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 31 毫秒
1.
The switch from HbA (α2β2A) to HbC (α2β2C) synthesis was induced by injection of erythropoietin into a lamb homozygous for HbA. Serial samples of bone marrow were analyzed to detect the initial commitment of erythroid stem cells (CFU-E) to form colonies which made HbC in vitro, and to detect the initial accumulation of βC-globin mRNA and the onset of HbC synthesis in erythroblasts in vivo. CFU-E-derived erythroid colonies were formed in plasma clot culture at a low erythropoietin concentration, and the relative amounts of βA- and βC-globin synthesized were determined after a 24 hr pulse of 3H-leucine, added after 84 hr in culture. RNA was extracted from nuclei and cytoplasm of “early” and “late” populations of bone marrow erythroblasts which had been fractionated by Ficoll-Hypaque density centrifugation. The concentration of βA- and βC-globin mRNA was determined by annealing to purified synthetic DNAs (cDNAs) complementary to βA and βC mRNA. No βC-globin was synthesized in erythroblasts or in CFU-E-derived erythroid colonies prior to the injection of erythropoietin. An increase in the concentration of CFU-E in the bone marrow and the appearance of βC-globin synthesis in CFU-E-derived colonies were detected 12 hr after the erythropoietin injection. In contrast, βC mRNA was not detected in either “early” or “late” erythroid cells until 36 hr later. The first measurable βC-globin mRNA was accompanied by the appearance of βC-globin synthesis in bone marrow erythroblasts. Our results suggest that the accumulation of βC-globin mRNA is a relatively late event following induction of HbA to HbC switching by erythropoietin. The expansion of the compartment of erythroid stem cells and the commitment of CFU-E to βC-globin synthesis appear to precede the detectable accumulation of βC mRNA by 24–36 hr.  相似文献   

2.
Circulating schistosome antigens (CSA) and circulating immune complexes (CIC) were investigated during the course of Schistosoma mansoni infection in mice. The radioimmunoprecipitation-polyethylene glycol (PEG) assay (RIPEGA) with [125I]anti-S. mansoni antibodies or [125I] anti-antigen “4” antibodies detected, respectively, total CSA and antigen “4” in serum and in 3% polyethylene glycol-precipitated CIC from infected mice. Complement fixation test and [125I] C1q-binding test revealed, respectively, an anticomplementary activity and the presence of C1q-binding CIC. All these substances appeared in infected mice at approximately the same period, i.e., between the 40th- and the 55th- day postinfection. No correlation was observed between the detection of anticomplementary active substances and C1q-binding CIC. In contrast, a close relationship was noticed between CSA and complement-activating material during the course of the infection. This suggests that substances with anticomplementary activity in serum from infected mice could be one or various CSA. A close correlation was also observed between C1q-binding CIC and free or “complexed” antigen “4.” This observation supports well the possibility that antigen “4” is one of the major complexed circulating antigen present in schistosomiasis. The immunoglobulins G1, G2a, M, and A were also characterized in 3% PEG-precipitated CIC from infected mice during the period in which we detected C1q-binding CIC. The roles played by specific S. mansoni CIC in either schistosomal nephropathy or protective mechanisms to a challenge infection in mice are discussed.  相似文献   

3.
The electric dipole transitions between pure spin and mixed spin electronic states are calculated at the XMC-QDPT2 and MCSCF levels of theory, respectively, for different intermolecular distances of the C6H6 and O2 collisional complex. The magnetic dipole transition moment between the mixed-spin ground (“triplet”) and the first excited (“singlet”) states is calculated by quadratic response at MCSCF level of theory. The obtained results confirm the theory of intensity borrowing and increasing the intensity of electronic transitions in the C6H6?+?O2 collision. The calculation of magnetically induced current density is performed for benzene molecule being in contact with O2 at the distances from 3.5 to 4.5 Å. The calculation shows that the aromaticity of benzene is rising due to the conjugation of π-MOs of both molecules. The C6H6?+?O2 complex becomes nonaromatic at the short distances (r?<?3.5 Å). The computation of static polarizability in the excited electronic states of the C6H6?+?O2 collisional complex at various distances supports the theory of red solvatochromic shift of the a?→?X band.
Graphical abstract The C6H6+ O2 collisional complex
  相似文献   

4.
Conformational preferences of modified nucleoside, N(4)-acetylcytidine, ac4C have been investigated using quantum chemical semi-empirical RM1 method. Automated geometry optimization using PM3 method along with ab initio methods HF SCF (6-31G**), and density functional theory (DFT; B3LYP/6-31G**) have also been made to compare the salient features. The most stable conformation of N(4)-acetyl group of ac4C prefers “proximal” orientation. This conformation is stabilized by intramolecular hydrogen bonding between O(7)···HC(5), O(2)···HC2′, and O4′···HC(6). The “proximal” conformation of N(4)-acetyl group has also been observed in another conformational study of anticodon loop of E. coli elongator tRNAMet. The solvent accessible surface area (SASA) calculations revealed the role of ac4C in anticodon loop. The explicit molecular dynamics simulation study also shows the “proximal” orientation of N(4)-acetyl group. The predicted “proximal” conformation would allow ac4C to interact with third base of codon AUG/AUA whereas the ‘distal’ orientation of N(4)-acetyl cytidine side-chain prevents such interactions. Single point energy calculation studies of various models of anticodon–codon bases revealed that the models ac4C(34)(Proximal):G3, and ac4C(34)(Proximal):A3 are energetically more stable as compared to models ac4C(34)(Distal):G3, and ac4C(34)(Distal):A3, respectively. MEPs calculations showed the unique potential tunnels between the hydrogen bond donor–acceptor atoms of ac4C(34)(Proximal):G3/A3 base pairs suggesting role of ac4C in recognition of third letter of codons AUG/AUA. The “distal” conformation of ac4C might prevent misreading of AUA codon. Hence, this study could be useful to understand the role of ac4C in the tertiary structure folding of tRNA as well as in the proper recognition of codons during protein biosynthesis process.  相似文献   

5.
The synthesis of four chiral NAD+ models 1 and their 1,4-dihydro analogs 2 is described. From the temperature dependence of the 1H-nmr spectra it is concluded that for these compounds two preferred conformations I and II, differing slightly in energy, exist. Both conformations are “folded” with the more or less parallel p-anisyl and pyridine groups mutually gauche, but in I the pyridine group is rotated by about 180° as compared with II, thus leading to a conspicuous difference in orientation of the substituent Z (NH2CO, C6H5NHSO2, (CH2)4NSO2, or (C4H8ON)SO2) in the pyridine ring toward the anisyl group. The most stable conformation (I) has Z closest to the center of the p-anisyl group. In 360-MHz spectra of the dihydropyridines at low temperature (?10°C), slow interconversion of I and II leads to the observation of an XY pattern for the C-4 methylene protons of the 1,4-dihydropyridine system. The anisochronity in this methylene group is caused mainly by the anisotropy of the neighboring p-anisyl group.  相似文献   

6.
Summary We present ab initio calculations of the Fermi contact term and experimental correlations of six coupling constants, 3JH N H , 1JC H , 2JCH , 1JC N, 2JC N and 1JCN, in a peptide as functions of the backbone dihedral angles, and . Given estimates of experimental uncertainties, we find semiquantitative experimental correlations for 3JH N H , 1JC N and 2JC N, qualitative correlations for 1JC H and 2JCH , but no experimental correlations of practical utility for 1JCN, owing to its complex dependence on at least four dihedral angles. Errors in the estimation of dihedral angles from X-ray crystallographic data for proteins, which result from uncertainties in atom-to-atom distances, place substantial limitations on the quantitative reliability of coupling constant calculations fitted to such data. In the accompanying paper [Edison, A.S. et al., J. Biomol. NMR, 4, 543–551] we apply the results of the coupling constant calculations presented here to the estimation of and angles in staphylococcal nuclease from experimental coupling constants.Abbreviations AO atomic orbital - BPTI basic pancreatic trypsin inhibitor (bovine) - CI-2 chymotrypsin inhibitor 2 - E.COSY exclusive correlation spectroscopy (Griesinger et al., 1986) - nJAB single bond (n=1), geminal (n=2), or vicinal (n=3) coupling constant between nuclei A and B - LCAO linear combination of atomic orbitals - NBO natural bond orbital - n lone pair orbitals - bonding orbitals - * antibonding orbitals - dihedral angle or molecular orbital wave function; r2, correlation coefficient - RHF restricted Hartree-Fock; rmsd, root-mean-square deviation - 3-21G and 6-31G* molecular orbital basis set designations (Hehre et al., 1986)  相似文献   

7.
The solution three-dimensional structure of the protonated [Leu7]-surfactin, an hepta-peptide extracted from Bacillus subtilis, has been determined from two-dimensional 1Hnmr performed in 2H6-dimethylsulfoxide and combined with molecular modeling. Experimental data included 9 coupling constants, 61 nuclear Overhauser effect derived distances, NH temperature coefficients, and 13C relaxation times. Two distance geometry (DISMAN) protocols converged toward models of the structure and the best of them were refined by restrained and unrestrained molecular dynamics (GROMOS). Two structures in accord with the set of experimental constraints are presented. Both are characterized by a “horse saddle” topology for ring atoms on which are attached the two polar Glu and Asp side chains showing an orientation clearly opposite to that of the C11–13 aliphatic chain. Amphipathic and surface properties of surfactin are certainly related to the existence of such minor polar and a major hydrophobic domains. The particular “claw” configuration of acidic residues observed in surfactin gives important clues for the understanding of its cation binding and transporting ability. © 1994 John Wiley & Sons, Inc.  相似文献   

8.
The trianionic heptadentate ligand, (Z)-3-(5′-chlorosalicylhydrazinocarbonyl) propenoic acid, has been synthesized and reacted with FeCl3·6H2O, to produce the complex [FeIII6(C12H8N2O5Cl)6(H2O)4(CH3OH)2]·8H2O·4CH3OH. In the self-assembly process the ligand was esterified and transferred into (Z)-methyl 3-(5′-chlorosalicylhydrazinocarbonyl) propenoate. In the crystal structure, the neutral Fe(III) complex contain a 18-membered metallacrown ring consisting of six Fe(III) and six trianionic ligands. The 18-membered metallacrown ring is formed by the succession of six structural moieties of the type [Fe(III)-N-N]. Due to the meridional coordination of the ligands to the Fe3+ ions, the ligands enforce the stereochemistry of the Fe3+ ions as a propeller configuration with alternating Λ/Δ forms. The metallacrown can be treated with SnCl2 or Zn powder to obtain purified ester.  相似文献   

9.
β-Cyclodextrin (β-CD; cyclomaltoheptaose; cyclohepta-amylose; C42H70O35) crystallises from aqueous solutions of HI and of MeOH in the form of stout prisms, which are isomorphous to each other with monoclinic space-group P21; cell constants for C42H70O35 · 2HI · 8 H2O: a = 21.25(3), b = 10.28(2), c = 15.30(2) Å, β = 113.25(9)°, and Z = 2; and for C42H70O35 · MeOH · 6.5 H2O: a = 21.03(3), b = 10.11(2), c = 15.33(2) Å, β = 111.02(9)°, and Z = 2. X-ray counter data were used to determine the structures of both crystals, which belong to the cage type, with β-CD molecules in nearly identical, “round” shapes. In the HI complex, one I- is located inside, and one outside, the β-CD cavity; in the MeOH complex, the MeOH is within the cavity. The cavity is closed at the O-2,O-3 side by adjacent β-CD molecules, and at the O-6 side by water molecules hydrogen-bonded to the guest and to surrounding β-CD molecules. Interstices between β-CD molecules are filled by water of hydration molecules in distorted co-ordination.  相似文献   

10.
Quasi-elastic light scattering studies on some polyelectrolyte systems exhibit a somewhat “bizarre” behavior in the profile of the apparent diffusion coefficient Dapp as a function of the salt concentration Cs. As Cs is decreased, Dapp first increases in accordance with polyelectrolyte theories, and then undergoes a precipitous drop in value by over an order of magnitude at a well-defined critical value Cs = C. This “transition” from Cs > C (ordinary) to Cs < C (extraordinary) is referred to as the “ordinary-extraordinary” (o-e) transition. Ghosh, Peitzsch, and Reed [(1992) Biopolymers, Vol. 32, pp. 1105–1122] proposed a “filterable aggregate” (FA) and “other particle” interpretation for the o-e transition and its reversibility in regard to ionic strength changes. The present communication examines in detail the FA model as applied to the o-e transition. It is shown that the FA model fails to account of the established characteristics of the o-e transition. © 1993 John Wiley & Sons, Inc.  相似文献   

11.
Actin filament bundles isolated from Limulus sperm were used for quantitative electron microscope studies of F-actin assembly. The assembly rate constants were calculated. In addition, the critical concentrations (Cos) for both filament ends were directly determined. In 75 mM KCI and 1–5 mM Mg++, the Cos were 0.1 μM and 0.5 μM for the barbed and pointed ends, respectively. Substitution of Ca++ (20–200 μM) for Mg++ resulted in Cos of 0.4 μM for both filament ends. Consistent with these findings, filament growth occurred only from the barbed ends of Limulus bundles “seeded” into F-actin solutions in KCI and Mg++. Finally, filaments originally grown from the pointed filament ends of Limulus bundles were gradually lost as the actin solution reached steady state. These results demonstrate that actin filaments can “treadmill” under physiological conditions, albeit at very slow rates.  相似文献   

12.
《Proteins》2018,86(3):273-278
Unusual local arrangements of protein in Ramachandran space are not well represented by standard geometry tools used in either protein structure refinement using simple harmonic geometry restraints or in protein simulations using molecular mechanics force fields. In contrast, quantum chemical computations using small poly‐peptide molecular models can predict accurate geometries for any well‐defined backbone Ramachandran orientation. For conformations along transition regions—ϕ from −60 to 60°—a very good agreement with representative high‐resolution experimental X‐ray (≤1.5 Å) protein structures is obtained for both backbone C−1‐N‐Cα angle and the nonbonded O−1…C distance, while “standard geometry” leads to the “clashing” of O…C atoms and Amber FF99SB predicts distances too large by about 0.15 Å. These results confirm that quantum chemistry computations add valuable support for detailed analysis of local structural arrangements in proteins, providing improved or missing data for less understood high‐energy or unusual regions.  相似文献   

13.
The processes of photosynthesis, chemosynthesis and sulphate reduction were quantitatively studied in the brackish meromitic lake Faro (Sicily) with the aid of C14 and S35. The layer of “red water” was situated at the depht of the chemocline (13–14 m), where the average concentration of H2S was 10 mg/l. The total biomass of bacterioplankton consisted in this layer mostly of a brown Chlorobium which reached a wet weight of 30 g/m3. The production of photosynthesis in this layer was 30–60 µg C/l/day. The microbial population in the “red water” was found adapted to an extremely low light intensity and to show a light optimum at the depth µg 9m where only 2,5% of outside light penetrates. The photoautotrophic microflora is consumed by infusoria found in mass in the “red water” layer. An active H2S-production was found in the water column in the upper part of the H2S-zone and in the bottom sediments. The data are discussed from the view point of the trophology of meromitic basins.  相似文献   

14.
Summary In order to examine the internal dynamic processes of the dodecamer d(CGCAAATTTGCG)2, the 13C-enriched oligonucleotide has been synthesized. The three central thymines were selectively 13C-labeled at the C1′ position and their spin-lattice relaxation parameters R(CZ), R(CX,Y), R(HZ→CZ), R(2HZCZ), R(2HZCX,Y) and R(H infZ supC ) were measured. Density functions were computed for two models of internal motions. Comparisons of the experimental data were made with the spin-lattice relaxation rates rather than with the density functions, whose values were altered by accumulation of the uncertainties of each relaxation rate measurement. The spin-lattice relaxation rates were computed with respect to the motions of the sugar around the C1′-N1 bond. A two-state jump model between the anti- and syn-conformations with P(anti)/P(syn)=91/9 or a restricted rotation model with Δχ=28° and an internal diffusion coefficient of 30×107 s-1 gave a good fit with the experimental data. Twist, tilt or roll base motions have little effect on 13C1′ NMR relaxation. Simulation of spin-relaxation rates with the data obtained at several temperatures between 7 and 32 °C, where the dodecamer is double stranded, shows that the internal motion amplitude is independent of the temperature within this range, as expected for internal motion. Using the strong correlation which exists in a B-DNA structure between the χ and δ angle, we suggest that the change in the glycosidic angle value should be indicative of a sugar puckering between the C1′-exo and C2′-endo conformations.  相似文献   

15.
Conformationally and configurationally restricted rotameric probes of phenylalanine have been incorporated in the sequence of substance P (SP)—Arg-Pro-Lys-Pro-Gln-Gln-Phe-Phe-Gly-Leu-Met-NH2—for analyzing the binding pockets of Phe7 (S7) and Phe8 (S8), in the neurokinin-1 receptor. These analogues of phenylalanine are (2S, 3R)- and (2S, 3S)-indanylglycines, E- and Z-α, β-dehydrophenylalanines, and 2(S)-α, β-cyclopropylphenylalanines [ΔE Phe, ΔZPhe, ▿E2(S)Phe, and ▿Z2(S)Phe]. Binding data obtained with either conformationally (Ing diastereoisomers) or configurationally (ΔEPhe, ΔZPhe) probes have unveiled large differences in the binding potencies of these rotameric probes. With the support of nmr data and energy calculations done on these SP-substituted analogues, we attempt to answer questions inherent to such study. First, none of these six probes prevents the formation of bioactive conformation(s) of the backbone of SP. Second, both diastereoisomers (S, S) and (S, R) of indanylglycine preferentially adopt, in the sequence of SP, the gauche (−) and trans side-chain orientations, respectively, as previously postulated from energy calculations with model peptides. However, in solution, the difference in energy between these rotamers included in the sequence of SP, compared to model peptides, is smaller since the other rotamer can be detected in [(2S, 3R) Ing7]SP. Finally, from this study we can hypothesize that the large variations observed in the affinities of Phe7 substituted analogues of SP must come from steric hindrance in the S7 binding site, which drastically restricts the space filling around the Cα (SINGLE BOND) Cβ bond of residue 7. © 1996 John Wiley & Sons, Inc.  相似文献   

16.
Over 40 years ago, Joliot et al. (Photochem Photobiol 10:309–329, 1969) designed and employed an elegant and highly sensitive electrochemical technique capable of measuring O2 evolved by photosystem II (PSII) in response to trains of single turn-over light flashes. The measurement and analysis of flash-induced oxygen evolution patterns (FIOPs) has since proven to be a powerful method for probing the turnover efficiency of PSII. Stemler et al. (Proc Natl Acad Sci USA 71(12):4679–4683, 1974), in Govindjee’s lab, were the first to study the effect of “bicarbonate” on FIOPs by adding the competitive inhibitor acetate. Here, we extend this earlier work by performing FIOPs experiments at various, strictly controlled inorganic carbon (Ci) levels without addition of any inhibitors. For this, we placed a Joliot-type bare platinum electrode inside a N2-filled glove-box (containing 10–20 ppm CO2) and reduced the Ci concentration simply by washing the samples in Ci-depleted media. FIOPs of spinach thylakoids were recorded either at 20-times reduced levels of Ci or at ambient Ci conditions (390 ppm CO2). Numerical analysis of the FIOPs within an extended Kok model reveals that under Ci-depleted conditions the miss probability is discernibly larger (by 2–3 %) than at ambient conditions, and that the addition of 5 mM HCO3 ? to the Ci-depleted thylakoids largely restores the original miss parameter. Since a “mild” Ci-depletion procedure was employed, we discuss our data with respect to a possible function of free or weakly bound HCO3 ? at the water-splitting side of PSII.  相似文献   

17.
The relative contribution of autotrophic carbon sources (aquatic macrophytes, flooded forest, phytoplankton) for heterotrophic bacterioplankton was evaluated in a floodplain lake of the Central Amazon. Stable carbon isotopes (13C) were used as tracers. Values of 13C of different autotrophic sources were compared to those of dissolved organic carbon (DOC) and those of bacterially produced CO2.The percentage of carbon derived from C4 macrophytes for bacterially produced CO2 was the highest, on average 89%. The average 13C value of CO2 from bacterial respiration was –18.5 ± 3.3. Considering a fractionation of CO2 of 3 by bacterial respiration, 13C value was –15.5, near C4 macrophyte 13C value (–13.1).The average value of total DOC 13C was –26.8 ± 2.4. The percentage of C4 macrophytes carbon for total DOC was on average 17%. Considering that bacteria consume mainly carbon from macrophytes, the dominance of C3 plants for total DOC probably reflects a faster consumption of the former source, rather than a major contribution of the latter source.Heterotrophic bacterioplankton in the floodplain may be an important link in the aquatic food web, transferring the carbon from C4 macrophytes to the consumers.  相似文献   

18.
Isotype analyses were performed on biochemical fractions isolated from leaves of Kalanchoe blossfeldiana Tom Thumb. during aging under long days or short days. Irrespective of the age or photoperiodic conditions, the intermediates of the starch-malate sequence (starch, phosphorylated compounds and organic acids) have a level of 13C higher than that of soluble sugars, cellulose and hemicellulose. In short days, the activity of the crassulacean acid metabolism pathway is predominant as compared to that of C3 pathway: leaves accumulate organic acids, rich in 13C. In long days, the activity of the crassulacean acid metabolism pathway increases as the leaves age, remaining, however, relatively low as compared to that of C3 pathway: leaves accumulate soluble sugars, poor in 13C. After photoperiodic change (long daysshort days), isotopic modifications of starch and organic acids suggest evidence for a lag phase in the establishment of the crassulacean acid metabolism pathway specific to short days. The relative proportions of carbon from a C3-origin (RuBPC acitivity as strong discriminating step, isotope discrimination in vivo=20) or C4-origin (PEPC activity as weak discriminating step, isotope discrimination in vivo=4) present in the biochemical fractions were calculated from their 13C values. Under long days, 30 to 70% versus 80 to 100% under short days, of the carbon of the intermediates linked to the starch-malate sequence, or CAM pathway (starch, phosphorylated compounds and organic acids), have a C4-origin. Products connected to the C3 pathway (free sugars, cellulose, hemicellulose) have 0 to 50% of their carbon, arising from reuptake of the C4 from malate, under long days versus 30 to 70% under short days.Abbreviations CAM crassulacean acid metabolism - CAM pathway pathway with malate accumulation by -carboxylation of PEP, arising from glycolysis of starch (starch-malate sequence) - C3-metabolism metabolism with primary carbon fixed by the Calvin and Benson pathway (C3-origin) - C4-metabolism metabolism with primary carbon fixed by the Hatch and Slack pathway (C4-origin) - C3-pathway pathway with RuBPC activity and the Calvin and Benson pathway, irrespective of the CO2-source, atmospheric or reuptake of the C4 from malate - 13C()=(Rsample-RPDR)103/RPDB where PDB Pee Dee belemnite (belemnite from the Pee Dee formation, South Carolina) and R=13C/12C - D isotope discrimination - PEP phosphoenolpyruvate - PEPC (EC 4.1.1.31) PEP carboxylase - PGA phosphoglyceric acid - Py.di-PK (EC 2.7.9.1) pyruvate, Pi-dikinase - RuBP ribulose bisphosphate - RuBPC (EC 4.1.1.39) RuBP carboxylase - SD short days - LD long days  相似文献   

19.
Synthesis of “reversed” methylenecyclopropane analogues of nucleoside phosphonates 6a,7a, 6b, and 7b is described. 1-Bromo-1-bromomethylcyclopropane 8 was converted to the bromocyclopropyl phosphonate 9 by Michaelis-Arbuzov reaction with triisopropyl phosphite. Base-catalyzed β-elimination and deacetylation gave the key Z- and E-hydroxymethylcyclopropyl phosphonates 10 and 11 separated by chromatography. The Mitsunobu type of alkylation of 10 or 11 with adenine or 2-amino-6-chloropurine afforded phosphonates 12a, 12b, 13a, and 13b. Acid hydrolysis furnished the adenine and guanine analogues 6a, 7a, 6b, and 7b. The E and Z configuration was assigned on the basis of NOE experiments with phosphonates 6b and 7b. All Z- and E-isomers were also distinguished by different chemical shifts of CH2O or CH2N (H4 or H4′). Significant differences of the chemical shifts of the cyclopropane C3(3’) carbons and coupling constants 3JP,C2(2’) or 3JP,C3(3’) selective for the Z- or E-isomers were also noted. Phosphonates 6a, 7a, 6b, and 7b are devoid of significant antiviral activity.  相似文献   

20.
《Inorganica chimica acta》1988,141(2):281-288
The crystal structures and 95Mo NMR spectra of two complexes formed between 2-α-hydroxybenzyl- benzimidazole (C6H5·CHOH·C7H5N2=HOBB), as its sodium salt, and MoO2Cl2 are reported. [MoO2- (OBB)2]·EtOH (OBB=deprotonated HOBB) crystallizes in space group P21/n, with a=12.8441(7), b=15.917(3), c=13.314(2) Å, β=97.163(8)° and Z =4. The structure was determined from 3096 observed reflections and refined to a final R value of 0.030. The complex is a six coordinate cis-dioxo species, the 95Mo spectrum of which shows a single sharp peak at 56 ppm in dimethylformamide (DMF). The second complex, [Mo2O5(OBB)2]·EtOH·H2O, crystallizes in space group Pbca, with a=22.482(4), b=16.442(3), c=18.407(3) Å and Z=8. The structure was determined from 2936 observed reflections and refined to a final R value of 0.061. The complex is a binuclear doubly bridged species in which one metal atom is six coordinate while the other is five coordinate. Its 95Mo NMR spectrum in DMF shows a sharp peak at 124 ppm and a second broader much weaker peak at 51 ppm.  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