首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 140 毫秒
1.

Background

Most low molecular diols are highly water-soluble, hygroscopic, and reactive with many organic compounds. In the past decades, microbial research to produce diols, e.g. 1,3-propanediol and 2,3-butanediol, were considerably expanded due to their versatile usages especially in polymer synthesis and as possible alternatives to fossil based feedstocks from the bioconversion of renewable natural resources. This study aimed to provide a new way for bacterial production of an acetylated diol, i.e. acetylbutanediol (ABD, 3,4-dihydroxy-3-methylpentan-2-one), by acetoin metabolism.

Methodology/Principal Findings

When Bacillus pumilus ATCC 14884 was aerobically cultured in a chemically defined medium with acetoin as the sole carbon and energy source, ABD was produced and identified by gas chromatography – chemical ionization mass spectrometry and NMR spectroscopy.

Conclusions/Significance

Although the key enzyme leading to ABD from acetoin has not been identified yet at this stage, this study proposed a new metabolic pathawy to produce ABD in vivo from using renewable resources – in this case acetoin, which could be reproduced from glucose in this study – making it the first facility in the world to prepare this new bio-based diol product.  相似文献   

2.

Background

Efficient microbial production of chemicals is often hindered by the cytotoxicity of the products or by the pathogenicity of the host strains. Hence 2,3-butanediol, an important drop-in chemical, is an interesting alternative target molecule for microbial synthesis since it is non-cytotoxic. Metabolic engineering of non-pathogenic and industrially relevant microorganisms, such as Escherichia coli, have already yielded in promising 2,3-butanediol titers showing the potential of microbial synthesis of 2,3-butanediol. However, current microbial 2,3-butanediol production processes often rely on yeast extract as expensive additive, rendering these processes infeasible for industrial production.

Results

The aim of this study was to develop an efficient 2,3-butanediol production process with E. coli operating on the premise of using cost-effective medium without complex supplements, considering second generation feedstocks. Different gene donors and promoter fine-tuning allowed for construction of a potent E. coli strain for the production of 2,3-butanediol as important drop-in chemical. Pulsed fed-batch cultivations of E. coli W using microaerobic conditions showed high diol productivity of 4.5 g l?1 h?1. Optimizing oxygen supply and elimination of acetoin and by-product formation improved the 2,3-butanediol titer to 68 g l?1, 76% of the theoretical maximum yield, however, at the expense of productivity. Sugar beet molasses was tested as a potential substrate for industrial production of chemicals. Pulsed fed-batch cultivations produced 56 g l?1 2,3-butanediol, underlining the great potential of E. coli W as production organism for high value-added chemicals.

Conclusion

A potent 2,3-butanediol producing E. coli strain was generated by considering promoter fine-tuning to balance cell fitness and production capacity. For the first time, 2,3-butanediol production was achieved with promising titer, rate and yield and no acetoin formation from glucose in pulsed fed-batch cultivations using chemically defined medium without complex hydrolysates. Furthermore, versatility of E. coli W as production host was demonstrated by efficiently converting sucrose from sugar beet molasses into 2,3-butanediol.
  相似文献   

3.
Lipase-catalyzed diacylation of 1,3-butanediol   总被引:1,自引:0,他引:1  
Summary A kinetic resolution of 1,3-butanediol was accomplished by lipase-catalyzed enantioselective diacylations in organic solvent. Diacylation of 1,3-butanediol was carried out using immobilized lipase SP382 (from Candida sp.) to produce (R) -1,3-diacetoxybutane with 85.8% e.e.. And then, this optically active product was chemically hydrolyzed to diol, and re-acylated with lipase SP382 to (R) -1,3-diacetoxybutane with over 98% e.e..  相似文献   

4.

Objectives

To investigate the outcomes of capsule lost on cell transformation efficiency and chemicals (1,3-propanediol, 2,3-butanediol, and 2-ketogluconic acid) production by Klebsiella pneumoniae.

Results

The cps gene cluster showed low sequence homology with pathogenic strains. The wza is a highly conserved gene in the cps cluster that encodes an outer membrane protein. A non-capsulated mutant was constructed by deletion of wza. Phenotype studies demonstrated that non-capsulated cells were less buoyant and easy to sediment. The transformation efficiency of the non-capsulated mutant reached 6.4 × 105 CFU μg?1 DNA, which is 10 times higher than that of the wild strain. 52.2 g 1,3-propanediol L?1, 30.7 g 2,3-butanediol L?1, and 175.9 g 2-ketogluconic acid L?1 were produced by non-capsulated mutants, which were 10–40% lower compared to wild strain. Furthermore, viscosities of the three fermentation broths decreased to approximately 1.3 cP from the range of 1.8–2.2 cP.

Conclusions

Non-capsulated K. pneumoniae mutants should allay concerns regarding biological safety, improve transformation efficiency, lower viscosity, and subsequently ameliorate the financial burden of the downstream process of chemicals production.
  相似文献   

5.
The ability of Hansenula miso IFO 0146 to utilize various alcohols and acidic salts as sole sources of carbon and the ability of resting cells to oxidize various alcohols and glucose were studied. Growing cells could utilize only ethanol, glycerol, acetate and lactate, while resting cells grown on ethanol medium could oxidize various alcohols such as 1,2-ethanediol, DL-1,2-propanediol, 1,3-propanediol, meso-2,3-butanediol, DL-1,3-butane-diol, and 1,4-butanediol. From 2 g of 1,2-ethanediol and DL-l,3-butanediol, 1.3 g of glycolic acid and 0.5 g of β-hydroxybutyric acid respectively were produced. The organism formed d-arabinitol from glycerol and glucose, respectively. From 100 ml of culture in medium containing 6 ml of ethanol and 3.0 g of (NH4)2HPO4 as carbon and nitrogen sources 3.40 g of dried cells were obtained.  相似文献   

6.

Objectives

To investigate the efficiency of a new cascade biocatalysis system for the conversion of R, S-β-amino alcohols to enantiopure vicinal diol and β-amino alcohol.

Results

An efficient cascade biocatalysis was achieved by combination of a transaminase, a carbonyl reductase and a cofactor regeneration system. An ee value of > 99% for 2-amino-2-phenylethanol and 1-phenyl-1, 2-ethanediol were simultaneously obtained with 50% conversion from R, S-2-amino-2-phenylethanol. The generality of the cascade biocatalysis was further demonstrated with the whole-cell approaches to convert 10–60 mM R, S-β-amino alcohol to (R)- and (S)-diol and (R)- and (S)-β-amino alcohol in 90–99% ee with 50–52% conversion. Preparative biotransformation was demonstrated at a 50 ml scale with mixed recombinant cells to give both (R)- and (S)-2-amino-2-phenylethanol and (R)- and (S)-1-phenyl-1, 2-ethanediol in > 99% ee and 40–42% isolated yield from racemic 2-amino-2-phenylethanol.

Conclusions

This cascade biocatalysis system provides a new practical method for the simultaneous synthesis of optically pure vicinal diol and an β-amino alcohol.
  相似文献   

7.
An asymmetric hydrogen-transfer biocatalyst consisting of mutated Rhodococcus phenylacetaldehyde reductase (PAR) or Leifsonia alcohol dehydrogenase (LSADH) was applied for some water-soluble ketone substrates. Among them, 4-hydroxy-2-butanone was reduced to (S)/(R)-1,3-butanediol, a useful intermediate for pharmaceuticals, with a high yield and stereoselectivity. Intact Escherichia coli cells overexpressing mutated PAR (Sar268) or LSADH were directly immobilized with polyethyleneimine or 1,6-diaminehexane and glutaraldehyde and evaluated in a batch reaction. This system produced (S)-1,3-butanediol [87% enantiomeric excess (e.e.)] with a space time yield (STY) of 12.5 mg h−1 ml−1 catalyst or (R)-1,3-butanediol (99% e.e.) with an STY of 60.3 mg h−1 ml−1 catalyst, respectively. The immobilized cells in a packed bed reactor continuously produced (R)-1,3-butanediol with a yield of 99% (about 49.5 g/l) from 5% (w/v) 4-hydroxy-2-butanoate over 500 h.  相似文献   

8.
Enantioselective syntheses of all the possible stereoisomers of 1-phenyl-2,3-butanediol (erythro isomer 1 and threo isomer 2) and of 3-hydroxy-4-phenyl-2-butanone 3, the odor components of wisteria flowers, was accomplished via Sharpless asymmetric epoxydation. The absolute configurations of 1–3 were determined by an HPLC analysis of the corresponding MTPA esters of synthetic samples.  相似文献   

9.
We have isolated cDNA clones representing mRNAs encoding chitinase and 1,3--glucanase in cotton (Gossypium hirsutum L.) leaves. The chitinase clones were sequenced and found to encode a 28,806 Da protein with 71% amino acid sequence similarity to the SK2 chitinase from potato (Solanum tuberosum). The 1,3--glucanase clones encoded a 37,645 Da protein with 57.6% identity to a 1,3--glucanase from soybean (Glycine max). Northern blot analyses showed that chitinase mRNA is induced in plants treated with ethaphon or salicylic acid, whereas the levels of 1,3--glucanase mRNA are relatively unaffected. Southern blots of cotton genomic DNA and genomic clones indicated chitinase is encoded by a small gene family of which two members, Chi 2;1 and Chi 2;2, were characterized. These genes share 97% sequence identity in their transcribed regions. The genes were found to have three exons which are 309, 154 and 550 bp long, and two introns 99 and 154 bp in length. The 5-flanking regions of Chi 2;1 and Chi 2;2 exhibit a large degree of similarity and may contain sequences important for gene response to chemical agents and fungal attack.  相似文献   

10.
Röse US  Tumlinson JH 《Planta》2004,218(5):824-832
Feeding of Helicoverpa zea larvae on cotton (Gossypium hirsutum L.) flower buds (squares) for 24 or 48 h induced the release of a number of terpenes [(E)--ocimene, linalool, (E)--farnesene, (E,E)--farnesene, (E)-4,8-dimethyl-1,3,7-nonatriene, (E,E)-4,8,12-trimethyl-1,3,7,11-tridecatetraene], isomeric hexenyl butyrates, 2-methylbutyrates, indole and (Z)-3-hexenyl acetate. These compounds are not released in significant amounts from undamaged squares and freshly damaged squares. The release of inducible compounds was not limited to the damaged squares themselves. The compounds were also released systemically from the upper undamaged leaves of the same plant after 72 h. However, the composition of the blend of systemically released volatiles differed from the blend released by damaged squares. The compounds that were systemically released from undamaged leaves in response to feeding on the squares were (Z)-3-hexenyl acetate, (E)--ocimene, linalool, (E)-4,8-dimethyl-1,3,7-nonatriene, (E)--farnesene, (E,E)--farnesene, and indole. This study shows that insect damage inflicted to the reproductive parts of a plant causes a systemic emission of volatiles from its vegetative parts.  相似文献   

11.
The effect of temperature on enantioselectivity and desymmetrization in the acetylation of 2-phenyl-1,3-propanediol (1a), 2-benzyl-1,3-propanediol (1b), 2-methyl-2-phenyl-1,3-propanediol (1c) and 2-benzyl-2-methyl-1,3-propanediol (1d) by a novel lipase (CSL) isolated from the yeast Cryptococcus spp. S-2 was studied. Desymmetrization of 1a, 1c and 1d by CSL-catalyzed acetylation was observed in the temperature range of ?20°C to 40°C, while diacetylation of 1b occurred considerably even at 0°C. The kinetic parameters of the selectivity indicated that the acetylation of 1a is an entropy controlled process whereas the reaction of 1c and 1d is mainly controlled by the enthalpy term. In the monoacetylation of the diol 1d, the preferred configuration in the enantiomeric induction by CSL was opposite to that of immobilized porcine pancreatic lipase (PPL).  相似文献   

12.
1,3-Fucosyltransferase solubilized from human liver has been purified 40 000-fold to apparent homogeneity by a multistage process involving cation exchange chromatography on CM-Sephadex, hydrophobic interaction chromatography on Phenyl Sepharose, affinity chromatography on GDP-hexanolamine Sepharose and HPLC gel exclusion chromatography. The final step gave a major protein peak that co-chromatographed with 1,3-fucosyltransferase activity and had a specific activity of 5–6 µmol min–1 mg–1 and anM r 44 000 deduced from SDS-PAGE and HPLC analysis. The purified enzyme readily utilized Gal1-4GlcNAc, NeuAc2-3Gal1-4GlcNAc and Fuc1-2Gal1-4GlcNAc, with a preference for sialylated and fucosylated Type 2 acceptors. Fuc1-2Gal1-4Glc and the Type 1 compound Gal1-3GlcNAc were very poor acceptors and no incorporation was observed with NeuAc2-6Gal1-4GlcNAc. A polyclonal antibody raised against the liver preparation reacted with the homologous enzyme and also with the blood group Lewis gene-associated 1,3/1,4-fucosyltransferase purified from the human A431 epidermoid carcinoma cell line. No cross reactivity was found with 1,3-fucosyltransferase(s) isolated from myeloid cells. Examination by Northern blot analysis of mRNA from normal liver and from the HepG2 cell line, together with a comparison of the specificity pattern of the purified enzyme with that reported for the enzyme expressed in mammalian cells transfected with theFuc-TVI cDNA, suggests a provisional identification ofFuc-TVI as the major 1,3-fucosyltransferase gene expressed in human liver.Died June, 1991  相似文献   

13.
A NAD-dependent (R)-2,3-butanediol dehydrogenase (EC 1.1.1.4), selectively catalyzing the oxidation at the (R)-center of 2,3-butanediol irrespective of the absolute configuration of the other carbinol center, was isolated from cell extracts of the yeast Saccharomyces cerevisiae. Purification was achieved by means of streptomycin sulfate treatment, Sephadex G-25 filtration, DEAE-Sepharose CL-6B chromatography, affinity chromatography on Matrex Gel Blue A and Superose 6 prep grade chromatography leading to a 70-fold enrichment of the specific activity with 44% yield. Analysis of chiral products was carried out by gas chromatographic methods via pre-chromatographic derivatization and resolution of corresponding diasteromeric derivatives. The enzyme was capable to reduce irreversibly diacetyl (2,3-butanediol) to (R)-acetoin (3-hydroxy-2-butanone) and in a subsequent reaction reversibly to (R,R)-2,3-butanediol using NADH as coenzyme. 1-Hydroxy-2-ketones and C5-acyloins were also accepted as substrates, whereas the enzyme was inactive towards the reduction of acetone and dihydroxyacetone. The relative molecular mass (M r) of the enzyme was estimated as 140 000 by means of gel filtration. On SDS-polyacrylamide gel the protein decomposed into 4 (identical) subunits of M r 35 000. Optimum pH was 6.7 for the reduction of acetoin to 2,3-butanediol and 7.2 for the reverse reaction.Abbreviations GC-MS gas chromatography-mass spectrometry - i.d. internal diameter - M r relative molecular mass - MTPA-Cl -methoxy--trifluoromethylphenyl acetic acid chloride - PEIC 1-phenylethylisocyanate  相似文献   

14.
2-(4-Phenylthiazol-2(3H)-ylidene)-malononitrile was synthesized by treating 1-phenyl-2-thiocyanatoethanone with malononitrile. Reaction of 2-(4-phenylthiazol-2(3H)-ylidene)-malononitrile with hydrazine hydrate afforded 4-(4-phenylthiazol-2-yl)-1H-pyrazole-3,5-diamine, reaction with benzylidenemalononitrile yielded 2-(5-benzylidene-4-phenyl-5H-thiazol-2-ylidene)-malononitrile, and coupling with benzenediazonium chloride gave 2-(4-phenyl-5-phenylazo-3H-thiazol-2-ylidene)-malononitrile. Diaminopyrazole reacted with enaminonitrile to yield the 3-(4-phenylthiazol-2-yl)pyrazolo[1,5-a]pyrimidine-2,7-diamine. All synthesized compounds showed significant antimicrobial activities with MIC range of 5–750 µg/mL. The results demonstrated a correlation of the hydrophobicity of the compounds with their antimicrobial activity. The most potent antimicrobial compound was 2-(4-phenylthiazol-2(3H)-ylidene)-malononitrile.  相似文献   

15.
1,3-Propanediol production by Klebsiella pneumoniae was studied in batch cultures under N2 flow and four airflow systems. Different byproducts were formed under different aeration conditions. An anaerobic/aerobic combined fed-batch culture was developed giving 70 g 1,3-propanediol l(-1) and 16 g 2,3-butanediol l(-1) with total diol yield of 0.6 mol(-1) glycerol.  相似文献   

16.
Phytophthora infestans excretes an endo--1,3-, an endo--1,4-, and a-1,3-glucanase (laminarinase), a-1,6-glucosidase and possibly small amounts of a-1,4-glucosidase. Ether extracts from the infected resistant cultivar Eba but not from the susceptible Bintje inhibited growth of the parasite. Solavetivone and rishitin, two phytoalexins, and the steroid glycoalkaloid tomatine inhibited growth of the fungus and also activities of some of the fungal glucanases, whereas phytuberin, another phytoalexin, and the two phenolic compounds scopoletin and chlorogenic acid inhibited neither fungal growth nor fungal glucanases. The phytoalexin lubimin strongly reduced fungal growth but did not reduce the activities of any of the fungal glucanases tested. A potential role for host derived fungal glucanase inhibitors as factors of resistance in thePhytophthora-potato system is discussed.  相似文献   

17.
NAD-dependent butanediol dehydrogenase (Bdh1p) from Saccharomyces cerevisiae reversibly transforms acetoin to 2,3-butanediol in a stereospecific manner. Deletion of BDH1 resulted in an accumulation of acetoin and a diminution of 2,3-butanediol in two S. cerevisiae strains under two different growth conditions. The concentrations of (2R,3R)-2,3-butanediol are mostly dependent on Bdh1p activity, while those of (meso)-2,3-butanediol are also influenced by the activity of NADP(H)-dependent oxidoreductases. One of them has been purified and shown to be d-arabinose dehydrogenase (Ara1p), which converts (R/S)-acetoin to meso-2,3-butanediol and (2S,3S)-2,3-butanediol. Deletion of BDH2, a gene adjacent to BDH1, whose encoded protein is 51% identical to Bdh1p, does not significantly alter the levels of acetoin or 2,3-butanediol in comparison to the wild-type strain. Furthermore, we have expressed Bdh2p with a histidine tag and have shown it to be inactive toward 2,3-butanediol. A whole-genome expression analysis with microarrays demonstrates that BDH1 and BDH2 are reciprocally regulated.Acetoin and 2,3-butanediol are minor products generated by Saccharomyces cerevisiae during alcohol fermentation. Their sensory impacts on wine are poorly documented. Acetoin may affect the wine bouquet, although its perception threshold in wine is relatively high, around 150 mg/liter (21, 31). On the other hand, 2,3-butanediol is odorless (33) and cannot be expected to appreciably affect the sensory quality of wine. However, the compound may contribute to the wine body (28).Acetaldehyde, pyruvate, and α-acetolactate are the main precursors of acetoin in S. cerevisiae. Acetoin can be formed from acetaldehyde and/or pyruvate through an anomalous reaction of pyruvate decarboxylase. Thus, although its main activity is to irreversibly decarboxylate pyruvate to acetaldehyde, it can also catalyze carbon-carbon bond formation, yielding acetoin from pyruvate and/or acetaldehyde (2, 4). In addition, α-acetolactate would produce acetoin through its nonenzymatic decarboxylation to diacetyl and subsequent reduction to acetoin through the action of several NADH- and NADPH-dependent oxidoreductases (12). However, the situation is more complex in wine fermentation, where other yeasts and bacteria display supplementary enzymatic activities capable of producing both acetoin and 2,3-butanediol (1, 27).We have previously characterized a butanediol dehydrogenase (Bdh1p) as a medium-chain dehydrogenase/reductase (MDR) that can reversibly transform R-acetoin and S-acetoin to (2R,3R)-2,3-butanediol and meso-2,3-butanediol, respectively, in a NAD(H)-dependent reaction (10). BDH2 is a gene adjacent to BDH1 whose uncharacterized protein product (Bdh2p) shares 51% sequence identity with Bdh1p. To evaluate the in vivo roles of Bdh1p and Bdh2p, we compared the levels of several extracellular metabolites in cultures of wild-type and deficient strains. The results show that, although Bdh1p is the main enzyme in 2,3-butanediol production [essentially the (2R,3R)-2,3-butanediol stereoisomer], some meso-2,3-butanediol is still produced by the bdh1Δ strains. We have characterized Ara1p as an oxidoreductase that can reduce racemic acetoin to meso-2,3-butanediol and (2S,3S)-2,3-butanediol in the presence of NADPH.Furthermore, we have overexpressed Bdh2p with a histidine tag at its carboxyl terminus and have shown it to be inactive toward acetoin and 2,3-butanediol. A microarray study indicated that BDH1 and BDH2 are reciprocally regulated under the conditions studied.  相似文献   

18.
Chloroperoxidase from Caldariomyces fumago catalyses the oxidation of 1,2-dihydronaphthalene to (1R,2R)-(+)-dihydroxytetrahydronaphthalene in homogenous citrate buffer/ionic liquid mixtures, using t-butyl hydroperoxide as O<>2<> donor. It tolerates up to 30 (v/v) 1,3-dimethylimidazolium methylsulfate or 1-butyl-3-methylimidazolium methylsulfate. The enzyme activity in these ionic liquid co-solvent systems is retained for 24h, but it falls to 3h using non-ionic organic solvents such as t-BuOH or acetone.  相似文献   

19.
1,3-Propanediol inhibition during glycerol fermentation to 1,3-propanediol by Clostridium butyricum CNCM 1211 has been studied. The initial concentration of the 1,3-propanediol affected the growth of the bacterium more than the glycerol fermentation. μ max was inversely proportional to the initial concentration of 1,3-propanediol (0–65 g l−1). For glycerol at 20 g l−1, the growth and fermentation were completely stopped at an initial 1,3-propanediol concentration of 65 g l−1. However, for an initial 1,3-propanediol concentration of 50 g l−1 and glycerol at 70 g l−1, the final concentration (initial and produced) of 1,3-propanediol reached 83.7 g l−1(1.1 M), with complete consumption of the glycerol. Therefore, during the fermentation, the strain tolerated a 1,3-propanediol concentration higher than the initial inhibitory concentration (65 g l−1). The addition of 1,2-propanediol or 2,3-butanediol (50 g l−1) in the presence of glycerol (50–100 g l−1), showed that 2-diols reduced the μ max in a similar way to 1,3-propanediol. The measurement of the osmotic pressure of glycerol solutions, diols and diol/glycerol mixtures did not indicate any differences between these compounds. The hypothesis of diol inhibition was discussed. Taking into account the strain tolerance of highly concentrated 1,3-propanediol during fermentation, the fermentation processes for optimising production were considered. Received: 15 November 1999 / Revision received: 1 February 2000 / Accepted: 4 February 2000  相似文献   

20.
Structure? activity relationships of nine thiophenes, 2,2′: 5′,2″‐terthiophene ( 1 ), 2‐chloro‐4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yn‐1‐yl acetate ( 2 ), 4‐(2,2′‐bithiophen‐5‐yl)but‐3‐yne‐1,2‐diyl diacetate ( 3 ), 4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yne‐1,2‐diyl diacetate ( 4 ), 4‐(2,2′‐bithiophen‐5‐yl)‐2‐hydroxybut‐3‐yn‐1‐yl acetate ( 5 ), 2‐hydroxy‐4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yn‐1‐yl acetate ( 6 ), 1‐hydroxy‐4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yn‐2‐yl acetate ( 7 ), 4‐(2,2′‐bithiophen‐5‐yl)but‐3‐yne‐1,2‐diol ( 8 ), and 4‐[5‐(penta‐1,3‐diyn‐1‐yl)thiophen‐2‐yl]but‐3‐yne‐1,2‐diol ( 9 ), isolated from the roots of Echinops transiliensis, were studied as larvicides against Aedes aegypti. Structural differences among compounds 3, 5 , and 8 consisted in differing AcO and OH groups attached to C(3″) and C(4″), and resulted in variations in efficacy. Terthiophene 1 showed the highest activity (LC50, 0.16 μg/ml) among compounds 1 – 9 , followed by bithiophene compounds 3 (LC50, 4.22 μg/ml), 5 (LC50, 7.45 μg/ml), and 8 (LC50, 9.89 μg/ml), and monothiophene compounds 9 (LC50, 12.45 μg/ml), 2 (LC50, 14.71 μg/ml), 4 (LC50, 17.95 μg/ml), 6 (LC50, 18.55 μg/ml), and 7 (LC50, 19.97 μg/ml). These data indicated that A. aegypti larvicidal activities of thiophenes increase with increasing number of thiophene rings, and the most important active site in the structure of thiophenes could be the tetrahydro‐thiophene moiety. In bithiophenes, 3, 5 , and 8 , A. aegypti larvicidal activity increased with increasing number of AcO groups attached to C(3″) or C(4″), indicating that AcO groups may play an important role in the larvicidal activity.  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-23

京公网安备 11010802026262号