Abstract
Magnetic relaxation in coordination compounds is largely dominated by the interaction of the spin with phonons. Although a comprehensive understanding of spin-phonon relaxation has been achieved for mononuclear complexes, only a qualitative picture is available for polynuclear compounds. Large zero-field splitting and exchange coupling values have been empirically found to strongly suppress spin relaxation and have been used as the main guideline for designing molecular compounds with long spin lifetime, also known as single-molecule magnets, but no microscopic rationale for these observations is available. Here we fill this critical knowledge gap by providing a full first-principles description of spin-phonon relaxation in an air-stable Co(II) dimer with both large single-ion anisotropy and exchange coupling. Simulations reproduce the experimental relaxation data with excellent accuracy and provide a microscopic understanding of Orbach and Raman relaxation pathways and their dependency on exchange coupling, zero-field splitting, and molecular vibrations. Theory and numerical simulations show that increasing cluster nuclearity to just four cobalt units would lead to a complete suppression of low-temperature Raman relaxation. These results hold a general validity for polynuclear single-molecule magnets, providing a deeper understanding of their relaxation and revised strategies for their improvement.
Short abstract
Spin relaxation in polynuclear coordination complexes is unraveled through ab initio open quantum system simulations, revealing new strategies to fully suppress Raman relaxation at low temperature.
Introduction
Coordination complexes with large magnetic anisotropies have been shown to display slow relaxation of their magnetic moment. Compounds of this class have been named single-molecule magnets (SMMs), in recognition of their similarities with hard magnets,1 and have attracted a large interest for potential applications such as nanomagnetism,2 spintronics3 and more recently quantum information science.4 At the same time, the chemical flexibility of SMMs has enabled an unprecedentedly detailed study of spin relaxation, a fundamental physical process key to magnetic resonance5 and magnetism6 at large, making them chemical systems of central importance across chemistry, physics, and materials science.
One of the main priorities for advancing the application of SMMs is mitigating the dramatic increase of the spin relaxation rate with temperature, T. The spin dynamics of a molecular spin is deeply affected by its interaction with the surrounding lattice, namely the spin-phonon coupling, which ultimately leads to a state with a zero expectation value of the magnetization.7 At high temperatures, the spin relaxation time, τ, follows the Arrhenius law
![]() |
1 |
where the effective magnetization reversal barrier Ueff can be taken as a measure of magnetic anisotropy.8 Very early on, it was recognized that in the absence of under-barrier processes, the effective energy barrier is Ueff = |D|S2, where D is the molecular axial zero-field splitting and S is the ground-state total spin value. For negative values of D, the maximum values of the spin projections, MS = ±S, are the lowest in energy, and up to 2S sequential phonon quanta have to be absorbed/emitted to go from MS = S → MS = −S and reverse the spin orientation. As the factor |D|S2 increases in value, phonons with high energy (and thus low thermal population) are needed to initiate this process, leading to a slower relaxation, as expressed by eq 1. This relaxation mechanism falls under the name of Orbach and maximizing D has proved to be the best strategy to reduce its detrimental effects on the spin.9,10 To this end, much effort has been devoted to the design of mononuclear coordination compounds and outstanding results have been achieved, e.g. Ueff ∼ 450 cm–111 and 1500 cm–112 in linear Co(II) and Dy(III) mononuclear complexes, respectively.
However, minimizing the effect of Orbach relaxation is not the sole challenge in the field, and other spin relaxation mechanisms are operative. For instance, Quantum Tunneling of the Magnetization (QTM) operates at low temperatures and is primarily responsible for the closure of the magnetic hysteresis curve at zero magnetic field. Exchange coupling, either of inter-13 or intramolecular14−17 origin, was recognized to significantly affect QTM and its study has been at the center stage of the field since the very beginning.
However, the maximization of both zero-field splitting and exchange coupling, in the attempt to quench both Orbach and QTM relaxation, has proved challenging.18 On the one hand, as the field moved to SMMs with large single-ion zero-field splittings, the phenomenology of spin-phonon relaxation has become much more complex and Raman and direct phonon-mediated relaxation mechanisms have started to play a role at low temperature and high magnetic field, respectively.19,20 Moreover, while early SMMs, like Mn12,1 and Fe4,21 were mostly based on ions with low zero-field splitting and large exchange coupling values, most compounds with large zero-field splitting fall in a weak or intermediate exchange coupling regime.22 Molecules of this class present a multitude of low-energy spin states available for relaxation, making the interpretation of experiments far from trivial and unequivocal.20,23
In recent years, the combination of ab initio computational methods and the theory of open quantum systems has made it possible to simulate spin-phonon relaxation in magnetic molecules,7,24 opening a window on the microscopic details of this process. Ab initio simulations of spin relaxation have played a key role in resolving conflictual interpretations of spin relaxation in mononuclear compounds, particularly in the interpretation of Raman relaxation,25,26 and resolving a large number of misinterpretations of experiments.27 A first-principles description of spin-phonon coupling and relaxation in SMMs has not only aided the interpretation of the experiments27,28 but has also provided insights into the nature of spin-phonon coupling and its relation to molecular structure.29−31 However, ab initio studies of spin relaxation have so far been exclusively performed for mononuclear compounds, and the study of exchange-coupled clusters has been limited to the computation of static magnetic properties32−38 As such our understanding of spin relaxation in polynuclear SMMs is still largely phenomenological. In order for the field of exchange-coupled SMMs to progress it is now imperative to achieve a complete understanding of their spin relaxation and address the urgent questions of the nature of spin-phonon coupling and relaxation mechanisms in these compounds.
In this contribution, we use ab initio open quantum systems theory to address spin relaxation in polynuclear clusters and complete the microscopic interpretation of Orbach and Raman spin-phonon relaxation in SMMs. For this purpose, we use the radical-bridged Co(II) dimer [K(18-crown-6]2][K(H2O)4][Co(bmsab)]2(μ-tmsab)] (Co2Rad), where bmsab is the dianion of 1,2-bis(methanesulfonamide) benzene and tmsab is the radical-trianion of 1,2,4,5-tetrakis(methanesulfonamide)benzene. This recently reported compound39 exhibits some of the largest Ueff and slowest relaxation among transition metal complexes. Importantly, unlike most Ln-based SMMs, this compound is air-stable, potentially representing an optimal building block for future technologies based on SMMs. Last but not least, the electronic structure of this compound has recently been exhaustively characterized,39 and spin relaxation in its mononuclear building block20 has been successfully explained with ab initio methods,40 thus offering an ideal starting point for our study.
Figure 1.
Co2Rad molecular structure. The three-dimensional structure of the anion Co2Rad is reported with the color code: Co in brown, H in white, C in green, S in yellow, O in red, and N in blue.
Methods
Geometry Optimization and Phonons
Cell and geometry optimization and simulations of Γ-point phonons are performed with periodic Density Functional Theory (DFT) using the CP2K software.41 Cell optimization are performed with a very tight force convergence criterion of 10–7 a.u. and SCF convergence criteria of 10–10 a.u. for the energy. A plane wave cutoff of 1000 Ry, DZVP-MOLOPT Gaussian basis sets, and Goedecker-Tetter-Hutter Pseudopotentials42 are employed for all atoms. The Perdew–Burke–Ernzerhof (PBE) functional and DFT-D3 dispersion corrections are used.43,44 Phonons are computed with a two-step numerical differentiation of atomic forces using a step of 0.01 Å.
Electronic Excitations and Spin Hamiltonian
The electronic structure of the dinuclear Co(II) complex is computed with complete active space self-consistent field (CASSCF) and multistate CAS perturbation theory to second order (MS-CASPT2) using the Molpro progam package.45,46 Computations use a CAS(19,14) (19 electrons distributed in 14 orbitals) as reported in the SI. The molecular orbitals are represented by def2-TZVPP basis sets47 for Co and N, and def2-SVP basis sets47 otherwise, along with the appropriate auxiliary basis sets for density fitting.47 The orbitals are optimized with 4 octet states, and 8 sextet, 8 quartet, and 8 doublet states included in the state-averaging and the combined first-order/second-order algorithm of Kreplin and co-workers is employed.48 This choice of states is suggested by recognizing that Co(II) ions with approximate D2d local symmetry only have two pairs of low-lying quartet states which couple to the doublet of the radical bridge. Tests with larger numbers of roots confirm that additional states would fall at least 4000 cm–1 higher in energy than the states considered, see SI. For the MS-CASPT2 computations, the local pair-natural-orbital (PNO) based implementation of Werner and co-workers is used,49,50 along with a shift of 0.45 Eh. The detailed input parameters are given in the SI. For the lowest states of each multiplet, the multistate corrections are negligible and the runs for determining the numerical derivatives of the exchange coupling are done without multistate corrections.
The g and zero-field splitting tensors of the Co(II) centers are determined in separate computations, where one center is diamagnetically substituted with Zn(II) and the bridging ligand is considered in its reduced diamagnetic state. Here, a CAS(7,5) is employed, which comprises the 3d orbitals of Co(II). Configuration-averaged Hartree–Fock orbitals (a state-averaging over all possible states weighted by their multiplicity)51 is used and all possible CAS configuration interaction (CASCI) states are determined (10 quartet and 40 doublet states). The spin–orbit Hamiltonian is set up for these CASCI state and augmented with the CASPT2 correlation energies and off-diagonal MS-CASPT2 coupling matrix elements. The Breit-Pauli spin–orbit Hamiltonian and the one-center approximation are used. The pseudospin formalism is employed to extract the g and zero-field splitting tensors.52
Spin-Phonon Coupling
Spin-phonon coupling coefficients are computed as
![]() |
2 |
where Qα is the dimensionless displacement vector associated with the phonon
mode α and N is the number of atoms in the
unit cell. Liα and
ωα are the corresponding eigenvectors of the
Hessian and the angular frequency. Only Γ-point phonons are
considered. The first-order derivatives of the spin Hamiltonian with
respect to the Cartesian coordinates Xi, , are computed by numerical differentiation.
Each molecular degree of freedom is sampled four times between ±0.1
Å. A sample of the profiles of D and J along some Cartesian degrees of freedom are provided in SI.
Spin-Phonon Relaxation
Once the eigenstates, |a⟩, and eigenvalues, Ea, of the spin Hamiltonian, , have been obtained, the spin dynamics
can be simulated by computing the transition rate among different
spin states, Wba. Spin
relaxation in Kramers systems with large magnetic anisotropy consists
of contributions from one- and two-phonon processes. Considering one-phonon
processes, the transition rate,
, between spin states reads53,54
![]() |
3 |
where ℏωba = Eb – Ea. The function G1–ph reads
![]() |
4 |
where is the Bose–Einstein distribution
accounting for the thermal population of phonons, kB is the Boltzmann constant, and the Dirac delta functions,
e.g. δ(ω – ωα), enforce
energy conservation during the absorption and emission of phonons
by the spin system, respectively. eq 3 accounts for the Orbach relaxation mechanism,7 where a series of phonon absorption processes
leads the spin from the fully polarized state Ms = S to an excited state
with an intermediate value of Ms before the spin can emit phonons back to Ms = −S.
Two-phonon transitions, W2–phba, are responsible for Raman relaxation and include the absorption of two phonons, emission of two phonons or absorption of one phonon and emission of a second one. The latter process is the one that determines the Raman relaxation rate at low temperatures and is modeled as24
![]() |
5 |
where the terms
![]() |
6 |
involve the contribution of all the spin states |c⟩ at the same time, often referred to as a virtual state. The function G2–ph fulfils a similar role as G1–ph for one-phonon processes and it includes contributions from the Bose–Einstein distribution and imposes energy conservation. For the absorption/emission of two phonons, G2–ph reads
![]() |
7 |
All two-phonon contributions to are included and their full equations are
reported elsewhere.24 All the parameters
appearing in eqs 3 and 5 are computed from first principles. The Dirac delta
functions appearing in eq 4 and 7 are replaced by Gaussians with a smearing
of 15 cm–1. As discussed elsewhere, this substitution
is a good approximation for a vanishing smearing and a full sampling
of the phonons’ Brillouin zone and corresponds to treating
the bath as harmonic.24,25,40 Tests on the consistency of predictions of the spin relaxation time
τ with respect to the Gaussian smearing are reported in SI.
The simulation of Kramers systems in zero external fields requires the use of the nondiagonal secular approximation,24 where population and coherence terms of the density matrix are not independent of one another. This is achieved by simulating the dynamics of the entire density matrix for one-phonon processes. The full expression of eq 3 is reported in the literature.24,54 On the other hand, an equation that accounts for the dynamics of the entire density matrix under the effect of two-phonon processes resulting from fourth-order time-dependent perturbation theory is not yet available. However, it is possible to remove the coupling between population and coherence terms by orienting the molecular easy axis along the quantization z-axis and by applying a small magnetic field to break Kramers degeneracy.24 Importantly, a magnetic field along the molecular easy axis has no effect on the spin-phonon relaxation rate, as noted previously.24 Here we employ the latter strategy to simulate Raman relaxation.
Once all the matrix elements Wn–phba have been computed, τ–1 can be predicted by simply diagonalizing Wn–phba and taking the eigenvalue corresponding to an eigenvector describing a population transfer between the states of the ground-state KD. This usually corresponds to the smallest nonzero eigenvalue. The software MolForge,24 freely available at github.com/LunghiGroup/MolForge, is used for all these simulations.
Results
Electronic Structure and Magnetic Interactions
The electronic structure of Co2Rad is determined with multireference computational methods as described in Methods. The spin-free energy levels obtained from MS-CASPT2 calculations are reported in Table I and fitted to a Heisenberg Hamiltonian of the form
![]() |
8 |
where ,
, and
are the pseudospin operators of the two
Co(II) centers and the radical bridge, respectively. Due to the inversion
symmetry of Co2Rad, the exchange coupling, J, is identical for the
two cobalt ions.
Table I. Spin Ladder of Co2Rada.
ΔE/cm–1 | S | assignment | Jeff/cm–1 |
---|---|---|---|
0.0 | 5/2 | 0 J | |
198.1 | 3/2 | 1/2 J | 396.3 |
394.5 | 1/2 | J | 394.5 |
783.9 | 5/2 | ||
783.7 | 5/2 | ||
919.2 | 1/2 | 2 J | 459.6 |
973.8 | 3/2 | ||
974.1 | 3/2 | ||
1119.1 | 3/2 | 5/2 J | 447.6 |
1165.6 | 1/2 | ||
1167.1 | 1/2 | ||
1319.9 | 5/2 | 3 J | 440.0 |
1383.6 | 7/2 | 7/2 J | 395.3 |
In accordance with previous findings,39 the present computations confirm a strong antiferromagnetic coupling between the Co(II) centers and the radical bridge, with J values in the range 395 to 440 cm–1 and thus leading to a ground-state with total spin S = 5/2. As the lowest multiplets usually dominate the magnetic properties and spin dynamics at low temperatures, we set J = 394.5 cm–1, corresponding to the value for the spin Hamiltonian model from the second lowest energy gap, between the quartet and the doublet. Note that the inclusion of dynamic correlation is vital for accurate exchange couplings, the J value extracted from CASCI states would only be 169 cm–1. The solution of the spin Hamiltonian with J = 396.3 cm–1 gives a spin ladder with total spin S = 5/2, 3/2, 1/2, 1/2, 3/2, 5/2, 7/2, which fits the states up to energy ∼400 cm–1. At higher energies, states originate from excited quartet states of the Co(II) centers and give rise to further spin ladders,39 as indicated in Tab. I. Including a direct exchange term between the two cobalt ions does improve the fit to the computed energy levels, see SI. We also did not consider anisotropic exchange contributions, which result from spin–orbit and spin–spin contributions and are expected to be in the 1 cm–1 range.55−57 Given the very large magnitude of the leading cobalt-radical isotropic coupling, we confine the exchange terms in the present study to those given in eq 8.
The effect of spin–orbit coupling is included through the zero-field splitting Hamiltonian
![]() |
9 |
where D represents the zero-field splitting tensor of the two Co(II) ions. As described in Methods, the projection to a pseudospin Hamiltonian for a single quartet state results in a highly axial zero-field splitting tensor with D = −114.1 cm–1 and E = 0.9 cm–1. The main anisotropy axis is directed along the long axis of the molecule, connecting the two Co(II) centers. A previous detailed spectroscopy study39 resulted in J = 390 cm–1 and D = −113 cm–1 (with E close to zero), confirming the accuracy of the present parametrization for J and D. Finally, in order to simulate the effect of a static magnetic field we also compute the Zeeman Hamiltonian
![]() |
10 |
The values of the g-tensors, g, are also extracted from ab initio simulations, while gb is set to the free-electron value. The total spin Hamiltonian is the sum of the three terms in eqs 8-10 and the numerical values of all its tensors are provided in SI. The eigenvalues of the total spin Hamiltonian are reported in the SI, where it can be seen that the system features a well-separated sextet with considerable zero-field splitting as the lowest state and a maximum MS quantum number for the ground-state Kramers doublet (KD).
Spin-Phonon Relaxation
The effect of phonons on spin dynamics can be treated within the framework of perturbation theory. In particular, the linear coupling terms of eq 2 are known to influence the dynamics of spin at both the second and fourth order of perturbation theory.7Figure 2 reports the comparison between simulated and experimental spin relaxation times.39 In the absence of an external magnetic field, the experimental relaxation time for Co2Rad is found to be almost insensitive to temperature up to ∼20 K. Above this threshold, a stark reduction is observed, coherently with reports in similar compounds.34 The relaxation mechanism is interpreted as driven by QTM and spin-phonon coupling in the two temperature regimes, respectively. The comparison with simulations shows good agreement only for the highest temperature measurement, but the two diverge otherwise. Here we are interested in understanding the role of phonons and we thus perform a comparison with experiments obtained in the presence of an external magnetic field39 to quench the effect of QTM and uncover the intrinsic limits to spin lifetime imposed by spin-phonon relaxation. AC magnetometry results show that relaxation time increases and an exponential regime is revealed for a field value of 0.2 T. Importantly, static fields of this magnitude (smaller than 0.6 T for Co2Rad(39)) only affect QTM, leaving spin-phonon relaxation unchanged, thus making it possible to compare experiments with simulations at or close to zero-field. For T lower than ∼15 K, relaxation time becomes too long to be tracked by AC magnetometry, and DC is instead used. Under the applied field the agreement between simulations and experimental values is excellent and clearly shows that AC tracks the Orbach regime, while the simulated Raman relaxation well describes slower DC values.
Figure 2.
Spin relaxation time. Computed Orbach and Raman relaxation times are reported with continuous and dashed green lines, respectively. Red square and circle symbols correspond to experimental relaxation times extracted from AC and DC magnetometry,39 respectively, in the presence of a magnetic field of 2000 Oe. The black triangles report the experimental relaxation time measured with AC magnetometry in zero external field.39
Aiming at interpreting the mechanism of relaxation for Co2Rad in both Orbach and Raman regimes, we perform new simulations where only a part of the phonons are included. The slope of the ln(τ) vs 1/T is canonically used to estimate the energy of the states involved in the Orbach process. This analysis returns a value of ∼300 cm–1, thus commensurate with the energy of the second excited KD. This analysis, however, is not able to discern between a one- or two-step relaxation mechanism. In the first case, a single phonon with ℏωα ∼ 300 cm–1 would induce a single transition between the ground-state KD and the second excited KD, while in the second scenario two phonons would be sequentially absorbed to reach the second KD. In the latter case, a first phonon with energy ∼230 cm–1 would induce a transition between ground-state KD and the first excited KD, and a second phonon with energy ∼70 cm–1 would promote a subsequent transition from the first excited KD to the second excited KD. The determination of which scenario takes place for Co2Rad is key to identifying which phonons are responsible for relaxation and thus guide synthetic efforts in engineering them. To this end, we compute the Orbach relaxation rate at 20 K by progressively removing phonons with energy higher than a cutoff ωc. We observe that the computed Orbach relaxation rates drastically drop only when ωc is set below 190–230 cm–1, and phonons with higher energy are excluded from the simulation (see SI). This makes it possible to demonstrate that phonons with energy ∼300 cm–1, resonant with the second excited KD, do not contribute to spin relaxation, confirming the assignment of the Orbach relaxation mechanism as promoted by two sequential absorption processes where phonons in resonance with the first excited KD ∼ 230 cm–1 initiate the relaxation process. This can also be seen from the top panel of Figure 3, where the arrows point to the most probable transitions among different spin states, showing that a one-step transition between the ground-state KD and the third excited state is improbable.
Figure 3.
Contributions to spin relaxation. Top panel: Computed transition rates due to one-phonon absorptions for the first five KDs. Only transitions with rates larger than 10–12 ps–1 have been considered. The color scale of the arrows is proportional to the logarithm of the rate. The x-axis reports the expectation value of the z component of the total spin angular momentum. Bottom panel: Computed Orbach and Raman relaxation times are reported with continuous and dashed lines, respectively. Green curves report results for Co2Rad, and purple curves for Co1.
The study of the matrix element of eq 5, supports an interpretation of Raman relaxation at T < 10 K as an intraground-state KD transition. In the giant spin representation of Co2Rad this corresponds to a transition among states with MS = ±5/2. We perform the simulations of the Raman relaxation rates by progressively removing phonons at high energy and identify the low-energy spectrum, below 50 cm–1, to be the only relevant contribution (see SI). We also perform the simulation of Raman rates at 7 K by progressively including more virtual transitions to KDs other than just the first excited one in the sum over c in eq 6. We note that the inclusion of the sole first KD produces a rate that is 1 order of magnitude faster, while the overall relaxation rate is well reproduced when the first two excited KDs are accounted for. However, the first five KDs are necessary to achieve full convergence (see SI). These are interesting observations for two reasons: 1) the contributions to τ of different virtual transitions do not simply add up and cancellation effects can arise, and 2) states beyond the fundamental spin multiplet might play a role.
We now turn to the study of the contribution of different magnetic interactions to the spin-phonon relaxation mechanism. First, we disentangle the contributions of single-ion anisotropy and exchange coupling on the overall molecular magnetic moments dynamics. In this attempt, we perform the simulations for a fictitious molecule Co1 where the magnetic moment on the radical and one of two Co ions have been quenched without affecting anything else. This is achieved by only considering the first term in eq 9 and its derivatives. Figure 3 reports the comparison between spin relaxation in Co1 and Co2Rad, highlighting the large impact of exchange coupling in slowing down Orbach relaxation by up to 2 orders of magnitude and Raman relaxation by up to 4 orders of magnitude. A similar conclusion was reported by Albold et al. in comparing the mononuclear prototype [Co(LA)2]2– (H2LA is 1,2-bis(methanesulfonamido)benzene) with the corresponding dimer34 and precursor of Co2Rad. Indeed the predicted dynamics of Co1 follows very closely the one of [Co(LA)2]2–.
Interestingly, we note that the contribution of exchange
coupling
is only operative at the level of the static Hamiltonian, therefore
in shaping the spin wave function, but has no contribution at the
level of spin-phonon relaxation, as often invoked in classical literature.58 Indeed, if the derivatives of J are omitted from eq 2, results remain unchanged. This is most likely due to the fact that
even though the derivatives (∂J/∂qα) are significant, the spin operator multiplying
them, e.g. , does not couple states with different Ms. We then perform the simulation
of spin-phonon relaxation as a function of the exchange coupling constant
to isolate the importance of the coupling strength. As can be seen
in Figure 4, both Orbach
and Raman relaxation times rapidly increase with J up to ∼200 cm–1. For larger values of J, only marginal, but still visible, increments of relaxation
times are observed. This can be explained by observing the bottom
panel of Figure 4,
where the lowest-energy states of the spin Hamiltonian are reported
as a function of J and color-coded according to the
expectation value of their total spin z-component,
⟨Sz⟩. While
the ground state always remains a pure MS = ±5/2, the first few excited KDs drastically
change their nature and only start settling for J > 200 cm–1. In this high-exchange regime, the
molecule approximately behaves as a 5/2 giant spin and the first three
KDs determine spin relaxation. The residual dependency of τ
for J > 400 cm–1 signals that
the
giant-spin approximation is not perfectly fulfilled. For J < 200 cm–1, the presence of extra low-energy
states offers additional relaxation pathways, cutting down the benefit
of large single-ion zero-field splitting and even producing faster
relaxation rates than in the mononuclear compound. The large rate
of transition to states outside the fundamental multiplet is also
visible in the top panel of Figure 3.
Figure 4.
Spin relaxation time vs exchange coupling strength. Computed Orbach (top panel) and Raman (middle panel) relaxation times are reported as a function of exchange coupling strength. The color code for the lines is reported as a graph sidebar. The energy of the low-lying spin states is reported as a function of exchange coupling strength (bottom panel). The color code is reported as a graph sidebar and corresponds to the computed expectation value of the z-component of the total spin.
Spin-Phonon Coupling
Now that the relaxation mechanism has been determined we can turn our attention to the nature of spin-phonon coupling. We start by plotting how this quantity is distributed over the phonons density of state in Figure 5. As it is common for molecular crystals of this complexity, a continuous and highly structured distribution of phonons is present from very low wavenumber,27,54,59 as also visible in Figure 5 up to ∼250 cm–1. A visual inspection of molecular displacements of phonons in this energy range shows atomic displacements characterized by complex motions delocalized over the entire unit cell. As we move to higher values of ℏωα, the intermolecular motions become less important and internal vibrations become more localized, also generating sharper peaks in the spin-phonon coupling distribution. Here we focus our attention on the two areas of this distribution that our previous analysis has found linked to spin dynamics, i.e. the lowest-energy phonons, responsible for Raman relaxation, and phonons around 200 cm–1, responsible for the first step of the Orbach relaxation. In the former case, we observe motions where the aromatic rings of the ligands remain rigid and tilt with respect to one another (middle panel of Figure 5). On the other hand, the modes in resonance with the first KD (bottom panel of Figure 5) exhibit a large twisting of the aromatic rings of the bmsab ligands and methyl rotations. Admixed to these complex motions, Co–N bonds and the NCoN angles are also sensibly modulated, inducing a non-negligible spin-phonon coupling.
Figure 5.
Spin-phonon coupling distribution. Top panel: The average spin-phonon coupling (Spc) of each phonon with spin is reported as their frequency. A Gaussian smearing of 10 cm–1 is applied to smooth out the distribution. Middle panel: molecular distortions associated with the first available phonon at the Γ-point and responsible for Raman relaxation. The equilibrium molecule is reported in blue and the distorted one is in red. Bottom panel: molecular distortions associated with a phonon of energy ∼200 cm–1 and responsible for the first step of Orbach relaxation. The equilibrium molecule is reported in blue and the distorted one is in red.
Relaxation in Clusters of Larger Nuclearity
Finally, we investigate the role of nuclearity in determining the relaxation time. To this end, we simulate a hypothetical trimer, Co3Rad2, where the original molecule is extended in a chain-like fashion, i.e., the first Co ion is exchange-coupled to the second one through a tmsab radical bridge, and the second Co ion is coupled to a third through a second tmsab radical bridge. The coordination of the first and third Co ions are completed by one bmsab ligand each, as for Co3Rad2 (see inset of Figure 6 and diagrams in the SI). We assume the spin Hamiltonian parameters and spin-phonon coupling coefficients to be identical to the dimer, as detailed in SI. Figure 6 reports the simulated dynamics of Co3Rad2 and shows that Raman relaxation slows down by 4 orders of magnitude. Orbach relaxation also improves, though to a smaller degree, and relaxation at 20 K slows down by a factor of 20. Interestingly, if a chain of four Co ions, Co4Rad3, is now considered, Raman relaxation times become so long that we cannot numerically estimate them, but Orbach relaxation remains virtually identical to Co3Rad2. The latter result is consistent with the notion that Ueff is not dramatically affected by nuclearity alone as the total effective ZFS of the ground spin multiplet scales as ∼1/S2,60,61 but it shows that Raman does not suffer from the same dependency and that it can be completely suppressed with a multi-ion strategy. Figure 6 also compares these results with a state-of-the-art mononuclear Dy SMM [CpDyCp(iPr)5)]+,12 showing that both Co3Rad2 and Co4Rad3 would support a slower Raman relaxation than this compound.
Figure 6.
Spin relaxation and nuclearity. Orbach and Raman relaxation times are reported with continuous and dashed lines, respectively. Green lines are used for Co2Rad, purple lines for Co3Rad2 and red lines for Co4Rad3. L1 = bmsab, L2 = tmsab. Black lines report the fit of experimental Orbach and Raman relaxation times the mononuclear Dy complex [CpDyCp(iPr)5)]+ from ref.12 QTM contributions are neglected to make a direct comparison with the computed spin-phonon relaxation times.
Discussion and Conclusions
The study of Co2Rad with ab initio methods has made it possible to shed light on important aspects of spin relaxation in polynuclear SMMs. First, we have demonstrated that relaxation follows a similar trend to the mononuclear case, with Orbach and Raman relaxation mechanisms operative at high and low temperatures, respectively. This result provides a robust theoretical ground for the interpretation of relaxation experiments in polynuclear SMMs. For Co2Rad, we have shown that both relaxation mechanisms are largely well described by an intra ground-state spin multiplet due to the large value of J for this complex. However, this scenario would rapidly change as J decreases below the value of the single-ion zero-field splitting. This analysis is in agreement with experiments that have observed a detrimental effect of coupling multiple ions17,62,63 and provides a quantitative framework for their interpretation. Importantly, while the effect of exchange on QTM has been known for a long time,13−17 we have here been able to show that large exchange coupling has a dramatic effect on Raman relaxation. To the best of our knowledge, this behavior has only been observed in experiments where a very large exchange coupling among magnetic centers is achieved,17,34,38 but never discussed in detail nor rationalized. These results also shed light on the limits of the giant-spin approximation to describe relaxation in SMMs. Our simulations show that electronic states beyond the fundamental S = 5/2 multiplet play a role in Raman relaxation and to a smaller degree in Orbach relaxation. A similar situation has also been recently observed for the mononuclear case of cobalt ions, where excitations beyond the fundamental quartet have been shown to contribute to Raman virtual transitions.64
The compound investigated here is the result of years of molecular engineering and unsurprisingly our simulations show that both strong exchange coupling and large single-ion zero-field act in a concerted way to suppress spin relaxation. As a consequence, the question of how to achieve further progress in transition-metal-based SMMs beckons. Arguably, several strategies lie ahead: 1) further increase both single-ion zero-field splitting and exchange coupling, 2) engineer lattice and molecular vibrations, and 3) increase clusters’ nuclearity.
While the present molecule already exhibits some of the largest Ueff in transition-metal-based SMMs, further engineering of coordination complexes to increase D and J remains one of the most efficient ways to improve their performances. Similarly to what has been achieved with Dy ions, introducing a large exchange coupling among linearly coordinated cobalt ions11 might reveal unprecedentedly long relaxation times. In this regard, some of the present authors have presented a large-scale computational study of Co(II) coordination complexes,65 suggesting that novel coordination environments able to support maximally large values of axial zero-field splitting are yet to be explored and that the potential of transition-metal-based SMMs is still untapped.
In terms of vibrational design, the visual inspection of the phonons of Co2Rad shows a high degree of rigidity, with the key vibrations driving relaxation being already rather localized on the ligands and often involving the rotation of methyl groups. The latter motion has also been found to play a key role in the relaxation of Vanadyl compounds59 and its substitution with a less flexible unit might represent a promising way to reduce low-energy vibrations. At the same time, we note that a consistent and quantitative definition of molecular rigidity and how it influences relaxation is yet to be achieved. We anticipate that additional theoretical efforts in this direction will be necessary to bring this design rule to fruition.
Finally, we have explored how increasing the nuclearity of Co2Rad could drastically influence its Raman relaxation mechanisms, slowing it down by several orders of magnitude. In particular, we have shown that while the upper limit for Ueff is achieved already at the level of three Co ions, Raman relaxation keeps scaling with nuclearity. This is a central result of our study and shows a clear way forward to controlling low-temperature relaxation. Given the chemical nature of this compound, we limited ourselves to the study of molecular chains, but the study could be expanded to more complex topologies. Interestingly, there is extensive literature on Co, or other ions, single-chain magnets using radical linkers.66,67 In this respect our analysis has 2-fold importance: 1) our proposed synthetic guidelines to unprecedentedly slow Raman relaxations are likely to be well within the reach of synthetic chemistry, and 2) the scope of our simulations serves as a stepping stone toward the rationalization of decades of research in the study of relaxation in 1D magnetic systems and an ab initio description of their Glauber’s dynamics.
We note that all these considerations have been done without considering QTM, and therefore only remain valid in the presence of external field and/or magnetic dilution. The theoretical modeling of QTM is still in its infancy68−70 and we will need further development of relaxation theories to also account for this mechanism. We can however reasonably expect that as the effective spin ground state of SMMs increases with nuclearity, QTM should also be suppressed accordingly, provided no transverse zero-filed splitting is introduced.
It is worth stressing that the results achieved for Co2Rad hold a general validity for polynuclear coordination compounds, including Ln-based ones, as well as beyond molecular systems. Indeed, while the chemistry of magnetic compounds can vary drastically across both molecular and solid-state materials, from a physical point of view their magnetic properties can often be described within the same framework, i.e. zero-field splitting, exchange coupling and a thermal bath of molecular crystal vibrations. As it has been already been shown for the mononuclear complexes and solid-state defects, these ingredients do not change qualitatively and the same underlying principles of relaxation remain valid across the full range of magnetic systems.27,71
In conclusion, we have provided a full ab initio description of the spin relaxation mechanism in a paradigmatic air-stable SMM where both single-ion zero-field splitting and exchange coupling have been maximized. Importantly, simulations demonstrate that further extending the nuclearity of this compound from two to just three or four Co(II) ions could potentially lead to a compound with unprecedentedly slow Raman spin relaxation. These results hold a general validity for both 3d and 4f molecules and we anticipate that they will provide a new blueprint for the engineering of novel polynuclear SMMs, as well as pave the way to the interpretation of spin dynamics in arbitrarily complex magnetic structures.
Acknowledgments
A.L. and S.M. acknowledge support from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. [948493]). A.K. and J.N. acknowledge support by the state of Baden-Württemberg through bwHPC and the German Research Foundation (DFG) through grant INST 40/575-1 FUGG (JUSTUS 2 cluster). J.v.S and D.H. acknowledge the support of the Landesgraduiertenforderung of the state of Baden-Württenberg (DFG SL104/10-1, SA1840-/9-1). A.L. and S.M. received computational resources by the Trinity College Research IT and the Irish Centre for High-End Computing (ICHEC). A.K. is also grateful for the support from the Stuttgart Center for Simulation Science (SimTech).
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acscentsci.4c02139.
The authors declare no competing financial interest.
Supplementary Material
References
- Sessoli R.; Gatteschi D.; Caneschi A.; Novak M. Magnetic bistability in a metal-ion cluster. Nature 1993, 365, 141–143. 10.1038/365141a0. [DOI] [Google Scholar]
- Gatteschi D.; Sessoli R.; Villain J.. Molecular nanomagnets; Oxford University Press, 2006. [Google Scholar]
- Sanvito S. Molecular spintronics. Chem. Soc. Rev. 2011, 40, 3336–3355. 10.1039/c1cs15047b. [DOI] [PubMed] [Google Scholar]
- Atzori M.; Sessoli R. The second quantum revolution: role and challenges of molecular chemistry. J. Am. Chem. Soc. 2019, 141, 11339–11352. 10.1021/jacs.9b00984. [DOI] [PubMed] [Google Scholar]
- Eaton S. S.; Eaton G. R. Relaxation Times of Organic Radicals and Transition Metal Ions. Distance measurements in biological systems by EPR 2002, 19, 29–154. 10.1007/0-306-47109-4_2. [DOI] [Google Scholar]
- Maehrlein S. F.; et al. Dissecting spin-phonon equilibration in ferrimagnetic insulators by ultrafast lattice excitation. Science advances 2018, 4, eaar5164 10.1126/sciadv.aar5164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lunghi A. Spin-phonon relaxation in magnetic molecules: Theory, predictions and insights. Computational Modelling of Molecular Nanomagnets 2023, 34, 219–289. 10.1007/978-3-031-31038-6_6. [DOI] [Google Scholar]
- Perfetti M.; Bendix J. The multiple faces, and phases, of magnetic anisotropy. Inorg. Chem. 2019, 58, 11875–11882. 10.1021/acs.inorgchem.9b00636. [DOI] [PubMed] [Google Scholar]
- Neese F.; Pantazis D. A. What is not required to make a single molecule magnet. Faraday Discuss. 2011, 148, 229–238. 10.1039/C005256F. [DOI] [PubMed] [Google Scholar]
- Rinehart J. D.; Long J. R. Exploiting single-ion anisotropy in the design of f-element single-molecule magnets. Chemical Science 2011, 2, 2078–2085. 10.1039/c1sc00513h. [DOI] [Google Scholar]
- Bunting P. C.; et al. A linear cobalt (II) complex with maximal orbital angular momentum from a non-aufbau ground state. Science 2018, 362, eaat7319 10.1126/science.aat7319. [DOI] [PubMed] [Google Scholar]
- Guo F.-S.; et al. Magnetic hysteresis up to 80 K in a dysprosium metallocene single-molecule magnet. Science 2018, 362, 1400–1403. 10.1126/science.aav0652. [DOI] [PubMed] [Google Scholar]
- Wernsdorfer W.; Aliaga-Alcalde N.; Hendrickson D. N.; Christou G. Exchange-biased quantum tunnelling in a supramolecular dimer of single-molecule magnets. Nature 2002, 416, 406–409. 10.1038/416406a. [DOI] [PubMed] [Google Scholar]
- Wilson A.; et al. Magnetization tunneling in high-symmetry single-molecule magnets: Limitations of the giant spin approximation. Phys. Rev. B 2006, 74, 140403. 10.1103/PhysRevB.74.140403. [DOI] [Google Scholar]
- Hill S.; et al. Magnetic quantum tunneling: insights from simple molecule-based magnets. Dalton Transactions 2010, 39, 4693–4707. 10.1039/c002750b. [DOI] [PubMed] [Google Scholar]
- Sorace L.; et al. Origin and spectroscopic determination of trigonal anisotropy in a heteronuclear single-molecule magnet. Phys. Rev. B 2013, 88, 104407. 10.1103/PhysRevB.88.104407. [DOI] [Google Scholar]
- Rinehart J. D.; Fang M.; Evans W. J.; Long J. R. Strong exchange and magnetic blocking in N 23--radical-bridged lanthanide complexes. Nat. Chem. 2011, 3, 538–542. 10.1038/nchem.1063. [DOI] [PubMed] [Google Scholar]
- Swain A.; Sharma T.; Rajaraman G. Strategies to quench quantum tunneling of magnetization in lanthanide single molecule magnets. Chem. Commun. 2023, 59, 3206–3228. 10.1039/D2CC06041H. [DOI] [PubMed] [Google Scholar]
- Zadrozny J. M.; Long J. R. Slow magnetic relaxation at zero field in the tetrahedral complex [Co(SPh)4]2–. J. Am. Chem. Soc. 2011, 133, 20732–20734. 10.1021/ja2100142. [DOI] [PubMed] [Google Scholar]
- Rechkemmer Y.; et al. A four-coordinate cobalt (II) single-ion magnet with coercivity and a very high energy barrier. Nat. Commun. 2016, 7, 10467. 10.1038/ncomms10467. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Accorsi S.; et al. Tuning anisotropy barriers in a family of tetrairon (III) single-molecule magnets with an S = 5 ground state. J. Am. Chem. Soc. 2006, 128, 4742–4755. 10.1021/ja0576381. [DOI] [PubMed] [Google Scholar]
- Demir S.; Jeon I.-R.; Long J. R.; Harris T. D. Radical ligand-containing single-molecule magnets. Coord. Chem. Rev. 2015, 289, 149–176. 10.1016/j.ccr.2014.10.012. [DOI] [Google Scholar]
- Pedersen K. S.; et al. Design of single-molecule magnets: insufficiency of the anisotropy barrier as the sole criterion. Inorg. Chem. 2015, 54, 7600–7606. 10.1021/acs.inorgchem.5b01209. [DOI] [PubMed] [Google Scholar]
- Lunghi A. Toward exact predictions of spin-phonon relaxation times: An ab initio implementation of open quantum systems theory. Science Advances 2022, 8, eabn7880 10.1126/sciadv.abn7880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lunghi A.; Sanvito S. The limit of spin lifetime in solid-state electronic spins. J. Phys. Chem. Lett. 2020, 11, 6273–6278. 10.1021/acs.jpclett.0c01681. [DOI] [PubMed] [Google Scholar]
- Briganti M.; et al. 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. 10.1021/jacs.1c05068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mondal S.; Lunghi A. Unraveling the contributions to spin–lattice relaxation in Kramers single-molecule magnets. J. Am. Chem. Soc. 2022, 144, 22965–22975. 10.1021/jacs.2c08876. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moseley D. H.; et al. Spin–phonon couplings in transition metal complexes with slow magnetic relaxation. Nat. Commun. 2018, 9, 2572. 10.1038/s41467-018-04896-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lunghi A.; Totti F.; Sanvito S.; Sessoli R. Intra-molecular origin of the spin-phonon coupling in slow-relaxing molecular magnets. Chemical Science 2017, 8, 6051–6059. 10.1039/C7SC02832F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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. 10.1021/jacs.1c01410. [DOI] [PubMed] [Google Scholar]
- Dey S.; Sharma T.; Rajaraman G. Unravelling the role of spin–vibrational coupling in designing high-performance pentagonal bipyramidal Dy(III) single ion magnets. Chemical Science 2024, 15, 6465–6477. 10.1039/D4SC00823E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bencini A.; Totti F. A few comments on the application of density functional theory to the calculation of the magnetic structure of oligo-nuclear transition metal clusters. J. Chem. Theory Comput. 2009, 5, 144–154. 10.1021/ct800361x. [DOI] [PubMed] [Google Scholar]
- Maurice R.; et al. Theoretical determination of the zero-field splitting in copper acetate monohydrate. Inorg. Chem. 2011, 50, 6229–6236. 10.1021/ic200506q. [DOI] [PubMed] [Google Scholar]
- Albold U.; Bamberger H.; Hallmen P. P.; van Slageren J.; Sarkar B. Strong exchange couplings drastically slow down magnetization relaxation in an air-stable cobalt (II)-radical single-molecule magnet (SMM). Angew. Chem., Int. Ed. 2019, 58, 9802–9806. 10.1002/anie.201904645. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sharma P.; Truhlar D. G.; Gagliardi L. Magnetic coupling in a tris-hydroxo-bridged chromium dimer occurs through ligand mediated superexchange in conjunction with through-space coupling. J. Am. Chem. Soc. 2020, 142, 16644–16650. 10.1021/jacs.0c06399. [DOI] [PubMed] [Google Scholar]
- Dey S.; Rajaraman G. Attaining record-high magnetic exchange, magnetic anisotropy and blocking barriers in dilanthanofullerenes. Chemical Science 2021, 12, 14207–14216. 10.1039/D1SC03925C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nguyen G. T.; Ungur L. The role of radical bridges in polynuclear single-molecule magnets. Chemistry A European J 2022, 28, e202200227 10.1002/chem.202200227. [DOI] [PubMed] [Google Scholar]
- Gould C. A.; et al. Ultrahard magnetism from mixed-valence dilanthanide complexes with metal-metal bonding. Science 2022, 375, 198–202. 10.1126/science.abl5470. [DOI] [PubMed] [Google Scholar]
- Hunger D.; et al. Electronic structure of mononuclear and radical-bridged dinuclear cobalt (II) single-molecule magnets. Nat. Commun. 2025, 16, 2157. 10.1038/s41467-025-57210-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lunghi A.; Sanvito S. Multiple spin–phonon relaxation pathways in a Kramer single-ion magnet. J. Chem. Phys. 2020, 153, 174113. 10.1063/5.0017118. [DOI] [PubMed] [Google Scholar]
- Kühne T. D. CP2K: An electronic structure and molecular dynamics software package-quickstep: Efficient and accurate electronic structure calculations. J. Chem. Phys. 2020, 152, 194103. 10.1063/5.0007045. [DOI] [PubMed] [Google Scholar]
- Goedecker S.; Teter M.; Hutter J. Separable dual-space gaussian pseudopotentials. Phys. Rev. B 1996, 54, 1703. 10.1103/PhysRevB.54.1703. [DOI] [PubMed] [Google Scholar]
- Perdew J. P.; Burke K.; Ernzerhof M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- Grimme S.; Antony J.; Ehrlich S.; Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
- Werner H.-J.et al. MOLPRO, version 2023.1, a package of ab initio programs (2023). See https://www.molpro.net.
- Werner H.-J.; et al. The Molpro quantum chemistry package. J. Chem. Phys. 2020, 152, 144107. 10.1063/5.0005081. [DOI] [PubMed] [Google Scholar]
- Weigend F.; Ahlrichs R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. 10.1039/b508541a. [DOI] [PubMed] [Google Scholar]
- Kreplin D. A.; Knowles P. J.; Werner H.-J. MCSCF optimization revisited. II. combined first- and second-order orbital optimization for large molecules. J. Chem. Phys. 2020, 152, 074102. 10.1063/1.5142241. [DOI] [PubMed] [Google Scholar]
- Menezes F.; Kats D.; Werner H.-J. Local complete active space second-order perturbation theory using pair natural orbitals (PNO-CASPT2). J. Chem. Phys. 2016, 145, 124115. 10.1063/1.4963019. [DOI] [PubMed] [Google Scholar]
- Kats D.; Werner H.-J. Multi-State Local Complete Active Space Second-Order Perturbation Theory Using Pair Natural Orbitals (PNO-MS-CASPT2). J. Chem. Phys. 2019, 150, 214107. 10.1063/1.5097644. [DOI] [PubMed] [Google Scholar]
- McWeeny R. SCF theory for excited states. Mol. Phys. 1974, 28, 1273–1282. 10.1080/00268977400102581. [DOI] [Google Scholar]
- Chibotaru L. F.; Ungur L. Ab initio calculation of anisotropic magnetic properties of complexes. I. Unique definition of pseudospin Hamiltonians and their derivation. J. Chem. Phys. 2012, 137, 064112. 10.1063/1.4739763. [DOI] [PubMed] [Google Scholar]
- 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. 10.1038/ncomms14620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lunghi A.; Sanvito S. How do phonons relax molecular spins?. Science Advances 2019, 5, eaax7163 10.1126/sciadv.aax7163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boeer A. B.; et al. A spectroscopic investigation of magnetic exchange between highly anisotropic spin centers. Angew. Chem., Int. Ed. 2011, 50, 4007–4011. 10.1002/anie.201100306. [DOI] [PubMed] [Google Scholar]
- Neese F. Importance of direct spinspin coupling and spin-flip excitations for the zero-field splittings of transition metal complexes: a case study. J. Am. Chem. Soc. 2006, 128, 10213–10222. 10.1021/ja061798a. [DOI] [PubMed] [Google Scholar]
- Suhr S.; et al. Air-stable dinuclear complexes of four-coordinate ZnII and NiII ions with a radical bridge: A detailed look at redox activity and antiferromagnetic coupling. Inorg. Chem. 2024, 63, 6042–6050. 10.1021/acs.inorgchem.4c00351. [DOI] [PubMed] [Google Scholar]
- Abragam A.; Bleaney B.. Electron paramagnetic resonance of transition ions; OUP, 2012. [Google Scholar]
- Garlatti E.; et al. The critical role of ultra-low-energy vibrations in the relaxation dynamics of molecular qubits. Nat. Commun. 2023, 14, 1653. 10.1038/s41467-023-36852-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Waldmann O. A criterion for the anisotropy barrier in single-molecule magnets. Inorg. Chem. 2007, 46, 10035–10037. 10.1021/ic701365t. [DOI] [PubMed] [Google Scholar]
- Cirera J.; Ruiz E.; Alvarez S.; Neese F.; Kortus J. How to build molecules with large magnetic anisotropy. Chemistry A European J 2009, 15, 4078–4087. 10.1002/chem.200801608. [DOI] [PubMed] [Google Scholar]
- Patrascu A. A.; et al. A chimeric design of heterospin 2p–3d, 2p–4f, and 2p–3d–4f complexes using a novel family of paramagnetic dissymmetric compartmental ligands. Chem. Commun. 2017, 53, 6504–6507. 10.1039/C7CC03236F. [DOI] [PubMed] [Google Scholar]
- Liu F.; et al. Single molecule magnet with an unpaired electron trapped between two lanthanide ions inside a fullerene. Nat. Commun. 2017, 8, 16098. 10.1038/ncomms16098. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mariano L. A.; Mondal S.; Lunghi A. Spin-vibronic dynamics in open-shell systems beyond the spin hamiltonian formalism. J. Chem. Theory Comput. 2024, 20, 323–332. 10.1021/acs.jctc.3c01130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mariano L. A.; Nguyen V. H. A.; Briganti V.; Lunghi A. Charting regions of cobalt’s chemical space with maximally large magnetic anisotropy: A computational high-throughput study. J. Am. Chem. Soc. 2024, 146, 34158. 10.1021/jacs.4c14076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Caneschi A.; et al. Cobalt (II)-nitronyl nitroxide chains as molecular magnetic nanowires. Angew. Chem., Int. Ed. 2001, 40, 1760–1763. . [DOI] [PubMed] [Google Scholar]
- Bernot K.; Bogani L.; Caneschi A.; Gatteschi D.; Sessoli R. A family of rare-earth-based single chain magnets: playing with anisotropy. J. Am. Chem. Soc. 2006, 128, 7947–7956. 10.1021/ja061101l. [DOI] [PubMed] [Google Scholar]
- Isakov S. V.; et al. Understanding quantum tunneling through quantum monte carlo simulations. Phys. Rev. Lett. 2016, 117, 180402. 10.1103/PhysRevLett.117.180402. [DOI] [PubMed] [Google Scholar]
- Aravena D. Ab initio prediction of tunneling relaxation times and effective demagnetization barriers in Kramers lanthanide single-molecule magnets. J. Phys. Chem. Lett. 2018, 9, 5327–5333. 10.1021/acs.jpclett.8b02359. [DOI] [PubMed] [Google Scholar]
- Mattioni A.; et al. Vibronic effects on the quantum tunnelling of magnetisation in kramers single-molecule magnets. Nat. Commun. 2024, 15, 485. 10.1038/s41467-023-44486-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mondal S.; Lunghi A. Spin-phonon decoherence in solid-state paramagnetic defects from first principles. npj Computational Materials 2023, 9, 120. 10.1038/s41524-023-01082-9. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.