Hemiporphyrazine-Involved Sandwich Dysprosium Double-Decker

Sep 19, 2018 - Dysprosium Compounds with Hula-Hoop-like Geometries: The Influence of Magnetic Anisotropy and Magnetic Interactions on Magnetic ...
0 downloads 0 Views 2MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Hemiporphyrazine-Involved Sandwich Dysprosium Double-Decker Single-Ion Magnets Wenbo Liu,† Suyuan Zeng,‡ Xin Chen,† Houhe Pan,† Dongdong Qi,† Kang Wang,*,† Jianmin Dou,*,‡ and Jianzhuang Jiang*,†

Inorg. Chem. Downloaded from pubs.acs.org by RENSSELAER POLYTECHNIC INST on 09/20/18. For personal use only.



Beijing Key Laboratory for Science and Application of Functional Molecular and Crystalline Materials, Department of Chemistry, University of Science and Technology Beijing, Beijing 100083, China ‡ Department of Chemistry, Liaocheng University, Liaocheng, 252059, China S Supporting Information *

ABSTRACT: Both heteroleptic (phthalocyaninato)(hemiporphyrazinato) and homoleptic bis(hemiporphyrazinato) dysprosium double-decker complexes, Dy[H(Hp)2] (1) and Dy[H(Pc)(Hp)] (2) (H2Pc = metal-free phthalocyanine; H2Hp = metal-free hemiporphyrazine), were designed, synthesized, and structurally characterized. The dysprosium center in both doubledeckers are octa-coordinated with a nearly ideal squareantiprismatic coordination geometry, which provides an increased molecular anisotropy for the dysprosium ion and ensures the strengthened magnetic properties of both single-ion magnets (SIMs) in terms of coordination geometry. Magnetic studies reveal that both double-deckers exhibit typical SIM behavior with a spin reversal energy barrier of 80.1 ± 6.3 K for 1 and 57.3 ± 3.8 K for 2 as well as the hysteresis loops emerging at 3 K. In particular, introduction of two Hp ligands with four pyridine nitrogen atoms coordinated with the dysprosium spin center endows Dy[H(Hp)2] (1) with the thus far highest energy barrier among the sandwich-type dysprosium SIMs with N4macrocyclic ligands, revealing the potential applications of sandwich-type lanthanide complexes with Hp ligands in molecularbased information storage.



INTRODUCTION Owing to the strong intermolecular π−π interaction(s) and special f electronic configuration of the lanthanide ion(s), sandwich-type tetrapyrrole, i.e., phthalocyanine/porphyrin, lanthanide multiple-decker complexes have been intensively studied as excellent organic semiconductors and typical single molecule magnets (SMMs).1−3 For the purpose of developing new advanced functional molecular materials, a number of conjugated macrocyclic ligands such as corroles, Schiff-base, and calix[4]arene were introduced onto the sandwich lanthanide structures in the past few years, resulting in a series of novel sandwich-type multiple-deckers including Dy2[Pc(OC 5 H 11 ) 8 ] 2 [Cor(FPh) 3 ], HEu 2 [Pc(R) 8 ] 2 [Cor(ClPh) 3 ], [KM(Pc)(L)CH3OH], and [Ln2(Hhms)2(NO3)4]·MeCN.4−6 Quite lately, a new conjugated ligand with a unique antiaromatic electronic structure, hemiporphyrazine (Hp), was also used to construct the sandwich lanthanide doubledecker complexes Ln[H(Pc)(Hp)] and Ln[H(Hp)2] (Ln = Eu, Lu; H2Pc = metal-free phthalocyanine; H2Hp = metal-free hemiporphyrazine), with their electronic structure systematically investigated.7 However, the functional properties of this novel series of sandwich compounds still remain unexplored. It has been revealed that due to the specific electronic configuration within the whole series of lanthanide ions, dysprosium was usually selected to construct corresponding © XXXX American Chemical Society

SMMs as well as single-ion magnets (SIMs) (SMM containing only one spin center) toward clarifying the structure− functionality relationship.8 As a result, in the present work, dysprosium was also employed to synthesize Dy[H(Hp)2] (1) and Dy[H(Pc)(Hp)] (2), Scheme 1. In these compounds, the Dy ions lie in the square-antiprismatic (SA) coordination geometry as revealed by single crystal X-ray diffraction analysis, which has been revealed to be the necessary requirement for good SIM performance.9 Meanwhile, unlike phthalocyanines and porphyrins with the tetrapyrrole structure, Hp bears two opposite facing pyridine rings as well as two opposite facing pyrrole units. It was already uncovered that the lone electron pair of pyridine nitrogen atom is more localized in comparison with the isoindole/pyrrole nitrogen atoms, leading to the higher electronic cloud density on the pyridine nitrogen atom. This provides an intensified coordination field strength, which was also proven to benefit the magnetic properties of the sandwich-type Tb/Dy SMMs including SIMs.10 As expected, magnetic studies reveal the typical SIM nature of both doubledeckers. In particular, owing to the introduction of two Hp ligands with four pyridine nitrogen atoms coordinated with the dysprosium spin center, Dy[H(Hp)2] (1) exhibits the largest Received: July 22, 2018

A

DOI: 10.1021/acs.inorgchem.8b02068 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Scheme 1. Molecular Structures of Dy[H(Hp)] (1) and Dy[H(Pc)(Hp)] (2)

Figure 1. Molecular structures of Dy[H(Pc)(Hp)] (2) (a) and Dy[H(Hp)2] (1) (b) in side (left) and top (right) view, respectively, Dy: red, N: blue, and C: off-white. For clarity, all the hydrogen atoms have been omitted.

This indicates the large molecular magnetic anisotropy of this dysprosium spin carrier-based SIM, which in turn ensures its strengthened magnetic behavior in terms of coordination geometry.9,10 This is also true for 1, in which the central Dy ion employs an octa-coordinated form complexed by four isoindole nitrogen and four pyridine nitrogen atoms from two Hp ligands, with the average skew angle φ of 38.07°. In the double-decker 2, the distances of the dysprosium center to the N4 mean planes of Hp and Pc are 1.406 and 1.388 Å, respectively, leading to a ring-to-ring distance of 2.794 Å. The Dy−N distances between Dy and the coordinated pyridine N atoms locate in the scope of 2.629−2.651 Å, while that between Dy and the coordinated isoindole N atoms of the Hp and Pc lies in the range of 2.369−2.431 Å. As for 1, Dy locates at 2.344−2.421 and 2.605−2.638 Å from the isoindole N and pyridine N atoms of the Hp ligands, respectively, giving a distance of 1.396 Å for Dy to the N4 mean plane of each Hp ligand and leading to a ring-to-ring distance of 2.792 Å. In particular, in both double-deckers, one pyridine moiety of the Hp ligand deviates significantly from the Dy ion, resulting in a dihedral angle between the NC5 mean plane (the pyridine ring) and N4 mean plane (Hp ligand), 14.17° for 1 and 29.11° for 2, owing to the existence of the H atom, which binds to the

energy barrier among the thus far reported sandwich-type dysprosium SIMs with N4-macrocyclic ligands.



RESULTS AND DISCUSSION Structural Studies. The protonated Dy[H(Hp)] (1) and Dy[H(Pc)(Hp)] (2) were prepared by using the previous procedure7 and characterized by a series of spectroscopic methods as detailed in the Supporting Information. Single crystals of 1 and 2 applicable to X-ray diffraction analysis were prepared by diffusing methanol vapor into their tetrahydrofuran solution of respective species. 1 and 2 crystallize in the same monoclinic system (P21/c group) with each unit cell containing four double-decker molecules, Tables S1 and S2 (Supporting Information). Figure 1 shows the molecular structures of 1 and 2 in side and top views, unambiguously revealing their sandwich-type double-decker molecular structure. As can be seen, the dysprosium center in 2 is octacoordinated by the six isoindole nitrogen atoms and two pyridine nitrogen atoms of the Pc ligand and the Hp ligand, with the skew angle φ amounting to 44.01°. This is similar to that of [Dy(Pc)2]−·[(C4H9)4N]+11a and Dy(Pc)2,11b 44.17° and 41.4°, respectively, resulting in a nearly ideal SA coordination geometry (φ = 45°) for the dysprosium center. B

DOI: 10.1021/acs.inorgchem.8b02068 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. In-phase (χ′) and out-of-phase (χ′′) frequency dependence in a zero dc filed for Dy[H(Hp)2] (1) [(a) χ′ and (b) χ′′] and Dy[H(Pc)(Hp)] (2) [(c) χ′ and (d) χ′′].

Figure 3. Comparison in the plots of ln(τ) vs. T−1 for the Dy[H(Hp)2] (1) (a) and Dy[H(Pc)(Hp)] (2) (b) based on the Cole−Cole fitting data under 0 Oe dc field. Hysteresis loops for the Dy[H(Hp)2] (1) (c) and Dy[H(Pc)(Hp)] (2) (d) at 2 and 3 K.

Magnetic Properties. Direct current (DC) magnetic measurements were taken on polycrystalline samples to study the static magnetic properties of 1 and 2 in 2−300 K with an applied field of 1000 Oe. It is worth noting that the experimental powder X-ray spectra of both polycrystalline samples are in line with the simulated ones from single crystal X-ray diffraction analysis, Figure S3 (Supporting Information). This result clearly reveals that the structure and composition of both polycrystalline samples are the same as those revealed by single crystal X-ray diffraction analysis mentioned above. The curves of the magnetic susceptibility for 1 and 2 display

nitrogen atom of the opposite pyridine moiety. This in turn reveals the protonated nature of both double-decker compounds.7 In addition, the crystal packing diagrams of 1 and 2 are shown in Figure S1 and S2 (Supporting Information) with corresponding data summarized in Table S2 (Supporting Information). As can be found, the dysprosium double-decker molecules in the single crystal of both 1 and 2 are well separated, with the nearest Dy···Dy distance being 9.178 Å for 1 and 9.318 Å for 2, which is enough to avoid the obvious intermolecular magnetic interaction in these two doubledeckers.3e,g,12 C

DOI: 10.1021/acs.inorgchem.8b02068 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. Electrostatic potential around the Dy3+ (2.5 Å radius sphere) in Dy[H(Pc)2] (left), Dy[H(Pc)(Hp)] (2) (middle), and Dy[H(Hp)2] (1) (right) caused by Pc and Hp ligands.

(Figure 3a,b). Since the relaxation process for the lanthanidebased SIMs usually combines the following four processes, i.e., Raman process, direct process, Orbach process, and QTM process, we therefore first tried to fit ln τ vs 1/T plots by considering all of these four relaxation processes by means of the following multiple relaxation equation [eq 1]:

evident temperature-dependent character, Figure S4 (Supporting Information). The χMT values of 14.23 and 13.11 cm3·K· mol−1 for 1 and 2 at 300 K are consistent with the expected value of 14.17 cm3·K·mol−1 expected for a free Dy3+ [6H15/2, S = 5/2, L = 5, g = 4/3].8,13 With lowering of the temperature, the χMT values for 1 and 2 decrease slowly until about 8 K and then decrease quickly to a minimum value of 5.40 and 9.42 cm3·K·mol−1 at 2 K, respectively, due to the excited Stark sublevels depopulation with the help of the crystal field effect.8,13 Moreover, 1 and 2 at 2 K and 50 kOe exhibit the maximum M(H) value of 3.33 Nβ and 4.62 Nβ, respectively, Figure S5 (Supporting Information). Both of which deviate obviously from the theoretically expected magnetization saturation data, 10 Nβ.8,13 This, in combination with the non-superposition field-dependence magnetization curves obtained at 2.0, 3.0, and 5.0 K for the 1 and 2, indicates the big magnetic anisotropy for the Dy ion in the two complexes.8,13 Toward further understanding the magnetic behavior of the two complexes, studies over the magnetic susceptibilities in both the variable-temperature and variable-frequency manner for 1 and 2 were performed under zero dc field (in a 2.0 Oe ac field and oscillating range of 1−999 Hz), Figures 2 and S6 (Supporting Information). These two complexes display not only the temperature- but also the frequency-dependent characteristics in both the in-phase (χ′) and out-of-phase (χ′′) signals, indicating the slow magnetization relaxation and in particular the SIM nature of both compounds. In the χ′′−T plots shown in Figure S6 (Supporting Information), clear χ′′ peaks can be found in the temperature range of 10−12 K for 1 and 9−11 K for 2, suggesting the existence of a thermalactivated relaxation process. While in the low temperature region, the increase of χ′′ is observed along with lowering the temperature, which is owing to the quantum tunneling (QTM) process. To better understand the magnetic processes of 1 and 2, the data from the frequency-dependent measurements were fitted by the generalized Debye model,14 giving the graphics of χ′′ versus χ′ (Cole−Cole plot) of 1 double-decker (2−14 K) and 2 double-decker (2−14 K), respectively, Figure S7 (Supporting Information). As usual, fitting of the experimental data gives the following sets of parameters, in detail α = 0.0294−0.216 for 1 and α = 0.088−0.187 for 2, respectively, Tables S3 and S4 (Supporting Information). The relative broad distribution of α values for the two SIMs suggests the existence of two or more magnetic relaxation processes in both compounds. The relaxation time (τ) has been exploited by fitting the frequency dependent result using the generalized employed Debye model, which is plotted by ln τ vs 1/T

i U y −1 τ −1 = τ0−1 expjjj− eff zzz + AT + CT n + τQTM T k {

(1)

However, counting the direct and Raman processes for the 1 in the fitting gives no reasonable result. This is also true for 2. As a result, the Orbach and QTM processes for both 1 and 2 were considered for fitting by the following equations [eq 2]. i U y −1 τ −1 = τ0−1 expjjj− eff zzz + τQTM T k {

(2)

Figure 3a,b shows the fitting of the ln τ vs 1/T plots, which agrees well with the data in the full temperature range, affording the parameters as Ueff = 80.1 ± 6.3 K, τ0 = (3.05 ± 0.66) × 10−7 s, and τQTM = 0.00481 s for 1 and Ueff = 57.3 ± 3.8 K, τ0 = (1.92 ± 0.62) × 10−6 s, and τQTM = 0.00131 s for 2. These results confirm the two magnetic relaxation processes (Orbach and QTM processes) in both double-decker SIMs. In addition, at 3 K the hysteresis loops emerged for both 1 and 2, Figure 3c,d, reconfirming their SIM nature. Moreover, the hysteresis loops of 1 are larger than 2 at the same temperature, in line with the larger energy barrier of 1 compared to 2, revealing its better SIM performance. Density Functional Theory (DFT) Calculations. It has been revealed that the energy barrier for the sandwich-type Dy SIMs with macrocyclic ligands are mainly determined by the coordination geometry and field strength on the spin centers.9,10,15 In the case of Dy[H(Hp)2] (1) and Dy[H(Pc)(Hp)] (2), their spin centers lie in the similar squareantiprismatic coordination geometry with octa-coordinated by eight N atoms. Therefore, the coordination field strength of the spin center plays the key role in the difference of the Ueff between 1 and 2. As detailed in the above-described crystal structure section, the dysprosium center in 2 is octacoordinated by the six isoindole nitrogen atoms and two pyridine nitrogen atoms from one Pc and one Hp ligands, and in 1 the central Dy ion is also octa-coordinated with four pyridine nitrogen and four isoindole nitrogen atoms from the two Hp ligands. As detailed in our previous report,7 the lone electron pair of N in pyridine is more localized in comparison with the isoindole nitrogen atom. This in turn results in a higher electronic cloud density on the pyridine nitrogen atom D

DOI: 10.1021/acs.inorgchem.8b02068 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

atmosphere for 4 h. The reaction mixture was then cooled to room temperature. The volatiles were removed under reduced pressure, and the residue was purified by chromatography on a silica-gel column with CH2Cl2/CH3OH (20:1 v/v) as eluent. The blue fraction containing corresponding target compound was collected and evaporated. Repeated chromatography followed by recrystallization from CH2Cl2 and CH3OH provided 2 as a blue powder in the yield of 27.1%. MS (MALDI−TOF) m/z: 1117.012 [M + H]+. Elemental analysis (%) calcd. for C58H34N16Dy·0.5CH2Cl2: C 60.57, H 3.04, N 19.32; found: C 60.15, H 3.02, N 18.81. UV−vis (DMSO) λ/nm (log ε): 328 (4.88), 351 (4.87), 423 (4.40), 562 (3.93), 696 (4.67).

in comparison with the isoindole nitrogen atom. As a consequence, the Hp ligand affords a more intensified coordination field strength than the Pc ligand. Additional support for this point comes from the DFT calculation result. As can be seen in Figure 4 and Table S5 (Supporting Information), the electrostatic potential around the Dy3+ ion from the isoindole N atoms of both Hp and Pc ligands in 1 and 2 are in the range of −1.01 to −3.72 au, which increases to ca. −11.50 au for the pyridine nitrogen atoms of Hp ligands in 1 and 2. As a total result, the more intensified coordination field strength for the dysprosium center in Dy[H(Hp)2] (1) than in Dy[H(Pc)(Hp)] (2) helps to generate larger splitting between the ground and excited mJ, leading to an elevated energy barrier for the former species over the latter one. This also rationalizes the superior magnetic properties of Dy[H(Hp)2] (1) over [Dy(Pc)2]−·[(C4H9)4N]+ in terms of the spin reversal energy barrier.3a Actually, owing to the intensified coordination field strength in Dy[H(Hp)2] (1), this SIM displays the highest energy barrier, 80.1 K, among the thus far reported sandwich-type Dy SIMs with N4-macrocyclic ligands, Table S6 (Supporting Information).



The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b02068. General procedures, spectroscopic characterization. details of the DFT calculation. Molecular packing of 1 and 2. Details of magnetic properties for 1 and 2. MALDITOF mass, electronic absorption, and IR spectra for 1 and 2. Crystal data and structure refinements, relaxation fitting parameters, elemental analytical and mass spectrometric data, and electronic absorption data for 1 and 2 (PDF)



CONCLUSION Briefly summarizing above, both heteroleptic (phthalocyaninato)(hemiporphyrazinato) and homoleptic bis(hemiporphyrazine) dysprosium double-decker complexes have been synthesized and structurally characterized. Both compounds display typical slow magnetic relaxation behavior expected for SIMs with a spin reversal energy barrier of 57.3 ± 3.8 and 80.1 ± 6.3 K, respectively. In particular, owing to the introduction of two Hp ligands with four pyridine nitrogen atoms coordinated with the dysprosium spin center, the energy barrier of Dy[H(Hp)2] (1) represents the largest one among the thus far reported sandwich-type dysprosium SIMs with N4macrocyclic ligands, revealing the potential applications of sandwich lanthanide compounds with Hp ligands in molecularbased information storage. This result will be helpful for the design and synthesis of novel macrocyclic ligand-based sandwich-type lanthanide SIMs/SMMs with good performance.



ASSOCIATED CONTENT

S Supporting Information *

Accession Codes

CCDC 1557721 and 1558646 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Authors

*(J.J.) E-mail: [email protected]. *(J.D.) E-mail: [email protected]. *(K.W.) E-mail: [email protected]. ORCID

Suyuan Zeng: 0000-0002-6421-8565 Jianzhuang Jiang: 0000-0002-4263-9211

EXPERIMENTAL SECTION

Notes

General Procedures. n-Pentanol was distilled freshly from sodium under Ar. Anhydrous 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU), phthalonitrile, and 2,6-diaminopyridine were purchased from Aldrich. The compounds Dy(acac)3·nH2O,16 Dy(Pc)(acac)· nH2O,17 and H2Hp18 were prepared according to the literature methods. All other reagents and solvents were of reagent grade and used as received. Synthesis of Dy[H(Hp)2] (1). A mixture of hemiporphyrazine (88.0 mg, 0.20 mmol), Dy(acac)3·nH2O (51 mg, 0.10 mmol), and DBU (0.1 mL) in n-pentanol (1 mL) was refluxed under N2 atmosphere for 4 h. The reaction mixture was cooled to room temperature. The volatiles were removed under reduced pressure, and the residue was purified by chromatography on a silica-gel column with CH2Cl2/CH3OH (20:1 v/v) as eluent. The brown fraction was collected and evaporated. Repeated chromatography followed by recrystallization from CH2Cl2 and CH3OH provided 1 as a brown powder in the yield of 19.1%. MS (MALDI−TOF) m/z: 1043.020 [M + H]+. Elemental analysis calcd. (%) for C52H29N16Dy·1.5CH3OH: C 59.04, H 3.24, N 20.59; found: C 59.19, H 3.23, N 20.12. UV−vis (DMSO) λ/nm (log ε): 363 (4.86), 398 (4.74). Synthesis of Dy[H(Pc)(Hp)] (2). A mixture of hemiporphyrazine (44.0 mg, 0.10 mmol), Dy(Pc)(acac)·nH2O (78.0 mg, 0.10 mmol), and DBU (0.1 mL) in n-pentanol (1 mL) was refluxed under N2

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support from the Natural Science Foundation of China (Nos. 21631003, 21671017, and 21401009), Fundamental Research Funds for the Central Universities (No. FRFBD-17-016A), Beijing Municipal Commission of Education, and University of Science and Technology Beijing is gratefully acknowledged.



REFERENCES

(1) (a) Jiang, J.; Ng, D. K. P. A Decade Journey in the Chemistry of Sandwich-Type Tetrapyrrolato-Rare Earth Complexes. Acc. Chem. Res. 2009, 42 (1), 79−88. (b) Wang, H.; Wang, B.-W.; Bian, Y.; Gao, S.; Jiang, J. Single-molecule magnetism of tetrapyrrole lanthanide compounds with sandwich multiple-decker structures. Coord. Chem. Rev. 2016, 306, 195−216. (c) Functional Phthalocyanine Molecular Materials, Structure and Bonding; Jiang, J., Vol. ed.; Mingos, D. M. P., Series ed.; Springer: Heidelberg, Germany, 2010. (d) Bian, Y.; Jiang, J. In 50 Years of Structure and Bonding − The Anniversary Volume; Mingos, D. M. P., Ed.; Springer International Publishing: Springer-

E

DOI: 10.1021/acs.inorgchem.8b02068 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Verlag Berlin Heidelberg, 2015; pp 159−199. (e) Liu, J. L.; Chen, Y. C.; Tong, M. L. Symmetry strategies for high performance lanthanidebased single-molecule magnets. Chem. Soc. Rev. 2018, 47 (7), 2431− 2453. (2) (a) Chen, Y.; Su, W.; Bai, M.; Jiang, J.; Li, X.; Liu, Y.; Wang, L.; Wang, S. High Performance Organic Field-Effect Transistors Based on Amphiphilic Tris(phthalocyaninato) Rare Earth Triple-Decker Complexes. J. Am. Chem. Soc. 2005, 127, 15700−15701. (b) Kan, J.; Chen, Y.; Qi, D.; Liu, Y.; Jiang, J. High-performance air-stable ambipolar organic field-effect transistor based on tris(phthalocyaninato) europium(III). Adv. Mater. 2012, 24 (13), 1755−1758. (c) Kong, X.; Zhang, X.; Gao, D.; Qi, D.; Chen, Y.; Jiang, J. Air-stable ambipolar field-effect transistor based on a solution-processed octanaphthoxy-substituted tris(phthalocyaninato) europium semiconductor with high and balanced carrier mobilities. Chem. Sci. 2015, 6 (3), 1967−1972. (d) Lu, G.; Kong, X.; Ma, P.; Wang, K.; Chen, Y.; Jiang, J. Amphiphilic (Phthalocyaninato) (Porphyrinato) Europium Triple-Decker Nanoribbons with Air-Stable Ambipolar OFET Performance. ACS Appl. Mater. Interfaces 2016, 8 (9), 6174−6182. (e) Lu, G.; Kong, X.; Sun, J.; Zhang, L.; Chen, Y.; Jiang, J. Solution-processed single crystal microsheets of a novel dimeric phthalocyanine-involved triple-decker for high-performance ambipolar organic field effect transistors. Chem. Commun. 2017, 53 (95), 12754−12757. (f) Lu, G.; Wang, K.; Kong, X.; Pan, H.; Zhang, J.; Chen, Y.; Jiang, J. Binuclear Phthalocyanine Dimer-Containing Yttrium Double-Decker Ambipolar Semiconductor with Sensitive Response toward Oxidizing NO2 and Reducing NH3. ChemElectroChem 2018, 5 (4), 605−609. (g) Dong, Z.; Kong, X.; Qi, D.; Zhao, S.; Li, X.; Chen, Y.; Jiang, J. Polymorphism in the self-assembled nanostructures of a tris(phthalocyaninato) europium derivative: Phase-dependent semiconducting and NO2 sensing behaviour. Org. Electron. 2018, 53, 127−134. (h) Lapkina, L. A.; Larchenko, V. E.; Kirakosyan, G. A.; Tsivadze, A. Y.; Troyanov, S. I.; Gorbunova, Y. G. Cation-Induced Dimerization of Crown-Substituted Phthalocyanines by Complexation with Rubidium Nicotinate As Revealed by X-ray Structural Data. Inorg. Chem. 2018, 57 (1), 82−85. (i) Birin, K. P.; Poddubnaya, A. I.; Isanbaeva, E. V.; Gorbunova, Y. G.; Tsivadze, A. Y. Crown-interlocked lanthanide diphthalocyaninates with switchable panchromatic absorption. J. Porphyrins Phthalocyanines 2017, 21 (04− 06), 406−415. (j) Oluwole, D. O.; Yagodin, A. V.; Mkhize, N. C.; Sekhosana, K. E.; Martynov, A. G.; Gorbunova, Y. G.; Tsivadze, A. Y.; Nyokong, T. First Example of Nonlinear Optical Materials Based on Nanoconjugates of Sandwich Phthalocyanines with Quantum Dots. Chem. - Eur. J. 2017, 23 (12), 2820−2830. (k) Birin, K. P.; Gorbunova, Y. G.; Tsivadze, A. Y. Efficient scrambling-free synthesis of heteroleptic terbium triple-decker (porphyrinato)(crown-phthalocyaninates). Dalton Trans. 2012, 41 (32), 9672−9681. (3) (a) Ishikawa, N.; Sugita, M.; Ishikawa, T.; Koshihara, S.-Y.; Kaizu, Y. Lanthanide Double-Decker Complexes Functioning as Magnets at the Single-Molecular Level. J. Am. Chem. Soc. 2003, 125 (29), 8694−8695. (b) Wang, H.; Qian, K.; Wang, K.; Bian, Y.; Jiang, J.; Gao, S. Sandwich-type tetrakis-(phthalocyaninato) dysprosiumcadmium quadruple-decker SMM. Chem. Commun. 2011, 47 (34), 9624−9626. (c) Wang, K.; Qi, D.; Wang, H.; Cao, W.; Li, W.; Liu, T.; Duan, C.; Jiang, J. Binuclear Phthalocyanine-Based Sandwich-Type Rare Earth Complexes: Unprecedented Two π-Bridged BiradicalMetal Integrated SMMs. Chem. - Eur. J. 2013, 19 (34), 11162−11166. (d) Wang, H.; Qian, K.; Qi, D.; Cao, W.; Wang, K.; Gao, S.; Jiang, J. Co-crystallized fullerene and a mixed (phthalocyaninato)(porphyrinato) dysprosium double-decker SMM. Chem. Sci. 2014, 5 (8), 3214−3220. (e) Cao, W.; Gao, C.; Zhang, Y.-Q.; Qi, D.; Liu, T.; Wang, K.; Duan, C.; Gao, S.; Jiang, J. Rational enhancement of the energy barrier of bis(tetrapyrrole) dysprosium SMMs via replacing atom of porphyrin core. Chem. Sci. 2015, 6 (10), 5947−5954. (f) Thiele, S.; Balestro, F.; Ballou, R.; Klyatskaya, S.; Ruben, M.; Wernsdorfer, W. Electrically driven nuclear spin resonance in singlemolecule magnets. Science 2014, 344 (6188), 1135−1138. (g) Chen, Y.; Ma, F.; Chen, X.; Dong, B.; Wang, K.; Jiang, S.; Wang, C.; Chen, X.; Qi, D.; Sun, H.; Wang, B.; Gao, S.; Jiang, J. A New

Bis(phthalocyaninato) Terbium Single-Ion Magnet with an Overall Excellent Magnetic Performance. Inorg. Chem. 2017, 56 (22), 13889− 13896. (h) Chen, Y.; Ma, F.; Chen, X.; Dong, B.; Wang, K.; Jiang, S.; Wang, C.; Chen, X.; Qi, D.; Sun, H.; Wang, B.; Gao, S.; Jiang, J. Novel bis(phthalocyaninato) rare earth complexes with the bulky and strong electron-donating dibutylamino groups: synthesis, spectroscopy, and SMM properties. Inorg. Chem. Front. 2017, 4, 1465−1471. (i) Wang, K.; Ma, F.; Qi, D.; Chen, X.; Chen, Y.; Chen, Y.-C.; Sun, H.-L.; Tong, M.-L.; Jiang, J. Chiral bis(phthalocyaninato) terbium double-decker compounds with enhanced single-ion magnetic behavior. Inorg. Chem. Front. 2018, 5 (4), 939−943. (j) Katoh, K.; Yamashita, S.; Yasuda, N.; Kitagawa, Y.; Breedlove, B.; Nakazawa, Y.; Yamashita, M. Control of the Spin Dynamics of Single-Molecule Magnets by using a Quasi One-Dimensional Arrangement. Angew. Chem., Int. Ed. 2018, 57, 1−7. (k) Morita, T.; Damjanovic, M.; Katoh, K.; Kitagawa, Y.; Yasuda, N.; Lan, Y.; Wernsdorfer, W.; Breedlove, B. K.; Enders, M.; Yamashita, M. Comparison of the Magnetic Anisotropy and Spin Relaxation Phenomenon of Dinuclear Terbium(III) Phthalocyaninato SingleMolecule Magnets Using the Geometric Spin Arrangement. J. Am. Chem. Soc. 2018, 140 (8), 2995−3007. (4) (a) Buckley, H. L.; Anstey, M. R.; Gryko, D. T.; Arnold, J. Lanthanide corroles: a new class of macrocyclic lanthanide complexes. Chem. Commun. 2013, 49 (30), 3104−3106. (b) Buckley, H. L.; Arnold, J. Recent developments in out-of-plane metallocorrole chemistry across the periodic table. Dalton Trans. 2015, 44 (1), 30−36. (c) Lu, G.; Yan, S.; Shi, M.; Yu, W.; Li, J.; Zhu, W.; Ou, Z.; Kadish, K. M. A new class of rare earth tetrapyrrole sandwich complexes containing corrole and phthalocyanine macrocycles: synthesis, physicochemical characterization and X-ray analysis. Chem. Commun. 2015, 51 (12), 2411−2413. (d) Alemayehu, A. B.; Vazquez-Lima, H.; McCormick, L. J.; Ghosh, A. Relativistic effects in metallocorroles: comparison of molybdenum and tungsten biscorroles. Chem. Commun. 2017, 53 (43), 5830−5833. (e) Lu, G.; He, C.; Wang, K.; Sun, J.; Qi, D.; Gong, L.; Wang, C.; Ou, Z.; Yan, S.; Zeng, S.; Zhu, W. Dysprosium Heteroleptic Corrole-Phthalocyanine TripleDecker Complexes: Synthesis, Crystal Structure, and Electrochemical and Magnetic Properties. Inorg. Chem. 2017, 56 (19), 11503−11512. (5) (a) Wang, H.; Cao, W.; Liu, T.; Duan, C.; Jiang, J. Synthesis, structure, and single-molecule magnetic properties of rare-earth sandwich complexes with mixed phthalocyanine and Schiff base ligands. Chem. - Eur. J. 2013, 19 (7), 2266−2270. (b) Ma, Q.; Zeng, S.; Feng, X.; Cao, W.; Wang, H.; Dou, J.; Jiang, J. A Mixed PorphyrinSchiff Base Dysprosium(III) Single-Molecule Magnet. Eur. J. Inorg. Chem. 2016, 2016, 4194−4198. (c) Gao, F.; Cui, L.; Liu, W.; Hu, L.; Zhong, Y. W.; Li, Y. Z.; Zuo, J. L. Seven-coordinate lanthanide sandwich-type complexes with a tetrathiafulvalene-fused Schiff base ligand. Inorg. Chem. 2013, 52 (19), 11164−11172. (d) Gao, F.; Zhang, X. M.; Cui, L.; Deng, K.; Zeng, Q. D.; Zuo, J. L. Tetrathiafulvalene-supported triple-decker phthalocyaninato dysprosium(III) complex: synthesis, properties and surface assembly. Sci. Rep. 2015, 4, 5928. (e) Gao, F.; Feng, X.; Yang, L.; Chen, X. New sandwich-type lanthanide complexes based on closed-macrocyclic Schiff base and phthalocyanine molecules. Dalton Trans. 2016, 45 (17), 7476−7482. (6) (a) Liu, C.-M.; Zhang, D.-Q.; Hao, X.; Zhu, D.-B. Syntheses, Crystal Structures, and Magnetic Properties of Two p-tertButylsulfonylcalix[4]arene Supported Cluster Complexes with a Totally Disordered Ln4(OH)4 Cubane Core. Cryst. Growth Des. 2012, 12 (6), 2948−2954. (b) Deacon, G. B.; Gardiner, M. G.; Junk, P. C.; Townley, J. P.; Wang, J. Rare-Earth Metalation of Calix[4]pyrrole/Calix[4]arene Free of Alkali-Metal Companions. Organometallics 2012, 31 (10), 3857−3864. (7) Liu, W.; Pan, H.; Wang, Z.; Wang, K.; Qi, D.; Jiang, J. Sandwich rare earth complexes simultaneously involving aromatic phthalocyanine and antiaromatic hemiporphyrazine ligands showing a predominantly aromatic nature. Chem. Commun. 2017, 53, 3765−3768. (8) (a) Liu, J.-L.; Chen, Y.-C.; Zheng, Y.-Z.; Lin, W.-Q.; Ungur, L.; Wernsdorfer, W.; Chibotaru, L. F.; Tong, M.-L. Switching the anisotropy barrier of a single-ion magnet by symmetry change from F

DOI: 10.1021/acs.inorgchem.8b02068 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry quasi-D5h to quasi-Oh. Chem. Sci. 2013, 4 (8), 3310−3316. (b) Guo, F. S.; Day, B. M.; Chen, Y. C.; Tong, M. L.; Mansikkamaki, A.; Layfield, R. A. A Dysprosium Metallocene Single-Molecule Magnet Functioning at the Axial Limit. Angew. Chem., Int. Ed. 2017, 56 (38), 11445−11449. (c) Liu, J.; Chen, Y. C.; Liu, J. L.; Vieru, V.; Ungur, L.; Jia, J. H.; Chibotaru, L. F.; Lan, Y.; Wernsdorfer, W.; Gao, S.; Chen, X. M.; Tong, M. L. A Stable Pentagonal Bipyramidal Dy(III) SingleIon Magnet with a Record Magnetization Reversal Barrier over 1000 K. J. Am. Chem. Soc. 2016, 138 (16), 5441−5450. (d) Chen, Y. C.; Liu, J. L.; Ungur, L.; Liu, J.; Li, Q. W.; Wang, L. F.; Ni, Z. P.; Chibotaru, L. F.; Chen, X. M.; Tong, M. L. Symmetry-Supported Magnetic Blocking at 20 K in Pentagonal Bipyramidal Dy(III) SingleIon Magnets. J. Am. Chem. Soc. 2016, 138 (8), 2829−2837. (e) Guo, Y. N.; Chen, X. H.; Xue, S.; Tang, J. Modulating magnetic dynamics of three Dy2 complexes through keto-enol tautomerism of the ovanillin picolinoylhydrazone ligand. Inorg. Chem. 2011, 50 (19), 9705−9713. (f) Guo, Y. N.; Xu, G. F.; Wernsdorfer, W.; Ungur, L.; Guo, Y.; Tang, J.; Zhang, H. J.; Chibotaru, L. F.; Powell, A. K. Strong axiality and Ising exchange interaction suppress zero-field tunneling of magnetization of an asymmetric Dy2 single-molecule magnet. J. Am. Chem. Soc. 2011, 133 (31), 11948−11951. (g) Wu, J.; Cador, O.; Li, X. L.; Zhao, L.; Le Guennic, B.; Tang, J. Axial Ligand Field in D4d Coordination Symmetry: Magnetic Relaxation of Dy SMMs Perturbed by Counteranions. Inorg. Chem. 2017, 56 (18), 11211− 11219. (h) Guo, M.; Wu, J.; Cador, O.; Lu, J.; Yin, B.; Le Guennic, B.; Tang, J. Manipulating the Relaxation of Quasi-D4d Dysprosium Compounds through Alternation of the O-Donor Ligands. Inorg. Chem. 2018, 57 (8), 4534−4542. (9) (a) Wang, H.; Wang, K.; Tao, J.; Jiang, J. Twist angle perturbation on mixed (phthalocyaninato)(porphyrinato) dysprosium(III) double-decker SMMs. Chem. Commun. 2012, 48 (24), 2973−2975. (b) Kan, J.; Wang, H.; Sun, W.; Cao, W.; Tao, J.; Jiang, J. Sandwich-type mixed tetrapyrrole rare-earth triple-decker compounds. Effect of the coordination geometry on the singlemolecule-magnet nature. Inorg. Chem. 2013, 52 (15), 8505−8510. (c) Shang, H.; Zeng, S.; Wang, H.; Dou, J.; Jiang, J. Peripheral Substitution: An Easy Way to Tuning the Magnetic Behavior of Tetrakis(phthalocyaninato) Dysprosium(III) SMMs. Sci. Rep. 2015, 5 (1), 8838. (10) (a) Takamatsu, S.; Ishikawa, N. A theoretical study of a drastic structural change of bis(phthalocyaninato)lanthanide by ligand oxidation: Towards control of ligand field strength and magnetism of single-lanthanide-ionic single molecule magnet. Polyhedron 2007, 26 (9−11), 1859−1862. (b) Ishikawa, N.; Mizuno, Y.; Takamatsu, S.; Ishikawa, T.; Koshihara, S.-Y. Effects of Chemically Induced Contraction of a Coordination Polyhedron on the Dynamical Magnetism of Bis(phthalocyaninato)disprosium, a Single-4f-Ionic Single-Molecule Magnet with a Kramers Ground State. Inorg. Chem. 2008, 47 (22), 10217−10219. (c) Chen, Y.; Liu, C.; Ma, F.; Qi, D.; Liu, Q.; Sun, H. L.; Jiang, J. Fabricating Bis(phthalocyaninato) Terbium SIM into Tetrakis(phthalocyaninato) Terbium SMM with Enhanced Performance through Sodium Coordination. Chem. - Eur. J. 2018, 24, 8066−8070. (d) Holmberg, R. J.; Polovkova, M. A.; Martynov, A. G.; Gorbunova, Y. G.; Murugesu, M. Impact of the coordination environment on the magnetic properties of singlemolecule magnets based on homo- and hetero-dinuclear terbium(III) heteroleptic tris(crownphthalocyaninate). Dalton Trans. 2016, 45 (22), 9320−9327. (11) (a) Loosli, C.; Liu, S.-X.; Neels, A.; Labat, G.; Decurtins, S. Crystal structures of tetrabutylammonium bis(phthalocyaninato)terbium(III) methanol solvate hydrate (1:1:3/2), [N(C4H9)4] [Tb(C8H4N2)2]·CH3OH·3/2H2O, and tetrabutylammonium bis(phthalocyaninato)dyspiosium(III) methanol solvate hydrate (1:1:1), [N(C4H9)4] [Dy(C8H4N2)2·CH3OH·H2O. Z. Kristallogr. New Cryst. Struct. 2006, 221, 135−141. (b) Katoh, K.; Yoshida, Y.; Yamashita, M.; Miyasaka, H.; Breedlove, B. K.; Kajiwara, T.; Takaishi, S.; Ishikawa, N.; Isshiki, H.; Zhang, Y. F.; Komeda, T.; Yamagishi, M.; Takeya, J. Direct Observation of Lanthanide(III)-Phthalocyanine Molecules on Au(111) by Using Scanning Tunneling Microscopy and

Scanning Tunneling Spectroscopy and Thin-Film Field-Effect Transistor Properties of Tb(III)- and Dy(III)-Phthalocyanine Molecules. J. Am. Chem. Soc. 2009, 131 (29), 9967−9976. (12) (a) Jiang, S. D.; Wang, B. W.; Su, G.; Wang, Z. M.; Gao, S. A mononuclear dysprosium complex featuring single-molecule-magnet behavior. Angew. Chem., Int. Ed. 2010, 49 (41), 7448−7451. (b) Liu, S. S.; Ziller, J. W.; Zhang, Y. Q.; Wang, B. W.; Evans, W. J.; Gao, S. A half-sandwich organometallic single-ion magnet with hexamethylbenzene coordinated to the Dy(III) ion. Chem. Commun. 2014, 50 (77), 11418−11420. (c) Cao, W.; Gao, C.; Zhang, Y.-Q.; Qi, D.; Liu, T.; Wang, K.; Duan, C.; Gao, S.; Jiang, J. Rational enhancement of the energy barrier of bis(tetrapyrrole) dysprosium SMMs via replacing atom of porphyrin core. Chem. Sci. 2015, 6 (10), 5947−5954. (13) (a) Benelli, C.; Gatteschi, D. Magnetism of Lanthanides in Molecular Materials with Transition-Metal Ions and Organic Radicals. Chem. Rev. 2002, 102, 2369−2388. (b) Tang, J.; Hewitt, I.; Madhu, N. T.; Chastanet, G.; Wernsdorfer, W.; Anson, C. E.; Benelli, C.; Sessoli, R.; Powell, A. K. Dysprosium triangles showing singlemolecule magnet behavior of thermally excited spin states. Angew. Chem., Int. Ed. 2006, 45 (11), 1729−1733. (14) (a) Cole, K. S.; Cole, R. H. Dispersion and absorption in dielectrics. I. Alternating-current characteristics. J. Chem. Phys. 1941, 9, 341−351. (b) Cole, K. S.; Cole, R. H. Dispersion and absorption in dielectrics. II. Direct-current characteristics. J. Chem. Phys. 1942, 10, 98−105. (c) Aubin, S. M. J.; Sun, Z.; Pardi, L.; Krzystek, J.; Folting, K.; Brunel, L.-C.; Rheingold, A. L.; Christou, G.; Hendrickson, D. N. Reduced Anionic Mn12 Molecules with Half-Integer Ground States as Single-Molecule Magnets. Inorg. Chem. 1999, 38 (23), 5329−5340. (15) (a) Sorace, L.; Benelli, C.; Gatteschi, D. Lanthanides in molecular magnetism: old tools in a new field. Chem. Soc. Rev. 2011, 40 (6), 3092−3104. (b) Feng, M.; Tong, M.-L. Single Ion Magnets from 3d to 5f: Developments and Strategies. Chem. - Eur. J. 2018, 24, 7574−7594. (c) Rinehart, J. D.; Long, J. R. Exploiting single-ion anisotropy in the design of f-element single-molecule magnets. Chem. Sci. 2011, 2 (11), 2078−2085. (16) Stites, J.; McCarty, C.; Quill, L. The Rare Earth Metals and their Compounds. VIII. An Improved Method for the Synthesis of Some Rare Earth Acetylacetonates. J. Am. Chem. Soc. 1948, 70, 3142− 3143. (17) Zhang, H.; Lai, Z.; Cui, X.; Jiang, J.; Machida, K.-I. Heteroleptic protonated bis(phthalocyaninato) rare earth compounds containing 1,4,8,11,15,18,22,25-octa(butyloxy)-phthalocyanine ligand. J. Alloys Compd. 2006, 408-412, 1041−1045. (18) (a) Collamati, I.; Cervone, E.; Scoccia, R. Synthesis and reactivity of some Fe(II) and Mn(II) complexes with hemiporphyrazine. Inorg. Chim. Acta 1985, 98 (1), 11−17. (b) Huber, S.; Seyfried, M.; Linden, A.; Luedtke, N. Excitonic luminescence of hemiporphyrazines. Inorg. Chem. 2012, 51 (13), 7032−7038.

G

DOI: 10.1021/acs.inorgchem.8b02068 Inorg. Chem. XXXX, XXX, XXX−XXX