Single Molecule Magnets
Transcription
Single Molecule Magnets
第34卷 第3期 2014年6月 物 理 学 进 展 PROGRESS IN PHYSICS Vol.34 No.3 Jun. 2014 Single Molecule Magnets Ren Min, Zheng Li-Min* School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210093, P. R. China Single molecule magnets (SMMs) refer to those molecules that show slow relaxation of magnetization below a characteristic blocking temperature (TB). Since the first report of SMM behavior in [Mn12 O12 (O2 CCH3 )16 (H2 O)4 ] · 4H2 O · 2CH3 CO2 H (Mn12 ) in 1993, great efforts have been devoted to the exploration of new SMMs and their potential applications in the information storage, spintronics and quantum computing etc. A number of mononuclear and polynuclear metal clusters have been discovered to show SMM behaviors. In this review, we focus on the SMMs of oxo-bridged Mn and Fe clusters, 3d single ion magnets and lanthanide-based SMMs. Key words: single molecule magnet; manganese; iron; lanthanide CLC number : O44 Document Code : A relaxation time can be extremely long, reaching years at low temperature (below 2 K),[1] which is reminiscent of bulk magnets. Remarkably, SMMs can also show a stepped magnetic hysteresis due to quantum tunneling of the magnetization (QTM).[2] Therefore, the magnetization relaxation of SMMs obeys two processes depending on the temperature. At high temperature region, the relaxation time τ is thermally activated following the Arrhenius law with an activation energy equal to U . At very low temperatures, QTM becomes the fastest pathway of relaxation which is temperature independent. A crossover occurs experimentally between these two regimes called thermally assisted QTM. In this intermediate range of temperature, the thermal barrier is short-cut by quantum tunneling, and an effective barrier, Ueff , is found smaller than U . In many SMM systems, this regime is the only one seen experimentally before that τ becomes temperature independent. Taking advantage of both the storage capacity and quantum phenomena, SMMs appear promising for the building of future integrated nanodevices, such as high density information storage,[3] quantum computing[4] and magnetic refregeration.[5] CONTENTS I. Introduction 119 II. Single molecule magnets based on 3d metal ions A. Mn-O clusters B. Fe-O clusters C. Single 3d-metal ions 120 120 123 124 III. Single molecule magnets based on 4f metal ions A. Lanthanide phthalocyaninates B. Lanthanide polyoxometalates C. Lanthanide organometallics D. Other mononuclear lanthanide SMMs E. Other polynuclear lanthanide SMMs 125 125 128 128 129 130 IV. Conclusion 131 References 131 I. INTRODUCTION Single molecule magnets (SMMs) refer to those molecules that show slow relaxation of magnetization below a characteristic blocking temperature (TB ). The slow relaxation is caused by a significant energy barrier to magnetization reversal, determined by the combined effect of a high-spin ground state ST and negative uniaxial anisotropy (D < 0) (Figure 1). The upper limit of the barrier (U ) is S 2 |D| or (S 2 −1/4)|D| for integer and half-integer spins, respectively. The intermolecular interactions must be minimal to avoid the long range magnetic ordering. The magnetization Since the first report of SMM behavior in compound [Mn12 O12 (O2 CCH3 )16 (H2 O)4 ]・4H2 O・2CH3 CO2 H (Mn12 ) in 1993,[6] a great effort has been devoted to the preparation and study of new systems with SMM behaviors. These efforts have led to the discovery of slow magnetic relaxation in a number of polynuclear metal clusters,[7] as well as mononuclear lanthanide,[8] actinide[9] and transition metal compounds.[10] Due to the limitation of space, the current review will only focus on the SMMs of oxo-bridged Mn and Fe clusters, 3d single ion magnets and lanthanide-based SMMs. Received date: 2013-2-19 *lmzheng@nju.edu.cn 文章编号: 1000-0542(2014)03-0119-17 119 120 Ren Min et al.: Single Molecule Magnets III 16+ FIG. 1. Left: The [MnIV core of 4 Mn8 (µ-O)12 ] [Mn12 O12 (O2 CCH3 )16 (H2 O)4 ]·4H2 O·2CH3 CO2 H. Color codes: MnIV purple red, MnIII brick red, O yellow. Right: Plot of the potential energy versus the magnetization direction for a SMM with an S = 10 ground state (adapted from ref. [1]) II. SINGLE MOLECULE MAGNETS BASED ON 3D METAL IONS A. Mn-O clusters A large amount of manganese clusters have been reported which show relatively large spin ground states, and large negative D values arising from the JahnTeller (JT) distorted MnIII ions.[11] The largest cluster is Mn84 , showing a SMM below 1.5 K with an energy barrier of 18 K.[12] However, larger cluster with high spin ground state cannot guarantee the higher energy barrier of a SMM. For example, a Mn19 cluster is not a SMM, although it displays a record spin ground state of S = 83/2.[13] The increase of the magnetic anisotropy is more important in raising the energy barrier. The record energy barrier for the magnetization reversal among the 3d metal SMMs is 86 K, held by [MnIII 6 O2 (Etsao)6 (O2 CPh(Me)2 )2 (EtOH)6 ].[14] The second highest barrier is observed in the Mn12 O12 family (up to 71 K). The family of [Mn12 O12 (O2 CR)16 (H2 O)4 ] ([Mn12 ]) is the first and most thoroughly studied SMMs to date.[7b] Table I. gives a list of the [Mn12 ]n− complexes, together with their ground state spin, anisotropic D values and energy barriers. Compound [Mn12 O12 (O2 CCH3 )16 (H2 O)4 ] (Mn12 Ac) contains IV a [MnIII The cluster has an 8 Mn4 (µ3 -O)12 ] core. overall D2d symmetry. The eight MnIII ions define the external octagon, whereas the four MnIV ions correspond to the internal tetrahedron (Fig. 1). It has a spin ground state of S = 10, arising from antiferromagnetic interactions between the S = 3/2 spins of MnIV ions and the S = 2 spins of MnIII ions with a negative axial zero-field splitting (D = −0.50 cm−1 ). Single crystal ac magnetic susceptibility shows that the magnetization of the Mn12 Ac is highly anisotropic with the easy axis parallel to the tetragonal axis of the cluster, arising from the near parallel alignment of the JT axes on the eight MnIII ions. Slow magnetization relaxation is observed with an effective energy barrier of 61 K for Mn12 Ac.[6] Stepped hysteresis loop appears at low temperature, attributing to the quantum tunneling of magnetization (QTM). For many [Mn12 ] clusters, two separate out-of-phase ac susceptibility signals are observed,[15] which can be explained by the presence of Jahn-Teller isomers within the same crystal.[16,17] The fast relaxing species can be suppressed when the symmetry of the cluster is high. Complexes [Mn12 O12 (O2 CCH2 Br)16 (H2 O)4 ]·4CH2 Cl2 (space group I41 /a)[19] and [Mn12 O12 (O2 CCH2 But )16 (MeOH)4 ]·MeOH (space group I − 4)[20] are best examples of [Mn12 ] clusters with high molecular symmetry. They show only one out-of-phase ac susceptibility signal and a “cleaner” hysteresis loop, which is invaluable for detailed QTM and HFEPR studies.[21] The energy barrier of the former increases to 74.4 K.[19] The neutral [Mn12 ] clusters can be reduced to produce one-electron, two-electron, and even three-electron reduced species with general formula [Mn12 O12 (O2 CR)16 (H2 O)x ]n− (n = 0, 1, 2, 3; x = 3, 4), using versatile carboxylate, mixed carboxylate and mixed carboxylate/non-carboxylate ligands. In all cases, the [Mn12 O12 ] core remains essentially the same (Fig. 1). The added electrons go to the peripheral MnIII atoms converting them to MnII , III II − IV III II giving a MnIV 4 Mn7 Mn ([Mn12 ] ), Mn4 Mn6 Mn2 2− IV III II 3− ([Mn12 ] ) and Mn4 Mn5 Mn3 ([Mn12 ] ) oxidation state description, respectively. The reduction of MnIII ions leads to the lowing of the molecular anisotropy, and hence the energy barrier. For example, the ground states and D values for a series of compounds (NPrn4 )z [Mn12 O12 (O2 CCHCl2 )16 (H2 O)4 ]z− ,[38] decreases from 10 and −0.45 cm−1 for the neutral [Mn12 ], to 19/2 and −0.35 cm−1 for [Mn12 ]− , 10 and −0.28 cm−1 for [Mn12 ]2− , and 17/2 and −0.25 cm−1 for [Mn12 ]3− , respectively. The energy barriers also decrease monotonously from 65 K for [Mn12 ], 45 K for [Mn12 ]− , 40 K for [Mn12 ]2− , to 26 K for [Mn12 ]3− . Other important families of Mn-O SMMs include Mn2 , Mn3 , Mn4 and Mn6 clusters. The first Mn2 -based SMM is [Mn(saltmen)(ReO4 )]2 , which contains a N2 O2 Schiff base ligand N,N’(1,1,2,2-tetramethylethylene)bis(salicylideneiminate) (saltmenH2 ) (Fig. 2, left).[43] Ferromagnetic interactions are propagated between the MnIII centers (2J/kB = +5.30 K), leading to an ST = 4 spin ground state with a uniaxial anisotropy (DMn = −4.0 K). Very small antiferromagnetic coupling of ca. −0.2 K was found to present between the dimers. This com- 121 Ren Min et al.: Single Molecule Magnets TABLE I. SMMs based on Mn12 O12 clusters Mn12 O12 [Mn12 O12 (O2 CMe)16 (H2 O)4 ]·2MeCOOH·4H2 O [Mn12 O12 (O2 CC6 H4 F-2)16 (H2 O)4 ] Space ST group 10 10 I41 /a I −4 P ca21 P 21 /c I −4 P 21 /n 9 10 10 10 10 10 10 10 10 10 D Ueff (cm−1 ) (K) -0.50 61 65.2 31.9 n.a 38 n.a 64 -0.44 65.4 -0.33 64.4 -0.38 74.4 -0.46 -0.43 62.6 -0.49 71.2 64.4 -0.42 62.5 τ0 (s) 2.1 × 10−7 2.3 × 10−9 3.0 × 10−10 2.0 × 10−10 7.7 × 10−9 2.4 × 10−9 1.5 × 10−9 3.3 × 10−9 I −4 P 21 /n P −1 C2/c P 2/c P −1 10 10 10 10 19/2 10 -0.65 -0.34 -0.4 -0.29 -0.34 -0.39 1.4 × 10−10 6.3 × 10−11 2.9 × 10−9 3.3 × 10−8 4.9 × 10−9 7.8 × 10−9 3.3 × 10−11 7.6 × 10−9 5.3 × 10−9 6.7 × 10−9 7.7 × 10−9 6.3 × 10−9 2.6 × 10−9 1.6 × 10−10 5.3 × 10−9 6.6 × 10−9 7.4 × 10−9 5.7 × 10−9 6.0 × 10−9 3.8 × 10−9 1.7 × 10−9 25 2.8 × 10−9 34 7.7 × 10−10 1.4 × 10−8 3.2 × 10−11 1.0 × 10−8 3.4 × 10−9 3.0 × 10−10 5.0 × 10−10 n.a. n.a. n.a. n.a. 2.1 × 10−9 3.0 × 10−9 2.3 × 10−9 3.1 × 10−9 1.6 × 10−8 9.0 × 10−9 35 36 36 37 4.7 × 10−9 [Mn12 O12 (O2 CC6 H4 -p-Me)16 (H2 O)4 ]·(HO2 CC6 H4 -p-Me) [Mn12 O12 (O2 CC6 H4 -p-Me)16 (H2 O)4 ]·3H2 O [Mn12 O12 (O2 CCHCHCH3 )16 (H2 O)4 ]·H2 O [Mn12 O12 (O2 CC6 H4 C6 H5 )16 (H2 O)4 ]·2C6 H5 C6 H4 COOH [Mn12 O12 (O2 CCH2 Br)16 (H2 O)4 ]·4CH2 Cl2 [Mn12 O12 (O2 CCH2 But )16 (CH3 OH)4 ]·CH3 OH [Mn12 O12 (O2 CCH2 But )16 (But OH)(H2 O)3 ]·2But OH [Mn12 O12 (O2 CCH2 But )16 (C5 H11 OH)4 ](C5 H11 OH: 1-pentanol) [Mn12 O12 (O2 CPhSCH3 )16 (H2 O)4 ]·8CHCl3 [Mn12 O12 (O2 CMe)16 (dpp)4 ]·6.1CH2 Cl2 ·0.4H2 O (dppH: diphenyl phosphate) [Mn12 O12 (O2 CCF3 )16 (H2 O)4 ]·2CF3 COOH·4H2 O [Mn12 O12 (O2 CCF3 )16 (H2 O)4 ]·CF3 COOH·7H2 O [Mn12 O12 (O2 CC6 F5 )16 (H2 O)4 ]·3CH2 Cl2 [NMe4 ]2 [Mn12 O12 (O2 CC6 F5 )16 (H2 O)4 ]·6C7 H8 [NMe4 ][Mn12 O12 (O2 CC6 F5 )16 (H2 O)4 ]·4.5CH2 Cl2 ·1/2H2 O [Mn12 O12 (O2 CPet )16 (MeOH)4 ] Pet COOH = 2,2-dimethylbutyric acid [Mn12 O12 (O2 CC4 H3 S)16 (H2 O)4 ]·6CH2 Cl2 ·2H2 O [Mn12 O12 (O2 CC4 H3 S)16 (HO2 CC4 H3 S)(H2 O)2 ]·5CH2 Cl2 C2/c I2/a Ibca P −1 P −1 10 10 0.65K 0.61K [Mn12 O12 (O2 CCHCl2 )8 (O2 CCH2 But )8 (H2 O)3 ]·CH2 Cl2 ·H2 O [Mn12 O12 (O2 CCHCl2 )8 (O2 CEt)8 (H2 O)3 ]·CH2 Cl2 [Mn12 O12 (O2 CC6 H5 )8 La4 (H2 O)4 ]·8CH2 Cl2 (H2 La : 10-(4acetylsulfanylmethyl-phenyl)-anthracene-1,8-dicarboxylic acid) [Mn12 O12 (NO3 )4 (O2 CCH2 But )12 (H2 O)4 ]·MeNO2 [Mn12 O12 (O2 CMe)8 (O3 SPh)8 (H2 O)4 ] [Mn12 O12 (Z)16 (H2 O)4 ][PF6 ]16 [Mn12 O12 (Z)16 (H2 O)4 ][W6 O19 ]8 [Mn12 O12 (Z)16 (H2 O)4 ][PW12 O40 ]16/3 [Mn12 O12 (Z)16 (H2 O)4 ][(H3 O)PW11 O39 Ni]4 [Mn12 O12 (Z)16 (H2 O)4 ][(H3 O)PW11 O39 Co]4 Z = O2 C-Ph-p-CH2 N(CH2 CH2 CH2 CH3 )3 [NBun4 ]2 [Mn12 O12 (OMe)2 (O2 CPh)16 (H2 O)2 ]·2H2 O·4CH2 Cl2 [NBun4 ]2 [Mn12 O12 (OMe)2 (O2 CPh)16 (H2 O)2 ]·2H2 O·CH2 Cl2 (NBun4 )2 [Mn12 O12 (OMe)2 (O2 CPh)16 (H2 O)2 ]·6CH2 Cl2 (PPh4 )[Mn12 O12 (O2 CEt)16 (H2 O)4 ] (PPh4 )[Mn12 O12 (O2 CPh)16 (H2 O)4 ] · 8CH2 Cl2 (PPh4 )[Mn12 O12 (O2 CPh)16 (H2 O)4 ] P −1 P −1 I41 /am d C2/c P −1 10 10 10 -0.45 -0.42 10 10 10 10 10 10 10 -0.46 -0.34 -0.44 -0.40 -0.40 52 53 P bca P −1 10 50.1 19/2 19/2 19/2 19/2 -0.40 -0.62 −0.44 n.a. 69.5 21.7 64 28 53 62 35 67.09 51.81 66.44 72 71 65.2 41.8 72 67 53 51 51 P 2/c 10 10 11 −0.09 −0.14 −0.22 55.1 57 57.5 55 28 50 25 45 40 26 24 54 50 53 51 28 27 18.5 30.3 34.7 P −4 11 −0.31 33.8 P 21 /c P −1 (m − MPYNN+ )[Mn12 O12 (O2 CPh)16 (H2 O)4 ]− 19/2 (NPrn4 )[Mn12 O12 (O2 CCHCl2 )16 (H2 O)4 ] (NPrn4 )2 [Mn12 O12 (O2 CCHCl2 )16 (H2 O)4 ] (NPrn4 )3 [Mn12 O12 (O2 CCHCl2 )16 (H2 O)4 ] [NMe4 ]3 [Mn12 O12 (O2 CCHCl2 )16 (H2 O)4 ] [Fe(C5 Me5 )2 ][Mn12 O12 (O2 CC6 F5 )16 (H2 O)4 ] · 2H2 O [Fe(C5 H5 )2 ][Mn12 O12 (O2 CC6 F5 )16 (H2 O)4 ] [Co(C5 Me5 )2 ][Mn12 O12 (O2 CC6 F5 )16 (H2 O)4 ] · 2CH2 Cl2 · C6 H14 [Co(C5 H5 )2 ][Mn12 O12 (O2 CC6 F5 )16 (H2 O)4 ] [Fe(C5 Me5 )2 ]2 [Mn12 O12 (O2 CC6 F5 )16 (H2 O)4 ] [Fe(C5 H5 )2 ]2 [Mn12 O12 (O2 CC6 F5 )16 (H2 O)4 ] (PPh4 )2 [Mn12 O12 (O2 CCHCl2 )16 (H2 O)4 ] · 4CH2 Cl2 · H2 O (PPh4 )2 [Mn12 O12 (O2 CCHCl2 )16 (H2 O)4 ] · 6CH2 Cl2 [Mn12 O12 (bet)16 (EtOH)4 ](PF6 )14 · 4CH3 CN · H2 O bet : (CH3 )3 N+ − CH2 − CO− 2 [Mn12 O12 (bet)16 (EtOH)3 (H2 O)](PF6 )13 (OH) · 6CH3 CN · EtOH · H2 O 19/2 10 17/2 17/2 21/2 21/2 Aba2 P 21 /c −0.35 −0.28 −0.25 −0.23 −0.38 21/2 21/2 −0.36 9.1 × 10−9 1.1 × 10−8 1.6 × 10−9 n.a. ref 6a 16 17 18 19 20 22 23 24 26 27 28 29 30 31 32 33 38 39 40 2.1 × 10−10 41,42 42 122 Ren Min et al.: Single Molecule Magnets FIG. 2. The molecular structure of compounds Mn(saltmen)(ReO4 )]2 (left, adapted from ref. [43]) and [MnIII 3 O(O2 CMe)3 (mpko)3 ]ClO4 (right, adapted from ref. [46]) pound shows slow relaxation of magnetization below 4 K with an activation energy Ueff of 16 K. Hysteresis appears below 2 K (for field sweep rates of 1.4 T/s) and becomes temperature independent below 0.6 K, indicating the QTM effect. By varying the apical ligands and the chemical features of the Schiff base, it is possible to modulate the MnIII . . . MnIII magnetic interactions and the overall magnetic behavior of the MnIII dimer. Those showing SMM behav2 iors include [Mn(saltmen)(X)]2 (X− = CH3 COO, N3 ), [Mn(salen)(NCO)]2 [salen2− = N,N’-ethylenebis(salicylide neiminate), [Mn(3,5-Brsalen)(3,5Brsalicylaldehyde)]2 [3,5-Brsalen2− = N,N’-ethylenebis(3,5-dibromosalicylideneiminate)][44] and [Mn2 (5MeOsaltmen)2 (DCNNQI)2 ] (DCNNQI = N,N’dicyano-1,4-naphthoquinonediiminate radical).[45] Their energy barriers are in the range of 19.0 ∼ 28.7 K. For the oxide-centered, triangular, [Mn3 O(O2 CR)6 L3 ]n+ (n=0, 1) complexes, ferromagnetic interactions can be “turned on” by posing relatively small, ligand-imposed structural distortions, defined by the Mn-N-O-Mn torsion angles (α). This leads to the first SMM with a triangular topology [MnIII 3 O(O2 CMe)3 (mpko)3 ]ClO4 ·3CH2 Cl2 (α = 11.2◦ ) (Fig. 2, right).[46] It has an S = 6 ground state and a negative axial anisotropy (D = −0.34 cm−1 ). The effective relaxation barrier is Ueff = 10.9 K. Below 0.3 K, the relaxation is temperature-independent, consistent with relaxation by ground-state QTM. A much higher effective barrier (37.5 K) is achieved for [MnIII 3 O(Mesalox)3 (2,4’-bpy)3 (ClO4 )]·0.5MeCN which shows a similar triangular topology but possessing a larger Mn-N-O-Mn torsion angle of 44.2◦ .[47] Brechin et al. carried out a systematic study by using R-saoH2 ligands to control the puckering or twisting of their [48] central cores of the MnIII Their energy 3 O triangles. barriers vary from 25.7 K to 57.0 K. A few Mn3 O SMMs using salox ligands were also reported.[49] FIG. 3. (a) Core structure of the [MnIV MnIII 3 O3 Cl] cubane; (b) The structure of the [Mn4 O3 Cl4 (O2 CEt)3 (py)3 ]2 dimer, denoted [Mn4 ]2 . The dashed lines are C-H · · · Cl hydrogen bonds and the dotted line is the close Cl · · · Cl approach. (c) The magnetization hysteresis loops of [Mn4 ]2 shown at different temperatures. Reproduced with permission from ref. [53]. Copyright 2002, Nature Publishing Group. Complexes [MnIV MnIII [L 3 O3 X(O2 CMe)3 L3 ] = pyridine (py), dibenzoylmethane (dbm)] contain a distorted cubane structure with a central 6+ [MnIV MnIII core.[50,51] The MnIV 3 (µ3 -O)3 (µ3 -X)] III . . . Mn antiferromagnetic exchange interactions leads to an S = 9/2 ground state. For compound [MnIV MnIII 3 O3 Cl(O2 CMe)3 (dbm)3 ], SMM behavior is observed at low temperature with the effective relaxation barrier of Ueff = 11.8 K. Although the ground state is a Kramer doublet, QTM is observed below 0.9 K, which is explained by internal magnetic field within the Mn4 complex due to the nuclear spins (IMn = 5/2, IH = 1/2).[52] A very interesting compound is [Mn4 O3 Cl4 (O2 CEt)3 (py)3 ]2 ([Mn4 ]2 ) which crystallizes in the hexagonal space group R − 3 with pairs of Mn4 molecules lying ‘head to head’ on a crystallographic S6 symmetry axis.[53] (Fig. 3a). This [Mn4 ]2 supramolecular arrangement is held together by six C-H · · · Cl hydrogen bonds between the pyridine rings on one [Mn4 ] and Cl ions on the other (Fig. 3b), thus a weak antiferromagnetic (AF) coupling is mediated between the Mn4 units resulting in an S = 0 ground state. Wernsdorfer et al. found that this AF coupling is manifested as an exchange bias of all tunnelling transitions, and the hysteresis loops consequently display unique features, such as the absence of a QTM step at zero field (Fig. 3c). The phenomenon is very important if SMMs are to be used for information storage. Another important family of Mn4 SMMs is mixed valent clusters containing a planar III MnII rhombus core. For example, compound 2 Mn2 II [ Mn2 MnIII 2 (O2 C Me)2 (H p d m)6 ] (Cl O4 )2 · 2.5 H2 O (pdmH2 is pyridine-2,6-dimethanol) shows ferromagnetic interactions between MnIII -MnII and MnIII -MnIII pairs resulting in an S = 9 ground state with negative axial ZFS (D = −0.31 cm−1 ).[54,55] It Ren Min et al.: Single Molecule Magnets exhibits a SMM behavior with the effective energy barrier of 16.7 K. A seriers of Mn4 (hmp)6 (hmp = 2-hydroxymethylpyridine) clusters with similar III MnII rhombus core showing SMM behaviors 2 Mn2 are also reported, with the energy barriers ranging from 9.2 K to 23.3 K.[56∼62] Other related complexes were reported using bridging ligands such as triethanolamine (teaH3 ) and N-butyldiethanolamine (bdeaH2 ). The highest energy barrier is found for III compound [MnII 2 Mn2 (bdea)2 (bdeaH)2 (O2 CPh)4 ] [63] (Ueff =26.7 K). Compound [MnIII 6 O2 (sao)6 (O2 CPh)2 (EtOH)4 ]·EtOH (saoH2 = salicylaldoxime) is among the first SMMs based on Mn6 family.[64] It has a nonplanar [MnIII 6 (µ3 O)2 (µ2 -OR)2 ]12+ core, made up of two off-set stacked 7+ triangular subunits bridged by two [MnIII 3 (µ3 -O)] central oximato oxygen atoms, with the remaining four sao2− ligands bridging in a near-planar η 1 : η 1 : 7+ η 1 : µ-fashion along the edges of the [MnIII 3 (µ3 -O)] triangles (Fig. 4). The ferromagnetic interaction between the antiferromagnetically coupled Mn3 O triangles leads to an S = 4 ground state. The energy barrier and τ0 values are 27.9 K and 2.3 × 10−8 s. Motivated by the triangular Mn3 O clusters in which the structural distortion could switch on the ferromagnetic interaction within the trimer, Brechin and co-workers carried out a systematic work to investigate whether the additional steric bulk of the derivatized oximates would enforce structural distortions.[65−69] They found a “magic area” (30.4◦ ∼ 31.3◦ ) of the torsion angles to predict pairwise exchange. When α > 31.3◦ , J > 0 (F). When α < 30.4◦ , J < 0 (AF). Thus the ground state of the Mn6 cluster can vary from 4 to 12 simply by controlling the structural distortion. A record energy barrier of 86.4 K among SMMs based on 3d transition metal clusters is observed for compound [MnIII 6 O2 (Etsao)6 (O2 CPh(Me)2 )2 (EtOH)6 ] (average torsion angle of 39.1◦ ) which shows blocking temperature (TB ) of ca. 4.5 K.[14] B. Fe-O clusters Since high spin FeIII has an 6 S ground state, large anisotropy cannot be realized for the single FeIII species. Very few FeIII complexes have been reported to show SMMs behaviors. The first one is [Fe8 O2 (OH)12 (tacn)6 ]Br8 ·9H2 O (Fe8 , tacn = 1,4,7triazacyclononane) (Fig. 5, left).[70] It also has an S = 10 ground state but with an Ising type magnetic anisotropy of about 1/3 that of Mn12 Ac. The low symmetry in Fe8 results in a sizeable transverse magnetic anisotropy.[71] Although the ac blocking temperature is only 3 K,[72] the Fe8 complex is a good candidate for the study of quantum effects on the magneti- 123 FIG. 4. The core (left) and molecular structure (right) of [MnIII 6 O2 (sao)6 (O2 CPh)2 (EtOH)6 ] (adapted from ref. [14]). FIG. 5. The structures of [Fe8 O2 (OH)12 (tacn)6 ]Br8 ·9H2 O (left, adapted from ref. [70]) and [FeIII 4 (OMe)6 (dpm)6 ] (right, adapted from ref. [74]). zation dynamics.[73] The propeller-like FeIII clusters are among 4 the simplest inorganic systems showing SMM behavior. The first one of this family has a molecular formula of [FeIII 4 (OMe)6 (dpm)6 ] (Hdpm = dipivaloylmethane).[74] The molecule has 2-fold symmetry. The four Fe atoms lie exactly on a plane, the inner Fe atom being in the center of the triangle (Fig. 5, right). Antiferromagnetic interactions are found between the central and peripheral Fe ions (J = −21.1 cm−1 ), while that between the neighboring peripheral ones is ferromagnetic (J 0 = 1.1 cm−1 ). This leads to a spin ground state of S = 5. The compound shows a uniaxial magnetic anisotropy with D = −0.20 cm−1 and E = 0.01 cm−1 . The energy barrier is 3.5 K, which is significantly smaller than the expected value (U = 7.1 K), attributed to the quantum tunneling due to the transverse anisotropies. Through site-specific ligand replacement of the methoxide bridges with a tripodal ligand R-C(CH2 OH)3 , complexes [FeIII 4 (L)2 (dpm)6 ] [H3 L= R-C(CH2 OH)3 ] was obtained.[75] The magnetic anisotropy and energy barriers of the FeIII SMMs 4 may be tuned by changing the organic groups in 124 Ren Min et al.: Single Molecule Magnets R-C(CH2 OH)3 .[76∼81] The new derivatives exhibit in general enhanced magnetic properties with respect to the parent cluster (Ueff = 11.1∼17.0 K). Their static and dynamic magnetic parameters correlate strongly with the helical pitch (γ) of the Fe(O2 Fe)3 core. The axial anisotropy |D| (evaluated from EPR spectra) and the effective anisotropy barrier Ueff (extracted from relaxation measurements) both increase with increasing helical pitch. A related III[82] [83] FeIII and a chiral FeIII compounds were 3 Cr 4 also reported with the energy barriers of 7.0 K and 4.1 K, respectively. The FeII -O clusters showing SMM behaviors are rare, including [FeII 4 (sae)4 (MeOH)4 ], [FeII [FeII 4 (3,5-Cl2 4 (5-Br-sae)4 (MeOH)4 ]·MeOH, [84] II sae)4 (MeOH)4 ] and [Fe9 (N3 )2 (O2 CMe)8 {(2py)2 CO2 }4 ].[85] Their energy barriers are 28.4 K, 30. 5K, 26.2 K and 41.7 K, respectively. C. Single 3d-metal ions Although much attention has been paid to metal clusters in searching for new SMMs with large barriers, mononuclear transition metal complexes become attractive very recently. But the number is still rather few so far. For example, iron(II) complexes of coordination numbers 4 and 3 have been characterized as having the magnetic signatures of orbital angular momentum.[86] Long et al. reported that the axial and transverse zero-field splitting (ZFS) parameter for the trigonal pyramidal complex K[(tpaMes )Fe] are D = −40 cm−1 and E = −0.4 cm−1 , respectively.[10a] The large magnitude of D stems from the presence of three electrons residing in the 1e orbital set (Fig. 6). The negative sign of D would indicate a significant intrinsic spin-reversal barrier of U = S 2 |D| = 227 K. Slow magnetization relaxation was observed under a dc field, giving an effective barrier of Ueff = 60.4 K (τ0 = 2 × 10−9 s). The absence of slow relaxation under zero applied field is attributed to QTM through spin-reversal barrier. The trigonal pyramidal iron(II) complexes [(tpaR )Fe]− can show the ability to systematically enhance the magnetic anisotropy of the S = 2 center via increasing the electron donating abilities of the tris(pyrrolyl-α-methyl)amine ligand.[10b] In the case of R = tert-butyl, the axial ZFS parameter becomes D = −48 cm−1 , and the effective energy barrier increases to 93.5 K. In complex [FeII (N(TMS)2 )2 (PCy3 )], the central FeII ion is coordinated by one PCy3 and two N(TMS)2 ligands in a trigonal planar arrangement. This leads to a negative ZFS (D = −7.6 cm−1 ). Slow magnetization relaxation is observed under an external dc field of 600 Oe, with the energy barrier of 42 K (τ0 = 6 × 10−7 s). In contrast, complex [FeII (N(TMS)2 )2 (depe)] in which the FeII ion has a distorted tetrahedral ge- FIG. 6. Structure of the trigonal pyramidal complex [(tpaR )Fe]− , R = tert-butyl (a); mesityl (b); phenyl (c); 2,6-difluorophenyl (d); and the splitting of the 3d orbital energies for a high-spin FeII center in a trigonal pyramidal ligand field (e) (adapted from ref. [10b]). ometry does not show SMM behavior.[87] In complex [5 CpFe(C6 H3 i Pr3 -2,6)] (5 Cp = C5 i Pr5 ), the Fe-C bond almost coincides with the C5 axis of the ring ligand. It shows a large negative axial ZFS (D = −51.4 cm−1 ). The Arrhenius fitting of the ac susceptibility data leads to effective energy barriers of 40.3 K (τ0 = 6 × 10−6 s) for the process probed at 750 Oe and 143.4 K (τ0 = 7.8 × 10−9 s) for the one probed at 2500 Oe.[88] Large magnetic anisotropies can also be achieved in two-coordinate complexes with a linear L-M-L geometry. Five compounds, Fe[N(SiMe3 )(Dipp)]2 (1) Fe[C(SiMe3 )3 ]2 (2), Fe[N(H)Ar’]2 (3), Fe[N(H)Ar*]2 (4), and Fe(OAr’)2 (5) feature a linear geometry at the FeII center, while the sixth compound, Fe[N(H)Ar# ]2 (6), is bent with an N-Fe-N angle of 140.9(2)◦ (Dipp = C6 H3 -2,6-Pri2 ; Ar’ = C6 H3 -2,6-(C6 H3 -2,6-Pri2 )2 ; Ar* = C6 H3 -2,6-(C6 H2 -2,4,6-Pri2 )2 ; Ar# = C6 H3 -2,6(C6 H2 -2,4,6-Me3 )2 ). Ac magnetic susceptibility data for all compounds revealed slow magnetic relaxation under an applied dc field, with the magnetic relaxation time following a general trend of 1 > 2 > 3 > 4 > 5 >> 6. Arrhenius plots were fit by employing a sum of tunneling, direct, Raman, and Orbach relaxation processes, resulting in spin reversal barriers of Ueff = 260, 210, 157, 150, and 62 K for 1∼5, respectively.[89] The mononuclear Co(II) complexes were also explored as SMMs. The first examples are [{ArNdCMe}2 (NPh)]Co(NCS)2 and [{ArNdCPh}2 (NPh)]Co(NCS)2 . Both have a distorted square-pyramidal geometry with the CoII centers lying above the basal plane. This leads to significant spin-orbit coupling for the d7 CoII ions and consequently to slow relaxation of the magnetization under a static dc field that is characteristic of SMM behavior. The effective energy barriers are 16 K (τ0 = 3.6 × 10−6 s) and 24 K (τ0 = 5.1 × 10−7 s), respectively.[90] Field-induced slow magnetization relaxation is also observed in compound cis[CoII (dmphen)2 (NCS)2 ]·0.25EtOH. The highly rhombically distorted octahedral environment is Ren Min et al.: Single Molecule Magnets FIG. 7. Left: Structure of the tetrahedral [Co(SPh)4 ]2− complex. Right: Electronic configuration and d-orbital energy level splitting for the molecule, with energies derived using the angular overlap model (adapted from ref. [92]). responsible for the strong axial and rhombic magnetic anisotropy of the high-spin CoII ion (D = +98 cm−1 , E = +8.4 cm−1 ). The activation energy resulting from the Arrhenius fitting of the ac data is 24.2∼26.0 K [τ0 = (3.0 − 4.4) × 10−7 s].[91] Noting that all the above mentioned mononuclear transition metal-based SMMs require application of a dc field to disrupt fast QTM process. Compound (Ph4 P)2 [Co(SPh)4 ] remains to be an exception (Fig. 7). In this complex, the CoII ion is tetrahedrally coordinated with an axial ZFS of D = −70 cm−1 . It displays SMM behavior in the absence of an applied magnetic field. The effective energy barrier is Ueff = 30 K (τ0 = 1.0 × 10−7 s).[92] III. SINGLE MOLECULE MAGNETS BASED ON 4f METAL IONS Lanthanides are widely used in magnet technology. The interest in f-element SMMs was boosted by the report of Ishikawa et al. in 2003 that the mononuclear bis-phthalocyanine compounds [Pc2 Ln] can show slow magnetic relaxation.[93] The origin of the magnetism is from both orbital and spin angular momentums of a single lanthanide ion, which is placed in a ligand field, giving the lowest sublevels a large |Jz | value and energy gaps from the rest of the sublevels. Due to the single ion features of them, these complexes are also called single-ion magnets. Table II. lists the pure 4f monomers of clusters that showing SMM behaviors. A. Lanthanide phthalocyaninates In TBA+ [Pc2 Ln]− [Ln = Tb, Dy, Ho, Er, Tm, Yb; Pc = dianion of phthalocyanine; TBA+ = N(C4 H9 )+ 4 ], the trivalent lanthanide ion is placed in an eight coor- 125 dinate square antiprism ligand field made by two Pc ligands with approximate D4d symmetry (Fig. 8a). The high-order axial coordination field results in a strong axial anisotropy along the C4 axis with potential to exhibit slow relaxation of magnetization. However, the slow magnetization relaxation is observed only in two of the complexes with the thermally activated barriers of 331 and 40 K for Tb and Dy species, respectively.[93] The energy barrier of Tb complex is significantly larger than those found in the 3d-metal based SMMs. Such a behavior is strongly related to the sublevel structures of the ground state multiplets of the complexes. The lowest substates in the Tb complex is Jz = ±6, which are the maximum and minimum values in the J = 6 ground state. The energy separation from the rest of the substates is more than 576 K (Fig. 9a). For the Dy complex, the lowest substates are characterized as Jz = 13/2 and −13/2, which are the second largest in the J = 15/2 ground state. The magnetic hysteresis measurements in the subkelvin temperature range show clear evidence of QTM for Tb, Dy, Ho compounds.[93,94] The quantum process is a result of the resonant quantum tunneling between entangled states of electron and nuclear spin systems.[93,94] For the Dy complex with half integer ground state, no tunneling should occur because of the Kramers theorem of spin parity. But the coupling of J = 15/2 with nuclear spin I = 5/2 lead to an integer total spin. Nevertheless, the step structure at H 6= 0 T is not clear because of the significantly reduced tunnel splitting gap.[95] The splitting of the crystal field levels is, however, very sensitive to the environment. For example, Ruben and co-workers employed solid state 1 H NMR to analyze the spin dynamics of the [Pc2 Tb]− complex in a series of different diamagnetically diluted preparations, TBA+ [Pc2 Tb]− ×9[TBA]Br and TBA+ [Pc2 Tb]− ×143[TBA]Br, using excess tetrabutylamonium bromide as matrix complement.[96] The activation energy increases from 840 K in the undiluted sample TBA+ [Pc2 Tb] to 922 K in the diamagnetically diluted samples. The observation emphasize that even the diamagnetic [TBA]Br matrix arrangement around the [Pc2 Tb]− complexes can alter the splitting of the crystal field levels, and hence the TbIII spin dynamics. The fact that small differences in molecular surrounding could trigger substantial modifications of the SMM property is of utmost importance for ongoing research on surface-deposited bis-phthalocyaninato terbium (III) molecules targeting the realization of single molecular data storage.[96] The double-decker phthalocyaninates can exist not only in the anionic form [Pc2 Ln]− , but also in the neutral form [Pc2 Ln]0 or cationic form [Pc2 Ln]+ . The [Pc2 Ln]0 molecule has two spin systems, i.e. an un- 126 Ren Min et al.: Single Molecule Magnets TABLE II. SMMs based on lanthanide and actinide ions and clusters Space group Ueff (K) [(C4 H9 )4 N][Pc2 Tb] 331 [(C4 H9 )4 N][Pc2 Dy] 40.3 [(C4 H9 )4 N][Pc2 Ho] n.a. (n − Bu4 N)+ [{Pc(Oet)8 }2 Dy]− 79.1 [(C4 H9 )4 N][Tb(Pc − R)2 ](R = −C15 H31 ) 640 [(C4 H9 )4 N][Tb(Pc − R)2 ](R = −C3 H7 ) 616 [(C4 H9 )4 N][Tb(Pc − R)2 ](R = −CH(CH3 )Ph) 666 [{Pc(OR)8 }2 Tb](R = CH3 (CH2 )11 OCH(CH3 )CH2 −) 690(cr) liquid-crystalline 607(dis) [(Pc)2 TbIII ]0 P 21 21 21 590 Dy(obPc)2 P 21 /n 63 [{Pc(OEt)8 }2 TbIII ]+ (SbCl6 )− [Pc(OEt)8 = dianion 791 of 2,3,9,10,16,17,23,24- octaethoxyphthalocyanine] [Pc(OEt)8 2 Dy]+ (SbCl6 )− 38.8 Tb2 (obPc)3 P −1 331 Dy2 (obPc)3 [Tb(obPc)2 ]Cd[Tb(obPc)2 ] 211 [Dy(obPc)2 ]Cd[Dy(obPc)2 ] [Pc(OC4 H9 )8 ]Dy[Pc(OC4 H9 )8 ]Cd[Pc(OC4 H9 )8 ]Dy[Pc(OC4 H9 )8 ] P 2/c 22.4 [Pc(OC4 H9 )8 ]Tb[Pc(OC4 H9 )8 ]Cd[Pc(OC4 H9 )8 ]Tb[Pc(OC4 H9 )8 ] n.a. 312.7 [Pc(OC4 H9 )8 ]Tb[Pc(OC4 H9 )8 ]Cd(Pc)Tb(Pc) n.a. 291.5 [Pc(OC4 H9 )8 ]Tb[Pc(OC4 H9 )8 ]Cd(Pc)Y(Pc) n.a. 228.9 Na9 [ErW10 O36 ] · 35H2 O P −1 55.2 Na9 [Ho(W5 O18 )2 ]·x H2 O P −1 n.a. K12 DyP5 W30 O110 ·nH2 O 24 K12 HoP5 W30 O110 ·nH2 O 0.8 (Cp∗ )Er(COT) P nma 323.2 197.4 [DyIII (COT” )2 Li(THF)(DME)] P 21 /c 18(0 Oe) 24 23 (100 Oe) 30 43 (200 Oe) 43 (600 Oe) [Dy(FTA)3 L1 ] P 32 54.4 [Dy(NTA)3 L2 ] C2/c 30.4 [Dy(acac)3 (H2 O)2 ] · H2 O · C2 H5 OH P 21 /n 66.1 Dy(acac)3 (phen) P 21 /c 63.8 [Dy(acac)3 (dpq)] P −1 136 [Dy(acac)3 (dppz)] P −1 187 [Dy(TTA)3 (bpy)] P 21 /n 58.0 [Dy(TTA)3 (phen)] P −1 85.0 Na[Dy(DOTA)(H2 O)] · 4H2 O P −1 61 Na[Er(DOTA)(H2 O)] · 4H2 O P −1 39 Na[Yb(DOTA)(H2 O)] · 4H2 O P −1 29 [Dy(9Accm)2 (NO3 )(dmf)2 ] Pn 23 Zn(L)Dy(sal)(NO3 )Br P 21 /n 336 94 [Zn3 Dy(LPr )(NO3 )3 (MeOH)3 ] · 4H2 O P −1 25.8 [ErIII ZnII 3(L1 )(Oac)(NO3 )2 (H2 O)1.5 (MeOH)0.5 ] P 21 21 21 3.7 24.6 [Dy(dbm)3 LR] · 2H2 O P 21 21 21 46.9 L = 2,5-bis(4,5-pinene-2-pyridyl)pyrazine [Dy2 (dbm)6 LR ] · 2H2 O P 21 89.1 [Dy2III (valdien)2 (NO3 )2 ] P −1 76 [Dy2 (ovph)2 (NO3 )2 (H2 O)2 ] · 2H2 O P −1 69 198 τ0 (s) ) 6.3× 10−8 6.3× 10−6 n.a. 4.5×10−6 6.35×10−11 1.34 ×10−10 2.22×10−11 1.5 ×10−9 1.6×10−5 6.9×10−9 8.3 ×10−6 1.1×10−10 4.7 ×10−8 3.6 × 10−7 5.5 × 10−9 8.4 × 10−9 5.9 × 10−8 1.6 × 10−8 n.a. 8.17 × 10−11 3.13 × 10−9 6 × 10−6 3 × 1016 3 × 10−5 6 × 10−6 3 × 10−7 3 × 10−7 8.7 × 10−6 4.5 × 10−6 8.1 × 10−7 5.7 × 10−6 3.1 × 10−8 1.3 × 10−8 3.4 × 10−7 3.8 × 10−7 7 × 10−11 2.5 × 10−8 4 × 10−7 1.3 × 10−6 1.1 × 10−5 1.1 × 10−9 1.2 × 10−6 5.3 × 10−7 1.4 × 10−7 5.9 × 10−8 6.0 × 10−7 5.3 × 10−7 7.3 × 10−9 ref 93 93 94 97 98 98 98 99 99 100,101 102 103 97 105 106,107 106 106 108 109 109 109 110 111 113 113 114 114 116 116 116 116 116 116 117 118 119 120 121 121 122 122 123 125 125 126 127 127 128 129 129 130 130 130 131 132 132 127 Ren Min et al.: Single Molecule Magnets Space group C2/m [Dy0.87 Yb1.13 (H2 cht)2 Cl4 (H2 O)(MeCN)] · MeCN [Dy2 (H2 cht)2 Cl4 (H2 O)(MeCN)] · MeCN {[(Me3 Si)2 N]2 (THF)Dy}2 (µ − η 2 : η 2 − N2 ) {[(Me3 Si)2 N]2 (THF)Tb}2 (µ − η 2 : η 2 − N2 )− [Dy3 (µ3 − OH)2 L3 Cl(H2 O)5 ]Cl3 · 4H2 O · 2MeOH · 0.7MeCN [Dy3 (HSA)5 (SA)2 (phen)3 ](H2 SA = salicylicacid) [Dy3 (HL)(H2 L)(NO3 )4 ](H4 L = N, N, N0 , N’-tetrakis(2-hydroxyethyl)-ethylene-diamine) [Ln4 (OH)2 L2 (acac)6 ] · 2H2 L · 2CH3 CN(H2 L = N, N0 -bis(salicylidene)-1,2-cyclohexanediamine) [Dy6 (µ3 − OH)4 L4 L02 (H2 O)9 Cl]Cl5 · 15H2 O [Dy5 O(Oi Pr)13 ] C2/m P −1 P −1 C2/c P na21 P −1 P −1 P bca [Ho5 O(Oi Pr)13 ] [Dy6 (ovph)4 (Hpvph)2 Cl4 (H2 O)2 (CO3 )2 ] · CH3 OH · H2 O · CH3 CN n.a.=not available paired π electron on the Pc ligand with S = 1/2 spin and a LnIII ion with 4f electrons. Compound [Pc2 Ln]0 crystallizes in orthorhombic space group P 21 21 21 . The twisted angle between the two Pc rings is 41.4◦ , causing a pseudo 4-fold axis perpendicular to the Pc rings and a distorted antiprismatic coordination envi00 ronment. Peaks of the out-of-phase ( χM ) component of the π-radical [Pc2 Ln]0 have been observed at 50, 43 and 36 K with ac magnetic fields of 103, 102 and 10 Hz, respectively, which are more than 10 K higher than the corresponding values of the anionic complex [Pc2 Tb]− with a closed-shell π-system. The energy barrier is increased to 590 K, about two times that of [Pc2 Tb]− compound.[100,101] The oxidation of neutral species [{Pc(OEt)8 }2 Tb]0 can result in compound [{Pc(OEt)8 }2 Tb]+ (SbCl6 )− , where Pc(OEt)8 is a dianion of 2,3,9,10,16,17,23,24octaethoxyphthalocyanine. The magnetization reversal barrier for [{Pc(OEt)8 }2 Tb]+ (SbCl6 )− increases to 791 K, 8% greater than that of [{Pc(OEt)8 }2 Tb]0 .[103] Such a significant increase is attributed to the increased multiplet splitting by the strengthened ligand field, resulting from the longitudinal contraction of the coordination space of the Tb3+ ion induced by the ligand oxidation. The lanthanide phthalocyanine can also form “triple-deckers” composed of three Pc ligands and two lanthanide ions. The lanthanide ions are placed along the fourfold symmetry axis with a separation of about 3.6 Å. In this case, the f-f or dipole-dipole interactions should be considered. By studying the magnetic properties of PcTbPcTbPc* ([Tb,Tb]), PcYPcTbPc* ([Y,Tb]) and PcTbPcYPc* ([Tb,Y]) (Pc* = dianion of 2, 3, 9, 10, 16, 17, 23, 24-octabutoxy-phthalocyanine), Ishikawa et al found that the effect of QTM, which governs the relaxation process of the mono-Tb com- C2 P 21 /n Ueff (K) 100 29 124 177 326 61.7 65 42.6 90.9 22 200 528 46.6 400 76 τ0 (s) ) 8 × 10−9 8.2 × 10−9 2.2 × 10−8 1.5 × 10−5 1.0 × 106 5.8 × 10−7 3.7 × 10−6 (1.5 kOe) 1.5 × 10−9 4.7 × 10−10 3.8 × 10−6 1.5 × 10−9 1.2 × 10−6 ref 133 133 133 134 135 136 141 142 142 143 143 144 145 145 146 147 plex in near zero magnetic fields, is removed in the [Tb, Tb] by presence of f-f interaction.[104] Yamashita and co-workers reported a related tripledecker compound [Tb2 (obPc)3 ] (obPc = dianion of 2,3,9,10,16,17,23,24-octabutoxyphthalocyanine) with an energy barrier of 331 K.[105] Interestingly this compound shows dual magnetic relaxation processes in the low temperature region in the presence of a dc magnetic field. To further investigate the influence of dipole-dipole (f-f) interactions on magnetic relaxation, a family of multipledecker phthalocyaninato dinuclear lanthanide (III) single-molecule magnets has been reported.[106,107] The results show the quadruple-decker terbium compound {[Tb(obPc)2 ]Cd[Tb(obPc)2 ]}, and the quintuple-decker compound {[Tb(obPc)2 ]Cd(obPc)Cd[Tb(obPc)2 ]} show clearly dual relaxation processes in ascent of dc field below 10 K, but the dysprosium compounds show single relaxation process. Jiang and co-workers reported the quadruple-decker complex { [ Pc (OC4 H 9 )8 ] Dy [ Pc (OC4 H 9 )8 ] Cd [ Pc (OC4 H 9 )8 ] Dy [ Pc (OC4 H 9 )8 ] } which shows SMM behaviors with the energy barrier of 22.4 K.[108] They further studied the magnetic properties of four related Tb complexes {[Pc(OC4 H9 )8 ]Tb[Pc(OC4 H9 )8 ] Cd [Pc (OC4 H9 )8 ]Tb[Pc(OC4 H9 )8 ]} and {[Pc(OC4 H9 )8 ]M1 } [Pc(OC4 H9 )8 ]Cd(Pc)M2 (Pc)} (M1 -Cd-M2 = Tb-CdTb, Tb-Cd-Y, Y-Cd-Tb).[109] The energy barriers of the Tb-Cd-Tb systems are found to be larger than that of the monoterbium analogue, indicating again the effect of the long-distance intramolecular f-f interactions in the axial direction on the magnetic properties of sandwich-type quadruple-decker complexes. Besides, the QTM can be suppressed by the long-distance intramolecular f-f interaction. 128 Ren Min et al.: Single Molecule Magnets FIG. 8. (a) Schematic diagram of [LnPc2 ]− (Ln = Tb, Dy, Ho, Er, Tm, or Yb). Reproduced with permission from ref. 93b. Copyright 2004, American Chemical Society. (b) Structure of the [ErW10 O36 ]9− POM and squareantiprismatic environment of Er (adapted from ref. [110]). B. Lanthanide polyoxometalates The family of lanthanide polyoxometalates (POMs) is formed by POMs encapsulating one or more lanthanide ions in order to give rise to a lanthanide complex in which the 4f-magnetic ions are subject to the crystal field created by the POM ligands. The first two families of POM-based SMMs are reported with the general formula [Ln(W5 O18 )2 ]9− (Ln = Tb, Dy, Ho and Er) and [Ln(SiW11 O39 )2 ]13− (Ln = Tb, Dy, Ho Er, Tm and Yb). Among them, only the [Er(W5 O18 )2 ]9− derivative exhibited slow relaxation of the magnetization associated with SMM behavior.[110,111] The anisotropy energy barrier was determined to be 55.2 K. Compound Na9 [Er(W5 O18 )2 ] is formed by two anionic [W5 O18 ]6− moieties sandwiching the center lanthanide ion. These anionic clusters are surrounded by sodium cations to balance the charge. Each anionic [W5 O18 ]6− moiety is twisted 44.2◦ with respect to the other, resulting in a slightly distorted squareantiprismatic environment around the Ln ion (Fig. 8b). The square-antiprism exhibits certain axial compression, in contrast to the axial elongation observed in “double-decker” [Pc2 Ln]− complexes. In fact, complex [Pc2 Er]− shows no slow relaxation phenomenon, with the low-lying ground states of MJ = ±1/2.[112] Interestedly, through the fitting of the susceptibility data of [ErW10 O36 ]9− , a Kramers doublet ground state with MJ = ±13/2 and two excited states with MJ = ±1/2 and ±15/2 is present (Fig. 9b). Such a difference in the energy level scheme seems to be associated with the different axial distortion of the Er coordination site in these two compounds. Very recently, another family of lanthanide POMs with a 5-fold symmetry has been reported by Coronado and co-workers, namely, K12 LnP5 W30 O110 ·nH2 O (Ln3+ = Tb, Dy, Ho, Er, Tm, and Yb).[113] In these structures, the lanthanide center can occupy two equivalent coordination sites, which show a very unusual 5-fold geometry formed by five phosphate oxygens and five bridging oxygens between tungsten atoms, resulting in a pentagonal antiprism coordination site. The shortest Ln-Ln distance is 13.2 Å. Compared with [ErW10 O36 ]9− , the C5 crystal field symmetry gives rise to remarkably large off-diagonal anisotropy parameters A56 , which mix magnetic states with different MJ values. Consequently, only Dy and Ho complexes exhibit magnetic hysteresis at low temperature corresponding to SMM behavior. The spin dynamics, especially at low temperatures, is dominated by fast tunneling processes and strongly affected by hyperfine interactions and external magnetic fields. The thermally activated energy barriers turn out to be very small. However, it can provide attractive candidates for the application as solid-state spin qubits. C. Lanthanide organometallics Many organometallic compounds show a doubledecker structure in which the metal center is sandwiched by two aromatic ligands. The uniaxial anisotropy is promoted by high symmetry coordination environment and delocalized ligands. The first organometallic lanthanide SMM was reported by Gao et al. in 2011,[114] namely (Cp*)Er(COT) (COT = cyclooctatetraenide, C8 H2− 8 ; Cp* = pentamethylcyclopentadienide, C5 Me− ). The Er(III) ion is sand5 wiched between the two aromatic rings, being closer to the COT center (1.66 Å) than to that of the Cp* ring (2.27 Å) (Fig. 10). Because of the different rings and the tilting between them, the Er(III) is situated in an environment of low point group symmetry of Cs . The COT group is crystallographically disordered in the temperature range of 10∼120 K. The disorder is assumed to be static by nature due to the coexistence of two stable conformers with different COT conformations in the crystal rather than dynamic position disorder. The local symmetry of Er(III) is approximated to be C∞v . The fine electronic structures are investigated with ligand field theory, and shows that the ground state is | ± 15/2i, with the first excited state | ± 13/2i, lying 273 K above. Ac magnetic measurements show that this compound has two thermally activated relaxation processes with the energy barriers of ∆E1 = 323 K (τ01 = 8.17×10−11 s) and ∆E2 = 197 K (τ02 = 3.13 × 10−9 s), attributed to the aforementioned two stable conformers in the crystal. Butterfly shaped hysteresis loops were recorded at an average scanning field speed of 550 Oe/min below 5 K. The isostructural Dy and Ho compounds also show SMM behaviors, but their energy barriers are much lower (24.3, 80.6 cm−1 ).[115] Murugesu and co-workers reported another organometallic lanthanide SMM with formula [DyIII (COT” )2 Li(THF)(DME)] (COT” = 1,4bis(trimethylsilyl)cyclooctatetraenyl dianion).[116] Ren Min et al.: Single Molecule Magnets 129 FIG. 9. Energy level diagrams of the ground-state multiplets for [LnPc2 ]− (left, reproduced with permission from ref. 93b. Copyright 2004, American Chemical Society) and [Ln(W10 O36 )]9− (right, reproduced with permission from ref. [111]. Copyright 2009, American Chemical Society). with the other fields, continues to illustrate that the low-frequency peak C is the predominant relaxation pathway. However, close inspection of the peak shape reveals that the peak signals are broad, suggesting an overlapping relaxation mechanism (A and C) with very similar relaxation times. The overlapping peaks are barely apparent. FIG. 10. Left: Schematic view of the title compound (Cp*)Er(COT). Right: Out-of-phase of ac susceptibility at various temperatures and frequencies in the absent of dc field. Reproduced with permission from ref. [114]. Copyright 2011, American Chemical Society. In this case, central DyIII ion is sandwiched by two COT” ligands. To accommodate the sterically bulky trimethylsilyl groups, the COT” rings are arranged in a staggered conformation. The Li ion interacts with one COT” ring. In zero dc field, the frequency-dependent ac magnetic susceptibility is observed. The data in the high temperature region can be fit by the Arrhenius law giving the relaxation barrier of Ueff = 18 K and a τ0 value of 6 × 10−6 s (Pathway A). Below 3.75 K, relaxation starts to become temperature-independent, indicative of a quantum regime (pathway B). Applying a 100 Oe field results in a reduction in the quantum tunneling along with the appearance of the secondary peak, indicating the coexistence of the new relaxation pathway C. At 200 Oe, the intensity of the secondary low frequency peak (C) increases and the intensity of the primary high-frequency peak (A) decreases. The data at the optimum field of 600 Oe, consistent D. Other mononuclear lanthanide SMMs Li et al. reported the first SMM based on mononuclear lanthanide β-diketone, e.g. [Dy(FTA)3 L] (FTA = 2-furyltrifluoro-acetonate, L = (S,S)-2,2’-bis(4benzyl-2-oxazoline)), which shows an energy barrier of Ueff = 54.4 K.[117] In this complex, the DyIII center is eight-coordinated by three β-diketonate anions and a N,N’-chelating chiral ligand L with a geometry between a bicapped square prism and a square antiprism. Distinct slow magnetic dynamic behaviors were found in two polymorphs of [Dy(NTA)3 L’] (NTA = 2-naphthyltrifluoro-acetonate, L’ = (1R, 2R)-1,2diphenylethane-1,2-diamine),[118] attributed to the different local environments of DyIII centers in the crystal. The polymorphic form with a distorted bicapped triangular prismatic coordination geometry in DyIII ion (C2v symmetry) shows typical features of the SMM behavior at zero field, while the polymorphic form with a distorted dodecahedral coordination geometry in DyIII ion (D2d symmetry) shows frequency 0 00 dependent in-phase (χ ) and out-of-phase (χ ) signals only by applying an external dc field.[118] Gao and co-workers reported the magnetic behavior of [Dy(acac)3 (H2 O)2 ] (acac = acetylacetonate),[119] where the DyIII ion is put on a distorted square- 130 Ren Min et al.: Single Molecule Magnets antiprismatic coordination geometry with approximately D4d local symmetry, similar to those in [LnPc2 ]− and [ErW10 O36 ]9− . This complex exhibits a crossover at 8 K between thermally activated relaxation above 8 K and quantum tunneling relaxation below this temperature. The energy barrier for the former is 66.1 K. When the coordination water is replaced by 1,10-phenanthroline (phen), the energy barrier of compound [Dy(acac)3 (phen)] is similar.[120] However, if the auxiliary groups are replaced by large aromatic groups, the ligand field around the DyIII in compounds [Dy(acac)3 (dpq)] and [Dy(acac)3 (dppz)] will be enhanced which promotes the separation of the lowest doubly degerate sublevels from the rest of the exited states.[121] By utilizing 4,4,4-trifluoro-1-(2-thienyl)-1,3-butanedionate (TTA) and capping ligands such as 2,2’-bipyridine (bipy) and phen, Tang and coworkers also reported two SMMs [Dy(TTA)3 (bipy)] (Ueff = 58 K) and [Dy(TTA)3 (phen)] (Ueff = 85 K).[122] Sessoli and co-workers reported the magnetic properties of a polycrystalline sample of Na[Dy(DOTA)(H2 O)]·4H2 O (H4 DOTA = 1,4,7,10tetraazacyclododecane- N, N’, N”, N’”-tetraacetic acid).[123] This compound behaves like a SMM and, more interestingly, shows a giant field dependence on the relaxation time of the magnetization. In zero static field a temperature-independent underbarrier mechanism is observed, while the application of a weak magnetic field induces a thermally activated regime with an effective barrier of ca. 60 K and an increase in the relaxation time of six orders of magnitude at 1.8 K compared to zero static field. In Na[Dy(DOTA)(H2 O)]·4H2 O, the DyIII ion is in an idealized square antiprismatic environment with a twist angle of 39◦ typical of the heaviest lanthanides. But the tetragonal symmetry is limited to the first coordination sphere. The magnetic anisotropy of this compound is then studied by combining measurements of the angular dependence of the single crystal magnetic susceptibility and luminescence characterization. It is found unprecedently that the easy axis is not linked to the idealized symmetry axis of the complex, but is almost perpendicular to it. The results demonstrate that even subtle structural details, like the position of hydrogen atoms and the consequent orientation of the nonbonding orbitals of the axial ligand can overcome the symmetry imposed by the coordination polyhedron.[124] E. Other polynuclear lanthanide SMMs In polynuclear lanthanide cluster complexes, the single-molecule magnet behavior is thought to originate largely from the strong anisotropy of the single FIG. 11. Molecular structures of compounds [DyIII 131) 2 (valdien)2 (NO3 )2 ] (left, adapted from ref. and [Dy2 (ovph)2 Cl2 (MeOH)3 ] (right, adapted from ref. [132] ). lanthanide center, with only weak contributions from intramolecular exchange coupling. However, compared with the isolated lanthanide ion, the magnetic behaviors of the polynuclear Ln systems are more complex due to the effect of the magnetic interactions and the usual non-collinearity of the main single-ion anisotropy axes of the different lanthanide ions. Compound [DyIII 2 (valdien)2 (NO3 )2 ] (H2 valdien = N1, N3-bis(3-methmethoxysalicylidene) diethylenetriamine) is a symmetrical Dy2 cluster (Fig. 11, left), exhibits SMM behavior with an anisotropic barrier Ueff = 76 K.[131] The step-like features in the hysteresis loops observed reveal an antiferromagnetic exchange coupling between the two dysprosium ions. The ab initio calculations reveal an exchange constant of JDy − JDy = −0.21 cm−1 . The asymmetrical dinuclear DyIII SMM, [Dy2 (ovph)2 Cl2 (MeOH)3 ]·MeCN (H2 ovph = pyridine-2-carboxylic acid [(2-hydroxy3-methoxyphenyl)methylene] hydrazide) (Fig. 11, right), reported by Tang and co-workers, shows a weak ferromagnetic coupling between the Dy centers, each with a nearly perfect pentagonal bipyramidal coordination environment.[132] The ab initio calculations indicate that the quantum tunneling pathways are strongly suppressed in low-lying exchange multiplets at low temperatures and the ferromagnetic coupling comes entirely from a ferromagnetic dipolar interaction (Jdip = 5.36 cm−1 ). Detailed magnetization dynamics studies reveal that the blockage of magnetization in the high-temperature regime occurs at individual Dy sites. Two closely spaced relaxation processes, attributed to the individual Dy ions, can be well resolved by the sum of two modified Debye functions. Tong et al reported a heterospin dinuclear complex [Dy0.87 Yb1.13 (H2 cht)2 Cl4 (H2 O)(MeCN)]·MeCN which shows shifts of the relaxation barriers with respect to the barriers observed in homospin Dy2 and Yb2 isostructural complexes.[133] The difference of activation energies in pure and mixed-metal complexes is due to small geometrical changes of the environment of the Ln ions. Ren Min et al.: Single Molecule Magnets In order to enhance the magnetic interactions between the 4f ions, Long et al. introduced N3− 2 radical as bridging ligand to produce dinuclear compounds {[(Me3 Si)2 N]2 (THF)Ln}2 (µ − η 2 : η 2 −N2 ) (Ln = Gd, Dy).[134] The fit of dc susceptibility for the GdIII congener using a spin-only Hamiltonian reveals the strongest magnetic coupling of J = −27 cm−1 yet observed for this ion. The incorporation of the highly anisotropic DyIII ion results in a molecule with a record magnetic blocking temperature of 8.3 K at conventional sweep rates. Further, synergizing the strong magnetic anisotropy of terbium(III) with the effective exchange-coupling ability of the N3− radical create 2 the hardest molecular magnet discovered to date. The terbium analogue exhibits magnetic hysteresis at 14 K and a 100 s blocking temperature of 13.9 K.[135] Those breakthroughs demonstrate that a joint contribution, combining strong magnetic coupling with single-ion anisotropy, may ultimately open up a new era in envisioned technological applications for SMMs. The triangular Dy3 cluster [Dy3 (µ3 − OH)2 L3 ]4+ (HL = o-vanillin) gives a unique example of SMMs (Fig. 12), showing unusual slow relaxation behavior to 8 K in spite of the almost non-magnetic ground state.[136] This intriguing behavior originates from the noncollinearity of the single ion magnetization axes of the DyIII ions, as revealed by single crystal magnetic studies,[137] muon spin lattice relaxation measurements,[138] as well as ab initio calculations.[139,140] The peculiar chiral nature of the ground non-magnetic doublet and the resonant quantum tunneling of the magnetization at the crossings of the discrete energy levels open new perspectives in quantum computation and data storage in molecular nanomagnets. This stimulates intensive investigations in utilizing triangular Dy3 units for creating new SMMs with higher energy barriers.[144] Considering that much of the fascinating physics of (TBA)[Tb(Pc)2 ] and other single-ion magnets, such as Na9 [Er(W5 O18 )2 ], is associated with their fourfold symmetry, McInnes and Winpenny et al studied the magnetic properties of an iso-propoxide-bridged dysprosium square-based pyramid [Dy5 O(Oi Pr)13 ] which has both fourfold symmetry and metal triangles. This compound shows an energy barrier of 528 K, which is by far the largest barrier yet observed for any d- or f -block clusters145 IV. CONCLUSION In this review, we summarized the progress of single molecule magnets of metal oxo clusters, single transition metal ions and lanthanide ions or clusters. The latter is of particular interest due to their high magnetization reversal barriers and high blocking temper- 131 FIG. 12. View of the structure of the triangular Dy3 cluster (adapted from ref. [136]). atures. Although much need to be understood about the lanthanide-based SMMs, such as the mechanism of the relaxation and how to control the magnetic anisotropy, these molecular materials show promising potential in spintronics and quantum computing. Progresses have been achieved in grafting the prototype Mn12 , Fe4 and Pc2 Ln systems on different substrates.[148] The XMCD technique was developed to give the information of the oxidation state of the metal ions in SMM monolayers, and to detect the magnetic signals arising from layers of a magnetic molecule.[149] These findings prove that the quantum spin dynamics can be observed in SMMs chemically grafted to surfaces, and offer a tool to reveal the organization of matter at the nanoscale. REFERENCES [1] Christou G, Gatteschi D, Hendrickson D N, et al. Mater. Res. Soc. Bull., 2000, 25: 66 [2] (a) Thomas L, Lionti F, Ballou R, et al. Nature, 1996, 383: 145; (b) Gatteschi D, Sessoli R. Angew. Chem. Int. Ed., 2003, 42: 268 [3] Mannini M, Pineider F, Sainctavit P, et al. Nat. Mater., 2009, 8: 194 [4] (a) Leuenberger M N, Loss D. Nature, 2001, 410: 789; (b) Ardavan A, Rival O, Morton J J L, et al. Phys. Rev. Lett., 2007, 98: 057201; (c) Stamp P C E, Gaita-Ariño A. J. Mater. Chem., 2009, 19: 1718 [5] Torres F, Hernández J M, Bohigas X, et al. J. Appl. Phys. Lett., 2000, 77: 3248 [6] (a) Sessoli R, Hui L, Schake A R, Wang S, et al. J. Am. Chem. Soc., 1993, 115: 1804; (b) Sessoli R, Gatteschi D, Caneschi A, et al. Nature, 1993, 365: 132 [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] Ren Min et al.: Single Molecule Magnets 141; (c) Gatteschi D, Sessoli R, Villain J. Molecular Nanomagnets, Oxford University Press, Oxford, 2006 (a) Gatteschi D, Sessoli R, Cornia A. Chem. Commun., 2000, 725-732; (b) Bagai R, Christou G. Chem. Soc. Rev., 2009, 38: 1011-1026; (c) M. Murrie, Chem. Soc. Rev., 2010, 39, 1986-1995; (d) Rogez G, Donnio B, Terazzi E, et al. Adv. Mater. 2009, 21: 4323-4333 (a) Sessoli R, Powell A K. Coord. Chem. Rev., 2009, 253: 2328; (b) Wang B W, Jiang S D, Wang X T, et al. Sci. China, Ser. B: Chem. 2009, 52: 1739; (c) Rinehart J D, Long J R. Chem. Sci., 2011, 2: 2078; (d) Sorace L, Benelli C, Gatteschi D. Chem. Soc. Rev., 2011, 40: 3092 (a) Rinehart J D, Long J R. J. Am. Chem. Soc., 2009, 131: 12558; (b) Rinehart J D, Meihaus K R, Long J R. J. Am. Chem. Soc., 2010, 132: 7572; (c) Magnani N, Apostolidis C, Morgenstern A, et al. Angew. Chem., Int. Ed., 2011, 50: 1696; (d) Antunes M A, Pereira L C J, Santos I C, et al. Inorg. Chem., 2011, 50: 9915 (a) Freedman D E, Harman W H, Harris T D, et al. J. Am. Chem. Soc., 2010, 132: 1224; (b) Harman W H, Harris T D, Freedman D E, et al. J. Am. Chem. Soc., 2010, 132: 18115 Kostakis G E, Ako A M, Powell A K. Chem. Soc. Rev., 2010, 39: 2238-2271 Tasiopoulos A J, Vinslava A, Wernsdorfer W, Abboud K A, et al. Angew. Chem. Int. Ed., 2004, 43: 2117-2121 Ako A M, Hewitt I J, Mereacre V, et al. Chem. Int. Ed., 2006, 45: 4926 -4929 Milios C J, Vinslava A, Wernsdorfer W, et al. J. Am. Chem. Soc., 2007, 129: 2754-2755 Eppley H J, Tsai H L, Devries N, et al. J. Am. Chem. Soc., 1995, 117: 301 Sun Z, Ruiz D, Dilley N R, et al. Chem. Commun., 1999, 1973-1974 Aubin S M J, Sun Z M, Eppley H J, et al. Inorg. Chem., 2001, 40: 2127 Ruiz-Molina D, Gerbier P, Rumberger E, et al. J. Mater. Chem., 2002, 12: 1152 Chakov N E, Lee S C, Harter A G, et al. J. Am. Chem. Soc., 2006, 128: 6975-6989 Barra A L, Caneschi A, Cornia A, et al. J. Am. Chem. Soc., 2007, 129: 10754 Wernsdorfer W, Murugesu M, Christou G. Phys. Rev. Lett., 2006, 96: 057208 Lampropoulos C, Redler G, Data S, et al. Inorg. Chem., 2010, 49: 1325 Zobbi L, Mannini M, Pacchioni M, et al. Chem. Commun., 2005, 1640-1642 Bian G Q, Kuroda-Sowa T, Konaka H, et al. Inorg. Chem., 2004, 43: 4790 Zhao H H, Berlinguette C P, Bacsa J, et al. Inorg. Chem., 2004, 43: 1359 Chakov N E, Soler M, Wernsdorfer W, et al. Inorg. Chem., 2005, 44: 5304 Chakov N E, Zakharov L N, Rheingold A L, et al. Inorg. Chem., 2005, 44: 4555 Lim J M, Do Y, Kim J. Eur. J. Inorg. Chem., 2006, 711 [29] Soler M, Artus P, Folting K, et al. Inorg. Chem., 2001, 40: 4902 [30] Pacchioni M, Cornia A, Fabretti A C, et al. Chem. Commun., 2004, 2604 [31] Artus P, Boskovic C, Yoo J, et al. Inorg. Chem., 2001, 40: 4199 [32] Chakov N E, Wernsdorfer W, Abboud K A, et al. D. N. Dalton Trans., 2003, 2243 [33] Forment-Aliaga A, Coronado E, Feliz M, et al. Inorg. Chem., 2003, 42: 8019 [34] Tasiopoulos A J, Wernsdorfer W, Abboud K A, et al. Angew. Chem. Int. Ed., 2004, 43: 6338 [35] Tasiopoulos A J, Wernsdorfer W, Abboud K A, Christou G. Inorg. Chem. 2005, 44: 6324 [36] Aubin S M J, Sun Z M, Pardi L, et al. Inorg. Chem., 1999, 38: 5329 [37] Takeda K, Awaga K. Phys. Rev. B, 1997, 56: 14560 [38] Bagai R, Christou G. Inorg. Chem., 2007, 46: 10810 [39] Kuroda-Sowa T, Lam M, Rheingold A L, et al. Inorg. Chem., 2001, 40: 6469 [40] Soler M, Wernsdorfer W, Abboud K A, et al. J. Am. Chem. Soc., 2003, 125: 3576 [41] Coronado E, Forment-Aliaga A, Gaita-Arino A, et al. Angew. Chem. Int. Ed., 2004, 43: 6152 [42] Clemente-Juan J M, Coronado E, Forment-Aliaga A, et al. Inorg. Chem., 2010, 49: 386 [43] Miyasaka H, Clérac R, Wernsdorfer W, et al. Angew. Chem. Int. Ed., 2004, 43: 2801-2805 [44] Lu Z L, Yuan M, Pan F, et al. Inorg. Chem., 2006, 45: 3538-3548 [45] Kachi-Terajima C, Miyasaka H, Sugiura K, et al. Inorg. Chem., 2006, 45: 4381-4390 [46] Stamatatos T C, Foguet-Albiol D, Stoumpos C C, et al. J. Am. Chem. Soc., 2005, 127: 15380-15381 [47] Yang C I, Wernsdorfer W, Cheng K H, et al. Inorg. Chem., 2008, 47: 10184-10186 [48] Inglis R, Taylor S M, Jones L F, et al. Dalton Trans., 2009, 9157-9168 [49] Feng P L, Koo C, Henderson J J, et al. Inorg. Chem., 2008, 47: 8610-8612; (b) Feng P L, Koo C, Henderson J J, et al. Inorg. Chem. 2009, 48: 3480-3492 [50] Aubin S M J, Wemple M W, Adams D M, et al. J. Am. Chem. Soc., 1996, 18: 7746 [51] Aubin S M J, Dilley N R, Pardi L, et al. J. Am. Chem. Soc., 1998, 120: 4991 [52] Kramers H A. Proc. R. Acad. Sci. Amsterdam, 1930, 33: 959 [53] Wernsdorfer W, Aliaga-Alcalde N, Hendrickson D N, et al. Nature, 2002, 416: 406 [54] Brechin E K, Yoo J, Nakano M, et al. Chem. Commun., 1999, 783 [55] Yoo J, Brechin E K, Yamaguchi A, et al. Inorg. Chem., 2000, 39: 3615 [56] Yoo J, Yamaguchi A, Nakano M, et al. Inorg. Chem., 2001, 40: 4604 [57] Lecren L, Wernsdorfer W, Li Y G, et al. J. Am. Chem. Soc., 2005, 127: 11311-11317 [58] Lecren L, Roubeau O, Coulon C, et al. J. Am. Chem. Soc., 2005, 127: 17353 [59] Lecren L, Roubeau O, Li R G, et al. Dalton Trans., 2008, 755 [60] Miyasaka H, Nakata K, Lecren L, et al. J. Am. Chem. Soc., 2006, 128: 3770-3783 Ren Min et al.: Single Molecule Magnets [61] Hiraga H, Miyasaka H, Nakata K, et al. Inorg. Chem., 2007, 46: 9661-9671 [62] Morimoto M, Miyasaka H, Yamashita M, et al. J. Am. Chem. Soc., 2009, 131: 9823 [63] Ako A M, Mereacre V, Hewitt I J, et al. J. Mater. Chem., 2006, 16: 2579 [64] Milios C J, Raptopoulou C P, Terzis A, et al. Angew. Chem. Int. Ed., 2004, 43: 210 [65] Milios C J, Inglis R, Vinslava A, et al. J. Am. Chem. Soc., 2007, 129: 12505 [66] Inglis R, Jones L F, Milios C J, et al. Dalton Trans., 2009, 3403-3412 [67] Moro F, Corradini V, Evangelisti M, et al. J. Phys. Chem. B, 2008, 112: 9729 [68] Milios C J, Vinslava A, Wood P A, et al. J. Am. Chem. Soc., 2007, 129: 8 [69] Jones L F, Inglis R, Cochrane M E, et al. Dalton Trans., 2008, 6205 [70] (a) Delfs C, Gatteschi D, Pardi L, et al. Inorg. Chem., 1993, 32: 3099-3103; (b) Barra A L, Gatteschi D, Sessoli R. Chem. Eur. J., 2000, 6: 16081614 [71] (a) Gatteschi D, Sessoli R, Cornia A. Chem.Commun., 2000, 725-732; (b) Wernsdorfer W, Sessoli R, Caneschi A, et al. Europhys. Lett., 2000, 50: 552-558 [72] Sangregorio C, Ohm T, Paulsen C, et al. Phys. Rev. Lett., 1997, 78: 4645-4648 [73] Takahashi S, Tupitsyn I S, van Tol J, et al. Nature, 2011, 76: 476 [74] arra A L, Caneschi A, Cornia A, et al. J. Am. Chem. Soc., 1999, 121: 5302 [75] Cornia A, Fabretti A C, Garrisi P, et al. Angew. Chem. Int. Ed., 2004, 43: 1136-1136 [76] Accorsi S, Barra A L, Caneschi A, et al. J. Am. Chem. Soc., 2006, 128: 4742-4755 [77] Barra A L, Bianchi F, Caneschi A, et al. Eur. J. Inorg. Chem., 2007, 4145-4152 [78] Gregoli L, Danieli C, Barra A L, et al. Chem. Eur. J., 2009, 15: 6456-6467 [79] Condorelli G G, Motta A, Pellegrino G, et al. Chem. Mater., 2008, 20: 2405-2411 [80] Bogani L, Danieli C, Biavardi E, et al. Angew. Chem. Int. Ed., 2009, 48: 746-750 [81] Rodriguez-Douton M J, Cornia A, Sessoli R, et al. Dalton Trans., 2010, 39: 5851-5859 [82] Tancini E, Rodriguez-Douton M J, Sorace L, et al. Chem. Eur. J., 2010, 16: 10482-10493 [83] Zhu Y Y, Guo X, Cui C, et al. Chem. Commun., 2011, 47: 8049-8051 [84] (a) Oshio H, Hoshino N, Ito T. J. Am. Chem. Soc., 2000, 122: 12602-12603; (b) Oshio H, Hoshino N, Ito T, et al. J. Am. Chem. Soc., 2004, 126: 8805-8812 [85] Boudalis A K, Donnadieu B, Nastopoulos V, et al. Angew. Chem. Int. Ed., 2004, 43: 2266 -2270 [86] (a) Andres H, Bominaar E L, Smith J M, et al. J. Am. Chem. Soc., 2002, 124: 3012; (b) Popescu C V, Mock M T, Stoian S A, et al. Inorg. Chem., 2009, 48: 8317 [87] Lin P H, Smythe N C, Gorelsky S I, et al. J. Am. Chem. Soc., 2011, 133: 15806-15809 [88] Weismann D, Sun Y, Lan Y, et al. Chem. Eur. J., 2011, 17: 4700-4704 133 [89] Zadrozny J M, Atanasov M, Bryan A M, et al. Chem. Sci., 2013, 4: 125-138 [90] Jurca T, Farghal A, Lin P H, et al. J. Am. Chem. Soc., 2011, 133: 15814-15817 [91] Vallejo J, Castro I, Ruiz-Garcı́a R, et al. J. Am. Chem. Soc., 2012, 134: 15704-15707 [92] Zadrozny J M, Long J R. J. Am. Chem. Soc., 2011, 133: 20732-20734 [93] (a) Ishikawa N, Sugita M, Ishikawa T, et al. J. Am. Chem. Soc., 2003, 125: 8694-8695; (b) Ishikawa N, Sugita M, Ishikawa T, et al. J. Phys. Chem. B, 2004, 108: 11265-11271 [94] Ishikawa N, Sugita M, Wernsdorfer W. J. Am. Chem. Soc., 2005, 127: 3650-3651 [95] Ishikawa N, Sugita M, Wernsdorfer W. Angew. Chem. Int. Ed., 2005, 44: 2931-2935 [96] Branzoli F, Carretta P, Filibian M, et al. J. Am. Chem. Soc., 2009, 131: 4387-4396 [97] Ishikawa N, Mizuno Y, Takamatsu S, et al. Inorg. Chem., 2008, 47: 10217-10219 [98] Gonidec M, Amabilino D B, Veciana J. Dalton Trans., 2012, 41: 13632-13639 [99] Gonidec M, Luis F, Vı́lchez A, et al. Angew. Chem. Int. Ed., 2010, 49: 1623-1626 [100] Ishikawa N, Sugita M, Tanaka N, et al. Inorg. Chem., 2004, 43: 5498-5500. [101] Katoh K, Yoshida Y, Yamashita M, et al. J. Am. Chem. Soc., 2009, 131: 9967-9976 [102] Katoh K, Umetsu K, Brian K B, et al. Sci. in China, 2012, 55: 918-925 [103] Takamatsu S, Ishikawa T, Koshihara S Y, et al. Inorg. Chem. 2007, 46: 7250-7252 [104] Ishikawa N, Otsuka S, Kaizu Y. Angew. Chem. Int. Ed., 2005, 44: 731-733 [105] Katoh K, Kajiwara T, Nakano M, et al. Chem. Eur. J., 2011, 17: 117-122 [106] Katoh K, Horii Y, Yasuda N, et al. Dalton Trans., 2012, 41: 13582 [107] Katoh K, Isshiki H, Komeda T, et al. Coord. Chem. Rev., 2011, 255: 2124-2148 [108] Wang H, Qian K, Wang K, et al. Chem. Commun., 2011, 47: 9624-9626 [109] Wang H, Liu T, Wang K, et al. Chem. Eur. J., 2012, 18: 7691-7694 [110] AlDamen M A, Clemente-Juan J M, Coronado E, et al. J. Am. Chem. Soc., 2008, 130: 8874-8875 [111] AlDamen M A, Cardona-Serra S, Clemente-Juan J M, et al. Inorg. Chem., 2009, 48: 3467-3479 [112] Ishikawa N, Sugita M, et al. J. Phys. Chem. B, 2004, 108: 11265-11271 [113] Cardona-Serra S, Clemente-Juan J M, Coronado E, et al. J. Am. Chem. Soc., 2012, 134: 14982-14990 [114] Jiang S D, Wang B W, Sun H L, et al. J. Am. Chem. Soc., 2011, 133: 4730-4733 [115] Jiang S D, Liu S S, Zhou L N, et al. Inorg. Chem., 2012, 51: 3079-3087 [116] Jeletic M, Lin P H, Le Roy J J, et al. J. Am. Chem. Soc., 2011, 133: 19286-19289 [117] Li D P, Wang T W, Li C H, et al. Chem. Commun., 2010, 46: 2929-2931 [118] Li D P, Zhang X P, et al. Chem. Commun., 2011, 47: 6867-6869 134 Ren Min et al.: Single Molecule Magnets [119] Jiang S D, Wang B W, et al. Angew. Chem. Int. Ed., 2010, 49: 7448-7451 [120] Chen G J, Gao C Y, Tian J L, et al. Dalton Trans., 2011, 40: 5579 [121] Chen G J, Guo Y N, Tian J L, et al. Chem. Eur. J., 2012, 18: 2484-2487 [122] Bi Y, Guo Y N, Zhao L, et al. Chem. Eur. J., 2011, 17: 12476-12481 [123] Car P E, Perfetti M, Mannini M, et al. Chem. Commun., 2011, 47: 3751 [124] Cucinotta G, Perfetti M, Luzon J, et al. Angew. Chem. Int. Ed., 2012, 51: 1606 [125] Boulon M E, Cucinotta G, Luzon J, et al. Angew. Chem. Int. Ed., 2013, 52: 350-354 [126] Menelaou M, Ouharrou F, Rodrı́guez L, et al. Chem. Eur. J., 2012, 18: 11545-11549 [127] Watanabe A, Yamashita A, et al. Chem. Eur. J., 2011, 17(27): 7428-7432 [128] Feltham H L C, Lan Y, Klöwer F, et al. Chem. Eur. J., 2011, 17(16): 4362-4365 [129] Yamashita A, Watanabe A, Akine S, et al. Angew. Chem. Int. Ed., 2011, 50(17): 4016-4019 [130] Li X L, Chen C L, Gao Y L, et al. Chem. Eur. J., 2012, 18: 14632-14637 [131] Habib F, Lin P H, Long J, et al. J. Am. Chem. Soc., 2011, 133: 8830-8833 [132] Guo Y N, Xu G F, Wernsdorfer W, et al. J. Am. Chem. Soc., 2011, 133: 11948-11951 [133] Leng J D, Liu J L, Zheng Y Z, et al. Chem. Commun., 2013, 49: 158 [134] Rinehart J D, Fang M, Evans W J, et al. Nat. Chem., 2011, 3: 538-542 [135] Rinehart J D, Fang M, Evans W J, et al. J. Am. Chem. Soc., 2011, 133: 14236-14239 [136] Tang J, Hewitt I, Madhu N T, et al. Angew. Chem. Int. Ed., 2006, 45: 1729-1733 [137] Luzon J, Bernot K, et al. Phys. Rev. Lett., 2008, 100: 247205 [138] Salman Z, Giblin S R, Lan Y, et al. Phys. Rev. B, 2010, 82: 174427 [139] Chibotaru L F, Ungur L, Soncini A. Angew. Chem. Int. Ed., 2008, 47: 4126-4129 [140] Ungur L, Van den Heuvel W, Chibotaru L F. New J. Chem., 2009, 33(6): 1224-1230 [141] Liu C S, Du M, Sãnudo E C, et al. Dalton Trans., 2011, 40: 9366-9369 [142] Wang Y X, Shi W, Li H, et al. Chem. Sci., 2012, 3: 3366 [143] Yan P F, Lin P H, Habib F, et al. Inorg. Chem., 2011, 50: 7059-7065 [144] Hewitt I J, Tang J, Madhu N T, Anson C E, et al. Angew. Chem. Int. Ed., 2010, 49: 6352-6356 [145] Blagg R J, Muryn C A, McInnes E J L, et al. Angew. Chem. Int. Ed., 2011, 50: 6530-6533 [146] Blagg R J, Tuna F, McInnes E J L, et al. Winpenny, Chem. Commun., 2011, 47: 10587-10589 [147] Guo Y N, Chen X H, Xue S, et al. Inorg. Chem., 2012, 51: 4035-4042 [148] (a) Bogani L, Wernsdorfer W. Nat. Mater. 2008, 7: 179; (b) Cornia A, Mannini M, Sainctavit P, et al. Chem. Soc. Rev., 2011, 40: 3076-3091; (c) Gonidec M, Biagi R, Corradini V, et al. J. Am. Chem. Soc., 2011, 133: 6603-6612; (d) Katoh K, Isshiki H, Komeda T, et al. Chem. Asian J. 2012, 7: 1154-1169 [149] Mannini M, Pineider F, Sainctavit P, et al. Adv. Mater., 2009, 21: 167-171 Ren Min et al.: Single Molecule Magnets 单分子磁体 任旻、郑丽敏 南京大学化学化工学院,配位化学国家重点实验室 摘 要 : 单分子磁体指的是那些在阻塞温度以下出现磁化强度慢弛豫的分子。 自从 1993 年报道了 第一例单分子磁体 [Mn12 O12 (O2 CCH3 )16 (H2 O)4 ] · 4H2 O · 2CH3 CO2 H (Mn12 ) 以来,人们在探索 新颖的单分子磁体及其在信息存储、自旋电子学和量子计算等方面的潜在应用方面付出了极大的努 力,已经合成得到许多具有单分子磁体性质的单核和多核金属簇合物。 本文将重点介绍基于氧桥联 的锰或铁簇合物、单核过渡金属化合物以及稀土配合物的单分子磁体。 关键 词: 单分子磁体;锰;铁;稀土 135