首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 546 毫秒
1.
 A crystal structure determination of a bilirubin analog with a sulfur instead of a C(10)–CH2 linking the two dipyrrinones is reported. Conformation-determining torsion angles and key hydrogen bond distances and angles are compared to those obtained from molecular dynamics calculations as well as to the corresponding data from X-ray determinations and molecular dynamics calculations of bilirubin. Like other bilirubins, the component dipyrrinones of the analog are present in the bis-lactam form with (Z)-configurated double bonds at C(4) and C(15). Despite the large differences in bond lengths and angles at –S–vs.–CH2–, the crystal structure shows considerable similarity to bilirubin: both pigments adopt a folded, intramolecularly hydrogen-bonded ridge-tile conformation stabilized by six hydrogen bonds – although the interplanar angle of the ridge-tile conformation of the title compound is smaller (∼ 86°) than that of bilirubin (∼ 98°). The collective data indicate that even with long C–S bond lengths and a smaller C–S–C bond angle at the pivot point on the ridge-tile seam, intramolecular hydrogen bonding persists.  相似文献   

2.
Summary. Crystal structures determined for three bilirubin analogs with gem-dimethyl groups at C(10) are reported, including the first X-ray structure of a bilirubin dimethyl ester. Conformation-determining torsion angles and key hydrogen bond distances and angles were compared to those from molecular dynamics calculations. Like other rubins, the component dipyrrinones of the three compounds were found to adopt the syn conformation, with Z-configuration double bonds at C(4) and C(15) and bis-lactam tautomeric structures of the end rings. No large differences in bond lengths and bond angles at C(10) were found, and the crystal structures of the two 10,10-dimethyl rubin acids showed considerable similarity to that of bilirubin: both pigments adopt a folded, intramolecularly hydrogen bonded ridge-tile conformation stabilized by six hydrogen bonds, with an interplanar angle in ridge-tile of 98° and 86°. In contrast, the dimethyl ester is intermolecularly hydrogen bonded in the crystal. Each molecule of the ester has its two syn-Z-dipyrrinones rotated into a conformation syn to the gem-dimethyl group, whereas in the acids they are anti.  相似文献   

3.
Summary. A symmetrical C(10)-selena-bilirubin analog, 8,12-bis(2-carboxyethyl)-7,13-dimethyl-2,3,17,18-tetraethyl-10-selenabiladiene-ac-1,19(21H,24H)-dione was synthesized from 8-(2-carboxyethyl)-2,3-diethyl-7-methyl-(10H)-dipyrrin-1-one in one step by reaction with diselenyl dichloride. The selena-rubin exhibited UV-vis and NMR spectroscopic properties similar to those of the parent mesobilirubin, and like bilirubin and mesobilirubin, it adopts an intramolecularly hydrogen-bonded conformation, shaped like a ridge-tile but with a steeper pitch. The longer C–Se bond lengths (2.2Å) and smaller bond angles at C–Se–C (88°), as compared to C–CH2–C (1.5Å, 106°), lead to an interplanar angle between the two dipyrrinones of only 72°, which is considerably less than that of bilirubin (100°) and close to that (74°) of its 10-thia-rubin analog. Despite the conformational distortion, the sensitivity of Se toward oxidation and the typically weak C–Se bond, the selena-rubin is metabolized in normal rats, like bilirubin, to acyl glucuronides, which are secreted into bile. In mutant (Gunn) rats lacking bilirubin glucuronosyl transferase (UGT1A1), glucuronide or other metabolites of the selena-rubin were not detected in bile, demonstrating the importance of hepatic glucuronidation for its biliary excretion.  相似文献   

4.
Summary. Crystal structure determinations of two novel bilirubin analogs with spirocyclohexyl and spirofluorenyl groups at C(10) are reported. Conformation-determining torsion angles and key hydrogen bond distances and angles are compared to those from molecular dynamics calculations, and to the corresponding data from X-ray determinations and molecular dynamics calculations of bilirubin. Like bilirubin, the component dipyrrinones of spirocyclohexyl and spirofluorenyl rubins are present in the bis-lactam form with (Z)-configuration double bonds at C(4) and C(15). Their crystal structures show considerable similarity to that of bilirubin: both pigments adopt a folded, intramolecularly hydrogen-bonded ridge-tile conformation stabilized by six hydrogen bonds. The interplanar angle of the spirocyclohexyl ridge-tile is nearly the same (94°) as that of bilirubin (95°), but the interplanar angle of the spirofluorenyl ridge-tile is noticeably smaller (84°). The hydrogen bond distances of both spiro-rubin crystal structures are generally longer by 0.1–0.2Å than those in bilirubin. Both new pigments exhibit excellent lipophilicity and, unlike bilirubin, are soluble in methanol.  相似文献   

5.
A symmetrical C(10)-selena-bilirubin analog, 8,12-bis(2-carboxyethyl)-7,13-dimethyl-2,3,17,18-tetraethyl-10-selenabiladiene-ac-1,19(21H,24H)-dione was synthesized from 8-(2-carboxyethyl)-2,3-diethyl-7-methyl-(10H)-dipyrrin-1-one in one step by reaction with diselenyl dichloride. The selena-rubin exhibited UV-vis and NMR spectroscopic properties similar to those of the parent mesobilirubin, and like bilirubin and mesobilirubin, it adopts an intramolecularly hydrogen-bonded conformation, shaped like a ridge-tile but with a steeper pitch. The longer C–Se bond lengths (2.2Å) and smaller bond angles at C–Se–C (88°), as compared to C–CH2–C (1.5Å, 106°), lead to an interplanar angle between the two dipyrrinones of only 72°, which is considerably less than that of bilirubin (100°) and close to that (74°) of its 10-thia-rubin analog. Despite the conformational distortion, the sensitivity of Se toward oxidation and the typically weak C–Se bond, the selena-rubin is metabolized in normal rats, like bilirubin, to acyl glucuronides, which are secreted into bile. In mutant (Gunn) rats lacking bilirubin glucuronosyl transferase (UGT1A1), glucuronide or other metabolites of the selena-rubin were not detected in bile, demonstrating the importance of hepatic glucuronidation for its biliary excretion.  相似文献   

6.
Summary. Bilirubin congeners with dipyrrinones conjoined to a diaceteylene unit (–CC–CC–) rather than to –CH2– were synthesized and examined spectroscopically. This new class of rubrified linear tetrapyrroles cannot easily fold or bend in the middle, but the dipyrrinones can rotate independently about the diacetylene unit. Thus, unlike bilirubin, which is bent in the middle and has a ridge-tile shape, the diacetylene unit orients the attached dipyrrinones along a linear path, and intramolecular hydrogen bonding between the dipyrrinones and opposing carboxylic acids preserves a twisted linear molecular shape when the usual propionic acids are replaced by hexanoic. In a bis-hexanoic acid rubin, the extended planes of the dipyrrinones intersect along the –(CC)2– axis at an angle of 102° for the conformation stabilized by intramolecular hydrogen bonding. With propionic acid chains, however, neither CO2H can engage an opposing dipyrrinone in intramolecular hydrogen bonding, and the energy-minimum conformation of this linear pigment, shows nearly co-planar dipyrrinones, with an intersection of an angle of 180° of the extended planes of the dipyrrinones. Spectroscopic evidence for such linearized and twisted (bis-hexanoic) and planar (bis-propionic) structures comes from the pigments NMR spectral data and their exciton UV-Vis and induced circular dichroism spectra.  相似文献   

7.
Summary.  Malonic ester derivatives of ethyl and methyl 3,5-dimethyl-4-(1′-iodoneopentyl)-1H-pyrrole-2-carboxylate exhibit restricted rotation about the pyrrole C(4)–C(1′) bond due to the bulky 1′-tert-butyl and malonic ester groups and the ortho effect at C(4) of the sterically crowded 3,5-dimethylpyrrole. The malonates belong to a rare class of atropisomers with restricted rotation about an sp3–sp2 C–C bond, and they undergo diastereomeric separation by TLC and crystallization: the diastereomers are stable in solution at room temperature. A crystal of one of the diastereomers, suitable for X-ray crystallography, gave the relative configuration of the chiral axis and stereogenic center at C(1′). Dynamic NMR studies of the purified diastereomers provide kinetic and thermodynamic parameters associated with the atropisomerism: ΔG  = 132–134 kJ/mol (∼32 kcal/mol) at 383 K in C2D2Cl4 solvent. Corresponding author. E-mail: lightner@scs.unr.edu Received July 1, 2002; accepted July 8, 2002  相似文献   

8.
 Malonic ester derivatives of ethyl and methyl 3,5-dimethyl-4-(1′-iodoneopentyl)-1H-pyrrole-2-carboxylate exhibit restricted rotation about the pyrrole C(4)–C(1′) bond due to the bulky 1′-tert-butyl and malonic ester groups and the ortho effect at C(4) of the sterically crowded 3,5-dimethylpyrrole. The malonates belong to a rare class of atropisomers with restricted rotation about an sp3–sp2 C–C bond, and they undergo diastereomeric separation by TLC and crystallization: the diastereomers are stable in solution at room temperature. A crystal of one of the diastereomers, suitable for X-ray crystallography, gave the relative configuration of the chiral axis and stereogenic center at C(1′). Dynamic NMR studies of the purified diastereomers provide kinetic and thermodynamic parameters associated with the atropisomerism: ΔG  = 132–134 kJ/mol (∼32 kcal/mol) at 383 K in C2D2Cl4 solvent.  相似文献   

9.
Novel bilirubin analogues with dipyrrinones conjoined to an acetylene rather than a methylene group were synthesized and examined spectroscopically. Despite the increased separation of the dipyrrinones forced by replacing a -CH(2)- by a -C(triple bond)C- unit, molecular dynamics calculations show that, like bilirubin, they may still engage in intramolecular hydrogen bonding to carboxylic acid groups when the propionic acid chains are slightly lengthened, e.g., butanoic acids. Unlike bilirubin, however, which is bent in the middle and has a ridge-tile shape, the acetylene orients the attached dipyrrinones along a linear path, and intramolecular hydrogen bonding preserves a twisted linear molecular shape. The extended planes of the dipyrrinones intersect along the -C(triple bond)C- axis at an angle of 136 degrees for the conformation stabilized by intramolecular hydrogen bonding in the bis-butyric acid rubin (1b). With shorter acid chains (propionic), only one CO(2)H can engage an opposing dipyrrinone in intramolecular hydrogen bonding, and in this energy-minimum conformation of the linear pigment 1a, the intersection of the extended planes of the dipyrrinones has an angle of 171 degrees. Spectroscopic evidence for such linearized and twisted structures was found in the pigments' NMR spectral data and their exciton UV-vis and induced circular dichroism spectra.  相似文献   

10.
A combined gas-phase electron diffraction and quantum chemical (B3LYP/6-311+G**, B3LYP/cc-pvtz, MP2/cc-pvtz) study of molecular structure of 2-nitrobenzenesulfonamide (2-NBSA) was carried out. Quantum chemical calculations showed that 2-NBSA has four conformers, two of which are stabilized by intramolecular hydrogen bond. The latter (with the S–N bond in a close to orthogonal position around the phenyl ring and differing from each other by staggered or eclipsed positions of the N–H and S=O bonds in the SO2NH2 group) presented in a saturated vapor over 2-NBSA at T = 433 (3) K in commensurable amounts. Experimental internuclear distances (Ǻ) for the staggered conformer are (?): r h1(C–H)av. = 1.071(9), r h1(C–C)av. = 1.390(4), r h1(C–S) = 1.789(8), r h1(S=O)av. = 1.427(6), r h1(S–N) = 1.644(6), r h1(N–O)av. = 1.221(4), r h1(C′–N) = 1.487(8), r h1(N–H)av. = 1.014. Calculations at B3LYP/cc-pvtz level were performed to determine the structure and the energies of the transition states between conformers. It was shown that the conformer structures of free molecule differ from those of a molecule stabilized by intermolecular hydrogen bonds in a crystal. Influence of a substituent X (X = –CH3, –NO2) on conformational features of the ortho-substituted benzenesulfonamide was established.  相似文献   

11.
Yellow N,N′-carbonyl-bridged dipyrrinones can generally be prepared from dipyrrinones simply by reaction with N,N′-carbonyldiimidazole in the presence of a strong, non-nucleophilic base. They are typically intensely fluorescent, with fluorescent quantum yields approaching 1.0. In an effort to shift the excitation wavelength, and thus the fluorescence emissions, strongly to the red, we prepared bridged dipyrrinones conjugated with thiobarbituric acid and Meldrum’s acid substituents at C-9. Such conjugation causes the dipyrrinones to have a magenta color (absorption wavelength shifted from ∼400 nm of a typical dipyrrinone to ∼550 nm of the dipyrrinone conjugate). For comparison, we also prepared analogs with formyl, carboxyl, acrylate, and acetyl substituents at C-9. Unexpectedly and uniquely, the 9-CHO substituent caused the fluorescence quantum yield to drop to ∼10−3 while carboethoxy substituent exerted only a minor influence. Correspondence: David A. Lightner, Department of Chemistry, University of Nevada, Reno, 89557 Nevada, USA.  相似文献   

12.
The conformational composition of gaseous MTMNB and the molecular structures of the rotational forms have been studied by electron diffraction at 130C aided by results from ab initio and density functional theory calculations. The conformational potential energy surface has been investigated by using the B3LYP/6-31G(d,p) method. As a result, six minimum-energy conformers have been identified. Geometries of all conformers were optimized using MP2/6-31G(d,p), B3LYP/6-31G(d,p), and B3LYP/cc-pVTZ methods. These calculations resulted in accurate geometries, relative energies, and harmonic vibrational frequencies for all conformers. The B3LYP/cc-pVTZ energies were then used to calculate the Boltzmann distribution of conformers. The best fit of the electron diffraction data to calculated values was obtained for the six conformer model, in agreement with the theoretical predictions. Average parameter values (ra in angstroms, angle α in degrees, and estimated total errors given in parentheses) weighted for the mixture of six conformers are r(C–C) = 1.507(5), r(C–C)ring, av = 1.397(3), r(C–S)av = 1.814(4), r(C–N) = 1.495(4), r(N–O)av = 1.223(3), ∠(C–C–C)ring = 116.0–122.5, ∠ C6–C4–C7 = 118.2(4), ∠ C–C–S = 113.6(6), ∠ C–S–C = 98.5(12), ∠ N–C–C4 = 121.9(3), ∠(O–N–C)av = 116.8(3), ∠ O–N–O = 127.0(4). Torsional angles could not be refined. Theoretical B3LYP/cc-pVTZ torsional angles for the rotation about C–N bond, φCN, were found to be 30.5–36.5 for different conformers. As to internal rotation about C–C and C–S bonds, values of φCC = 68–118 and φCS = 66–71 were obtained for the three most stable conformers with gauche orientation with respect to these bonds. Some conclusions of this work were presented in a short communication in Russ. J. Phys. Chem. 2005, 79, 1701.  相似文献   

13.
The existence of a short C–H ⋯ π (alkyl–alkyne) interaction in the structure of a strained and relatively rigid tolanophane is expected to hinder the rotation about the C–C sp3 single bond. Variable-temperature NMR experiments (performed in three solvents, CDCl3, THF-d8, and acetone-d6) and ab initio density functional calculations were carried out to investigate its dynamic nature. An energy barrier of 48.6 kJ/mol is determined at coalescence (210 K) with acetone-d6 which is in good agreement with calculation result (54 kJ/mol). Correspondence: Hossein Reza Darabi, Chemistry and Chemical Engineering Research Center of Iran, Pajoohesh Blvd., km 17, Karaj Hwy, 14968-13151 Tehran, Iran.  相似文献   

14.
The molecular structure of triphenylsilane has been investigated by gas-phase electron diffraction and theoretical calculations. The electron diffraction intensities from a previous study (Rozsondai B, Hargittai I, J Organomet Chem 334:269, 1987) have been reanalyzed using geometrical constraints and initial values of vibrational amplitudes from calculations. The free molecule has a chiral, propeller-like equilibrium conformation of C 3 symmetry, with a twist angle of the phenyl groups τ = 39° ± 3°; the two enantiomeric conformers easily interconvert via three possible pathways. The low-frequency vibrational modes indicate that the three phenyl groups undergo large-amplitude torsional and out-of-plane bending vibrations about their respective Si–C bonds. Least-squares refinement of a model accounting for the bending vibrations gives the following bond distances and angles with estimated total errors: r g(Si–C) = 1.874 ± 0.004 ?, 〈r g(C–C)〉 = 1.402 ± 0.003 ?, 〈r g(C–H)〉 = 1.102 ± 0.003 ?, and ∠aC–Si–H = 108.6° ± 0.4°. Electron diffraction studies and MO calculations show that the lengths of the Si–C bonds in H4−n SiPh n molecules (n = 1–4) increase gradually with n, due to π → σ*(Si–C) delocalization. They also show that the mean lengths of the ring C–C bonds are about 0.003 ? larger than in unsubstituted benzene, due to a one hundredth angstrom lengthening of the Cipso–Cortho bonds caused by silicon substitution. A small increase of r(Si–H) and decrease of the ipso angle with increasing number of phenyl groups is also revealed by the calculations.  相似文献   

15.
Summary.  The synthesis, crystal structure determination, conformational analysis, and spectroscopic properties of 3,3′-diethyl-4,4′-dimethyl-2,2′-dipyrryl ketone (1) are reported. The dipyrryl ketone is a model for the dipyrrole core of 10-oxobilirubin, a presumed metabolite in alternate pathways of excretion of the yellow pigment of jaundice, bilirubin. In the crystal, 1 adopts a helical conformation, with a molecule of one helicity being hydrogen-bonded to two molecules of the opposite helicity. Thus, 1 self-assembles via hydrogen bonding into supramolecular double-stranded arrays, where molecules of the same helicity comprise one strand and are paired through hydrogen bonding to molecules of opposite helicity in the second strand. In the observed molecular conformation each pyrrole ring and adjacent carbonyl group are rotated into an sc conformation (torsion angle ∼29 °), with each N-H pointing in the same direction as the C*O. Molecular mechanics/dynamics calculations predict the sc,sc conformation, absent hydrogen bonding, to be the most stable, but only by a few tenths of a kj/mol. In CHCl3, 1 is monomeric according to vapor pressure osmometry studies (). 1H NMR NH chemical shifts in CDCl3 suggest a predominantly anti orientation of the C=O and pyrrole NHs, which is opposite to the orientation observed in the crystal. Received February 4, 2000. Accepted February 14, 2000  相似文献   

16.
Yellow N,N′-carbonyl-bridged dipyrrinones can generally be prepared from dipyrrinones simply by reaction with N,N′-carbonyldiimidazole in the presence of a strong, non-nucleophilic base. They are typically intensely fluorescent, with fluorescent quantum yields approaching 1.0. In an effort to shift the excitation wavelength, and thus the fluorescence emissions, strongly to the red, we prepared bridged dipyrrinones conjugated with thiobarbituric acid and Meldrum’s acid substituents at C-9. Such conjugation causes the dipyrrinones to have a magenta color (absorption wavelength shifted from ∼400 nm of a typical dipyrrinone to ∼550 nm of the dipyrrinone conjugate). For comparison, we also prepared analogs with formyl, carboxyl, acrylate, and acetyl substituents at C-9. Unexpectedly and uniquely, the 9-CHO substituent caused the fluorescence quantum yield to drop to ∼10−3 while carboethoxy substituent exerted only a minor influence.  相似文献   

17.
Geometrical parameters, harmonic vibrational frequencies, atomic charge distributions, bonding character, and relative stability of the CH3NgF (Ng = He, Ar, Kr, or Xe) species were investigated at the MP2 level of theory. CH3HeF was also predicted stable at the CCSD(T) level. All the four CH3NgF species have C 3v symmetry. Ng–F bond lengths of the CH3NgF species are all longer than those of the corresponding HNgF species. The calculated infrared intensities of the C–Ng and Ng–F stretching vibrations are much larger than those of the other vibrations, which is advantageous for the experimental spectroscopic identification of the species. The atoms in molecules (AIM) topological analysis indicated that the three Ng–F (Ng = He, Ar, or Kr) bonds are dominated by electrostatic interaction whereas the two C–Ng (Ng = Ar or Kr) bonds are dominated by covalent interaction. In contrast, the bond length analysis seems to indicate that both the Ng–F and C–Ng bonds are dominated by covalent interaction. According to the MP2 calculations, CH3HeF and CH3ArF are higher in energy than the dissociation limits CH3 + He + F and CH3 + Ar + F by 15.10 and 2.64 kcal/mol whereas CH3KrF and CH3XeF are lower in energy than CH3 + Kr + F and CH3 + Xe + F by 16.80 and 38.44 kcal/mol, respectively.  相似文献   

18.
Based on the continuum dielectric model, this work has established the relationship between the solvent reorganization energy of electron transfer (ET) and the equilibrium solvation free energy. The dipole-reaction field interaction model has been proposed to describe the electrostatic solute-solvent interaction. The self-consistent reaction field (SCRF) approach has been applied to the calculation of the solvent reorganization energy in self-exchange reactions. A series of redox couples, O2/O 2, NO/NO+, O3/O 3, N3/N 3, NO2/NO+ 2, CO2/CO 2, SO2/SO 2, and ClO2/ClO 2, as well as (CH2)2C-(-CH2-) n -C(CH2)2 (n=1 ∼ 3) model systems have been investigated using ab initio calculation. For these ET systems, solvent reorganization energies have been estimated. Comparisons between our single-sphere approximation and the Marcus two-sphere model have also been made. For the inner reorganization energies of inorganic redox couples, errors are found not larger than 15% when comparing our SCRF results with those obtained from the experimental estimation. While for the (CH2)2C–(–CH2–) n –C(CH2)2 (n=1 ∼ 3) systems, the results reveal that the solvent reorganization energy strongly depends on the bridge length due to the variation of the dipole moment of the ionic solute, and that solvent reorganization energies for different systems lead to slightly different two-sphere radii. Received: 19 April 2000 / Accepted: 6 July 2000 / Published online: 27 September 2000  相似文献   

19.
A new bilirubinoid analog (1) with two methoxy beta-substituents on the lactam ring of each dipyrrinone was synthesized and examined spectroscopically. It is more soluble in CH3OH and CHCl3 than bilirubin, which is insoluble in CH3OH but soluble in CHCl3. The solubility of 1 is approximately 10 microg/mL in CH3OH (vs < or =1 microg/mL for bilirubin) and approximately 3 mg/mL in CHCl3 (vs approximately 0.6 mg/mL for bilirubin). Vapor pressure osmometry indicates that 1, like bilirubin, is monomeric in CHCl3, and NMR studies show that the most stable structure has the syn-4Z,syn-15Z configuration, with the pigment's dipyrrinones engaged in intramolecular hydrogen bonding to the propionic acid carboxyl groups. And, like bilirubin, Z,Z-1 adopts a conformation that is bent in the middle into a ridge-tile shape. For the first time, a crystal structure of a bilirubin E-isomer has been obtained. Crystallization of 1 under dim room lighting gave an X-ray quality crystal of the anti-4E,syn-15Z-(photo) isomer, in which only the Z-dipyrrinone half is engaged in intramolecular hydrogen bonding to a propionic acid. Hydrogen bonding is nearly completely disengaged in the E-dipyrrinone half; yet, the ridge-tile conformation persists.  相似文献   

20.
The reaction of [PtMe3(bpy)(Me2CO)](BF4) (2) (prepared from [PtMe3I(bpy)] (1) plus Ag(BF4)) with MeSSMe resulted in the formation of [PtMe3(bpy)(MeSSMe-κS)](BF4) (3). A single-crystal X-ray diffraction analysis revealed in the octahedral Pt(IV) complex (configuration index: OC-6-33), a conformation of the monodentately κS bound MeSSMe ligand (C–S–S–C 92.7(4)°) being very close to that in non-coordinated MeSSMe, thus allowing some hyperconjugative interaction stabilizing the S–S bond. The reaction of [K(18C6)][(PtMe3)2(μ-I)(μ-pz)2] (4; 18C6 = 18-crown-6, Hpz = pyrazole) with Ag(BF4) and MeSSMe resulted in the formation of dinuclear complexes [(PtMe3)2(μ-pz)2(μ-MeSSMe)] existing at room temperature in acetone solution as different fast interconverting isomers. At –40 °C, two isomers with a μ-1κS:2κS (5a) and a μ-1κS:2κS′ (5b) coordinated MeSSMe ligand in the ratio 2:1 could be identified 1H NMR spectroscopically. DFT calculations of type 5 complexes revealed the existence of two conformers with a μ-MeSSMe-1κS:2κS ligand, which differ mainly in the C–S–S–C dihedral angle (66.4 vs. 180.0° 6a/6a′). They have essentially the same energy and a very low activation barrier in acetone as solvent (1.3 kcal/mol) for their mutual interconversion. A further equilibrium structure was identified to be an isomer having a μ-MeSSMe-1κS:2κS′ ligand (6b) that proved to be only 1.9 kcal/mol higher in energy than 6a/6a′.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号