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 15 毫秒
1.
Integrase is one of three enzymes expressed by HIV and represents a validated target for therapy. A previous study of the diketoacid-based chemotype suggested that there are two aryl-binding domains on integrase. In this study, modifications to the indole-based diketoacid chemotype are explored. It is demonstrated that the indole group can be replaced with secondary but not tertiary (e.g., N-methyl) aniline-based amides without sacrificing in vitro inhibitory activity. The difference in activity between the secondary and tertiary amides is most likely due to the opposite conformational preferences of the amide bonds, s-trans for the secondary-amide and s-cis for the tertiary-amide. However, it was found that the conformational preference of the tertiary amide can be reversed by incorporating the amide nitrogen atom into an indoline heterocycle, resulting in very potent integrase inhibitors.  相似文献   

2.
Claisen-Schmidt condensation of 3-formyl-9-methylcarbazole with various amides of 3-aminoacetophenone afforded N-{3-[3-(9-methyl-9H-carbazol-3-yl)-acryloyl]-phenyl}-benzamide/amide derivatives. All compounds were investigated for their in vitro xanthine oxidase (XO), tyrosinase and melanin production inhibitory activity. Most of the target compounds had more potent XO inhibitory activity than the standard drug (IC(50)=4.3-5.6μM). Interestingly, compound 7q bearing cyclopropyl ring was found to be the most potent inhibitor of XO (IC(50)=4.3μM). Molecular modelling study gave an insight into its binding modes with XO. Compounds 7a, 7d, 7e, 7g, and 7k were found to be potent inhibitors of tyrosinase (IC(50)=14.01-17.52μM). These results suggest the possible use of these compounds for the design and development of novel XO and tyrosinase inhibitors.  相似文献   

3.
J D O'Neil  B D Sykes 《Biochemistry》1988,27(8):2753-2762
The coat protein of bacteriophage M13 is inserted into the inner membrane of Escherichia coli where it exists as an integral membrane protein during the reproductive cycle of the phage. The protein sequence consists of a highly hydrophobic 19-residue central segment flanked by an acidic 20-residue N-terminus and a basic 11-residue C-terminus. We have measured backbone amide hydrogen exchange of the protein solubilized in perdeuteriated sodium dodecyl sulfate using 1H nuclear magnetic resonance (NMR) spectroscopy. Direct proton exchange-out measurements in D2O at 24 degrees C were used to follow the exchange of the slowest amides in the protein. Multiple exponential fitting of the exchange data showed that these amides (29 +/- 3 at pH 4.5) exchanged in two kinetic sets with exchange rates [(1.2 +/- 0.4) x 10(-4) s-1 and (4.1 +/- 1.2) x 10(-7) s-1] that differed by more than 100-fold, the slower kinetic set being retarded 10(5)-fold relative to poly(DL-alanine). The exchange rate constant for the slowest set of amides exhibited an unusual pD dependence, being proportional to [OD-]1/2. It is shown that this is an artifact of the multiple exponential fitting of the data, and a new method of presentation of exchange data as a function of pD is introduced. Steady-state saturation-transfer techniques were also used to measure exchange. These methods showed that 15-20 amides in the protein are very stable at 55 degrees C and that about 30 amides have exchange rates retarded by at least 10(5)-fold at 24 degrees C. Saturation-transfer studies also showed that the pH dependence of exchange in the hydrophilic termini was unusual. This is explained as being due to long-range electrostatic effects arising both from the protein itself and also from the anionic detergent molecules. Hydrogen exchange studies on the products of proteinase K digestion of the protein localized the slowly exchanging amides to the hydrophobic core of the protein. Relaxation [Henry, G.D., Weiner, J.H., & Sykes, B.D. (1986) Biochemistry 25, 590-598] and solid-state NMR experiments [Leo, G.C., Colnago, L.A., Valentine, K.G., & Opella, S.J. (1987) Biochemistry 26, 854-862] have previously shown that the majority of the protein backbone is rigid on the picosecond to microsecond time scale, except for the extreme ends of the molecule which are mobile.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
Parallel synthesis of the C-terminal-modified DAPT (1) derivatives was accomplished utilizing our novel resin 7. Condensation reaction of the N-acylamino acid 10 with the amines 11a-o proceeded smoothly to give the corresponding amides 6a-o without any epimerization. Among the analogues, the benzophenonemethyl amide derivative 6o showed 30 times more potent activity than the original DAPT (1).  相似文献   

5.
C Sanders  B D Sykes  L B Smillie 《Biochemistry》1988,27(18):7000-7008
The side chain and backbone mobilities of chicken gizzard tropomyosin (TM) and its nonpolymerizable derivative have been investigated by H NMR spectroscopy and amide hydrogen exchange kinetics and compared to those of rabbit cardiac TM and its nonpolymerizable derivative. Analysis of the 300-MHz H NMR spectra of native chicken gizzard and rabbit cardiac TMs and their nonpolymerizable derivatives showed that the line widths of the aromatic and histidine residues were within a factor of 2 for all four proteins, demonstrating that the side chain mobility of these residues is similar in the different TMs. Direct proton exchange-out kinetics were determined in D2O in the pD range 1.5-3.0 at 25 degrees C by H NMR spectroscopy. Multiple exponential fitting of the exchange data indicated the presence in gizzard TM of at least three kinetically distinct classes of amide hydrogens at pD 1.7 with average population sizes of 147, 74, and 61, whose rates were retarded by a factor of 10, 10(3), and 10(5), respectively, relative to the random-coil peptide poly(DL-alanine). Measurement of the direct exchange kinetics of both rabbit cardiac and nonpolymerizable gizzard TMs showed that their rate constants and population sizes were within experimental error of those for the gizzard protein, except that the fast exchanging class for cardiac TM was increased in size while that of the nonpolymerizable gizzard TM was reduced, relative to that for gizzard TM. Comparison of the exchange-out kinetics for the cardiac and gizzard proteins at pH 2.0 and 55 degrees C, where only the two slowly exchanging amide hydrogen sets are measured, again demonstrated the similarity of their kinetic parameters.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
The reactivity of previously reported peroxo adducts [Fe(mu-O2)(mu-L)(O2CPhCy)2(1-Bu-Im)2] (1), and [Fe(mu-O2)(mu-L)(O2CPhCy)2(py)2] (2), where L is a dinucleating ligand based on the m-xylylenediamine bis(Kemp's triacid imide), toward a variety of substrates is described. These studies were performed to probe the electronic properties of 1 and 2 and evaluate their potential as selective hydrocarbon oxidants. Compound 1 is nucleophilic at -77 degrees C, reacting with phenols and carboxylic acids to liberate hydrogen peroxide, whereas the less electron-rich pyridine analogue 2 is unreactive toward both reagents. By contrast, neither reacts at -77 degrees C with electrophilic reagents such as olefins or triphenylphosphine, or with weak hydrogen atom donors such as dimethylbenzylamine. When solutions of 1 are warmed to room temperature in solvents such as THF, toluene, and cyclopentane, mixtures of alcohol and ketone products derived from the solvent are formed. A detailed investigation of cyclopentane oxidation strongly points to a radical autoxidation pathway. These results are discussed in the context of the selective hydroxylation chemistry that occurs at the carboxylate-bridged diiron centers in soluble methane monooxygenase.  相似文献   

7.
In view of the important role arginine plays in living organisms as the free amino acid and, especially, as a residue in peptides and proteins, the homologous beta-homoarginines are central in our investigations of beta-peptides (Fig. 1). The preparation of beta2-homoarginine derivatives suitably protected for solution- or solid-phase peptide syntheses is described with full experimental detail (9 and 12 in Scheme 1). The readily available Fmoc-beta3 hArg(Boc)2-OH is used for manual solid-phase synthesis of beta3-oligoarginines (on Rink amide or Rink amide AM resin) either by single amino acid coupling (Scheme 3) or, much better, by dimer-fragment coupling (Scheme 4). In this way, beta3-oligoarginine amides composed of 4, 6, 7, 8, and 10 residues, both with and without fluorescein labelling, were synthesized (Schemes 2-4), purified by preparative HPLC and identified by high-resolution mass spectrometry. The free amino acids (R)- and (S)-H-beta2 hArg-OH and (S)-H-beta3 hArg-OH were tested for their ability to function as substrates for NO synthase (iNOS); the beta3-oligoarginine amides (5, 6, and 7 residues) were tested for antibacterial (against six pathogens) and hemolytic (against rat and human erythrocytes) activities. All test results were negative: none of the free beta-homoarginines induced NO formation (Fig. 3), and there was no lysis of erythrocytes (concentrations up to 100 microM; Table 1), and no significant antibiotic activity (MIC > or = 64 microg/ml; Table 2). Cell-penetration studies with the fluorescence-labelled, peptidase-resistant beta3-oligoarginine amides were carried out with HeLa cells and human foreskin keratinocytes (HFKs). The results obtained with fluorescence microscopy are: i) the longer-chain beta-oligoarginine amides (8 and 10 residues; Figs. 4-6) enter the cells and end up in the nuclei, especially in the nucleoli, irrespective of temperature (37 degrees and 4 degrees with HFKs) or pretreatment with NaN3 (with HFKs), indicating a non-endocytotic and non-energy-dependent uptake mechanism; ii) the beta-tetraarginine derivative occupies the cell surface but does not enter the cells (with HeLa); iii) the cell-growth rate of the HFKs is not affected by a 1-microM concentration of the fluorescence-labelled beta-octaarginine amide (Fig. 7), i.e., there is no antiproliferative effect. In vivo experiments with mouse skin and the beta-octaarginine derivative show migration of the beta-peptide throughout the epidermis (Fig. 8). As a contribution to understanding the mechanism, we have also studied the behavior of fluorescence-labelled beta-octa- and beta-decaarginine amides (TFA salts) towards giant unilamellar vesicles (GUVs) built of neutral (POPC) or anionic (POPC/POPG mixtures) phospholipids: the beta-oligoarginine amides bind tightly to the surface of anionic GUVs but do not penetrate the lipid bilayer (Fig. 9) as they do with living cells. In contrast, a beta-heptapeptide FL-22, which had been used as a negative control sample for the cell-penetration experiments, entered the GUVs of negative surface charge. Thus, the mechanisms of cell and GUV-model penetration appear to be different. Finally, the possible applications and implications of the 'protein transduction' by beta-oligoarginines are discussed.  相似文献   

8.
Bovine pancreatic trypsin was treated with ethylene glycol bis(succinic acid N-hydroxysuccinimide ester). Approximately 8 of 14 lysines per trypsin molecule were modified. This derivative (EG trypsin) was more stable than native between 30 degrees and 70 degrees C: T50 values were 59 degrees C and 46 degrees C, respective. EG trypsin's half-life of 25 min at 55 degrees C was fivefold greater than native's. EG trypsin had a decreased rate of autolysis and retained more activity in aqueous mixtures of 1,4-dioxan, dimethylformamide, dimethylsulfoxide, and acetonitrile. EG trypsin had lower Km values for both amide and ester substrates; its kcat values for two amides (benzoyl-L-arginine p-nitroanilide and benzyloxycarbonyl glycyl-glycyl-arginyl-7-amino-4-methyl coumarin) increased, whereas its kcat value for an ester (thiobenzoyl benzoyloxycarbonyl-L-lysinate) decreased slightly. The specific activity (kcat/Km) of EG trypsin was increased for both amide and ester substrates. EG trypsin gave higher yields and reaction rates than native in kinetically controlled synthesis of benzoyl argininyl-leucinamide in acetonitrile and in t-butanol. Highest peptide yields occurred with EG trypsin in 95% acetonitrile, where 90% of the substrate was converted to product. No peptide synthesis occurred in 95% DMF with either form of trypsin.  相似文献   

9.
The steady-state and time-resolved fluorescence spectroscopy was applied to determine the influence of an alkyl substituent(s) (methyl or ethyl, n-propyl, iso-propyl, n-butyl, sec-butyl, or t-butyl) on amide nitrogen atom on photophysical properties of tyrosine and N-acetyltyrosine amides in water. Generally, the amide group strongly quenches the fluorescence of tyrosine, however, the size and number of substituents on amide nitrogen atom modify the quenching process only in small degree. The fluorescence intensity decays of all amides studied are bi-exponential. The contribution of both components (alphai) to the fluorescence decay undergoes irregular change. An introduction of alkyl substituent on amide nitrogen atom causes an increase of the fluorescence lifetime of tyrosine derivative compared to the unsubstituted amide for both N-acetyltyrosine and tyrosine with the protonated amino group. Calculated, basing on the fluorescence quantum yield (QY) and average lifetime, the radiative rate constants (kf) are similar, which indicates that the substituent(s) does not have substantial influence on radiative process of the deactivation of the excited state of the phenol chromophore for all compounds studied regardless the amino group status as well as the number and type of substituent (linear or branched). The comparison of the ground-state rotamer populations of tyrosine amides and N-acetyltyrosine amides with different alkyl substituent on amide nitrogen atom obtained from 1H NMR with the value of pre-exponential factors indicates that not the rotamer populations, but specific hydration of a whole molecule of the amino acid including chromophore and amino acid moiety, seems to be the main reason of the heterogenous fluorescence intensity decay of tyrosine derivatives.  相似文献   

10.
M F Jeng  S W Englander  G A El?ve  A J Wand  H Roder 《Biochemistry》1990,29(46):10433-10437
Hydrogen exchange and two-dimensional nuclear magnetic resonance (2D NMR) techniques were used to characterize the structure of oxidized horse cytochrome c at acid pH and high ionic strength. Under these conditions, cytochrome c is known to assume a globular conformation (A state) with properties resembling those of the molten globule state described for other proteins. In order to measure the rate of hydrogen-deuterium exchange for individual backbone amide protons in the A state, samples of oxidized cytochrome c were incubated at 20 degrees C in D2O buffer (pD 2.2, 1.5 M NaCl) for time periods ranging from 2 min to 500 h. The exchange reaction was then quenched by transferring the protein to native conditions (pD 5.3). The extent of exchange for 44 amide protons trapped in the refolded protein was measured by 2D NMR spectroscopy. The results show that this approach can provide detailed information on H-bonded secondary and tertiary structure in partially folded equilibrium forms of a protein. All of the slowly exchanging amide protons in the three major helices of native cytochrome c are strongly protected from exchange at acid pH, indicating that the A state contains native-like elements of helical secondary structure. By contrast, a number of amide protons involved in irregular tertiary H-bonds of the native structure (Gly37, Arg38, Gln42, Ile57, Lys79, and Met80) are only marginally protected in the A state, indicating that these H-bonds are unstable or absent. The H-exchange results suggest that the major helices of cytochrome c and their common hydrophobic domain are largely preserved in the globular acidic form while the loop region of the native structure is flexible and partly disordered.  相似文献   

11.
A novel series of tropane derivatives containing a fluorinated tertiary amino or amide at the 2β position was synthesized, labeled with the positron-emitter fluorine-18 (t(1/2)=109.8 min), and tested as potential in vivo dopamine transporter (DAT) imaging agents. The corresponding chlorinated analogs were prepared and employed as precursors for radiolabeling leading to the fluorine-18-labeled derivatives via a one-step nucleophilic aliphatic substitution reaction. In vitro binding results showed that the 2β-amino compounds 6b, 6d and 7b displayed moderately high affinities to DAT (K(i)<10nM). Biodistribution studies of [(18)F]6b and [(18)F]6d showed that the brain uptakes in rats were low. This is likely due to their low lipophilicities. Further structural modifications of these tropane derivatives will be needed to improve their in vivo properties as DAT imaging agents.  相似文献   

12.
The D-amino acid amidase-producing bacterium was isolated from soil samples using an enrichment culture technique in medium broth containing D-phenylalanine amide as a sole source of nitrogen. The strain exhibiting the strongest activity was identified as Delftia acidovorans strain 16. This strain produced intracellular D-amino acid amidase constitutively. The enzyme was purified about 380-fold to homogeneity and its molecular mass was estimated to be about 50 kDa, on sodium dodecyl sulfate polyacrylamide gel electrophoresis. The enzyme was active preferentially toward D-amino acid amides rather than their L-counterparts. It exhibited strong amino acid amidase activity toward aromatic amino acid amides including D-phenylalanine amide, D-tryptophan amide and D-tyrosine amide, yet it was not specifically active toward low-molecular-weight D-amino acid amides such as D-alanine amide, L-alanine amide and L-serine amide. Moreover, it was not specifically active toward oligopeptides. The enzyme showed maximum activity at 40 degrees C and pH 8.5 and appeared to be very stable, with 92.5% remaining activity after the reaction was performed at 45 degrees C for 30 min. However, it was mostly inactivated in the presence of phenylmethanesulfonyl fluoride or Cd2+, Ag+, Zn2+, Hg2+ and As3+ . The NH2 terminal and internal amino acid sequences of the enzyme were determined; and the gene was cloned and sequenced. The enzyme gene damA encodes a 466-amino-acid protein (molecular mass 49,860.46 Da); and the deduced amino acid sequence exhibits homology to the D-amino acid amidase from Variovorax paradoxus (67.9% identity), the amidotransferase A subunit from Burkholderia fungorum (50% identity) and other enantioselective amidases.  相似文献   

13.
An l-aminopeptidase of Pseudomonas putida, used in an industrial process for the hydrolysis of d,l-amino acid amide racemates, was purified to homogeneity. The highly l-enantioselective enzyme resembled thiol reagent-sensitive alkaline serine proteinases and was strongly activated by divalent cations. It possessed a high substrate specificity for dipeptides and alpha-H amino acid amides, e.g., l-phenylglycine amide.  相似文献   

14.
Nocardia globerula NHB-2 exhibited an intracellular acetonitrile hydrolysing activity (AHA) when cultivated in nutrient broth supplemented with glucose (10.0 g/l) and yeast extract (1.0 g/l), at pH 8.0, 30 degrees C for 21 hr. Maximum AHA was recorded in the culture containing 0.1 M of sodium phosphate buffer, (pH 8.8) at 45 degrees C for 15 min with 600 micromol of acetonitrile and resting cells of N. globerula NHB-2 equivalent to 1.0 ml culture broth. This activity was stable up to 40 degrees C and was completely inactivated at or above 60 degrees C. About five-fold increase in AHA was observed after optimization of culture and reaction conditions. Under the optimized conditions, this organism hydrolyzed various nitriles and amides such as propionitrile, benzonitrile. acetamide, and acrylamide to corresponding acids. This nitrile/amide hydrolysing activity of N. globerula NHB-2 has potential applications in enzymatic synthesis of organic acids and bioremediation of nitriles and amides contaminated soil and water system.  相似文献   

15.
The influence of steric effects on the helical geometry and the interconversion of type II to type I polyproline in water was examined by the synthesis and analysis of proline dimers and hexamers containing up to three (2S,5R)-5-tert-butylproline residues. In the dimers, the bulky 5-tert-butyl substituent was found to exert a significant influence on the local prolyl amide geometry such that the predominant trans-isomer in N-(acetyl)prolyl-prolinamide (1) was converted to 63% cis isomer in N-(acetyl)prolyl-5-tert-butylprolinamide (2) as measured by (1)H-nmr spectroscopy. Similarly, the presence of a 5-tert-butyl group on the C-terminal residue in the polyproline hexamer Ac-Pro(5)-t-BuPro-NH(2) (4) produced a local 5-tert-butylprolyl amide population containing 61% cis isomer in water. In spite of the presence of a local prolyl cis amide geometry, the downstream prolyl amides in 4 remained in the trans isomer as determined by NOESY spectroscopy. Conformational analysis by (13)C-nmr and CD spectroscopy indicated that Ac-Pro(6)-NH(2) (3) adopted the all-trans amide polyproline type II helix in water. As the amount of 5-tert-butylproline increased from one to three residues in hexamers 4-6, a gradual destabilization of the polyproline type II helical geometry was observed by CD spectroscopy in water; however, no spectrum was obtained, indicative of a complete conversion to a polyproline type I helix. The implications of these results are discussed with respect to the previously proposed theoretical mechanisms for the helical interconversion of polyproline, which has been suggested to occur by either a cooperative C- to N-terminal isomerization of the prolyl amide bonds or via a conformational intermediate composed of dispersed sequences of prolyl amide cis and trans isomers.  相似文献   

16.
Amidation of highly methoxylated citrus pectin with primary amines   总被引:3,自引:0,他引:3  
Partially amidated pectin derivatives (N-alkyl pectinamides) were prepared from highly methoxylated citrus pectin by treatment with different primary amines in methanol. The characterisation of reaction products was made by elemental analysis, photometry and diffuse reflectance FTIR spectroscopy. N-alkyl pectinamides (secondary amides) had two intense infrared bands (amide I and amide II) shifted to lower wave numbers in comparison with the corresponding bands of commercial amidated pectins (primary amides). In some cases aminolysis of HM pectin caused the appearance of infrared bands from N-substituents. Multiple Gaussian decomposition of the characteristic bands in an IR spectrum in the region of 1850–1500 cm−1 were applied for evaluation of the degrees of amidation and methylation. The aminolysis of pectins appears to be an interesting way to produce pectin derivatives with new properties.  相似文献   

17.
Bile acid amides and oxazolines were synthesized by a sequence of steps involving the reaction of the free bile acid with formic acid to yield the formyloxy derivative, preparation of the formyloxy acid chloride, condensation of the acid chloride with 2-amino-2-methyl-1-propanol to give the amide and, finally, cyclization of the amide with thionyl chloride to give the oxazoline. The oxazolines were characterized by physical constants, thin layer and gas-liquid chromatography and identified by elemental analysis and gas-liquid chromatography-mass spectrometry. Some of the bile acid oxazoline derivatives alter the activity of bacterial 7-dehydroxylases in vitro, and inhibit the growth of certain anaerobic bacteria in pure culture.  相似文献   

18.
This paper presents stable carboxypeptidase A (CPA)-glyoxyl derivatives, to be used in the controlled hydrolysis of proteins. They were produced after immobilizing-stabilizing CPA on cross-linked 6% agarose beads, activated with low and high concentrations of aldehyde groups, and different immobilization times. The CPA-glyoxyl derivatives were compared to other agarose derivatives, prepared using glutaraldehyde as activation reactant. The most stabilized CPA-glyoxyl derivative was produced using 48 h of immobilization time and high activation grade of the support. This derivative was approximately 260-fold more stable than the soluble enzyme and presented approximately 42% of the activity of the soluble enzyme for the hydrolysis of long-chain peptides (e.g., cheese whey proteins previously hydrolyzed with immobilized trypsin and chymotrypsin) and of the small substrate N-benzoylglycyl-l-phenylalanine (hippuryl-l-Phe). These results were much better than those achieved using the conventional support, glutaraldehyde-agarose. Amino acid analysis of the products of the acid hydrolysis of CPA (both soluble and immobilized) showed that approximately four lysine residues were linked on the glyoxyl agarose beads, suggesting the existence of an intense multipoint covalent attachment between the enzyme and the support. The maximum temperature of hydrolysis was increased from 50 degrees C (soluble enzyme) to 70 degrees C (most stable CPA-glyoxyl derivative). The most stable CPA-glyoxyl derivative could be efficiently used in the hydrolysis of long-chain peptides at high temperature (e.g., 60 degrees C), being able to release 2-fold more aromatic amino acids (Tyr, Phe, and Trp) than the soluble enzyme, under the same operational conditions. This new CPA derivative greatly increased the feasibility of using this protease in the production of protein hydrolysates that must be free of aromatic amino acids.  相似文献   

19.
A series of cinnamic acyl 1,3,4-thiadiazole amide derivatives (6a-10e) have been designed and synthesized, and their biological activities were also evaluated as potential antiproliferation and tubulin polymerization inhibitors. Among all the compounds, 10e showed the most potent activity in vitro, which inhibited the growth of MCF-7 and A549 cell lines with IC(50) values of 0.28 and 0.52μg/mL, respectively. Compound 10e also exhibited significant tubulin polymerization inhibitory activity (IC(50)=1.16μg/mL). Docking simulation was performed to insert compound 10e into the crystal structure of tubulin at colchicine binding site to determine the probable binding model. Based on the preliminary results, compound 10e with potent inhibitory activity in tumor growth may be a potential anticancer agent.  相似文献   

20.
Hydrogen-deuterium exchange of the primary amide, isobutyramide, was investigated as a model for asparagine and glutamine-NH2 exchange in a protein. A simple amide was chosen since the structures of several well-characterized proteins show most of these residues to be exposed to solvent. Isobutyramide-exchange data were obtained in 1:1 D2O:dioxane solutions using a near-infrared method. The rate data were strictly pseudo-first order and yielded an average of 95% exchange of the primary amide hydrogens. In analogy with secondary amides, the pD dependence of the rate constants was characteristic of specific acid and base catalysis. In addition, analysis of the rate-pD profile for isobutyramide indicated a significant uncatalyzed exchange reaction. Temperature-dependence studies of the first-order rate constants at a fixed pD yielded an apparent activation energy of 19.3 kcal/mole. Predicted half-life times for the exposed primary amide hydrogens in proteins, based on these exchange parameters, indicate that asparagine and glutamine side chains generally would contribute to the overall rate data only below 15°C and then only for approximately 1 pD unit around the point of minimum reaction velocity.  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