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 62 毫秒
1.
The amidase from Rhodococcus rhodochrous J1, which hydrolyses amide to acid and ammonia, was found to catalyze the synthesis of hydrazide using hydrazine as a substrate. This is the first report on the hydrazide synthesis through enzymatic reactions. The enzyme also acted on benzoic acid in the presence of hydrazine, yielding benzoic hydrazide. Together with the finding that benzoic hydrazide was converted into benzoic acid (when it was used as a substrate in the absence of hydrazine), these unique characteristics suggest that the reaction route for the formation of the acid from the hydrazide and that of the hydrazide from the acid are reversible to each other via the acyl-enzyme. Not only aromatic hydrazides but also aliphatic hydrazides were synthesized from the corresponding amides and hydrazine.  相似文献   

2.
Deuterium NMR spectra for the phenyl ring deuterons have been obtained for D-phenylalanine, L-phenylalanine, phenylacetic acid, and phenyl propionic acid in randomly oriented crystals of carboxypeptidase A as a function of water content. The spectra are analyzed using a two-site jump model for phenyl ring pi-flips when the ligand is bound to the protein, and the model includes the possibility that the ligand may exchange with isotropic or unbound environments within the crystal. Although the binding pocket may impose local dynamical constraints, a complete pi-flip motion is consistent with the spectra of all ligands at all water contents. The rate constants for the pi-flip at 298 K are found to be 7.5 x 10(5) S-1, 1.9 x 10(6) S-1, 4.0 x 10(6) S-1, and 4.0 x 10(6) S-1 for L-phenylalanine, D-phenylalanine, phenyl propionic acid, and phenylacetic acid, respectively, at water activity of 0.98. The pi-flip rate for the ligand bound to the enzyme increases with water content. Assuming that the activation barrier may be written, delta G+2 = delta G+2o + baw, where aw is the water activity, and the value of b is -1.9 kcal/mol for phenylacetic acid and phenyl propionic acid, -1.3 kcal/mol for L-phenylalanine, and -2.1 kcal/mol for D-phenylalanine. Phenylacetic acid crystals were studied as an example of a phenyl ring motion that is highly constrained by a known and symmetrical packing environment. The deuterium spectra are complex and are not consistent with pi-flip motions, but they are consistent with a superposition of ring jump motions of 24 degrees, 34 degrees, and 72 degrees, with probabilities in the ratio of 1:1:2. Because of the limited space for motion imposed by the tight packing in the crystal, these motions must be highly cooperative and probably locally coherent; however, the spectra by themselves do not prove this intuitively reasonable hypothesis.  相似文献   

3.
Volatile compounds of unifloral Salvia officinalis L. honey has been investigated for the first time. The botanical origin of ten unifloral Salvia honey samples has been ascertained by pollen analysis (the honey samples displayed 23-60% of Salvia pollen). Fifty-four volatile compounds were identified by GC and GC/MS in ten Salvia honey extracts obtained by ultrasound-assisted extraction (USE) with pentane/Et(2)O 1 : 2. The yield of isolated volatiles varied from 25.7 to 30.5 mg kg(-1). Salvia honey could be distinguished on the basis of the high percentage of benzoic acid (6.4-14.8%), and especially phenylacetic acid (5.7-18.4%). Minor, but floral-origin important volatiles were identified such as shikimate pathway derivatives, 'degraded-carotenoid-like' structures (3,5,5-trimethylcyclohex-2-ene derivatives) and 2,6,6-trimethylcyclohex-2-ene derivatives. Compounds from other metabolic pathways such as aliphatic acids and higher linear hydrocarbons, as well as heterocycles (pyrans, furans, and pyrroles), were also present. Most of the identified compounds do not constitute specific Salvia honey markers, due to their presence in honeys of other botanical origins; however, their ratio in different honeys could be useful to distinguish floral origin. Salvia-honey volatile markers were: benzoic acid, phenylacetic acid, p-anisaldehyde, alpha-isophorone, 4-ketoisophorone, dehydrovomifoliol, 2,6,6-trimethyl-4-oxocyclohex-2-ene-1-carbaldehyde, 2,2,6-trimethylcyclohexane-1,4-dione, and coumaran.  相似文献   

4.
Non-steroidal anti-inflammatory agents (NSAIDs) are widely used for pain relief. However, they have been associated with harmful and sometimes fatal side effects. Usually, the target organs are the GI tract and liver. In this study, we have investigated the physicochemical requirements of 21 NSAIDs for glucuronidation and cytotoxicity by quantitative structure-toxicity relationships (QSTRs) in isolated rat hepatocytes. Furthermore, we have investigated the contrast in physicochemical variables that correlated with NSAID-induced hepatocyte cytotoxicity when glucuronidation was inhibited with borneol. The competitive inhibition of hepatocyte p-nitrophenol glucuronidation by NSAIDs was determined by HPLC. Glucuronidation-inhibited hepatocytes were more susceptible to NSAID-induced cytotoxicity. Also, we found a parabolic correlation between lipophilicity and the inhibition of glucuronidation for a subset of NSAIDs. For NSAIDs with a benzoic acid moiety, cytotoxicity also correlated parabolically with lipophilicity, but correlated linearly with the HOMO-LUMO gap, and the first-order valence connectivity index. The cytotoxicity of NSAIDs with a phenylacetic acid (or propionic acid) substructure also correlated with lipophilicity, but not with the HOMO-LUMO gap. Our findings indicated that the inhibition of glucuronidation resulted in increased NSAID cytotoxicity, suggesting that acyl-glucuronide metabolites were acutely less cytotoxic. Also, comparative QSTRs revealed that benzoic acid NSAIDs may form cytotoxic radical metabolites (parameterized by the HOMO-LUMO gap) or alter mitochondrial respiration (parameterized by the connectivity index), whereas phenylacetic acid derived NSAIDs may form different cytotoxic metabolites, since they did not correlate with these parameters. In summary, we have used QSTRs as a tool to distinguish the cytotoxic mechanism of two groups of NSAIDs, which, if analyzed together as one group, did not reveal such mechanism-based differences.  相似文献   

5.
Novel mercapto-1,3,4-oxadiazole and -1,2,4-triazole derivatives were synthesized by various pathways starting from 4-(4-halogeno-phenylsulfonyl)benzoic acid hydrazides which were reacted with carbon disulfide or isothiocyanates. The heterocyclic mercaptans prepared in this way were assayed as inhibitors of three physiologically relevant isoforms of the zinc enzyme carbonic anhydrase (CA, EC 4.2.1.1), i.e., the cytosolic CA I and II, and the tumor-associated, transmembrane isozyme CA IX. Interesting biological activity was detected for some of the new mercaptans, with inhibition constants in the low micromolar range.  相似文献   

6.
Novel mercapto-1,3,4-oxadiazole and -1,2,4-triazole derivatives were synthesized by various pathways starting from 4-(4-halogeno-phenylsulfonyl)benzoic acid hydrazides which were reacted with carbon disulfide or isothiocyanates. The heterocyclic mercaptans prepared in this way were assayed as inhibitors of three physiologically relevant isoforms of the zinc enzyme carbonic anhydrase (CA, EC 4.2.1.1), i.e., the cytosolic CA I and II, and the tumor-associated, transmembrane isozyme CA IX. Interesting biological activity was detected for some of the new mercaptans, with inhibition constants in the low micromolar range.  相似文献   

7.
It was found that a new compound of phenylalanine metabolites (2-hydroxy-3-phenylpropenoic acid) and phenylacetic acid were formed in the cultured Czapek medium containing phenylalanine by Aspergillus sojae. 2-Hydroxy-3-phenylpropenoic acid (HPPA) was formed from phenylalanine (d- and l-form) via phenyllactic acid (d- and l-form), and degraded to benzoic acid, p-hydroxybenzoic acid, protocatechuic acid, and catechol in this order.

On the other hand, phenylacetic acid was formed from phenylpyruvic acid, and converted to homogentisic acid via o-hydroxyphenylacetic acid. From these results, a metabolic pathway of phenylalanine in Asp. sojae was proposed.  相似文献   

8.
Summary Fluorescent peptide substrates of the resonance energy transfer type, employing the donor-acceptor pair 5-[(2′-aminoethyl)amino]naphthalene sulphonic acid (EDANS) and 4-[[4′-(dimethylamino)phenyl]azo]benzoic acid (DABCYL), are widely used for continuous monitoring of the activity of various proteases. We describe here a flexible synthetic scheme which allows the synthesis of peptides having EDANS and DABCYL in any position of the sequence; the synthesis is entirely performed on solid phase with standard methods and makes use only of EDANS, DABCYL and commercially available amino acid derivatives. Moreover, we show that our scheme can be equally applied to the synthesis of EDANS/DABCYL-derivatised depsipeptides.  相似文献   

9.
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   

10.
A method based on reversed-phase HPLC is reported for the separation and quantification of various urinary aromatic metabolites: hippuric, phenylaceturic, salicyluric, benzoic, phenylacetic, salicylic. 3-phenylpropionic and cinnamic acids and several phenols in ruminant urine. In this method, a Nova-Pak C18 (4 μm) 150×3.9 mm I.D. column, two solvents [A: 15°b methanol in 20 mM acetic acid (pH 3.3); B: methanol] in a gradient mode at a flow-rate of 0.8 ml/min, and UV detection at 210 nm were used. Quantification of the total (free and conjugated) benzoic, phenylacetic and salicylic acids present in urine was achieved by hydrolysis of the samples in 3 M HCl at 100°C for 24 h prior to HPLC analysis. The lowest detection concentration was 50 μmol/I. This method is useful for scanning the profile of aromatic metabolites in urine of ruminants, which provides information on the diets the animals receive.  相似文献   

11.
As a result of screening of n-alkylbenzenes assimilating yeasts, it was shown that the yeasts which grew well on n-alkane (C15) showed also good growth on n-alkylbenzenes (from C7 to C19 of side chain). Among four Candida strains selected, C. tropicalis S131Y1 produced 4-(o-hydroxyphenyl)-butyric acid, o-hydroxy phenylacetic acid and phenylacetic acid from n-alkylbenzenes with even-carbon side chain and cinnamic acid from n-alkylbenzenes with odd-carbon side chain.

On the other hand, another three strains produced only phenylacetic acid from n-alkyl-benzenes with even-carbon side chain. In addition, as for the products from n-alkylbenzenes with odd-carbon side chain, two of three strains, C. parapsilosis IFO-1068 and C hydrocarbofumarica Et 15-2 produced benzoic acid.

From these oxidation products and oxygen uptake experiment, a new metabolic pathway of K-alkylbenzenes was assumed.  相似文献   

12.
In this study a series of isatin-based, pH-sensitive aryl imine derivatives with differing aromatic substituents and substitution patterns were synthesised and their acid-catalysed hydrolysis evaluated. These derivatives were functionalised at the C3 carbonyl group of a potent N-substituted isatin cytotoxin and were stable at physiological pH but readily cleaved at pH 4.5. Observed rates of hydrolysis for the embedded imine-acid moiety were in the order para-phenylpropionic acid>phenylacetic acid (para>meta)>benzoic acid (meta>para). The ability to fine-tune hydrolysis rates in this way has potential implications for optimising imine linked, tumour targeting cytotoxin-protein conjugates.  相似文献   

13.
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   

14.
Uptake of phenylacetic acid, the side-chain precursor of benzylpenicillin, was studied in Penicillium chrysogenum Wisconsin 54-1255 and in a strain yielding high levels of penicillin. In penicillin fermentations with the high-yielding strain, 100% recovery of phenylacetic acid in benzylpenicillin was found, whereas in the Wisconsin strain only 17% of the supplied phenylacetic acid was incorporated into benzylpenicillin while the rest was metabolized. Accumulation of total phenylacetic acid-derived carbon in the cells was nonsaturable in both strains at high external concentrations of phenylacetic acid (250-3500 microM), and in the high-yielding strain at low phenylacetic acid concentrations (2. 8-100 microM), indicating that phenylacetic acid enters the cells by simple diffusion, as concluded earlier for P. chrysogenum by other authors. However, at low external concentrations of phenylacetic acid saturable accumulation appeared in the Wisconsin strain. HPLC-analyses of cell extracts from the Wisconsin strain showed that phenylacetic acid was metabolized immediately after entry into the cells and different [14C]-labeled metabolites were detected in the cells. Up to approximately 50% of the accumulated phenylacetic acid was metabolized during the transport-assay period, the conversion having an impact on the uptake experiments. Nevertheless, accumulation of free unchanged phenylacetic acid in the cells showed saturation kinetics, suggesting the possible involvement of a high-affinity carrier in uptake of phenylacetic acid in P. chrysogenum Wisconsin 54-1255. At high concentrations of phenylacetic acid, contribution to uptake by this carrier is minor in comparison to simple diffusion and therefore, of no importance in the industrial production of penicillin.  相似文献   

15.
Brian RC 《Plant physiology》1967,42(9):1209-1214
The adsorption of chloro-aromatic acids to monomolecular layers of oat squashes is reported in earlier papers but it was not possible by the technique used, to measure unambiguously the adsorption of unsubstituted and 2,6-dichloro-aromatic acids. This has now been achieved by a modification of the earlier method and involves assessments of competitive adsorption between the unknown acid and a standard acid, using measurements of surface potential.

Benzoic and phenoxyacetic acids were not adsorbed but phenylacetic acid was weakly adsorbed. The second ring in naphthalene and naphthoxyacetic acids greatly increased adsorption. Substitution of the 2 and 6 positions in the phenyl and phenoxyacetic acids resulted in low adsorption but 2,6-disubstituted phenoxybutyric and benzoic acids were more highly adsorbed.

The adsorption values from earlier work are combined and discussed in relation to the growth-regulating activity of the acids. It is conciuded that there is no direct relation embracing all acids between adsorption and activity, notable exceptions being those substituted by chlorine in the 3-position of the aromatic ring. However, for a number of acids it is suggested that activity is limited not only by their ability to interact at enzyme sites but also by the amount of acid immobilised by adsorption when moving to these sites. It is also concluded that the hydrophilic/lipophilic balance of a growth regulator sometimes used as a guide to its activity, is an unreliable indication of interfacial behaviour.

  相似文献   

16.
Aryl metabolite biosynthesis was studied in the white rot fungus Bjerkandera adusta cultivated in a liquid medium supplemented with L-phenylalanine. Aromatic compounds were analyzed by gas chromatography-mass spectrometry following addition of labelled precursors ((14)C- and (13)C-labelled L-phenylalanine), which did not interfere with fungal metabolism. The major aromatic compounds identified were benzyl alcohol, benzaldehyde (bitter almond aroma), and benzoic acid. Hydroxy- and methoxybenzylic compounds (alcohols, aldehydes, and acids) were also found in fungal cultures. Intracellular enzymatic activities (phenylalanine ammonia lyase, aryl-alcohol oxidase, aryl-alcohol dehydrogenase, aryl-aldehyde dehydrogenase, lignin peroxidase) and extracellular enzymatic activities (aryl-alcohol oxidase, lignin peroxidase), as well as aromatic compounds, were detected in B. adusta cultures. Metabolite formation required de novo protein biosynthesis. Our results show that L-phenylalanine was deaminated to trans-cinnamic acid by a phenylalanine ammonia lyase and trans-cinnamic acid was in turn converted to aromatic acids (phenylpyruvic, phenylacetic, mandelic, and benzoylformic acids); benzaldehyde was a metabolic intermediate. These acids were transformed into benzaldehyde, benzyl alcohol, and benzoic acid. Our findings support the hypothesis that all of these compounds are intermediates in the biosynthetic pathway from L-phenylalanine to aryl metabolites. Additionally, trans-cinnamic acid can also be transformed via beta-oxidation to benzoic acid. This was confirmed by the presence of acetophenone as a beta-oxidation degradation intermediate. To our knowledge, this is the first time that a beta-oxidation sequence leading to benzoic acid synthesis has been found in a white rot fungus. A novel metabolic scheme for biosynthesis of aryl metabolites from L-phenylalanine is proposed.  相似文献   

17.
A simple, rapid and reproducible reversed-phase high-performance liquid chromatographic method for the simultaneous determination of benzoic acid (BA), phenylacetic acid (PAA) and their respective glycine conjugates hippuric acid (HA) and phenaceturic acid (PA) in sheep urine is described. The procedure involves only direct injection of a diluted urine sample, thus obviating the need for an extraction step or an internal standard. The compounds were separated on a Nova-Pak C18 column with isocratic elution with acetate buffer (25 mM, pH 4.5)—methanol (95:5). A flow-rate of 1.0 ml/min, a column temperature of 35°C and detection at 230 nm were employed. These conditions were optimized by investigating the effects of pH, molarity, methanol concentration in the mobile phase and column temperature on the resolution of the metabolites. The total analysis time was less than 15 min per sample. At a signal-to-noise ratio of 3 the detection limits for ten-fold diluted urine were 1.0 μg/ml for BA and HA and 5.0 μg/ml for PAA and PA with a 20-μl injection.  相似文献   

18.
Several aryl and heteroaryl hydrazides were synthesized and evaluated for their inhibitory effects on glutamic acid decarboxylase (GAD), GABA-alpha-oxoglutarate aminotransferase (GABA-T), and monoamine oxidase (MAO) enzyme systems in chick brain 24 h after their intramuscular administration (0.75 mmol/kg). All compounds produced a reduction in GAD, GABA-T, and MAO activity. Structure-activity relationships indicated that the ring structure had a greater influence on the degree of GAD and GABA-T inhibition than did the N'-terminal group. In contrast, structural requirements for MAO inhibition were much more restrictive. The intramuscular administration of benzoic acid hydrazide to chicks 24 h prior to their being exposed to oxygen at high pressure provided significant protection against the onset of the hyperbaric oxygen-induced seizures.  相似文献   

19.
We examined how molecular structure of 24 anthranilate and benzoic acid derivatives correlated with drinking behavior in European starlings Sturnus vulgaris. The effectiveness of bird repellents was associated with basicity, the presence of an electron-donating group in resonance with an electron-withdrawing carboxylic group on a phenyl ring, and a heterocyclic ring in the same pi cloud plane as the phenyl ring. Of the benzoic acid derivatives tested in this study, methyl, ethyl, dimethyl, and linalyl anthranilate as well as anthranilic acid and 4-ketobenztriazine were repellent to birds. Water consumption was significantly reduced relative to control levels at concentrations as low as 0.05% (weight/volume) for the best repellents. Further statistical tests showed that reduction in consumption for the best repellents was absolute, not significantly different from zero consumption. Anthranilic acid isomers were moderately good repellents. The ability to generate a model predicting repellency allows for the efficient identification and development of ecologically sound, nonlethal, taxa-specific repellents to be used for the protection of wildlife in agricultural and industrial applications.  相似文献   

20.
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 [1-13C]naphthalene or deuterated D8-naphthalene, all metabolites mentioned derived from the introduced labeled naphthalene. When a [13C]bicarbonate-buffered growth medium was used in conjunction with unlabeled naphthalene, 13C 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   

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