Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 93
Filter
Add more filters










Publication year range
1.
Acta Crystallogr C Struct Chem ; 76(Pt 5): 427-432, 2020 May 01.
Article in English | MEDLINE | ID: mdl-32367823

ABSTRACT

The monoclinic crystal structure of Na2SO3(H2O)7 is characterized by an alternating stacking of (100) cationic sodium-water layers and anionic sulfite layers along [100]. The cationic layers are made up from two types of [Na(H2O)6] octahedra that form linear 1∞[Na(H2O)4/2(H2O)2/1] chains linked by dimeric [Na(H2O)2/2(H2O)4/1]2 units on both sides of the chains. The isolated trigonal-pyramidal sulfite anions are connected to the cationic layers through an intricate network of O-H...O hydrogen bonds, together with a remarkable O-H...S hydrogen bond, with an O...S donor-acceptor distance of 3.2582 (6) Å, which is about 0.05 Šshorter than the average for O-H...S hydrogen bonds in thiosalt hydrates and organic sulfur compounds of the type Y-S-Z (Y/Z = C, N, O or S). Structural relationships between monoclinic Na2SO3(H2O)7 and orthorhombic Na2CO3(H2O)7 are discussed in detail.

2.
Org Biomol Chem ; 16(19): 3672-3680, 2018 05 15.
Article in English | MEDLINE | ID: mdl-29717748

ABSTRACT

Racemic ethyl hydrogen (1-hydroxy-2-methylsulfanyl-1-phenylethyl)phosphonate was resolved with (R)-1-phenylethylamine. The (R)-configuration of the (-)-enantiomer was determined by chemical correlation. Esterification of the (-)-enantiomer with a substituted diazomethane derived from 3-hydroxy-1,3,5(10)-estratrien-17-one delivered two epimeric phosphonates separated by HPLC. Methylation with methyl fluorosulfate at the sulfur atom and treatment with a strong base induced an α-hydroxyphosphonate-phosphate rearrangement with formation of dimethyl sulphide and two enantiomerically pure enol phosphates. Their oily nature interfered with a single crystal X-ray structure analysis to determine the stereochemistry at the phosphorus atom.

3.
Chemistry ; 23(64): 16346-16356, 2017 Nov 16.
Article in English | MEDLINE | ID: mdl-28898483

ABSTRACT

Members of a series of iridacyclopentenes of composition [TpMe2 Ir(k2 -C,C-CH2 CR'=CRCH2 )(CO)] (TpMe2 =hydrotris(3,5-dimethylpyrazolyl)borate; R=R'=H, 1; R=Me, R'=H, 2; R=R'=Me, 3) have been subjected to common organic chemistry procedures for hydrogenation, cyclopropanation, epoxidation, water addition through hydroboration, cis-dihydroxylation, and ozonolysis. The stability of metallacycles 1-3, imparted by the presence of the co-ligands TpMe2 and CO, directs the reactivity towards the C=C double bonds, and furthermore the stereochemistry of the products formed is strongly dictated by the steric demands of the TpMe2 ligand. While the products obtained in some of the above-mentioned reactions are the expected ones from an organic chemistry point of view, in other cases the results differ from the outcomes of similar reactions carried out with the all-carbon counterparts.

4.
Monatsh Chem ; 147(10): 1713-1719, 2016.
Article in English | MEDLINE | ID: mdl-27729712

ABSTRACT

ABSTRACT: In the present work, we have prepared a series of octahedral Fe(II) complexes of the type trans-[Fe(PNP)(CO)2Cl]+-PNP are tridentate pincer-type ligands based on 2,6-diaminopyridine. These complexes are formed irrespective of the size of the substituents at the phosphorus sites and whether cis-[Fe(PNP)(Cl2)(CO)] or trans-[Fe(PNP)(Cl2)(CO)] are reacted with CO in the presence of 1 equiv of silver salts. X-ray structures of representative complexes are presented. Based on simple bonding considerations the selective formation of trans-dicarbonyl Fe(II) complexes is unexpected. In fact, DFT calculations confirm that trans-dicarbonyl complexes are indeed thermodynamically disfavored over the respective cis-dicarbonyl compounds, but are favored for kinetic reasons.

5.
Bioorg Med Chem ; 24(21): 5326-5339, 2016 11 01.
Article in English | MEDLINE | ID: mdl-27614920

ABSTRACT

Positron emission tomography (PET) using fluorine-18 (18F)-labeled 2-nitroimidazole radiotracers has proven useful for assessment of tumor oxygenation. However, the passive diffusion-driven cellular uptake of currently available radiotracers results in slow kinetics and low tumor-to-background ratios. With the aim to develop a compound that is actively transported into cells, 1-(6'-deoxy-6'-[18F]fluoro-ß-d-allofuranosyl)-2-nitroimidazole (ß-[18F]1), a putative nucleoside transporter substrate, was synthetized by nucleophilic [18F]fluoride substitution of an acetyl protected labeling precursor with a tosylate leaving group (ß-6) in a final radiochemical yield of 12±8% (n=10, based on [18F]fluoride starting activity) in a total synthesis time of 60min with a specific activity at end of synthesis of 218±58GBq/µmol (n=10). Both radiolabeling precursor ß-6 and unlabeled reference compound ß-1 were prepared in multistep syntheses starting from 1,2:5,6-di-O-isopropylidene-α-d-allofuranose. In vitro experiments demonstrated an interaction of ß-1 with SLC29A1 and SLC28A1/2/3 nucleoside transporter as well as hypoxia specific retention of ß-[18F]1 in tumor cell lines. In biodistribution studies in healthy mice ß-[18F]1 showed homogenous tissue distribution and excellent metabolic stability, which was unaffected by tissue oxygenation. PET studies in tumor bearing mice showed tumor-to-muscle ratios of 2.13±0.22 (n=4) at 2h after administration of ß-[18F]1. In ex vivo autoradiography experiments ß-[18F]1 distribution closely matched staining with the hypoxia marker pimonidazole. In conclusion, ß-[18F]1 shows potential as PET hypoxia radiotracer which merits further investigation.


Subject(s)
Hypoxia/diagnostic imaging , Imidazoles/analysis , Imidazoles/chemistry , Monosaccharides/analysis , Monosaccharides/chemistry , Neoplasms/diagnostic imaging , Positron-Emission Tomography , Radiopharmaceuticals/analysis , Radiopharmaceuticals/chemical synthesis , Animals , Cell Line, Tumor , Dose-Response Relationship, Drug , Humans , Hypoxia/pathology , Imidazoles/chemical synthesis , Imidazoles/pharmacokinetics , Mice , Molecular Structure , Monosaccharides/chemical synthesis , Monosaccharides/pharmacokinetics , Neoplasms/pathology , Radiopharmaceuticals/chemistry , Radiopharmaceuticals/pharmacokinetics , Structure-Activity Relationship , Tissue Distribution
6.
Chemistry ; 22(38): 13715-23, 2016 Sep 12.
Article in English | MEDLINE | ID: mdl-27535720

ABSTRACT

The reaction between the bis(ethylene) complex [TpRh(C2 H4 )2 ], 1, (Tp=hydrotris(pyrazolyl)borate), and dimethyl acetylenedicarboxylate (DMAD) has been studied under different experimental conditions. A mixture of products was formed, in which TpRh(I) species were prevalent, whereas the presence of trapping agents, like water or acetonitrile, allowed for the stabilization and isolation of octahedral TpRh(III) compounds. An excess of DMAD gave rise to a small amount of the [2+2+2] cyclotrimerization product hexamethyl mellitate (6). Although no catalytic application of 1 was achieved, mechanistic insights shed light on the formation of stable rhodium species representing the resting state of the catalytic cycle of rhodium-mediated [2+2+2] cyclo(co)trimerization reactions. Metallacyclopentene intermediate species, generated from the activation of one alkyne and one ethylene molecule from 1, and metallacyclopentadiene species, formed by oxidative coupling of two alkynes to the rhodium centre, are crucial steps in the pathways leading to the final organometallic and organic products.

7.
Organometallics ; 34(15): 3820-3832, 2015 Aug 10.
Article in English | MEDLINE | ID: mdl-26294806

ABSTRACT

The ortho-deprotonation of halide-substituted ferrocenes by treatment with lithium tetramethylpiperidide (LiTMP) has been investigated. Iodo-, bromo-, and chloro-substituted ferrocenes were easily deprotonated adjacent to the halide substituents. The synthetic applicability of this reaction was, however, limited by the fact that, depending on the temperature and the degree of halide substitution, scrambling of both iodo and bromo substituents at the ferrocene core took place. Iodoferrocenes could not be transformed selectively into ortho-substituted iodoferrocenes since, in the presence of LiTMP, the iodo substituents scrambled efficiently even at -78 °C, and this process had occurred before electrophiles had been added. Bromoferrocene and certain monobromo-substituted derivatives, however, could be efficiently ortho-deprotonated at low temperature and reacted with a number of electrophiles to afford 1,2- and 1,2,3-substituted ferrocene derivatives. For example, 2-bromo-1-iodoferrocene was synthesized by ortho-deprotonation of bromoferrocene and reaction with the electrophiles diiodoethane and diiodotetrafluoroethane, respectively. In this and related cases the iodide scrambling process and further product deprotonation due to the excess LiTMP could be suppressed efficiently by running the reaction at low temperature and in inverse mode. In contrast to the low-temperature process, at room temperature bromo substituents in bromoferrocenes scrambled in the presence of LiTMP. Chloro- and 1,2-dichloroferrocene could be ortho-deprotonated selectively, but in neither case was scrambling of a chloro substituent observed. As a further application of this ortho-deprotonation reaction, a route for the synthesis of 1,3-disubstituted ferrocenes was developed. 1,3-Diiodoferrocene was accessible from bromoferrocene in four steps. On a multigram scale an overall yield of 41% was achieved. 1,3-Diiodoferrocene was further transformed into symmetrically 1,3-disubstituted ferrocenes (1,3-R2Fc; R = CHO, COOEt, CN, CH=CH2).

8.
Dalton Trans ; 44(1): 281-94, 2015 Jan 07.
Article in English | MEDLINE | ID: mdl-25376759

ABSTRACT

Treatment of anhydrous FeX2 (X = Cl, Br) with 2 equiv. of the sterically little demanding N,N'-bisphosphino-2,6-diaminopyridine based PNP ligands--featuring Ph, biphenol (BIPOL), Me, Et, nPr, and nBu substituents at the phosphorus sites and H, Me, and Ph substituents at the N-linkers--afforded diamagnetic cationic octahedral complexes of the general formula [Fe(κ(3)-P,N,P-PNP)(κ(2)-P,N-PNP)X](+) featuring a κ(2)-P,N bound PNP ligand. With the sterically more encumbered N-methylated ligand PNP(Me)-Ph the related complex [Fe(κ(3)-P,N,P-PNP(Me)-Ph)(κ(2)-P,N-PN(HMe)-Ph)Cl](+) rather than [Fe(κ(3)-P,N,P-PNP(Me)-Ph)Cl2] was formed. This reaction was accompanied by P-N bond cleavage, thereby forming the κ(2)-P,N-bound N-diphenylphosphino-N,N'-methyl-2,6-diaminopyridine ligand. In contrast, with the N-phenylated ligands PNP(Ph)-Et and PNP(Ph)-nPr, despite small Et and nPr substituents at the phosphorus sites, complexes [Fe(κ(3)-P,N,P-PNP(Ph)-Et)Cl2] and [Fe(κ(3)-P,N,P-PNP(Ph)-nPr)Cl2] were formed, revealing that sterics can be also controlled by substituent variations at the amino N-sites. Depending on the solvent, complexes featuring κ(2)-P,N-bound ligands undergo facile rearrangement reactions to give dicationic complexes of the type [Fe(κ(3)-P,N,P-PNP)2](2+) where both PNP ligands are bound in a κ(3)-P,N,P-fashion. In the presence of either Ag(+) or Na(+) salts as halide scavengers this reaction takes place within a few minutes. The pendant PR2 arm of the κ(3)-κ(2)-complexes is readily oxidized to the corresponding phosphine sulfides upon treatment with elemental sulfur. This was exemplarily shown for [Fe(κ(3)-P,N,P-PNP-nPr)(κ(2)-P,N-PNS-nPr)Cl](+). Halide abstraction afforded the dicationic bis-chelated octahedral Fe(II) complex [Fe(κ(3)-P,N,P-PNP)2](2+) together with the free SNP ligand rather than [Fe(κ(3)-P,N,P-PNP-nPr)(κ(3)-S,P,N-PNS-nPr)](2+).

9.
Dalton Trans ; 43(39): 14669-79, 2014 Oct 21.
Article in English | MEDLINE | ID: mdl-25142749

ABSTRACT

In the present study a series of six-coordinate neutral 16e halocarbonyl Mo(ii) complexes of the type [Mo(PNP(Me)-iPr)(CO)X2] (X = I, Br, Cl), featuring the PNP pincer ligand N,N'-bis(diisopropylphosphino)-N,N'-dimethyl-2,6-diaminopyridine (PNP(Me)-iPr), were prepared and fully characterized. The synthesis of these complexes was accomplished by different methodologies depending on the halide ligands. For X = I and Br, [Mo(PNP(Me)-iPr)(CO)I2] and [Mo(PNP(Me)-iPr)(CO)Br2] were obtained by reacting [Mo(PNP(Me)-iPr)(CO)3] with stoichiometric amounts of I2 and Br2, respectively. In the case of X = Cl, [Mo(PNP(Me)-iPr)(CO)Cl2] was afforded by the reaction of [Mo(CO)4(µ-Cl)Cl]2 with 1 equiv. of PNP(Me)-iPr. The equivalent procedure also worked for X = Br. The modification of the 2,6-diaminopyridine scaffold by introducing NMe instead of NH spacers between the aromatic pyridine ring and the phosphine moieties changed the steric properties of the PNP-iPr ligand significantly. While in the present case exclusively neutral six-coordinate complexes of the type [Mo(PNP(Me)-iPr)(CO)X2] were obtained, with the parent PNP-iPr ligand, i.e. featuring NH spacers, cationic seven-coordinate complexes of the type [Mo(PNP-iPr)(CO)3X]X were afforded. Upon treatment of [Mo(PNP(Me)-iPr)(CO)X2] (X = Br, Cl) with Ag(+) in CH3CN, the cationic complexes [Mo(PNP(Me)-iPr)(CO)(CH3CN)X](+) were formed. Halide abstraction from [Mo(PNP(Me)-iPr)(CO)Cl2] in THF-CH2Cl2 afforded [Mo(PNP(Me)-iPr)(CO)(THF)Cl](+). In keeping with the facile synthesis of monocationic complexes preliminary ESI-MS and DFT/B3LYP studies revealed that one halide ligand in complexes [Mo(PNP(Me)-iPr)(CO)X2] is labile forming cationic fragments [Mo(PNP(Me)-iPr)(CO)X](+) which react with molecular oxygen in parallel pathways to yield mono and dioxo Mo(iv) and Mo(vi) species. Structures of representative complexes were determined by X-ray single crystal analyses.

10.
Dalton Trans ; 43(39): 14517-9, 2014 Oct 21.
Article in English | MEDLINE | ID: mdl-25138445

ABSTRACT

Treatment of anhydrous FeCl2 with 2 equiv. of the pincer ligand PNP-Ph afforded the diamagnetic cationic octahedral complex [Fe(κ(3)-P,N,P-PNP)(κ(2)-P,N-PNP)Cl](+) featuring a κ(2)-P,N-bound PNP ligand. Preliminary reactivity studies revealed that the κ(2)-P,N-bound PNP ligand is labile reacting with CO to afford trans-[Fe(PNP-Ph)(CO)2Cl](+).

11.
Organometallics ; 33(8): 1945-1952, 2014 Apr 28.
Article in English | MEDLINE | ID: mdl-24795493

ABSTRACT

Two representative Walphos analogues with an achiral 2,2″-biferrocenediyl backbone were synthesized. These diphosphine ligands were tested in the rhodium-catalyzed asymmetric hydrogenation of several alkenes and in the ruthenium-catalyzed hydrogenation of two ketones. The results were compared with those previously obtained on using biferrocene ligands with a C2-symmetric 2,2″-biferrocenediyl backbone as well as with those obtained with Walphos ligands. The application of one newly synthesized ligand in the hydrogenation of 2-methylcinnamic acid gave (R)-2-methyl-3-phenylpropanoic acid with full conversion and with 92% ee. The same ligand was used to transform 2,4-pentanedione quantitatively and diastereoselectively into (S,S)-2,4-pentanediol with 98% ee.

12.
Dalton Trans ; 43(29): 11152-64, 2014 Aug 07.
Article in English | MEDLINE | ID: mdl-24695811

ABSTRACT

Several new octahedral iron(ii) complexes of the type [Fe(PN(R)-Ph)2X2] (X = Cl, Br; R = H, Me) containing bidentate PN(R)-Ph (R = H, Me) (1a,b) ligands based on 2-aminopyridine were prepared. (57)Fe Mössbauer spectroscopy and magnetization studies confirmed in all cases their high spin nature at room temperature with magnetic moments very close to 4.9µB reflecting the expected four unpaired d-electrons in all these compounds. While in the case of the PN(H)-Ph ligand an S = 2 to S = 0 spin crossover was observed at low temperatures, complexes with the N-methylated analog PN(Me)-Ph retain an S = 2 spin state also at low temperatures. Thus, [Fe(PN(H)-Ph)2X2] (2a,3a) and [Fe(PN(Me)-Ph)2X2] (2b,3b) adopt different geometries. In the first case a cis-Cl,P,N-arrangement seems to be most likely, as supported by various experimental data derived from (57)Fe Mössbauer spectroscopy, SQUID magnetometry, UV/Vis, Raman, and ESI-MS as well as DFT and TDDFT calculations, while in the case of the PN(Me)-Ph ligand a trans-Cl,P,N-configuration is adopted. The latter is also confirmed by X-ray crystallography. In contrast to [Fe(PN(Me)-Ph)2X2] (2b,3b), [Fe(PN(H)-Ph)2X2] (2a,3a) is labile and undergoes rearrangement reactions. In CH3OH, the diamagnetic dicationic complex [Fe(PN(H)-Ph)3](2+) (5) is formed via the intermediacy of cis-P,N-[Fe(κ(2)-P,N-PN(H)-Ph)2(κ(1)-P-PN(H)-Ph)(X)](+) (4a,b) where one PN ligand is coordinated in a κ(1)-P-fashion. In CH3CN the diamagnetic dicationic complex cis-N,P,N-[Fe(PN(H)-Ph)2(CH3CN)2](2+) (6) is formed as a major isomer where the two halide ligands are replaced by CH3CN.

13.
Article in English | MEDLINE | ID: mdl-24098156

ABSTRACT

The crystal structure of tamarugite [sodium aluminium bis-(sulfate) hexa-hydrate] was redetermined from a single crystal from Mina Alcaparossa, near Cerritos Bayos, southwest of Calama, Chile. In contrast to the previous work [Robinson & Fang (1969 ▶). Am. Mineral. 54, 19-30], all non-H atoms were refined with anisotropic displacement parameters and H-atoms were located by difference Fourier methods and refined from X-ray diffraction data. The structure is built up from nearly regular [Al(H2O)6](3+) octa-hedra and infinite double-stranded chains [Na(SO4)2](3-) that extend parallel to [001]. The Na(+) cation has a strongly distorted octa-hedral coordination by sulfate O atoms [Na-O = 2.2709 (11) - 2.5117 (12) Å], of which five are furnished by the chain-building sulfate group S2O4 and one by the non-bridging sulfate group S1O4. The [Na(SO4)2](3-) chain features an unusual centrosymmetric group formed by two NaO6 octa-hedra and two S2O4 tetra-hedra sharing five adjacent edges, one between two NaO6 octa-hedra and two each between the resulting double octa-hedron and two S2O4 tetra-hedra. These groups are then linked into a double-stranded chain via corner-sharing between NaO6 octa-hedra and S2O4 tetra-hedra. The S1O4 group, attached to Na in the terminal position, completes the chains. The [Al(H2O)6](3+) octa-hedron (〈Al-O〉 = 1.885 (11) Å) donates 12 comparatively strong hydrogen bonds (O⋯O = 2.6665 (14) - 2.7971 (15) Å) to the sulfate O atoms of three neighbouring [Na(SO4)2](3-) chains, helping to connect them in three dimensions, but with a prevalence parallel to (010), the cleavage plane of the mineral. Compared with the previous work on tamarugite, the bond precision of Al-O bond lengths as an example improved from 0.024 to 0.001 Å.

14.
Acta Crystallogr C ; 69(Pt 10): 1085-90, 2013 Oct.
Article in English | MEDLINE | ID: mdl-24096490

ABSTRACT

The title compound, tricaesium sodium iron(III) µ3-oxido-hexa-µ2-sulfato-tris[aquairon(III)] pentahydrate, Cs2.91Na1.34Fe(3+)0.25[Fe3O(SO4)6(H2O)3]·5H2O, belongs to the family of Maus's salts, K5[Fe3O(SO4)6(H2O)3]·6H2O, which is based on the triaqua-µ3-oxido-hexa-µ-sulfato-triferrate(III) anion, [Fe3O(SO4)6(H2O)3](5-), with Fe in a characteristically distorted octahedral coordination environment, sharing a common corner via an oxide O atom. Cs in four different cation sites, Na in three different cation sites and five water molecules link the anions in three dimensions and set up a crystal structure in which those parts parallel to (001) and within 0.05 < z < 0.95 have a distinct trigonal pseudosymmetry, whereas the cation arrangement and bonding near z ∼ 0 generate a clear-cut noncentrosymmetric polar edifice with the monoclinic space group C2. The structure shows some cation disorder in the region near z ∼ 1/2 where one Na atom in octahedral coordination is partly substituted by Fe(3+), and a Cs atom is substituted by small amounts of Na on a separate nearby site. One Na atom, located on a twofold axis at z = 0 and tetrahedrally coordinated by four sulfate O atoms of two [Fe3O(SO4)6(H2O)3](5-) units, plays a key role in generating the noncentrosymmetric structure. Three of the seven different cation sites are on twofold axes (one Na(+) site and two Cs(+) sites), and all other atoms of the structure are in general positions.

15.
Org Biomol Chem ; 11(46): 8092-102, 2013 Dec 14.
Article in English | MEDLINE | ID: mdl-24163003

ABSTRACT

A practical synthesis of novel coordinating chiral ionic liquids with an amino alcohol structural motif was developed starting from commercially available amino alcohols. These basic chiral ionic liquids could be successfully applied as catalysts in the asymmetric alkylation of aldehydes and gave high enantioselectivities of up to 91% ee.

16.
J Organomet Chem ; 735(100): 80-87, 2013 07 01.
Article in English | MEDLINE | ID: mdl-23825866

ABSTRACT

Treatment of [FeCp(CO)2Cl] with 1 equiv of the amidophosphine ligands Li[R2PNR'] (R = Ph, iPr, R' = iPr, tBu, Cy) afforded complexes of the type [FeCp(CO)(κ2(C,P)-(C = O)-NiPr-PPh2)] (1a), [FeCp(CO)(κ2(C,P)-(C = O)-NtBu-PPh2)] (1b), and [FeCp(CO)(κ2(C,P)-(C = O)-NCy-PiPr2)] (1c) in 40-50% yields. Complex 1a was also formed when [FeCp(CO)2(PPh2NHiPr)]+ (2) was reacted with 1 equiv of KOtBu. These complexes feature a four-membered carboxamido-phospha-ferracycle as a result of an intramolecular nucleophilic attack of the amidophosphine ligand on coordinated CO. Upon treatment of 1a with the electrophile [Me3O]BF4 the aminocarbene complex [FeCp(CO)(κ2(C,P) = C(OMe)-NiPr-PPh2)]+ (3) was obtained bearing an aza-phospha-carbene moiety. Upon treatment of cis,trans,cis-[Fe(CO)2(Ph2PNHiPr)2(Br)2] (4a) and cis,trans,cis-[Fe(CO)2(Ph2PNHtBu)2(Br)2] (4b) with KOtBu the carboxamido-phospha-ferracycles trans-[Fe(CO)2(κ2(C,P)-(C = O)-NiPr-PPh2)(Ph2PNHiPr)Br] (5a) and trans-[Fe(CO)2(κ2(C,P)-(C = O)-NtBu-PPh2)(Ph2PNHtBu)Br] (5b) were formed in moderate yield. Finally, representative structures were determined by X-ray crystallography.

17.
Organometallics ; 32(10): 3042-3052, 2013 May 24.
Article in English | MEDLINE | ID: mdl-23794778

ABSTRACT

In the present study the Mo(0) and W(0) complexes [M(PNP)(CO)3] as well as seven-coordinate cationic hydridocarbonyl Mo(II) and W(II) complexes of the type [M(PNP)(CO)3H]+, featuring PNP pincer ligands based on 2,6-diaminopyridine, have been prepared and fully characterized. The synthesis of Mo(0) complexes [Mo(PNP)(CO)3] was accomplished by treatment of [Mo(CO)3(CH3CN)3] with the respective PNP ligands. The analogous W(0) complexes were prepared by reduction of the bromocarbonyl complexes [W(PNP)(CO)3Br]+ with NaHg. These intermediates were obtained from the known dinuclear complex [W(CO)4(µ-Br)Br]2, prepared in situ from W(CO)6 and stoichiometric amounts of Br2. Addition of HBF4 to [M(PNP)(CO)3] resulted in clean protonation at the molybdenum and tungsten centers to generate the Mo(II) and W(II) hydride complexes [M(PNP)(CO)3H]+. The protonation is fully reversible, and upon addition of NEt3 as base the Mo(0) and W(0) complexes [M(PNP)(CO)3] are regenerated quantitatively. All heptacoordinate complexes exhibit fluxional behavior in solution. The mechanism of the dynamic process of the hydrido carbonyl complexes was investigated by means of DFT calculations, revealing that it occurs in a single step. The structures of representative complexes were determined by X-ray single-crystal analyses.

18.
Acta Crystallogr Sect E Struct Rep Online ; 69(Pt 5): i30-i31, 2013 May 01.
Article in English | MEDLINE | ID: mdl-23723753

ABSTRACT

The title compound, Li3AsS4·8H2O, is built up from infinite cationic [Li3(H2O)8](3+) chains which extend along [001] and are cross-linked by isolated tetra-hedral AsS4 (3-) anions via O-H⋯S hydrogen bonds. Two Li and two As atoms lie on special positions with site symmetries -1 (1 × Li) and 2 (1 × Li and 2 × As). The [Li3(H2O)8](3+) chain contains four independent Li atoms of which two are in octa-hedral and two in tetra-hedral coordination by water O atoms. An outstanding feature of this chain is a linear group of three edge-sharing LiO6 octa-hedra to both ends of which two LiO4 tetra-hedra are attached by face-sharing. Such groups of composition Li5O16 are linked into branched chains by means of a further LiO4 tetra-hedron sharing vertices with four adjacent LiO6 octa-hedra. The Li-O bonds range from 1.876 (5) to 2.054 (6) Šfor the LiO4 tetra-hedra and from 2.026 (5) to 2.319 (5) Šfor the LiO6 octa-hedra. The two independent AsS4 (3-) anions have As-S bond lengths ranging from 2.1482 (6) to 2.1677 (6) Š[ = 2.161 (10) Å]. The eight independent water mol-ecules of the structure donate 16 relatively straight O-H⋯S hydrogen bonds to all S atoms of the AsS4 tetra-hedra [ = 3.295 (92) Å]. Seven water mol-ecules are in distorted tetra-hedral coordination by two Li and two S; one water mol-ecule has a flat pyramidal coordination by one Li and two S. At variance with related compounds like Schlippe's salt, Na3SbS4·9H2O, there are neither alkali-sulfur bonds nor O-H⋯O hydrogen bonds in the structure.

19.
Acta Crystallogr Sect E Struct Rep Online ; 69(Pt 5): m234-5, 2013 May 01.
Article in English | MEDLINE | ID: mdl-23723755

ABSTRACT

In the mononuclear title iridium(III) complex, [Ir(C4H8)(C15H22BN6)(C3H9P)], which is based on the [tris-(3,5-dimethyl-pyrazol-1-yl)hydro-borato]iridium moiety, Ir[Tp(Me2)], the Ir(III) atom is coordinated by a chelating butane-1,4-diyl fragment and a trimethyl-phosphane ligand in a modestly distorted octa-hedral coordination environment formed by three facial N, two C and one P atom. The iridium-butane-1,4-diyl ring has an envelope conformation. This ring is disordered because alternately the second or the third C atom of the butane-1,4-diyl fragment function as an envelope flap atom (the occupancy ratio is 1:1). In the crystal, mol-ecules are organized into densely packed columns extending along [101]. Coherence between the mol-ecules is essentially based on van der Waals inter-actions.

20.
Acta Crystallogr Sect E Struct Rep Online ; 69(Pt 4): m224-5, 2013 Apr 01.
Article in English | MEDLINE | ID: mdl-23634017

ABSTRACT

In the title compound, [Ir(C15H22BN6)(C8H7O)Cl]·CHCl3, the Ir atom is formally trivalent and is coordinated in a slightly distorted octa-hedral geometry by three facial N atoms, one C atom, one O atom and one Cl atom. The Ir=Ccarbene bond is strong and short and exerts a notable effect on the trans-Ir-N bond, which is about 0.10 Šlonger than the two other Ir-N bonds. The chloro-form solvent mol-ecule is anchored via a weak C-H⋯Cl hydrogen bond to the Cl atom of the Ir complex mol-ecule. In the crystal, the constituents adopt a layer-like arrangement parallel to (010) and are held together by weak inter-molecular C-H⋯Cl hydrogen bonds, as well as weak Cl⋯Cl [3.498 (2) Å] and Cl⋯π [3.360 (4) Å] inter-actions. A weak intra-molecular C-H⋯O hydrogen bond is also observed.

SELECTION OF CITATIONS
SEARCH DETAIL