 , 25 = [ UO 2 ( salen ) ] 2 Mn ( Py ) 3 } 6 , 26 = [ Fe ( C ( SiMe 3 ) 3 ) 2 ] , 27 = [ ( sIPr ) CoNDmp ] , 28 = Co ( C ( SiMe 2 ONaph ) 3 ) 2 , 29 = ( Cp iPr 5 )2 Dy 2 I 3 , and 30 = ( Cp iPr 5 )2 Tb 2 I 3 . Notice that [ Dy 4 K 2 O ( OtBu ) 12 (11) (in purple) and Co ( C ( SiMe 2 ONaph ) 3 ) 2 (28) (in brown) have similar parameters, so they overlap each other. We have also added a linearly fitted line to show the trend for the optimization of the SMMs in the previous work. This trending line points to the further optimization expected in the future, going towards a higher blocking temperature and a larger effective energy barrier.
Molecules 30 02522 g003
Figure 4. A summary of the experimental characterization methods for SMMs. SQUID = superconducting quantum interference device; NS = neutron scattering; EPR = electron paramagnetic resonance; XMCD = X-ray magnetic circular dichroism; SP-STM = spin-polarized scanning tunneling microscopy.
Figure 4. A summary of the experimental characterization methods for SMMs. SQUID = superconducting quantum interference device; NS = neutron scattering; EPR = electron paramagnetic resonance; XMCD = X-ray magnetic circular dichroism; SP-STM = spin-polarized scanning tunneling microscopy.
Molecules 30 02522 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wu, W.; Huang, T.; Zhu, J.; Zou, T.; Wang, H. Exploring Single-Molecular Magnets for Quantum Technologies. Molecules 2025, 30, 2522. https://doi.org/10.3390/molecules30122522

AMA Style

Wu W, Huang T, Zhu J, Zou T, Wang H. Exploring Single-Molecular Magnets for Quantum Technologies. Molecules. 2025; 30(12):2522. https://doi.org/10.3390/molecules30122522

Chicago/Turabian Style

Wu, Wei, Tianhong Huang, Jianhua Zhu, Taoyu Zou, and Hai Wang. 2025. "Exploring Single-Molecular Magnets for Quantum Technologies" Molecules 30, no. 12: 2522. https://doi.org/10.3390/molecules30122522

APA Style

Wu, W., Huang, T., Zhu, J., Zou, T., & Wang, H. (2025). Exploring Single-Molecular Magnets for Quantum Technologies. Molecules, 30(12), 2522. https://doi.org/10.3390/molecules30122522

Article Metrics

Back to TopTop