Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 10 de 10
Filter
Add more filters










Publication year range
1.
Inorg Chem ; 45(20): 8318-25, 2006 Oct 02.
Article in English | MEDLINE | ID: mdl-16999432

ABSTRACT

The novel complexes [Zn(L)Cl] (1), [Cd(L)Cl] (2), [Hg(L)Cl] (3), {[Hg(L)Cl].NaOH.2H2O} (3.NaOH.2H2O), and {[Hg3(HL)2Cl6].2H2O} (4) (L = -SCH2CH2NH2) were prepared and investigated by means of IR spectroscopy and single-crystal X-ray diffraction. The crystal structures of 1, 2, and 3.NaOH.2H2O show chelating N,S-coordination of the cysteaminate ligand, bridging S, and terminally coordinating Cl. Apart from these common features, the coordination geometries and modes of intermolecular association are different. 1 forms a cyclic tetramer with a Zn4S4 ring, and 3.NaOH.2H2O contains one-dimensional [Hg(L)Cl]n chains with S-bridged Hg atoms. Zn and Hg atoms in 1 and 3.NaOH.2H2O are tetracoordinate with a distorted tetrahedral M(ClNS2) geometry (M = Zn, Hg). Each Cd atom of 2 binds to three S atoms and vice versa, such that layers of distorted Cd3S3 hexagons are formed. 2 is the first example for a compound exhibiting a group 12-group 16 layer structure, which can be described as an analogue of a graphite layer. Additionally, each Cd atom binds to a chlorine atom and a nitrogen atom from a cysteaminate ligand resulting in pentacoordination with a distorted trigonal bipyramidal Cd(ClNS3) geometry. 4 contains two differently coordinate Hg atoms. One displays a distorted trans-octahedral Hg(Cl4S2) geometry, while the other is coordinated by four Cl atoms and one S atom and additionally forms a long Hg...Cl contact.

2.
Inorg Chem ; 44(22): 8087-96, 2005 Oct 31.
Article in English | MEDLINE | ID: mdl-16241159

ABSTRACT

2-Ammoniumethanethiolate, (-)SCH(2)CH(2)NH(3)(+), the first structurally characterized zwitterionic ammoniumthiolate, is the stable form of cysteamine (HL) in the solid state and in aqueous solution. Reactions of ZnCl(2), Cd(Oac)(2), and HgCl(2) with cysteamine and NaOH in a 1:2:2 ratio, respectively, lead to the homoleptic complexes ML(2). Their single-crystal X-ray structures demonstrate basic differences in the coordination chemistry of Zn(II), Cd(II), and Hg(II). While chelating N,S-coordination modes are found for all metal ions, Zn(II) forms a mononuclear complex with a distorted tetrahedral Zn(N(2)S(2)) coordination mode, whereas Hg(II) displays a dimer with Hg(N(2)S(2)) coordinated monomers being connected by two long Hg...S contacts. Solid-state (199)Hg NMR spectra of HgL(2) and [Hg(HL)(2)]Cl(2) reveal a low-field shift of the signals with increasing coordination number. Strong and nearly symmetric Cd-S-Cd bridges in solid CdL(2) lead to a chain structure, Cd(II) displaying a distorted square pyramidal Cd(N(2)S(3)) coordination mode. The ab initio [MP2/LANL2DZ(d,f)] structures of isolated ML(2) show a change from a distorted tetrahedral to bisphenoidal coordination mode in the sequence Zn(II)-Cd(II)-Hg(II). A natural bond orbital analysis showed a high ionic character for the M-S bonds and suggests that the S-M-S fragment is best described by a 3c4e bond. The strength of the M...N interactions and the stability of ML(2) toward decomposition to M and L-L decreases in the sequence Zn > Cd > Hg. Ab initio calculations further suggest that a tetrahedral S-M-S angle stabilizes Zn(II) against substitution by Cd(II) and Hg(II) in a M(N(2)S(2)) environment. Such geometry is provided in zinc-finger proteins, as was found by a database survey.

3.
Dalton Trans ; (19): 3221-8, 2005 Oct 07.
Article in English | MEDLINE | ID: mdl-16172648

ABSTRACT

The molecular structures of Se(SCH(3))(2) and Te(SCH(3))(2) were investigated using gas-phase electron diffraction (GED) and ab initio and DFT geometry optimisations. While parameters involving H atoms were refined using flexible restraints according to the SARACEN method, parameters that depended only on heavy atoms could be refined without restraints. The GED-determined geometric parameters (r(h1)) are: rSe-S 219.1(1), rS-C 183.2(1), rC-H 109.6(4) pm; angleS-Se-S 102.9(3), angleSe-S-C 100.6(2), angleS-C-H (mean) 107.4(5), phiS-Se-S-C 87.9(20), phiSe-S-C-H 178.8(19) degrees for Se(SCH(3))(2), and rTe-S 238.1(2), rS-C 184.1(3), rC-H 110.0(6) pm; angleS-Te-S 98.9(6), angleTe-S-C 99.7(4), angleS-C-H (mean) 109.2(9), phiS-Te-S-C 73.0(48), phiTe-S-C-H 180.1(19) degrees for Te(SCH(3))(2). Ab initio and DFT calculations were performed at the HF, MP2 and B3LYP levels, employing either full-electron basis sets [3-21G(d) or 6-31G(d)] or an effective core potential with a valence basis set [LanL2DZ(d)]. The best fit to the GED structures was achieved at the MP2 level. Differences between GED and MP2 results for rS-C and angleS-Te-S were explained by the thermal population of excited vibrational states under the experimental conditions. All theoretical models agreed that each compound exists as two stable conformers, one in which the methyl groups are on the same side (g(+)g(-) conformer) and one in which they are on different sides (g(+)g(+) conformer) of the S-Y-S plane (Y = Se, Te). The conformational composition under the experimental conditions could not be resolved from the GED data. Despite GED R-factors and ab initio and DFT energies favouring the g(+)g(+) conformer, it is likely that both conformers are present, for Se(SCH(3))(2) as well as for Te(SCH(3))(2).


Subject(s)
Selenium Compounds/chemistry , Sulfides/chemistry , Tellurium/chemistry , Electrons , Molecular Structure
4.
Dalton Trans ; (21): 3765-71, 2004 Nov 07.
Article in English | MEDLINE | ID: mdl-15510304

ABSTRACT

Selenium(II) dimethanethiolate, Se(SMe)(2), was synthesized by reaction of SeO(2) with HSMe. Basic spectroscopic data for Se(SMe)(2) and selenium(II) bis(2-methyl-2-propanethiolate), Se(S(t)Bu)(2), were recorded and interpreted with the support of ab initio calculations. Both compounds are thermodynamically unstable relatively to selenium and the corresponding disulfide. The UV/vis spectra of both compounds are qualitatively similar, the two bands being attributed to n(Se)-sigma*(Se-S) transitions. The bands at 369 and 397 cm(-1) in the IR spectra of Se(SMe)(2) and Se(S(t)Bu)(2), respectively, are assigned to nu(as)(SeS(2)). The (77)Se NMR shifts of Se(SMe)(2)(784 ppm) and Se(S(t)Bu)(2)(556 ppm) differ substantially from each other and show positive temperature gradients. Calculations at the GIAO-HF/962+(d) level reproduced the difference of the (77)Se NMR chemical shifts between Se(SMe)(2) and Se(S(t)Bu)(2). At the same level, the effect of conformational changes on (77)Se shifts were studied for Se(SMe)(2). In the solid state Se(SMe)(2) forms long intermolecular SeS contacts while Se(S(t)Bu)(2) does not. Both compounds exhibit anti-conformations of the methyl and tert-butyl groups with respect to the SeS(2) plane. MP2/LANL2DZ(d) geometry optimizations, single point energy and frequency calculations performed for Se(SMe)(2) show, that syn- (C(s)) and anti-conformers (C(2)) represent minima on the potential energy surface, the latter being by 8 kJ mol(-1) lower in energy than the former. Both conformers are stabilized by intramolecular pi-type n(S(1))-sigma*(Se-S(2)) orbital interactions. The energy of the transition state for the mutual conversion of the two conformers was calculated to be 31 kJ mol(-1) above that of the syn conformer, allowing a rapid interconversion of the two conformers at room temperature. Intermolecular interactions between Se(SMe)(2) molecules were also studied by means of calculations at the MP2/LANL2DZ(d) level. For Se(S(t)Bu)(2) MP2/LANL2DZ(d) geometry optimizations and single point energy calculations revealed a C(2)-symmetric anti- and a C(1) symmetric syn-conformer, the latter being 21 kJ mol(-1) higher in energy than the former. Se(SMe)(2) and Se(S(t)Bu)(2) exchange thiolate groups with other selenium(II) dithiolates, tellurium(II) dithiolates and with thiols, if catalytic amounts of p-CH(3)C(6)H(4)SO(3)H are added.

5.
Inorg Chem ; 43(18): 5529-36, 2004 Sep 06.
Article in English | MEDLINE | ID: mdl-15332803

ABSTRACT

The novel compounds PbCl(2).(SCH(2)CH(2)NH(3)) (1), Pb(SCH(2)CH(2)NH(2))(2).2PbCl(SCH(2)CH(2)NH(2)) (2), and Pb(SCH(2)CH(2)NH(2))(2) (3) were synthesized by reaction of PbO or PbCl(2) with [HSCH(2)CH(2)NH(3)]Cl and NaOH, and were characterized by elemental analysis, IR-, and UV/vis-spectroscopy. Single-crystal X-ray diffraction revealed different coordination modes for the two Pb atoms in 2. The Pb atom in the Pb(SCH(2)CH(2)NH(2))(2) unit forms two covalent Pb-S and two intramolecular dative Pb...N bonds, leading to a pseudo trigonal bipyramidal configuration with a stereochemically active lone pair. The Pb atom in the PbCl(SCH(2)CH(2)NH(2)) unit, the first moiety structurally characterized of the PbCl(SR) type (R = organic group), forms covalent Pb-Cl and Pb-S bonds, an intramolecular dative Pb...N bond, and two intermolecular Pb...S contacts, giving a pseudo octahedral configuration with a stereochemically active lone pair as well. Despite the Pb(SCH(2)CH(2)NH(2))(2) moiety exhibiting C(2) symmetry in 2, and C(1) symmetry in 3, its structural parameters are rather similar in the two compounds. The influence of the Pb...N bond on molecular structure and thermodynamic stability were estimated by means of quantum chemical ab initio methods. Although an analysis of the wave function in terms of natural bond orbitals (NBO) revealed that n(N) and n(p)(S) compete for the empty p-orbital of the Pb(II) atom, the sigma-type n(N)-6p(Pb) interaction is stronger than the pi-type n(p)(S)-6p(Pb) interaction and hence determines the conformation of the compounds.


Subject(s)
Cysteamine/chemistry , Cysteamine/chemical synthesis , Lead/chemistry , Models, Chemical , Molecular Structure
6.
Inorg Chem ; 41(18): 4739-47, 2002 Sep 09.
Article in English | MEDLINE | ID: mdl-12206699

ABSTRACT

New methods of preparing tellurium(II) dithiolates, Te(SR)(2), are presented. Te(SCH(2)CH(2)OAc)(2), 1, was made from Te(SCH(2)CH(2)OH)(2) by acetylation of the hydroxyl groups. Te(SCH(2)CH(2)SAc)(2), 2, [Te(SCH(2)CH(2)NH(3))(2)]Cl(2), 3, and Te(SC(6)H(4)(o-NH(2)))(2), 4, were synthesized by ligand exchange reactions of Te(S(t)Bu)(2) with 2 equiv of HSCH(2)CH(2)SAc, [HSCH(2)CH(2)NH(3)]Cl, and HSC(6)H(4)(o-NH(2)), respectively. Of all compounds, 4 exhibits the strongest thermal sensitivity toward decomposition and the largest low-field shift of the (125)Te NMR signal, two features that are attributed to weak Te.N interactions. The structural parameters of the CSTeSC unit exhibit very similar values for all four compounds, while the torsion angles of the side chains differ between the molecules, a feature rationalized by ab initio studies. In the solid state, different kinds of intermolecular aggregation and contacts to the Te atoms are present. 1 and 2 crystallize in the same space group (orthorhombic, Pbcn) and exhibit C(2) symmetric molecules, with two intermolecular Te.S contacts, leading to a trapezoidal coordination mode of the Te atoms. SCCE and C(S)CEC (with E = O, S) torsion angles represent the major differences between 1 and 2, which are attributed to their unlike intermolecular hydrogen bridges. In the solid state structure of 3, [Te(SCH(2)CH(2)NH(3))(2)](2+) cations and Cl(-) anions form a three-dimensional network via N-H...Cl and C-H...Cl hydrogen bonds (triclinic, P(-)1). Two neighboring [Te(SCH(2)CH(2)NH(3))(2)](2+) cations are linked via two Te...S contacts, and each Te atom forms one additional Te...Cl contact, resulting in a slightly distorted trapezoidal coordination mode. In the solid state structure of 4, adjacent molecules form Te...Te and Te...N contacts as well as hydrogen bridges. Two chemically different Te atoms are present, both of which are tetracoordinate with distorted sawhorse configurations. The absence of intramolecular Te...O, Te...S, or Te...N contacts in 1, 2, and 4, respectively, is attributed to the conformational rigidity of the CSTeS unit, where conformation ruling coordination is the case.

7.
Spectrochim Acta A Mol Biomol Spectrosc ; 58(5): 911-25, 2002 Mar 15.
Article in English | MEDLINE | ID: mdl-11942397

ABSTRACT

Quantum-chemical calculations at HF, MP2 and B3LYP levels with 6-31G* and 6-311G** basis sets are reported for disilylamine, NH(SiH3)2. The equilibrium structure is found to vary with both level and basis set, all but one of the structures exhibiting a small lack of planarity of the HNSi2 system. The barrier to inversion, however, is found to be very low, at most 38 cm(-1). Vibration frequencies and intensities are calculated. The frequencies are scaled, where possible, either using updated infrared data or with the aid of factors transferred from N(CH3)(SiH3)2. Unobserved frequencies due to the v(s)NSi2, deltaNSi2 and delta(perpendicular)NH modes are predicted near 610, 210 and 360 cm(-1), respectively. The lower silyl torsion lies below 40 cm(-1). The appearance of a single broad vSiH band in gas-phase samples of both NH(SiH3)2 and NH(SiH3)(SiD3) is suggestive of signal averaging due to internal rotation. The frequencies v(is)SiH, infrared intensities and Raman scattering activities of the bands due to an isolated SiH bond in an otherwise deuterated species are calculated and correlated with the torsional angle of this bond and with the Mulliken charge on the hydrogen atom. The strength of the bond is a minimum, and the infrared intensity and Raman scattering activity are maxima, when the bond direction is roughly orthogonal to the skeletal plane. A major part of the frequency and intensity variations is attributed to n(p)(N)-sigma*(Si-H)) hyperconjugation which, NBO calculations show, reaches a maximum for this conformation. However, systematic smaller variations are found for SiH bonds lying in the skeletal plane, which reflect the proximity of the other silyl group and only partly correlate with Mulliken charge. vSiH-vSiH interaction force constants, f', are calculated for pairs of SiH bonds in different silyl groups and compared with the corresponding dipole-dipole potential energy, the latter calculated using a classical treatment of the interaction between point dipoles arising from delta mu/delta r for the SiH bonds involved. The gradient of the correlation is very close to that expected from the theory, but a negative intercept indicates the presence of additional factors.


Subject(s)
Silicon Compounds/chemistry , Silicon/chemistry , Biophysical Phenomena , Biophysics , Models, Molecular , Quantum Theory , Spectrophotometry, Infrared , Spectrum Analysis, Raman
8.
J Med Chem ; 45(2): 382-9, 2002 Jan 17.
Article in English | MEDLINE | ID: mdl-11784142

ABSTRACT

2-Amino-3-benzoylthiophenes are allosteric enhancers (AE) of agonist activity at the A(1) adenosine receptor. The present report describes syntheses and assays of the AE activity at the human A(1)AR (hA(1)AR) of a panel of compounds consisting of nine 2-amino-3-aroylthiophenes (3a-i), eight 2-amino-3-benzoyl-4,5-dimethylthiophenes (12a-h), three 3-aroyl-2-carboxy-4,5-dimethylthiophenes (15a-c), 10 2-amino-3-benzoyl-5,6-dihydro-4H-cyclopenta[b]thiophenes (17a-j), 14 2-amino-3-benzoyl-4,5,6,7-tetrahydrobenzo[b]thiophenes (18a-n), and 15 2-amino-3-benzoyl-5,6,7,8-tetrahydro-4H-cyclohepta[b]thiophenes (19a-o). An in vitro assay employing the A(1)AR agonist [(125)I]ABA and membranes from CHO-K1 cells stably expressing the hA(1)AR measured, as an index of AE activity, the ability of a candidate AE to stabilize the agonist-A(1)AR-G protein ternary complex. Compounds 3a-i had little or no AE activity, and compounds 12a-h had only modest activity, evidence that AE activity depended absolutely on the presence of at least a methyl group at C-4 and C-5. Compounds 17a-c lacked AE activity, suggesting the 2-amino group is essential. Polymethylene bridges linked thiophene C-4 and C-5 of compounds 17a-j, 18a-n, and 19a-o. AE activity increased with the size of the -(CH(2))(n)()- bridge, n = 3 < n = 4 < n = 5. The 3-carbethoxy substituents of 17a, 18a, and 19a did not support AE activity, but a 3-aroyl group did. Bulky (or hydrophobic) substituents at the meta and para positions of the 3-benzoyl group and also 3-naphthoyl groups greatly enhanced activity. Thus, the hA(1)AR contains an allosteric binding site able to accommodate 3-aroyl substituents that are bulky and/or hydrophobic but not necessarily planar. A second region in the allosteric binding site interacts constructively with alkyl substituents at thiophene C-4 and/or C-5.


Subject(s)
Purinergic P1 Receptor Agonists , Thiophenes/pharmacology , Allosteric Regulation , Animals , CHO Cells , Cricetinae , Humans , Hydrophobic and Hydrophilic Interactions , Quantitative Structure-Activity Relationship , Thiophenes/chemistry
9.
Inorg Chem ; 38(16): 3725-3729, 1999 Aug 09.
Article in English | MEDLINE | ID: mdl-11671134

ABSTRACT

The reaction between Te(O(i)Pr)(4) and HSR offers a new and effective route to tellurium dithiolates, Te(SR)(2). Te(S(i)Pr)(2) (1) and Te(S(t)Bu)(2) (2) are stable compounds whereas Te(SPh)(2) (3) slowly decomposes at room temperature to give Te and Ph(2)S(2). IR spectra of 1-3 and ab initio calculations (HF/3-21G(d) and MP2 with double-zeta polarization effective core potential basis set) show nu(as)(Te-S) and nu(s)(Te-S) to be around 340 and 380 cm(-)(1), respectively. UV spectra exhibit similar lambda(max) (346-348 nm) for all three compounds, with the greater extinction coefficient of 3 accounting for its different and more intense color. Analysis of the molecular orbitals of the model compound Te(SCH(3))(2) shows that the phototransition is likely to be of n(p)(Te)-sigma(Te-S) type, thus rationalizing the instability of 3 when irradiated. Single-crystal X-ray diffraction of 1-3 revealed the following basic structural parameters: 1 d(av)(Te-S) 239.4(1) and d(av)(S-C) 183.8(5) pm, angle(STeS) 99.61(4) and angle(av)(TeSC) 105.8(3) degrees, tau(CSTeS) 77.0(2) and 90.3(2) degrees; 2 d(Te-S) 239.1(1) and d(S-C) 186.4(2) pm, angle(STeS) 103.88(2) and angle(TeSC) 107.6(1) degrees, tau(CSTeS) 78.01(8) degrees; 3 d(Te-S) 240.6(2) and d(S-C) 177.4(7) pm, angle(STeS) 100.12(6) angle(TeSC) 103.2(2) degrees, tau(CSTeS) 69.0(3) and tau(CCSTe) 81.6(6) degrees. Geometries of model compounds Te(SH)(2) and Te(SCH(3))(2) optimized at the MP2 level exhibit d(Te-S), angle(STeS), and tau(XSTeS) (X = H, C) values similar to those of 1-3. Natural bond orbital analysis revealed n(p)(S(1))-sigma(Te-S(2)) hyperconjugation as the cause for the CSTeS torsion angles being close to 90 or -90 degrees. Thermochemical calculations on the HF and MP2 level proved Te(SH)(4) to be unstable with respect to Te(SH)(2) and HSSH, thus rationalizing the reduction of Te(IV) to Te(II) when Te(O(i)Pr)(4) or TeO(2) are reacted with thiols. NMR spectra reveal ligand exchange reactions between different tellurium(II) dithiolates and between Te(SR)(2) and HSR'. These types of reaction offer other routes to tellurium(II) dithiolates.

10.
Inorg Chem ; 38(21): 4700-4704, 1999 Oct 18.
Article in English | MEDLINE | ID: mdl-11671194

ABSTRACT

The reaction between LiAlH(4) and pyridine, 4-methylpyridine, or 3,5-dimethylpyridine results in hydride transfer to the pyridine ring to give tetrakis(pyridine)lithium tetrakis(1,4-dihydropyrid-1-yl)aluminate(III), 1, tetrakis(4-methylpyridine)lithium tetrakis(1,4-dihydro-4-methylpyrid-1-yl)aluminate(III), 2, or tetrakis(3,5-dimethylpyridine)lithium tetrakis(3,5-dimethyl-1,4-dihydropyrid-1-yl)aluminate(III), 3, respectively. We claim that 1, instead of lithium tetrakis(1,4-dihydropyrid-1-yl)aluminate(III), is the compound which is known as Lansbury's reagent. Treatment of trimethylamine-alane, AlH(3).NMe(3), with pyridine yields tris(1,4-dihydropyrid-1-yl)(pyridine)aluminum, 4. It could be shown that AlH(3).NMe(3) initially reduces pyridine to 1,2-dihydropyridine, which is subsequently converted into its 1,4-isomer. The X-ray crystal structures of 1-4 were determined. While the differences between Al-N distances within each of the compounds 1-3 are not significant, 4 exhibits two distinctly different types of Al-N bonds, the dative bond between Al and N(pyridine), d(Al-N) = 1.959(2) Å, and the covalent bonds between Al and N(1,4-dihydropyrid-1-yl), d(av)(Al-N) = 1.833 Å.

SELECTION OF CITATIONS
SEARCH DETAIL
...