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 31 毫秒
1.
It is known that the enzymatic activity of papain (EC 3.4.22.2) toward α-N-benzoyl-l-arginine p-nitroanilide can be substantially increased by hydroxynitrobenzylation of Trp-177 through reaction of the enzyme with the active site-directed reagent, 2-chloromethyl-4-nitrophenyl (N-carbobenzoxy)glycinate (S.-M. T. Chang and H. R. Horton, 1979, Biochemistry18, 1559–1563). To determine the effect of such hydroxynitrobenzylation on the nucleophilicity of the essential thiol group at the active site of the enzyme, rates of inactivation by SN2 reactions of Cys-25 with chloroacetamide and chloroacetate and by Michael addition of Cys-25 to N-ethylmaleimide were monitored. The kinetics revealed that, at pH 6.5, the reactivities of the sulfhydryl group of hydroxynitrobenzylated papain with chloroacetamide and with N-ethylmaleimide are 24 and 27% greater than those of the sulfhydryl group of native papain. At pH 7.1, the rate enhancements are 34 and 39%, respectively. These increases in reactivity of Cys-25 as an attacking nucleophile appear to account for the increased catalytic activity of hydroxnitrobenzyl-papain toward an oligopeptide substrate, α-N-benzoyl-l-phenylalanyl-l-valyl-l-arginine p-nitroanilide, and toward an ester substrate, N-carbobenzoxyglycine p-nitrophenyl ester. However, the presence of the hydroxynitrobenzyl reporter group provides substantially greater improvement (250%) in enzymatic efficiency toward α-N-benzoyl-l-arginine p-nitroanilide, apparently by blocking nonproductive binding of this substrate to the enzyme. Fluorescence changes accompanying the various chemical modifications are interpreted in terms of a charge-transfer interaction between the imidazolium ion of His-159 and the indole moiety of Trp-177 in the active form of native papain, which should help to stabilize the catalytically essential mercaptide-imidazolium ion-pair (Cys-25, His-159).  相似文献   

2.
A method is presented for purifying papain from extracts of papaya latex. The procedure involves precipitation of the extract of papaya latex with sodium chloride followed by affinity chromatography of the redissolved precipitate. Precipitation of the protein from the latex extract is necessary to separate the papain from material which interferes with the binding of papain to the affinity column. During affinity chromatography, the affinity column is overloaded to insure absence in the final product of impurities which are capable of binding to the affinity column.The papain prepared by this procedure yielded an amino acid analysis and an N-terminal amino acid analysis expected for a sample of pure papain. No Met was detected on amino acid analysis nor was the presence of N-terminal residues other than He detected. On polyacrylamide disc gel electrophoresis at pH 4.3, papain prepared by the method described in this work was indistinguishable from crystalline papain which was prepared by the method of Kimmel and Smith, and further purified by affinity chromatography. Both disc gel patterns consisted of a single band and a trailing shadow which was less than 5% of the main band. In routine spectrophotometric assays, the specific activity toward N,α-benzoyl-l-arginine ethyl ester of papain prepared by the procedure described in this work was indistinguishable from crystalline papain prepared by the method of Kimmel and Smith, and further purified by affinity chromatography. Values of 24 sec?1' and 15 mm were obtained from the turnover number and Km for the papain-catalyzed hydrolysis of N,α-benzoyl-l-arginine ethyl ester at 25 °C, pH 6.00, Γ2 0.30 using a pH stat.  相似文献   

3.
The growth of the blue-green alga Anacystis nidulans was inhibitedby relatively low concentrations of proflavine. The durationof the lag phase increased linearly with increasing proflavineconcentrations. The alga neither acquired increased toleranceto the drug during the course of six subcultures in proflavine-supplementedmedia nor was there any shortening of lag-phase or increasein final population density following repeated subculture inproflavine medium. The strain obtained after six subcultures in proflavine wasfound to have become more sensitive to streptomycin than theuntreated stock strain. A comparison of sensitivity to proflavine of two streptomycin-resistantstrains and of untreated stock strain of the alga revealed thatthe streptomycin-resistant strains were less sensitive to proflavinethan the untreated strain. The resistance of a strain resistant to 200 µg streptomycin/mlwas lost after it had been cultured once in streptomycin-freebasal medium that had been supplemented with 1 µg proflavineµml.The results obtained provide some indirect evidence for theexistence of episome in A. nidulans. 1Present temporary address: Division of Biosciences, NationalResearch Council, Ottawa-2, Canada.  相似文献   

4.
A new reagent, 6,6′-diselenobis-(3-nitrobenzoic acid) (DSNB) has been synthesized and is shown to be useful for quantitative estimation of sulfhydryl groups in proteins. This reagent, which is a selenium analog of Ellman's reagent, reacts specifically and quantitatively with thiol groups of proteins to yield a selenenyl sulfide and the dianion of 3-nitro-6-selenobenzoic acid. The molar absorption coefficient of the chromophoric dianion is 10,000 at 432 nm in dilute aqueous solutions. The titration can best be performed at pH 8.20 where >98% of 3-nitro-6-selenobenzoic acid is in the form of the intensely colored dianion. Sulfhydryl content determinations by this reagent of reduced and denatured ribonuclease, reduced and denatured lysozyme, native papain, and native and denatured thymidylate synthetase are compared with those from corresponding 5,5′-dithiobis-(2-nitrobenzoic acid) (DTNB) titrations. The reagent was found to inactivate thymidylate synthetase, an enzyme with essential sulfhydryl groups. Unlike DTNB which undergoes alkaline decomposition of pH values greater than 9, DSNB was found to be stable to hydrolysis, even in 0.05 m NaOH.  相似文献   

5.
Both Escherichia coli B and a proflavine-resistant mutant, E. coli B/Pr, took up the same amounts of proflavine when suspended in buffer containing the dye. In growth media, however, sensitive cells took up more proflavine than did resistant cells. Adding growth media or any one of several constituents of these media, including amino acids, glycerol, pyruvic acid, and metabolizable sugars, to resistant cells that had taken up proflavine in buffer caused them to lose the dye, but had less or no effect on sensitive cells. Certian salts caused an equal release of proflavine from resistant and sensitive cells. Proflavine released from resistant cells by glucose was not changed chemically. The effects of temperature and metabolic inhibitors suggest that proflavine uptake is a passive process but that its release may be an active one, dependent on metabolism. Glucose had more effect on the proflavine binding of E. coli B grown in a minimal medium than on that of cells grown in a complex medium. E. coli B was less susceptible to proflavine when growing in a minimal medium. The effects of other synthetic media on proflavine susceptibility of E. coli B were also studied. Deoxyribonucleic acid and envelopes from sensitive and resistant cells bound the same amounts of proflavine, and no difference was seen in the site of dye binding when sensitive and resistant cells that had taken up proflavine in buffer were sonically broken and fractionated. The results suggest that sensitive and resistant cells are equally permeable to proflavine but differ in the ease with which metabolites cause them to release bound proflavine. So far, however, these differences do not account completely for the ability of resistant cells to grow in high proflavine concentrations.  相似文献   

6.
In resting grains of Triumph barley (Hordeum vulgare L. cv Triumph) about 40% of the β-amylase could be extracted with a saline solution, the remaining 60% being in a bound form. During seedling growth (20°C), the bound form was released mainly between days 1 and 3. When a preparation containing bound β-amylase was incubated with an extract made of endosperms separated from germinating grains, release of bound β-amylase took place and could be studied in vitro. The release was almost completely prevented by leupeptin and antipain, specific inhibitors of a group of SH-proteinases, but it was not inhibited by pepstatin A or EDTA, which inhibit some other barley proteinases. It is thus very likely that in a whole grain, at least the bulk of the bound β-amylase is released by the proteolytic action of one or several SH-proteinases. When the bound β-amylase was released by papain, its molecular weight was about 5000 daltons smaller than that of β-amylase released by dithiothreitol. This indicates that the release is due to removal of a sequence of β-amylase itself. A similar decrease in size took place during seedling growth. Bound β-amylase showed some activity against native starch and it hydrolyzed maltotetraose at a rate that was about 70% of the rate the same amount of bound β-amylase gave after release. Bound β-amylase is thus not inactive and it is likely that the slower rate of hydrolysis is due to steric hindrances which prevent substrates from reaching the active site.  相似文献   

7.
A method is described of using photography to measure the concentrations of a small ligand (proflavine) above and below the boundary of a macromolecule (DNA, both native and denatured) sedimenting in the ultracentrifuge. The measurements are used to determine the extent of the binding of proflavine to DNA, and the results compared with those obtained by a spectrophotometric method. The results obtained by the two methods agree within 10%, thus validating the spectrophotometric technique under these conditions. The variation of the sedimentation coefficient with the extent of binding of proflavine was also studied. The results are discussed in relation to previously observed changes in the viscosity of the solutions.  相似文献   

8.
Targeting papain family cysteine proteases is one of the novel strategies in the development of chemotherapy for a number of diseases. Novel cysteine protease inhibitors derived from 1-pyridylimidazo[1,5-a]pyridine representing pharmacologically important class of compounds are being reported here for the first time. The derivatives were initially designed and screened in silico by molecular docking studies against papain to explore the possible mode of action. The molecular interaction between the compounds and cysteine protease (papain) was found to be very similar to the interactions observed with the respective epoxide inhibitor (E-64c) of papain. Subsequently, compounds were synthesized to validate their efficacy in wet lab experiments. When characterized kinetically, these compounds show their Ki and IC50 values in the range of 13.75 to 99.30 µM and 13.40 to 96.50 µM, respectively. The thermodynamics studies suggest their binding with papain hydrophobically and entropically driven. These inhibitors also inhibit the growth of clinically important different types of Gram positive and Gram negative bacteria having MIC50 values in the range of 0.6–1.4 µg/ml. Based on Lipinski’s rule of Five, we also propose these compounds as potent antibacterial prodrugs. The most active antibacterial compound was found to be 1-(2-pyridyl)-3-(2-hydroxyphenyl)imidazo[1,5-a]pyridine (3a).  相似文献   

9.
In papain-treated rat renal brush-border membrane vesicles, cystine uptake was enhanced under sodium gradient conditions. This effect was not observed when sodium was equilibrated across the vesicle membrane or when sodium was completely absent from the incubation medium. The increased rate of cystine uptake occurred within the first two minutes of incubation and coincided with the period of increased flux of sodium known to occur after papain treatment. Under sodium gradient conditions, the Vmax of cystine uptake by treated vesicles was 65% greater while the Km was 25% lower than the value observed in untreated membranes. The increased cystine uptake after papain treatment occurred when medium cystine was in the electroneutral form. In the absence of a sodium gradient, cystine uptake by control membranes was insensitive to changes in membrane potential and this was unaltered after papain treatment. Exposure of the membranes to papain also resulted in a profound decrease in cystine binding which occurs in native membranes incubated with cystine. The fact that cystine uptake is unchanged under sodium equilibration and even enhanced under sodium gradient conditions suggests that the component of cystine binding is not essential for cystine transport and may represent non-specific binding to membrane proteins.  相似文献   

10.
DeJong ES  Chang CE  Gilson MK  Marino JP 《Biochemistry》2003,42(26):8035-8046
Rev is an essential regulatory HIV-1 protein that binds the Rev responsive element (RRE) within the env gene of the HIV-1 RNA genome, activating the switch between viral latency and active viral replication. Previously, we have shown that selective incorporation of the fluorescent probe 2-aminopurine (2-AP) into a truncated form of the RRE sequence (RRE-IIB) allowed the binding of an arginine-rich peptide derived from Rev and aminoglycosides to be characterized directly by fluorescence methods. Using these fluorescence and nuclear magnetic resonance (NMR) methods, proflavine has been identified, through a limited screen of selected small heterocyclic compounds, as a specific and high-affinity RRE-IIB binder which inhibits the interaction of the Rev peptide with RRE-IIB. Direct and competitive 2-AP fluorescence binding assays reveal that there are at least two classes of proflavine binding sites on RRE-IIB: a high-affinity site that competes with the Rev peptide for binding to RRE-IIB (K(D) approximately 0.1 +/- 0.05 microM) and a weaker binding site(s) (K(D) approximately 1.1 +/- 0.05 microM). Titrations of RRE-IIB with proflavine, monitored using (1)H NMR, demonstrate that the high-affinity proflavine binding interaction occurs with a 2:1 (proflavine:RRE-IIB) stoichiometry, and NOEs observed in the NOESY spectrum of the 2:1 proflavine.RRE-IIB complex indicate that the two proflavine molecules bind specifically and close to each other within a single binding site. NOESY data further indicate that formation of the 2:1 proflavine.RRE-IIB complex stabilizes base pairing and stacking within the internal purine-rich bulge of RRE-IIB in a manner analogous to what has been observed in the Rev peptide.RRE-IIB complex. The observation that proflavine competes with Rev for binding to RRE-IIB by binding as a dimer to a single high-affinity site opens the possibility for rational drug design based on linking and modifying it and related compounds.  相似文献   

11.
Catalytic and inhibitor binding properties of bovine α-chymotrypsin, in which the Met-192 residue has been converted by treatment with chloramine T to the sulfoxide derivative (Met(O)192 α-chymotrypsin), have been examined relative to the native enzyme (α-chymotrypsin), between pH 4.5 and 8.0 (μ = 0.1), and/or 5.0°C and 40.0°C. Values of kcat, k+2 and/or k+3 for the hydrolysis of all the substrates examined (i.e., tMetAcONp, ZAlaONp, ZLeuONp, ZLysONp and ZTyrONp) catalyzed by native and Met(O)192 α-chymotrypsin are similar, as well as values of Km for the hydrolysis of ZLeuONp, ZLysONp and ZTyrONp. On the other hand, Ks and Km values for the hydrolysis of ZAlaONp and tMetAcONp are decreased by about 5-fold. Met-192 oxidation does not affect the kinetic and thermodynamic parameters for the (de)stabilization of the complex formed between the proteinase and the bovine basic pancreatic trypsin inhibitor. On the other hand, the recognition process between between α-chymotrypsin and the recombinant proteinase inhibitor eglin c from the leech Hirudo medicinalis is influenced by the oxidation event. Considering known molecular models, the observed catalytic and inhibitor binding properties of native and Met(O)192 α-chymotrypsin were related to the inferred stereochemistry of the proteinase-substrate and proteinase-inhibitor contact region(s).  相似文献   

12.
Acridine orange and proflavine form complexes with the dinucleoside monophosphate, 5-iodocytidylyl(3′–5′)guanosine. The acridine orange-iodoCpG2 crystals are monoclinic, space group P21, with unit cell dimensions a = 14.36 A?, b = 19.64 A?, c = 20.67 A?, β = 102.5 °. The proflavine-iodoCpG crystals are monoclinic, space group C2, with unit cell dimensions a = 32.14 A?, b = 22.23 A?, c = 18.42 A?, β = 123.3 °. Both structures have been solved to atomic resolution by Patterson and Fourier methods, and refined by full matrix least-squares.Acridine orange forms an intercalative structure with iodoCpG in much the same manner as ethidium, ellipticine and 3,5,6,8-tetramethyl-N-methyl phenanthrolinium (Jain et al., 1977, Jain et al., 1979), except that the acridine nucleus lies asymmetrically in the intercalation site. This asymmetric intercalation is accompanied by a sliding of base-pairs upon the acridine nucleus and is similar to that observed with the 9-aminoacridine-iodoCpG asymmetric intercalative binding mode described in the previous papers (Sakore et al., 1977, Sakore et al., 1979). Basepairs above and below the drug are separated by about 6.8 Å and are twisted about 10 °; this reflects the mixed sugar puckering pattern observed in the sugar-phospate chains: C3′ endo (3′–5′) C2′ endo (i.e. each cytidine residue has a C3′ endo sugar comformation, while each guanosine residue has a C2′ endo sugar conformation), alterations in glycosidic torsional angles and other small but significant conformational changes in the sugar-phosphate backbone.Proflavine, on the other hand, demonstrates symmetric intercalation with iodoCpG. Hydrogen bonds connect amino groups on proflavine with phosphate oxygen atoms on the dinucleotide. In contrast to the acridine orange structure, base-pairs above and below the intercalative proflavine molecule are twisted about 36 °. The altered magnitude of this angular twist reflects the sugar puckering pattern that is observed: C3′ endo (3′–5′) C3′ endo. Since proflavine is known to unwind DNA in much the same manner as ethidium and acridine orange (Waring, 1970), one cannot use the information from this model system to understand how proflavine binds to DNA (it is possible, for example, that hydrogen bonding observed between proflavine and iodoCpG alters the intercalative geometry in this model system).Instead, we propose a model for proflavine-DNA binding in which proflavine lies asymmetrically in the intercalation site (characterized by the C3′ endo (3′–5′) C2′ endo mixed sugar puckering pattern) and forms only one hydrogen bond to a neighboring phosphate oxygen atom. Our model for proflavine-DNA binding, therefore, is very similar to our acridine orange-DNA binding model. We will describe these models in detail in this paper.  相似文献   

13.
Proflavine formed a complex with transforming deoxyribonucleic acid (DNA) from Haemophilus influenzae, with optimal formation at a ratio of proflavine to DNA of 0.06. The rate of dissociation of the complex by dialysis increased in the order: native, denatured, renatured DNA. The transforming activity of the DNA was reduced by its interaction with proflavine. This inactivation was dependent on the physical state of the DNA, the proflavine concentration, and the temperature. DNA that had been denatured and renatured was most sensitive; native DNA was much less sensitive. The inactivation remained after dialysis and was stable to prolonged storage. It is concluded that the inactivation of transforming DNA by proflavine takes place by a mechanism different from that of DNA-proflavine complex formation.  相似文献   

14.
Fluorescent quantum dots (QDs) have been widely applied in biological and biomedical areas, but relatively little is known about the interaction of QDs with some natural enzymes. Herein, the interactions between 3-mercaptopropionic acid-capped CdTe QDs (MPA-QDs) and papain were systematically investigated by UV–Vis absorption spectra, fluorescence spectra and circular dichroism (CD) spectra under the physiological conditions. The fluorescence spectra results indicated that MPA-QDs quenched the fluorescence intensity of papain. The modified Stern–Volmer quenching constant K a at different temperatures and the corresponding thermodynamic parameters ΔH, ΔG and ΔS were also calculated. The binding of MPA-QDs and papain is a result of the formation of QDs-papain complex and the electrostatic interactions play a major role in stabilizing the complex. The CD technique was further used to analyze the conformational changes of papain induced by MPA-QDs and the results indicated that the biological activity of papain was affected by MPA-QDs dramatically.  相似文献   

15.
A series of proflavine derivatives for use to further image Aβ amyloid deposits were synthesized and characterized. Aged 3xTg-AD (23 months old) mice hippocampus sections incubated with these derivatives revealed preferential labeling of amyloid plaques. Furthermore an in vitro binding study showed an inhibitory effect, although moderate, of these compounds on Aβ40 fibril formation. This study highlights the potential of proflavine as a molecular scaffold for designing new Aβ imaging agents, its native fluorescence allowing in vitro neuropathological staining in AD damaged brain sections.  相似文献   

16.
Binding of proflavine to calf thymus DNA is investigated by differential scanning calorimetry and spectrophotometry. It is shown that proflavine interacts with DNA by three binding modes. At high DNA—ligand concentration ratios (P/D), proflavine prefers to intercalate into GC-sites but can also insert into other sites. At low P/D ratios, proflavine interacts with DNA by the external binding mode. The parameters of proflavine-DNA complexation have been calculated from spectrophotometric concentration dependences. Thermodynamic parameters of DNA melting have been calculated from differential scanning calorimetry data.  相似文献   

17.
The results of an investigation on the interaction of proflavine and of ethidium bromide with DNA (calf thymus) in dilute aqueous solution are reported. The binding of the two dyes by DNA has been studied by means of microcalorimetric and of equilibrium dialysis measurements. Data on the thermodynamics of dimerization of both proflavine and ethidium bromide in aqueous solution obtained on the basis of spectroscopic and/or calorimetric experiments are also reported.The enthalpy data show that dye-dimerization and dye “strong” interaction with DNA are energetically favourable and quite similar while only in the latter case the phenomenon is also entropy driven. This is taken as further evidence in support of the concept that “strong” interaction-of both proflavine and ethidium bromide with DNA means dye molecules intercalation into the native, double helical structure of the biopolymer.  相似文献   

18.
Comparative data on the properties of four thiol proteinase inhibitors, and of four serine proteinase inhibitors (two subtilisin and two trypsin inhibitors) isolated from seeds of Vigna are presented. They were similar in their molecular weights (5000–15,000) and dissociation constants (10?8–10?9m). The range of isoelectric points of the thiol proteinase inhibitors was 6.5 to 10.6, and of the serine proteinase inhibitors was 5.0 to 5.9. The amino acid compositions of one papain isoinhibitor, one of subtilisin, and one of trypsin are presented. Papain inhibitor A1 and subtilisin inhibitor 2a were low in cystine. All of the inhibitors were stable upon heating to 80 °C for 5 min at low pH. The subtilisin inhibitor did not bind to catalytically inactive subtilisin derivatives, whereas the papain inhibitor was stoichiometrically bound to the Hg or thioacetamide derivatives of papain. Incubation of the subtilisin inhibitor with catalytic amounts of subtilisin led to the formation of a modified form with the same inhibitor activity as the native inhibitor but with a different electrophoretic mobility. There was no indication of a similar modification of the papain inhibitor by papain. Separate sites are present on the trypsin-chymotrypsin inhibitors for trypsin and chymotrypsin. The papain inhibitors have the same binding sites for papain and ficin.  相似文献   

19.
Sun Z  Henson CA 《Plant physiology》1990,94(1):320-327
The initial hydrolysis of native (unboiled) starch granules in germinating cereal kernels is considered to be due to α-amylases. We report that barley (Hordeum vulgare L.) seed α-glucosidases (EC 3.2.1.20) can hydrolyze native starch granules isolated from barley kernels and can do so at rates comparable to those of the predominant α-amylase isozymes. Two α-glucosidase charge isoforms were used individually and in combination with purified barley α-amylases to study in vitro starch digestion. Dramatic synergism, as much as 10.7-fold, of native starch granule hydrolysis, as determined by reducing sugar production, occurred when high pl α-glucosidase was combined with either high or low pl α-amylase. Synergism was also found when low pl α-glucosidase was combined with α-amylases. Scanning electron micrographs revealed that starch granule degradation by α-amylases alone occurred specifically at the equatorial grooves of lenticular granules. Granules hydrolyzed by combinations of α-glucosidases and α-amylases exhibited larger and more numerous holes on granule surfaces than did those granules attacked by α-amylase alone. As the presence of α-glucosidases resulted in more areas being susceptible to hydrolysis, we propose that this synergism is due, in part, to the ability of the α-glucosidases to hydrolyze glucosidic bonds other than α-1,4- and α-1,6- that are present at the granule surface, thereby eliminating bonds which were barriers to hydrolysis by α-amylases. Since both α-glucosidase and α-amylase are synthesized in aleurone cells during germination and secreted to the endosperm, the synergism documented here may function in vivo as well as in vitro.  相似文献   

20.
Kinases play central roles in many cellular processes, transferring the terminal phosphate groups of nucleoside triphosphates (NTPs) onto substrates. In the absence of substrates, kinases can also hydrolyse NTPs producing NDPs and inorganic phosphate. Hydrolysis is usually much less efficient than the native phosphoryl transfer reaction. This may be related to the fact that NTP hydrolysis is metabolically unfavorable as it unproductively consumes the cell's energy stores. It has been suggested that substrate interactions could drive changes in NTP binding pocket, activating catalysis only when substrates are present. Structural data show substrate-induced conformational rearrangements, however there is a lack of corresponding functional information. To better understand this phenomenon, we developed a suite of isothermal titration calorimetry (ITC) kinetics methods to characterize ATP hydrolysis by the antibiotic resistance enzyme aminoglycoside-3′-phosphotransferase-IIIa (APH(3′)-IIIa). We measured Km, kcat, and product inhibition constants and single-turnover kinetics in the presence and absence of non-substrate aminoglycosides (nsAmgs) that are structurally similar to the native substrates. We found that the presence of an nsAmg increased the chemical step of cleaving the ATP γ-phosphate by at least 10- to 20-fold under single-turnover conditions, supporting the existence of interactions that link substrate binding to substantially enhanced catalytic rates. Our detailed kinetic data on the association and dissociation rates of nsAmgs and ADP shed light on the biophysical processes underlying the enzyme's Theorell-Chance reaction mechanism. Furthermore, they provide clues on how to design small-molecule effectors that could trigger efficient ATP hydrolysis and generate selective pressure against bacteria harboring the APH(3′)-IIIa.  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