A series of heterometallic 3d-4f polyoxometalates as single-molecule magnets

Shurong Li Zhenzhang Weng Linpeng Jiang Rongjia Wei Haifeng Su Lasheng Long Lansun Zheng Xiangjian Kong

Citation:  Shurong Li, Zhenzhang Weng, Linpeng Jiang, Rongjia Wei, Haifeng Su, Lasheng Long, Lansun Zheng, Xiangjian Kong. A series of heterometallic 3d-4f polyoxometalates as single-molecule magnets[J]. Chinese Chemical Letters, 2023, 34(3): 107251. doi: 10.1016/j.cclet.2022.02.056 shu

A series of heterometallic 3d-4f polyoxometalates as single-molecule magnets

English

  • Heterometallic 3d-4f cluster compounds have attracted widespread interest due to their interesting optical, electrical, magnetic, and catalytic properties resulting from the contribution of 3d and 4f electrons [1-9]. Considerable efforts in this field have resulted in the fabrication of a large number of 3d-4f clusters with fascinating structures, and unique physical and chemical properties based on various organic ligands. In these clusters, the organic ligands play an important protecting and bridging role in the separation and preparation of cluster structures [10-12]. 3d-4f Clusters protected by inorganic ligands have higher thermal stabilities and unique rigid structures compared to 3d-4f clusters protected by organic ligands [13, 14].

    Lacunary polyoxometalates (POMs) are a class of inorganic multidentate ligands with well-defined vacant sites and high negative charges. These are commonly used to prepare stable metal oxygen clusters or cluster-of-cluster aggregates due to their high nucleophilicity and high reactivity [15-19]. Several 3d metal clusters based on the trivacant POM {XW9O34} (heteroatom x = P, Si, Sb) have been reported [20-29]. However, the synthesis of POM-based 3d-4f clusters remains challenging due to the coordination competition between 3d metal ions and 4f ions in inorganic POM ligands. Moreover, the strong reaction between 4f ions and the oxygen-rich POM makes precipitation easy but crystallization difficult [13, 14, 30-34]. Therefore, reducing the reaction rate of lanthanide ions with POMs may be an effective strategy for the synthesis of POM-based 3d-4f clusters [35-37]. In this study, we have obtained three {FeW9O34}-based 3d-4f clusters, formulated—Na8K2[Fe2Ln2(H2O)4(B-α-FeW9O34)2]·nH2O (Ln2Fe4, where Ln = Dy and n = 17 for cluster 1; Ln = Ho and n = 15 for cluster 2; and Ln = Y and n = 14 for cluster 3). These clusters have classical sandwich-type structures, where the [(B-α-FeW9O34)11−] unit is generated via the transformation of the [(B-α-SbW9O33)9−] precursor (Figs. 1a and b). Ln2Fe4 is the first 3d-4f cluster assembled from POMs containing d-metal heteroatoms. Interestingly, Dy2Fe4 exhibits single-molecule magnet (SMM) behavior with a high energy barrier.

    Figure 1

    Figure 1.  The structures of (a) [B-α-SbW9O33]9− unit; (b) [B-α-FeW9O34]11− unit. (c, d) Ball-and-stick/polyhedral view of [(B-α-FeW9O34)2Fe2Dy2(H2O)4]10−. (e) The metal core of 2Dy3+ and 4Fe3+ ions in cluster 1. (f-h) The coordination geometries of DyO8/FeO6/FeO4. Color key: W/WO6, blue; Fe/FeO6/FeO4, green; Dy/DyO8, purple; Sb, black and O, pink.

    Clusters 1–3 were synthesized from Na9[B-α-SbW9O33]·19.5H2O, Ln2O3 (Ln = Dy/Ho/Y), FeCl3·6H2O and KH2PO4 under hydrothermal conditions (details in Supporting information). Single-crystal X-ray diffraction analysis showed that the isomorphic clusters 13 crystallized in the triclinic crystal system with a P-1 space groups (Table S1 in Supporting information). The valence bonding theory calculations of the three clusters (Tables S2-S4 in Supporting information) and the Mössbauer spectrum of cluster 3 indicate that the Fe ions are all high-spin Fe3+ (Fig. S3, Tables S5 and S6 in Supporting information) [38]. As shown in Figs. 1c and d, anion cluster 1 [(B-α-FeW9O34)2Fe2Dy2(H2O)4]10− exhibits a classic sandwich structural unit, which can be observed as a [Fe2Dy2(H2O)4]12+ unit sandwiched by two inorganic [(B-α-FeW9O34)11−] ligands. The [(B-α-FeW9O34)11−] unit is formed via the in situ transformation of the [(B-α-SbW9O33)9−] precursor. The Ln2Fe4 described in this study is the first 3d-4f cluster containing a trilacunary [(B-α-FeW9O34)11−] unit [39-43]. The formation of the [(B-α-FeW9O34)11−] species was investigated using single-crystal structure analysis, infrared spectroscopy, and energy dispersive spectroscopy (Figs. S1, S4-S6 in Supporting information). The metal core Fe4Dy2 in cluster 1 has exactly two tetrahedral structures with shared edges, with Fe⋯Fe distances of 3.1301(33)–3.3196(29) Å and Fe⋯Dy distances of 3.4987(22)–3.7597(26) Å (Fig. 1e, Fig. S18 and Table S12 in Supporting information). Continuous shape measurement (CShM) using Alvarez's SHAPE program (Table S8 in Supporting information) [44, 45] shows that Dy3+ in the [Fe2Dy2(H2O)4]12+ unit is 8-coordinated with triangular dodecahedral geometry (Fig. 1f), which is different from that reported for Fe4Dy2 [46-50]. The Dy-O length is between 2.243(12) Å and 2.514(15) Å (Table S7 in Supporting information). Six-coordinated Fe3+ has a classical octahedral geometry (Fig. 1g) with an Fe-O length of 1.991(10) Å–2.048(10) Å, whereas 4-coordinated Fe3+ in [(FeW9O34)11−] has a special tetrahedral geometry (Fig. 1h) with an Fe-O length of 1.834(10) Å–1.856(10) Å (Table S2). Clusters 2 and 3 have the same coordination pattern and similar bond lengths and bond angles as cluster 1 (Tables S3, S4, S7 and S8 in Supporting information). Based on elemental analysis, thermogravimetric analysis (Fig. S2 in Supporting information), ion chromatography, and charge balance theory, there are 8 Na+ and 2 K+ counter cations. As shown in Figs. S7 and S8 (Supporting information), the [(B-α-FeW9O34)2Fe2Dy2]10− anions are connected through hydrated Na+ and K+ counter ions to form a two-dimensional framework. The minimum asymmetric units of clusters 1-3 are shown in Figs. S9-S11 (Supporting information).

    The use of insoluble lanthanide substances also plays a key role in the formation of clusters 13. Clusters 1–3 can be obtained using insoluble lanthanide salts such as Ln2(CO3)3 and Ln2(SO4)3 to replace Ln2O3 (Ln = Dy, Ho and Y). However, such clusters cannot be obtained using soluble LnCl3 or Ln(NO3)3 salts. Ln2O3 can slowly release Ln3+ ions under mild acidic conditions (pH 5.5) and medium-high temperature hydrothermal conditions (140 ℃). The low equilibrium concentration of Ln3+ ions during the reaction can effectively slow the reaction rate between Ln3+ and POMs to prevent precipitation. Therefore, combining in situ POM transformation with the slow release of Ln3+ ions might be an effective synthetic strategy for the synthesis of POM-based 3d-4f clusters.

    The variable-temperature magnetic susceptibilities (χmT) of clusters 1–3 were measured between 2 K and 300 K under a 1000 Oe applied magnetic field. The relevant magnetic data obtained from these measurements are listed in Table S9 (Supporting information). The experimental χmT values for clusters 1–3 at 300 K (33.05, 32.20 and 5.63 cm3 K/mol, respectively) are smaller than the expected values (45.83, 45.63 and 17.50 cm3 K/mol, respectively), which may be due to antiferromagnetic interactions in the clusters [51]. As shown in Fig. 2, with decreasing temperature, the χmT values of clusters 2 and 3 gradually decrease to 21.69 cm3 K/mol and 1.09 cm3 K/mol at 2 K, respectively, mainly due to the thermal depopulation of the excited states at the mJ sublevels of the Fe3+ and Ho3+ ions. The χmT values of cluster 1 gradually decreased to 25.22 cm3 K/mol at 9 K followed by a sudden rebound to 25.99 cm3 K/mol at 2 K. This indicates that the presence of weak dipole interactions and/or ferromagnetic interactions between Dy3+ and Fe3+ below 9 K, including the decrease above 9 K can be attributed to the thermal depopulation of the excited states at the mJ sublevels of the Fe3+ and Dy3+ ions [52]. Fitting the χm−1-T data individually in the appropriate temperature interval for these three clusters according to the Curie-Weiss law gives C = 34.77 cm3/mol, θ = −17.57 K (1); C = 34.57 cm3/mol, θ = −16.87 K (2); and C = 34.45 cm3/mol, θ = −1554.21 K (3) (Figs. S12-S14 in Supporting information), which further indicates the presence of antiferromagnetic interactions. As shown in Figs. S15-S17 (Supporting information), the inter-cluster magnetic interactions are negligible due to the long distances. Cluster 3 has only Fe-Fe magnetic interactions, whereas cluster 1 (2) has Fe-Fe and Fe-Dy (Fe-Ho) magnetic interactions (Fig. S18 in Supporting information). Therefore, the difference in χmT values between isomorphic clusters 1, 2 and 3 can be attributed to the contribution of Fe-Dy (Fe-Ho) magnetic interactions (Fig. S19 in Supporting information). The field-dependent magnetization curves of 1–3 at 2 K are shown in Fig. S20 (Supporting information). The experimental values at 7 T are much smaller than their theoretical saturation values. Furthermore, the M vs. H/T curves at different temperatures are not superimposed on a single master curve (Fig. S20, Table S9), which suggests the existence of significant magnetic anisotropy and/or low-lying excited states [53].

    Figure 2

    Figure 2.  χmT vs. T in the range of 2–300 K under an applied field of 1000 Oe for clusters 13. Inset: Enlarged view of the upturn of χmT (cluster 1).

    AC susceptibility measurements of clusters 13 were performed to explore their dynamic magnetic properties at indicated frequencies under zero DC field and a 3 Oe AC field at 2 K. As shown in Figs. S21 (Supporting information), the out-of-phase susceptibilities (χ") of clusters 2 and 3 exhibit temperature dependence at 100–1500 Hz, but no significant peaks are observed due to the quantum tunneling effect. The in-phase susceptibilities (χ') and the out-of-phase susceptibilities (χ") of cluster 1 exhibit obvious temperature dependence, and the out-of-phase susceptibilities (χ") have obvious peaks at 0.1–1500 Hz. The AC magnetic susceptibility of cluster 1 was further tested at the specified temperatures in the frequency range of 0.1–1000 Hz under zero DC field and a 3 Oe AC field. Figs. 3a and b show the frequency dependence plots of χ' and χ''. A series of single frequency-dependent peaks were observed between 2 K and 9 K, indicating the SMM behavior of cluster 1. The Cole-Cole plots were fitted using the generalized Debye model (Fig. 3c). The large α values under zero DC field indicates the presence of multiple relaxation processes (Table S10 in Supporting information) [54]. The lnτ vs. T−1 plots are almost linear when T > 7 K, which can be fitted by considering only the Orbach relaxation process using Arrhenius' law (τ = τ0exp(Ueff/kBT)). The obtained parameters are Ueff/kB = 43.68 K and τ0=1.4 × 10−6 s, which is consistent with the reported ranges for other SMMs (10−6–10−11 s). However, when T < 7 K, the lnτ vs. T−1 plots deviate from linearity, indicating the presence of other magnetic relaxation processes. Combining the Obrach and quantum tunneling of magnetization (QTM) relaxation processes using the equation τ−1 = τ0−1 exp(−Ueff/kBT) + τQTM−1 gives the best fit (R2 = 0.99865), where τQTM is the rate of QTM. The fit gives τ0 = 6.4 × 10−7 s, τQTM = 6.0 × 10−3 s, and Ueff/kB= 50.24 K (Fig. 3d). To slow relaxation, the AC magnetic susceptibilities were measured under specific DC fields at 2 K. Under DC fields from 1000 Oe to 4000 Oe, the out-of-phase AC signal values increased significantly compared to those at 0 Oe (Fig. S22 in Supporting information), indicating that QTM is suppressed. Simultaneously, the frequency corresponding to the peak of the AC signal effectively decreases after applying the field and remains stable above 2500 Oe (Fig. S23 in Supporting information). The AC susceptibilities of cluster 1 were measured under an optimal 2500 Oe DC field. The temperature-dependent and frequency-dependent out-of-phase susceptibilities (χ") are significantly higher than that at zero field. The peak frequency is lower, and the peak temperature is slightly higher under the same frequency conditions. The Cole-Cole plot obtained by fitting the same generalized Debye model shows a semicircular shape (Fig. 4 and Fig. S24 in Supporting information). Similarly, the lnτ vs. T−1 plot is almost linear in the high-temperature region; considering only the Orbach relaxation process fitted according to Arrhenius' law: Ueff/kB = 64.55 K, and τ0 = 1.7 × 10−7 s. In the low-temperature region, the lnτ vs. T−1 plot deviates from linearity. As shown in Fig. 4d, combining the Obrach, direct and Raman relaxation processes using the equation τ−1 = τ0−1exp(-Ueff/kBT) + AT + CTn, results in τ0 = 9.3 × 10−8s, Ueff/kB = 80.21 K, A = 1.97 K−1 s−1, C = 1.51 × 10−3 s−1 K−n, n = 6.6 (R2 = 0.99762). If the QTM relaxation process is also considered, the fitted τQTM = −8.05 × 1014 s, suggesting that QTM can be considered completely suppressed at 2500 Oe [33, 55].

    Figure 3

    Figure 3.  (a) In-phase susceptibility (χ') and (b) out-of-phase susceptibility (χ'') vs. v under zero Oe DC field for 1. (c) Cole-Cole plots for AC susceptibilities under zero DC field. The colored solid lines are the best fit to the generalized Debye model. (d) Arrhenius plots of lnτ vs. T–1 for 1 under zero DC field.

    Figure 4

    Figure 4.  (a) In-phase (χ') and (b) out-of-phase susceptibility (χ'') vs. v at indicated temperatures under a 2500 Oe DC field for cluster 1. (c) Cole-Cole plots for AC susceptibilities under a 2500 Oe DC field. The colored solid lines are the best fit to the generalized Debye model. (d) Arrhenius plots of lnτ vs. T–1 for cluster 1.

    POMs are redox reversible and maintain high structural stability while rapidly gaining and losing multiple electrons [56]. The cyclic voltammograms and controlled-potential coulometry experiments showed that the isomorphic clusters 1-3 have three redox processes, peak shapes, and potentials similar to those of the Fe6W18 cluster, which indicates the gradual reduction of the four Fe centers. Based on the peak reduction currents, the three redox processes are correspond to the two-electron process, the one-electron process and the one-electron process, respectively (Figs. S25 and S26 in Supporting information). Clusters 1-3 are structurally similar to the reported cluster Fe6W18, which have only two Fe ions replaced by lanthanide ions at the periphery of the sandwich part [43]. Therefore, the Fe heteroatoms in clusters 1–3 (Ln2Fe4) can participate in the electrochemical redox process, whereas the Fe heteroatoms in the Fe6W18 cluster are electrochemically inert. Notably, among the POMs containing d-metal heteroatoms, only a few heteroatoms exhibit electrochemical activity [57]. The stability of cluster 1-3 in solution was confirmed by ESI-MS (Figs. S27-S29 in Supporting information).

    In summary, three {FeW9O34}-based 3d-4f clusters were obtained via a strategy that employs the slow release of lanthanide ions. The sandwich-like Ln2Fe4 clusters are the first 3d-4f clusters assembled from POMs containing d-metal heteroatoms. The in situ formation of [B-α-FeW9O34]11− and the slow release of Ln3+ play important roles in the formation of Ln2Fe4. Interestingly, the Dy2Fe4 cluster exhibits single molecule magnet properties with an 80 K energy barrier under an optimal DC field. The assembly of lacunary polyoxometalate-based 3d-4f SMMs is currently in progress.

    The authors declare that there is no interest for this manuscript.

    This work was supported by the National Natural Science Foundation of China (Nos. 21871224, 92161104, 92161203 and 21721001) and Innovation Laboratory for Sciences and Technologies of Energy Materials of Fujian Province (IKKEM No. RD2021040301).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2022.02.056.


    1. [1]

      L.Z. Cai, Q.S. Chen, C.J. Zhang, et al., J. Am. Chem. Soc. 137 (2015) 10882–10885. doi: 10.1021/jacs.5b05320

    2. [2]

      R. Chen, C.L. Chen, M.H. Du, et al., Chem. Commun. 57 (2021) 3611–3614. doi: 10.1039/d0cc08132a

    3. [3]

      R. Chen, Z.H. Yan, X.J. Kong, et al., Angew. Chem. Int. Ed. 57 (2018) 16796–16800. doi: 10.1002/anie.201811211

    4. [4]

      W.P. Chen, G.J. Zhou, Z.L. Gou, et al., Chin. Chem. Lett. 32 (2021) 838–841. doi: 10.1016/j.cclet.2020.05.018

    5. [5]

      M.H. Du, S.H. Xu, G.J. Li, et al., Angew. Chem. Int. Ed. 61 (2022) e202116296.

    6. [6]

      N.F. Li, Q.F. Lin, Y.M. Han, et al., Chin. Chem. Lett. 32 (2021) 3803–3806. doi: 10.1016/j.cclet.2021.04.042

    7. [7]

      H.J. Lun, M.H. Du, D.H. Wang, et al., Inorg. Chem. 59 (2020) 7900–7904. doi: 10.1021/acs.inorgchem.0c00613

    8. [8]

      X. Wang, M.H. Du, H. Xu, et al., Inorg. Chem. 60 (2021) 5925–5930. doi: 10.1021/acs.inorgchem.1c00333

    9. [9]

      H.L. Zhang, Y.Q. Zhai, L. Qin, et al., Matter 2 (2020) 1481–1493. doi: 10.1016/j.matt.2020.02.021

    10. [10]

      R. Chen, Z.F. Hong, Y.R. Zhao, et al., Inorg. Chem. 58 (2019) 15008–15012. doi: 10.1021/acs.inorgchem.9b02112

    11. [11]

      H. Han, Y.S. Ding, X. Zhu, et al., Inorg. Chem. Front. 7 (2020) 4070–4076. doi: 10.1039/d0qi00792g

    12. [12]

      M.G.F. Vaz, M. Andruh, et al., Coord. Chem. Rev. 427 (2021) 213611–213634. doi: 10.1016/j.ccr.2020.213611

    13. [13]

      V. Das, R. Kaushik, F. Hussain, Coord. Chem. Rev. 143 (2020) 213271–213291.

    14. [14]

      J.W. Zhao, Y.Z. Li, L.J. Chen, et al., Chem. Commun. 52 (2016) 4418–4445. doi: 10.1039/C5CC10447E

    15. [15]

      J. Cai, X.Y. Zheng, J. Xie, et al., Inorg. Chem. 56 (2017) 8439–8445. doi: 10.1021/acs.inorgchem.7b01104

    16. [16]

      C. Dey, J. Cluster Sci. 33 (2022) 1839–1856. doi: 10.1007/s10876-021-02110-8

    17. [17]

      C. Li, N. Mizuno, K. Yamaguchi, et al., J. Am. Chem. Soc. 141 (2019) 7687–7692. doi: 10.1021/jacs.9b02541

    18. [18]

      J.X. Liu, X.B. Zhang, Y.L. Li, et al., Coord. Chem. Rev. 414 (2020) 213260–213276. doi: 10.1016/j.ccr.2020.213260

    19. [19]

      H.Y. Wang, X.Y. Zheng, L.S. Long, et al., Inorg. Chem. 60 (2021) 6790–6795. doi: 10.1021/acs.inorgchem.1c00622

    20. [20]

      B.S. Bassil, M. Ibrahim, R. Al-Oweini, et al., Angew. Chem. Int. Ed. 50 (2011) 5961–5964. doi: 10.1002/anie.201007617

    21. [21]

      B. Godin, Y.G. Chen, J. Vaissermann, et al., Angew. Chem. Int. Ed. 44 (2005) 3072–3075. doi: 10.1002/anie.200463033

    22. [22]

      X.B. Han, Y.G. Li, Z.M. Zhang, et al., J. Am. Chem. Soc. 137 (2015) 5486–5493. doi: 10.1021/jacs.5b01329

    23. [23]

      X.B. Han, Z.M. Zhang, T. Zhang, et al., J. Am. Chem. Soc. 136 (2014) 5359–5366. doi: 10.1021/ja412886e

    24. [24]

      L. Huang, S.S. Wang, J.W. Zhao, et al., J. Am. Chem. Soc. 136 (2014) 7637–7642. doi: 10.1021/ja413134w

    25. [25]

      M. Ibrahim, Y. Lan, B.S. Bassil, et al., Angew. Chem. Int. Ed. 50 (2011) 4708–4711. doi: 10.1002/anie.201100280

    26. [26]

      Z.J. Liu, X.L. Wang, C. Qin, et al., Coord. Chem. Rev. 313 (2016) 94–110. doi: 10.1016/j.ccr.2015.12.006

    27. [27]

      S.S. Mal, U. Kortz, Angew. Chem. Int. Ed. 44 (2005) 3777–3780. doi: 10.1002/anie.200500682

    28. [28]

      L. Qiao, M. Song, A. Geng, et al., Chin. Chem. Lett. 30 (2019) 1273–1276. doi: 10.1016/j.cclet.2019.01.024

    29. [29]

      Y. Sakai, K. Yoza, C.N. Kato, et al., Chem. Eur. J. 9 (2003) 4077–4083. doi: 10.1002/chem.200305182

    30. [30]

      X. Feng, W. Zhou, Y. Li, et al., Inorg. Chem. 51 (2012) 2722–2724. doi: 10.1021/ic202418y

    31. [31]

      M. Ibrahim, V. Mereacre, N. Leblanc, et al., Angew. Chem. Int. Ed. 54 (2015) 15574–15578. doi: 10.1002/anie.201504663

    32. [32]

      Y.D. Lin, R. Ge, C.B. Tian, et al., Chem. Commun. 57 (2021) 8624–8627. doi: 10.1039/d1cc03262c

    33. [33]

      E. Tanuhadi, E. Al-Sayed, G. Novitchi, et al., Inorg. Chem. 59 (2020) 8461–8467. doi: 10.1021/acs.inorgchem.0c00890

    34. [34]

      Y. Chen, Z.W. Guo, X.X. Li, et al., CCS Chem. 3 (2021) 1232–1241.

    35. [35]

      S.R. Li, H.Y. Wang, H.F. Su, et al., Small. Methods 5 (2020) 2000777.

    36. [36]

      H.H. Wu, S. Yao, Z.M. Zhang, et al., Dalton. Trans. 42 (2013) 342–346. doi: 10.1039/C2DT32318D

    37. [37]

      L.L. Li, G.J. Cao, J.W. Zhao, et al., Inorg. Chem. 55 (2016) 5671–5683. doi: 10.1021/acs.inorgchem.6b00754

    38. [38]

      G. Abbas, Y. Lan, V. Mereacre, et al., Inorg. Chem. 48 (2009) 9345–9355. doi: 10.1021/ic901248r

    39. [39]

      J. Bai, F. Li, Z. Sun, et al., Inorg. Chim. Acta 458 (2017) 1–7. doi: 10.1016/j.ica.2016.12.010

    40. [40]

      D. Barats, G. Leitus, R. Popovitz-Biro, et al., Angew. Chem. Int. Ed. 47 (2008) 9908–9912. doi: 10.1002/anie.200803966

    41. [41]

      M.P.E.M. Limanski, E. Droste, K. Burgemeister, et al., J. Cluster Sci. 13 (2002) 369–379. doi: 10.1023/A:1020551016077

    42. [42]

      P.T. M, J.P. Wang, Y. Shen, et al., Cryst. Growth. Des. 7 (2007) 603–605. doi: 10.1021/cg060751y

    43. [43]

      J.D. Compain, P. Mialane, A. Dolbecq, et al., Angew. Chem. Int. Ed. 48 (2009) 3007–3081.

    44. [44]

      M. Llunell, D. Casanova, J. Girera, P. Alemany, S. Alvarez, SHAPE, Version 2.0, Shape Software, Barcelona, 2010.

    45. [45]

      H. Zabrodsky, S. Peleg, D. Avnir, J. Am. Chem. Soc. 115 (1993) 8278–8289. doi: 10.1021/ja00071a042

    46. [46]

      M.N. Akhtar, V. Mereacre, G. Novitchi, et al., Chem. Eur. J. 15 (2009) 7278–7282. doi: 10.1002/chem.200900758

    47. [47]

      A. Baniodeh, C.E. Anson, A.K. Powell, Chem. Sci. 4 (2013) 4354–4361. doi: 10.1039/c3sc52041b

    48. [48]

      L.J. Chen, F. Zhang, X. Ma, et al., Dalton 44 (2015) 12598–12612. doi: 10.1039/C5DT00696A

    49. [49]

      S.F.M. Schmidt, C. Koo, V. Mereacre, J. Park, D.W. Heermann, et al., Inorg. Chem. 56 (2017) 4796–4806. doi: 10.1021/acs.inorgchem.6b02682

    50. [50]

      J.J. Jiang, Y.H. Chen, L.L. Liu, et al., Inorg. Chem. 58 (2019) 15853–15863. doi: 10.1021/acs.inorgchem.9b02273

    51. [51]

      X.Y. Zheng, H. Zhang, Z. Wang, et al., Angew. Chem. Int. Ed. 56 (2017) 11475–11479. doi: 10.1002/anie.201705697

    52. [52]

      G. Peng, Y.Y. Zhang, B. Li, et al., Dalton. Trans. 47 (2018) 17349–17356. doi: 10.1039/c8dt03613f

    53. [53]

      Y. Huo, Y.C. Chen, S.G. Wu, et al., Inorg. Chem. 58 (2019) 1301–1308. doi: 10.1021/acs.inorgchem.8b02788

    54. [54]

      K.C. Mondal, A. Sundt, Y.H. Lan, et al., Angew. Chem. Int. Ed. 51 (2012) 7550–7554. doi: 10.1002/anie.201201478

    55. [55]

      F. Shao, B. Cahier, Y.T. Wang, et al., Chem. Asian. J. 15 (2020) 391–397. doi: 10.1002/asia.201901511

    56. [56]

      J.J. Chen, M. Symes, L. Cronin, Nat. Chem. 10 (2018) 1042–1047. doi: 10.1038/s41557-018-0109-5

    57. [57]

      Y.Y. Liu, S.F. Lu, H.N. Wang, Adv. Energy Mater. 7 (2017) 1601224. doi: 10.1002/aenm.201601224

  • Figure 1  The structures of (a) [B-α-SbW9O33]9− unit; (b) [B-α-FeW9O34]11− unit. (c, d) Ball-and-stick/polyhedral view of [(B-α-FeW9O34)2Fe2Dy2(H2O)4]10−. (e) The metal core of 2Dy3+ and 4Fe3+ ions in cluster 1. (f-h) The coordination geometries of DyO8/FeO6/FeO4. Color key: W/WO6, blue; Fe/FeO6/FeO4, green; Dy/DyO8, purple; Sb, black and O, pink.

    Figure 2  χmT vs. T in the range of 2–300 K under an applied field of 1000 Oe for clusters 13. Inset: Enlarged view of the upturn of χmT (cluster 1).

    Figure 3  (a) In-phase susceptibility (χ') and (b) out-of-phase susceptibility (χ'') vs. v under zero Oe DC field for 1. (c) Cole-Cole plots for AC susceptibilities under zero DC field. The colored solid lines are the best fit to the generalized Debye model. (d) Arrhenius plots of lnτ vs. T–1 for 1 under zero DC field.

    Figure 4  (a) In-phase (χ') and (b) out-of-phase susceptibility (χ'') vs. v at indicated temperatures under a 2500 Oe DC field for cluster 1. (c) Cole-Cole plots for AC susceptibilities under a 2500 Oe DC field. The colored solid lines are the best fit to the generalized Debye model. (d) Arrhenius plots of lnτ vs. T–1 for cluster 1.

  • 加载中
计量
  • PDF下载量:  4
  • 文章访问数:  288
  • HTML全文浏览量:  35
文章相关
  • 发布日期:  2023-03-15
  • 收稿日期:  2022-01-26
  • 接受日期:  2022-02-21
  • 修回日期:  2022-02-16
  • 网络出版日期:  2022-02-24
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章