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 630 毫秒
1.
A general method of unequivocally identifying and obtaining sequence information on cross-linked peptides derived by proteolytic digestion of cross-linked proteins has been developed. The method is based on isotopic labeling of alpha-amino groups with 2, 4-dinitrofluorobenzene (DNFB) coupled with electrospray ionization mass spectrometry. Proteins containing covalent cross-link(s) are reductively methylated to convert lysine residues to dimethyl lysine. The methylated protein is partially hydrolyzed and the liberated alpha-amino termini are derivatized with an equimolar mixture of DNFB and [(2)H(3)]DNFB. Dinitrophenyl (DNP)-labeled peptides may be fractionated into mono- and bis-DNP pools by chromatography on phenyl media. The bis-DNP peptides are further separated by reverse-phase HPLC and analyzed by electrospray ionization mass spectrometry. The molecular ions of cross-linked peptides are unambiguously identified as 1:2:1 triplets in the mass spectrum resulting from the binomial distribution of isotopic label in the bis-DNP derivative. Sequence information can be elucidated from the unique product ion patterns which are generated from in-source fragmentation at an elevated cone voltage. Analysis of the disulfide cross-linked peptide (VTCG)(2) was undertaken as a proof of concept and the generality of the method was demonstrated by isolating and sequencing the isopeptide bond of polyubiquitin.  相似文献   

2.
The activity of recombinant murine G9a toward lysine 9 of histone H3 was investigated. GST fusion proteins containing various lengths of the histone H3 amino-terminal tail were used as substrates in the presence of recombinant G9a enzyme and AdoMet cosubstrate. The minimal substrate methylated by G9a contained seven amino acids (TARKSTG) of the histone H3 tail. Furthermore, mutational analysis of the minimal substrate was performed to identify the amino acids essential for G9a-mediated methylation. All amino acids except Thr-11 were indispensable for the methylation reaction. Steady-state kinetic analysis of the wild-type and histone H3 point mutants, lysine 4 changed to alanine (K4A) or lysine 27 changed to alanine (K27A), with purified G9a revealed similar catalytic efficiency but a reduction in turnover number (k(cat)) from 78 to 58 h(-)(1). G9a methylated synthetic peptide substrates containing the first 13 amino acids of histone H3 efficiently, although methylation, acetylation, or mutation of proximal Lys-4 amino acids reduced Lys-9 methylation. The k(cat) for wild-type peptide substrate vs Lys-4 acetyl- or trimethyl-modified peptide were 88 and 32 h(-)(1), respectively, and the K(m) for the peptides varied from 0.6 to 2.2 muM, resulting in a large difference (15-91) in catalytic efficiency. Ser-10 or Thr-11 phosphorylation resulted in poor methylation by G9a. Immunoprecipitation of unmodified and Ser-10 and Thr-11 phosphorylated histone H3 displayed mostly Lys-4 dimethylation. Dimethylated Lys-9 was reduced in Ser-10 and Thr-11 immunoprecipitated phosphorylated histones as compared to nonphosphorylated H3. In an immunocytochemical assay, GFP fusion SUV39H1 or G9a did not colocalize with phosphorylated histone H3. Thus, Ser-10/Thr-11 phosphorylation impairs Lys-9 methylation. These data suggest that the sequence context of the modified residue affects G9a activity and the modification in the proximal amino acids influences methylation.  相似文献   

3.
2,2,2-Trifluoroethanol (TFE) is widely used to induce helix formation in peptides and proteins, but the mechanism behind this effect is still poorly understood. Several recent papers have proposed that TFE acts by selectively desolvating the peptide backbone groups of the helix state. Infrared (IR) spectroscopy of the amide I band of polypeptides can be used to probe both secondary structure and backbone solvation, making this technique well suited for addressing the effect of TFE on polypeptide conformation. In this paper, we report the IR spectra as a function of TFE concentration for an alanine-rich peptide based on the repeat (AAKAA)(n)(). The IR spectra confirm that TFE desolvates the helical state of the peptide to a greater extent than the random coil state. Moreover, using a series of specifically (13)C-labeled peptides, the precise residues desolvated in the presence of TFE were identified. The residues most desolvated by TFE are the alanines located at position i - 4 in the sequence, where i is a lysine residue. This pattern of desolvation is consistent with molecular dynamics simulations which predict strong interactions between the lysine side chain at position n and the backbone carbonyl of the alanine at position i - 4. This is the first direct spectroscopic evidence of specific desolvation of helix backbone atoms in model alanine-rich peptides.  相似文献   

4.
The methyl-accepting chemotaxis proteins (MCPs) are integral membrane proteins that undergo reversible methylation during adaptation of bacterial cells to environmental attractants and repellents. The numerous methylated forms of each MCP are seen as a pattern of multiple bands on polyacrylamide gels. We have characterized the methylation sites in MCPI by analyzing methyl-accepting tryptic peptides. At least two different tryptic peptides accept methyl esters; one methyl-accepting peptide contains methionine and lysine and may be methylated a maximum of four times. The second methyl-accepting tryptic peptide contains arginine and may be methylated twice. Base-catalyzed demethylations of tryptic peptides and analysis of the charge differences between the different methylated forms of MCPI show that MCPI molecules may be methylated a total of six times. The two methyl esters on the methyl-accepting arginine peptide appear to be preferentially methylated in most of the forms of MCPI in attractant-stimulated cells. The ability to acquire six methylations on MCPI allows the bacterial cells to adapt to a broad range of attractant and repellent concentrations.  相似文献   

5.
Konkol L  Hirai TJ  Adams JA 《Biochemistry》2000,39(1):255-262
Based on the X-ray structure of the insulin receptor kinase [Hubbard, S. R. (1997) EMBO J. 16, 5572-5581], Arg-1130 in the oncoprotein v-Fps, a nonreceptor tyrosine protein kinase, is predicted to interact with the P+1 glutamate in substrate peptides. To determine whether this residue is an important recognition element in v-Fps, Arg-1130 was substituted with leucine (R1130L) and glutamic acid (R1130E). The ability of these mutants to phosphorylate the peptide EAEIYXAIE, where X is glutamic acid, alanine, or lysine, was assessed. A comparison of the rates of peptide phosphorylation under limiting substrate concentrations (i.e., k(cat)/K(m) conditions) indicates that substrate specificity is altered by the electrostatic environment of the P+1 pocket. When the pocket displays a positive charge (Arg-1130; wild type), no charge (R1130L), or a negative charge (R1130E), v-Fps prefers to phosphorylate the glutamate peptide over the lysine peptide by a 200:1, 9:1, or 1:1 margin. While k(cat)/K(m) for the glutamate peptide is 50-fold higher for wild type compared to R1130E, k(cat)/K(m) for the lysine peptide is 3-fold higher for R1130E compared to wild type, a 150-fold change in relative substrate specificity. Analysis of the individual steps in the kinetic mechanism using viscosometric techniques indicates that the wild-type enzyme binds the glutamate peptide 3-fold better than the alanine peptide and, at least, 10-fold better than the lysine peptide. For R1130L, this margin range is reduced substantially, and for R1130E, no binding preference is observed. Nonetheless, the lysine peptide binds, at least, 4-fold better to R1130E than to wild type, and the glutamate peptide binds 3-fold poorer to R1130E than to wild type. The mutants lower the phosphoryl transfer rate by 4-30-fold for the three peptides, suggesting that Arg-1130 helps to position the tyrosine for optimum catalysis. The data indicate that a single mutation in v-Fps can alter significantly the relative substrate specificity by about 2 orders of magnitude with, at least, 50% of this effect occurring through relative changes in peptide binding affinity.  相似文献   

6.
The transmembrane PTPase HPTP beta differs from its related family members in having a single rather than a tandemly duplicated cytosolic catalytic domain. We have expressed the 354-amino acid, 41-kDa human PTP beta catalytic fragment in Escherichia coli, purified it, and assessed catalytic specificity with a series of pY peptides. HPTP beta shows distinctions from the related LAR PTPase and T cell CD45 PTPase domains: it recognizes phosphotyrosyl peptides of 9-11 residues from lck, src, and PLC gamma with Km values of 2, 4, and 1 microM, some 40-200-fold lower than the other two PTPases. With kcat values of 30-205 s-1, the catalytic efficiency, kcat/Km, of the HPTP beta 41-kDa catalytic domain is very high, up to 5.7 x 10(7) M-1 s-1. The peptides corresponding to PLC gamma (766-776) and EGFR (1,167-1,177) phosphorylation sites were used for structural variation to assess pY sequence context recognition by HPTP beta catalytic domain. While exchange of the alanine residue at the +2 position of the PLC gamma (Km of 1 microM) peptide to lysine or aspartic acid showed little or no effect on substrate affinity, replacement by arginine increased the Km 35-fold. Similarly, the high Km value of the EGFR pY peptide (Km of 104 microM) derives largely from the arginine residue at the +2 position of the peptide, since arginine to alanine single mutation at the -2 position of the EGFR peptide decreased the Km value 34-fold to 3 microM. Three thiophosphotyrosyl peptides have been prepared and act as substrates and competitive inhibitors of these PTPase catalytic domains.  相似文献   

7.
The synthetic pentadecapeptide FN-C/H II (KNNQKSEPLIGRKKT-NH(2)) has the sequence of the carboxy-terminal heparin-binding domain of module III(14) of fibronectin. Interaction of FN-C/H II with bovine lung heparin has been studied by (1)H and (23)Na NMR spectroscopy and by heparin affinity chromatography. FN-C/H II binds to heparin from pD <2 up to pD approximately 10; at higher pD, the binding decreases as the lysine side-chain ammonium groups are titrated. Na(+) counterions are displaced from the counterion condensation volume that surrounds sodium heparinate by FN-C/H II, which provides direct evidence that the binding involves electrostatic interactions. The pK(A) values for each of the five ammonium groups of FN-C/H II increase upon binding to heparin which, together with chemical shift data, indicates that the binding involves both delocalized and direct electrostatic interactions between ammonium groups of FN-C/H II and carboxylate and/or sulfate groups of heparin. NMR data also provide evidence for the direct interaction of the guanidinium group of the arginine side chain with anionic sites on heparin. The affinity of heparin for FN-C/H II and for 13 analogue peptides in which lysine and arginine residues were systematically substituted with alanine increases as the number of basic residues increases. The relative contribution of each lysine and arginine to the affinity of heparin for FN-C/H II is R(12) > K(13) > K(14) > K(1) > K(5). Nuclear Overhauser enhancement (NOE) data indicate that, while FN-C/H II is largely unstructured in aqueous solution, the bound peptide interconverts among overlapping, turn-like conformations over the L(9) - T(15) segment of the peptide. NOE data for the interaction of FN-C/H II with a heparin-derived hexasaccharide, together with the number of Na(+) ions displaced from heparin by FN-C/H II as determined by (23)Na NMR, indicates that the peptide binds to a hexasaccharide segment of heparin. Identical NMR and heparin affinity chromatography results were obtained for the interaction of FN-C/H II and its D-amino acid analogue peptide with heparin, which is of interest for the potential use of peptides as therapeutic agents for diseases in which cell adhesion plays a critical role.  相似文献   

8.
1. The p.m.r. spectra of the larger CNBr-cleavage peptides of troponin I from rabbit fast-twitch skeletal muscle corresponded largely to those of fairly flexible solution structures. 2. On addition of troponin C to each of the CNBr-cleavage peptides in turn, perturbations of side chains were noted only for peptides CN5 (residues 1-21) and CN4 (residues 96-116). 3. In the presence of Ca2+, troponin C induced perturbations of the side chains of threonine-11, alanine, isoleucine and arginine residues of peptide CN5. 4. In the presence of Ca2+, troponin C induced perturbations of the side chains of phenylalanine, lysine and leucine residues of peptide CN4. 5. Irrespective of the presence or absence of Ca2+, specific interaction with actin was observed only with peptide CN4. In this case the side chains of arginine residues were perturbed. 6. It is concluded that actin interacts with the C-terminal region of peptide CN4, whereas troponin C interacts with the N-terminal region of peptide CN4 and with peptide CN5.  相似文献   

9.
Lysine-specific murine histone H3 methyltransferase, G9a, was expressed and purified in a baculovirus expression system. The primary structure of the recombinant enzyme is identical to the native enzyme. Enzymatic activity was favorable at alkaline conditions (>pH 8) and low salt concentration and virtually unchanged between 25 and 42 degrees C. Purified G9a was used for substrate specificity and steady-state kinetic analysis with peptides representing un- or dimethylated lysine 9 histone H3 tails with native lysine 4 or with lysine 4 changed to alanine (K4AK9). In vitro methylation of the H3 tail peptide resulted in trimethylation of Lys-9 and the reaction is processive. The turnover number (k(cat)) for methylation was 88 and 32 h(-1) on the wild type and K4AK9 histone H3 tail, respectively. The Michaelis constants for wild type and K4AK9 ((K(m)(pep))) were 0.9 and 1.0 microM and for S-adenosyl-L-methionine (K(m)(AdoMet)) were 1.8 and 0.6 microM, respectively. Comparable kinetic constants were obtained for recombinant histone H3. The conversion of K4AK9 di- to trimethyl-lysine was 7-fold slower than methyl group addition to unmethylated peptide. Preincubation studies showed that G9a-AdoMet and G9a-peptide complexes are catalytically active. Initial velocity data with peptide and S-adenosyl-L-methionine (AdoMet) and product inhibition studies with S-adenosyl-L-homocysteine were performed to assess the kinetic mechanism of the reaction. Double reciprocal plots and preincubation studies revealed S-adenosyl-L-homocysteine as a competitive inhibitor to AdoMet and mixed inhibitor to peptide. Trimethylated peptides acted as a competitive inhibitor to substrate peptide and mixed inhibitor to AdoMet suggesting a random mechanism in a Bi Bi reaction for recombinant G9a where either substrate can bind first to the enzyme, and either product can release first.  相似文献   

10.
A tyrosine protein kinase (TPK-I), isolated from rat spleen [Brunati, A. M. & Pinna, L. A. (1988) Eur. J. Biochem. 172, 451-457] and recently identified as the product of the lyn oncogene [Brunati, A. M., Donella-Deana, A., Ralph, S., Marchiori, F., Borin, G., Fischer, S. & Pinna, L. A. (1991) Biochim. Biophys. Acta 1901, 123-126], is stimulated by a variety of effectors, including poly(lysine), heparin and very high NaCl concentrations. The efficacy of these compounds is variably dependent on the nature of the phosphoacceptor peptide substrates. Here we show that poly(lysine), but neither NaCl nor heparin, specifically enhances the phosphorylation efficiency of lyn TPK for the peptide EDNEYTA (src peptide). It reproduces the main autophosphorylation site of pp60c-src (Tyr416), which is entirely conserved in lyn TPK. The favourable effect of poly(lysine) is accounted for by both a dramatic drop of the Km value (70 microM versus 670 microM) and more than a threefold increase in Vmax. The effect is not so evident with a variety of different peptides, either unrelated to the src peptide (e.g. angiotensin II, AAYAA) or derived from the src peptide by single or multiple substitutions of the residues located on the N-terminal side of tyrosine. In particular, the responsiveness to poly(lysine) is dramatically reduced whenever alanine is replaced for either asparagine at position -2 or glutamic acid at position -1, either in the src heptapeptide or in its shorter derivative, the pentapeptide NEYTA. In sharp contrast, the favourable effect of 2 M NaCl, which is invariably accounted for only by an increased Vmax, is especially evident with peptides like angiotensin II and AAYAA, whose phosphorylation is less sensitive to poly(lysine) stimulation. The phosphorylation of the src peptides are actually inhibited rather than stimulated by 2 M NaCl. Consistent with this, lyn TPK autophosphorylation is also dramatically stimulated by poly(lysine) while being inhibited by 2 M NaCl. These data show that poly(lysine) deeply alters the selectivity of lyn TPK by conferring to it an enhanced activity and an especially high affinity toward peptides that reproduce the conserved autophosphorylation site of the TPK of the src family. It is suggested that endogenous compound, whose effect is mimicked by poly(lysine) in vitro, may play a relevant role in determining the specificity of lyn TPK in vivo and possibly of other TPK of the src family.  相似文献   

11.
12.
A S Acharya  B N Manjula 《Biochemistry》1987,26(12):3524-3530
The mode of derivatization of amino groups of proteins by glyceraldehyde, an aldotriose, depends on the presence or absence of reducing agent. In the presence of sodium cyanoborohydride, the Schiff base adducts of the aldehyde with the amino groups are reduced, and dihydroxypropylation of amino groups takes place (reductive mode). The reductively glycated lysine residue, N epsilon-(2,3-dihydroxypropyl)lysine, is a substituted alpha-amino alcohol. This alpha-amino alcoholic function of the derivatized lysine should be susceptible to periodate oxidation, and this oxidation is anticipated to result in the regeneration of the lysine residue. This aspect has been now investigated. Indeed, on mild periodate oxidation (15 mM periodate, 15 min at room temperature) of dihydroxypropylated ribonuclease A, nearly 95% of its N epsilon-(2,3-dihydroxypropyl)lysine residues were regenerated to lysine residues. The removal of the dihydroxypropyl groups by periodate oxidation could be accomplished within a wide pH range with little variation in the recovery of lysines. The possible usefulness of this reversible chemical modification procedure in the primary structural studies of proteins was investigated with a tryptic peptide of dihydroxypropylated streptococcal M5 protein, namely, DHP-T4. This 12-residue tryptic peptide contains one internal N epsilon-(dihydroxypropyl)lysine. The dihydroxypropylated peptide released most of its dihydroxypropyl groups on mild periodate oxidation. Redigestion of the periodate-treated peptide with trypsin generated the two expected peptides, demonstrating the generation of a trypsin-susceptible site. Reductive dihydroxypropylation of amino groups of RNase A resulted in the loss of its enzyme activity, the extent of inactivation increasing with the concentration of the glyceraldehyde used.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Autolysin-defective pneumococci continue to synthesize both peptidoglycan and teichoic acid polymers (Fischer and Tomasz, J. Bacteriol. 157:507-513, 1984). Most of these peptidoglycan polymers are released into the surrounding medium, and a smaller portion becomes attached to the preexisting cell wall. We report here studies on the degree of cross-linking, teichoic acid substitution, and chemical composition of these peptidoglycan polymers and compare them with normal cell walls. peptidoglycan chains released from the penicillin-treated pneumococci contained no attached teichoic acids. The released peptidoglycan was hydrolyzed by M1 muramidase; over 90% of this material adsorbed to vancomycin-Sepharose and behaved like disaccharide-peptide monomers during chromatography, indicating that the released peptidoglycan contained un-cross-linked stem peptides, most of which carried the carboxy-terminal D-alanyl-D-alanine. The N-terminal residue of the released peptidoglycan was alanine, with only a minor contribution from lysine. In addition to the usual stem peptide components of pneumococcal cell walls (alanine, lysine, and glutamic acid), chemical analysis revealed the presence of significant amounts of serine, aspartate, and glycine and a high amount of alanine and glutamate as well. We suggest that these latter amino acids and the excess alanine and glutamate are present as interpeptide bridges. Heterogeneity of these was suggested by the observation that digestion of the released peptidoglycan with the pneumococcal murein hydrolase (amidase) produced peptides that were resolved by ion-exchange chromatography into two distinct peaks; the more highly mobile of these was enriched with glycine and aspartate. The peptidoglycan chains that became attached to the preexisting cell wall in the presence of penicillin contained fewer peptide cross-links and proportionally fewer attached teichoic acids than did their normal counterparts. The normal cell wall was heavily cross-linked, and the cross-linked peptides were distributed equally between the teichoic acid-linked and teichoic acid-free fragments.  相似文献   

14.
15.
Solid state (2)H NMR spectroscopy was employed to study peptides related to the transmembrane domain of the human epidermal growth factor receptor, for insight into the interaction of its cytoplasmic juxtamembrane domain with the membrane surface. Since such receptors have clusters of (+)charged amino acids in this region, the effect of (-)charged phosphatidylserine at the concentration found naturally in the cytoplasmic leaflet (15 mol%) was considered. Each peptide contained 34 amino acids, which included the hydrophobic 23 amino acid stretch thought to span the membrane and a ten amino acid segment beyond the 'cytoplasmic' surface. Non-perturbing deuterium probe nuclei were located within alanine side chains in intramembranous and extramembranous portions. (2)H NMR spectra were recorded at 35 degrees C and 65 degrees C in fluid lipid bilayers consisting of (zwitterionic) 1-palmitoyl-2-oleoylphosphatidylcholine, with and without 15 mol% (anionic) phosphatidylserine. The cationic extramembranous portion of the receptor backbone was found to be highly rotationally mobile on a time scale of 10(-4)-10(-5) s in both types of membrane - as was the alpha-helical intramembranous portion. Deuterium nuclei in alanine side chains (-CD(3)) detected modest changes in peptide backbone orientation and/or dynamics related to the presence of 1-stearoyl-2-oleoylphosphatidylserine: in the case of the extramembranous portion of the peptide these seemed related to lipid charge. Temperature effects on the peptide backbone external to the membrane were qualitatively different from effects on the helical transmembrane domain - likely reflecting the different physical constraints on these peptide regions and the greater flexibility of the extramembranous domain. Effects related to lipid charge could be detected in the spectrum of CD(3) groups on the internally mobile side chain of Val(650), six residues beyond the membrane surface.  相似文献   

16.
A S-adenosylmethionine:protein-lysine N-methyltransferase (EC 2.1.1.43) has been purified from rat brain cytosol 7,080-fold with a yield of 8%, using octopus calmodulin as a substrate. It contains a lysine residue that is not fully methylated. The enzyme was purified by ammonium sulfate fractionation, Sephacryl S-200 gel filtration, and phosphocellulose and octopus calmodulin-Sepharose affinity chromatographies. Among protein substrates, it was highly specific toward octupus calmodulin. The Km values for octopus calmodulin and S-adenosyl-L-methionine were found to be 2.2 X 10(-8) M and 0.8 X 10(-6) M, respectively. The molecular weight was estimated to be 57,000 by gel filtration and the pH optimum was between 7.5 and 8.5. The enzyme was stimulated in the presence of 10(-7) M Mn2+ and 10(-4) M Ca2+. HPLC of the acid hydrolysate of methyl-3H-labeled calmodulin showed the formation of epsilon-N-mono, epsilon-N-di, and epsilon-N-trimethyllysine. Reverse-phase HPLC of tryptic peptides of the methyl-3H-labeled calmodulin demonstrated that the labeled N-methyllysine lies in the 107-126 peptide. These findings suggest that this enzyme methylated a specific lysine residue of octopus calmodulin.  相似文献   

17.
In order to facilitate a novel means for coupling proteins to metal oxides, peptides were identified from a dodecamer peptide yeast surface display library that bound a model metal oxide material, the C, A, and R crystalline faces of synthetic sapphire (alpha-Al(2)O(3)). Seven rounds of screening yielded peptides enriched in basic amino acids compared to the naive library. While the C-face had a high background of endogenous yeast cell binding, the A- and R faces displayed clear peptide-mediated cell adhesion. Cell detachment assays showed that cell adhesion strength correlated positively with increasing basicity of expressed peptides. Cell adhesion was also shown to be sensitive to buffer ionic strength as well as incubation with soluble peptide (with half maximal inhibition of cell binding at approximately 5 microM peptide). Next, dodecamer peptides cloned into yeast showed that lysine led to stronger interactions than arginine, and that charge distribution affected adhesion strength. We postulate binding to arise from peptide geometries that permit conformation alignment of the basic amino acids towards the surface so that the charged groups can undergo local electrostatic interactions with the surface oxide. Lastly, peptide K1 (-(GK)(6)) was cloned onto the c-terminus of maltose binding protein (MBP) and the resultant mutant protein showed a half-maximal binding at approximately 10(-7)-10(-6) M, which marked a approximately 500- to 1,000-fold binding improvement to sapphire's A-face as compared with wild-type MBP. Targeting proteins to metal oxide surfaces with peptide tags may provide a facile one-step alternative coupling chemistry for the formation of protein bioassays and biosensors.  相似文献   

18.
Tyrosinase inhibition by peptides may find its application in food, cosmetics or medicine. In order to identify novel tyrosinase inhibitory peptides, protein-based peptide libraries made by SPOT synthesis were used to screen for peptides that show direct interaction with tyrosinase. One of the peptide libraries studied consists of overlapping, octameric peptides derived from industrial proteins as beta-casein, alpha-lactalbumin, beta-lactoglobulin, ovalbumin, gliadin, glycinin, and beta-conglycinin. On-membrane activity staining resulted in a set of peptides that are not only able to bind to tyrosinase, but are able to inhibit tyrosinase as well. Peptides containing aspartic or glutamic acid residues usually do not bind very well to tyrosinase. Strong tyrosinase-binding peptides always contain one or more arginine residues, often in combination with phenylalanine, while lysine residues can be found equally among nonbinding peptides as well as moderate tyrosinase-binding peptides. The presence of the hydrophobic, aliphatic residues valine, alanine or leucine appears to be important for tyrosinase inhibition. Therefore, good tyrosinase inhibitory peptides preferably contain arginine and/or phenylalanine in combination with valine, alanine and/or leucine.  相似文献   

19.
A series of synthetic peptide analogs of the cardiac troponin inhibitory subunit (TN-1) phosphorylation site sequence, Arg12-Pro-Ala-Pro-Ala-Val-Arg18-Arg19-Ser20-Asp21-Arg22-Ala, have been tested as substrates for the catalytic subunit of the cyclic AMP-dependent protein kinase (EC 2.7.1.37, ATP:protein phosphotransferase). As substrates, these peptides were generally inferior to the pyruvate kinase analog peptide Leu-Arg-Arg-Ala-Ser-Leu-Gly or its COOH-terminal amide analog. Replacing Arg-19 with alanine had only a minor effect on the kinetics of phosphorylation of the TN-1 peptide analog. In contrast, replacement of Arg-22 and Arg-18 with alanine resulted in marked enhancement and reduction of the Vmax, respectively. The results of this study have demonstrated that synthetic peptide analogs of the local phosphorylation site sequences of natural substrates may differ widely in their capacity to act as substrates for this protein kinase. In the case of the TN-1 peptide analogs, the contribution of the 4 arginine residues can be distinguished in terms of their influence on the kinetics of phosphorylation.  相似文献   

20.
Archaeal flagellins are made initially as preproteins with short, positively charged leader peptides. Analysis of all available archaeal preflagellin sequences indicates that the -1 position is always held by a glycine while the -2 and -3 positions are almost always held by charged amino acids. To evaluate the importance of these and other amino acids in the leader peptides of archaeal flagellins for processing by a peptidase, Methanococcus voltae mutant FlaB2 preflagellin genes were generated by PCR and the proteins tested in a methanogen preflagellin peptidase assay that detects the removal of the leader peptide from preflagellin. When the -1 position was changed from glycine to other amino acids tested, no cleavage was observed by the peptidase, with the exception of a change to alanine at which poor, partial processing was observed. Amino acid substitutions at the -2 lysine position resulted in a complete loss of processing by the peptidase, while changes at the -3 lysine resulted in partial processing. A mutant preflagellin with a leader peptide shortened from 12 amino acids to 6 amino acids was not processed. When the invariant glycine residue present at position +3 was changed to a valine, no processing of this mutant preflagellin was observed. The identification of critical amino acids in FlaB2 required for proper processing suggests that a specific preflagellin peptidase may cleave archaeal flagellins by recognition of a conserved sequence of amino acids.  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