Next Article in Journal
Ni(NH3)2(NO3)2—A 3-D Network through Bridging Nitrate Units Isolated from the Thermal Decomposition of Nickel Hexammine Dinitrate
Previous Article in Journal
The Versatile SALSAC Approach to Heteroleptic Copper(I) Dye Assembly in Dye-Sensitized Solar Cells
Previous Article in Special Issue
Ferromagnetic Cluster Spin Wave Theory: Concepts and Applications to Magnetic Molecules
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecular Engineering of High Energy Barrier in Single-Molecule Magnets Based on [MoIII(CN)7]4− and V(II) Complexes

by
Vladimir S. Mironov
Shubnikov Institute of Crystallography of Federal Scientific Research Centre “Crystallography and Photonics” RAS, Leninsky Prospect 59, 119333 Moscow, Russia
Inorganics 2018, 6(2), 58; https://doi.org/10.3390/inorganics6020058
Submission received: 1 May 2018 / Revised: 25 May 2018 / Accepted: 28 May 2018 / Published: 30 May 2018
(This article belongs to the Special Issue Single-Molecule Magnets)

Abstract

:
Molecular engineering of high energy barrier Ueff in single-molecule magnets (SMMs) of general composition MoIIIkVIIm based on orbitally-degenerate pentagonal-bipyramidal [MoIII(CN)7]4− complexes with unquenched orbital momentum and high-spin V(II) complexes is discussed. In these SMMs, the barrier originates exclusively from anisotropic Ising-type exchange interactions −Jxy(SixSjx + SiySjy) − JzSizSjz in the apical cyano-bridged pairs MoIII–CN–VII, which produce a double-well energy profile with a doubly degenerate ground spin state ±MS. It is shown that the spin-reversal barrier Ueff is controlled by anisotropic exchange parameters Jz, Jxy, and the number n of apical MoIII–CN–VII groups in a SMM cluster, Ueff ~ 0.5|JzJxy|n; it can reach a value of many hundreds of wavenumbers (up to 741 cm−1). This finding provides a very efficient straightforward strategy for further scaling Ueff to high values (>1000 cm−1) by means of enhancing exchange parameters Jz, Jxy, and increasing the number of [MoIII(CN)7]4− complexes in a SMM molecule.

Graphical Abstract

1. Introduction

Single-molecule magnets (SMMs) are individual high-spin molecules featuring slow spin relaxation and preserving their magnetic moment below characteristic blocking temperature TB [1,2,3,4,5,6,7]. The slow relaxation of SMMs is caused by a double-well potential with the two lowest MS = +S and MS = −S spin states separated by an energy barrier Ueff, which arises from the combined effect of the easy-axis (Ising-type) magnetic anisotropy (quantified by the axial zero-field splitting parameter D) and high-spin ground state S. The spin-reversal barrier Ueff is expressed by |D|S2 or |D|(S2 − 1/4) for integer and half-integer spins, respectively. Since the discovery of the first representative (Mn12Ac) in 1993 [1,2], SMMs have attracted increasing attention due to their huge forward-looking application potential in the design of high-density information storage [8], quantum computing [9,10,11,12], and molecular spintronics [13,14]. Currently, development of SMMs is one of the most rapidly growing lines of research in the field of molecular magnetism. However, rapid advance of the SMM-based technology is hampered by a low blocking temperature TB, which is normally below a few Kelvins [1,2,3,4,5,6,7]. Designing of SMMs with higher TB and Ueff characteristics has remained a highly challenging task over the past two decades. Earlier efforts toward increasing Ueff and TB were mainly focused on obtaining large polynuclear transition metal complexes (mostly based on MnIII ions) with a high ground-state spin S. However, as these studies progressed, it was realized that the barrier Ueff and temperature TB do not generally rise with increasing spin and nuclearity of a magnetic molecule because the overall magnetic anisotropy D is inversely proportional to the square of the spin, D ~ S−2 [15,16,17]. In fact, the barrier Ueff = |D|S2 is largely independent on spin S; consequently, the barrier Ueff can be increased only by increasing the molecular magnetic anisotropy D. Basically, magnetic anisotropy D originates from a combined effect of spin-orbit coupling and some orbital angular momentum; it may operate as a first-order perturbation or as a second-order perturbation, depending on the nature of the ground state. In SMMs based on high-spin 3d ions, magnetic anisotropy D originates from single-ion contributions resulting from zero field splitting (ZFS) on individual magnetic ions. Being the product of a second-order effect, the ZFS energy is relatively small, typically within few tens of cm−1 or less for spin-only 3d-ions (such as MnIII, FeIII, and NiII); thus, it leads generally to a rather small magnetic anisotropy D and low barriers Ueff [15,16,17]. Much stronger first-order magnetic anisotropy is produced by magnetic ions with the unquenched orbital momentum, such as lanthanide and actinide ions [18,19,20,21]. Indeed, all lanthanide ions with open 4f-shell (except Gd3+) exhibit large unquenched orbital momentum, which, in concert with strong spin-orbit coupling and weak crystal-field splitting of 4f states, produces extremely strong single-ion magnetic anisotropy of Ln3+ ions; these features taken together ultimately lead to high Ueff and TB values of Ln-based SMMs. With the discovery of the first Ln-based SMM, the double decker mononuclear complex [(Pc)2Tb] (Pc = phthalocyanine) in 2003 [22], lanthanide complexes have opened a new avenue to high-performance SMMs [23,24,25,26,27,28]. Even mononuclear lanthanide complexes exhibit slow magnetic relaxation at low temperatures with a high barrier Ueff; they are referred to as single-ion magnets (SIMs) [18,19,20,21,22,23,24,25,26,27,28,29]. The last few years have seen impressive progress in designing advanced SIMs with new record blocking temperature, such as TB = 20 K [30], and spin-reversal barrier Ueff = 1261 cm−1 (1815 K) [31]). Recently, new record SMM characteristics were reported for a mononuclear dysprosium complex [Dy(Cp(t-Bu3)2]+, Ueff/kB = 1760 K, TB = 60 K [32,33]. These values are an order of magnitude higher than those known for SMMs based on polynuclear transition metal complexes, Ueff = 62 cm−1 and TB = 4.5 K for a Mn6 complex with S = 12 [34]).
More recently, there has also been a growing interest in mononuclear 3d complexes with unquenched first-order orbital momentum, which is provided by a special less-common coordination of the metal atom. Especially interesting are complexes with a large uniaxial magnetic anisotropy (Ising spin units), which are capable to behave as SIMs with a large energy barrier originating from the first-order spin-orbit splitting of the orbitally degenerate ground state [35,36,37,38]. In recent years, there have been numerous reports on monometallic 3d complexes with SIM behavior [35,36]. Among them, the largest Ueff barrier has been obtained for linear two-coordinated Fe(I) [37] and Co(II) [38] complexes (246 and 413 cm−1, respectively); high energy barriers were also reported for other low-coordinated 3d transition metal complexes [39,40,41,42].
These results vividly show that magnetic centers with unquenched orbital angular momentum provide great opportunities to maximize magnetic anisotropy and to achieve a high barrier and blocking temperature. However, it should be emphasized that the current record-breaking SMM characteristics are based exclusively on single-center magnetic anisotropy of high-spin monometallic complexes, which is already close to the physical limits for 4f and 3d ions. In fact, the maximum feasible barrier Ueff in 4f-SIMs is specified by the total crystal-field splitting energy of the ground J-multiplet of Ln3+ ions, which is normally within several hundreds of cm−1 and very rarely reaches 1000 cm−1 or higher [43]. Therefore, the current record barrier around 1250 cm−1 for Dy-based SIMs [32,33] is close to the maximum CF splitting energy for lanthanide ions and thus it is difficult to increase the barrier far beyond this value [44]. The same is also true for mononuclear high-spin 3d complexes with unquenched orbital momentum, in which maximum value of Ueff is controlled by the spin-orbit splitting energy of the ground orbital manifold; this energy is limited by ca. 1000 cm−1 due to weak spin-orbit coupling for 3d electrons.
Alternative strategy toward high-performance SMMs is based on pair-ion contributions to the molecular magnetic anisotropy D rather than on single-ion contributions [45]. These pair-ion contributions are associated with anisotropic exchange interactions produced by low-spin (S = 1/2) orbitally-degenerate 4d and 5d complexes with unquenched orbital angular momentum, such as pentagonal-bipyramidal (PBP) complexes [MoIII(CN)7]4− [45,46,47] and [ReIV(CN)7]3− [46,48,49], and octahedral hexacyano complexes [RuIII(CN)6]3− [50] and [OsIII(CN)6]3− [51]. It is noteworthy that, unlike high-spin 4f and 3d complexes, these complexes exhibit no single-ion magnetic anisotropy in the usual sense, i.e., as the ZFS of the ground-state spin multiplet 2S + 1 (which occurs at S > 1/2). Instead, their single-ion magnetic anisotropy can be seen only in an anisotropic g-tensor and in anisotropic magnetic susceptibility; moreover, it can completely disappear for magnetically isotropic [RuIII(CN)6]3− and [OsIII(CN)6]3− octahedral complexes [50,51]. In this case, the unquenched orbital momentum of low-spin orbitally degenerate 4d and 5d complexes affects the overall magnetic anisotropy through highly anisotropic exchange interactions with other spin carriers [45,46,47,49]. Thus, being incorporated into polynuclear heterometallic spin clusters together with high-spin 3d ions, these 4d/5d complexes form pair-ion 4d/5d–3d anisotropic exchange interactions, which can produce high magnetic anisotropy D and spin reversal barrier Ueff [45,46,47,49]. It is important to note, however, that the anisotropic exchange interactions in themselves do not necessarily lead to a SMM behavior because of the condition for a molecular spin cluster to have a negative magnetic anisotropy with a small transverse component (D < 0, E ~ 0) [52,53]. For this, two additional conditions are needed: (i) the anisotropic exchange interactions must have the Ising-type character, −Jxy(SixSjx + SiySjy) − JzSizSjz with |Jz| >> |Jxy|; and (ii) the anisotropic exchange parameters Jz, Jxy must be large in absolute value. Actually, among known orbitally-degenerate complexes, only PBP complexes [MoIII(CN)7]4− and [ReIV(CN)7]3− fully meet these conditions [45,46,47,48,49,50,51,52,53,54]. As shown in our previous works, these complexes have a unique property to form uniaxial anisotropic spin coupling Heff = −Jxy(SMxSMnx + SMySMny) − JzSMzSMnz (M = MoIII, ReIV) with high-spin MnII ions (S = 5/2) attached in both the apical and equatorial coordination positions; remarkably, the uniaxial symmetry of the spin Hamiltonian Heff retains even for the actual low symmetry M–CN–Mn exchange-coupled pairs [47,49]. The apical pairs M–CN–Mn exhibit Ising-type anisotropic exchange interactions (|Jz| > |Jxy|), while exchange interaction in the equatorial pairs is nearly isotropic with a small easy-axis (xy) anisotropic component (|Jz| < |Jxy|, |JzJxy| << |Jz|) [47,49]. Ising-type exchange interactions of apical pairs are especially important since they generate uniaxial magnetic anisotropy of the necessary quality (i.e., D < 0, E ~ 0) and form a barrier. Thus, in the MoIIIMnII2 trinuclear complex composed of central [MoIII(CN)7]4− complex and two MnII ions in the apical positions, Ising-type exchange interactions −Jxy(SMoxSMnx + SMoySMny) − JzSMozSMnz of two apical MoIII–CN–MnII pairs (with Jz = −34, Jxy = −11 cm−1) form a double-well potential with the energy barrier of 40.5 cm−1 (58.5 K) and blocking temperature TB = 3.2 K [55]. Quite recently, another related MoIIIMnII2 SMM complex with very close characteristics (Ueff = 44.9 cm−1, TB = 2.5 K, Jz = −35.4, Jxy = −11.4 cm−1) was reported [56]. Remarkably, in these systems, the barrier Ueff is controlled exclusively by the anisotropic exchange parameters (namely, by the difference |JzJxy|) and by the number n of the apical exchange-coupled pairs MoIII–CN–MnII, Ueff ≈ |JzJxy|n [47] (Figure 1).
This opens new avenues for a significant increase in the barrier and the blocking temperature both by enhancing exchange interactions and by increasing nuclearity of the SMM cluster [45,46,47]. To further develop this strategy, this article develops theoretical frameworks for obtaining high SMM characteristics in some hypothetical 4d–3d heterometallic polynuclear complexes based on [MoIII(CN)7]4− heptacyanometallate and VII ions. In these systems, high Ueff and TB values could potentially be obtained due to much stronger exchange interactions in MoIII–CN–VII cyano-bridged pairs, i.e., J > 200 cm−1 for VII [57,58,59] vs. J ~ 40 cm−1 for MnII [55,56]. It is important to emphasize that this system fully meets the specific requirements for orbitally-degenerate complexes, which are necessary for obtaining a high barrier. In this respect, it is relevant to note that anisotropic exchange interactions were also observed in other heterometallic cyano-bridged transition-metal complexes [53,54]; among 3d-based cyano-bridged complexes, anisotropic exchange was first reported in a {LCuII–NC–FeIII(CN)5} ferromagnetically coupled cluster [60]. However, as is shown below, most of the d-complexes featuring anisotropic exchange are of little interest in increasing SMM characteristics, either due to small exchange parameters or because of low symmetry of the anisotropic spin Hamiltonian. As for lanthanide-based SMMs, 4f-3d anisotropic exchange contributions to the Ueff barrier and hysteresis have recently been shown to be important in heterometallic {CrIII2DyIII2} SMM clusters [61,62]. At the same time, it is clear that lanthanide ions are generally not good candidates for the efficient implementation of the claimed strategy owing to inherently weak exchange interactions resulting from the core-like nature of 4f electrons.
The aim of this work is to establish main regularities in the variation of the spin-reversal barrier Ueff in heterometallic MoIII–VII molecular spin clusters, depending on their size, composition and topology of anisotropic and isotropic exchange linkages. It is of special interest to explore the interplay between anisotropic exchange interactions in the apical MoIII–CN–VII pairs and isotropic exchange interactions in the equatorial pairs and to analyze their impact on the overall magnetic anisotropy and barrier. Some specific approaches to engineering high energy barrier in terms of anisotropic exchange interactions are discussed.

2. Results and Discussion

2.1. Electronic Structure of [MoIII(CN)7]4− Complex

This section discusses the basic mechanism of anisotropic exchange interactions between [MoIII(CN)7]4− complexes and VII ions. In fact, this mechanism is similar to that considered earlier for the spin coupling between [MoIII(CN)7]4− and MnII [47] complexes, since in both cases strong exchange anisotropy in the MoIII–CN–M(3d) pairs is caused by the unquenched angular orbital angular momentum of the PBP complex [MoIII(CN)7]4−. Although the electronic structure and magnetic properties of [MoIII(CN)7]4− were previously discussed in detail in Ref. [45,47], it is still appropriate to provide here an overview of its main characteristics, which are essential for understanding specific details of the origin of strong exchange anisotropy. In the PBP ligand surrounding, the two lowest degenerate orbitals 4dxz and 4dyz are populated by three electrons to produce an orbital doublet 2Φxz,yz = 2Φ(ML = ±1), whose two components correspond to the ML = ±1 projection of the unquenched angular orbital momentum L on the polar z-axis of the [Mo(CN)7]4− bipyramid. The 2Φ(ML = ±1) wave functions are represented by two degenerate electronic configurations (d+1)2(d−1)1 and (d−1)2(d+1)1 with d±1 = (dxz ± idyz)/√2 being complex d orbitals with definite projection of the orbital momentum (ml = ±1) on the polar axis of the pentagonal bipyramid (Figure 2). Alternatively, the real wave functions are expressed by 2Φxz = (dyz)2(dxz)1 and 2Φyz = (dxz)2(dyz)1. The orbital doublet 2Φ undergoes spin-orbit splitting (ζMoLS) into the ground φ(±1/2) and excited χ(±1/2) Kramers doublets (Figure 2). In the strong pentagonal-bipyramidal ligand field, the φ(±1/2) wave functions are well described by the single-determinant electronic configurations (d+1)2(d−1)↑ and (d−1)2(d+1)↓ with unquenched orbital momentum (Lz = ±1, Figure 2) [45,47]. As a result, [Mo(CN)7]4− complex exhibits highly anisotropic g-tensor (such as gz = 3.89 and gx = gy = 1.77 [63].
In distorted [MoIII(CN)7]4− complexes, the orbital momentum is reduced or quenched due to low-symmetry ligand field splitting δ of the ground orbital doublet 2Φ±. However, weak or moderate low-symmetry distortions (when δ < ζMo) do not significantly change the overall picture, leaving the orbital angular momentum unquenched. Moreover, these low-symmetry distortions of the [Mo(CN)7]4− bipyramid do not affect the axial symmetry of the anisotropic exchange Hamiltonian Heff = −Jxy(SMoxSVx + SMoySVy) − JzSMozSVz (see below).

2.2. Anisotropic Spin Coupling MoIII–CN–VII

Analysis of anisotropic exchange interactions between [MoIII(CN)7]4− complexes and VII ions basically follows the theoretical model developed for MoIII–CN–MnII exchange coupled pairs in MoIIIMnII2 SMM [47]. With the spin-orbit coupling (SOC) on MoIII switched off (ζMo = 0), the exchange interaction between MoIII and VII is described by an isotropic orbitally dependent spin Hamiltonian Horb = A + RSMoSV, which acts in the space of the 2Φ±(Mo) × 4A2(V) wave functions of the MoIII–CN–VII pair (with × being the anti-symmetrized product and 4A2(V) the ground state of high-spin VII ion). Here, SMo and SV are spin operators of MoIII and VII, while A and R are, respectively, spin-independent and spin-dependent orbital operators acting on the orbital variables only. In general case, the A and R operators are written as a Hermitian 2 × 2 matrices; it is noteworthy that the spin-dependent orbital R matrix is diagonalized by a rotation of the coordinate frame around the polar z-axis [47]. Therefore, without loss of generality, Horb can be written as
H orb = ( A 11 A 12 A 21 A 22 ) ( J 1 0 0 J 2 ) S M o S V ,
where J1 and J2 are orbital exchange parameters referring to the real wave functions 2Φxz = (dyz)2(dxz)1 and 2Φyz = (dxz)2(dyz)1. Then, when the SOC on MoIII is turned on, the isotropic orbitally dependent operator Horb in Equation (1) transforms into an anisotropic spin Hamiltonian Heff that describes effective spin coupling between the ground Kramers doublet φ(±1/2) and the true spin S = 3/2 of VII ions. The spin Hamiltonian Heff is obtained by projection of the full Hamiltonian A + RSMoSV + ζMoLMoSMo acting in the space of the 2Φ±(Mo) × 4A2(V) wave functions onto the restricted space of the φ(m) × 4A2(V) wave functions (where m = ±1/2 is the projection of the fiction spin S = 1/2 of the ground Kramers doublet φ(±1/2)). At this point it is important to emphasize a remarkable property of the pentagonal bipyramidal complex [MoIII(CN)7]4− that the SOC operator ζ4dLMoSMo is diagonal within the space of the 2Φ±(Mo) × 4A2(V) wave functions since it does not mix the two states 2Φ(ML = +1) and 2Φ(ML = −1) of MoIII; thus the ζ4dLMoSMo operator can be replaces by its z-component ζ4dLzMoSzMo. This implies that the full spin Hamiltonian A + RSMoSV + ζ4dLzMoSzMo of the MoIII–CN–VII pair commutes with the operator SzMo + SzV of the total spin projection on the pentagonal z axis of [MoIII(CN)7]4−. Thus, the projection of the total spin MS(Mo) + MS(V) of the MoIII–CN–VII pair on the z axis is a good quantum number, while the total spin SMo + SV is not such due to strong first-order spin-orbit coupling on MoIII. Therefore, regardless of the actual symmetry of the MoIII–CN–VII pair (which is generally low), the effective anisotropic spin Hamiltonian Heff always exhibit uniaxial symmetry, i.e., −Jxy(SMoxSVx + SMoySVy) − JzSMozSVz (see also ref. [47] for more detail).
For the regular (D5h) or moderately distorted structure of [MoIII(CN)7]4−, Heff is calculated by equating the matrix elements of Heff and A + RSMoSV in the space of wave functions |m, MS> = φ(m) × |SV,MS>, <m, MS|Heff|m′, MS′> = <m, MS|A + RSMoSV|m′, MS′>. This results in
H eff = C J x y ( S Mo x S V x + S Mo y S V y ) J z S Mo z S V z ,
where Jz = (J1 + J2)/2, Jxy = (J1J2)/2, and |J1| ≥ |J2|; here, C = (A11 + A22)/2 is a constant stemming from the spin-independent orbital operator A in Equation (1); it can be neglected in further calculations. It is also important to note that the uniaxial symmetry of Heff retains even for moderately distorted PBP [MoIII(CN)7]4− complexes, which frequently occur in many [MoIII(CN)7]4− based molecular magnets [64,65,66]. Indeed, the SOC operator remains diagonal (ζ4dLzMoSzMo) in the space of wave functions 2Φ±(Mo) × 4A2(V), provided that the distortions of the PBP structure of [MoIII(CN)7]4− slightly admix its high-lying excited 4d3 states with ML = ±2 and ML = 0 to the ground state 2Φ(ML = ±1); in fact, due to large energy gap (~20,000 cm−1 >> δ, Figure 1), this holds true even for pronounced departures from the strict (D5h) PBP geometry of the [MoIII(CN)7]4− complex. At the same time, without violating the uniaxial character of the spin Hamiltonian Heff in Equation (2), distortions tend to quench the orbital momentum of MoIII and reduce the relative anisotropy of the −Jxy(SMoxSVx + SMoySVy) − JzSMozSVz spin Hamiltonian (which is measured by the Jz/Jxy ratio), especially when δ > ζMo.
Thus, the uniaxial symmetry of the anisotropic exchange Hamiltonian Heff is an intrinsic property of [MoIII(CN)7]4− complexes, which is extremely useful for creating the uniaxial symmetry of the magnetic anisotropy of SMM clusters (D < 0, E = 0).

2.3. Estimate of Anisotropic Exchange Parameters

It is important to evaluate anisotropic exchange parameters Jz and Jxy in MoIII–CN–VII cyano-bridged pairs, especially given their strong impact on the barrier, Ueff ~ |JzJxy| (see Figure 1) [47]. The MoIII ion has diffuse and high-energy 4d magnetic orbitals, which may provide strong exchange interactions with high-spin 3d metal ions [54]. Especially intense exchange interactions are expected between [MoIII(CN)7]4− complexes and vanadium(II) ions, which, similarly to MoIII ions, also tend to form strong exchange interactions due to more diffuse 3d orbitals of early transition metal ions [67]. However, data on the exchange parameters in [MoIII(CN)7]4−–VII based molecular magnets are scarce or absent [68]. The magnitude of exchange interactions in cyano-bridged pairs MoIII–CN–M(3d) can be estimated from available experimental data for the related hexacyano complex [MoIII(CN)6]3−, J = −122 cm−1 [57] and J = −228 cm−1 [58] for MoIII–CN–VII cyano-bridged pairs; very large exchange parameters for the MoIII–CN–VII linkages in a Prussian blue type structure were also predicted from a broken symmetry DFT study (J ≈ 350 cm−1) [59]. It is noteworthy, however, that these exchange parameters refer to the spin coupling between high-spin octahedral [MoIII(CN)6]3− complex (SMo = 3/2) and high-spin VII ions (SV = 3/2). For the MoIII–CN–VII pairs involving low-spin [MoIII(CN)7]4− PBP complex (S = 1/2), these exchange parameters should be rescaled for a smaller spin value. Approximately, it can be done in terms of the t and U parameters of the superexchange theory, which are, respectively, electron transfer parameter between magnetic t2-orbitals of Mo and V and metal-to-metal charge-transfer energy. With two active antiferromagnetic pathways 4dxz–CN–3dxz and 4dyz–CN–3dyz for the high-spin ground state 4A2 of [MoIII(CN)6]3−, this gives J = −8t2/9U for the high-spin [MoIII(CN)6]3− complex; given that J = −228 cm−1 [58], we have t2/U = 256.5 cm−1, which corresponds to the electron transfer parameter of t ~ 3500 cm−1 at a typical charge-transfer energy of U = 50,000 cm−1 (6 eV). In the low-spin [MoIII(CN)7]4− PBP complex, only one active superexchange pathway lefts for each of two orbital states 2Φxz and 2Φyz (4dxz–CN–3dxz and 4dyz–CN–3dyz, respectively), which results in two equal orbital exchange parameters J1 = J2 = −4t2/3U for the linear apical MoIII–CN–VII pairs and J1 = 4t2/3U, J2 ~ 0 for the equatorial pairs (see ref. [47] for more detail). Therefore, in the linear MoIII–CN–VII pair, the orbital exchange parameters for the low-spin MoIII ion are approximately 1.5 times larger than that for the high-spin MoIII, i.e., 342 cm−1 vs. 228 cm−1. With the equations for the anisotropic exchange parameters, Jz = (J1 + J2)/2 and Jxy = (J1J2)/2, we have antiferromagnetic Ising spin coupling for the apical pair, Jz = 342 cm−1, Jxy = 0, and isotropic spin coupling Jz = Jxy = −171 cm−1 for the equatorial pairs. However, in reality, the pairs are distorted and thus generally J1 is not equal to J2 in the apical pairs [45,47]. In further model calculations, we can conventionally assume J1 = 350, J2 = 250 cm−1 for the orbital exchange parameters in distorted apical pairs MoIII–CN–VII, which corresponds to the anisotropic exchange parameters Jz = −300 and Jxy = −50 cm−1 in Equation (2). Similarly, for equatorial pairs, the isotropic exchange parameter (J = Jz = Jxy) can be roughly rounded to −150 cm−1. More quantitative calculations of the exchange parameters in [MoIII(CN)7]4− based systems are beyond the scope of this article; they will the subject of a separate study.

2.4. Spin Energy Spectra of MoIIIkVIIm Clusters: Engineering of High Energy Barrier

This Section presents results of a comparative study of spin energy diagrams (E vs. MS) for polynuclear clusters MoIIIkVIIm (k = 1–4, m = 1–5) with highly anisotropic magnetic interactions. These clusters are composed of alternating cyano-bridged [MoIII(CN)7]4− and V(II) complexes (Figure 3, Figure 4, Figure 5 and Figure 6) and involve variable number of apical and equatorial MoIII–CN–VII groups featuring, respectively, anisotropic (Ising-type) and isotropic exchange interactions. This study aimed to investigate the influence of the composition, structure and topology of MoIIIkVIIm clusters on the spin energy diagrams and, especially, on the height of the barrier Ueff. In this respect, it is especially interesting to investigate the dependence of the barrier on the number n of apical groups with Ising interactions, which are the main source of molecular magnetic anisotropy in MoIIIkVIIm clusters. For this purpose, the spin energy spectra of the MoIIIkVIIm clusters are calculated in terms of a spin Hamiltonian
H ^ = < i j > ( J x y ( i j ) ( S M o ( i ) x S V ( j ) x + S M o ( i ) y S V ( j ) y ) + J z ( i j ) S M o ( i ) z S V ( j ) z ) ,
where the sum <ij> runs over all Mo(i)–CN–V(j) pairs in the cluster; note that, since the Mo–Mo and V–V exchange interactions are neglected, here i and j indexes numerate, respectively, Mo and V centers. According to the estimates in Section 2.3, the exchange parameters are set to Jxy = −50 and Jz = −300 cm−1 for apical pairs and to Jxy = Jz = −150 cm−1 for equatorial pairs.
Primary calculations were performed for MoIIIVIIm clusters composed of the single central [MoIII(CN)7]4− complex and several VII complexes attached in the apical and equatorial positions (Figure 3 and Figure 4). It is important to note that in these clusters the projection of the total spin MS onto the z-axis (which is parallel to the pentagonal axis of the [MoIII(CN)7]4− complex) is a good quantum number due to uniaxial symmetry of the spin Hamiltonian (3), which commutes with the Sz operator of the total spin. As a result, each spin quantum state has a definite MS spin projection on the polar z-axis of [MoIII(CN)7]4− and thus spin energy spectra can be visualized in terms of the E vs. MS diagrams. Figure 3 shows spin energy diagrams calculated for MoIIIVIIm clusters with one apical pair and progressively increasing number of equatorial pairs (m = 1–5).
In all cases (Figure 3a–d), the spin energy diagrams show a butterfly-shaped figure with a double-well profile for the lower part of the energy spectrum with doubly degenerate ground energy level represented by two spin quantum states MS = +S and MS = −S, which are separated by the barrier Ueff. Note that the spin energy diagrams of MoIIIVIIm clusters differ considerably from the conventional double-well pattern DS2 in ordinary SMMs. In this case, similar to the parabolic energy profile DS2, the value of the barrier Ueff can be associated with the energy position of the lowest spin state with the smallest spin projection (i.e., MS = 0 or MS = ±1/2 for clusters with integer and half-integer spin, respectively, Figure 3 and Figure 4). There is an interesting feature that in the MoIIIVIIm clusters the barrier Ueff is rather insensitive to the number m of the equatorial Mo–CN–V pairs (with some exceptions for the smallest cluster MoIIIVII2, Figure 3a). This can be explained by the fact that there is the only (constant) source of the magnetic anisotropy coming from the single apical pair, while addition of the equatorial pairs with isotropic exchange interactions add nothing to the overall magnetic anisotropy D and to the barrier Ueff. On the other hand, albeit the barrier varies only slightly with the increasing number of equatorial MoIII–CN–VII pairs, the ground-state spin S increases significantly, which improves the overall SMM performance due to larger separation between the two lowest spin states |MS = ±S> in the MS scale.
Addition of the second apical group in the MoIIIVIIm clusters increases the barrier approximately twice (from ca. 150 to 300 cm−1, Figure 4a–d), but the overall picture remains the same: the E vs. MS diagrams retain a butterfly-shaped pattern with a double-well potential, the barrier Ueff varies rather little with increasing number m of the equatorial groups and the ground-state spin MS increases considerably (from 5/2 to 7, Figure 4). This suggests that each apical group MoIII–CN–VII contributes additively to the barrier Ueff.
Next, important information on the variation of the spin energy diagrams and barrier Ueff on the structure of the MoIIIkVIIm clusters was obtained from comparative calculations for the series of MoIII2VII4 clusters with the central MoIII2VII2 square and two side VII complexes connected via various apical and equatorial positions to the [MoIII(CN)7]4− complexes (Figure 5). In these calculations, an idealized structure of MoIIIkVIIm clusters is adopted with the parallel orientation of the local pentagonal axes of two [MoIII(CN)7]4− complexes to ensure uniaxial symmetry of the total spin Hamiltonian in Equation (3). In this case, z-projection MS of the total spin is good quantum number, so each quantum spin state has a definite value of MS.
These calculations show that all the MoIII2VII4 clusters have a double-well potential with the MS = ±5 ground spin state and a large barrier Ueff, (up to ~500 cm−1, Figure 5b) which correlates directly with the number n of apical groups; in fact, Ueff is roughly proportions to n (Figure 5). Similar regularities were also found for larger MoIII4VII4 clusters with cubic and ladder structures (Figure 6). Thus, in the ladder cluster MoIII4VII4 with six apical Mo–CN–V groups (n = 6), the barrier reaches a value of Ueff = 741 cm−1. Consequently, these results indicate the possibility of further increase of the barrier in larger MoIIIkVIIm clusters with alternating [MoIII(CN)7]4− and VII units, in which the number n of apical groups may be considerably larger.
To establish a more quantitative correlation between Ueff and anisotropic exchange parameters, numerous calculations were performed for various MoIIIkVIIm clusters with variable exchange parameters Jz and Jxy for the apical and equatorial MoIII–CN–VII groups. Analysis of these results has revealed that the barrier is approximately given by Ueff ~ 0.5|JzJxy|n, provided that the isotropic exchange interactions (i.e., with Jz = Jxy) for the equatorial pairs are strong enough as compared to the anisotropic Ising-type exchange interactions in the apical pairs (Jz > Jxy). The meaning of this condition is that the isotropic exchange interactions of equatorial pairs integrate the local magnetic anisotropies of the apical pairs into the overall magnetic anisotropy of the MoIIIkVIIm cluster, and thus they must be strong enough to carry out this function. Fortunately, for the specific cyano-bridged MoIIIkVIIm clusters, this condition is well satisfied due to the relation between anisotropic exchange parameters (Jz, Jxy) and orbital exchange parameters J1, J2 involved in the orbitally-dependent spin Hamiltonian (2), Jz = (J1 + J2)/2 and Jxy = (J1J2)/2.
These data indicate the possibility of the direct scaling of the barrier Ueff in molecular cyano-bridged clusters based on alternating [MoIII(CN)7]4− and VII complexes by means of increasing the cluster size and proper organization of its structure and composition. In principle, from the ratio Ueff ~ 0.5|JzJxy|n with |JzJxy| = 200–300 cm−1, one can expect to obtain a barrier above 1000 cm−1 in large MoIIIkVIIm clusters. The main conditions for obtaining a high barrier are a large number n of apical groups and large absolute values of the anisotropic exchange parameters Jz and Jxy. However, it is important to note that, in the large polynuclear MoIIIkVIIm clusters with a fixed number of apical groups n, the maximum barrier Ueff could be obtained only if there are a sufficient number of equatorial groups in the cluster. Thus, the equatorial MoIII–CN–VII groups carry out an important function in the formation of a high barrier: although their isotropic exchange interactions do not directly contribute to the magnetic anisotropy of the MoIIIkVIIm cluster, they ensure the overall magnetic connectivity of the molecular spin cluster and thus help to convert the local Ising-type exchange anisotropy −Jxy(SixSjx + SiySjy) − JzSizSjz of individual apical pairs MoIII–CN–VII into the value of the barrier Ueff.
A unique feature of the PBP [MoIII(CN)7]4− complexes as molecular building blocks lies in the fact that their uniaxial anisotropic exchange interactions −Jxy(SixSjx + SiySjy) − JzSizSjz automatically create a molecular magnetic anisotropy D with a purely axial symmetry, without undesirable transverse component (E = 0), which is very important for obtaining a high barrier Ueff. In other words, the symmetry of the molecular magnetic anisotropy is higher than the geometric symmetry of the molecule. This simplifies drastically the problem of molecular design of nanomagnets, since it eliminates the need to obtain the axial geometric symmetry of a SMM cluster during its chemical synthesis.
However, it is important to note that this study does not fully address the problem of designing high-temperature SMMs—its goal is to demonstrate theoretically a great potential of the new approach in engineering high energy barriers Ueff, which exploits highly anisotropic exchange interactions of PBP [MoIII(CN)7]4− complexes. Of course, many complicated features of these systems (such as distortions, non-collinear orientation of magnetic axes, specific molecular building blocks and structures, synthetic approaches, etc.) remain so far beyond the limits of this theoretical model. Thus, the paper discusses some new general principles for constructing a high barrier rather than specific recommendations for the synthesis of advanced SMM. Practical development of this strategy will be based on new PBP 4d and 5d complexes with unquenched orbital momentum (instead of [MoIII(CN)7]4− and [ReIV(CN)7]3− heptacyanide complexes), in which the pentagonal coordination of the metal ion in the equatorial plane is enforced by a five-membered chelate ring of the planar pentadentate macrocyclic ligands (see our recent paper [69]); this work is underway.

3. Method and Computation Details

Spin energy spectra of MoIIIkVIIm clusters were calculated using specially designed computational approaches and routines based of anisotropic spin Hamiltonian in Equation (3). More specifically, a Fortran routine first calculates the set of matrix elements of the spin exchange Hamiltonian (3) in the full basis set of multicenter spin functions generated by the product of single-ion spin functions of Mo(III) and V(II) ions involved in the MokVm cluster. Next, the energies of spin states and eigenvectors are obtained by diagonalization of the full matrix of the spin Hamiltonian (3); then, the eigenvectors are sorted according their E energies and MS spin projections to visualize the spin energy diagrams shown in Figure 3, Figure 4, Figure 5 and Figure 6.

4. Conclusions

In summary, anisotropic Ising-type exchange infractions −Jxy(SixSjx + SiySjy) − JzSizSjz with large exchange parameters Jz and Jxy generated by PBP [MoIII(CN)7]4− complexes and VII ions connected via apical positions represent a very efficient tool for constructing high-performance SMMs. It is shown that in mixed 4d–3d heterometallic SMM clusters composed of alternating [MoIII(CN)7]4− and VII complexes the spin-reversal barrier is controlled by anisotropic exchange parameters and the number n of apical MoIII–CN–VII groups, Ueff ~ 0.5|JzJxy|n. This finding provides a very efficient straightforward strategy for scaling Ueff and TB values by means of enhancing exchange parameters J and Jxy and increasing the number of PBP [MoIII(CN)7]4− complexes in a SMM molecule. Some ideas on molecular engineering high energy barrier Ueff and further developments toward high-T SMMs are discussed, which are illustrated in Figure 3, Figure 4, Figure 5 and Figure 6. Practical implementation of this approach will be focused on a new family of PBP 4d and 5d complexes with pentadentate macrocyclic ligands [69].

Acknowledgments

This work was partially supported by the Russian Foundation for Basic Research (Project 18-03-00543a) and partially supported by the Federal Agency of Scientific Organizations (Agreement No 007-GZ/CH3363/26) in part of development of computational methods.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Sessoli, R.; Gatteschi, D.; Caneschi, A.; Novak, M.A. Magnetic bistability in a metal-ion cluster. Nature 1993, 365, 141–143. [Google Scholar] [CrossRef]
  2. Sessoli, R.; Tsai, H.L.; Schake, A.R.; Wang, S.Y.; 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]
  3. Gatteschi, D.; Sessoli, R. Quantum tunneling of magnetization and related phenomena in molecular materials. Angew. Chem. Int. Ed. 2003, 42, 268–297. [Google Scholar] [CrossRef] [PubMed]
  4. Coulon, C.; Miyasaka, H.; Clérac, R. Single-Molecule Magnets and Related Phenomena. In Structure and Bonding; Winpenny, R., Ed.; Springer: Berlin, Germany, 2006; Volume 122. [Google Scholar]
  5. Gatteschi, D.; Sessoli, R.; Villain, J. Molecular Nanomagnets; Oxford University Press: Oxford, UK, 2006. [Google Scholar]
  6. Gao, S. Molecular Nanomagnets and Related Phenomena; Springer: New York, NY, USA, 2015; Volume 164. [Google Scholar]
  7. Milios, C.J.; Winpenny, R.E.P. Cluster-based single-molecule magnets. In Molecular Nanomagnets and Related Phenomena; Gao, S., Ed.; Springer: Berlin, Germany, 2015; Volume 164, pp. 1–109. [Google Scholar]
  8. Ardavan, A.; Blundell, S.J. Storing quantum information in chemically engineered nanoscale magnets. J. Mater. Chem. 2009, 19, 1754–1760. [Google Scholar] [CrossRef]
  9. Leuenberger, M.N.; Loss, D. Quantum computing in molecular magnets. Nature 2001, 410, 789–793. [Google Scholar] [CrossRef] [PubMed]
  10. Aromi, G.; Aguilà, D.; Gamez, P.; Luis, F.; Roubeau, O. Design of magnetic coordination complexes for quantum computing. Chem. Soc. Rev. 2012, 41, 537–546. [Google Scholar] [CrossRef] [PubMed]
  11. Wedge, C.J.; Timco, G.A.; Spielberg, E.T.; George, R.E.; Tuna, F.; Rigby, S.; McInnes, E.J.L.; Winpenny, R.E.P.; Blundell, S.J.; Ardavan, A. Chemical engineering of molecular qubits. Phys. Rev. Lett. 2012, 108, 107204. [Google Scholar] [CrossRef] [PubMed]
  12. Pedersen, K.S.; Ariciu, A.; McAdams, S.; Weihe, H.; Bendix, J.; Tuna, F.; Piligkos, S. Toward Molecular 4f Single-Ion Magnet Qubits. J. Am. Chem. Soc. 2016, 138, 5801–5804. [Google Scholar] [CrossRef] [PubMed]
  13. Rocha, A.R.; García-Suarez, V.M.; Bailey, S.W.; Lambert, C.J.; Ferrer, J.; Sanvito, S. Towards Molecular Spintronics. Nat. Mater. 2005, 4, 335–339. [Google Scholar] [CrossRef] [PubMed]
  14. Bogani, L.; Wernsdorfer, W. Molecular spintronics using single-molecule magnets. Nat. Mater. 2008, 7, 179–186. [Google Scholar] [CrossRef] [PubMed]
  15. Waldmann, O. A criterion for the anisotropy barrier in single-molecule magnets. Inorg. Chem. 2007, 46, 10035–10037. [Google Scholar] [CrossRef] [PubMed]
  16. 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]
  17. Neese, F.; Pantazis, D.A. What is not required to make a single molecule magnet. Faraday Discuss. 2011, 148, 229–238. [Google Scholar] [CrossRef] [PubMed]
  18. Sessoli, R.; Powell, A.K. Strategies towards single molecule magnets based on lanthanide ions. Coord. Chem. Rev. 2009, 253, 2328–2341. [Google Scholar] [CrossRef]
  19. Sorace, L.; Benelli, J.R.; Gatteschi, D. Lanthanides in molecular magnetism: old tools in a new field. Chem. Soc. Rev. 2011, 40, 3092–3105. [Google Scholar] [CrossRef] [PubMed]
  20. Rinehart, J.D.; Long, J.R. Exploiting single-ion anisotropy in the design of f-element single-molecule magnets. Chem. Sci. 2011, 2, 2078–2085. [Google Scholar] [CrossRef]
  21. Layfield, R.A.; Murugesu, M. Lanthanides and Actinides in Molecular Magnetism; Wiley-VCH: Weinheim, Germany, 2015. [Google Scholar]
  22. Ishikawa, N.; Sugita, M.; Ishikawa, T.; Koshihara, S.; 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]
  23. Luzon, J.; Sessoli, R. Lanthanides in molecular magnetism: So fascinating, so challenging. Dalton Trans. 2012, 41, 13556–13567. [Google Scholar] [CrossRef] [PubMed]
  24. Woodruff, D.N.; Winpenny, R.E.P.; Layfield, R.A. Lanthanide Single-Molecule Magnets. Chem. Rev. 2013, 113, 5110–5148. [Google Scholar] [CrossRef] [PubMed]
  25. Layfield, R.A. Organometallic single-molecule magnets. Organometallics 2014, 33, 1084–1099. [Google Scholar] [CrossRef]
  26. Tang, J.; Zhang, P. Lanthanide Single-Molecule Magnets; Springer: Berlin, Germany, 2015. [Google Scholar]
  27. Liddle, S.T.; van Slageren, J. Improving f-element single molecule magnets. Chem. Soc. Rev. 2015, 44, 6655–6669. [Google Scholar] [CrossRef] [PubMed]
  28. Habib, F.; Murugesu, M. Lessons learned from dinuclear lanthanide nano-magnets. Chem. Soc. Rev. 2013, 42, 3278–3288. [Google Scholar] [CrossRef] [PubMed]
  29. Rinehart, J.D.; Fang, M.; Evans, W.J.; Long, J.R. A N23-Radical-Bridged Terbium Complex Exhibiting Magnetic Hysteresis at 14 K. J. Am. Chem. Soc. 2011, 133, 14236–14239. [Google Scholar] [CrossRef] [PubMed]
  30. 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] [PubMed]
  31. Ding, Y.S.; Chilton, N.F.; Winpenny, R.E.P.; 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] [PubMed]
  32. Guo, F.S.; Day, B.M.; Chen, Y.C.; Tong, M.L.; Mansikkamäki, A.; Layfield, R.A. A dysprosium metallocene single-molecule magnet functioning at the axial limit. Angew. Chem. Int. Ed. 2017, 56, 11445–11449. [Google Scholar] [CrossRef] [PubMed]
  33. Goodwin, C.A.P.; 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] [PubMed]
  34. 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] [PubMed]
  35. Craig, G.A.; Murrie, M. 3D single-ion magnets. Chem. Soc. Rev. 2015, 44, 2135–2147. [Google Scholar] [CrossRef] [PubMed]
  36. Frost, J.M.; Harriman, K.L.M.; Murugesu, M. The rise of 3-d single-ion magnets in molecular magnetism: Towards materials from molecules. Chem. Sci. 2016, 7, 2470–2491. [Google Scholar] [CrossRef] [PubMed]
  37. Zadrozny, J.M.; Xiao, D.J.; Long, J.R.; Atanasov, M.; Neese, F.; Grandjean, F.; Long, G.J. Mossbauer Spectroscopy as a Probe of Magnetization Dynamics in the Linear Iron(I) and Iron(II) Complexes [Fe(C(SiMe3)3)2]1−/0. Inorg. Chem. 2013, 52, 13123–13131. [Google Scholar] [CrossRef] [PubMed]
  38. Yao, X.-N.; Du, J.-Z.; Zhang, Y.-Q.; Leng, X.-B.; Yang, M.-W.; Jiang, S.-D.; Wang, Z.-X.; Ouyang, Z.-W.; Deng, L.; Wang, B.-W.; et al. Two-Coordinate Co(II) Imido Complexes as Outstanding Single-Molecule Magnets. J. Am. Chem. Soc. 2017, 139, 373–380. [Google Scholar] [CrossRef] [PubMed]
  39. Bar, A.K.; Pichon, C.; Sutter, J.-P. Magnetic anisotropy in two- to eight-coordinated transition–metal complexes: Recent developments in molecular magnetism. Coord. Chem. Rev. 2016, 308, 346–380. [Google Scholar] [CrossRef]
  40. Ruamps, R.; Batchelor, L.J.; Guillot, R.; Zakhia, G.; Barra, A.L.; Wernsdorfer, W.; Guihéry, N.; Mallah, T. Ising-type magnetic anisotropy and single molecule magnet behaviour in mononuclear trigonal bipyramidal Co(II) complexes. Chem. Sci. 2014, 5, 3418–3424. [Google Scholar] [CrossRef]
  41. Ruamps, R.; Maurice, R.; Batchelor, L.J.; Boggio-Pasqua, M.; Guillot, R.; Barra, A.L.; Liu, J.; Bendeif, E.-E.; Pillet, S.; Hill, S.; et al. Giant Ising-Type Magnetic Anisotropy in Trigonal Bipyramidal Ni(II) Complexes: Experiment and Theory. J. Am. Chem. Soc. 2013, 135, 3017–3026. [Google Scholar] [CrossRef] [PubMed]
  42. Novikov, V.V.; Pavlov, A.A.; Nelyubina, Y.V.; Boulon, M.-E.; Varzatskii, O.A.; Voloshin, Y.Z.; Winpenny, R.E.P. A Trigonal Prismatic Mononuclear Cobalt(II) Complex Showing Single-Molecule Magnet Behavior. J. Am. Chem. Soc. 2015, 137, 9792–9795. [Google Scholar] [CrossRef] [PubMed]
  43. Morrison, C.A.; Leavitt, R.P. Spectroscopic Properties of Triply Ionized Lanthanides in Transparent Host Crystals. In Handbook on the Physics and Chemistry of Rare Earths; Elsevier: Amsterdam, The Netherlands, 1982; Volume 5, pp. 461–692. [Google Scholar]
  44. Ungur, L.; Chibotaru, L.F. Strategies toward High-Temperature Lanthanide-Based Single-Molecule Magnets. Inorg. Chem. 2016, 55, 10043–10056. [Google Scholar] [CrossRef] [PubMed]
  45. Mironov, V.S.; Chibotaru, L.F.; Ceulemans, A. Mechanism of a Strongly Anisotropic MoIII–CN–MnII Spin–Spin Coupling in Molecular Magnets Based on the [Mo(CN)7]4− Heptacyanometalate: A New Strategy for Single-Molecule Magnets with High Blocking Temperatures. J. Am. Chem. Soc. 2003, 125, 9750–9760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Mironov, V.S. New approaches to the problem of high-temperature single-molecule magnets. Dokl. Phys. Chem. 2006, 408, 130–136. [Google Scholar] [CrossRef]
  47. Mironov, V.S. Origin of Dissimilar Single-Molecule Magnet Behavior of Three MnII2MoIII Complexes Based on [MoIII(CN)7]4− Heptacyanomolybdate: Interplay of MoIII–CN–MnII Anisotropic Exchange Interactions. Inorg. Chem. 2015, 54, 11339–11355. [Google Scholar] [CrossRef] [PubMed]
  48. Bennett, M.V.; Long, J.R. New Cyanometalate Building Units: Synthesis and Characterization of [Re(CN)7]3− and [Re(CN)8]3−. J. Am. Chem. Soc. 2003, 125, 2394–2395. [Google Scholar] [CrossRef] [PubMed]
  49. Samsonenko, D.G.; Paulsen, C.; Lhotel, E.; Mironov, V.S.; Vostrikova, K.E. [MnIII(Schiff base)]3[ReIV(CN)7], highly anisotropic 3D coordination framework: Synthesis, crystal structure, magnetic investigations, and theoretical analysis. Inorg. Chem. 2014, 53, 10217–10231. [Google Scholar] [CrossRef] [PubMed]
  50. Bendix, J.; Steenberg, P.; Sotofte, I. Isolation and Molecular Structure of Hexacyanoruthenate(III). Inorg. Chem. 2003, 42, 4510–4512. [Google Scholar] [CrossRef] [PubMed]
  51. Albores, P.; Slep, L.D.; Baraldo, L.M.; Baggio, R.; Garland, M.T.; Rentschler, E. Crystal Structure and Electronic and Magnetic Properties of Hexacyanoosmate(III). Inorg. Chem. 2006, 45, 2361–2363. [Google Scholar] [CrossRef] [PubMed]
  52. Mironov, V.S. Trigonal bipyramidal spin clusters with orbitally degenerate 5d cyano complexes [OsIII(CN)6]3–, prototypes of high-temperature single-molecule magnets. Dokl. Phys. Chem. 2007, 415, 199–204. [Google Scholar] [CrossRef]
  53. Dreiser, J.; Pedersen, K.S.; Schnegg, A.; Holldack, K.; Nehrkorn, J.; Sigrist, M.; Tregenna-Piggott, P.; Mutka, H.; Weihe, H.; Mironov, V.S.; et al. Three-Axis Anisotropic Exchange Coupling in the Single-Molecule Magnets NEt4[MnIII2(5-Brsalen)2(MeOH)2MIII(CN)6] (M = Ru, Os). Chem. Eur. J. 2013, 19, 3693–3701. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, X.-Y.; Avendaño, C.; Dunbar, K.R. Molecular magnetic materials based on 4d and 5d transition metals. Chem. Soc. Rev. 2011, 40, 3213–3238. [Google Scholar] [CrossRef] [PubMed]
  55. 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]
  56. Wu, D.-Q.; Shao, D.; Wei, X.-Q.; Shen, F.-X.; Shi, L.; Kempe, D.; Zhang, Y.-Z.; Dunbar, K.R.; Wang, X.-Y. Reversible On–Off Switching of a Single-Molecule Magnet via a Crystal-to-Crystal Chemical Transformation. J. Am. Chem. Soc. 2017, 139, 11714–11717. [Google Scholar] [CrossRef] [PubMed]
  57. Freedman, D.E.; Jenkins, D.M.; Long, J.R. Strong magnetic exchange coupling in the cyano-bridged coordination clusters [(PY5Me2)4V4M(CN)6]5+ (M = Cr, Mo). Chem. Commun. 2009, 4829–4831. [Google Scholar] [CrossRef] [PubMed]
  58. Pinkowicz, D.; Southerland, H.; Wang, X.-Y.; Dunbar, K.R. Record Antiferromagnetic Coupling for a 3d/4d Cyanide-Bridged Compound. J. Am. Chem. Soc. 2014, 136, 9922–9924. [Google Scholar] [CrossRef] [PubMed]
  59. Ruiz, E.; Rodríguez-Fortea, A.; Alvarez, S.; Verdaguer, M. Is it possible to get high Tc magnets with prussian blue analogues? A theoretical prospect. Chem. Eur. J. 2005, 11, 2135–2144. [Google Scholar] [CrossRef] [PubMed]
  60. Gunter, M.J.; Berry, K.J.; Murray, K.S. A Model for the Cyanide Form of Oxidized Cytochrome Oxidase: An Iron(III)/Copper(II) Porphyrin Complex Displaying Ferromagnetic Coupling. J. Am. Chem. Soc. 1984, 106, 4227–4235. [Google Scholar] [CrossRef]
  61. Langley, S.K.; Wielechowski, D.P.; Vieru, V.; Chilton, N.F.; Moubaraki, B.; Abrahams, B.F.; Chibotaru, L.F.; Murray, K.S. A {CrIII2DyIII2} Single-Molecule Magnet: Enhancing the Blocking Temperature through 3d Magnetic Exchange. Angew. Chem. Int. Ed. 2013, 52, 12014–12019. [Google Scholar] [CrossRef] [PubMed]
  62. Langley, S.K.; Wielechowski, D.P.; Vieru, V.; Chilton, N.F.; Moubaraki, B.; Chibotaru, L.F.; Murray, K.S. Modulation of slow magnetic relaxation by tuning magnetic exchange in {Cr2Dy2} single molecule magnets. Chem. Sci. 2014, 5, 3246–3256. [Google Scholar] [CrossRef]
  63. Hursthouse, M.B.; Maijk, K.M.A.; Soares, A.M.; Gibson, J.F.; Griffith, W.P. The X-ray crystal structure of NaK3[Mo(CN)7]·2H2O and the structure of its anion in aqueous solution. Inorg. Chim. Acta 1980, 45, L81–L82. [Google Scholar] [CrossRef]
  64. Larionova, J.; Clérac, R.; Sanchiz, J.; Kahn, O.; Golhen, S.; Ouahab, L. Ferromagnetic Ordering, Anisotropy, and Spin Reorientation for the Cyano-Bridged Bimetallic Compound Mn2(H2O)5Mo(CN)7·4H2O (α Phase). J. Am. Chem. Soc. 1998, 120, 13088–13095. [Google Scholar] [CrossRef]
  65. Larionova, J.; Kahn, O.; Gohlen, S.; Ouahab, L.; Clérac, R. Structure, Ferromagnetic Ordering, Anisotropy, and Spin Reorientation for the Two-Dimensional Cyano-Bridged Bimetallic Compound K2Mn3(H2O)6[Mo(CN)7]2·6H2O. J. Am. Chem. Soc. 1999, 121, 3349–3356. [Google Scholar] [CrossRef]
  66. Wang, Q.-L.; Southerland, H.; Li, J.-R.; Prosvirin, A.V.; Zhao, H.; Dunbar, K.R. Crystal-to-Crystal Transformation of Magnets Based on Heptacyanomolybdate(III) Involving Dramatic Changes in Coordination Mode and Ordering Temperature. Angew. Chem. Int. Ed. 2012, 51, 9321–9324. [Google Scholar] [CrossRef] [PubMed]
  67. Holmes, S.M.; Girolami, G.S. Sol–Gel Synthesis of KVII[CrIII(CN)6]·2H2O: A Crystalline Molecule-Based Magnet with a Magnetic Ordering Temperature above 100 °C. J. Am. Chem. Soc. 1999, 121, 5593–5594. [Google Scholar] [CrossRef]
  68. Tomono, K.; Tsunobuchi, Y.; Nakabayashi, K.; Ohkoshi, S. Vanadium(II) Heptacyanomolybdate(III)-Based Magnet Exhibiting a High Curie Temperature of 110 K. Inorg. Chem. 2010, 49, 1298–1300. [Google Scholar] [CrossRef] [PubMed]
  69. Mironov, V.S.; Bazhenova, T.A.; Manakin, Y.V.; Lyssenko, K.A.; Talantsev, A.D.; Yagubskii, E.B. A new Mo(IV) complex with the pentadentate (N3O2) Schiff-base ligand: the first non-cyanide pentagonal–bipyramidal paramagnetic 4d complex. Dalton Trans. 2017, 46, 14083–14087. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Origin of SMM behavior of MnII–MoIII–MnII clusters [55,56]. Anisotropic exchange interactions in two apical groups Mo–CN–Mn are described by the uniaxial spin Hamiltonian Heff = −Jxy(SixSjx + SiySjy) − JzSizSjz of the Ising type (|Jz|> |Jxy|), which result in a double-well energy profile with the ground doubly degenerate quantum spin states MS = +9/2 and MS = −9/2 separated by the energy barrier Ueff. The latter is controlled by the anisotropic exchange parameters, Ueff ≈ 2|JzJxy| [47].
Figure 1. Origin of SMM behavior of MnII–MoIII–MnII clusters [55,56]. Anisotropic exchange interactions in two apical groups Mo–CN–Mn are described by the uniaxial spin Hamiltonian Heff = −Jxy(SixSjx + SiySjy) − JzSizSjz of the Ising type (|Jz|> |Jxy|), which result in a double-well energy profile with the ground doubly degenerate quantum spin states MS = +9/2 and MS = −9/2 separated by the energy barrier Ueff. The latter is controlled by the anisotropic exchange parameters, Ueff ≈ 2|JzJxy| [47].
Inorganics 06 00058 g001
Figure 2. Ground-state electronic structure of the PBP [MoIII(CN)7]4− complex (D5h). The ligand-field energy splitting diagram of 4d orbitals of MoIII is indicated. Three 4d electrons of MoIII occupy the two lowest degenerate orbitals 4dxz and 4dyz to produce a low-spin orbital doublet 2Фxz,yz = 2Ф(ML = ±1) with unquenched orbital momentum ML = ±1. Spin-orbit coupling (ζ4d) on Mo splits the orbital doublet 2Ф(ML = ±1) to produce magnetically anisotropic (Ising-type) ground Kramers doublet φ(±1/2) and excited Kramers doublet χ(±1/2). The orbital composition of the ground-state wave functions φ(±1/2) is shown.
Figure 2. Ground-state electronic structure of the PBP [MoIII(CN)7]4− complex (D5h). The ligand-field energy splitting diagram of 4d orbitals of MoIII is indicated. Three 4d electrons of MoIII occupy the two lowest degenerate orbitals 4dxz and 4dyz to produce a low-spin orbital doublet 2Фxz,yz = 2Ф(ML = ±1) with unquenched orbital momentum ML = ±1. Spin-orbit coupling (ζ4d) on Mo splits the orbital doublet 2Ф(ML = ±1) to produce magnetically anisotropic (Ising-type) ground Kramers doublet φ(±1/2) and excited Kramers doublet χ(±1/2). The orbital composition of the ground-state wave functions φ(±1/2) is shown.
Inorganics 06 00058 g002
Figure 3. Energy diagrams of the spin levels (E vs. MS) of (a) MoIIIVII2, (b) MoIIIVII3, (c) MoIIIVII4, and (d) MoIIIVII5 clusters with single apical group Mo–CN–V. Spin energy spectra exhibit a double-well character with doubly degenerate ±MS ground spin state, which is inherent in single-molecule magnets. Spin energy spectra are calculated using the anisotropic spin Hamiltonian Heff = −Jxy(SixSjx + SiySjy) − JzSizSjz with the exchange parameters indicated in the text (Jz = −300 cm−1 and Jxy = −50 cm−1 for the apical Mo–CN–V pair and Jz = Jxy = −150 cm−1 for equatorial pairs).
Figure 3. Energy diagrams of the spin levels (E vs. MS) of (a) MoIIIVII2, (b) MoIIIVII3, (c) MoIIIVII4, and (d) MoIIIVII5 clusters with single apical group Mo–CN–V. Spin energy spectra exhibit a double-well character with doubly degenerate ±MS ground spin state, which is inherent in single-molecule magnets. Spin energy spectra are calculated using the anisotropic spin Hamiltonian Heff = −Jxy(SixSjx + SiySjy) − JzSizSjz with the exchange parameters indicated in the text (Jz = −300 cm−1 and Jxy = −50 cm−1 for the apical Mo–CN–V pair and Jz = Jxy = −150 cm−1 for equatorial pairs).
Inorganics 06 00058 g003
Figure 4. Energy diagrams of the spin levels (E vs. MS) of (a) MoIIIVII2, (b) MoIIIVII3, (c) MoIIIVII4, and (d) MoIIIVII5 clusters with two polar Mo–CN–V groups and variable number of equatorial groups.
Figure 4. Energy diagrams of the spin levels (E vs. MS) of (a) MoIIIVII2, (b) MoIIIVII3, (c) MoIIIVII4, and (d) MoIIIVII5 clusters with two polar Mo–CN–V groups and variable number of equatorial groups.
Inorganics 06 00058 g004
Figure 5. Energy diagrams of the spin levels (E vs. MS) of four MoIII2VII4 clusters with the different number n of apical Mo–CN–V groups, (a) n = 2, (b) n = 4, (c,d) n = 3. The full spin energy diagrams are depicted in the inserts. The height of the barrier Ueff correlates directly with the number n of apical Mo–CN–V groups.
Figure 5. Energy diagrams of the spin levels (E vs. MS) of four MoIII2VII4 clusters with the different number n of apical Mo–CN–V groups, (a) n = 2, (b) n = 4, (c,d) n = 3. The full spin energy diagrams are depicted in the inserts. The height of the barrier Ueff correlates directly with the number n of apical Mo–CN–V groups.
Inorganics 06 00058 g005
Figure 6. Energy diagrams of the spin levels (E vs. MS) of MoIII4VII4 clusters with (a) four (cube, n = 4 and (b) with six (ladder, n = 6) apical Mo–CN–V groups calculated using anisotropic spin Hamiltonian (3) with the exchange parameters indicated in the text.
Figure 6. Energy diagrams of the spin levels (E vs. MS) of MoIII4VII4 clusters with (a) four (cube, n = 4 and (b) with six (ladder, n = 6) apical Mo–CN–V groups calculated using anisotropic spin Hamiltonian (3) with the exchange parameters indicated in the text.
Inorganics 06 00058 g006

Share and Cite

MDPI and ACS Style

Mironov, V.S. Molecular Engineering of High Energy Barrier in Single-Molecule Magnets Based on [MoIII(CN)7]4− and V(II) Complexes. Inorganics 2018, 6, 58. https://doi.org/10.3390/inorganics6020058

AMA Style

Mironov VS. Molecular Engineering of High Energy Barrier in Single-Molecule Magnets Based on [MoIII(CN)7]4− and V(II) Complexes. Inorganics. 2018; 6(2):58. https://doi.org/10.3390/inorganics6020058

Chicago/Turabian Style

Mironov, Vladimir S. 2018. "Molecular Engineering of High Energy Barrier in Single-Molecule Magnets Based on [MoIII(CN)7]4− and V(II) Complexes" Inorganics 6, no. 2: 58. https://doi.org/10.3390/inorganics6020058

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop