Next Article in Journal / Special Issue
4′-(5-N-Propylthiophen-2-yl)-2,2′:6′,2″-terpyridine
Previous Article in Journal
N-(6-Chloro-3-nitropyridin-2-yl)-5-(1-methyl-1H-pyrazol-4-yl)isoquinolin-3-amine
Previous Article in Special Issue
Synthesis and Structural Characterization of (E)-4-[(2-Hydroxy-3-methoxybenzylidene)amino]butanoic Acid and Its Novel Cu(II) Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

Cis-Bis(2,2′-Azopyridinido)dicarbonylruthenium(II)

1
Department of Natural Sciences and Informatics, Fukushima University, 1 Kanayagawa, Fukushima 960-1296, Japan
2
Graduate School of Science and Engineering, Fukushima University, 1 Kanayagawa, Fukushima 960-1296, Japan
*
Authors to whom correspondence should be addressed.
Molbank 2021, 2021(1), M1182; https://0-doi-org.brum.beds.ac.uk/10.3390/M1182
Submission received: 28 December 2020 / Accepted: 15 January 2021 / Published: 18 January 2021

Abstract

:
An [Ru(apy)2Cl2] precursor (apy = 2,2′-azopyridine) in 2-methoxyethanol was heated under a pressurized CO atmosphere to afford a diradical complex, [Ru(apy·)2(CO)2], containing one-electron-reduced azo anion radical ligands. The electronic states of the complex were characterized by spectroscopic techniques and computational studies. Magnetic measurements revealed the existence of antiferromagnetic interactions in the diradical complex.

1. Introduction

The oxidation number of the central metal atom in a complex generally determines the electronic states of the complex. In a metal complex containing a redox-active non-innocent ligand, conversely, ligands as well as the metal centers can also control the electronic states of the complex. For example, azo compounds can directly accept one or two electrons because they have a low-lying azo-centered vacant π* molecular orbital (Scheme 1) [1]. Due to this property, many transition metal complexes, particularly those in groups 7 and 8, containing azopyridyl ligands, e.g., 2,2′-azopyridine (apy) and 2-phenylazopyridine (pap), (Figure 1a), have been reported [2,3,4,5,6,7]. Based on numerous studies, metal complexes containing non-innocent azo ligands are expected to be applied to multifunctional materials which surpass conventional magnetic materials, such as metal oxides [8,9].
We have previously reported the synthesis and properties of a monocarbonylruthenium(II) complex containing two azopyridine molecules [10]. All the azopyridyl ligands in the monocarbonyl complex were neutral (nonradical, singlet state: S = 0). In contrast, the newly synthesized dicarbonylruthenium(II) complex in this work (Figure 1b) experimentally and theoretically revealed that both the azopyridyl ligands are coordinated as anion radicals (triplet state: S = 1). Additionally, the magnetic properties of the diradical complex were examined.

2. Results and Discussion

2.1. Synthesis and Characterization of the Diradical Complex

We previously reported the detailed characterization, including X-ray crystallography, of the monocarbonyl complex ([Ru(pap)2(CO)Cl]+) [10]. In this report, we confirmed that the two azo ligands were neutral in the monocarbonyl complex. Meanwhile, two carbonyl stretching frequencies in the newly prepared dicarbonyl complex were observed at 2060 and 1995 cm−1 in the IR spectra (Figure S1). These values are 30–40 cm−1 lower than those of similar Ru complexes ([Ru(N-N)2(CO)2]2+; N-N = bidentate pyridyl ligands) [11,12,13]. The reason is that the reduction of the azo moiety increases the electron density of the complex, resulting in a red shift of the CO bands [5]. Additionally, no peaks assignable to the present complex were observed in the mass spectra, and its 1H-NMR spectrum was silent. It is, therefore, interpreted that the introduction of two CO molecules, which function as ligand and reducing agents, led to the selective formation of the diradical neutral complex, in which both the azo ligands underwent one-electron reduction (Equation (1)).
[RuII(apy)2Cl2] + 2CO → [RuII(apy·)2(CO)2] + 2Cl

2.2. Electronic Structure Analysis

We performed quantum chemical calculations for both the diradical and the corresponding nonradical complexes to evaluate their stability (Figure S2). The adiabatic energy gap between the two complexes in methanol were calculated (1530.0478 Hartree for the nonradical state and 1530.3652 Hartree for the diradical one) [14]. The results indicate that the diradical state is 30.75 kcal/mol more stable than the nonradical one. Figure 2 shows the spin densities in the optimized structures. Since the spin densities of both azo sites in the diradical state are 0.623, the azo site is the primary spin-bearing center.
Figure 3a shows the estimated absorption spectra of the nonradical (S = 0) and diradical complexes (S = 1) using time-dependent density functional theory (TDDFT) calculations. It is presumed that the nonradical complex exhibits absorption maximum around 400 nm, while the diradical one exhibits absorptions around 350 and 470 nm. Since the latter values are consistent with the spectral data of the prepared complex (λmax = 355 and 487 nm; Figure 3b), computational results also suggested that the present complex contains two anion radical ligands (apy·).

2.3. Magnetic Properties

To experimentally confirm that the present complex is a radical complex with unpaired spins, we carried out measurements of the dependence of the magnetic moment on the external magnetic field (M-H properties). Since M = χH (M: magnetic moment, H: external applied magnetic field, χ: magnetic molar susceptibility) is satisfied within the temperature range measured, the complex is the radical complex, which displays paramagnetic properties. Additionally, the Curie–Weiss law defined by Equation (2) (C: Curie constant, T: absolute temperature, θ: Weiss constant) is satisfied in the low-temperature region bearing on the temperature dependence of the magnetic susceptibility (the applied magnetic field of 10000 Oe) as shown in Figure 4; the Weiss constant is negative (θ = −10.7 K), suggesting that antiferromagnetic interactions exist between the spins in each radical [5].
χ = C/(Tθ)
In summary, we have found that the electronic state of an azo ligand can be controlled by the number of coordinated carbonyls, i.e., the monocarbonyl complex has neutral azo ligands, whereas the corresponding dicarbonyl one led to anion radicals. The latter complex exhibited paramagnetic and antiferromagnetic interactions between the spins in each radical. Further studies using analogous compounds are underway, directed toward the development of multifunctional materials in detail.

3. Materials and Methods

3.1. Measurements

The infrared (IR) spectra were obtained using a JASCO FT-IR 4100 spectrometer (JASCO Corporation, Hachioji, Japan). The elemental analysis data were obtained on a Perkin-Elmer 2400II series CHN analyzer (Perkin-Elmer, Inc., Waltham, MA, USA). The absorption spectra were obtained on a JASCO V-560 spectrophotometer (JASCO Corporation, Hachioji, Japan). The magnetic susceptibilities of the complex were recorded on a SQUID magnetometer (MPMS, Quantum Design, Inc., San Diego, CA, USA) in the temperature range of 5–300 K. DFT and TDDFT calculations were performed using the quantum chemical program, Gaussian 16 [15]. The geometries of the complex in its singlet (S = 0) or triplet (S = 1) states were fully optimized by restricted (singlet) or unrestricted (triplet) DFT methods employing the B3LYP function [16,17], with a 6-31G(d) basis set for the light elements [18,19] and a LanL2DZ basis set [20] for the Ru atom. The solvent effect of methanol was evaluated using an implicit solvent model, and a conductor-like polarizable continuum model (CPCM). Vibrational analyses were performed at the same calculation level employed for geometry optimization. The spin density analyses were also calculated at the optimized geometry. Excited-state calculations based on the B3LYP-optimized geometries were conducted for the TDDFT formalism in methanol using the CPCM.

3.2. Synthesis of the Complex

The compound, 2,2′-azopyridine (apy), and the starting material ([Ru(apy)2Cl2]) were prepared according to known procedures or through modification of published methods [21,22].
A solution of [Ru(apy)2Cl2] (38 mg, 0.071 mmol) in 2-methoxyethanol (40 mL) was heated to 100 °C under an atmosphere of CO (2 MPa) for 72 h. The volume was reduced to 5 mL using a rotary evaporator, and a black precipitate was formed by the addition of diethyl ether. The product was collected by filtration, washed with diethyl ether, and dried in vacuo. The crude product was recrystallized from dichloromethane and diethyl ether. Yield: 33 mg (88%). Anal. Calc. for C22H16N8O2Ru·2.5CH2Cl2: C, 39.88; H, 2.87; N, 15.19. Found: C, 39.82; H, 2.89; N, 15.28%. IR (KBr): ν = 2060 and 1995 cm−1 (C≡O). UV-vis (dimethyl sulfoxide): λmax/nm (ε/M−1 cm−1) 355 (6800) and 487 (2900).

Supplementary Materials

The following are available online, Figure S1: The IR spectrum of [Ru(apy·)2(CO)2] (KBr method), Figure S2: Optimized structures of [Ru(apy·)2(CO)2] and [Ru(apy)2(CO)2]2+ (in MeOH).

Author Contributions

Conceptualization, T.T. and D.O.; synthesis, S.K.; formal analysis, S.K. and T.K.; investigation, T.T.; writing—original draft preparation, T.T.; writing—review and editing, D.O.; supervision, D.O.; funding acquisition, D.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JSPS KAKENHI, grant numbers JP17K05799 and JP20K05536.

Data Availability Statement

The data presented in this study are available in this article and supporting supplementary material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kaim, W. Complexes with 2,2′-azobispyridine and related ‘S-frame’ bridging ligands containing the azo function. Coord. Chem. Rev. 2001, 219–221, 463–488. [Google Scholar] [CrossRef]
  2. Shivakumar, M.; Pramanik, K.; Bhattacharyya, I.; Chakravorty, A. Chemistry of metal-bound anion radicals. A family of mono- and bis(azopyridine) chelates of bivalent ruthenium. Inorg. Chem. 2000, 39, 4332–4338. [Google Scholar] [CrossRef] [PubMed]
  3. Das, D.; Mondal, T.K.; Mobin, S.M.; Lahiri, G.K. Sensitive valence structures of [(pap)2Ru(Q)]n (n = +2, +1, 0, –1, –2) with two different redox noninnocent ligands, Q = 3,5-di-tert-butyl-N-aryl-1,2-benzoquinonemonoimine and pap = 2-phenylazopyridine. Inorg. Chem. 2009, 48, 9800–9810. [Google Scholar] [CrossRef] [PubMed]
  4. Oyama, D.; Asuma, A.; Hamada, T.; Takase, T. Novel [Ru(polypyridine)(CO)2Cl2] and [Ru(polypyridine)2(CO)Cl]+-type complexes: Characterizing the effects of introducing azopyridyl ligands by electrochemical, spectroscopic and crystallographic measurements. Inorg. Chim. Acta 2009, 362, 2581–2588. [Google Scholar] [CrossRef]
  5. Paul, N.; Samanta, S.; Goswami, S. Redox induced electron transfer in doublet azo-anion diradical rhenium(II) complexes. Characterization of complete electron transfer series. Inorg. Chem. 2010, 49, 2649–2655. [Google Scholar] [CrossRef]
  6. Ghosh, P.; Samanta, S.; Roy, S.K.; Demeshko, S.; Meyer, F.; Goswami, S. Introducing a new azoaromatic pincer ligand. Isolation and characterization of redox events in its ferrous complexes. Inorg. Chem. 2014, 53, 4678–4686. [Google Scholar] [CrossRef]
  7. Oyama, D.; Mun, B.; Takase, T. Redox-induced reversible intramolecular carbon-nitrogen bond formation of an azopyridylruthenium complex: Control of carbonyl ligand photoreactivity caused by structural change of the complex. J. Organomet. Chem. 2015, 799–800, 173–178. [Google Scholar] [CrossRef]
  8. Zapata-Rivera, J.; Maynau, D.; Calzado, C.J. Evaluation of the magnetic interactions in salts containing [Ni(dmit)2] radical anions. Chem. Mater. 2017, 29, 4317–4329. [Google Scholar] [CrossRef]
  9. Farcaş, A.-A.; Beu, T.A.; Bende, A. Light-induced spin transitions in Ni(II)-based macrocyclic-ligand complexes: A DFT study. J. Photochem. Photobiol. 2019, 376, 316–323. [Google Scholar] [CrossRef]
  10. Oyama, D.; Takatsuki, Y.; Fujita, R. Azopyridylruthenium(II) complexes containing Ru-C bonds: Synthesis, characterization and reactivity. Trends Inorg. Chem. 2010, 12, 31–40. [Google Scholar]
  11. Kobayashi, K.; Tanaka, K. Reactivity of CO2 activated on transition metals and sulfur ligands. Inorg. Chem. 2015, 54, 5085–5095. [Google Scholar] [CrossRef] [PubMed]
  12. Oyama, D.; Abe, R.; Takase, T. CO-ligand photodissociation in two Ru(II) complexes affected by different polypyridyl supporting ligands. Chem. Lett. 2017, 46, 1412–1414. [Google Scholar] [CrossRef] [Green Version]
  13. Akatsuka, K.; Abe, R.; Takase, T.; Oyama, D. Coordination chemistry of Ru(II) complexes of an asymmetric bipyridine analogue: Synergistic effects of supporting ligand and coordination geometry on reactivities. Molecules 2020, 25, 27. [Google Scholar] [CrossRef] [Green Version]
  14. Luckhaus, D.; Yamamoto, Y.-I.; Suzuki, T.; Signorell, R. Genuine binding energy of the hydrated electron. Sci. Adv. 2017, 3, e1603224. [Google Scholar] [CrossRef] [Green Version]
  15. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 (Revision, A.03); Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  16. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  17. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. 1988, B37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  18. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self-consistent molecular orbital methods. XII. Further extensions of Gaussian-type basis sets for use in molecular orbital studies of organic molecules. J. Chem. Phys. 1972, 56, 2257–2261. [Google Scholar] [CrossRef]
  19. Francl, M.M.; Pietro, W.J.; Hehre, W.J.; Binkley, J.S.; Gordon, M.S.; DeFrees, D.J.; Pople, J.A. Self-consistent molecular orbital methods. XXIII. A polarization-type basis set for second-row elements. J. Chem. Phys. 1982, 77, 3654–3665. [Google Scholar] [CrossRef] [Green Version]
  20. Wadt, W.R.; Hay, P.J. Ab initio effective core potentials for molecular calculations. Potentials for main group elements Na to Bi. J. Chem. Phys. 1985, 82, 284–298. [Google Scholar] [CrossRef]
  21. Baldwin, D.A.; Lever, A.B.P.; Parish, R.V. Complexes of 2,2′-azopyridine with iron(II), cobalt(II), nickel(II), copper(I), and copper(II). Infrared study. Inorg. Chem. 1969, 8, 107–115. [Google Scholar] [CrossRef]
  22. Hotze, A.C.G.; Caspers, S.E.; DeVos, D.; Kooijman, H.; Spek, A.L.; Flamigni, A.; Bacac, M.; Sava, G.; Haasnoot, J.G.; Reedijk, J. Structure-dependent in vitro cytotoxicity of the isomeric complexes [Ru(L)2Cl2] (L=o-tolylazopyridine and 4-methyl-2-phenylazopyridine) in comparison to [Ru(azpy)2Cl2]. J. Biol. Inorg. Chem. 2004, 9, 354–364. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Chemical structures of (a) azopyridines; (b) The Ru-complex presented in this study.
Figure 1. Chemical structures of (a) azopyridines; (b) The Ru-complex presented in this study.
Molbank 2021 m1182 g001
Scheme 1. Two-step electron transfers in azo compounds.
Scheme 1. Two-step electron transfers in azo compounds.
Molbank 2021 m1182 sch001
Figure 2. Spin density analyses for [Ru(apy·)2(CO)2], calculated at the DFT/B3LYP/LanL2DZ/6-31G(d) level of theory.
Figure 2. Spin density analyses for [Ru(apy·)2(CO)2], calculated at the DFT/B3LYP/LanL2DZ/6-31G(d) level of theory.
Molbank 2021 m1182 g002
Figure 3. (a) The theoretical absorption spectra calculated for the singlet and triplet spin configurations of [Ru(apy)2(CO)2]2+/0 at the TDDFT/B3LYP/LanL2DZ/6-31G(d) level of theory; (b) The absorption spectrum of [Ru(apy·)2(CO)2].
Figure 3. (a) The theoretical absorption spectra calculated for the singlet and triplet spin configurations of [Ru(apy)2(CO)2]2+/0 at the TDDFT/B3LYP/LanL2DZ/6-31G(d) level of theory; (b) The absorption spectrum of [Ru(apy·)2(CO)2].
Molbank 2021 m1182 g003
Figure 4. Temperature dependence of the inverse molar magnetic susceptibility for [Ru(apy·)2(CO)2].
Figure 4. Temperature dependence of the inverse molar magnetic susceptibility for [Ru(apy·)2(CO)2].
Molbank 2021 m1182 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Takase, T.; Kainuma, S.; Kanno, T.; Oyama, D. Cis-Bis(2,2′-Azopyridinido)dicarbonylruthenium(II). Molbank 2021, 2021, M1182. https://0-doi-org.brum.beds.ac.uk/10.3390/M1182

AMA Style

Takase T, Kainuma S, Kanno T, Oyama D. Cis-Bis(2,2′-Azopyridinido)dicarbonylruthenium(II). Molbank. 2021; 2021(1):M1182. https://0-doi-org.brum.beds.ac.uk/10.3390/M1182

Chicago/Turabian Style

Takase, Tsugiko, Shuya Kainuma, Takatoshi Kanno, and Dai Oyama. 2021. "Cis-Bis(2,2′-Azopyridinido)dicarbonylruthenium(II)" Molbank 2021, no. 1: M1182. https://0-doi-org.brum.beds.ac.uk/10.3390/M1182

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

Article Metrics

Back to TopTop