To see the other types of publications on this topic, follow the link: Fe(bpy)3(PF6)2.

Journal articles on the topic 'Fe(bpy)3(PF6)2'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Fe(bpy)3(PF6)2.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Hoshikawa, Ryusei, Kosuke Yoshida, Ryoji Mitsuhashi, Masahiro Mikuriya, Takashi Okuno, and Hiroshi Sakiyama. "Structure Controlling Factors of Oxido-Bridged Dinuclear Iron(III) Complexes." Molecules 26, no. 4 (February 8, 2021): 897. http://dx.doi.org/10.3390/molecules26040897.

Full text
Abstract:
Oxido bridges commonly form between iron(III) ions, but their bond angles and symmetry vary with the circumstances. A large number of oxido-bridged dinuclear iron(III) complexes have been structurally characterized. Some of them belong to the C2 point group, possessing bent Fe–O–Fe bonds, while some others belong to the Ci symmetry, possessing the linear Fe–O–Fe bonds. The question in this study is what determines the structures and symmetry of oxido-bridged dinuclear iron(III) complexes. In order to gain further insights, three oxido-bridged dinuclear iron(III) complexes were newly prepared with 2,2′-bipyridine (bpy) and 1,10-phenanthroline (phen) ligands: [Fe2OCl2(bpy)4][PF6]2 (1), [Fe2O(NO3)2(bpy)4][PF6]2·0.6MeCN·0.2(2-PrOH) (2), and [Fe2OCl2(phen)4][PF6]2·MeCN·0.5H2O (3). The crystal structures of 1, 2, and 3 were determined by the single-crystal X-ray diffraction method, and all of them were found to have the bent Fe–O–Fe bonds. Judging from the crystal structure, some intramolecular interligand hydrogen bonds were found to play an important role in fixing the structures. Additional density functional theory (DFT) calculations were conducted, also for a related oxido-bridged dinuclear iron(III) complex with a linear Fe–O–Fe bond. We conclude that the Fe–O–Fe bridge tends to bend like a water molecule, but is often stretched by interligand steric repulsion, and that the structures are mainly controlled by the intramolecular interligand interactions.
APA, Harvard, Vancouver, ISO, and other styles
2

Tang, Li-Hua, Mao-Ling Cheng, Zhe Qian, Ai-Quan Jia, and Qian-Feng Zhang. "Heteroleptic ruthenium(II) 2,2′-bipyridine complexes incorporating substituted pyrazol-1-yl-pyridazine ancillaries." Zeitschrift für Naturforschung B 75, no. 4 (April 28, 2020): 383–91. http://dx.doi.org/10.1515/znb-2019-0233.

Full text
Abstract:
AbstractCondensation of 3,6-dichloropyridazine or 3,6-dichloro-4,5-dimethyl- pyridazine with 3-methyl-1H-pyrazole or 4-methyl-1H-pyrazole with the assistance of sodium metal in tetrahydrofuran at reflux afforded three 3,6-bis(pyrazolyl)- pyridazine-type ligands: 3,6-bis(3-methylpyrazolyl)pyridazine (L1), 3,6-bis(4-methyl- pyrazolyl)pyridazine (L2) and 4,5-dimethyl-3,6-bis(4-methylpyrazolyl)pyridazine (L3). Reactions of cis-[RuCl2(bpy)2] · 2H2O (bpy = 2,2′-bipyridine) and L1, L2 or L3 in the presence of NH4PF6 produced the heteroleptic cationic ruthenium(II) complexes [Ru(L1)(bpy)2](PF6)2 (1), [Ru(L2)(bpy)2](PF6)2 (2) and [Ru(L3)(bpy)2](PF6)2 (3), respectively. The three complexes have been characterized by UV/Vis and luminescence spectroscopy. The crystal structures of 1 · EtOH, 2 · EtOH and 3 have been determined by single-crystal X-ray diffraction.
APA, Harvard, Vancouver, ISO, and other styles
3

Luis, Ena T., Graham E. Ball, Alyssa Gilbert, Hasti Iranmanesh, Connor W. Newdick, and Jonathon E. Beves. "Efficient microwave-assisted synthesis and characterization of key ruthenium(II) polypyridyl complexes [Ru(bpy)3](PF6)2, [Ru(phen)3](PF6)2, [Ru(bpy)2(phen)](PF6)2 and [Ru(phen)2(bpy)](PF6)2." Journal of Coordination Chemistry 69, no. 11-13 (June 19, 2016): 1686–94. http://dx.doi.org/10.1080/00958972.2016.1194404.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Marchenko, Nataliia, Pascal G. Lacroix, Valerii Bukhanko, Marine Tassé, Carine Duhayon, Martial Boggio-Pasqua, and Isabelle Malfant. "Multistep Photochemical Reactions of Polypyridine-Based Ruthenium Nitrosyl Complexes in Dimethylsulfoxide." Molecules 25, no. 9 (May 8, 2020): 2205. http://dx.doi.org/10.3390/molecules25092205.

Full text
Abstract:
The photorelease of nitric oxide (NO·) has been investigated in dimethylsulfoxide (DMSO) on two compounds of formula [Ru(R-tpy)(bpy)(NO)](PF6)3, in which bpy stands for 2,2′-bipyridine and R-tpy for the 4′-R-2,2′:6′,2″-terpyridine with R = H and MeOPh. It is observed that both complexes are extremely sensitive to traces of water, leading to an equilibrium between [Ru(NO)] and [Ru(NO2)]. The photoproducts of formula [Ru(R-tpy)(bpy)(DMSO)](PF6)2 are further subjected to a photoreaction leading to a reversible linkage isomerization between the stable Ru-DMSO(S) (sulfur linked) and the metastable Ru-DMSO(O) (oxygen linked) species. A set of 4 [Ru(R-tpy)(bpy)(DMSO)]2+ complexes (R = H, MeOPh, BrPh, NO2Ph) is investigated to characterize the ratio and mechanism of the isomerization which is tentatively related to the difference in absorbance between the Ru-DMSO(S) and Ru-DMSO(O) forms. In addition, the X-ray crystal structures of [Ru(tpy)(bpy)(NO)](PF6)3 and [Ru(MeOPh-tpy)(bpy)(DMSO(S))](PF6)2 are presented.
APA, Harvard, Vancouver, ISO, and other styles
5

Tong, Bihai, Jiayan Qiang, Qunbo Mei, Hengshan Wang, Qianfeng Zhang, and Zhao Han. "Tuning the Photophysical Properties of Cyclometalated Ir(III) Complexes by a Trifluoroacetyl Group." Zeitschrift für Naturforschung B 67, no. 3 (March 1, 2012): 213–18. http://dx.doi.org/10.1515/znb-2012-0306.

Full text
Abstract:
Four cationic Ir(III) complexes, [Ir(dpq)2(bpy)]PF6 (1), [Ir(dpq)2(phen)]PF6 (2), [Ir(tfapq)2- (bpy)]PF6 (3), and [Ir(tfapq)2(phen)]PF6 (4) (dpqH = 2,4-diphenylquinoline, tfapqH = 2-(4ʹ-trifluoroacetylphenyl)- 4-phenylquinoline, bpy = 2,2ʹ-bipyridine, phen = 1,10-phenanthroline) have been synthesized and fully characterized. The structure of 4 was also confirmed by single-crystal X-ray diffraction. The electron-acceptor character of the trifluoroacetyl unit leads to a reduced HOMO-LUMO gap and consequently a red-shift of the UV/Vis absorption and luminescence spectra. The solvophobic character of the trifluoroacetyl unit gives rise to a molecule assembly in solution.
APA, Harvard, Vancouver, ISO, and other styles
6

Yersin, Hartmut, and Cornelius Kratzer. "Energy transfer and harvesting in [Ru1−xOsx(bpy)3](PF6)2 and {Λ-[Ru(bpy)3]Δ-[Os(bpy)3]}(PF6)4." Coordination Chemistry Reviews 229, no. 1-2 (July 2002): 75–93. http://dx.doi.org/10.1016/s0010-8545(02)00106-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Biner, M., H. B. Buergi, A. Ludi, and C. Roehr. "Crystal and molecular structures of [Ru(bpy)3](PF6)3 and [Ru(bpy)3](PF6)2 at 105 K." Journal of the American Chemical Society 114, no. 13 (June 1992): 5197–203. http://dx.doi.org/10.1021/ja00039a034.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Luo, Hong, Zhi-Ping Wang, An-Guo Zhang, and Ke-Zhi Wang. "Synthesis, and Acid–Base and DNA-Binding Properties of a Thiophen-Appended Ruthenium Complex." Australian Journal of Chemistry 64, no. 2 (2011): 206. http://dx.doi.org/10.1071/ch10316.

Full text
Abstract:
2-(5-Phenylthiophen-2-yl)-1H-imidazo[4,5-f][1,10]phenanthroline (Hptip) and its RuII complex [Ru(bpy)2(Hptip)](PF6)2 (where bpy = 2,2′-bipyridine) have been synthesized and characterized by elemental analysis, 1H NMR spectroscopy, and mass spectrometry. The acid–base properties of the complex were studied by UV-visible and luminescence spectrophotometric pH titrations, and ground- and excited-state acidity ionization constants were derived. The DNA-binding properties of [Ru(bpy)2(Hptip)](PF6)2 were also investigated by means of UV-vis and emission spectroscopy, salt effects, steady-state emission quenching by [Fe(CN)6]4–, DNA competitive binding with ethidium bromide, DNA melting experiments, and viscosity measurements. Density functional theoretical calculations were also carried out in order to understand the DNA binding properties.
APA, Harvard, Vancouver, ISO, and other styles
9

Mazzeo, Francesca, Fabian Brunner, Alessandro Prescimone, Edwin C. Constable, and Catherine E. Housecroft. "Intra-Cation versus Inter-Cation π-Contacts in [Cu(P^P)(N^N)][PF6] Complexes." Crystals 10, no. 1 (December 18, 2019): 1. http://dx.doi.org/10.3390/cryst10010001.

Full text
Abstract:
A series of [Cu(POP)(N^N][PF6] and [Cu(xantphos)(N^N][PF6] compounds has been prepared and characterized in which POP = bis[2-(diphenylphosphanyl)phenyl]ether (IUPAC PIN oxydi(2,1-phenylene)bis(diphenylphosphane), xantphos = 4,5-bis(diphenylphosphanyl)-9,9-dimethyl-9H-xanthene (IUPAC PIN (9,9-dimethyl-9H-xanthene- 4,5-diyl)bis(diphenylphosphane)) and the N^N ligands are 4-(4-bromophenyl)-6,6′-dimethyl-2,2′- bipyridine (1), 5,5′-bis(3-methoxyphenyl)-6-methyl-2,2′-bipyridine (2), and 6-benzyl-2,2′-bipyridine (3). The single crystal structures of [Cu(xantphos)(1)][PF6]·CH2Cl2, [Cu(xantphos)(2)][PF6]·CH2Cl2 and [Cu(POP)(3)][PF6]·0.5H2O were determined by X-ray diffraction. Each complex contains a copper(I) ion in a distorted tetrahedral environment with chelating N^N and P^P ligands. In the [Cu(xantphos)(1)]+ and [Cu(xantphos)(2)]+ cations, there are face-to-face π-stackings of bpy and PPh2 phenyl rings (i.e., between the ligands); in addition in [Cu(xantphos)(2)][PF6]·CH2Cl2, inter-cation π-embraces lead to the formation of infinite chains as a primary packing motif. In [Cu(POP)(3)][PF6]·0.5H2O, centrosymmetric pairs of [Cu(POP)(3)]+ cations engage in C–H…π (phenyl to bpy) and offset face-to-face (bpy…bpy) contacts. The electrochemical and photophysical properties of the compounds containing ligands 1 and 2 are reported. They are green or yellow emitters in the solid-state (λem in the range 535–577 nm) with values for the photoluminescence quantum yield (PLQY) in the range 19%–41%.
APA, Harvard, Vancouver, ISO, and other styles
10

Pérez-Cordero, Eduardo, Rosanna Buigas, Nancy Brady, Luis Echegoyen, Claudia Arana, and Jean-Marie Lehn. "[M(bpy)3] (M = Fe, Ru, Os): New Crystalline Materials from the reductive electrocrystallization of [M(bpy)3](PF6)2." Helvetica Chimica Acta 77, no. 5 (August 10, 1994): 1222–28. http://dx.doi.org/10.1002/hlca.19940770504.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Busemann, Anja, Ingrid Flaspohler, Xue-Quan Zhou, Claudia Schmidt, Sina K. Goetzfried, Vincent H. S. van Rixel, Ingo Ott, Maxime A. Siegler, and Sylvestre Bonnet. "Ruthenium-based PACT agents based on bisquinoline chelates: synthesis, photochemistry, and cytotoxicity." JBIC Journal of Biological Inorganic Chemistry 26, no. 6 (August 10, 2021): 667–74. http://dx.doi.org/10.1007/s00775-021-01882-8.

Full text
Abstract:
AbstractThe known ruthenium complex [Ru(tpy)(bpy)(Hmte)](PF6)2 ([1](PF6)2, where tpy = 2,2’:6’,2″-terpyridine, bpy = 2,2’-bipyridine, Hmte = 2-(methylthio)ethanol) is photosubstitutionally active but non-toxic to cancer cells even upon light irradiation. In this work, the two analogs complexes [Ru(tpy)(NN)(Hmte)](PF6)2, where NN = 3,3'-biisoquinoline (i-biq, [2](PF6)2) and di(isoquinolin-3-yl)amine (i-Hdiqa, [3](PF6)2), were synthesized and their photochemistry and phototoxicity evaluated to assess their suitability as photoactivated chemotherapy (PACT) agents. The increase of the aromatic surface of [2](PF6)2 and [3](PF6)2, compared to [1](PF6)2, leads to higher lipophilicity and higher cellular uptake for the former complexes. Such improved uptake is directly correlated to the cytotoxicity of these compounds in the dark: while [2](PF6)2 and [3](PF6)2 showed low EC50 values in human cancer cells, [1](PF6)2 is not cytotoxic due to poor cellular uptake. While stable in the dark, all complexes substituted the protecting thioether ligand upon light irradiation (520 nm), with the highest photosubstitution quantum yield found for [3](PF6)2 (Φ[3] = 0.070). Compounds [2](PF6)2 and [3](PF6)2 were found both more cytotoxic after light activation than in the dark, with a photo index of 4. Considering the very low singlet oxygen quantum yields of these compounds, and the lack of cytotoxicity of the photoreleased Hmte thioether ligand, it can be concluded that the toxicity observed after light activation is due to the photoreleased aqua complexes [Ru(tpy)(NN)(OH2)]2+, and thus that [2](PF6)2 and [3](PF6)2 are promising PACT candidates. Graphic abstract
APA, Harvard, Vancouver, ISO, and other styles
12

Breu, Josef, and Andrea Zwicknagel. "Chirale Erkennung bei Tris(diimin)-Metallkomplexen, 10. Vergleich der intermolekularen Wechselwirkungs- und Packungsmuster in der Reihe [Cr(bpy)3]n+(PF6)n (n = 0 – 3) / Chiral Recognition among Tris(diimine)-metal Complexes, 10. Comparison of Intermolecular Interactions and Packing Patterns in the Series [Cr(bpy)3]n+(PF6)n (n = 0–3)." Zeitschrift für Naturforschung B 59, no. 9 (September 1, 2004): 1015–25. http://dx.doi.org/10.1515/znb-2004-0911.

Full text
Abstract:
Due to their conformational rigidity, the corrugated, chiral molecular structure, and the variability in the central metal and its oxidation state, [M(bpy)3]n+ complexes are particularly well suited to study chiral recognition and to identify intermolecular interaction patterns in the crystalline state. For [Cr(bpy)3]n+(PF6)n (n=0 - 3) four oxidation states are readily accessible which allows to investigate the influence of the cation/anion ratio on the observed packing patterns.The crystal structures of all four oxidation states are governed by so-called ‘π-π-interactions’. Apparently, in molecular salts the Madelung energy is less important as compared to classical inorganic salts.Interestingly, [Cr(bpy)3](PF6) and [Cr(bpy)3](PF6)2 comprise the same homochiral layers. However, while the former crystallises as true racemate, the latter spontaneously resolves into a conglomerate. This two-dimensional building block of homochiral layers is the most popular structural motif in this class of compounds which has been observed in a great variety of racemic and homochiral stackings.
APA, Harvard, Vancouver, ISO, and other styles
13

Cuello-Garibo, Jordi-Amat, Catriona C. James, Maxime A. Siegler, and Sylvestre Bonnet. "Ruthenium-based PACT compounds based on an N,S non-toxic ligand: a delicate balance between photoactivation and thermal stability." Chemistry Squared 1 (December 1, 2017): 2. http://dx.doi.org/10.28954/2017.csq.12.002.

Full text
Abstract:
In photoactivated chemotherapy, the photocleavable protecting group that prevents the bioactive compound from interacting with biomolecules in the dark is sometimes cytotoxic, which makes interpretation of phototoxicity challenging. For ruthenium polypyridyl complexes new, non-toxic protecting ligands that prevent a toxic metal complex from binding to biomolecules in the dark, but that can be efficiently photosubstituted upon visible light irradiation to recover the high toxicity of the metal complex, are necessary. In this work, we report on the synthesis, stereochemical characterization and cytotoxicity of a series of polypyridyl complexes; [Ru(bpy)2(mtpa)](PF6)2 ([1](PF6)2, bpy = 2,2’-bipyridine), [Ru(bpy)(dmbpy)(mtpa)](PF6)2 ([2](PF6)2, dmbpy = 6,6’-dimethyl-2,2’-bipyridine), and [Ru(dmbpy)2(mtpa)](PF6)2 ([3](PF6)2) based on the non-toxic 3-(methylthio)propylamine protecting ligand (mtpa). The number of methyl groups had a crucial effect on the photochemistry and cytotoxicity of these complexes. The non-strained complex [1]2+ was not capable of fully releasing mtpa and was not phototoxic in lung cancer cells (A549). In the most strained complex [3]2+, thermal stability was lost, leading to poor photoactivation in vitro and a generally high toxicity also without light activation. The heteroleptic complex [2]2+ with intermediate strain showed, upon blue light irradiation, efficient mtpa photosubstitution and increased cytotoxicity in cancer cells, but photosubstitution was not selective. Overall, fine-tuning of the lipophilicity and steric strain of ruthenium complexes appears as an efficient method to obtain phototoxic ruthenium-based photoactivated chemotherapeutic prodrugs, at the cost of synthetic simplicity and photosubstitution selectivity.
APA, Harvard, Vancouver, ISO, and other styles
14

Oshio, Hiroki, Hironori Onodera, Osamu Tamada, Hideto Mizutani, Takashi Hikichi, and Tasuku Ito. "Cyanide-Bridged Fe−Fe and Fe−Co Molecular Squares: Structures and Electrochemistry of [Fe(μ-CN)4(bpy)8](PF6)4⋅4 H2O, [FeCo(μ-CN)4(bpy)8](PF6)4⋅3 CHCl3⋅2 CH3CN, and [FeCo(μ-CN)4(bpy)8](PF6)6⋅2 CHCl3⋅4 CH3NO2." Chemistry - A European Journal 6, no. 14 (July 17, 2000): 2523–30. http://dx.doi.org/10.1002/1521-3765(20000717)6:14<2523::aid-chem2523>3.0.co;2-t.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Pan, Zhong Wei, Zi Cheng Cai, Meng Jin Bian, Ting Lian, Zeng Yi Su, Zhen Zhen Zhang, and Du Qing Huang. "Temperature Dependent Solid-Liquid Extraction Behavior of Rare Earths Using N-Butyl Pyridinium Hexafluorophosphate with Benzoyl Acetone." Advanced Materials Research 734-737 (August 2013): 906–10. http://dx.doi.org/10.4028/www.scientific.net/amr.734-737.906.

Full text
Abstract:
Temperature dependent solid-liquid extraction (SLE) behavior of six rare earth elements (REs) was investigated using N-butyl pyridinium hexafluorophosphate ([BPy]PF6) as an ionic liquid solvent and benzoyl acetone (HA) as an extractant at 80 °C. Parameters including the amount of [BPy]PF6, HA concentration, extraction time, extraction temperature and pH were investigated and optimized. The extracted species was neutral REA(n-1)Cl (n-1 = 3 or 2) in the REn+-[BPy]PF6-HA SLE system. The extraction percentage of REn+could be 100% at stated pH and HA concentration. The recovery of REn+extracted into [BPy]PF6can be achieved using the mixture of 0.1 mol·L-1ClCH2CO2H and 0.2 mol·L-1HCl as stripping agents. REs were extracted into solid ionic liquid phase with smaller volume and concentrated to some extent. Furthermore the temperature dependent SLE allows to recover [BPy]PF6after the extraction procedures. These results indicate that the proposed procedure can be used for the preconcentration and separation of REs using [BPy]PF6with high melting point.
APA, Harvard, Vancouver, ISO, and other styles
16

Shahroosvand, Hashem, Parisa Abbasi, Mohsen Ameri, and Mohammad Reza Riahi Dehkordi. "Dye-Sensitized Nanocrystalline ZnO Solar Cells Based on Ruthenium(II) Phendione Complexes." International Journal of Photoenergy 2011 (2011): 1–10. http://dx.doi.org/10.1155/2011/634147.

Full text
Abstract:
The metal complexes ( (phen)2(phendione))(PF6)2(1), [ (phen)(bpy)(phendione))(PF6)2(2), and ( (bpy)2(phendione))(PF6)2(3) (phen = 1,10-phenanthroline, bpy = 2,2′-bipyridine and phendione = 1,10-phenanthroline-5,6-dione) have been synthesized as photo sensitizers for ZnO semiconductor in solar cells. FT-IR and absorption spectra showed the favorable interfacial binding between the dye-molecules and ZnO surface. The surface analysis and size of adsorbed dye on nanostructure ZnO were further examined with AFM and SEM. The AFM images clearly show both, the outgrowth of the complexes which are adsorbed on ZnO thin film and the depression of ZnO thin film. We have studied photovoltaic properties of dye-sensitized nanocrystalline semiconductor solar cells based on Ru phendione complexes, which gave power conversion efficiency of (η) of 1.54% under the standard AM 1.5 irradiation (100 mW cm−2) with a short-circuit photocurrent density () of 3.42 mA cm−2, an open-circuit photovoltage () of 0.622 V, and a fill factor (ff) of 0.72. Monochromatic incident photon to current conversion efficiency was 38% at 485 nm.
APA, Harvard, Vancouver, ISO, and other styles
17

Liao, Guojian, Zhengyuan Ye, Yunlu Liu, Bin Fu, and Chen Fu. "Octahedral ruthenium (II) polypyridyl complexes as antimicrobial agents against mycobacterium." PeerJ 5 (April 27, 2017): e3252. http://dx.doi.org/10.7717/peerj.3252.

Full text
Abstract:
Tuberculosis is one of the world’s deadliest infectious disease with 1.5 millions deaths annually. It is imperative to discover novel compounds with potent activity against M. tuberculosis. In this study, susceptibilities of M. smegmatis to the octahedral ruthenium(II) polypyridyl complexes, 1 {[(bpy)3Ru] (PF6)2 (bpy = 2,2′-bipyridine)}, 2 {[(phen)2Ru(dppz)](PF6)2 (phen = 1,10-phenanthroline, dppz = dipyridophenazine)} and 3 {[(phen)3Ru](PF6)2} were measured by broth microdilution and reported as the MIC values. Toxicities of complex 3 to LO2 and hepG2 cell lines were also measured. Complex 2 inhibited the growth of M. smegmatis with MIC value of 2 µg/mL, while complex 3 was bactericidal with MIC value of 26 µg/mL. Furthermore, the bactericidal activity of complex 3 was dependent on reactive oxygen species production. Complex 3 showed no cytotoxicity against LO2 and hepG2 cell lines at concentration as high as 64 µg/mL, paving the way for further optimization and development as a novel antibacterial agent for the treatment of M. tuberculosis infection.
APA, Harvard, Vancouver, ISO, and other styles
18

Prescimone, Alessandro, Javier Sanchez-Benitez, Konstantin V. Kamenev, John E. Warren, Alistair R. Lennie, Mark Murrie, Simon Parsons, and Euan K. Brechin. "High-Pressure Study of Oxo-bridged Mixed-Valent MnIII/MnIV Dimers High-Pressure Study of Oxo-bridged Mixed-Valent MnIII/MnIV Dimers." Zeitschrift für Naturforschung B 65, no. 3 (March 1, 2010): 221–30. http://dx.doi.org/10.1515/znb-2010-0302.

Full text
Abstract:
A combination of high-pressure single crystal X-ray diffraction and high-pressure SQUID magnetometry has been used to study two oxo-bridged mixed-valent MnIII/MnIV dimers. [Mn2O2(bpy)4](ClO4)3·3CH3CN, (1·3CH3CN; bpy = 2,2ʹ-bipyridine) has been compressed to 2.0 GPa whilst [Mn2O2(bpy)4](PF6)3·2CH3CN·1H2O, (2·2CH3CN·1H2O) could be measured crystallographically up to 4.55 GPa. The PF6 salt of [Mn2O2(bpy)4]3+ has never been reported before while 1 has been reported as a hydrate and in a different crystallographic space group. The application of hydrostatic pressure imposes significant distortions and modifications in the structures of both complexes. In particular, in complex 1·3CH3CN the Mn-Mn separation is reduced by the contraction of some of the Mn-O bond lengths, whilst in 2·2CH3CN·1H2O the Mn-O-Mn bridging angles and the Mn-O bond lengths are substantially unchanged. Interestingly 2·2CH3CN·1H2O also shows a constant contraction in nearly all the Mn-N bonds. The magnetic behaviour of the complexes has been measured up to 0.87 GPa for 1·3CH3CN and 0.84 GPa for 2·2CH3CN·1H2O.
APA, Harvard, Vancouver, ISO, and other styles
19

Ruhlmann, L., C. Costa-Coquelard, J. Hao, S. Jiang, C. He, L. Sun, and I. Lampre. "Association of ruthenium complexes [Ru(bpy)3]2+ or [Ru(bpy)2(Mebpy-py)]2+ with Dawson polyanions α-[P2W18O62]6– or α2-[FeIII(H2O)P2W17O61]7–." Canadian Journal of Chemistry 86, no. 11 (November 1, 2008): 1034–43. http://dx.doi.org/10.1139/v08-139.

Full text
Abstract:
The formation of electrostatically linked complexes between [Ru(bpy)3]Cl2 or [Ru(bpy)2(Mebpy-py)](PF6)2 and (TBA)6α-[P2W18O62] or (TBA)5K2α2-[FeIII(H2O)P2W17O61] in solution were studied by steady-state absorption and luminescence spectroscopy. The stoichiometry and global association constants of the formed complexes were obtained by the Job method. The formed complexes were characterized by 1H and 31P NMR spectroscopy and their electrochemical properties compared to the precursor compounds were investigated by using the cyclic voltammetry method.Key words: electrochemistry, polyoxometalate, ruthenium complex, electrostatic complex.
APA, Harvard, Vancouver, ISO, and other styles
20

Toma, S. H., M. Nakamura, and H. E. Toma. "The Effect of -Cyclodextrin Inclusion on the Morphology of [Ru(bpy)2Cl(BPEB)](PF6) Films by Scanning Force Microscopy." Microscopy and Microanalysis 11, S03 (December 2005): 142–45. http://dx.doi.org/10.1017/s1431927605051093.

Full text
Abstract:
Molecular level organization has been a subject of great relevance in supramolecular chemistry and nanotechnology. Supramolecular chemists count on the ability of molecules to form several kinds of organization, allowing the development of nanoscaled devices. In this way, the scanning probe microscopy provides a great tool for characterization, manipulation and interfacing such devices [1]. Regarding the ruthenium complexes [Ru(bpy)2Cl(BPEB)](PF6) and {[Ru(bpy)2Cl]2(BPEB)}(PF6)2, where bpy = 2,2'-bipyridine, the presence of the BPEB (1,4-bis[4-pyridyl)ethenyl]benzene) ligand has an important role as a recognition site for van der Waals interactions (Figure 1). On the other hand, cyclodextrins are macrocyclic molecules bearing a hydrophobic cavity that can support several types of guest molecules [2-3]. In this work we are showing the influence of the recognition site of the BPEB ligand and the formation of an inclusion compound in the patterning structures of films deposited over mica substrates, by SFM microscopy.
APA, Harvard, Vancouver, ISO, and other styles
21

Habibagahi, Fatemeh, and Robert J. Crutchley. "Mixed-valence properties of a dinuclear ruthenium complex bridged by bis(phenylcyanamido)tetrazine." Canadian Journal of Chemistry 92, no. 11 (November 2014): 1081–85. http://dx.doi.org/10.1139/cjc-2014-0307.

Full text
Abstract:
The novel bridging ligand 3,6-bis(phenylcyanamido)-1,2,4,5-tetrazine (tdpcH2) and its dinuclear complex [{Ru(ttpy)(bpy)}2(μ-tdpc)][PF6]2 were prepared and characterized by elemental analysis and 1H NMR spectroscopy. Cyclic voltammetry and vis-NIR and IR spectroelectrochemistry of [{Ru(ttpy)(bpy)}2(μ-tdpc)]2+ showed that [{Ru(ttpy)(bpy)}2(μ-tdpc)]3+ is a Class II mixed-valence system with metal−metal coupling of 400 cm−1 assuming a transition dipole moment length of 21 Å. DFT calculations of tdpc2− suggested that the stability of the HOMO results in weak metal−metal coupling via the hole-superexchange mechanism.
APA, Harvard, Vancouver, ISO, and other styles
22

Huttner, Dominikus, and Josef Breu. "Simulation of Nucleation for the Molecular Salt [Ru(bpy)3](PF6)2." Zeitschrift für anorganische und allgemeine Chemie 630, no. 11 (September 2004): 1730. http://dx.doi.org/10.1002/zaac.200470076.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Schenker, Sabine, Andreas Hauser, Wei Wang, and I. Y. Chan. "Matrix effects on the high-spin→low-spin relaxation in [M1−Fe (bpy)3](PF6)2 (M=Cd, Mn and Zn, bpy=2,2′-bipyridine)." Chemical Physics Letters 297, no. 3-4 (November 1998): 281–86. http://dx.doi.org/10.1016/s0009-2614(98)01136-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Potočňák, Ivan, Lucia Váhovská, and Peter Herich. "Low-dimensional compounds containing cyanide groups. XXV. Synthesis, spectroscopic properties and crystal structures of two ionic iron(II) complexes with tricyanomethanide: tris(1,10-phenanthroline-κ2 N,N′)iron(II) bis(tricyanomethanide) and tris(2,2′-bipyridine-κ2 N,N′)iron(II) bis(tricyanomethanide) sesquihydrate." Acta Crystallographica Section C Structural Chemistry 70, no. 5 (April 8, 2014): 432–36. http://dx.doi.org/10.1107/s2053229614006512.

Full text
Abstract:
Two new diamagnetic coordination compounds, [Fe(phen)3][C(CN)3]2, (I), and [Fe(bpy)3][C(CN)3]2·1.5H2O, (II), have been synthesized and characterized by single-crystal X-ray diffraction analysis, and IR and UV–Vis spectroscopy (phen is 1,10-phenanthroline, C12H8N2, and bpy is 2,2′-bipyridine, C10H8N2). Both compounds are ionic with distorted octahedral [Fe(phen)3]2+ or [Fe(bpy)3]2+ complex cations, with average Fe—N distances of 1.977 (2) and 1.971 (3) Å, respectively, and two uncoordinated planar tricyanomethanide, or [C(CN)3]−, counter-anions balancing the positive charges of the cations. Solvent water molecules and tcm anions in (II) are linked via O—H...N hydrogen bonds into negatively charged layers and the interlayer space is filled by [Fe(bpy)3]2+ cations. The structures of (I) and (II) are stabilized by π–π interactions between the stacked aromatic rings of the phen ligands of two adjacent cations and by O—H...N hydrogen bonds, respectively, and also by π–π stacking interactions between phen and tcm units in (I).
APA, Harvard, Vancouver, ISO, and other styles
25

Henke, Wade C., Julie A. Hopkins, Micah L. Anderson, Jonah P. Stiel, Victor W. Day, and James D. Blakemore. "4,5-Diazafluorene and 9,9’-Dimethyl-4,5-Diazafluorene as Ligands Supporting Redox-Active Mn and Ru Complexes." Molecules 25, no. 14 (July 13, 2020): 3189. http://dx.doi.org/10.3390/molecules25143189.

Full text
Abstract:
4,5-diazafluorene (daf) and 9,9’-dimethyl-4,5-diazafluorene (Me2daf) are structurally similar to the important ligand 2,2’-bipyridine (bpy), but significantly less is known about the redox and spectroscopic properties of metal complexes containing Me2daf as a ligand than those containing bpy. New complexes Mn(CO)3Br(daf) (2), Mn(CO)3Br(Me2daf) (3), and [Ru(Me2daf)3](PF6)2 (5) have been prepared and fully characterized to understand the influence of the Me2daf framework on their chemical and electrochemical properties. Structural data for 2, 3, and 5 from single-crystal X-ray diffraction analysis reveal a distinctive widening of the daf and Me2daf chelate angles in comparison to the analogous Mn(CO)3(bpy)Br (1) and [Ru(bpy)3]2+ (4) complexes. Electronic absorption data for these complexes confirm the electronic similarity of daf, Me2daf, and bpy, as spectra are dominated in each case by metal-to-ligand charge transfer bands in the visible region. However, the electrochemical properties of 2, 3, and 5 reveal that the redox-active Me2daf framework in 3 and 5 undergoes reduction at a slightly more negative potential than that of bpy in 1 and 4. Taken together, the results indicate that Me2daf could be useful for preparation of a variety of new redox-active compounds, as it retains the useful redox-active nature of bpy but lacks the acidic, benzylic C–H bonds that can induce secondary reactivity in complexes bearing daf.
APA, Harvard, Vancouver, ISO, and other styles
26

Yersin, H., G. Hensler, and E. Gallhuber. "On the zero-phonon structure of single-crystal [Ru(bpy)3](PF6)2." Inorganica Chimica Acta 132, no. 2 (September 1987): 187–91. http://dx.doi.org/10.1016/s0020-1693(00)81740-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Yersin, H., E. Gallhuber, and G. Hensler. "Highly resolved polarized absorption spectra of single-crystal [Ru(bpy)3](PF6)2." Chemical Physics Letters 134, no. 5 (March 1987): 497–501. http://dx.doi.org/10.1016/0009-2614(87)87180-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Yersin, Hartmut, Gerold Hensler, and Erich Gallhuber. "Energy transfer and highly resolved emission of [Ru1-xOsx(bpy)3] (PF6)2." Journal of Luminescence 40-41 (February 1988): 676–77. http://dx.doi.org/10.1016/0022-2313(88)90385-7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

KRAUSZ, Elmars, and Grainne MORAN. "MAGNETIC FIELD EFFECTS IN THE LUMINESCENCE OF THE Ru(bpY)3(PF6)2 DOPED WITH Os(bpy)32+." Journal of the Magnetics Society of Japan 11, S_1_ISMO (1987): S1_23–26. http://dx.doi.org/10.3379/jmsjmag.11.s1_23.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Krausz, Elmars, and Grainne Moran. "Low temperature non-thermalised luminescence of Ru(bpy)32+ doped in Cd(bpy)3(PF6)2 single crystals." Journal of Luminescence 40-41 (February 1988): 272–73. http://dx.doi.org/10.1016/0022-2313(88)90190-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Schönherr, T., J. Degen, E. Gallhuber, G. Hensler, and H. Yersin. "Geometrical distortions in excited A′2 states of single-crystal [Ru(bpy)3](PF6)2." Chemical Physics Letters 158, no. 6 (June 1989): 519–24. http://dx.doi.org/10.1016/0009-2614(89)87382-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Arata, Shinobu, Takefumi Hamamatsu, Tetsuya Sato, Tomotaka Iihoshi, Naohide Matsumoto, and Seiichiro Iijima. "Spin Crossover Iron(II) Complexes of (2-Methylimidazol-4-yl)methylideneamino-2-ethylpyridine: 2:1 [Fe(HLMe)2](PF6)2and 3:1 [Fe(HLMe)3](PF6)2." Chemistry Letters 36, no. 6 (June 5, 2007): 778–79. http://dx.doi.org/10.1246/cl.2007.778.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Meyer, Marco, Fabian Brunner, Alessandro Prescimone, Edwin C. Constable, and Catherine E. Housecroft. "Desymmetrizing Heteroleptic [Cu(P^P)(N^N)][PF6] Compounds: Effects on Structural and Photophysical Properties, and Solution Dynamic Behavior." Molecules 26, no. 1 (December 29, 2020): 125. http://dx.doi.org/10.3390/molecules26010125.

Full text
Abstract:
The preparation, characterization and electrochemical and photophysical properties of a series of desymmetrized heteroleptic [Cu(P^P)(N^N)][PF6] compounds are reported. The complexes incorporate the chelating P^P ligands bis(2-(diphenylphosphanyl)phenyl)ether (POP) and (9,9-dimethyl-9H-xanthene-4,5-diyl)bis(diphenylphosphane) (xantphos), and 6-substituted 2,2′-bipyridine (bpy) derivatives with functional groups attached by –(CH2)n– spacers: 6-(2,2′-bipyridin-6-yl)hexanoic acid (1), 6-(5-phenylpentyl)-2,2′-bipyridine (2) and 6-[2-(4-phenyl-1H-1,2,3,triazol-1-yl)ethyl]-2,2′-bipyridine (3). [Cu(POP)(1)][PF6], [Cu(xantphos)(1)][PF6], [Cu(POP)(2)][PF6], [Cu(xantphos)(2)][PF6], and [Cu(xantphos)(3)][PF6] have been characterized in solution using multinuclear NMR spectroscopy, and the single crystal structure of [Cu(xantphos)(3)][PF6].0.5Et2O was determined. The conformation of the 6-[2-(4-phenyl-1H-1,2,3,triazol-1-yl)ethyl]-substituent in the [Cu(xantphos)(3)]+ cation is such that the α- and β-CH2 units reside in the xanthene ‘bowl’ of the xantphos ligand. The 6-substituent desymmetrizes the structure of the [Cu(P^P)(N^N)]+ cation and this has consequences for the interpretation of the solution NMR spectra of the five complexes. The NOESY spectra and EXSY cross-peaks provide insight into the dynamic processes operating in the different compounds. For powdered samples, emission maxima are in the range 542–555 nm and photoluminescence quantum yields (PLQYs) lie in the range 13–28%, and a comparison of PLQYs and decay lifetimes with those of [Cu(xantphos)(6-Mebpy)][PF6] indicate that the introduction of the 6-substituent is not detrimental in terms of the photophysical properties.
APA, Harvard, Vancouver, ISO, and other styles
34

Chen, Chong, Fule Wu, Jiao Ji, Ai-Quan Jia, and Qian-Feng Zhang. "Synthesis, structural characterization and catalytic activity of chlororuthenium(II) complexes with substituted Schiff base/phosphine ancillary ligands." Zeitschrift für Naturforschung B 75, no. 9-10 (November 26, 2020): 851–57. http://dx.doi.org/10.1515/znb-2020-0110.

Full text
Abstract:
AbstractTreatment of [(η6-p-cymene)RuCl2]2 with one equivalent of chlorodiphenylphosphine in tetrahydrofuran at reflux afforded a neutral complex [(η6-p-cymene)RuCl2(κ1-P-PPh2OH)] (1). Similarly, the reaction of [Ru(bpy)2Cl2·2H2O] (bpy = 2,2′-bipyridine) and chlorodiphenylphosphine in methanol gave a cationic complex [Ru(bpy)2Cl(κ1-P-PPh2OCH3)](PF6) (2), while treatment of [RuCl2(PPh3)3] with [2-(C5H4N)CH=N(CH2)2N(CH3)2] (L1) in tetrahydrofuran at room temperature afforded a ruthenium(II) complex [Ru(PPh3)Cl2(κ3-N,N,N-L1)] (3). Interaction of the chloro-bridged complex [Ru(CO)2Cl2]n with one equivalent of [Ph2P(o-C6H4)CH=N(CH2)2N(CH3)2] (L2) led to the isolation of [Ru(CO)Cl2(κ3-P,N,N-L2)] (4). The molecular structures of the ruthenium(II) complexes 1–4 have been determined by single-crystal X-ray crystallography. The properties of the ruthenium(II) complex 4 as a hydrogenation catalyst for acetophenone were also tested.
APA, Harvard, Vancouver, ISO, and other styles
35

Miller, Jack M., and Kesagapillai Balasanmugam. "Characterization of metal complexes of 1,10-phenanthroline, 2,2′-bipyridine, and their derivatives by fast atom bombardment mass spectrometry." Canadian Journal of Chemistry 67, no. 9 (September 1, 1989): 1496–500. http://dx.doi.org/10.1139/v89-228.

Full text
Abstract:
A series of 1,10-phenanthroline (phen), 5-nitro-1,10-phenanthroline (5-NO2-phen), 2,9-dimethyl- 1,10-phenanthroline (dmp), 2,2′-bipyridine (bpy), and 4,4′-dimethy-2,2′-bipyridine (dmb) metal complexes has been characterized using fast atom bombardment mass spectrometry (FAB MS). Intact cations are observed in the positive ion FAB mass spectra of all tetracoordinated complexes, [Ag(L)2]NO3, [Cu(L)2]2 SO4•5H2O, [Cu(L)2]SO4, and [Tl(L)2](ClO4), where L=bpy, phen, or dmp. Structurally significant fragment ions (ML2+, ML+, M+, and (L+H)+, where M+=metal ion) are also observed. Species such as Ag[Ag(L)2]NO3+, where L=bpy or phen, and Tl[Tl(phen)2](ClO4)+ are also seen. The hexacoordinate complexes, [Co(phen)3]Cl2, [Ru(bpy)Cl2], [Fe(phen)2(CN)2], [Fe(L)3](ClO4), where L=5-NO2-phen, dmb, or bpy, and [Fe(L)3]SO4, where L=bpy or phen, show intact cations in the positive ion spectra; [Ni(bpy)3]X2, where X=Cl−,Br−, NO3−, ClO4−, SCN−, and tartrate2−, and [Co(bpy)3](ClO4) do not. Generally, fragment ions such as ML3+, ML2X+, ML2+, MLX+, and (L+H)+ are observed, where X=singly charged intact anion or HSO4−. Comparison of FAB results with those of laser mass spectrometry (LMS) show that they are comparable, through reduction processes are more important for FAB. LMS always gave the intact cation while FAB was not consistent. Keywords: FAB mass spectrometry, phenanthroline complexes, bipyridine complexes, laser desorption mass spectrometry.
APA, Harvard, Vancouver, ISO, and other styles
36

Krausz, E., H. Riesen, and AD Rae. "Crystal Structures of and Polarized Absorption-Spectroscopy in Racemic and Resolved [Ru(bpy)3] (ClO4)2 and [Zn(bpy)3] (ClO4)2Bpy=2,2'-Bipyridine." Australian Journal of Chemistry 48, no. 5 (1995): 929. http://dx.doi.org/10.1071/ch9950929.

Full text
Abstract:
[Zn( bpy )3] (ClO4)2 and [ Ru ( bpy )3] (ClO4)2 are isomorphous in both their racemic and resolved crystal forms. The resolved materials are monohydrates and have a C 2, Z = 8, structure with two independent formula units on general sites in the asymmetric unit. The cations have the same chirality. The inherent threefold axis of each cation lies approximately parallel to the c axis. The unrelated racemic form has a C2/c, Z = 4, structure which is a commensurate modulation of a P3c1, Z = 2, parent structure, typified by the room-temperature structure of [ Ru ( bpy )3] (PF6)2. A primary, secondary and tertiary axis of P3c1 become the c, b and a axes respectively of C2/c, retaining a third of the symmetry elements of P3c1. The crystals grow as multiply contacted twins. This structure bas just one spectroscopic site with the cation lying on a twofold axis that passes through the metal and one of the bidendate ligands and relates the other two ligands to each other. This feature is particularly useful in the study of the optical spectroscopy of the metal-to- ligand charge transfer excitations of [ Ru ( bpy )3]2+ and related systems. A comparison of structural and spectral data indicates that the positions of the anions have a dominant influence on the relative energies of the metal-to- ligand excitations. An energy difference between excitations involving the two (lower-energy) equivalent ligands and the third ligand of the order of 800 cm-1 is indicated in both singlet and triplet regions for the racemic perchlorate. The absorption spectra of [ Ru ( bpy )3]2+and [Os( bpy )3]2+ in a number of crystalline hosts are compared and discussed.
APA, Harvard, Vancouver, ISO, and other styles
37

Ferguson, J., F. Herren, E. R. Krausz, M. Maeder, and J. Vrbancich. "Electronic spectroscopy of M(bpy)2+3 (M = Fe, Ru, Os), Cr(bpy)3+3 and related compounds." Coordination Chemistry Reviews 64 (May 1985): 21–39. http://dx.doi.org/10.1016/0010-8545(85)80039-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Liu, Ming, Jianmei Hu, Meng Yu, Weiwei Fan, Wenwen Ding, and Yong Wang. "The syntheses, crystal structures, electrochemical and magnetic properties of tri-nuclear cyanide-bridged complexes [cis-MII(bpy)2(CN)2]2MnIII(salcy) (PF6) (M = Fe, Ru, Os)." New Journal of Chemistry 43, no. 7 (2019): 3162–68. http://dx.doi.org/10.1039/c8nj04142c.

Full text
Abstract:
Three cyanide-bridged heteronuclear complexes, [cis-M(bpy)2(CN)2]2Mn(salcy) (PF6) (M = Fe, 1; M = Ru, 2; M = Os, 3; bis(salicylideneiminato) dianion) were synthesized and fully characterized. Complexes 1 and 2 show low temperature antiferromagnetic order.
APA, Harvard, Vancouver, ISO, and other styles
39

AGHABOZORG, Hossein, Fahimeh MOHAMAD PANAH, and Elahe SADR-KHANLOU. "Crystal Structure of [Fe(bpy)3][Fe(pydc)2]2(pydcH2)1/2·6.5H2O Complex (bpy = 2,2′-bipyridine, pydc = pyridine-2,6-dicarboxylate)." Analytical Sciences: X-ray Structure Analysis Online 23 (2007): x139—x140. http://dx.doi.org/10.2116/analscix.23.x139.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Schumann, Hans, and Liliana Eguren. "Synthese neuer Phosphan- und Fluorophosphankomplexe durch Ligandmodifikation in der Koordinationssphäre von kationischen Cyclopentadienyleisen-bis(phosphan)-Komplexen / Synthesis of New Phosphane and Fluorophosphane Complexes by Ligand Modification in the Coordination Sphere of Cationic Cyclopentadienyliron Bis(phosphane) Complexes." Zeitschrift für Naturforschung B 46, no. 7 (July 1, 1991): 887–95. http://dx.doi.org/10.1515/znb-1991-0707.

Full text
Abstract:
The reaction of nitrile solvate complexes [(C5H5)(P(OCH3)3)2(NCCH3)Fe]PF6 (Ia) and [(C5H5)(DPPE)(NCCH3)Fe]PF6 (Ib, DPPE = [(C6H5)2PCH2]2) with halogenophosphanes, -arsines or -stibines L′ afford the cationic complexes [(C5H5)(L2)(L′)Fe]PF6 (II, L2 = (P(OCH3)3)2, L′ = PCl3, P(CH3)Cl2, P(t-C4H9)Cl2, P(C6H5)Cl2, P(C6H5)2Cl, PBr3, AsCl3, SbCl3; III, L2 = DPPE, L′ = PCl3, AsCl3, SbCl3) in high yield. Spectroscopic data are given together with the characterization of [(C5H5)(P(OCH3)3)2(P(t-C4H9)Cl2)Fe]PF6 (IIc) by single crystals X-ray diffraction analysis. Through reduction of coordinated P– Cl ligands in selected complexes II and III complexes bearing PH3, P(C6H5)H2 and P(C6H5)2H ligands are available. With anhydrous KF in acetonitrile solution in the presence of dibenzo-18-crown-6, the coordinated P– Cl ligands in selected complexes II and HIII undergo Cl/F-exchange with formation of related complexes with P(CH3)ClF, P(CH3)F2, PF2Cl and PF3 ligands.
APA, Harvard, Vancouver, ISO, and other styles
41

Izarova, Natalia V., Maxim N. Sokolov, Alexander Rothenberger, Lukasz Ponikiewski, Dieter Fenske, and Vladimir P. Fedin. "Synthesis and crystal structure of a new metal–organic coordination polymer [Fe(4,4′-bpy)3(H2O)2](PF6)2·2(4,4′-bpy)·5 H2O with nanosized channels clathrate large organic molecules." Comptes Rendus Chimie 8, no. 6-7 (June 2005): 1005–10. http://dx.doi.org/10.1016/j.crci.2004.11.006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Braun, Dieter, and Hartmut Yersin. "Energy Transfer between Different Sites in Neat Single-Crystal [Ru(bpy)3](PF6)2." Inorganic Chemistry 34, no. 7 (March 1995): 1967–68. http://dx.doi.org/10.1021/ic00111a054.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Hensler, Gerold, Erich Gallhuber, and Hartmut Yersin. "Fine structure in the emission spectrum of [Ru(bpy)3] (PF6)2 single crystals." Inorganica Chimica Acta 113, no. 2 (March 1986): 91–94. http://dx.doi.org/10.1016/s0020-1693(00)82227-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Field, Ryan, Lai Chung Liu, Wojciech Gawelda, Cheng Lu, and R. J. Dwayne Miller. "Spectral Signatures of Ultrafast Spin Crossover in Single Crystal [FeII (bpy)3 ](PF6 )2." Chemistry - A European Journal 22, no. 15 (February 25, 2016): 5118–22. http://dx.doi.org/10.1002/chem.201600374.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Braun, Dieter, Erich Gallhuber, and Hartmut Yersin. "Zeeman splittings of the two lowest excited states of [Ru(bpy)3](PF6)2." Chemical Physics Letters 171, no. 1-2 (July 1990): 122–26. http://dx.doi.org/10.1016/0009-2614(90)80061-h.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Nguyen, Luong Lam, Régis Guillot, Jérôme Laisney, Lionel Rechignat, Salma Bedoui, Gabor Molnár, Eric Rivière, and Marie-Laure Boillot. "Fe(Me2-bpy)2(NCSe)2spin-crossover micro- and nanoparticles showing spin-state switching above 250 K." New Journal of Chemistry 39, no. 3 (2015): 1603–10. http://dx.doi.org/10.1039/c4nj01257g.

Full text
Abstract:
Nano- and microparticles or polycrystalline powders of the Fe(Me2-bpy)2(NCSe)2spin-crossover complex were easily elaborated from the diamagnetic precursor [Fe(Me2-bpy)3](NCSe)2·S by precipitation in an anti-solvent or by solid-state thermolysis.
APA, Harvard, Vancouver, ISO, and other styles
47

Znovjyak, Kateryna, Igor O. Fritsky, Iryna A. Golenya, Tatiana Y. Sliva, and Matti Haukka. "Crystal structure and magnetic properties of (tris{4-[1-(2-methoxyethyl)imidazol-2-yl]-3-azabut-3-enyl}amine)iron(II) bis(hexafluoridophosphate)." Acta Crystallographica Section E Crystallographic Communications 75, no. 3 (February 12, 2019): 358–61. http://dx.doi.org/10.1107/s2056989019001531.

Full text
Abstract:
In the complex cation of the title compound, [Fe(C27H41N10O3)](PF6)2, the tripodal tris{4-[1-(2-methoxyethyl)imidazol-2-yl]-3-azabut-3-enyl}amine ligand is coordinated to an FeII ion through the nitrogen atoms of three imidazole and three imino groups. The Fe atom exhibits a distorted octahedral geometry. In the crystal, L and D antipodes are arranged in layers in the bc plane. Weak C...F and C—H...F/O contacts exist between the ligands of the complex cation and the PF6 − anions, generating a three-dimensional network. At 120 K, the FeII ion is in a low-spin state, with an average Fe—N bond distance of 1.970 (2) Å. On heating, the FeII ion converts to the high-spin state, as demonstrated by magnetic susceptibility measurements.
APA, Harvard, Vancouver, ISO, and other styles
48

Riesen, H., and E. Krausz. "Orientational dependences of the Zeeman effect in the lowest-excited states of [Zn(bpy)3](ClO4)2:Ru(II) and [Ru(bpy)3](PF6)2." Chemical Physics Letters 217, no. 5-6 (January 1994): 613–18. http://dx.doi.org/10.1016/0009-2614(93)e1420-l.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Alexander, Bruce D., Trevor J. Dines, and Rayne W. Longhurst. "DFT calculations of the structures and vibrational spectra of the [Fe(bpy)3]2+ and [Ru(bpy)3]2+ complexes." Chemical Physics 352, no. 1-3 (September 2008): 19–27. http://dx.doi.org/10.1016/j.chemphys.2008.05.010.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

MacBeth, Cora E., Seth B. Harkins, and Jonas C. Peters. "Synthesis and characterization of cationic iron complexes supported by the neutral ligands NPi-Pr3, NArPi-Pr3, and NSt-Bu3." Canadian Journal of Chemistry 83, no. 4 (April 1, 2005): 332–40. http://dx.doi.org/10.1139/v05-017.

Full text
Abstract:
This paper compares the local geometries, spin states, and redox properties of a series of iron complexes supported by neutral, tetradentate NP3 (tris(phosphine)amine) and NS3 (tris(thioether)amine) ligands. Our consideration of an Fe-mediated N2 fixation scheme similar to that proposed by Chatt for molybdenum motivates our interest in systems of these types. This report specifically describes the synthesis and characterization of cationic Fe(II) chloride complexes supported by the neutral ligands NPi-Pr3 (NPi-Pr3 = [N(CH2CH2P-i-Pr2)3]), NArPi-Pr3 (NArPi-Pr3 = [N(2-diisopropylphosphine-4-methylphenyl)3]), and NSt-Bu3 (NSt-Bu3 = [N(CH2CH2S-t-Bu)3]). The solid-state structures, electrochemistry, and magnetic properties of these complexes are reported. Whereas the NPi-Pr3 and NArPi-Pr3 ligands provide pseudotetrahedral S = 2 ferrous cations [Fe(NPi-Pr3)Cl]PF6 (1[PF6]) and [Fe(NArPi-Pr3)Cl]BPh4 (2[BPh4]) featuring a long Fe—N bond distance, the NSt-Bu3 ligand gives rise to a trigonal bipyramidal structure with a S = 1 ground state and a much shorter Fe–N interaction. The complexes 1[BPh4] and 2[BPh4] can be reduced under CO to give rise to the five-coordinate Fe(I) monocarbonyls [Fe(NPi-Pr3)CO]BPh4 (4[BPh4]) and [Fe(NArPi-Pr3)CO]BPh4 (5[BPh4]). The solid-state structures and electrochemistry of 4[BPh4] and 5[BPh4] are described, as is the EPR spectrum of 4[BPh4]. The synthesis and characterization of the hydride–dinitrogen complex [Fe(NPi-Pr3)(N2)(H)]PF6 (6[PF6]) has also been accomplished and its properties are also reported.Key words: nitrogenase, iron, polydentate phosphines, thioether ligands, N2 chemistry, nitrogen, Fe(I).
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography