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 46 毫秒
1.
Daunorubicin and doxorubicin are efficient agents for cancer treatment. Their clinical efficacy is, however, hampered by their indiscriminant toxicity. This problem may be circumvented by encapsulating the drugs in liposomes and selectively targeting the tumor cells using tumor targeting agents. Furthermore, the antitumor effect could be enhanced by attaching the Auger electron emitter, (125)I, to daunorubicin and doxorubicin derivatives. In this context a number of ester, amide, and amine derivatives of daunorubicin and doxorubicin were synthesized. Benzoic acid ester derivatives of daunorubicin were synthesized by nucleophilic esterification of the 14-bromodaunorubicin with the potassium salt of the corresponding benzoic acid, resulting in good yields. Nicotinic acids and benzoic acids, activated with a succinimidyl group, were coupled to the amino group of daunorubicin to give the corresponding amide derivatives. Amine derivatives were obtained by the reductive amination of aromatic aldehydes with daunorubicin hydrochloride. The stannylated ester and amide derivatives were used as precursors for radioiodination. Radiolabeling with (125)I was performed using chloramine-T as an oxidant. The optimized labeling resulted in high radiolabeling yields (85-95%) of the radioiodinated daunorubicin and doxorubicin derivatives. Radioiodination of the amines was conducted at the ortho position of the activated phenyl rings providing moderate radiochemical yields (55-75%).  相似文献   

2.
Glycine conjugation of a series of benzoic acid derivatives was investigated in mouse kidney mitochondria. The chlorine and methyl substitutions in the para- and meta-positions of the benzene ring yielded an increase in glycine conjugation. The acids with a methoxy group showed a low degree of glycine conjugation. In addition, the acids with nitro or amino groups were conjugated to a slight extent with glycine. The in vitro conjugation of salicylic acid with glycine occurred not in liver but in kidney. The specificity of the renal medium chain acyl-CoA synthetase catalyzing the first reaction of glycine conjugation was also examined. The enzyme accepted not only medium chain fatty acids but also aromatic and arylacetic acids. The highest activity was shown with hexanoic acid. High activities were observed for benzoic acid derivatives with alkyl and alkoxyl groups in the para- and meta-positions of the benzene ring. An ortho-substituted acid exhibited no activity. In addition, the enzyme was less active with valproic acid, tranexamic acid, indomethacin and ketoprofen. The enzyme was inhibited by diflunisal, 2-hydroxydodecanoic acid and salicylic acid, which did not act as substrates. There was a poor correlation between the activity of the medium chain acyl-CoA synthetase and glycine conjugation of eleven substituted benzoic acids. These findings suggest that the present medium chain acyl-CoA synthetase is involved in glycine conjugation of the substituted acids in mouse kidney mitochondria, but there may be a larger contribution of another isoenzyme.  相似文献   

3.
Two large classes of phenolic acids were comprised in this review: benzoic acid derivatives and cinnamic acid derivatives. They have been found to be very extended in fruits and vegetables at different concentrations. For example, hydroxycinnamic acids concentration was higher than that found for hydroxybenzoic acids. Concerning their consumption, hydroxycinnamic acids provide larger contributions to the total polyphenol intake than benzoic acid derivatives or flavonoids. This phenolic acid intake is led by the coffee intake since it has very rich concentrations in hydroxycinnamic acids. Moreover, several experimental and epidemiological studies report the protection of phenolic acids against various degenerative diseases. However, despite all these interesting attributions and even if phenolic acids are the main polyphenols consumed, their bioavailability has not received as attention as that flavonoids. This concept is an essential step to understand the health-promoting properties of phenolic acids and to serve as tool to design in vivo and in vitro experiments to know their biological properties. Therefore, a compilation of bioavailability data of phenolic acids have been presented here paying attention to the two types of phenolic acid bioavailability, direct and indirect derived from the direct phenolic acid and flavonoid consumption, respectively. Then, a new relevant concept which may be named as total bioavailability of phenolic acids includes the direct absorption and metabolism of phenolic acids from food consumption and phenolic acids bioavailability as a result of the cleavage on the main skeleton ring of flavonoids by the gut microflora.  相似文献   

4.
The present study was designed to synthesize and evaluate pyrrolo-isoxazole benzoic acid derivatives as potential acetylcholinesterase (AChE) inhibitors for the management of Alzheimer's disease. The synthesis of pyrrolo-isoxazole benzoic acid derivatives involved ring opening cyclization of p-aminobenzoic acid with maleic anhydride to yield maleanilic acid, which in turn afforded N-arylmaleimide via ring closed cyclization. Azomethine-N-oxides were obtained by condensation of N-arylhydroxylamine with differently substituted benzaldehydes followed by refluxing of N-arylmaleimide with differently substituted azomethine-N-oxides to pyrrolo-isoxazole benzoic acid derivatives as cis- and trans-stereoisomers. The synthesized compounds were evaluated in vitro for AChE inhibitory activity in rat brain homogenate with donepezil as standard AChE inhibitor. Thereafter, the most potent test compound was evaluated for in vitro butyrylcholinesterase inhibitory activity and in vivo memory evaluation in scopolamine (0.4mg/kg)-induced amnesia in mice by employing Morris water maze test. All pyrrolo-isoxazole benzoic acid derivatives demonstrated potent AChE inhibitory activity. Most of compounds exhibited similar activity to donepezil and four of them (7h, 7i, 8i, and 8h, IC(50)=19.1±1.9-17.5±1.5nM) displayed higher inhibitory activity as compared to donepezil (21.5±3.2nM) with compound 8ia (IC(50)=17.5±1.5nM) being the most active one. The test compound 8ia also ameliorated scopolamine-induced amnesia in mice in terms of restoration of time spent in target quadrant (TSTQ) and escape latency time (ELT). It may be concluded that pyrrolo-isoxazole benzoic acid derivatives may be employed as potential AChE inhibitors.  相似文献   

5.
The carbonyl cofactor of bovine serum amine oxidase, recently identified as pyrroloquinoline quinone [Ameyama, Hayashi, Matsushita, Shinagawa & Adachi (1984) Agric. Biol. Chem. 48, 561-565; Lobenstein-Verbeek, Jongejan, Frank & Duine (1984) FEBS Lett. 170, 305-309], reacts stoichiometrically and irreversibly with hydrazides of phenylacetic acid and of benzoic acid. With the phenylacetic hydrazides a reversible intermediate step was detected by competition with substrate, carbonylic reagents or phenylhydrazine, a typical inhibitor of the enzyme. All hydrazides form an intense broad band with maximum absorbance in a narrow wavelength range (350-360 nm), irrespective of the acyl group, suggesting that the transition is located on the organic cofactor. A different situation is found with some phenylhydrazines, where extended conjugation can occur between the cofactor and the phenyl pi-electron system via the azo group, as shown by the lower energy and higher intensity of the transition. In this case the transition is sensitive to substituents in the phenyl ring. The c.d. spectrum of the adducts is influenced by the type of hydrazide (derived from phenylacetic acid or benzoic acid), by pH and by NN-diethyldithiocarbamate binding to copper, probably as a result of shifts of equilibria between hydrazone-azo tautomers.  相似文献   

6.
Dihydroxy stilbene derivatives were designed based on lithospermic acid B and were prepared from 4-(chloromethyl)benzoic acid. The inhibitory activities of the novel compounds against protein tyrosine phosphatase 1B (PTP1B) were evaluated. 3,4-Dihydroxy stilbene carbonyl compounds (7, 11b, 27b) inhibited PTP1B with IC50 values comparable to molybdate, while the conjugation-extended compound (15b) showed inhibition 3-fold better than preclinical RK682. The introduction of electron withdrawing groups or amides into the second phenyl ring, or extension of the conjugation into the stilbene molecule may increase stability of the generated radicals.  相似文献   

7.
Considering the potential of selective adenosine A3 receptor subtype ligands in the development of prospective therapeutic agents, an attempt has been made to explore physicochemical requirements of 1,2,4-triazolo[4,3-a]quinoxalin-1-one derivatives for A3 receptor binding. In this study, lipophilicity (logP), physicochemical substituent constants (pi, MR, sigma p) of phenyl ring substituents, and Wang-Ford charges of common atoms of the quinoxaline nucleus (calculated from molecular electrostatic potential surface of energy-minimized geometry using AM1 technique) were used as independent variables along with suitable dummy parameters. The best multiple linear regression (MLR) equation obtained from factor analysis (FA-MLR) as the preprocessing step could explain and predict 72.6% and 65.3%, respectively, of the variance of the binding affinity. The same equation also emerged as the best equation in the population of 100 equations obtained from genetic function approximation (GFA-MLR). The results suggested that presence of an electron-withdrawing group at the para position of the phenyl ring would be favorable for the binding affinity. Again, the presence of a nitro group at position R1 increases the binding affinity. When factor scores were used as predictor variables in the principal component regression analysis, the resultant model showed 78.6% explained variance and 63.1% predicted variance. The best equation derived from G/PLS could explain and predict 74.4% and 64.8%, respectively. The results have suggested the importance of Wang-Ford charges of atoms C15 and C19, apart from positive contributions of electron-withdrawing para substituents of the variance of the phenyl ring and nitro group at the R1 position.  相似文献   

8.
The first committed step in lipid A biosynthesis is catalyzed by uridine diphosphate-(3-O-(R-3-hydroxymyristoyl))-N-acetylglucosamine deacetylase (LpxC), a zinc-dependent deacetylase, and inhibitors of LpxC may be useful in the development of antibacterial agents targeting a broad spectrum of Gram-negative bacteria. Here, we report the design of amphipathic benzoic acid derivatives that bind in the hydrophobic tunnel in the active site of LpxC. The hydrophobic tunnel accounts for the specificity of LpxC toward substrates and substrate analogues bearing a 3-O-myristoyl substituent. Simple benzoic acid derivatives bearing an aliphatic 'tail' bind in the hydrophobic tunnel with micromolar affinity despite the lack of a glucosamine ring like that of the substrate. However, although these benzoic acid derivatives each contain a negatively charged carboxylate 'warhead' intended to coordinate to the active site zinc ion, the 2.25A resolution X-ray crystal structure of LpxC complexed with 3-(heptyloxy)benzoate reveals 'backward' binding in the hydrophobic tunnel, such that the benzoate moiety does not coordinate to zinc. Instead, it binds at the outer end of the hydrophobic tunnel. Interestingly, these ligands bind with affinities comparable to those measured for more complicated substrate analogue inhibitors containing glucosamine ring analogues and hydroxamate 'warheads' that coordinate to the active site zinc ion. We conclude that the intermolecular interactions in the hydrophobic tunnel dominate enzyme affinity in this series of benzoic acid derivatives.  相似文献   

9.
The chloroform extract of the heartwood of Pterocarpus santalinus yielded a mixture of red pigments which could be separated by polyamide column chromatography into two major compounds, santalin-A and santalin-B. Both gave the same permethyl ether, C38H36O10 which had 8 methoxyls and formed a number of derivatives typical of anhydrobenzopyranols. IR and UV spectra confirmed the same. NMR and MS suggested the presence of homoveratrayl group supported by the formation of veratraldehyde in alkali degradation. Permanganate oxidation gave 2,4-dimethoxy benzoic acid, veratric acid and 3,4,6-trimethoxy phthalic acid. On a basic fluorone skeleton, the substituents in the A ring are indicated by 2,4-dihydroxy-5-methoxy benzaldehyde, an alkali fission product and, further, 2,4-dimethoxy phenyl and homoveratryl units are located in ring C based on NMR, MS and biogenetic considerations. The residues constitute another benzene ring fused to ring C leading to the complete structures of the permethyl ether as (VII) which explains all its degradations and which constitutes a highly condensed biflavonoid of a new type.  相似文献   

10.
M J Behe  W S Englander 《Biochemistry》1979,18(19):4196-4201
The ability of a variety of phenyl derivatives to inhibit sickle cell hemoglobin gelation was placed on a quantitative scale by parallel equilibrium and kinetic assays. Modifications of the phenyl ring studied include polar, nonpolar, and charged substituents, added aromatic rings, and loss of aromaticity. Other noncovalent inhibitors previously reported to have high potency were measured and placed on the same quantitative scale. Some phenyl derivatives were found to be as effective an any other known noncovalent antigelling agent. The phenyl compounds penetrate easily into red cells, and their potency is tolerant to chemical modification, which holds out the possibility of designing low-toxicity derivatives. On the negative side, the level of potency obtainable appears to be inadequate for clinical use. The best phenyl inhibitors display a functionally defined inhibitory constant (K1) of 75 mM, and it can be estimated that inhibitor concentrations over 20 mM would be necessary to obtain minimal clinically significant benefit. Furthermore, with the variety of modifications tested here, no impressive increase in activity could be achieved over that found in the simplest phenyl compounds.  相似文献   

11.
Herein we describe the SAR of 1,5-biaryl pyrrole derivatives, with substituents in the 6-position of the benzoic acid moiety, as EP(1) receptor antagonists. Substitution at this position was well tolerated and led to the identification of several analogues with high affinity for the EP(1) receptor that displayed good efficacy in the established FCA model of inflammatory pain. Furthermore, several analogues were prepared which combined substitution at the 5- and 6-positions as well as derivatives with an aromatic ring fused to the 5- and 6-positions.  相似文献   

12.
Six derivatives of the general formula 2- or 4-(7-trifluoromethylquinolin-4-ylamino) benzoic acid N'-(nitrooxyacetyl or propionyl) hydrazide and an oxime of the formula 1-[4-(7-trifluoromethylquinolin-4-ylamino)phenyl]ethanone oxime were synthesized and tested for their in vivo anti-inflammatory, analgesic, and ulcerogenic properties, as well as their in vitro nitric oxide release ability. Compound 2-(7-trifluoromethylquinolin-4-ylamino)benzoic acid N'-(2-nitrooxy propionyl)hydrazide 12 showed an anti-inflammatory activity comparable to that of indomethacin in the carrageenan-induced rat paw edema test, and equipotency to glafenine in the acetic acid mice induced writhing model at 100mg/kg p.o., respectively. All the final compounds showed no tendency to induce stomach ulceration in rats; nitric oxide seems to contribute to their excellent safety profile.  相似文献   

13.
We have investigated the structural features of 5,10-dideaza-5,6,7,8-tetrahydrofolate (DDATHF) that determine the activity of this compound as an inhibitor of glycinamide ribonucleotide formyltransferase (GARFT) purified from mouse L1210 cells. 5-Deazatetrahydrofolate was as good an inhibitor of GARFT as DDATHF, indicating that isosteric replacement of nitrogen by carbon at the 5-position of tetrahydrofolate is sufficient for inhibition of GARFT. 5,10-Dideazafolic acid, 5,8,10-trideazatetrahydrofolate, and 2-desamino-5,10-dideazatetrahydrofolate were poor inhibitors of GARFT, indicating that a reduced pyridopyrimidine ring, N-8, and the 2-amino group of DDATHF, respectively, play an important role in the binding of tetrahydrofolate analogues to this enzyme. DDATHF analogues in which the phenyl ring was replaced either by a cyclohexyl ring or by methylene groups retained activity as inhibitors. 5,10-Dideazatetrahydrohomofolate was about 6 times more potent as an inhibitor of GARFT than DDATHF, but 5,10-dideazatetrahydronorfolate had about one-fifth of the activity of DDATHF. An analogue of DDATHF in which the glutamic acid side chain was replaced by aspartic acid (which was not a substrate for polyglutamation and was only weakly cytotoxic) was equiactive with DDATHF as an inhibitor of purified GARFT. Surprisingly, 5,10-dideazatetrahydropteroic acid was about as active as DDATHF as an inhibitor of GARFT, an indication that the glutamic acid in the side chain of DDATHF does not play a role in this ligand-enzyme interaction. The polyglutamate derivatives of DDATHF bound up to 100 times tighter to GARFT than DDATHF itself; longer chain polyglutamates conformed to Goldstein's zone B behavior under experimental conditions and were projected to be in zone C, i.e., stoichiometric inhibition, in vivo. We conclude that the presence of carbon at the 5-position of tetrahydrofolate analogues is sufficient for inhibition of GARFT, that N-8 and the 2-amino group are involved in binding of DDATHF to GARFT, probably through hydrogen bonds, and that the structures of the phenyl ring and amino acid side chain of DDATHF analogues are not primary determinants of GARFT inhibition by monoglutamate forms of these compounds. We also conclude that polyglutamation plays a major role in the potent cytotoxicity of DDATHF.  相似文献   

14.
The lead serum and glucocorticoid-related kinase 1 (SGK1) inhibitors 4-(5-phenyl-1H-pyrrolo[2,3-b]pyridin-3-yl)benzoic acid (1) and {4-[5-(2-naphthalenyl)-1H-pyrrolo[2,3-b]pyridin-3-yl]phenyl}acetic acid (2) suffer from low DNAUC values in rat, due in part to formation and excretion of glucuronic acid conjugates. These PK/glucuronidation issues were addressed either by incorporating a substituent on the 3-phenyl ring ortho to the key carboxylate functionality of 1 or by substituting on the group in between the carboxylate and phenyl ring of 2. Three of these analogs have been identified as having good SGK1 inhibition potency and have DNAUC values suitable for in vivo testing.  相似文献   

15.
Anaerobic naphthalene degradation by a sulfate-reducing enrichment culture   总被引:5,自引:0,他引:5  
Anaerobic naphthalene degradation by a sulfate-reducing enrichment culture was studied by substrate utilization tests and identification of metabolites by gas chromatography-mass spectrometry. In substrate utilization tests, the culture was able to oxidize naphthalene, 2-methylnaphthalene, 1- and 2-naphthoic acids, phenylacetic acid, benzoic acid, cyclohexanecarboxylic acid, and cyclohex-1-ene-carboxylic acid with sulfate as the electron acceptor. Neither hydroxylated 1- or 2-naphthoic acid derivatives and 1- or 2-naphthol nor the monoaromatic compounds ortho-phthalic acid, 2-carboxy-1-phenylacetic acid, and salicylic acid were utilized by the culture within 100 days. 2-Naphthoic acid accumulated in all naphthalene-grown cultures. Reduced 2-naphthoic acid derivatives could be identified by comparison of mass spectra and coelution with commercial reference compounds such as 1,2,3, 4-tetrahydro-2-naphthoic acid and chemically synthesized decahydro-2-naphthoic acid. 5,6,7,8-Tetrahydro-2-naphthoic acid and octahydro-2-naphthoic acid were tentatively identified by their mass spectra. The metabolites identified suggest a stepwise reduction of the aromatic ring system before ring cleavage. In degradation experiments with [1-(13)C]naphthalene or deuterated D(8)-naphthalene, all metabolites mentioned derived from the introduced labeled naphthalene. When a [(13)C]bicarbonate-buffered growth medium was used in conjunction with unlabeled naphthalene, (13)C incorporation into the carboxylic group of 2-naphthoic acid was shown, indicating that activation of naphthalene by carboxylation was the initial degradation step. No ring fission products were identified.  相似文献   

16.
The recombinant bioluminescent bacterium, DNT5, containing a nagR-nagAa::luxCDABE fusion, was tested in a multi-channel continuous monitoring system to evaluate its ability to detect benzoic acid derivatives. Seven chemicals, benzoic acid, salicylic acid, 2,5-dihydroxy benzoic acid, 3,5-dihydroxy benzoic acid, benzene, naphthalene and phenol, were used to characterize the responses of DNT5. This strain responded uniquely to each chemical, and these responses were then evaluated based upon the structures of each chemical. The greatest bioluminescent responses were to salicylic acid and benzoic acid, followed by 2,5-dihydroxy benzoic acid and 3,5-dihydroxy benzoic acid, but DNT5 was unresponsive when exposed to benzene, phenol and naphthalene, suggesting it has a strong preference for benzoic acid derivatives with few or no ring-substituted groups.  相似文献   

17.
A series of anilides and phenyl esters of piperidine-3-carboxylic acid (nipecotic acid) were synthesized and tested for the ability to inhibit aggregation of human platelet rich-plasma triggered by adenosine 5'-diphosphate (ADP) and adrenaline. As a rule, amides were about two times more active than the corresponding esters, and derivatives bearing substituents at the para position of the phenyl ring were significantly more active than the meta-substituted ones. Among the tested compounds, 4-hexyloxyanilide of nipecotic acid (18a) was found to be the most active one, its IC(50) value being close to that of the most active bis-3-carbamoylpiperidines reported in literature (ca. 40 micro M) and aspirin (ca. 60 microM) in ADP- and adrenaline-induced aggregation, respectively. Compared with the isomeric 4-hexyloxyanilides of piperidine-2-carboxylic (pipecolinic) and piperidine-4-carboxylic (isonipecotic) acids, compound 18a showed higher activity, and a Hansch-type quantitative structure-activity relationship (QSAR) study highlighted lipophilicity and increase in electron density of the phenyl ring as the properties which mainly increase the antiplatelet activity (r(2)=0.74, q(2)=0.64). The interaction of nipecotoyl anilides with phosphatidylinositol, a major component of the inner layer of the platelet membranes, was investigated by means of flexible docking calculation methods to give an account of a key event underlying their biological action.  相似文献   

18.
The spectral properties of diastereomeric esters and amides (1b-20b), derived from optically pure 2-(1-naphthyl)-2-phenylacetic acids (1-NPA), were systematically investigated. It was found that all compounds prepared exhibit the NMR spectral nonequivalence (Deltadelta) with regular sign distribution of particular groups according to the predicted model. Further, the analysis of data revealed that the phenyl ring is responsible for a shielding effect (upfield shift) instead of a naphthyl one. This conclusion is supported by the crystallographic analysis showing the almost ap-arrangement of the acid methine hydrogen atom and carbonyl group. In this arrangement, the phenyl ring faces toward the ester part of the molecule while the naphthyl one is orthogonal to the phenyl plane. Therefore, the mutual position of phenyl and alkyl groups with respect to the central molecule co-planarity thus determines the chemical shifts of the alcohol/amine substituents. The relative magnitude of the Deltadelta corresponds to those of Mosher's derivatives.  相似文献   

19.
Aza TSAO-T derivatives bearing a substituted dihydroisothiazole dioxide ring with a phenyl group at 5' position were prepared. Biological evaluation showed that phenyl group gives rise to a dramatical decrease of the inhibitory effect.  相似文献   

20.
The phenylacetic acid transport system (PATS) of Pseudomonas putida U was studied after this bacterium was cultured in a chemically defined medium containing phenylacetic acid (PA) as the sole carbon source. Kinetic measurement was carried out, in vivo, at 30 degrees C in 50 mM phosphate buffer (pH 7.0). Under these conditions, the uptake rate was linear for at least 3 min and the value of Km was 13 microM. The PATS is an active transport system that is strongly inhibited by 2,4-dinitrophenol, 4-nitrophenol (100%), KCN (97%), 2-nitrophenol (90%), or NaN3 (80%) added at a 1 mM final concentration (each). Glucose or D-lactate (10 mM each) increases the PATS in starved cells (140%), whereas arsenate (20 mM), NaF, or N,N'-dicyclohexylcarbodiimide (1 mM) did not cause any effect. Furthermore, the PATS is insensitive to osmotic shock. These data strongly suggest that the energy for the PATS is derived only from an electron transport system which causes an energy-rich membrane state. The thiol-containing compounds mercaptoethanol, glutathione, and dithiothreitol have no significant effect on the PATS, whereas thiol-modifying reagents such as N-ethylmaleimide and iodoacetate strongly inhibit uptake (100 and 93%, respectively). Molecular analogs of PA with a substitution (i) on the ring or (ii) on the acetyl moiety or those containing (iii) a different ring but keeping the acetyl moiety constant inhibit uptake to different extents. None of the compounds tested significantly increase the PA uptake rate except adipic acid, which greatly stimulates it (163%). The PATS is induced by PA and also, gratuitously, by some phenyl derivatives containing an even number of carbon atoms on the aliphatic moiety (4-phenyl-butyric, 6-phenylhexanoic, and 8-phenyloctanoic acids). However, similar compounds with an odd number of carbon atoms (benzoic, 3-phenylpropionic, 5-phenylvaleric, 7-phenylheptanoic, and 9-phenylnonanoic acids) as well as many other PA derivatives do not induce the system, suggesting that the true inducer molecule is phenylacetyl-coenzyme A (PA-CoA). Furthermore, after P. putida U is cultured in the same medium containing other carbon sources (glucose or octanoic, benzoic, or 4-hydroxyphenylacetic acid) in the place of PA, the PATS and PA-CoA are not detected; neither the PATS nor PA-CoA is found in cases in which mutants (PA- and PCL-) lacking the enzyme which catalyzed the initial step of the PA degradation (phenylacetyl-CoA ligase) are used. PA-CoA has been extracted from bacteria and identified as a true PA catabolite by high-performance liquid chromatography and also enzymatically with pure acyl-CoA:6-aminopenicillanic acid acyltransferase from Penicillium chrysogenum.  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