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 765 毫秒
1.
The binding of substrate and product analogs to phenylalanine ammonia-lyase (EC 4.3.1.5) from maize has been studied by a protection method. The ligand dissociation constants, KL, were estimated from the variation with [L] of the pseudo-first-order rate constants for enzyme inactivation by nitromethane. The phenylalanine analogs d- and l-2-aminooxy-3-phenylpropionic acid showed KL, values over 20,000-fold lower than the Km for l-phenylalanine. From these and other KL values it is deduced that when the enzyme binds l-phenylalanine the structural free energy stored in the protein is higher than when it binds the superinhibitors. Models for binding d- and l-phenylalanine and the superinhibitors are described. The enantiomeric pairs are considered to have similar KL values because they pack into the active site in a mirror-image relationship. If the elimination reaction approximates to the least-motion course deduced on stereoelectronic grounds, the mirror-image packing of the superinhibitors into the active site mimics the conformation inferred for a transition state in the elimination. It appears, therefore, that structural changes take place in the enzyme as the transition state conformation is approached causing stored free energy to be released. This lowers the activation free energy for the elimination reaction and accounts for the strong binding by the above analogs.  相似文献   

2.
Cyclic dipeptide cyclo(l- or d-Glu-l-His) carrying an anionic site and a nucleophilic site has been synthesized and used as a catalyst for the solvolysis of cationic esters in aqueous alcohols. In the solvolysis of 3-acyloxy-N-trimethylanilinium iodide (S+n, n = 2 and 10) and Cl?H3N+(CH2)11COOPh(NO2), no efficient nucleophilic catalysis was observed. On the other hand, in the solvolysis of Gly-OPh(NO2)·HCl, Val-OPh(NO2)·HCl and Leu-OPh(NO2)·HCl a very efficient general base-type catalysis by cyclo(l-Glu-l-His) was observed. In particular, with the latter two substrates the catalysis by cyclo(l-Glul-His) was more efficient than that by imidazole, although the catalysis was not enantiomer-selective. The diastereomeric cyclic dipeptide cyclo(d-Glu-l-His) was almost inactive under the same conditions. Confomation of cyclo(l- or d-Glu-l-His) in aqueous solution was investigated and the structure/catalysis relationship is discussed.  相似文献   

3.
Observation of random copolypeptides of γ-benzyl-l-glutamate with l-phenylalanine, l-valine and l-alanine was carried out in an electron microscope with samples cast from dilute solution. The relationship between the morphology and the molecular conformation in solution was studied with mixed solvents composed of chloroform and trifluoroacetic acid; these show a preference for α-helix and random coil, respectively. From the solutions in which molecules take α-helical conformation, fibrous films of nematic structure were formed. From random coil solutions discrete precipitates with folded molecules such as lamellar single crystals, piles of lamellae and structureless particles were formed. A copolypeptide containing l-valine in sufficiently large quantity to form β-structure also showed a variation in morphology with solvent, from films to discrete precipitates. It is suggested that the change in stiffness of the molecules contributes to the morphological variation.  相似文献   

4.
We have developed a new fluorescence assay for dipeptidylpeptidase IV using a tripeptide, l-prolyl-l-prolyl-l-alanine, which might be one of the potential natural substrates. The principle of the assay is based on the measurement of fluorescent adduct between alanine liberated from the tripeptide by enzymatic hydrolosis and o-phthaldialdehyde in the presence of 2-mercaptoethanol in aqueous alkaline medium. This new assay is sensitive enough to measure the enzyme activity in as little as 0.01 μl of human serum and in crevicular fluid obtained from human gingival sulcus. The Km value for the tripeptide was 1.7 · 10?5 M which is less than one-tenth of that obtained with other chromogenic or fluorogenic substrates. The interference by serum was overcome by simply incorporating the same amount of serum in the standards.  相似文献   

5.
A partition chromatographic procedure utilizing a cationic exchange resin column in the Li+ form and 90% ethanol as the mobile phase was employed to quantify 3-deoxy-d-manno-octulosonic acid (KDO) and l-glycero-d-manno-heptose in the lipopolysaccharides (LPS) of Re and RdP? rough mutants of Salmonella minnesota. In a standard mixture of monosaccharides, KDO eluted shortly after the void volume and heptose eluted after the neutral hexoses. Mild acid treatment of either the Re or RdP? LPS with 0.16 n methanesulfonic acid in the presence of Dowex 50-X8 resin (H+ form) released more than 80% of the KDO residues within 15 min. The heptose of the RdP? LPS, first detected after 90 min of hydrolysis, increased gradually to a maximum level at 12 h. A secondary gradual increase in KDO became apparent during the heptose release. The weight contents of these two monosaccharides based upon aheir maximum values detected during hydrolysis were 20.3 ± 0.6% KDO, for the Re LPS, and 13.8 ± 0.4% KDO and 12.0 ± 0.4% heptose, for the RdP? LPS. The relationship between the kinetics of release of KDO and heptose and the nature of the linkages involving these two monosaccharides are discussed.  相似文献   

6.
2,3-O-Isopropylidene-d-ribose diethyl dithioacetal, prepared from d-ribose, was converted in three steps into the corresponding dimethyl acetal, which was monotosylated at O-5, and the ester oxidized at C-4 with pyridinium chlorochromate; addition of methyl phenylphosphinate to the resulting pentos-4-ulose derivative then provided (4R,S)-4,5-anhydro-2,3-O-isopropylidene-4-C-[(R,S)-(methoxy)phenylphosphinyl]-d-erythro-pentose dimethyl acetal. Hydrogenation of this compound in the presence of Raney Ni, followed by reduction with SDMA, hydrolysis, and acetylation, yielded the title compounds (seven kinds), the structures of which were established on the basis of their 400-MHz, 1H-n.m.r. and mass spectra. A general dependence of the 2JPH and 3JPH values on the OPCH and PCCH dihedral angles provided an effective method for the assignment of the configurations and conformations of these 4-deoxy-4-phosphinyl-pentofuranoses.  相似文献   

7.
l-Alanylglycyl-l-alanylglycyl-l-alanylglycyl-l-serylglycine and its pentachlorophenyl ester methanesulphonate have been synthesized as monomers for the preparation of silk fibroin model polypeptide. The former octapeptide was polymerized with diphenylphosphorylazide (DPPA) and triethylamine in DMSO or in HMPA—pyridine, and the latter octapeptide pentachlorophenylester was polymerized by adding triethylamine in DMSO to give poly(l-alanylglycyl-l-alanylglycyl-l-alanylglycyl-l-serylglycine). This sequential polypeptide gave a similar i.r. pattern to the crystalline part of Bombyx mori silk fibroin, which indicated antiparallel β-conformation. Dialysis of the solution of this polymer in 60%, aqueous LiBr against water gave mainly the polymer of α-form. O.r.d. measurements suggest that this polypeptide exists as a random structure in dichloroacetic acid on in 60% aqueous LiBr.  相似文献   

8.
A sensitive and specific radioassay for l-glutamine-d-fructose-6-phosphate aminotransferase (EC 5.3.1.19) activity is presented. Picomoles of product are measurable, and the assay can be applied to systems having limited quantities of available protein, particularly in extracts of either cell or organ cultures. The assay is at least 10,000 times more sensitive under K1 concentrations of fructose 6-phosphate than the modified Elson-Morgan colorimetric assay and 20 times more sensitive under saturating conditions of fructose 6-phosphate. As little as 0.5 μg of cell-extract protein will yield measurable product. In contrast, 280 μg of crudeextract protein from colon is required with the modified Elson-Morgan colorimetric assay.  相似文献   

9.
In vitro biosynthesis of iodothyronines from diiodo-3,5-L-tyrosyl-diiodo-3,5-L-tyrosine A comparative study of two types of in vitro synthesis of iodothyronines has been done from 3,5-diiodotyrosine and from diiodo-3,5-L-tyrosyl-diiodo-3,5-L-tyrosine (Tyr(I)2-Tyr(I)2) (equimolecular in tyrosyl rings).Incubations are made with rat thyroid gland minces in Eagle's medium or with thyroid microsomal fraction.Synthesis of thyroid hormones from Tyr(I)2-Tyr(I)2 is faster and more important than from diiodo-3,5-L-tyrosine (Tyr(I)2).A mechanism of iodothyronine formation via Tyr(I)2 - Tyr(I)2 and different from the one occuring for Tyr(I)2 is suggested.  相似文献   

10.
2,5-Anhydro-d-altritol (2a) and the previously-unknown 2,5-anhydro-d-iditol (3a) have been prepared from 2,5-anhydro-d-mannitol (1a). The preparation of 3a from the intermediate epoxide 7b is particularly sensitive to pH, and a mechanism is proposed to explain this. Attention is drawn to the limitations of the trifluoroperacetic acid-disodium hydrogenphosphate procedure for the epoxidation of alkenes of diminished reactivity.  相似文献   

11.
A preparation of l-phenylalanine ammonia-lyase (EC 4.1.3.5.) from soybean (Glycine max L. cv. Kanrich) showed negative cooperativity with respect to l-phenylalanine and competitive inhibition by d-phenylalanine. A two-protomer partially concerted model for inhibition kinetics is described. If cooperativity is associated with ligand binding but not kcat, plots of v against log [S] at constant [I] are symmetrical. Such curves may be fitted by graphical or iterative least-squares methods. The experimental results conform to this restricted model. The three-substrate and three-inhibitor dissociation constants were estimated by a stepwise procedure. For substrate only the first and second dissociation constants were 12 and 78 μm, respectively, with a symmetry point value of 30.5 μm. To a first approximation, site occupancy determines the cooperativity. As d- and l-phenylalanine produce equivalent effects, they are assumed to pack into the same induced space. As ligand binding at one site has little influence on the relative d:l binding at the other and does not influence kcat, cooperativity probably reflects changes in regions remote from the active site such as the interface between the protomers. The regulatory range in [S] of the enzyme in vivo may be indicated by the linearity range of the semilog plot for the isolated enzyme. The observed range corresponds to a 100-fold change in [S] compared to a 10-fold change for Michaelis-Menten kinetics.  相似文献   

12.
Solvents play a stabilising role with the more stable conformations obtained in polar solvents than in vacuo. We investigate to what extent the structural propensities of the pentacyclo-undecane (PCU) cage polypeptide chain of the type Ac-3Aib-Cage-3Aib-NHMe are influenced in implicit water and in explicit solvents: methanol (MEOH), dimethyl sulphoxide (DMSO) and TIP3P water. The sampling of the α-helical conformations of the PCU cage polypeptide was investigated using the in-house modified PARM94 force-field parameters. Analysis of 50 ns molecular dynamics (MD) simulations revealed a tendency of the PCU cage polypeptide to assume bent structures, especially in polar solvents. The choice of solvents was designed to relate the simulations to physiological conditions. The individual amino-isobutyric acid residues predominantly sampled the right-handed and left-handed 310-helical conformations, indicating that the helical conformations are preferred in all four environments (in vacuo, MEOH, water and DMSO). Additionally, the 100 ns replica exchange MD (REMD) simulations of the PCU cage polypeptide in implicit water revealed more conformational variety present than in explicit solvents, and is more consistent with previous theoretical studies on the PCU cage residue. The present theoretical results may help in rationalising experimental results on these PCU cage polypeptides, and definitely show the importance of a dynamical approach for a correct interpretation and prediction of the conformational behaviour of the PCU cage molecules in different environments.  相似文献   

13.
Three-dimensional X-ray diffraction data were used to determine the crystal structure of sodium β-d-glucuronate monohydrate, a model system for investigating the factors involved in the binding of sodium ions to d-glucuronate residues of glycosaminoglycans. Crystals of the salt are monoclinic, space group P21, with a = 9.206(3) Å, b = 7.007(2) Å, c = 7.378(3) Å, β = 96.84(3)°, and Z = 2. Intensity data for 858 reflections were measured with an automated diffractometer. A trial structure, obtained by direct methods, was refined by least squares to R = 0.035. An outstanding feature of the crystal packing is the interaction of d-glucuronate anions with sodium ions. The sodium ion is coordinated to three symmetry-related d-glucuronate anions and to one water molecule. The d-glucuronate anion binds sodium cations through the three following sites: one that involves a carboxyl oxygen atom combined with ring oxygen O-5; one that includes a single carboxyl oxygen atom, and one composed of the O-3–O-4 pair of hydroxyl groups.  相似文献   

14.
The CNDO/2 method using the tight binding approximation for polymers was applied to poly(l-proline I) and poly(l-proline II). The calculations were also carried out for poly(l-alanines) and model molecules which have the same backbone geometrics as those of poly(l-prolines). The results obtained show that both forms of poly(l-proline I) and poly(l-proline II) have nearly the same energy in agreement with experimental results. From the analysis of the total energy, it was found that the intrasegment energy of poly(l-proline II) was lower than that of poly(l-proline I) while the intersegment energy of poly(l-proline I) was lower than that of poly(l-proline II). This result can be considered to correspond well with the experimental fact that poly(l-proline II) is more stable in good or polar solvents and poly(l-proline I) in poor or non-polar solvents. The analysis of the total energy of poly(l-proline) leads us to the conclusion that the α and β carbons play an important role in determining the relative stability between poly(l-proline I) and poly(l-proline II) and the γ carbon does hvae a marked effect on the electronic structures of the polymers in question. This conclusion was also confirmed by comparison of the electronic structures of poly(l-prolines) with those of poly(l-alanines) and model compounds concerned.  相似文献   

15.
Transport of l-proline into Saccharomyces cerevisiae K is mediated by two systems, one with a KT of 31 μM and Jmax of 40 nmol · s?1 · (g dry wt.)?1, the other with KT > 2.5 mM and Jmax of 150–165 nmol · s?1 · (g dry wt.)?1, The kinetic properties of the high-affinity system were studied in detail. It proved to be highly specific, the only potent competitive inhibitors being (i) l-proline and its analogs l-azetidine-2-carboxylic acid, sarcosine, d-proline and 3,4-dehydro-dl-proline, and (ii) l-alanine. The other amino acids tested behaved as noncompetitive inhibitors. The high-affinity system is active, has a sharp pH optimum at 5.8–5.9 and, in an Arrhenius plot, exhibits two inflection points at 15°C and 20–21°C. It is trans-inhibited by most amino acids (but probably only the natural substrates act in a trans-noncompetitive manner) and its activity depends to a considerable extent on growth conditions. In cells grown in a rich medium with yeast extract maximum activity is attained during the stationary phase, on a poor medium it is maximal during the early exponential phase. Some 50–60% of accumulated l-proline can leave cells in 90 min (and more if washing is done repeatedly), the efflux being insensitive to 0.5 mM 2,4-dinitrophenol and uranyl ions, to pH between 3 and 7.3, as well as to the presence of 10–100 mM unlabeled l-proline in the outside medium. Its rate and extent are increased by 1% d-glucose and by 10 μg nystatin per ml.  相似文献   

16.
Two forms of p-nitrophenyl α-D-mannosidase and p-nitrophenyl α-D-galactosidase were purified from the protein bodies of mature Lupinus angustifolius seeds. A MW of 300 000 was calculated for both α-mannosidase A and B with Km = 1.92 and 2.70 mM and activation energies of 10.9 and 10.8 kcal/mol, respectively. α-Galactosidase I and II had MWs of 70800 and 17000 with Km = 0.282 and 0.556 mM and activation energies 17.7 and 11.5 kcal/mol, respectively. The enzymes had acid pH optima and were inhibited by various metal ions, carbohydrates and glycoproteins. They were able to release free sugar from several putative natural substrate oligosaccharides and the Lupinus storage glycoprotein, α-conglutin.  相似文献   

17.
A series of aroyl- and aryl-hydrazide derivatives was prepared from d-glycero-d-gulo-heptono-1,4-lactone (1). The reactivity of the NH proton in these hydrazides, in terms of their dissociation constants (pKa), was determined from their electronic spectra, and correlated to the Hammett σ values of the substituents. Comparable reactivities of the NH protons for the compounds, and the effect of the substituent, were studied by n.m.r. spectroscopy. Decomposition of the aroylhydrazides with copper(II) sulfate or nitrous acid resulted in the regeneration of 1.  相似文献   

18.
This study investigated the anti-inflammatory effects of novel pseudotripeptides (GPE 13) as potential candidates to counteract neuroinflammation processes in Alzheimer’s disease.GPE 13 pseudotripeptides are synthetic derivatives of Gly-l-Pro-l-Glu (GPE), the N-terminal tripeptide of IGF-1, obtained through the introduction of isosteres of the amidic bond (aminomethylene unit) to increase the metabolic stability of the native tripeptide. The results showed that all synthetic derivatives possessed higher half-lives (t1/2?>?4?h) than GPE (t1/2?=?30?min) in human plasma and had good water solubility. The biological results demonstrated that GPE 13 had protective properties in several experimental models of treated THP-1 cells. Notably, the novel pseudotripeptides influenced inflammatory cytokine expression (IL-1β, IL-18, and TNF-α) in Aβ25–35-, PMA-, and LPS-treated THP-1 cells. In PMA-differentiated THP-1 macrophages, both GPE 1 and GPE 3 reduced the expression levels of all selected cyto-chemokines, even though GPE 3 showed the best neuroprotective properties.  相似文献   

19.
Although 2,3,4-tri-O-acetyl-β-D-xylopyranosyl chloride is shown by n.m.r. data to be 80 percent in the 1C4 conformation in chloroform solution, it crystallizes in the normal 4C1 conformation as shown by a three-dimensional, X-ray structure analysis. The crystals are orthorhombic, space group P212121. The phase problem was solved by the heavy-atom method. The parameters were refined to an R-value of 0.039 for 1101 structure factors. With the chlorine atom being in equatorial position in the 4C1 conformation, the C-1O-6 bond is not shortened and the C-1Cl-1 bond is not lengthened. These results are in agreement with comparable values for cis-2,3-dichloro-1,4-dioxane.  相似文献   

20.
Concentration and temperature dependences of the 1H nmr spectra of N-acetyl-L -proline N-methylamide were observed in various solvents [CCl4, CDCl3, (CD3)2CO, (CD3)2SO, H2O, and D2O]. The fraction of the cis isomer (with respect to the bond between the acetyl carbonyl carbon and prolyl nitrogen atoms) depends greatly on the solvent used; the fraction of the cis isomer is higher in polar solvents than in nonpolar solvents. It depends also on concentration and temperature in nonpolar solvents but not in polar solvents. In nonpolar solvents the trans isomer mostly exists in the γ-turn structure with an intramolecular hydrogen bond and the cis isomer tends to form molecular aggregates by intermolecular hydrogen bonds. In polar solvents both the cis and trans isomers exist in monomeric forms which interact with solvent molecules. The pH dependences of the N-methyl proton resonances indicate that the γ-turn structure of the trans isomer is present also in aqueous solution, though its population is difficult to determine. Apparent enthalpy and entropy changes for the conversion of the trans isomer to cis isomer are evaluated for various solvents. The results are discussed in terms of the intra- and intermolecular hydrogen bondings.  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