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.
The key enzyme of the fermentation of glutamate by Acidaminococcus fermentans, 2-hydroxyglutarylcoenzyme A dehydratase, catalyzes the reversible syn-elimination of water from (R)-2-hydroxyglutaryl-coenzyme A, resulting in (E)-glutaconylcoenzyme A. The dehydratase system consists of two oxygen-sensitive protein components, the activator (HgdC) and the actual dehydratase (HgdAB). Previous biochemical and spectroscopic studies revealed that the reduced [4Fe–4S]+ cluster containing activator transfers one electron to the dehydratase driven by ATP hydrolysis, which activates the enzyme. With a tenfold excess of titanium(III) citrate at pH 8.0 the activator can be further reduced, yielding about 50% of a superreduced [4Fe–4S]0 cluster in the all-ferrous state. This is inferred from the appearance of a new Mössbauer spectrum with parameters δ = 0.65 mm/s and ΔE Q = 1.51–2.19 mm/s at 140 K, which are typical of Fe(II)S4 sites. Parallel-mode electron paramagnetic resonance (EPR) spectroscopy performed at temperatures between 3 and 20 K showed two sharp signals at g = 16 and 12, indicating an integer-spin system. The X-band EPR spectra and magnetic Mössbauer spectra could be consistently simulated by adopting a total spin S t = 4 for the all-ferrous cluster with weak zero-field splitting parameters D = ?0.66 cm?1 and E/D = 0.17. The superreduced cluster has apparent spectroscopic similarities with the corresponding [4Fe–4S]0 cluster described for the nitrogenase Fe-protein, but in detail their properties differ. While the all-ferrous Fe-protein is capable of transferring electrons to the MoFe-protein for dinitrogen reduction, a similar physiological role is elusive for the superreduced activator. This finding supports our model that only one-electron transfer steps are involved in dehydratase catalysis. Nevertheless we discuss a common basic mechanism of the two diverse systems, which are so far the only described examples of the all-ferrous [4Fe–4S]0 cluster found in biology.  相似文献   

2.
3.
Pyranose oxidase and pyranosone dehydratase (aldos-2-ulose dehydratase), enzymes which convert in coupled reactions d-glucose to -pyrone cortalcerone, peaked coincidently during idiophasic growth of Phanerochaete chrysosporium under agitated conditions. The enzymes were purified from mycelial extracts of the fungus and separated from each other by hydrophobic interaction chromatography on Phenyl-Sepharose and Phenyl-Superose. Two pyranosone dehydratase activity peaks, PD I and PD II, were resolved. The major PD I fraction, consisting about 74% of the total dehydratase activity, was further purified by anion exchange chromatography on Mono Q to yield apparently pure enzyme as judged by SDS-PAGE and gel filtration on Superose 12. Isoelectric focusing indicated microheterogeneity of the protein by the presence of at least five protein bands with pI 5.1–5.3. PD II had a pI of 5.75. Overall PD I purification was 60.7-fold with 50% yield. The enzyme acted on several osones (glycosuloses), with the preferred substrate being d-glucosone. d-Xylosone and 6-deoxy-d-glucosone were dehydrated at C-3-C-4 to give the corresponding 5-hydroxy-2,3-dioxoalcanals (4-deoxy-2,3-glycosdiuloses), new enzymatically produced sugar derivatives. The latter labile compounds were trapped as diphenylhydrazine or o-phenylenediamine derivatives and spectroscopically identified. The analogous d-glucosone dehydration product did not accumulate due to its further transformation. pH optimum of PD I activity was 6.0 and its pH stability was optimal at pH 7-11. The enzyme was sensitive to Me2+ chelating agents and some heavy metal ions (Hg2+, Cu2+).Abbreviations DMAB 3-dimethylaminobenzoic acid - DTT dithiothreitol - MBTH 3-methyl-2-benzothiazolinone hydrazone-hydrochloride - PD pyranosone dehydratase - PMSF phenylmethylsulfonyl fluoride - POD pyranose oxidase  相似文献   

4.
d-Galacturonic acid, the main monomer of pectin, is an attractive substrate for bioconversions, since pectin-rich biomass is abundantly available and pectin is easily hydrolyzed. l-Galactonic acid is an intermediate in the eukaryotic pathway for d-galacturonic acid catabolism, but extracellular accumulation of l-galactonic acid has not been reported. By deleting the gene encoding l-galactonic acid dehydratase (lgd1 or gaaB) in two filamentous fungi, strains were obtained that converted d-galacturonic acid to l-galactonic acid. Both Trichoderma reesei Δlgd1 and Aspergillus niger ΔgaaB strains produced l-galactonate at yields of 0.6 to 0.9 g per g of substrate consumed. Although T. reesei Δlgd1 could produce l-galactonate at pH 5.5, a lower pH was necessary for A. niger ΔgaaB. Provision of a cosubstrate improved the production rate and titer in both strains. Intracellular accumulation of l-galactonate (40 to 70 mg g biomass−1) suggested that export may be limiting. Deletion of the l-galactonate dehydratase from A. niger was found to delay induction of d-galacturonate reductase and overexpression of the reductase improved initial production rates. Deletion of the l-galactonate dehydratase from A. niger also delayed or prevented induction of the putative d-galacturonate transporter An14g04280. In addition, A. niger ΔgaaB produced l-galactonate from polygalacturonate as efficiently as from the monomer.  相似文献   

5.
Castellaniella (ex Alcaligenes) defragrans strain 65Phen mineralizes monoterpenes in the absence of oxygen. Soluble cell extracts anaerobically catalyzed the isomerization of geraniol to linalool and the dehydration of linalool to myrcene. The linalool dehydratase was present in cells grown on monoterpenes, but not if grown on acetate. We purified the novel enzyme ∼1800-fold to complete homogeneity. The native enzyme had a molecular mass of 160 kDa. Denaturing gel electrophoresis revealed one single protein band with a molecular mass of 40 kDa, which indicated a homotetramer as native conformation. The aerobically purified enzyme was anaerobically activated in the presence of 2 mm DTT. The linalool dehydratase catalyzed in vitro two reactions in both directions depending on the thermodynamic driving forces: a water secession from the tertiary alcohol linalool to the corresponding acyclic monoterpene myrcene and an isomerization of the primary allylalcohol geraniol in its stereoisomer linalool. The specific activities (Vmax) were 140 nanokatals mg−1 for the linalool dehydratase and 410 nanokatals mg−1 for the geraniol isomerase, with apparent Km values of 750 μm and 500 μm, respectively. The corresponding open reading frame was identified and revealed a precursor protein with a signal peptide for a periplasmatic location. The amino acid sequence did not affiliate with any described enzymes. We suggest naming the enzyme linalool dehydratase-isomerase according to its bifunctionality and placing it as a member of a new protein family within the hydrolyases (EC 4.2.1.X).  相似文献   

6.
Tetrahydrofuran/toluene solutions of (nacnac)TiCl2 (nacnac = [ArNC(tBu)]2CH, Ar = 2,6-iPr2C6H3) react readily with KC8 to afford the titanium imide (ArN(tBu)CCHC(tBu))TiNAr(THF)Cl (1) in 67% isolated yield. Complex 1 forms from the two-electron reductive C-N bond cleavage of the β-diketiminate ligand. Likewise, reduction of (nacnac)TiCl(NHAr) (2), prepared in 85% yield from (nacnac)TiCl2 and LiNHAr, with KC8 results in formation of the imide-anilide analogue (ArN(tBu)CCHC(tBu))TiNAr(NHAr) (3) in 88% yield. Another reductant such as LitBu (3 equiv.) reacts cleanly with the precursor (nacnac)TiCl2 to afford the alkylidene-ate complex [Li(Et2O)][(ArN(tBu)CCHC(tBu))TiNAr(Et2O)] (4), in 81% yield. Complexes 1-4 have been characterized by 1H and 13C NMR spectra as well as single-crystal X-ray diffraction analysis. Plausible mechanisms to formation of compounds 1, 3 and 4 are also presented and discussed.  相似文献   

7.
A new anti-Prelog short-chain dehydrogenase/reductase (SDR) encoding gene lcsdr was cloned from Lactobacillus composti DSM 18527, and heterologously expressed in Escherichia coli. LcSDR is nicotinamide adenine dinucleotide phosphate (NADPH)-dependent and has a molecular weight of approximately 30 kDa. The optimal pH and temperature were 6.5 and 30?°C, respectively. The maximal reaction rate Vmax was 133.9 U mg?1; the Michaelis–Menten constant K m of LcSDR were 0.345 mM for acetophenone (1a), and 0.085 mM for NADPH. Through introducing an EsGDH-catalyzed NADPH regeneration system, a biocatalytic process for (R)-1-phenylethanol ((R)-1b) was developed with outstanding time–space yield. Under the optimized conditions, 50 g l?1 1a was converted to (R)-1b in 2 h with a yield of 93.8%, enantiomeric excess of product (e.e.p) above 99% and space–time yield of 562.8 g l?1 d?1.  相似文献   

8.
The pathway of d-xylose degradation in archaea is unknown. In a previous study we identified in Haloarcula marismortui the first enzyme of xylose degradation, an inducible xylose dehydrogenase (Johnsen, U., and Schönheit, P. (2004) J. Bacteriol. 186, 6198–6207). Here we report a comprehensive study of the complete d-xylose degradation pathway in the halophilic archaeon Haloferax volcanii. The analyses include the following: (i) identification of the degradation pathway in vivo following 13C-labeling patterns of proteinogenic amino acids after growth on [13C]xylose; (ii) identification of xylose-induced genes by DNA microarray experiments; (iii) characterization of enzymes; and (iv) construction of in-frame deletion mutants and their functional analyses in growth experiments. Together, the data indicate that d-xylose is oxidized exclusively to the tricarboxylic acid cycle intermediate α-ketoglutarate, involving d-xylose dehydrogenase (HVO_B0028), a novel xylonate dehydratase (HVO_B0038A), 2-keto-3-deoxyxylonate dehydratase (HVO_B0027), and α-ketoglutarate semialdehyde dehydrogenase (HVO_B0039). The functional involvement of these enzymes in xylose degradation was proven by growth studies of the corresponding in-frame deletion mutants, which all lost the ability to grow on d-xylose, but growth on glucose was not significantly affected. This is the first report of an archaeal d-xylose degradation pathway that differs from the classical d-xylose pathway in most bacteria involving the formation of xylulose 5-phosphate as an intermediate. However, the pathway shows similarities to proposed oxidative pentose degradation pathways to α-ketoglutarate in few bacteria, e.g. Azospirillum brasilense and Caulobacter crescentus, and in the archaeon Sulfolobus solfataricus.d-Xylose, a constituent of the polymer xylan, is the major component of the hemicellulose plant cell wall material and thus one of the most abundant carbohydrates in nature. The utilization of d-xylose by microorganisms has been described in detail in bacteria and fungi, for which two different catabolic pathways have been reported. In many bacteria, such as Escherichia coli, Bacillus, and Lactobacillus species, xylose is converted by the activities of xylose isomerase and xylulose kinase to xylulose 5-phosphate as an intermediate, which is further degraded mainly by the pentose phosphate cycle or phosphoketolase pathway. Most fungi convert xylose to xylulose 5-phosphate via xylose reductase, xylitol dehydrogenase, and xylulose kinase. Xylulose 5-phosphate is also an intermediate of the most common l-arabinose degradation pathway in bacteria, e.g. of E. coli, via activities of isomerase, kinase, and epimerase (1).Recently, by genetic evidence, a third pathway of xylose degradation was proposed for the bacterium Caulobacter crescentus, in analogy to an alternative catabolic pathway of l-arabinose, reported for some bacteria, including species of Azospirillum, Pseudomonas, Rhizobium, Burkholderia, and Herbasprillum (2, 3). In these organisms l-arabinose is oxidatively degraded to α-ketoglutarate, an intermediate of the tricarboxylic acid cycle, via the activities of l-arabinose dehydrogenase, l-arabinolactonase, and two successive dehydration reactions forming 2-keto-3-deoxy-l-arabinoate and α-ketoglutarate semialdehyde; the latter compound is further oxidized to α-ketoglutarate via NADP+-specific α-ketoglutarate semialdehyde dehydrogenase (KGSADH).2 In a few Pseudomonas and Rhizobium species, a variant of this l-arabinose pathway was described involving aldolase cleavage of the intermediate 2-keto-3-deoxy-l-arabinoate to pyruvate and glycolaldehyde, rather than its dehydration and oxidation to α-ketoglutarate (4). Because of the presence of some analogous enzyme activities in xylose-grown cells of Azosprillum and Rhizobium, the oxidative pathway and its variant was also proposed as a catabolic pathway for d-xylose. Recent genetic analysis of Caulobacter crecentus indicates the presence of an oxidative pathway for d-xylose degradation to α-ketoglutarate. All genes encoding xylose dehydrogenase and putative lactonase, xylonate dehydratase, 2-keto-3-deoxylonate dehydratase, and KGSADH were found to be located on a xylose-inducible operon (5). With exception of xylose dehydrogenase, which has been partially characterized, the other postulated enzymes of the pathway have not been biochemically analyzed.The pathway of d-xylose degradation in the domain of archaea has not been studied so far. First analyses with the halophilic archaeon Haloarcula marismortui indicate that the initial step of d-xylose degradation involves a xylose-inducible xylose dehydrogenase (6) suggesting an oxidative pathway of xylose degradation to α-ketoglutarate, or to pyruvate and glycolaldehyde, in analogy to the proposed oxidative bacterial pentose degradation pathways. Recently, a detailed study of d-arabinose catabolism in the thermoacidophilic crenarchaeon Sulfolobus solfataricus was reported. d-Arabinose was found to be oxidized to α-ketoglutarate involving d-arabinose dehydrogenase, d-arabinoate dehydratase, 2-keto-3-deoxy-d-arabinoate dehydratase, and α-ketoglutarate semialdehyde dehydrogenase (3).In this study, we present a comprehensive analysis of the complete d-xylose degradation pathway in the halophilic archaeon Haloferax volcanii. This halophilic archaeon was chosen because it exerts several suitable properties for the analyses. For example, it can be cultivated on synthetic media with sugars, e.g. xylose, an advantage for in vivo labeling studies in growing cultures. Furthermore, a shotgun DNA microarray of H. volcanii is available (7) allowing the identification of xylose-inducible genes, and H. volcanii is one of the few archaea for which an efficient protocol was recently described to generate in-frame deletion mutants.Accordingly, the d-xylose degradation pathway was elucidated following in vivo labeling experiments with [13C]xylose, DNA microarray analyses, and the characterization of enzymes involved and their encoding genes. The functional involvement of genes and enzymes was proven by constructing corresponding in-frame deletion mutants and their analysis by selective growth experiments on xylose versus glucose. The data show that d-xylose was exclusively degraded to α-ketoglutarate involving xylose dehydrogenase, a novel xylonate dehydratase, 2-keto-3-deoxyxylonate dehydratase, and α-ketoglutarate semialdehyde dehydrogenase.  相似文献   

9.
Anaerobically prepared cell-free extracts from Fusobacterium nucleatum contain 2-hydroxyglutaryl-CoA dehydratase with a specific activity of 20 nkat mg-1. The enzyme was purified 24-fold to a specific activity of 480 nkat mg-1 by anion exchange chromatography, gel filtration and chromatography on Blue-Sepharose. The activity of the purified enzyme was strictly dependent on the reductant Ti(III)citrate and stimulated 25-fold by 0.15 mM ATP and 5 mM MgCl2. ATP is hydrolysed to ADP during incubation with 2-hydroxyglutaryl-CoA dehydratase in the presence or absence of the substrate. The enzyme is extremely sensitive towards oxygen and is inhibited by 10 M chloramphenicol, 10 M 2,4-dinitrophenol or 0.15 mM hydroxylamine. The pure enzyme consists of three subunits (49 kDa), (39 kDa) and (24 kDa) in approximately equal amounts. In this respect the enzyme differs from the related 2-hydroxy-glutaryl-CoA dehydratase from Acidaminococcus fermentans and lactyl-CoA dehydratase from Clostridium propionicum both of which are composed of only two subunits with sizes comparable to those of and but require an additional protein for activity. The relative molecular mass of the native enzyme of about 100 kDa suggests a trimeric -structure. The homogeneous enzyme contains riboflavin (0.5 mol/112 kDa), iron and sulfur (3.5 mol/112 kDa each). Polyclonal antibodies directed against the 2-hydroxyglutaryl-CoA dehydratase from A. fermentans did not crossreact with cell free extracts or purified dehydratase from F. nucleatum. A comparison of the N-terminal amino acid sequences of the dehydratase subunits from A. fermentans and F. nucleatum, however, showed some similarities in the -subunits.Non-standard abbreviations DTT dithiothreitol - PAGE polyaccrylamide gel electrophoresis - VIS visible  相似文献   

10.
11.
12.

Objectives

To reduce the unpleasant odor during 1-deoxynojirimycin (DNJ) production, the genes of leucine dehydrogenase (bcd) and phosphate butryltransferase (ptb) were deleted from Bacillus amyloliquefaciens HZ-12, and the concentrations of branched-chain short fatty acids (BCFAs) and DNJ were compared.

Results

By knockout of the ptb gene, 1.01 g BCFAs kg?1 was produced from fermented soybean by HZ-12Δptb. This was a 56% decrease compared with that of HZ-12 (2.27 g BCFAs kg?1). Moreover, no significant difference was found in the DNJ concentration (0.7 g kg?1). After further deletion of the bcd gene from HZ-12Δptb, no BCFAs was detected in fermented soybeans with HZ-12ΔptbΔbcd, while the DNJ yield decreased by 26% compared with HZ-12.

Conclusions

HZ-12Δptb had decreased BCFAs formation but also maintained the stable DNJ yield, which contributed to producing DNJ-rich products with decreased unpleasant smell.
  相似文献   

13.
14.
The preparation ofΔ 1, 4 , 17-dione fromΔ 4 , 17-dione with the aid ofFusarium lateritium 403 is described, the yield being 80%, referred to the original steroid. The undesirable 1-dehydrotestololactone is formed under the given conditions only in traces. If progesterone was used as the starting steroid the yield of the undesirable 1-dehydrotestololactone is 40%, referred to the progesterone used. Dehydroepiandrosterone was not transformed by theFusarium lateritium strain to steroid metabolites. During the preparation of 1-dehydrotestosterone fromΔ 4 -androstene-3, 17-dione, using two successive microbial procedures (dehydrogenation of the A ring in position 1–2 and reduction of the keto group at C17 giving rise to the corresponding 17β-hydroxy derivative), the isolation yield was 55–60%, referred to the starting steroid.  相似文献   

15.

Objectives

To prepare (R)-phenyl-1,2-ethanediol ((R)-PED) with high enantiomeric excess (ee p) and yield from racemic styrene oxide (rac-SO) at high concentration by bi-enzymatic catalysis.

Results

The bi-enzymatic catalysis was designed for enantioconvergent hydrolysis of rac-SO by a pair of novel epoxide hydrolases (EHs), a Vigna radiata EH3 (VrEH3) and a variant (AuEH2A250I) of Aspergillus usamii EH2. The simultaneous addition mode of VrEH3 and AuEH2A250I, exhibiting the highest average turnover frequency (aTOF) of 0.12 g h?1 g?1, was selected, by which rac-SO (10 mM) was converted into (R)-PED with 92.6% ee p and 96.3% yield. Under the optimized reaction conditions: dry weight ratio 14:1 of VrEH3-expressing E. coli/vreh3 to AuEH2A250I-expressing E. coli/Aueh2 A250I and reaction at 20 °C, rac-SO (10 mM) was completely hydrolyzed in 2.3 h, affording (R)-PED with 98% ee p. At the weight ratio 0.8:1 of rac-SO to two mixed dry cells, (R)-PED with 97.4% ee p and 98.7% yield was produced from 200 mM (24 mg/ml) rac-SO in 10.5 h.

Conclusions

Enantioconvergent hydrolysis of rac-SO at high concentration catalyzed by both VrEH3 and AuEH2A250I is an effective method for preparing (R)-PED with high ee p and yield.
  相似文献   

16.
Previously we have characterized a threonine dehydratase mutant TDF383V (encoded by ilvA1) and an acetohydroxy acid synthase mutant AHASP176S, D426E, L575W (encoded by ilvBN1) in Corynebacterium glutamicum IWJ001, one of the best l-isoleucine producing strains. Here, we further characterized an aspartate kinase mutant AKA279T (encoded by lysC1) and a homoserine dehydrogenase mutant HDG378S (encoded by hom1) in IWJ001, and analyzed the consequences of all these mutant enzymes on amino acids production in the wild type background. In vitro enzyme tests confirmed that AKA279T is completely resistant to feed-back inhibition by l-threonine and l-lysine, and that HDG378S is partially resistant to l-threonine with the half maximal inhibitory concentration between 12 and 14 mM. In C. glutamicum ATCC13869, expressing lysC1 alone led to exclusive l-lysine accumulation, co-expressing hom1 and thrB1 with lysC1 shifted partial carbon flux from l-lysine (decreased by 50.1 %) to l-threonine (4.85 g/L) with minor l-isoleucine and no l-homoserine accumulation, further co-expressing ilvA1 completely depleted l-threonine and strongly shifted carbon flux from l-lysine (decreased by 83.0 %) to l-isoleucine (3.53 g/L). The results demonstrated the strongly feed-back resistant TDF383V might be the main driving force for l-isoleucine over-synthesis in this case, and the partially feed-back resistant HDG378S might prevent the accumulation of toxic intermediates. Information exploited from such mutation-bred production strain would be useful for metabolic engineering.  相似文献   

17.
d-Serine is a physiological co-agonist of the N-methyl-d-aspartate receptor. It regulates excitatory neurotransmission, which is important for higher brain functions in vertebrates. In mammalian brains, d-amino acid oxidase degrades d-serine. However, we have found recently that in chicken brains the oxidase is not expressed and instead a d-serine dehydratase degrades d-serine. The primary structure of the enzyme shows significant similarities to those of metal-activated d-threonine aldolases, which are fold-type III pyridoxal 5′-phosphate (PLP)-dependent enzymes, suggesting that it is a novel class of d-serine dehydratase. In the present study, we characterized the chicken enzyme biochemically and also by x-ray crystallography. The enzyme activity on d-serine decreased 20-fold by EDTA treatment and recovered nearly completely by the addition of Zn2+. None of the reaction products that would be expected from side reactions of the PLP-d-serine Schiff base were detected during the >6000 catalytic cycles of dehydration, indicating high reaction specificity. We have determined the first crystal structure of the d-serine dehydratase at 1.9 Å resolution. In the active site pocket, a zinc ion that coordinates His347 and Cys349 is located near the PLP-Lys45 Schiff base. A theoretical model of the enzyme-d-serine complex suggested that the hydroxyl group of d-serine directly coordinates the zinc ion, and that the ϵ-NH2 group of Lys45 is a short distance from the substrate Cα atom. The α-proton abstraction from d-serine by Lys45 and the elimination of the hydroxyl group seem to occur with the assistance of the zinc ion, resulting in the strict reaction specificity.  相似文献   

18.

Key message

Variability in traits of 15, diverse 6-year-old candidate plus trees of Jatropha curcas was determined to identify the best gain heritable traits correlating with oil yield for Jatropha improvement.

Abstract

Study was carried out on 15 6-year-old candidate plus trees of Jatropha curcas adapted on semi-arid wasteland with an objective to assess variation in morphological, physiological and oil quality characters. Heritable and non-heritable components of the total variability of the characters were determined by genotypic (GCV) and phenotypic (PCV) co-efficient of variation, heritability and genetic advance (GA) and the best gain traits for Jatropha improvement through selection and breeding were assessed. Further, association among the traits were assessed and germplasm were separated into different clusters. Significant variation was found among the different genotypes for all the characters. The photosynthetic and transpiration rate correlated with oil content, seed and oil yield. The chlorophyll pigments correlated positively with the photosynthetic rate and oil content. The seed oil content varied considerably from 27.68 % (JCN01) to 37.49 % (JCN14) and had high heritability, but it had low PCV and GCV and moderate GA. The oil yield plant?1 had high genetic variability and varied significantly from 0.07 (JCN15) to 0.47 kg plant?1 (JCN09/IC 565733). Though the different fatty acids differed significantly with different germplasm and also had high heritability, they had low PCV, GCV and GA. Seed weight, fruit weight, seed weight fruit?1 and seed yield plant?1 strongly correlated with oil yield and had moderate to high GCV, PCV, coupled with high heritability and GA. Germplasm were separated into four distinct clusters with a maximum inter distance found between cluster II and IV, and minimum between cluster I and III. The study helped to identify the superior germplasm among diverse genotypes of J. curcas that can serve as parents with desirable characters like high oil yield, low stomatal conductance and high water use efficiency for further breeding purposes.
  相似文献   

19.
The polyether bridged diphosphines,
(n = 1,2) have been prepared in 60–70% yield by reduction of the corresponding diphosphinedioxides with Si2Cl6 or (i-Bu)2AlH. These diphosphinedioxides have been prepared in 75–90% yield by reaction of two equivalents of the appropriate
with one equivalent of di- and triethylene glycol ditosylate.In general, reaction between the diphosphines, Rh(COD)acac and HClO4 gives a mixture of species, cis-[Rh(COD)(PP)] [ClO4] being the main complex. This complex reacts with CO to η3-trans- [Rh(CO)(POP)] [ClO4].  相似文献   

20.
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