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 19 毫秒
1.
CD spectra have been measured in the vacuum-uv region to about 135 nm for cyclo(L alanyl-glycine), cyclo(L alanyl-L -alanine), and cyclo(L -prolyl-L -proline). In addition to the amide CD bands usually observed at wavelengths longer than 180 nm, a Independent systems calculations show that these intense short-wavelenght bands can be attributed to σσ* transitions of the backbone. Transitions of the amide chromophore expected at wavelenghts shorter than 180 nm cannot account for the observed CD.  相似文献   

2.
The optical spectroscopic characterization of γ‐turns in solution is uncertain and their distinction from β‐turns is often difficult. This work reports systematic ECD and vibrational circular dichroism (VCD) spectroscopic studies on γ‐turn model cyclic tetrapeptides cyclo(Ala‐β‐Ala‐Pro‐β‐Ala) ( 1 ), cyclo(Pro‐β‐Ala‐Pro‐β‐Ala) ( 2 ) and cyclo(Ala‐β‐Ala‐Ala‐β‐Ala) ( 3 ). Conformational analysis performed at the 6‐31G(d)/B3LYP level of theory using an adequate PCM solvent model predicted one predominant conformer for 1‐3 , featuring two inverse γ‐turns. The ECD spectra in ACN of 1 and 2 are characterized by a negative n→π* band near 230 nm and a positive π→π* band below 200 nm with a long wavelength shoulder. The ECD spectra in TFE of 1‐3 show similar spectra with blue‐shifted bands. The VCD spectra in ACN‐d3 of 1 and 2 show a +/?/+/? amide I sign pattern resulting from four uncoupled vibrations in the case of 1 and a sequence of two positive couplets in the case of 2 . A ?/+/+/? amide I VCD pattern was measured for 3 in TFE‐d2. All three peptides give a positive couplet or couplet‐like feature (+/?) in the amide II region. VCD spectroscopy, in agreement with theoretical calculations revealed that low frequency amide I vibrations (at ~1630 cm?1 or below) are indicative of a C7 H‐bonded inverse γ‐turns with Pro in position 2, while γ‐turns encompassing Ala absorb at higher frequency (above 1645 cm?1). Chirality, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

3.
Terminally blocked (L-Pro-Aib)n and Aib-(L-Pro-Aib)n sequential oligopeptides are known to form right-handed β-bend ribbon spirals under a variety of experimental conditions. Here we describe the results of a complete CD and ir characterization of this subtype of 310-helical structure. The electronic CD spectra were obtained in solvents of different polarity in the 260-180 nm region. The vibrational CD and Fourier transform ir (FTIR) spectra were measured in deuterochloroform solution in the amide I and amide II (1750-1500 cm?1) regions. The critical chain length for full development of the β-bend ribbon spiral structure is found to be five to six residues. Spectral effects related to concentration-induced stabilization of the structures of the longer peptides were seen in the resolution-enhanced FTIR spectra. Comparison to previous studies of (Aib)n and (Pro)n oligomers indicate that the low frequency of the amide I mode is due to the interaction of secondary and tertiary amide bonds and not to a strong difference in conformation from a regular 310-helix. © 1995 John Wiley & Sons, Inc.  相似文献   

4.
Comparative CD and Fourier transform ir (FTIR) spectroscopic data on N-Boc protected linear peptides with or without the (Pro-Gly) β-turn motif (e.g., Boc-Tyr-Pro-Gly-Phe-Leu-OH and Boc-Tyr-Gly-Pro-Phe-Leu-OH) are reported herein. The CD spectra, reflecting both backbone and aromatic contributions, were not found to be characteristic of the presence of β-turns. In the amide I region of the FTIR spectra, analyzed by self-deconvolution and curve-fitting methods, the β-turn band shewed up between 1639 and 1633 cm?1 in trifluoroethanol (TFE) but only for models containing the (Pro-Gly) core. This band war-also present in the spectra in chloroform but absent in dimethylsulfoxide. These findings, in agreement with recent ir data on cyclic models and 310-helical polypeptides and protein in D2O [see S. J. Prestrelski, D. M. Byler, and M. P. Thompson (1991), International Journal of Peptide and Protein Research, Vol. 37, pp. 508–512; H. H. Mantsch, A. Perczel. M. Hollósi, and G. D. Fasman (1992), FASEB Journal, Vol. 6, p. A341; H. H. Mantsch. A. Perczel, M. Hollósi, and G. Fasman (1992), Biopolymers. Vol. 33, pp. 201–207; S. M. Miick, G. V. Martinez, W. R. Fiori, A. P. Tedd, and G. L. Millhauser (1992). Nature, Vol. 359, pp. 653–655], suggest that the amide I band, with a major contribution from the acceptor C ? O of the 1 ← 4 intramolecular H bond of β-turns, appears near or below 1640 cm?1, rather than above 1660 cm?1. In TFE, bands between 1670 and 1660 cm?1 are mainly due to “free” carbonyls, that is, C ? O's of amides that are solvated but not involved in the characteristic H bonds of periodic secondary structures or β-turns. © 1994 John Wiley & Sons, Inc.  相似文献   

5.
The refolding kinetics of horse cyanometmyoglobin induced by concentration jump of urea was investigated by five optical probe stopped-flow methods: absorption at 422 nm, tryptophyl fluorescence at around 340 nm, circular dichroism (CD) at 222 nm, CD at 260 nm, and CD at 422 nm. In the refolding process, we detected three phases with rate constants of > 1 × 102 s?1, (4.5–9.3) S?1, and (2–5) × 10?3 s?1. In the fastest phase, a substantial amount of secondary structure (40%) is formed within the dead time of the CD stopped-flow apparatus (10.7 ms). The kinetic intermediate populated in the fastest phase is shown to capture a hemindicyanide, suggesting that a “heme pocket precursor” recognized by hemindicyanide must be constructed within the dead time. In the middle phase, most of secondary and tertiary structures, especially around the captured hemindicyanide, have been constructed. In the slowest phase, we detected a minor structural rearrangement accompanying the ligand-exchange reaction in the fifth coordination of ferric iron. We present a possible model for the refolding process of myoglobin in the presence of the heme group. © 1994 Wiley-Liss, Inc.  相似文献   

6.
Conformation of cyclo (Sar-Sar-Gly)2, cyclo(Sar)6, and cyclo(Sar-Gly-Gly)2 was investigated by nmr spectroscopy. cyclo(Sar-Sar-Gly)2, were shown to assume various conformations in dimethysulfoxide. It was attributed to the distribution of cis as well as trans Gly-Sar or Sar-Sar amide links along the peptide backbone. In particular, cyclo(Sar-Sar-Gly)2 took five or six different conformations: one or three C2-symmetric conformations and four or three asymmetric conformations, respectively. Three of nine NH resonance signals were ascribed to the internally hydrogen-bonded glycine residues. cyclo(Sar-Sar-Gly)2 and cyclo(Sar)6 showed a spectral change on the addition of alkali thiocyanates, indicating a conformational change induced by a complex formation with the alkali cations. The complex nmr spectrum due to a hybridization of different conformations changed with the salt addition into a simple nmr spectrum, suggesting a preponderence of a new, single conformation. On the basis of the spectral change, the strength for the cations binding the cyclic peptides was found to be in the order of K+ > Na+ > Rb+ > Cs+ for cyclo(Sar-Gly-Gly)2 and K+ > Rb+ > Cs+ for cyclo(Sar)6. On the other hand, cyclo(Sar-Gly-Gly)2 in dimethylsulfoxide assumed a single C2 conformation having two glycyl peptide protons shielded from solvent and the other two exposed to solvent. This conformation did not change with the salt addition. Finally, the conformations of several cyclic peptides containing the sarcosine residue such as cyclo(Sar)6 cyclo(Sar-Sar-Gly)2 cyclo(Pro-Sar-Gly)2, and cyclo (Sar-Gly-Gly)2 were compared. It appeared that proline and glycine residues reduced the conformational multiplicity of the cyclic peptide backbone, and the ability to bind alkali metal cations decreased in the above order.  相似文献   

7.
Abstract

A group of unnatural 1-(2-deoxy-β-D-ribofuranosyl)-2,4-difluorobenzenes possessing a 5-I or 5-CF3 substituent, that were originally designed as thymidine mimics, were coupled via their 5′-OH group to a cyclosaligenyl (cycloSal) ring system having a variety of C-3 substituents (Me, OMe, H). The 5′-O-cycloSal-pronucleotide concept was designed to effect a thymidine kinase-bypass, thereby providing a method for the intracellular delivery and generation of the 5′-O-monophosphate for nucleosides that are poorly phosphorylated. The 5′-O-cycloSal pronucleotide phosphotriesters synthesized in this study were obtained as a 1:1 mixture of two diastereomers that differ in configuration (S P or R P) at the asymmetric phosphorous center. The (S P)- and (R P)-diastereomers for the 5′-O-3-methylcycloSal- and 5′-O-3-methoxycycloSal derivatives of 1-(2-deoxy-β-D-ribofuranosyl)-2,4-difluoro-5-iodobenzene were separated by silica gel flash column chromatography. This class of cycloSal pronucleotide compounds generally exhibited weak cytotoxic activities in a MTT assay (CC50 values in the 10?3 to 10?4 M range), against a number of cancer cell lines (143B, 143B-LTK, EMT-6, Hela, 293), except for cyclosaligenyl-5′-O-[1′-(2,4-difluoro-5-iodophenyl)-2′-deoxy-β-D-ribofuranosyl]phosphate that was more potent (CC50 values in the 10?5 to 10?6 M range), than the reference drug 5-iodo-2′-deoxyuridine (IUDR) which showed CC50 values in the 10?3 to 10?5 M range.  相似文献   

8.
Enzymatic acylation of rutin and esculin with aromatic, aliphatic and aryl-aliphatic acids using Candida antarctica lipase in tert amyl alcohol as solvent was investigated under low water content. Whatever the acyl donor used, the conversion yields and initial rates for esculin were higher than for rutin. For a given flavonoid, the performance of these reactions depended on the acyl donor structures. For aliphatic acids, conversion yields and initial rates of both flavonoids were respectively in the ranges of 68–90% and of 9.5×10?2–72×10?2 mmol l?1 h?1. For aromatic acids, the reaction occurred only with the aryl subgroup (cinnamic, hydrocinnamic, 3,4-dihydroxyhydrocinnamic and 4-hydroxyphenyl acetic acids) and was drastically influenced by the presence of side chain and substitution patterns of the aromatic ring. Except for hydrocinnamic acid (75%, 23.4×10?2 mmol l?1 h?1), with these acids the conversion yields and initial rates were lower and in the range of 10–45% and of 0.7×10?2 to 12.1×10?2 mmol l?1 h?1. Unsaturation of the side chain of the hydrocinnamic acid decreased the esculin conversion rate from 75 to 13% and initial rate from 23.4 to 1.76×10?2 mmol l?1 h?1. The presence of hydroxyl or nitro-groups on the aromatic ring of the aryl aliphatic acid also reduced conversion yields and initial rates. Even without a spacer, the non-phenolic ring acid (quinic acid) was reactive and lead to conversion yields of about 20 and 23% respectively for rutin and esculin.  相似文献   

9.
The Miyazawa-Blout-Krimm (M-B-K) treatment of polypeptide absorption in the infrared is extended to the calculation of circular dichroism (CD), linear dichroism, and oriented CD for the amide I and amide II transitions. Matrix methods are applied to the α helix and β structures using measured values for the strengths and directions of the transition dipole moments and empirical values from M-B-K for the coupling constants. Relatively small aggregates, a 36-residue helix, and 8-chain × 4-residue β sheets, are large enough to show calculated absorption agreeing with M-B-K results, which are based on infinite lattices. In all cases the predicted CD is an approximately conservative couple. The strongest CD should appear in the α helix, Δε/ε ?± 10?3 for both transitions. The amide II transition should show moderate CD couples in both β structures, Δε/ε ? (+2 to ?1) × 10?4. The amide I transitions in β structures should show weak CD couples, Δε/ε = (+3 to ?2) × 10?5, except that the negative branch in the antiparallel structure may be detectable (Δε/ε ? ?2 × 10?4) because absorption is very low at its wavelength peak. CD on oriented samples should be enhanced over the unoriented cases, giving values as large as Δε/ε = 3 × 10?3 because particular directions of observation allow the light to avoid much of the absorption in the sample. If all three structures are considered as helices, then the larger distance of the transition dipoles from the axis in the α helix, and the orientations of the transitions in the different structures, are the factors that, in terms of our previous theoretical work [Snir and Schellman (1973) J. Phys. Chem. 77 , 1653] satisfactorily explain the calculated results. Simple dipole–dipole interaction is calculated to make a substantial contribution to the coupling between groups.  相似文献   

10.
Abstract

The interaction between glycated human serum albumin (gHSA) and folic acid (FA) was investigated by various spectroscopic techniques, such as fluorescence, circular dichroism, UV–vis absorption spectroscopy and electrophoretic light scattering technique. These methods characterize the binding properties of an albumin–folic acid system. The binding constants values (Ka) at 300 and 310 K are about 104 M?1. The standard enthalpy change (ΔH) and the standard entropy change (ΔS) were calculated to be ~?20?kJ mol?1 and ~16 J mol?1 K?1, respectively, which indicate characteristic electrostatic interactions between gHSA and folic acid. The CD studies showed that there are no significant conformational changes in the secondary structure of the protein. Moreover, the zeta potential measurements proved that under physiological conditions the gHSA–folic acid complex shows instability. No significant changes in the secondary structure of the protein and reversible drug binding are the desirable effect from pharmacological point of view.

Communicated by Ramaswamy H. Sarma  相似文献   

11.
Abstract. The influence of spines on nocturnal acid accumulation was examined for two Crassulacean acid metabolism plants, the barrel cactus Ferocactus acanthodes (Lemaire) Britton & Rose var. lecontei (Engelm.) Lindsay and the cylindropuntia Opuntia bigelovii Engelm. var. bigelovii. The ambient photo-synthetically active radiation (PAR) required for 90% saturation of nocturnal acid accumulation by O. bigelovii was 23 mol m?2 d?1 for stems that were 32% shaded by spines and 16 mol m?2 d?1 when the spines were removed. For F. acanthodes, 90% saturation occurred for an ambient PAR of 45 mol m?2 d?1 where the stem was 78% shaded by spines and 21 mol m?2 d?1 when the spines were removed. For comparison, the ambient PAR averaged 21 mol m?2 d?1 at various times of the year and stem positions for O. bigelovii fully exposed to solar irradiation. Although spines can reflect some PAR toward the stem surface, their presence decreases the PAR incident on the stem surface, thereby decreasing the nocturnal acid accumulation. For O. bigelovii, periodically removing the spines led to a 60% greater increase in stem volume over 21/2 years. The effect of variation in spine-shading on stem surface temperature was determined for F. acanthodes using computer simulations and an energy budget model. A substantial reduction in daytime stem temperatures and a smaller increase in night-time temperatures as the spine-shading increased had very little influence on nocturnal acid accumulation for both winter and summer days. Thus, the main effect of spines on the metabolism of cacti is apparently to reduce the PAR incident on the stem surface and thereby to reduce productivity.  相似文献   

12.
Abstract

The synthesis of cycloSal-FdUMP 3a-d as a new prodrug approach for FdU 1 is described. Phosphotriesters 3 release the FdUMP 2 selectively by a controlled, chemically induced tandem reaction in hydrolysis studies. The biological activity (IC50) of cycloSal-phosphotriesters 3 was evaluated in FM3A/O cells and FM3A/TK? cells.  相似文献   

13.
An nmr spectral comparison of a model cyclic pentapeptide cyclo(Gly-Pro-Gly-D-Phe-Pro) with an analogous pseudopeptide has been made. The pseudopeptide contains a ψ[CH2S] amide bond replacement at the only amide linkage that, in the model, is not involved in an intramolecular hydrogen bond. Both proton and carbon-13 nmr spectral evidence confirms the retention of β- and γ-turns in the pseudopeptide in chloroform. Characteristic chemical shifts, temperature dependence, and glycine α-resonances support this interpretation. However, evidence of a more flexible conformation involving cistrans proline isomerism is seen on addition of dimethylsulfoxide.  相似文献   

14.
For better operational control of the completely mixed oxygen activated sludge process (CMOAS), a study concerning the kinetics, performance, and operational stability of the Ramanathan-Gaudy model was conducted. Short-term experiments were conducted at various dilution rates (1/9, 1/6, 1/3, 1/1.5, and 1/1.0 hr?1) by using two recycle solids concentration values (5000 and 10,000 mg/liter). The influent substrate was an actual industrial organic wastewater (soft drink waste) and its concentration was maintained at 1000 mg/liter COD. The hydraulic recycle ratio, α, was maintained at 0.30. It was found that for CMOAS system with constant recycle cell concentration, a “steady state” with respect to reactor biological solids and effluent COD at different dilution rates could be attained. No appreciable dilute-out of reactor biological solids and substrate was observed up to the dilution rate of 1 hr?1 for both systems of different XR (5000 and 10,000 mg/liter). For the system of XR = 5000 mg/liter, except the dilution rate of hr?1, the effluent filtrate COD was lower than 100 mg/liter, the aerator biological solids concentration was about 1550 mg/liter, and the COD removal efficiency was higher than 90% for all dilution rates. For the system of XR = 10,000 mg/liter, the effluent filtrate COD was lower than 71 mg/liter, the aerator biological solids concentration was about 2750 mg/liter, and the COD removal efficiency was higher than 90% throughout all the dilution rates selection in the present study. The value of the Sludge Volume Index (SVI) was the range of 37.0 to 58.5 and provided good settleability of sludge. The sludge yield was 0.53 for the system of XR = 5000 mg/liter and 0.57 for the system of XR = 10,000 mg/liter. The carbohydrate and the protein content of the cells were 10.1–21.6% and 35.6–50.6%, respectively. For predicting the reactor biological solid and effluent COD of the CMOAS system by using the Ramanathan-Gaudy model, two sets of values for the biological kinetic constants should be considered since it provided the best fit of predicted values of the observed values. In the present study, μm = 0.4 hr?1, ks = 92 mg/liter for 1/3 ? D ? 1, and μm = 0.05 hr?1, ks = 11.1 mg/liter for 1/9 ? D < 1/3 were used to calculate the predicted values of reactor biological solid and effluent filtrate COD.  相似文献   

15.
The kinetics of double-helix formation by poly U and the complementary monomer N-6,9-dimethyladenine (m6m9A) has been measured using a new fast temperature-jump apparatus. The cooperative binding kinetics are complicated by the extensive self-association of the monomers, but a satisfactory analysis using average relaxation times was possible in terms of three different models. Application of a model which considers only monomer binding yields the upper limit for the binding rate constant of an m6m9A monomer next to an already bound monomer on a poly U strand: (2 ± 0.4) × 108 M?1sec?1. A lower limit is found by using a model which allows for binding of all m6m9A stacks to poly U with equal rate constants: (3 ± 0.3) × 107 M?1sec?1. A third model with “weighted” rate constants consistent with the data: (7.5 ± 1.0) × 107 M?1sec?1. The rate of cooperative binding of m6m9A to the trimer UpUpU has also been measured. The rate constants obtained with the trimer agree with those obtained with the polymer for each of the three models within experimental error.  相似文献   

16.
Solid-phase synthesis was used for the preparation of pyroglutamyl-histidyl-p-nitrophenylalanyl-phenylalanyl-alanyl-leucine amide (I) and glycyl-glycyl-histidyl-p-nitrophenylalanyl-phenylalanyl-alanyl-leucine amide (II), two water-soluble and sensitive chromophoric substrates of chicken pepsin, hog pepsin A, and bovine spleen cathepsin D. The kinetic constants of hydrolysis of the p-nitrophenylalanyl-phenylalanyl bond of the substrates were measured by difference spectrophotometry at 308 nm (Δ? = 860 m?1 cm?1) and by ninhydrin colorimetry (substrate I, ?570 = 2.31 × 104m?1 cm?1). The pH optimum of cleavage is 5 for the pepsins and 3.7 for cathepsin D. Since all three proteinases still have a significant activity at pH 5.5–6 a new, simple assay was designed for submicrogram quantities of pepsins in the presence of pepsinogens without interference of the latter. The method is particularly suitable for the analyses of the zymogen activation mixtures.  相似文献   

17.
 Fourier transform infrared (FTIR) spectroscopy is used to compare the thermally induced conformational changes in horse, bovine and tuna ferricytochromes c in 50 mM phosphate/0.2 M KCl. Thermal titration in D2O at pD 7.0 of the amide II intensity of the buried peptide NH protons reveals tertiary structural transitions at 54  °C in horse and at 57  °C in bovine c. These transitions, which occur well before loss of secondary structure, are associated with the alkaline isomerization involving Met80 heme-ligand exchange. In tuna c, the amide-II-monitored alkaline isomerization occurs at 35  °C, followed by a second amide II transition at 50  °C revealing a hitherto unreported conformational change in this cytochrome. Amide II transitions at 50  °C (tuna) and 54  °C (horse) are also observed during the thermal titration of the CN-ligated cytochromes (where CN displaces the Met80 ligand), but a well-defined 35  °C amide II transition is absent from the titration curve of the CNadduct of tuna c. The different mechanisms suggested by the FTIR data for the alkaline isomerization of tuna and the mammalian cytochromes c are discussed. After the alkaline isomerization, loss of secondary structure and protein aggregation occur within a 5  °C range with T m values at 74  °C (bovine c), 70  °C (horse c) and 65  °C (tuna c), as monitored by changes in the amide I′ bands. The FTIR spectra were also used to compare the secondary structures of the ferricytochromes c at 25  °C. Curve fitting of the amide I (H2O) and amide I′ (D2O) bands reveals essentially identical secondary structure in horse and bovine c, whereas splitting of the α-helical absorption of tuna c indicates the presence of less-stable helical structures. CN adduct formation results in no FTIR-detectable changes in the secondary structures of either tuna or horse c, indicating that Met80 ligation does not influence the secondary structural elements in these cytochromes. The data provided here demonstrate for the first time that the selective thermal titration of the amide II intensity of buried peptide NH protons in D2O is a powerful tool in protein conformational analysis. Received: 1 April 1999 / Accepted: 24 August 1999  相似文献   

18.
Simple approximate expressions have been derived from the theory of Zimm and Bragg for use in the analysis of experimental data on the helix-coil transition in polypeptide. On the basis of the resulting expressions practical procedures are proposed to determine two basic parameters characterizing a thermally induced transition, i.e., helix initiation parameter σ and enthalpy change for helix formation, ΔH. They have been applied to the data for poly(β-benzyl L -aspartate) (PBLA) with the result: σ = 1.6 × 10?4 and ΔH = ?450 cal/mole for PBLA in m-cresol; σ = 0.6 × 10?4 and ΔH = 260 cal/mole for PBLA in chloroform containing 5.7 vol-% of dichloroacetic acid. This result gives evidence that σ may change not only from one polypeptide to another but also for a given polypeptide in different solvents. The change in limiting viscosity number [η] accompanying the transition was measured in the same solvents. The curve of [η] versus helical content had a relatively monotonic shape for the chloroformdichloroacetic acid solutions as compared with that for the m-cresol solutions, indicating that [η] depended largely on σ. Provided that [η] is a direct measure of the mean-square radius of gyration, 〈S2〉, the results are consistent with the theoretical predictions of Nagai and of Miller and Flory for 〈S2〉.  相似文献   

19.
β-lactoglobulin (β-LG) is a member of lipocalin superfamily of transporters for small hydrophobic molecules such as retinoids. We located the binding sites of retinol and retinoic acid on β-LG in aqueous solution at physiological conditions, using FTIR, CD, fluorescence spectroscopic methods, and molecular modeling. The retinoid-binding sites and the binding constants as well as the effect of retinol and retinoic acid complexation on protein stability and secondary structure were determined. Structural analysis showed that retinoids bind strongly to β-LG via both hydrophilic and hydrophobic contacts with overall binding constants of K retinol- β -LG?=?6.4 (±?.6)?×?106?M?1 and K retinoic acid- β -LG?=?3.3 (±?.5)?×?106?M?1. The number of retinoid molecules bound per protein (n) is 1.1 (±?.2) for retinol and 1.5 (±?.3) for retinoic acid. Molecular modeling showed the participation of several amino acids in the retinoid–protein complexes with the free binding energy of ?8.11?kcal/mol for retinol and ?7.62?kcal/mol for retinoic acid. Protein conformation was altered with reduction of β-sheet from 59 (free protein) to 52–51% and a major increase in turn structure from 13 (free protein) to 24–22%, in the retinoid–β-LG complexes, indicating a partial protein destabilization.  相似文献   

20.
(1) Chitin-UDP acetylglucosaminyltransferase (E.C. 2.4.1.16., chitin synthetase) in the cell-free system from phytopathogenic fungus Piricularia oryzae, and effects of various polyoxins and related compounds on the enzyme activity were studied. Polyoxins A~M, polyoxin A derivatives, polyoxin C derivatives, 5′-amino-5′-deoxyuridine, uridine and thymidine inhibited equally the incorporation of N-acetylglucosamine (GlcNAc) from UDP-N-acetylglucosamine (UDP-GlcNAc) into chitin.

(2) Competition between the above inhibitors and UDP-GlcNAc was observed by kinetic studies. The Km for UDP-GlcNAc was determined to be 3.3 × 10?3 m and the Ki values for polyoxins A~M, except polyoxin C, were found to be in the range of 3.3 × 10?5 m to 3.4 × 10?6 m. For polyoxin C, 5′-amino-5′-deoxyuridine and uridine, the Ki values of 2.7 × 10?3 m, 8.0 × 10?3 m and 3.0 × 10?3 m were given, respectively. The inhibitor constants for other related compounds were also calculated.

(3) The values of binding affinity, ?ΔG, for formation of substrate- or inhibitor-enzyme complexes were calculated from the Km or Ki values. In addition, partial binding affinities, ?Δg, for certain moieties or groups of polyoxins were estimated from the ?ΔG. For instance, the ?ΔG values for UDP-GlcNAc and polyoxin L were 5.7 kcal/mole and 9.2 kcal/mole, respectively. And the ?Δg values for the nucleoside moiety (part I), the carbamylpolyoxamic acid moiety (part II) and the carboxyl group at C5′ position of polyoxin L were 5.2, 3.5 and 0.7 kcal/mole, respectively.

(4) From the results obtained, the mechanism of action and relation between chemical structure and competitive inhibition of chitin synthetase were discussed.

  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