Journal articles on the topic 'CATIONIC LIGANDS'

To see the other types of publications on this topic, follow the link: CATIONIC LIGANDS.

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 'CATIONIC LIGANDS.'

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

Lamberts, Kevin, Andreas Möller, and Ulli Englert. "Enantiopure and racemic alanine as bridging ligands in Ca and Mn chain polymers." Acta Crystallographica Section B Structural Science, Crystal Engineering and Materials 70, no. 6 (December 1, 2014): 989–98. http://dx.doi.org/10.1107/s2052520614021398.

Full text
Abstract:
Under accelerated and controlled evaporation, chain polymers crystallize from aqueous solutions of CaIIand MnIIhalides with enantiopure L-alanine or racemic DL-alanine. In all ten solids thus obtained zwitterionic amino acid ligands bridge neighbouring cations. The exclusively O-donor-based coordination sphere around the metal cations is completed by aqua ligands; the halides remain uncoordinated and act as counter-anions for the cationic strands. Despite the differences in ionic radii and electronic structure between the main group and the transition metal cation, their derivatives with L-alanine share a common structure type. In contrast, the solids derived from DL-alanine differ and adopt structures depending on the metal cation and the halide. Homochiral chains of either chirality or heterochiral chains with different arrangements of crystallographic inversion centres along the polymer strands are encountered. On average, the six-coordinated CaIIcations, devoid of any ligand field effect, show more pronounced deviation from idealized octahedral geometry than thed-block cation MnII.
APA, Harvard, Vancouver, ISO, and other styles
2

Ogura, Yusaku, Masahiro Nakano, Hajime Maeda, Masahito Segi, and Taniyuki Furuyama. "Cationic Axial Ligand Effects on Sulfur-Substituted Subphthalocyanines." Molecules 27, no. 9 (April 26, 2022): 2766. http://dx.doi.org/10.3390/molecules27092766.

Full text
Abstract:
Herein, we report the synthesis of sulfur-substituted boron(III) subphthalocyanines (SubPcs) with cationic axial ligands. Subphthalocyanines were synthesized by a condensation reaction using the corresponding phthalonitriles and boron trichloride as a template. An aminoalkyl group was introduced on the central boron atom; this process was followed by N-methylation to introduce a cationic axial ligand. The peripheral sulfur groups shifted the Q band of SubPcs to a longer wavelength. The cationic axial ligands increased the polarity and enhanced the hydrophilicity of SubPcs. The effect of axial ligands on absorption and fluorescence properties is generally small. However, a further red shift was observed by introducing cationic axial ligands into the sulfur-substituted SubPcs. This change is similar to that in sulfur-substituted silicon(IV) phthalocyanines. The unique effect of the cationic axial ligand was extensively investigated by theoretical calculations and electrochemistry. In particular, the precise oxidation potential was determined using ionization potential measurements. Thus, the results of the present study provide a novel strategy for developing functional dyes and pigments based on SubPcs.
APA, Harvard, Vancouver, ISO, and other styles
3

Wright, Stephen H. "Molecular and cellular physiology of organic cation transporter 2." American Journal of Physiology-Renal Physiology 317, no. 6 (December 1, 2019): F1669—F1679. http://dx.doi.org/10.1152/ajprenal.00422.2019.

Full text
Abstract:
Organic cation transporters play a critical role in mediating the distribution of cationic pharmaceuticals. Indeed, organic cation transporter (OCT)2 is the initial step in the renal secretion of organic cations and consequently plays a defining role in establishing the pharmacokinetics of many cationic drugs. Although a hallmark of OCTs is their broad selectivity, this characteristic also makes them targets for unwanted, adverse drug-drug interactions (DDIs), making them a focus for efforts to develop models of ligand interaction that could predict and preempt these adverse interactions. This review discusses the molecular characteristics of these transporters as well as the evidence that established the OCTs as key players in the distribution of organic cations. However, the primary focus is the present understanding of the complexity of ligand interaction with OCTs, particularly OCT2, including evidence for the presence of multiple ligand-binding sites and the influence of substrate structure on the affinity of the transporter for inhibitory ligands. This leads to a discussion of the complexities associated with the development of protocols for assessing the inhibitory potential of new molecular entities to perpetrate unwanted DDIs, the criteria that should be considered in the interpretation of the results of such protocols, and the challenges associated with development of models capable of predicting unwanted DDIs.
APA, Harvard, Vancouver, ISO, and other styles
4

Babukutty, Leslie, Ethan Moskovic, Davina Wadler, Thomas Strekas, and Robert Engel. "Polycations XX: New Monodentate Cationic Ligands and Their Coordination with Ruthenium for the Construction of Complexes Expressing Enhanced Interaction with DNA." Organic Chemistry International 2012 (October 15, 2012): 1–7. http://dx.doi.org/10.1155/2012/282137.

Full text
Abstract:
Prior investigations from this laboratory concerned with the preparation of new types of organic cations for a variety of biological and nonbiological applications have been extended to the preparation of cation-bearing ligands with nitrogen coordinating sites for use in complexation reactions with ruthenium cores. The syntheses of new cationic ligands as well as ruthenium complexes bearing them are reported here. The introduction of these new types of ligands is intended to provide to the complexes an enhanced ability to interact with DNA, and thereby to have the potential to be enhanced antitumor agents. Preliminary observations of their interactions with DNA are presented.
APA, Harvard, Vancouver, ISO, and other styles
5

Zeng, Juxiang, Guodong Tang, and Jun Qian. "Synthesis and crystal structure of (1,10-phenanthroline-κ2 N,N′)[2-(1H-pyrazol-1-yl)phenyl-κ2 N 2,C 1]iridium(III) hexafluoridophosphate with an unknown number of solvent molecules." Acta Crystallographica Section E Crystallographic Communications 76, no. 6 (May 5, 2020): 803–6. http://dx.doi.org/10.1107/s2056989020005861.

Full text
Abstract:
The cationic complex in the title compound, [Ir(C9H7N2)2(C12H8N2)]PF6, comprises two phenylpyrazole (ppz) cyclometallating ligands and one 1,10-phenanthroline (phen) ancillary ligand. The asymmetric unit consists of one [Ir(ppz)2(phen)]+ cation and one [PF6]− counter-ion. The central IrIII ion is six-coordinated by two N atoms and two C atoms from the two ppz ligands as well as by two N atoms from the phen ligand within a distorted octahedral C2N4 coordination set. In the crystal structure, the [Ir(ppz)2(phen)]+ cations and PF6 − counter-ions are connected with each other through weak intermolecular C—H...F hydrogen bonds. Additional C—H...π interactions between the rings of neighbouring cations consolidate the three-dimensional network. Electron density associated with additional disordered solvent molecules inside cavities of the structure was removed with the SQUEEZE procedure in PLATON [Spek (2015). Acta Cryst. C71, 9–18]. The given chemical formula and other crystal data do not take into account the unknown solvent molecule(s). The title compound has a different space-group symmetry (C2/c) from its solvatomorph (P21/c) comprising 1.5CH2Cl2 solvent molecules per ion pair.
APA, Harvard, Vancouver, ISO, and other styles
6

Mbiangué, Yves Alain, Alfred Bijvédé, Patrice Kenfack Tsobnang, Emmanuel Wenger, and Claude Lecomte. "Crystal structure of diammonium bis[tris(oxamide dioxime-κ2 N,N′)nickel(II)] bis[tris(oxalato-κ2 O,O′)chromate(III)] 6.76-hydrate." Acta Crystallographica Section E Crystallographic Communications 76, no. 11 (October 9, 2020): 1732–36. http://dx.doi.org/10.1107/s2056989020013390.

Full text
Abstract:
The asymmetric unit of the title compound, (NH4)2[Ni(C2H6N4O2)3]2[Cr(C2O4)3]2·6.76H2O, comprises two NH4 + cations, two [Ni(C2H6N4O2)3]2+ cations and two [Cr(C2O4)3]3– anions, as well as eight water molecules of crystallization of which only one is fully occupied. In the cationic and anionic complexes, the central atoms (NiII and CrIII) are each surrounded by three bidentate ligands (N-chelating oxamide dioxime and O-chelating oxalate, respectively), resulting in distorted octahedral coordination spheres. In the crystal, O—H...O hydrogen bonds between the oxamide dioxime ligands as donor groups and the oxalate ligands as acceptor groups alternately connect the cationic and anionic complexes into infinite pillars extending parallel to [100]. Moreover, N—H...O hydrogen bonds between the same ligands connect neighboring pillars, thus delineating channels that accommodate the charge-balancing NH4 + cations as well as the water molecules of crystallization. Although the H atoms could not be localized for these two species, the corresponding N...O and O...O distances indicate hydrogen bonds of medium strength.
APA, Harvard, Vancouver, ISO, and other styles
7

Santiago, Tomás G., Carmen Urbaneja, Eleuterio Álvarez, Elena Ávila, Pilar Palma, and Juan Cámpora. "Neutral, cationic and anionic organonickel and -palladium complexes supported by iminophosphine/phosphinoenaminato ligands." Dalton Transactions 49, no. 2 (2020): 322–35. http://dx.doi.org/10.1039/c9dt04062e.

Full text
Abstract:
Ligand exchange and oxidative addition reactions allow the synthesis of Ni(ii) and Pd(ii) complexes with deprotonable iminophosphine ligands. The acid–base behavior of iminophosphine ligands coordinated to organometallic Ni(ii) fragments is analyzed.
APA, Harvard, Vancouver, ISO, and other styles
8

Rahimi, Naser, and Davit Zargarian. "Cationic tetra- and pentacoordinate complexes of nickel based on POCN- and POCOP-type pincer ligands: synthesis, characterization, and ligand exchange studies." New Journal of Chemistry 45, no. 33 (2021): 15063–73. http://dx.doi.org/10.1039/d1nj01355f.

Full text
Abstract:
The closely related pincer ligands POCN and POCOP display different electron donating properties and different degrees of resistance to ligand exchange reactions proceeding via cationic reaction intermediates.
APA, Harvard, Vancouver, ISO, and other styles
9

Wagler, Jörg, Uwe Böhme, Erica Brendler, and Gerhard Roewer. "First X-Ray Structure of a Cationic Silicon Complex with Salen-Type Ligand: An Unusual Compound with Two Different Si-N Dative Bonds." Zeitschrift für Naturforschung B 59, no. 11-12 (December 1, 2004): 1348–52. http://dx.doi.org/10.1515/znb-2004-11-1255.

Full text
Abstract:
Abstract Two novel compounds containing pentacoordinate alkylsiliconium-cations with an 〈O, N, O 〉- chelating ligand of salen-type were prepared by reacting the trimethylsilyl derivatives of the ligand with alkyltrichlorosilanes. Pentacoordination of the silicon atom is found in solid state as well as in solution. Crystals of compound 2a, ethylene-N, N’ -bis(2-oxy-4-methoxybenzophenoneiminato) methylsiliconium chloride, were obtained from chloroform solution. This complex crystallizes in monoclinic space group P21/n. The chloride ion is surrounded by three chloroform molecules in the solid state. The siliconium cation has trigonal bipyramidal geometry in the solid state, although the signals of two chemically equal half-sides of the salen-type ligand were revealed in 1H and 13C NMR spectra of its chloroform solution. Therefrom two different Si-N dative bonds within the same molecule arise. The reaction of methyltrichlorosilane with two non-linked “half-ligands” of the salen-type leads also to a siliconium complex with similar cationic coordination sphere motif.
APA, Harvard, Vancouver, ISO, and other styles
10

Schenk, Wolfdieter A., Ute Karl, and Michael R. Horn. "Kationische Schwefeldioxid-Komplexe des Eisens und Rutheniums vom Halbsandwich-Typ [1] / Cationic Halfsandwich-Type Sulfur Dioxide Complexes of Iron and Ruthenium [1]." Zeitschrift für Naturforschung B 44, no. 12 (December 1, 1989): 1513–18. http://dx.doi.org/10.1515/znb-1989-1208.

Full text
Abstract:
Cationic halfsandwich-type complexes of sulfur dioxide, [C5R5M(PR3)2(SO2)]+ (R = H, Me, M = Fe, Ru, (PR3)2 = mono- or bidentate phosphorus ligands) and [C5Me5Fe(CO)(PR3)(SO2)]+, are obtained by ligand exchange from labile cationic (M = Fe) or neutral (M = Ru) precursors. The new compounds are characterized by IR, 1H131H, 13C and 31P NMR spectroscopy. Their stability increases with increasing electron density at the metal.
APA, Harvard, Vancouver, ISO, and other styles
11

Zhou, Qin-Qin, Lei Jin, Wei-Qiang Liao, and Yi Zhang. "A potential molecular ferroelectric material: poly[[bis(1-aza-4-azoniabicyclo[2.2.2]octane)di-μ3-chlorido-tetra-μ2-chlorido-dichloridotricadmium(II)] dihydrate]." Acta Crystallographica Section C Crystal Structure Communications 69, no. 3 (February 16, 2013): 237–40. http://dx.doi.org/10.1107/s0108270113003892.

Full text
Abstract:
The title compound, {[Cd3(C6H13N2)2Cl8]·2H2O}n, consists of pendant protonated cationic diamine ligands bonded to an anionic one-dimensional coordination polymer chloridocadmate scaffold. Each coordination chain features two kinds of CdIIcentre, each with distorted octahedral coordination geometry. One CdIIcation lies on a centre of inversion and is coordinated by six bridging chloride ligands, while the other is coordinated by four bridging chloride ligands, one terminal chloride ligand and a 1-aza-4-azoniabicyclo[2.2.2]octane aza N atom. This gives a reversible corner-sharing half-cubic linear polymer that lies along the crystallographicadirection. The chains interact through hydrogen bonding with solvent water, with each water molecule accepting one N—H...O interaction from a cation and donating to two O—H...Cl interactions with anionic chains, thus linking three separate chains and completing the packing structure.
APA, Harvard, Vancouver, ISO, and other styles
12

Cuza, Emmelyne, Samia Benmansour, Nathalie Cosquer, Françoise Conan, Sébastien Pillet, Carlos J. Gómez-García, and Smail Triki. "Spin Cross-Over (SCO) Anionic Fe(II) Complexes Based on the Tripodal Ligand Tris(2-pyridyl)ethoxymethane." Magnetochemistry 6, no. 2 (June 7, 2020): 26. http://dx.doi.org/10.3390/magnetochemistry6020026.

Full text
Abstract:
Reactions of Fe(II) with the tripodal chelating ligand 1,1,1-tris(2-pyridyl)ethoxymethane (py3C-OEt) and (NCE)− co-ligands (E = S, Se, BH3) give a series of mononuclear complexes formulated as [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCE)3]2·2CH3CN, with E = S (1) and BH3 (2). These compounds are the first Fe(II) spin cross-over (SCO) complexes based on the tripodal ligand tris(2-pyridyl)ethoxymethane and on the versatile co-ligands (NCS)− and (NCBH3)−. The crystal structure reveals discrete monomeric isomorph structures formed by a cationic [Fe(py3C-OEt)2]2+ complex and by two equivalent anionic [Fe(py3C-OEt)(NCE)3]− complexes. In the cations the Fe(II) is facially coordinated by two py3C-OEt tripodal ligands whereas in the anion the three nitrogen atoms of the tripodal ligand are facially coordinated and the N-donor atoms of the three (NCE)− co-ligands occupy the remaining three positions to complete the distorted octahedral environment of the Fe(II) centre. The magnetic studies show the presence of gradual SCO for both complexes: A one-step transition around 205 K for 1 and a two-step transition for compound 2, centered around 245 K and 380 K.
APA, Harvard, Vancouver, ISO, and other styles
13

Bröring, Martin, Carsten D. Brandt, Silke Köhler, and Maximilian Sieber. "Crystallographic evidence for structural diversity in cationic (tripyrrinato)cobalt(II) complexes." Journal of Porphyrins and Phthalocyanines 09, no. 10 (October 2005): 683–90. http://dx.doi.org/10.1142/s1088424605000794.

Full text
Abstract:
Cobalt(II) complexes of tripyrrin ligands can be transformed into cations by anion exchange using NaBAr F , a salt of a weakly coordinating anion. The complete characterization of these highly sensitive cations (probably solvate complexes) could not be achieved, but the coordination to donor ligands like nitriles or trialkylphosphanes stabilizes the cations sufficiently for the isolation of some single crystals. As the structural analyses of these donor-stabilized cobalt(II) tripyrrins reveal, either four- or five coordinate structures are formed depending on the type and size of ligand used.
APA, Harvard, Vancouver, ISO, and other styles
14

Kilic, Ahmet, Eyyup Yasar, and Emine Aytar. "Neutral boron [(L1-3)BPh2] and cationic charged boron [(L1a-3a)BPh2] complexes for chemical CO2 conversion to obtain cyclic carbonates under ambient conditions." Sustainable Energy & Fuels 3, no. 4 (2019): 1066–77. http://dx.doi.org/10.1039/c8se00633d.

Full text
Abstract:
Three neutral salen ligands (L1-3) with three cationic charged salen ligands (L1a-3a) and the corresponding neutral boron [(L1-3)BPh2] and cationic charged boron [(L1a-3a)BPh2] compounds have been successfully synthesized and characterized under ambient conditions.
APA, Harvard, Vancouver, ISO, and other styles
15

Son, Jung-Su, Sung-Chul Lim, Hochun Lee, and Seung-Tae Hong. "Crystal structure ofcatena-poly[[[diaquacobalt(II)]-bis(μ-hex-3-enedinitrile-κ2N:N′)] bis(tetrafluoridoborate)]." Acta Crystallographica Section E Crystallographic Communications 71, no. 6 (May 23, 2015): m135—m136. http://dx.doi.org/10.1107/s2056989015009548.

Full text
Abstract:
In the structure of the title salt, [Co(C6H6N2)2(H2O)2](BF4)2, the CoIIatom is located on an inversion centre. The transition metal is in a slightly distorted octahedral coordination environment, defined by the cyano N atoms of four hex-3-enedinitrile ligands in equatorial positions and the O atoms of two water molecules in axial positions. The bridging mode of the hex-3-enedinitrile ligands leads to the formation of cationic chains extending parallel to [1-10]. The BF4−counter-anion is disordered over two sets of sites [occupancy ratio = 0.512 (19):0.489 (19)]. It is located in the voids between the cationic chains and is connected to the aqua ligands of the latter through O—H...F hydrogen bonds. One methylene H atom of the hex-3-enedinitrile ligand forms another and weak C—H...O hydrogen bond with a water O atom of a neighbouring chain, thus consolidating the three-dimensional network structure.
APA, Harvard, Vancouver, ISO, and other styles
16

Wang, Yanlan, Anna Monfredini, Pierre-Alexandre Deyris, Florent Blanchard, Etienne Derat, Giovanni Maestri, and Max Malacria. "All-metal aromatic cationic palladium triangles can mimic aromatic donor ligands with Lewis acidic cations." Chem. Sci. 8, no. 11 (2017): 7394–402. http://dx.doi.org/10.1039/c7sc03475j.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Terwingen, Steven van, and Ulli Englert. "Argentophilic Interactions in Two AgI Complexes of 3-(2-(Pyridin-4-yl)ethyl)pentane-2,4-dione, a Promising Ditopic Ligand." Crystals 9, no. 8 (August 9, 2019): 414. http://dx.doi.org/10.3390/cryst9080414.

Full text
Abstract:
Reactions of 3-(2-(pyridin-4-yl)ethyl)pentane-2,4-dione (HacacPyen) with AgPF6 and AgBF4 afforded cationic silver complexes [Ag(HacacPyen)2]+ with essentially linear coordination of the Ag I cation by two pyridine N atoms. Rather unexpectedly, the HacacPyen ligands in the PF6- salt 1 adopt the diketo form, in contrast to the uncoordinated HacacPyen molecule, whereas the corresponding BF4- salt 2 and the majority of 3-substituted acetylacetones crystallizes as the enol tautomer. In both compounds 1 and 2, complex cations aggregate via short Ag...Ag interactions to pairs. These contacts amount to 3 . 21 Å in 1 and 3 . 26 Å or 3 . 31 Å in 2. As they are unsupported by any additional bridging ligands and correspond to the closest interionic interactions between neighbouring complex cations, they may be addressed as argentophilic interactions. The PF6- anions in 1 and the BF4- counter ions in 2 are involved in long and presumably electrostatic Ag...F contacts of ca. 2 . 9 Å. Additional coordination between Ag I and keto O atoms of symmetry-equivalent ligands occurs in 1 and leads to an extended two-periodic supramolecular structure.
APA, Harvard, Vancouver, ISO, and other styles
18

Ju-Nam, Yon, Wanisa Abdussalam-Mohammed, and Jesus J. Ojeda. "Highly stable noble metal nanoparticles dispersible in biocompatible solvents: synthesis of cationic phosphonium gold nanoparticles in water and DMSO." Faraday Discussions 186 (2016): 77–93. http://dx.doi.org/10.1039/c5fd00131e.

Full text
Abstract:
In this work, we report the synthesis of novel cationic phosphonium gold nanoparticles dispersible in water and dimethyl sulfoxide (DMSO) for their potential use in biomedical applications. All the cationic-functionalising ligands currently reported in the literature are ammonium-based species. Here, the synthesis and characterisation of an alternative system, based on phosphonioalkylthiosulfate zwitterions and phosphonioalkylthioacetate were carried out. We have also demonstrated that our phosphonioalkylthiosulfate zwitterions readily disproportionate into phosphonioalkylthiolates in situ during the synthesis of gold nanoparticles produced by the borohydride reduction of gold(iii) salts. The synthesis of the cationic gold nanoparticles using these phosphonium ligands was carried out in water and DMSO. UV-visible spectroscopic and TEM studies have shown that the phosphonioalkylthiolates bind to the surface of gold nanoparticles which are typically around 10 nm in diameter. The resulting cationic-functionalised gold nanoparticles are dispersible in aqueous media and in DMSO, which is the only organic solvent approved by the U.S. Food and Drug Administration (FDA) for drug carrier tests. This indicates their potential future use in biological applications. This work shows the synthesis of a new family of phosphonium-based ligands, which behave as cationic masked thiolate ligands in the functionalisation of gold nanoparticles. These highly stable colloidal cationic phosphonium gold nanoparticles dispersed in water and DMSO can offer a great opportunity for the design of novel biorecognition and drug delivery systems.
APA, Harvard, Vancouver, ISO, and other styles
19

McKinty, Adam M., and Douglas W. Stephan. "Ru alkylidene compounds bearing tridentate, dianionic ligands: Lewis acid activation and olefin metathesis." Dalton Transactions 45, no. 9 (2016): 3844–52. http://dx.doi.org/10.1039/c5dt04481b.

Full text
Abstract:
The series of tridentate complexes were prepared and shown to react with BCl3 to give the complexes affording neutral and cationic complexes sequentially. The generated five coordinate cations were evaluated in standard preliminary tests for olefin metathesis catalysis.
APA, Harvard, Vancouver, ISO, and other styles
20

Xanthopoulos, Konstantinos, Zafeiria Anagnostou, Sophocles Chalkiadakis, Duane Choquesillo-Lazarte, Gellert Mezei, Jan K. Zaręba, Jerzy Zoń, and Konstantinos D. Demadis. "Platonic Relationships in Metal Phosphonate Chemistry: Ionic Metal Phosphonates." Crystals 9, no. 6 (June 11, 2019): 301. http://dx.doi.org/10.3390/cryst9060301.

Full text
Abstract:
Phosphonate ligands demonstrate strong affinity for metal ions. However, there are several cases where the phosphonate is found non-coordinated to the metal ion. Such compounds could be characterized as salts, since the interactions involved are ionic and hydrogen bonding. In this paper we explore a number of such examples, using divalent metal ions (Mg2+, Ca2+, Sr2+ and Ni2+) and the phosphonic acids: p-aminobenzylphosphonic acid (H2PABPA), tetramethylenediamine-tetrakis(methylenephosphonic acid) (H8TDTMP), and 1,2-ethylenediphosphonic acid (H4EDPA). The compounds isolated and structurally characterized are [Mg(H2O)6]·[HPABPA]2·6H2O, [Ca(H2O)8]·[HPABPA]2, [Sr(H2O)8]·[HPABPA]2, [Mg(H2O)6]·[H6TDTMP], and [Ni(H2O)6]·[H2EDPA]·H2O. Also, the coordination polymer {[Ni(4,4’-bpy)(H2O)4]·[H2EDPA]·H2O}n was synthesized and characterized, which contains a bridging 4,4’-bipyridine (4,4’-bpy) ligand forming an infinite chain with the Ni2+ cations. All these compounds contain the phosphonate anion as the counterion to charge balance the cationic charge originating from the metal cation.
APA, Harvard, Vancouver, ISO, and other styles
21

Henwood, Adam F., Ashu K. Bansal, David B. Cordes, Alexandra M. Z. Slawin, Ifor D. W. Samuel, and Eli Zysman-Colman. "Solubilised bright blue-emitting iridium complexes for solution processed OLEDs." Journal of Materials Chemistry C 4, no. 17 (2016): 3726–37. http://dx.doi.org/10.1039/c6tc00151c.

Full text
Abstract:
Combining a sterically bulky, electron-deficient cyclometalating CN ligands with an electron rich, highly rigidified NN ligand gives an iridium complex, that achieves extraordinarily bright blue emission (ΦPL = 90%; λmax = 459 nm in MeCN) for a cationic iridium complex.
APA, Harvard, Vancouver, ISO, and other styles
22

Delpech, Fabien, Ilia A. Guzei, and Richard F. Jordan. "Cationic Indium Alkyl Complexes Incorporating Aminotroponiminate Ligands." Organometallics 21, no. 6 (March 2002): 1167–76. http://dx.doi.org/10.1021/om010578x.

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

Endo, Koji, and Robert H. Grubbs. "Cationic ruthenium alkylidene catalysts bearing phosphine ligands." Dalton Transactions 45, no. 8 (2016): 3627–34. http://dx.doi.org/10.1039/c5dt04506a.

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

Tanaka, Kimiya, Makoto Tanabe, Tomohito Ide, and Kohtaro Osakada. "Cationic Hydridotriplatinum Complex with Bridging Germylene Ligands." Organometallics 33, no. 10 (May 7, 2014): 2608–12. http://dx.doi.org/10.1021/om500309k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Korolev, Andrey V., Eiji Ihara, Ilia A. Guzei, Victor G. Young,, and Richard F. Jordan. "Cationic Aluminum Alkyl Complexes Incorporating Aminotroponiminate Ligands." Journal of the American Chemical Society 123, no. 34 (August 2001): 8291–309. http://dx.doi.org/10.1021/ja010242e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Ihara, Eiji, Victor G. Young,, and Richard F. Jordan. "Cationic Aluminum Alkyl Complexes Incorporating Aminotroponiminate Ligands." Journal of the American Chemical Society 120, no. 32 (August 1998): 8277–78. http://dx.doi.org/10.1021/ja9817444.

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

Louie, Brenda M., Steven J. Rettig, Alan Storr, and James Trotter. "Rhenium and manganese carbonyl compounds incorporating tridentate chelating pyrazolyl gallate ligands." Canadian Journal of Chemistry 63, no. 8 (August 1, 1985): 2261–72. http://dx.doi.org/10.1139/v85-373.

Full text
Abstract:
The coordinating properties of a variety of unsymmetrical uninegative tridentate chelating "pyrazolylgallate" ligands have been studied using the tricarbonyl moieties "M(CO)3", where M = Mn or Re, as acceptor species. A series of monomeric, pseudo octahedral complexes has been characterized and a fac mode of coordination established for the tridentate gallate ligands from 1H nmr, ir measurements, and X-ray structure determinations. Nitrosylation of a selection of the rhenium tricarbonyl compounds has yielded a number of cationic rhenium mononitrosyl dicarbonyl species. The reactivity of these cations towards reducing agents has been investigated.
APA, Harvard, Vancouver, ISO, and other styles
28

Parshina, Yu P., T. A. Kovylina, A. N. Konev, A. A. Belikov, P. O. Baber, A. D. Komarova, E. A. Romaeva, and L. N. Bochkarev. "Norbornene-Substituted Cationic Iridium(III) Complex and Water-Soluble Luminescent Polymers Based on It: Synthesis, Photophysical and Cytotoxic Properties." Russian Journal of General Chemistry 92, no. 12 (December 2022): 2666–75. http://dx.doi.org/10.1134/s1070363222120167.

Full text
Abstract:
Abstract A norbornene-substituted cationic iridium(III) complex containing 1-phenylisoquinoline cyclometalating ligands and an additional phenylimidazophenanthroline ligand was synthesized. On the base of this complex, water-soluble polymers were obtained by ring-opening metathesis polymerization (ROMP). The resulting polymers showed oxygen-dependent phosphorescence in the orange spectral region and high cytotoxicity against HCT116 cancer cells.
APA, Harvard, Vancouver, ISO, and other styles
29

Shutkov, Ilya A., Yulia N. Okulova, Dmitrii M. Mazur, Nikolai A. Melnichuk, Denis A. Babkov, Elena V. Sokolova, Alexander A. Spasov, Elena R. Milaeva, and Alexey A. Nazarov. "New Organometallic Ru(II) Compounds with Lonidamine Motif as Antitumor Agents." Pharmaceutics 15, no. 5 (April 29, 2023): 1366. http://dx.doi.org/10.3390/pharmaceutics15051366.

Full text
Abstract:
The combination of one molecule of organic and metal-based fragments that exhibit antitumor activity is a modern approach in the search for new promising drugs. In this work, biologically active ligands based on lonidamine (a selective inhibitor of aerobic glycolysis used in clinical practice) were introduced into the structure of an antitumor organometallic ruthenium scaffold. Resistant to ligand exchange reactions, compounds were prepared by replacing labile ligands with stable ones. Moreover, cationic complexes containing two lonidamine-based ligands were obtained. Antiproliferative activity was studied in vitro by MTT assays. It was shown that the increase in the stability in ligand exchange reactions does not influence cytotoxicity. At the same time, the introduction of the second lonidamine fragment approximately doubles the cytotoxicity of studied complexes. The ability to induce apoptosis and caspase activation in tumour cell MCF7 was studied by employing flow cytometry.
APA, Harvard, Vancouver, ISO, and other styles
30

Chen, Xun, Steven Stout, Uwe Mueller, George Boykow, Richard Visconti, Phieng Siliphaivanh, Kerrie Spencer, et al. "Label-Free, LC-MS-Based Assays to Quantitate Small-Molecule Antagonist Binding to the Mammalian BLT1 Receptor." SLAS DISCOVERY: Advancing the Science of Drug Discovery 22, no. 9 (August 1, 2017): 1131–41. http://dx.doi.org/10.1177/2472555217719748.

Full text
Abstract:
We have developed and validated label-free, liquid chromatography–mass spectrometry (LC-MS)-based equilibrium direct and competition binding assays to quantitate small-molecule antagonist binding to recombinant human and mouse BLT1 receptors expressed in HEK 293 cell membranes. Procedurally, these binding assays involve (1) equilibration of the BLT1 receptor and probe ligand, with or without a competitor; (2) vacuum filtration through cationic glass fiber filters to separate receptor-bound from free probe ligand; and (3) LC-MS analysis in selected reaction monitoring mode for bound probe ligand quantitation. Two novel, optimized probe ligands, compounds 1 and 2, were identified by screening 20 unlabeled BLT1 antagonists for direct binding. Saturation direct binding studies confirmed the high affinity, and dissociation studies established the rapid binding kinetics of probe ligands 1 and 2. Competition binding assays were established using both probe ligands, and the affinities of structurally diverse BLT1 antagonists were measured. Both binding assay formats can be executed with high specificity and sensitivity and moderate throughput (96-well plate format) using these approaches. This highly versatile, label-free method for studying ligand binding to membrane-associated receptors should find broad application as an alternative to traditional methods using labeled ligands.
APA, Harvard, Vancouver, ISO, and other styles
31

Edwards, Gavin L., David St C. Black, Glen B. Deacon, and Laurence PG Wakelin. "Effect of charge and surface area on the cytotoxicity of cationic metallointercalation reagents." Canadian Journal of Chemistry 83, no. 6-7 (June 1, 2005): 969–79. http://dx.doi.org/10.1139/v05-110.

Full text
Abstract:
Reaction of a series of nitrogen donor ligands (1-phenylpyrazoles, 2-phenylpyridine, benzo[h]quinoline, 1-(2′-pyridyl)indole, 1-phenylindazole, and 2-phenylindazole) with palladium(II) and platinum(II) salts gave complexes where ortho-metallation had occurred resulting in bidentate binding to the metal centres through N and C atoms. These cyclometallated products were isolated as µ-chloro dimers. Subsequent treatment of these µ-chloro dimers with chelating diamines such as 1,2-ethanediamine converted them into 14 cationic (1+) complexes. Analogous coordination mixed ligand complexes (charge 2+) were prepared by reaction of dichloro(1,2-ethanediamine-N,N′)palladium(II) with aromatic diamines such as 2-(1′-pyrazolyl)pyridine, 2,2′-bipyridine, and 1,10-phenanthroline. The complexes exhibited growth inhibitory activity against L1210 mouse leukæmia cells in vitro over a wide concentration range; in general, the cyclometallated complexes were more active than the mixed ligand complexes, although one cyclometallated organoplatinum complex was less active than the mixed ligand analogue. Substitution around the periphery of the aromatic ligands also resulted in increased activity. One complex, derived from 1-(2'-pyridyl)indole, was tested in vivo and showed no significant antitumour inhibition against P388 leukæmia at doses below toxic levels. Key words: anticancer, metallointercalator, cyclometallation, palladium, platinum, cytotoxicity.
APA, Harvard, Vancouver, ISO, and other styles
32

Friese, Seth J., Samantha A. Labb, Conner J. Masteran, Savannah G. Albright, Bakr Ali, Hayley A. Chapman, Yijie Cheng, et al. "Synthesis of a Water-Soluble, Soft N-Donor BTzBP Ligand Containing Only CHON." Synlett 31, no. 14 (June 24, 2020): 1384–88. http://dx.doi.org/10.1055/s-0040-1707163.

Full text
Abstract:
A hydrophilic ligand that contains only C, H, O, and N substituents and uses a 6,6′-bis(1H-1,2,3-triazol-4-yl)-2,2′-bipyridine (BTzBP) structural core has been synthesized. The effect of adding water-soluble groups onto extractant ligands has been extensively studied to facilitate the efficient partitioning of 4f and transuranic 5f elements for the treatment of spent nuclear fuel. Soft, N-donor ligands exhibit greater binding affinities for the trivalent actinides over the trivalent lanthanides, making BTzBP ligands an ideal candidate in the search for extractants to be used on an industrial scale. To date, hydrophobic BTzBPs have been shown to exhibit physical and chemical properties that might be conducive to nuclear waste processing conditions. However, hydrophilic BTzBPs have yet to be reported. Herein, we show the synthesis of a hydrophilic BTzBP ligand featuring cationic water solubilizing groups attached to the bipyridal rings.
APA, Harvard, Vancouver, ISO, and other styles
33

Tikhonov, D. B., L. G. Magazanik, and E. I. Nagaeva. "Ligands of Acid-Sensing Ion Channel 1a: Mechanisms of Action and Binding Sites." Acta Naturae 11, no. 1 (March 15, 2019): 4–13. http://dx.doi.org/10.32607/20758251-2019-11-1-4-13.

Full text
Abstract:
The proton-gated cationic channels belonging to the ASIC family are widely distributed in the central nervous system of vertebrates and play an important role in several physiological and pathological processes. ASIC1a are most sensitive to acidification of the external medium, which is the reason for the current interest in their function and pharmacology. Recently, the list of ASIC1a ligands has been rapidly expanding. It includes inorganic cations, a large number of synthetic and endogenous small molecules, and peptide toxins. The information on the mechanisms of action and the binding sites of the ligands comes from electrophysiological, mutational and structural studies. In the present review, we attempt to present a systematic view of the complex pattern of interactions between ligands and ASIC1a.
APA, Harvard, Vancouver, ISO, and other styles
34

Xu, Li-Qing, Li-Ping Lu, and Miao-Li Zhu. "Analogy to a Chinese knot in an ion-pair copper(II)–neodymium(III) complex based on a hexadentate Schiff base condensation product of 5-bromosalicylaldehyde and glycylglycine." Acta Crystallographica Section C Crystal Structure Communications 69, no. 4 (March 14, 2013): 376–79. http://dx.doi.org/10.1107/s0108270113006367.

Full text
Abstract:
Self-assembly of CuCl2, NdCl3, 5-bromosalicylaldehyde and glycylglycine yields the ion-pair copper(II)–neodymium(III) complex, poly[[decaaquabis[μ3-2-({2-[(5-bromo-2-oxidobenzylidene)amino]acetyl}azanidyl)acetato]bis[μ2-2-({2-[(5-bromo-2-oxidobenzylidene)amino]acetyl}azanidyl)acetato]tetracopper(II)dineodymium(III)] bis{[2-({2-[(5-bromo-2-oxidobenzylidene)amino]acetyl}azanidyl)acetato]cuprate(II)} tetradecahydrate], {[Cu4Nd2(C11H8BrN2O4)4(H2O)10][Cu(C11H8BrN2O4)]2·14H2O}n. The anion is planar and mononuclear, showing an approximately square-planar coordination of the metal atom, while the cation is a hexanuclear centrosymmetric transition metal–lanthanide (Cu–Nd) heterometallic complex, with the independent copper cations in square-planar and square-pyramidal coordinations. The asymmetric unit comprises one half of this cation, one anion and seven solvent water molecules. The positions of the six metal centres in the cation reproduce a Chinese knot arrangement. The dipeptidic Schiff base releases three H atoms and can act as a tetradentate, a pentadentate or a hexadentate ligand. Longer interactions between the pentadentate ligands and the Jahn–Teller CuIIcation link the hexanuclear aggregates into cationic chains in the [010] direction in which 14- and 22-membered subloops occur. Extensive hydrogen bonding in three dimensions involves both the coordinated and the solvent water molecules.
APA, Harvard, Vancouver, ISO, and other styles
35

Egorova, B. V., O. A. Fedorova, and S. N. Kalmykov. "Cationic radionuclides and ligands for targeted therapeutic radiopharmaceuticals." Russian Chemical Reviews 88, no. 9 (September 3, 2019): 901–24. http://dx.doi.org/10.1070/rcr4890.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Liu, Shengming, Miguel-Angel Munoz-Hernandez, and David A. Atwood. "Bimetallic and cationic aluminum with N3O2 chelate ligands." Journal of Organometallic Chemistry 596, no. 1-2 (February 2000): 109–14. http://dx.doi.org/10.1016/s0022-328x(99)00575-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Gopinathan, Sarada, S. S. Deshpande, and C. Gopinathan. "Cationic Ruthenium(II) Carbonyl Complexes with Nitrile Ligands." Zeitschrift f�r anorganische und allgemeine Chemie 527, no. 8 (August 1985): 203–7. http://dx.doi.org/10.1002/zaac.19855270823.

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

Zhang, Chenghui, Libo Sun, Chuanqi Zhang, Song Wan, Zhiqiang Liang, and Jiyang Li. "Novel photo- and/or thermochromic MOFs derived from bipyridinium carboxylate ligands." Inorganic Chemistry Frontiers 3, no. 6 (2016): 814–20. http://dx.doi.org/10.1039/c6qi00013d.

Full text
Abstract:
Two new Zn-MOFs based on 1-(3,5-dicarboxyphenyl)-4,4′-bipyridinium bromide have been synthesized. Compound 1 has a 3D interpenetrated cationic framework with photo- and thermochromism, while compound 2 possesses a 2D layered cationic network with only photochromism.
APA, Harvard, Vancouver, ISO, and other styles
39

Laurie, J. C. V., L. Duncan, R. C. Haltiwanger, R. T. Weberg, and Mary Rakowski DuBois. "Activation of hydrogen by cationic cyclopentadienylmolybdenum dimers with sulfido ligands. 1. Cationic complexes derived from protonation of 1,2-alkenedithiolate ligands." Journal of the American Chemical Society 108, no. 20 (October 1986): 6234–41. http://dx.doi.org/10.1021/ja00280a021.

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

Knauer, Lena, Michael Knorr, Lydie Viau, and Carsten Strohmann. "Crystal structure of the coordination polymer catena-poly[[[(acetonitrile-κN)copper(I)]-μ3-1,3-dithiolane-κ3 S:S:S′] hexafluoridophosphate]." Acta Crystallographica Section E Crystallographic Communications 76, no. 1 (January 1, 2020): 38–41. http://dx.doi.org/10.1107/s205698901901627x.

Full text
Abstract:
The polymeric title compound, [Cu2(C2H3N)2(C3H6S2)2](PF6)2, represents an example of a one-dimensional coordination polymer resulting from the reaction of [Cu(MeCN)4][PF6] with 1,3-dithiolane. The cationic one-dimensional ribbon consists of two copper(I) centers each ligated by one acetonitrile molecule and interconnected through two bridging 1,3-dithiolane ligands. One S-donor site of each ligand is κ1-bound to Cu, whereas the second S atom acts as a four-electron donor, bridging two Cu atoms in a κ4-bonding mode. The positive charge of each copper cation is compensated for by a hexafluoridophosphate counter-ion. In the crystal, the polymer chains are linked by a series of C—H...F hydrogen bonds, forming a supramolecular framework.
APA, Harvard, Vancouver, ISO, and other styles
41

Ren, YiXin, Andrea M. Bruck, and Lisa F. Szczepura. "Octa-μ3-selenido-pentakis(triethylphosphane-κP)(trimethylacetonitrile-κN)-octahedro-hexarhenium(III) bis(hexafluoridoantimonate) trimethylacetonitrile monosolvate." Acta Crystallographica Section E Structure Reports Online 70, no. 7 (June 4, 2014): m242—m243. http://dx.doi.org/10.1107/s1600536814011982.

Full text
Abstract:
The crystal structure of the title compound, [Re6Se8{NCC(CH3)3}(Et3P)5](SbF6)2·NCC(CH3)3, contains a face-capped octahedral [Re6(μ3-Se)8]2+cluster core. The pseudo-centrosymmetric [Re6Se8]2+cluster core is bonded through the Re atoms to five triethylphosphane ligands and one trimethylacetonitrile ligand. No significant interactions are observed between the cationic cluster, the SbF6−anions and the trimethylacetonitrile solvent molecule.
APA, Harvard, Vancouver, ISO, and other styles
42

Du, Bin, Si-Chun Yuan, and Jian Pei. "Cationic Iridium Dendrimers: Synthesis and Photophysical Properties." Australian Journal of Chemistry 64, no. 9 (2011): 1211. http://dx.doi.org/10.1071/ch11143.

Full text
Abstract:
Two dendrimers, D1 and D2, containing the cationic iridium complexes (C1 and C2) as cores and truxene-functionalized chromophores as the branches, have been developed by a convergent synthetic strategy. The cationic complexes employ 3-(pyridin-2-yl)-1H-1,2,4-triazole and 2-(pyridin-2-yl)-benzimidazole derivatives as the ancillary ligands. To avoid the change in emission colour arising from the iridium complex, the conjugation between the dendron and the ligand is decoupled by separating them using the alkyl chain. An investigation of their photoluminescent features reveals that efficient energy transfer happens from the dendrons to the core in the solid state. Likewise, the charged dendritic structure is demonstrated to be an efficient method to improve the compatibility between the polar charged iridium complexes and typical hydrophobic hosts with the additional benefit of excellent solution processability. Both dendrimers exhibit strong solvatochromic behaviours in solvents and exclusive green and yellow-orange light in the solid state.
APA, Harvard, Vancouver, ISO, and other styles
43

Lee, Byungwoo, and Jung Hoon Song. "Surface Decomposition and Healing in Solution-Phase Ligand Exchange for Efficient Colloidal Quantum Dot Solar Cells." Science of Advanced Materials 14, no. 1 (January 1, 2022): 141–46. http://dx.doi.org/10.1166/sam.2022.4189.

Full text
Abstract:
The efficiency of solution-processed colloidal quantum dot (CQD) thin-film solar cells has been significantly improved recently. However, there is still potential for efficiency improvements, which can be realized through a deeper understanding of surface chemical treatment for CQDs. In this study, we developed CQD thin-film solar cells with an improved power conversion efficiency (PCE) of 11.97%. We accomplished this by simply controlling the species and their molar concentrations used in the surface chemical treatment in the solutionphase ligand exchange process. Iodine treatment that generates conductive CQDs induces surface defects via cationic decomposition of the CQD surface; healing the CQD surface requires an excess of lead cations in the solution-phase ligand exchange process. Accordingly, we found that manufacturing CQD thin-film solar cells that have high efficiencies requires well-controlled surface treatment to effectively exchange surface ligands and simultaneously minimize surface defects on the CQD surface.
APA, Harvard, Vancouver, ISO, and other styles
44

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
45

Liu, Ruihuan, Dongxin Ma, Chen Zhang, and Lian Duan. "Sublimable cationic iridium(iii) complexes with large steric hindrance for high-performance organic light-emitting diodes." Dalton Transactions 48, no. 26 (2019): 9669–75. http://dx.doi.org/10.1039/c9dt00495e.

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

Milocco, Francesca, Serhiy Demeshko, Franc Meyer, and Edwin Otten. "Ferrate(ii) complexes with redox-active formazanate ligands." Dalton Transactions 47, no. 26 (2018): 8817–23. http://dx.doi.org/10.1039/c8dt01597j.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Battistin, Federica, Alessio Vidal, Gabriele Balducci, and Enzo Alessio. "Investigating the reactivity of neutral water-soluble Ru(ii)–PTA carbonyls towards the model imine ligands pyridine and 2,2′-bipyridine." RSC Advances 10, no. 45 (2020): 26717–27. http://dx.doi.org/10.1039/d0ra05898j.

Full text
Abstract:
The reactivity of selected Ru(ii)-PTA carbonyls with potentially labile ligands (i.e. H2O, dmso and/or Cl) towards the model imine ligands pyridine and 2,2′-bipyridine was investigated, yielding several neutral and cationic water-soluble derivatives.
APA, Harvard, Vancouver, ISO, and other styles
48

Carballo, Rosa, Ana Belén Lago, Arantxa Pino-Cuevas, Olaya Gómez-Paz, Nuria Fernández-Hermida, and Ezequiel M. Vázquez-López. "Neutral and Cationic Chelidonate Coordination Polymers with N,N′-Bridging Ligands." Chemistry 3, no. 1 (February 11, 2021): 256–68. http://dx.doi.org/10.3390/chemistry3010019.

Full text
Abstract:
The biomolecule chelidonic acid (H2chel, 4-oxo-4H-pyran-2,6-dicarboxylic acid) has been used to build new coordination polymers with the bridging N,N′-ligands 4,4′-bipyridine (4,4-bipy) and 1,2-bis(4-pyridyl)ethane (bpe). Four compounds have been obtained as single crystals: 1D cationic coordination polymers [M(4,4-bipy)(OH2)4]2+ with chelidonate anions and water molecules in the second coordination sphere in 1∞[Zn(4,4-bipy)(H2O)4]chel·3H2O (2) and in the two pseudopolymorphic 1∞[Cu(4,4-bipy)(H2O)4]chel·nH2O (n = 3, 4a; n = 6, 4b), and the 2D neutral coordination polymers 2∞[Zn(chel)(4,4-bipy)(H2O)]·2H2O (1) and 2∞[Zn(chel)(bpe)(H2O)]·H2O (3) where the chelidonate anion acts as a bridging ligand. The effects of the hydrogen bonds on the crystal packing were analyzed. The role of the water molecules hosted within the crystals lattices was also studied.
APA, Harvard, Vancouver, ISO, and other styles
49

Cabeza, Javier A, Ignacio del Río, María C Goite, Enrique Pérez-Carreño, and Vanessa Pruneda. "Cationic Heterocycles as Ligands: Synthesis and Reactivity with Anionic Nucleophiles of Cationic Triruthenium Clusters Containing C-MetalatedN-Methylquinoxalinium orN-Methylpyrazinium Ligands." Chemistry - A European Journal 15, no. 30 (July 27, 2009): 7339–49. http://dx.doi.org/10.1002/chem.200901079.

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

Spiridonova, Yulia S., Yulia A. Nikolaeva, Anna S. Balueva, Elvira I. Musina, Igor A. Litvinov, Igor D. Strelnik, Vera V. Khrizanforova, Yulia G. Budnikova, and Andrey A. Karasik. "Synthesis and Structure of Iron (II) Complexes of Functionalized 1,5-Diaza-3,7-Diphosphacyclooctanes." Molecules 25, no. 17 (August 19, 2020): 3775. http://dx.doi.org/10.3390/molecules25173775.

Full text
Abstract:
In order to synthesize new iron (II) complexes of 1,5-diaza-3,7-diphosphacyclooctanes with a wider variety of the substituents on ligands heteroatoms (including functionalized ones, namely, pyridyl groups) and co-ligands, it was found that these ligands with relatively small phenyl, benzyl, and pyridin-2-yl substituents on phosphorus atoms in acetonitrile formed bis-P,P-chelate cis-complexes [L2Fe(CH3CN)2]2+ (BF4)2−, whereas P-mesityl-substituted ligand formed only monoligand P,P-complex [LFe(CH3CN)4]2+ (BF4)2−. 3,7-dibenzyl-1,5-di(1′-(R)-phenylethyl)-1,5-diaza-3,7-diphosphacyclooctane reacted with hexahydrate of iron (II) tetrafluoroborate in acetone to give an unusual bis-ligand cationic complex of the composition [L2Fe(BF4)]+ BF4−, where two fluorine atoms of the tetrafluoroborate unit occupied two pseudo-equatorial positions at roughly octahedral iron ion, according to X-ray diffraction data. 1,5-diaza-3,7-diphosphacyclooctanes replaced tetrahydrofurane and one of the carbonyl ligands of cyclopentadienyldicarbonyl(tetrahydrofuran)iron (II) tetrafluoroborate to form heteroligand complexes [CpFeL(CO)]+BF4−. The structural studies in the solid phase and in solutions showed that diazadiphosphacyclooctane ligands of all complexes adopted chair-boat conformations so that their nitrogen atoms were in close proximity to the central iron ion. The redox properties of the obtained complexes were performed by the cyclic voltammetry method.
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