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 15 毫秒
1.
Deuterated oleates have been synthesized by semihydrogenation of acetylenic intermediates. [11-2H2]Oleate was prepared by two-carbon chain extension of the C16 alcohol obtained from [1-2H2]octyl bromide and 7-octyn-1-ol. [8-2H2] and [7-2H2]oleates were both prepared from dimethyl suberate, tetradeutero intermediate C16 alcohols were synthesized from [1,8-2H4] and [2,7-2H4]octane diols by monobromination, conversion to deuterated 9-decyn-1-ols and reaction with octyl bromide. Oxidation gave [8-2H2]-9-octadecynoate and [2,7-2H2]-9-octadecynoate, after semihydrogenation of the latter, deuterons at C-2 were removed by exchange with aqueous alkali. [6-2H2] and [5-2H2]oleates were obtained from methyl 5-tetradecynoate, semihydrogenation, deuterium exchange at C-2 and two malonate extensions gave [6-2H2]oleate; reduction with lithium aluminum deuteride, two malonate extensions and semihydrogenation gave the [5-2H2] ester. [4-2H2] and [3-2H2]oleates were both obtained from methyl 7-cis-hexadecenoate, exchange of the α protons and chain extension gave the [4-2H2] ester and reduction with lithium aluminum deuteride and chain extension gave the [3-2H2] ester.  相似文献   

2.
H2 oxidation,O2 uptake and CO2 fixation in hydrogen treated soils   总被引:2,自引:0,他引:2  
Dong  Z.  Layzell  D.B. 《Plant and Soil》2001,229(1):1-12
In many legume nodules, the H2 produced as a byproduct of N2 fixation diffuses out of the nodule and is consumed by the soil. To study the fate of this H2 in soil, a H2 treatment system was developed that provided a 300 cm3 sample of a soil:silica sand (2:1) mixture with a H2 exposure rate (147 nmol H2 cm–3hr–1) similar to that calculated exist in soils located within 1–4 cm of nodules (30–254 nmol H2 cm–3hr–1). After 3 weeks of H2 pretreatment, the treated soils had a Km and Vmax for H2 uptake (1028 ppm and 836 nmol cm–3 hr–1, respectively) much greater than that of control, air-treated soil (40.2 ppm and 4.35 nmol cm–3 hr–1, respectively). In the H2 treated soils, O2, CO2 and H2 exchange rates were measured simultaneously in the presence of various pH2. With increasing pH2, a 5-fold increase was observed in O2 uptake, and CO2 evolution declined such that net CO2 fixation was observed in treatments of 680 ppm H2 or more. At the H2 exposure rate used to pretreat the soil, 60% of the electrons from H2 were passed to O2, and 40% were used to support CO2 fixation. The effect of H2 on the energy and C metabolism of soil may account for the well-known effect of legumes in promoting soil C deposition.  相似文献   

3.
Dissolved H2 and CO2 were measured by an improved manual headspace-gas chromatographic method during fermentative H2 production with N2 sparging. Sparging increased the yield from 1.3 to 1.8 mol H2/mol glucose converted, although H2 and CO2 were still supersaturated regardless of sparging. The common assumption that sparging increases the H2 yield because of lower dissolved H2 concentrations may be incorrect, because H2 was not lowered into the range necessary to affect the relevant enzymes. More likely, N2 sparging decreased the rate of H2 consumption via lower substrate concentrations.  相似文献   

4.
Methyl [17-2H2]oleate was prepared by stepwise reduction from 17-oxooleate in 24% yield. Methyl [18-2H3], [16-2H2], [14-2H2] and [12-2H2] oleates were synthesized from appropriately deuterated octylbromides by conversion to deuterated 7-hexadecyn-1-ols and chain extention to deuterated stearolates followed by semihydrogenation; overall yields were about 17%.  相似文献   

5.
《Inorganica chimica acta》1988,144(1):105-108
Reaction of the [Pt(dppe)] 2+ cation (where dppe = Ph 2PCH 2CH 2PPh 2) with the activated acetylene dimethyl acetylenedicarboxylate in methanol solution results in the formation of the bis-vinyl ether complex Pt(dppe)[(MeO 2C)CC(OMe)(CO 2Me)] 2. The compound is characterized by infrared, 1H and 31P( 1H) NMR spectroscopy, elemental analysis and X-ray crystallography.  相似文献   

6.
《Inorganica chimica acta》2001,312(1-2):111-116
The first structurally characterized, quadruply bonded complexes containing chiral diamine ligands, [Mo2(O2CCF3)2(S,S-dach)2(CH3CN)2][BF4]2 (1), and [Mo2(O2CCF3)2(R,R-dach)2(CH3CN)2][BF4]2 (2); (dach=1,2-diaminocyclohexane) were prepared by reactions of [Mo2(O2CCF3)2(CH3CN)6][BF4]2 with S,S-dach and R,R-dach, respectively, in CH3CN. Their UV–Vis and circular dichroism (CD) spectra have been recorded and their structures determined by X-ray crystallography. Crystals of complexes 1 and 2 conform to the space groups P2 with two independent half molecules in the asymmetric unit. The two molecules have a similar structure consisting of a Mo2 unit bridged by two cis-trifluoroacetate ligands and chelated by two dach ligands. Two acetonitrile molecules are coordinated to the Mo centers along the MoMo bond. The absorption wavelength at 507 nm for both 1 and 2 can be assigned to δxy→δxy* transitions. The solution CD spectra of these two complexes show two prominent bands at 525 and 385 nm and form mirror images of each other. The solid CD spectra of complexes 1 and 2 show marked red-shift in the absorption energies as compared with those measured in solution. The one-electron static coupling mechanism was invoked to explain the CD spectra for these complexes and the second lowest energy bands were assigned to be δxy→δx2y2 transitions.  相似文献   

7.
《Free radical research》2013,47(3):157-161
Many copper and iron complexes can be reduced by O-2 as well as by H2O2. According to the rates of reduction and the concentration of O-2 and H2O2, the metal complexes may serve either as catalyst of O-2 dismutation or as catalysts of the reaction between O-2 and H2O2 to form OH' radical (Haber-Weiss reaction). Various factors which influence whether metal complexes protect the biological systems from superoxide toxicity or enhance it are discussed.  相似文献   

8.
Nitric oxide (NO) has been shown to both enhance hydrogen peroxide (H2O2) toxicity and protect cells against H2O2 toxicity. In order to resolve this apparent contradiction, we here studied the effects of NO on H2O2 toxicity in cultured liver endothelial cells over a wide range of NO and H2O2 concentrations. NO was generated by spermine NONOate (SpNO, 0.001–1 mM), H2O2 was generated continuously by glucose/glucose oxidase (GOD, 20–300 U/l), or added as a bolus (200 μM). SpNO concentrations between 0.01 and 0.1 mM provided protection against H2O2-induced cell death. SpNO concentrations >0.1 mM were injurious with low H2O2 concentrations, but protective at high H2O2 concentrations. Protection appeared to be mainly due to inhibition of lipid peroxidation, for which SpNO concentrations as low as 0.01 mM were sufficient. SpNO in high concentration (1 mM) consistently raised H2O2 steady-state levels in line with inhibition of H2O2 degradation. Thus, the overall effect of NO on H2O2 toxicity can be switched within the same cellular model, with protection being predominant at low NO and high H2O2 levels and enhancement being predominant with high NO and low H2O2 levels.  相似文献   

9.
To improve understanding of the unimolecular decomposition mechanism of 1,2,4-butanetriol trinitrate (BTTN) in the gas phase, density functional theory calculations were performed to determine various decomposition pathways at the B3LYP/6-311G** level. Two main mechanisms for the unimolecular decomposition of BTTN were found. In the first, homolysis of one of the O–NO2 bonds occurs to form ?NO2 and CH2ONO2CHONO2CH2CH2O?, which subsequently decomposes to form CH3CHO + ?CHO + 3NO2 + HCHO. In the second, successive HONO elimination reactions yield three HONO and OHCCH2CHONO2CH2ONO2 fragments, which subsequently decompose to form CH3CHO + 2CO + 3HONO. We also found that the first pathway has a slightly lower activation energy than the second. The results show that the pathway involving O–NO2 cleavage is slightly more energetically favorable than that involving HONO elimination.  相似文献   

10.
Hydrothermal reactions of lead(II) acetate and HO2C(CH2)3N(CH2PO3H2)2 at 170 and 140 °C, respectively, resulted in two different lead diphosphonates, namely, Pb2[NH(CH2PO3)2] · 2H2O (1), in which the butyric acid moiety of the HO2C(CH2)3N(CH2PO3H2)2 has been cleaved and a novel layered compound, Pb3[HO2C(CH2)3NH(CH2PO3)2]2 · 2H2O (2). Their crystal structures have been determined by single crystal X-ray diffraction. In compound 1, the interconnection of the lead(II) ions by bridging amino-diphosphonate ligands leads to the formation of a 3D network. Compound 2 features an unusual triple-layer structure with the non-coordinated butyric acid moieties as pendant groups between the layers.  相似文献   

11.
Adding one equivalent of H2O2 to compounds of stoichiometry MoCl2(O)2(OPR3)2, OPR3 = OPMePh2 or OPPh3, leads to the formation of oxo-peroxo compounds MoCl2(O)(O2)(OPR3)2. The compound MoCl2(O)(O2)(OPMePh2)2 crystallized with an unequal disorder, 63%:37%, between the oxo and peroxo ligands, as verified by single-crystal X-ray diffractometry, and can be isolated in reasonable yields. MoCl2(O)(O2)(OPPh3)2, was not isolated in pure form, co-crystallized with MoCl2(O)2(OPPh3)2 in two ratios, 18%:82% and 12%:88%, respectively, and did not contain any disorder in the arrangement of the oxo and peroxo groups. These complexes accomplish the isomerization of various allylic alcohols. A mechanism of this reaction has been constructed based on 18O isotopic studies and involves exchange between the alcohol and metal bonded O atoms.  相似文献   

12.
The kinetics of the reactions between anhydrous HCl and trans-[MoL(CNPh)(Ph2PCH2CH2PPh2)2] (L=CO, N2 or H2) have been studied in thf at 25.0 °C. When L=CO, the product is [MoH(CO)(CNPh)(Ph2PCH2CH2PPh2)2]+, and when L=H2 or N2 the product is trans-[MoCl(CNHPh)(Ph2PCH2CH2PPh2)2]. Using stopped-flow spectrophotometry reveals that the protonation chemistry of trans-[MoL(CNPh)(Ph2PCH2CH2PPh2)2] is complicated. It is proposed that in all cases protonation occurs initially at the nitrogen atom of the isonitrile ligand to form trans-[MoL(CNHPh)(Ph2PCH2CH2PPh2)2]+. Only when L=N2 is this single protonation sufficient to labilise L to dissociation, and subsequent binding of Cl gives trans-[MoCl(CNHPh)(Ph2PCH2CH2PPh2)2]. At high concentrations of HCl a second protonation occurs which inhibits the substitution. It is proposed that this second proton binds to the dinitrogen ligand. When L=CO or H2, a second protonation is also observed but in these cases the second protonation is proposed to occur at the carbon atom of the aminocarbyne ligand, generating trans-[MoL(CHNHPh)(Ph2PCH2CH2PPh2)2]2+. Addition of the second proton labilises the trans-H2 to dissociation, and subsequent rapid binding of Cl and dissociation of a proton yields the product trans-[MoCl(CNHPh)(Ph2PCH2CH2PPh2)2]. Dissociation of L=CO does not occur from trans-[Mo(CO)(CHNHPh)(Ph2PCH2CH2PPh2)2]2+, but rather migration of the proton from carbon to molybdenum, and dissociation of the other proton produces [MoH(CO)(CNPh)(Ph2PCH2CH2PPh2)2]+.  相似文献   

13.
Ten fungi, Aspergillus niger, A. flavus, A. ochraceus, A. ruber, A. repens, A. amstelodami, Alternaria tenuis, Penicillium brevi-compactum, Cladosporium herbarum, and Chaetomium dolicotrichum, were isolated from moldy flue-cured tobacco and grown in various mixtures of N2-O2 or CO2-O2. A 1 to 5% concentration of O2 in an N2 atmosphere caused the greatest change in growth of the nine species, and a 10 to 20% concentration of O2 for A. flavus. All species, except A. amstelodami and A. ruber, grew faster in air than in mixtures containing 10% O2. High O2 concentrations generally inhibited furrow production in the mycelial mats. In an atmosphere of 5 to 40% O2 in the N2 atmosphere, furrows formed in mycelial mats between 5 and 40% O2 in the species except for A. ruber, A. repens, and A. amstelodami, which produced none in any concentration. As O2 decreased below 20%, spore production was progressively decreased, colony color faded to white, and cleistothecia formation was suppressed. In CO2-O2 mixtures radial growth of all species increased with each quantitative decrease of CO2. All species except A. niger grew faster in air than in 10% CO2. In contrast to N2-O2 mixtures, the fungi formed furrows, sporulation and cleistothecial formation were suppressed, and colony color changed to white in higher O2 concentrations.  相似文献   

14.
Four series of borosilicate glasses modified by alkali oxides and doped with Tb3+ and Sm3+ ions were prepared using the conventional melt quenching technique, with the chemical composition 74.5B2O3 + 10SiO2 + 5MgO + R + 0.5(Tb2O3/Sm2O3) [where R = 10(Li2O /Na2O/K2O) for series A and C, and R = 5(Li2O + Na2O/Li2O + K2O/K2O + Na2O) for series B and D]. The X‐ray diffraction (XRD) patterns of all the prepared glasses indicate their amorphous nature. The spectroscopic properties of the prepared glasses were studied by optical absorption analysis, photoluminescence excitation (PLE) and photoluminescence (PL) analysis. A green emission corresponding to the 5D47F5 (543 nm) transition of the Tb3+ ions was registered under excitation at 379 nm for series A and B glasses. The emission spectra of the Sm3+ ions with the series C and D glasses showed strong reddish‐orange emission at 600 nm (4G5/26H7/2) with an excitation wavelength λexci = 404 nm (6H5/24F7/2). Furthermore, the change in the luminescence intensity with the addition of an alkali oxide and combinations of these alkali oxides to borosilicate glasses doped with Tb3+ and Sm3+ ions was studied to optimize the potential alkali‐oxide‐modified borosilicate glass.  相似文献   

15.
16.
Protease-activated receptor-2 (PAR2) is one of four protease-activated G-protein-coupled receptors. PAR2 is expressed on multiple cell types where it contributes to cellular responses to endogenous and exogenous proteases. Proteolytic cleavage of PAR2 reveals a tethered ligand that activates PAR2 and two major downstream signaling pathways: mitogen-activated protein kinase (MAPK) and intracellular Ca2+ signaling. Peptides or peptidomimetics can mimic binding of the tethered ligand to stimulate signaling without the nonspecific effects of proteases. The most commonly used peptide activators of PAR2 (e.g. SLIGRL-NH2 and SLIGKV-NH2) lack potency at the receptor. However, although the potency of 2-furoyl-LIGRLO-NH2 (2-f-LIGRLO-NH2) underscores the use of peptidomimetic PAR2 ligands as a mechanism to enhance pharmacological action at PAR2, 2-f-LIGRLO-NH2 has not been thoroughly evaluated. We evaluated the known agonist 2-f-LIGRLO-NH2 and two recently described pentapeptidomimetic PAR2-specific agonists, 2-aminothiazol-4-yl-LIGRL-NH2 (2-at-LIGRL-NH2) and 6-aminonicotinyl-LIGRL-NH2 (6-an-LIGRL-NH2). All peptidomimetic agonists stimulated PAR2-dependent in vitro physiological responses, MAPK signaling, and Ca2+ signaling with an overall rank order of potency of 2-f-LIGRLO-NH2 ≈ 2-at-LIGRL-NH2 > 6-an-LIGRL-NH2 ≫ SLIGRL-NH2. Because PAR2 plays a major role in pathological pain conditions and to test potency of the peptidomimetic agonists in vivo, we evaluated these agonists in models relevant to nociception. All three agonists activated Ca2+ signaling in nociceptors in vitro, and both 2-at-LIGRL-NH2 and 2-f-LIGRLO-NH2 stimulated PAR2-dependent thermal hyperalgesia in vivo. We have characterized three high potency ligands that can be used to explore the physiological role of PAR2 in a variety of systems and pathologies.  相似文献   

17.
《Inorganica chimica acta》2006,359(4):1275-1281
Two new complexes of composition [Cu(2-NO2bz)2(3-pyme)2(H2O)2] (1) and/or [Cu{3,5-(NO2)2bz}2(3-pyme)2] (2) (3-pyme = 3-pyridylmethanol, ronicol or 3-pyridylcarbinol, 2-NO2bz = 2-nitrobenzoate and 3,5-(NO2)2bz = 3,5-dinitrobenzoate) have been prepared and studied by elemental analysis, electronic, infrared and EPR spectroscopy, magnetic susceptibility measurements and the structure of both complexes has been solved. Complex (1) shows an unusual molecular type of structure consisting of the [Cu(2-NO2bz)2(3-pyme)2(H2O)2] molecules held together by hydrogen bonds and van der Waals interactions. Complex (2) exhibits a polymeric chain-like structure [Cu{3,5-(NO2)2bz}2(3-pyme)2]n with copper atoms doubly bridged by two 3-pyridylmethanol molecules and the polymeric molecules are held together by van der Waals interactions. Complex (1) exhibits a magnetic moment μeff = 1.84 B.M. at 300 K that remains nearly constant within the temperature region (5–300 K). Further cooling results in lowering the magnetic moment to μeff = 1.82 B.M. at 1.8 K. The magnetic susceptibility temperature dependence obeys Curie–Weiss law with Curie constant of 0.423 cm3 K mol−1 and with Weiss constant of −0.06 K. The magnetic moment of (2) exhibits a small increase with a decrease in the temperature (μeff = 1.80 B.M. at 300 K and μeff = 1.85 B.M. at 1.8 K) with Curie constant of 0.409 cm3 K mol−1 and with Weiss constant of +1.1 K, which can indicate a very weak ferromagnetic interaction between the copper atoms within the chain. Applying the molecular field model resulted in obtaining zJ′ values −0.08 cm−1 for complex (1), and −0.07 cm−1 for complex (2), respectively, that could characterize intermolecular and interchain interactions transmitted through π–π stacking.  相似文献   

18.
19.
《Inorganica chimica acta》2006,359(9):2933-2941
The molecular structures of the thermodynamically unstable head-to-head isomers, HH-[Pd2(Ph2Ppy)2Cl2] and HH-[PtPd(Ph2Ppy)2I2], have been determined by single crystal X-ray diffraction. The two complexes have proved to be isostructural. The severe distortions of the bond angles from the ideal square planar geometry around the metal centers ligating the trans phosphorus donor atoms are indicative of a more pronounced internal strain in the HH isomers as compared to the HT counterparts. The enhanced internal strain is thought to be the major driving force responsible for the spontaneous conversion of the head-to-head isomers to their head-to-tail congeners. 13C NMR spectra in solution phase as well as solid-state 31P MAS NMR spectra have proved to be informative regarding the orientation of the asymmetric Ph2Ppy ligands.  相似文献   

20.
Regulation of H2 utilization, as monitored by the hydrogenase-mediated3H2 exchange reaction, was examined among phytoplankton communitiesin situ and populations in culture. During a 2-year study in the Chowan River, North Carolina, at least 2 major groups of phytoplankton dominated3H2 exchange rates. They included N2 fixing cyanobacteria and NO3 }- utilizing genera. Utilization of3H2 by N2 fixers was mainly dark-mediated, whereas3H2 utilization associated with periods of NO3 }- abundance revealed an increasing dependence on light. Inhibitors of N2 fixation (C2H2 and NH4 +) negatively affected3H2 utilization, substantiating previous findings that close metabolic coupling of both processes exists among N2 fixing cyanobacteria. Conversely, NO3 }- stimulated3H2 utilization among N2 and non-N2 fixing genera, particularly under illuminated conditions. A variety of environmental factors were shown to control3H2 exchange. In addition to the nitrogen sources discussed above, dissolved O2, photosynthetically available radiation (PAR), temperature, and pH changes altered3H2 exchange rates. It is likely that other factors not addressed here could also affect3H2 exchange rates. At least 2 ecological benefits from H2 utilization in natural phytoplankton can be offered. They include the simultaneous generation of adenosine triphosphate (ATP) and consumption of O2 during the oxidation of H2 via an oxyhydrogen or “Knallgas” reaction. Both processes could help sustain phytoplankton, and particularly cyanobacterial, bloom intensity under natural conditions when O2 supersaturation is common in surface waters. H2 utilization appeared to be a general feature of natural and laboratory phytoplankton populations. The magnitudes of3H2 utilization rates were directly related to community biomass. Although it can be shown that utilization rates are controlled by specific environmental factors, the potential relationships between H2 utilization and phytoplankton primary production remain poorly understood.  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