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.
The acid-catalyzed hydrogen exchange rate constants kH, and the base-catalyzed rate constants kOH, have been determined (in the preceding paper) for the 25 most rapidly exchanging NH groups of bovine pancreatic trypsin inhibitor. Most of these NH groups are at the protein-solvent interface. The correlation of kH, but not kOH, with the static accessibility and hydrogen bonding of the peptide carbonyl O atom indicates that the mechanism of acid catalysis in proteins involves O-protonation. Agreement between the ionic strength dependence observed for kH and kOH and the ionic strength dependence calculated for an O-protonation mechanism supports this conclusion. N-protonation for acid catalysis, as well as N-deprotonation for base catalysis, have traditionally been assumed in the mechanism of the chemical step in peptide amide proton exchange. A preference for the alternative O-protonation mechanism has far-reaching implications in the interpretation of protein hydrogen exchange kinetics. With an O-protonation mechanism, acid-catalyzed rates of surface NH groups are primarily a function of the average solvent accessibility of the carbonyl O atoms in the dynamic solution structure, while base-catalyzed rates of surface NH groups measure solvent accessibility of the peptide N. The relative dynamic accessibilities of peptide O atoms, as measured by relative values of kH (corrected for electrostatic effects), correlate with O static accessibilities in the crystal structure. A lower correlation of static accessibility of N atoms with kOH is observed for surface NH groups in peptide groups in which the carbonyl O is not hydrogen bonded. For some surface NH groups, the observed pH of minimum rate, pHmin, deviates widely from the pHmin of model compounds. This is explained as the combined result of electrostatic effects and of the differences in accessibility of the carbonyl O and N atoms that result in a change in the relative values of kH and kOH as compared to those of model peptides. A mechanism whereby exchange of interior sites is catalyzed by interactions of catalysis ions with protein surface atoms via charge transfer is suggested.  相似文献   

2.
The acid and base catalytic rate constants, kH, obs and kOH, obs and the pH at the minimum rate, pHmin, of 25 rapidly exchanging protons in bovine pancreatic trypsin inhibitor have been determined. Here we report the labeling procedure giving 1H nuclear magnetic resonance spectral resolution of seven additional rapidly exchanging NH protons and the pH dependence of their chemical shifts. Values of kH,obs kOH,obs and pHmin are given for Ala16, Gly28 and Arg53 NH groups, the only backbone amide protons with static accessibility of more than zero in the crystal structure not previously reports, and for Gly56 NH, buried at the C terminus of an alpha-helix. All four protons reported here have pH min greater than or equal to 3. Conclusions of the previous study predict that peptide protons with pHmin higher than those of model compounds have greater static accessibility of the peptide O than of the peptide N atom. The locations in the crystal structure of the four NH groups whose exchange rates are reported here are in qualitative agreement with these predictions. The ionic strength dependence of Ala16 at pH 5.5 shows a sharp increase in the exchange rate with decreasing salt concentration, as expected for base-catalyzed exchange in a positive electrostatic field.  相似文献   

3.
Solvent exchange of 18O-labeled buried water in bovine pancreatic trypsin inhibitor (BPTI), trypsin, and trypsin-BPTI complex is measured by high-precision isotope ratio mass spectrometry. Buried water is labeled by equilibration of the protein in 18O-enriched water. Protein samples are then rapidly dialyzed against water of normal isotope composition by gel filtration and stored. The exchangeable 18O label eluting with the protein in 10-300 s is determined by an H2O-CO2 equilibration technique. Exchange of buried waters with solvent water is complete before 10-15 s in BPTI, trypsin, and BPTI-trypsin, as well as in lysozyme and carboxypeptidase measured as controls. When in-exchange dialysis and storage are carried out at pH greater than or equal to 2.5, trypsin-BPTI and trypsin, but not free BPTI, have the equivalent of one 18O atom that exchanges slowly (after 300 s and before several days). This oxygen is probably covalently bound to a specific site in trypsin. When in-exchange dialysis and storage are carried out at pH 1.1, the equivalent of three to seven 18O atoms per molecule is associated with the trypsin-BPTI complex, apparently due to nonspecific covalent 18O labeling of carboxyl groups at low pH. In addition to 18O exchange of buried waters, the hydrogen isotope exchange of buried NH groups H bonded to buried waters was also measured. Their base-catalyzed exchange rate constants are on the order of NH groups that in the crystal are exposed to solvent (static accessibility greater than 0) and hydrogen-bonded main chain O, and their pH min is similar to that for model compounds. The pH dependence of their exchange rate constants suggests that direct exchange with water may significantly contribute to their observed exchange rate.  相似文献   

4.
The constant fragment of the immunoglobulin light chain (type lambda) has two tryptophyl residues at positions 150 and 187. Trp-150 is buried in the interior, and Trp-187 lies on the surface of the molecule. The hydrogen-deuterium exchange kinetics of the indole NH proton of Trp-150 were studied at various pH values at 25 degrees C by 1H nuclear magnetic resonance. Exchange rates were approximately first order in hydroxyl ion dependence above pH 8, were relatively independent of pH between pH 7 and 8, and decreased below pH 7. On the assumption that the exchange above pH 8 proceeds through local fluctuations of the protein molecule, the exchange rates between pH 7 and 8 through global unfolding were estimated. The exchange rate constant within this pH range at 25 degrees C thus estimated was consistent with that of the global unfolding of the constant fragment under the same conditions as those reported previously [Kikuchi, H., Goto, Y., & Hamaguchi, K. (1986) Biochemistry 25, 2009-2013]. The activation energy for the exchange process at pH 7.8 was the same as that for the unfolding process by 2 M guanidine hydrochloride. The exchange rates of backbone NH protons were almost the same as that of the indole NH proton of Trp-150 at pH 7.1. These observations also indicated that the exchange between pH 7 and 8 occurs through global unfolding of the protein molecule and is rate-limited by the unfolding.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
A previous comprehensive analysis of the pH dependence of native-state amide hydrogen (NH) exchange in turkey ovomucoid third domain (OMTKY3) yielded apparent opening and closing rate constants (k(op) and k(cl)) at 14 NH groups involved in global conformational changes. This analysis has been extended to 18 additional slowly exchanging NH groups. Quench-flow experiments were performed to monitor NH exchange in native OMTKY3 from neutral to very alkaline pH ( approximately 12) conditions. Above pH 10 the mechanism of exchange switched from one governed by a rapid equilibrium preceding the chemistry of exchange (i.e. EX2 exchange), to one where exchange was limited by the rate of opening (i.e. EX1 exchange). Kinetics of solvent exposure are now known for nearly all backbone NH groups in native OMTKY3, yielding rate constants that span five orders of magnitude, 0.004 to 200 s(-1).  相似文献   

6.
W Gallagher  F Tao  C Woodward 《Biochemistry》1992,31(19):4673-4680
Hydrogen exchange rate constants for the 17 slowest exchanging amide NH groups in bovine pancreatic trypsin inhibitor (BPTI) were measured in solution and in form II and form III crystals. All 17 amide hydrogens are buried and intramolecularly hydrogen bonded in the crystal structure, except Lys 41 which is buried and hydrogen bonded to a buried water. Large-scale crystallization procedures were developed for these experiments, and rate constants for both crystal and solution exchange were measured by 1H NMR spectroscopy of exchange-quenched samples in solution. Two conditions of pH and temperature, pH 9.8 and 35 degrees C, and pH 9.4 and 25 degrees C, bring two groups of hydrogens into the experimental time window (minutes to weeks). One consists of the 10 slowest exchanging hydrogens, all of which are associated with the central beta-sheet of BPTI. The second group consists of seven more rapidly exchanging hydrogens, which are distributed throughout the molecule, primarily in a loop or turn. In both groups, most hydrogens exchange more slowly in crystals, but there is considerable variation in the degree to which the exchange is depressed in crystals. Many differences observed for the more rapidly exchanging hydrogens can be attributed to local surface effects arising from intermolecular contacts in the crystal lattice. Within the slower group, however, a very large effect on exchange of Ile 18 and Tyr 35 appears to be selectively transmitted through the matrix of the molecule.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
K Akasaka  T Inoue  H Hatano  C K Woodward 《Biochemistry》1985,24(12):2973-2979
The hydrogen isotope exchange kinetics of the 10 slowest exchanging resonances in the 1H NMR spectrum of Streptomyces subtilisin inhibitor (SSI) have been determined at pH 7-11 and 30-60 degrees C. These resonances are assigned to peptide amide protons in the beta-sheet core that comprises the extensive protein-protein interface of the tightly bound SSI dimer. The core protons are atypical in that their exchange rates are orders of magnitude slower than those for all other SSI protons. When they do exchange at temperatures greater than 50 degrees C, they do so as a set and with a very high temperature coefficient. The pH dependence of the exchange rate constants is also atypical. Exchange rates are approximately first order in hydroxyl ion dependence at pH less than 8.5 and greater than 9.5 and pH independent between pH 8.5 and 9.5. The pH dependence and temperature dependence of the SSI proton exchange rates are interpreted by the two-process model [Woodward, C. K., & Hilton, B. D. (1980) Biophys. J. 32, 561-575]. The results suggest that in the average solution structure of SSI, an unusual mobility of secondary structural elements at the protein surface is, in a sense, compensated by an unusual rigidity and inaccessibility of the beta-sheet core at the dimer interface.  相似文献   

8.
Electrostatic interactions at the protein surface yield over a billion-fold range of amide hydrogen exchange rates. This range is equivalent to the maximal degree of attenuation in exchange rates that have been shown to occur for amides buried within the protein interior. Continuum dielectric analysis of Ala-Ala, Ala-Gly, Gly-Ala and trans-Pro-Ala peptide conformer acidities predicts that the relative orientation of the two neighboring peptide groups can account for a million-fold variation in hydroxide-catalyzed hydrogen exchange rates. As in previous protein studies, an internal dielectric value of 3 was found to be applicable to simple model peptides, presumably reflecting the short lifetime of the peptide anion intermediate. Despite the million-fold range in conformer acidities, the small differences in the experimental exchange rates for these peptides are accurately predicted. Ala-Ala conformers with an extended N-terminal residue and the C-terminal residue in the α conformation are predicted to account for over 60% of the overall hydrogen exchange reaction, despite constituting only 12% of the protein coil population.  相似文献   

9.
The kinetics of hydrogen-deuterium exhcange in the polypeptide chain elongation factor Tu (EF Tu) from Escherichia coli and that from Thermus thermophilus HB8 has been examined in aqueous solutions at various pH and temperatures by means of infrared absorption measurements. The free EF-Tu from E. Coli has a greater reaction rate at all pH values and at every temperature than that of the GTP-bound or GDP-bound EF-Tu. The free EF-Tu from T. thermophilus, on the other hand, has an alomst equal reaction rate to that of EF-Tu-GDP in the temperature range 38-55 degrees C. For the peptide NH groups belonging to a medium-labile kinetic class, a small but definite difference in the rate of exchange reaction was observed between EF-Tu-GDP and EF-Tu-GTP for both E. coli and T. thermophilus. For less labile peptide NH groups, on the other hand, the rate of the exchange reaction with EF-Tu-GDP from T. thermophilus is only slightly affected by the pH of the solution at 38 degrees C and 45 degrees C, while the rate constant(k) with E. coli EF-Tu-GDP is pH-dependent (log k oc pH). For T. thermophilus EF-Tu, heat stability measurements, kinetics of the rates of GDP and GTP dissociation, and circular dichroic measurements have also been made. The molecular basis for the thermostability of T. thermophilus EF-Tu is discussed.  相似文献   

10.
T Li  J E Johnson    G J Thomas  Jr 《Biophysical journal》1993,65(5):1963-1972
We describe a novel approach to investigating exchange kinetics in biological assemblies. The method makes use of a Raman multichannel analyzer coupled with a dialysis flow cell. We employ this methodology to determine exchange rates of labile hydrogens in both the packaged RNA genome and protein subunits of bean pod mottle virus (BPMV). In the BPMV assembly, which is similar to human picornaviruses, the x-ray structure indicates that about 20% of the ssRNA chain is ordered at the threefold vertices of the icosahedral capsid, although the nucleotide bases in the ordered segments are not known (Chen et al., 1989). Here, we compare exchange profiles of the native virus with those of the empty capsid, model nucleic acids and aqueous solvent to reveal the following exchange characteristics of BPMV RNA and protein: (i) Base-specific retardation of exchange is observed in the packaged RNA. (ii) Retardation is greatest for uracil residues, for which the first-order exchange rate constant (kU = 0.18 +/- 0.02 min-1) is 40% lower than that of either the H2O solvent or adenine or cytosine groups of RNA (ksolv approximately kA approximately kC = 0.30 +/- 0.02 min-1). (iii) Retardation of exchange is also observed for the guanine residues of packaged RNA. (iv) No appreciable exchange of amide NH groups of capsid subunits occurs within the time of complete exchange (t approximately 10 min) of packaged RNA or bulk solvent. Thus, the present results identify sites in both the protein subunits (amide NH) and RNA nucleotides (amino NH2 and imino NH) which are resistant to solvent-catalyzed hydrogen exchange. We propose that retardation of exchange of labile sites of the RNA nucleotides is a consequence of the organization of the RNA chromosome within the virion. Our findings support a model for BPMV in which surface and buried domains of capsid subunits are extensively and rigidly hydrogen-bonded, and in which uracil and guanine exocyclic donor groups of packaged RNA are the principal targets for subunit interaction at the threefold vertices of the capsid.  相似文献   

11.
The kinetics of the hydrogen-deuterium exchange reaction in a trypsin inhibitor (Kunitz) from soya bean have been followed by infrared absorption measurements in aqueous solutions at various temperatures and pH values. It was found that, in every case, 49% of the total peptide hydrogen atoms exchange relatively slowly. This amount corresponds to 83 peptide groups per molecule, and this is considered to be equal to the number of peptide NH groups involved in hydrogen bonds with the carbonyls of other peptide groups in the protein molecule in its native form. Each rate constant (k) determined at pH 2.75 for this category of the NH groups is in good agreement with the value expected from an idea that the breaking of the peptide-peptide hydrogen bonds takes place very slowly, and that this is the rate-determining process in the hydrogen-deuterium exchange reaction. Thus, by ultraviolet absorption measurements at 297 nm, the equilibrium constant of the native and denatured forms has been determined in the temperature range from 42 to 53.5 °C, as well as the reaction rate of reaching equilibrium from an off-equilibrium state. From these data the rate constant (k1) of the denaturation reaction is determined, and the k1 value is found to be practically equal to the hydrogen exchange rate constant (k). The Arrhenius plot of this rate constant (k) gives a straight line in the 25 to 55 °C region, and this gives a value of 48.6 kcal/mol for the activation energy of the denaturation reaction. The rate of this reaction is found to be very low at 25 °C; its half-life is about eleven days. Infrared absorption spectra observed in the amide I region suggest that the very slow denaturation of this protein is accompanied by a conformation change from an α-helix to a β-form. The number of the peptide groups involved in this αβ change is estimated to be 9 ± 3.  相似文献   

12.
In an effort to develop a more versatile quenched hydrogen exchange method for studies of peptide conformation and protein-ligand interactions, the mechanism of amide proton exchange for model peptides in DMSO-D2O mixtures was investigated by NMR methods. As in water, H-D exchange rates in the presence of 90% or 95% DMSO exhibit characteristic acid- and base-catalyzed processes and negligible water catalysis. However, the base-catalyzed rate is suppressed by as much as four orders of magnitude in 95% DMSO. As a result, the pH at which the exchange rate goes through a minimum is shifted up by about two pH units and the minimum exchange rate is approximately 100-fold reduced relative to that in D2O. The solvent-dependent decrease in base-catalyzed exchange rates can be attributed primarily to a large increase in pKa values for the NH group, whereas solvent effects on pKW seem less important. Addition of toluene and cyclohexane resulted in improved proton NMR chemical shift dispersion. The dramatic reduction in exchange rates observed in the solvent mixture at optimal pH makes it possible to apply 2D NMR for NH exchange measurements on peptides under conditions where rates are too rapid for direct NMR analysis. To test this solvent-quenching method, melittin was exchanged in D2O (pH 3.2, 12 degrees C), aliquots were quenched by rapid freezing, lyophilized, and dissolved in quenching buffer (70% DMSO, 25% toluene, 4% D2O, 1% cyclohexane, 75 mM dichloroacetic acid) for NMR analysis. Exchange rates for 21 amide protons were measured by recording 2D NMR spectra on a series of samples quenched at different times. The results are consistent with a monomeric unfolded conformation of melittin at acidic pH. The ability to trap labile protons by solvent quenching makes it possible to extend amide protection studies to peptide ligands or labile protons on the surface of a protein involved in macromolecular interactions.  相似文献   

13.
The kinetics of H/2H chemical exchange of the amide proton has been suggested as one of the tools available for investigating hydrogenbond stabilizing interactions in gangliosides.The amide proton/deuterium (NH/2H) exchange rates in GM2 ganglioside were studied by1H-NMR spectroscopy on 12 samples prepared following different procedures. In samples passed through a sodium salt Chelex-100 cation exchange resin column prior to being analysed theN-acetylneuraminic acid NH exchange occurred in less than 10 min and that of ceramide NH in 30 min. TheN-acetylgalactosamine acetamido NH exchange was slower, the half-life of the signal ranging from 15 min to 3.5 h. Contact of the Chelex-treated GM2 samples with water, through a dialysis process, modified the NH/2H exchange rate values, theN-acetylgalactosamine acetamido NH exchange becoming faster than that of ceramide NH and similar to that ofN-acetylneuraminic acid NH. Our results indicate that the deuterium/proton exchange rate strongly depends on sample preparation (ion content and minor contaminants present in water). The three-dimensional model involving theN-acetylgalactosamine acetamido NH and theN-acetylneuraminic acid carboxyl group hydrogen-bonding, which is supported by experimental evidence, cannot be confirmed by NH-exchange measurement.  相似文献   

14.
The exchange kinetics for the slowly exchanging amide hydrogens in three defensins, rabbit NP-2, rabbit NP-5, and human HNP-1, have been measured over a range of pH at 25°C using 1D and 2D NMR methods. These NHs have exchange rates 102 to 105 times slower than rates from unstructured model peptides. The observed distribution of exchange rates under these conditions can be rationalized by intramolecular hydrogen bonding of the individual NHs, solvent accessibility of the NHs, and local fluctuations in structure. The temperature dependencies of NH chemical shifts (NH temperature coefficients) were measured for the defensins and these values are consistent with the defensin structure. A comparison is made between NH exchange kinetics, NH solvent accessibility, and NH temperature coefficients of the defensins and other globular proteins. Titration of the histidine side chain in NP-2 was examined and the results are mapped to the three-dimensional structure. © 1994 Wiley-Liss, Inc.  相似文献   

15.
The azide complex of horseradish peroxidase was studied by high resolution 1H and 15N NMR spectroscopy and by the temperature-jump method. The heme peripheral methyl proton peaks and the ligand 15N resonance were resolved to show that binding of azide by horseradish peroxidase occurs only in acidic solution below pH 6.5. It was also found that the chemical exchange rate of azide with the ferric enzyme was much faster on the 1H and 15N NMR time scale. This was further substantiated by kinetics of azide binding by horseradish peroxidase where the chemical exchange rate was confirmed to be in the microseconds range at pH 5.0 and 23 degrees C. This rate is salient in usual ligand exchange reactions in hemoproteins so far reported. pH dependences of the first order association and dissociation rate constants were also studied by the temperature-jump method to suggest a strong linkage of the azide binding with a proton uptake of an amino acid residue on the enzyme. These results were compared with the case of horse metmyoglobin and were interpreted to indicate that a heme-linked ionizable group on the enzyme facilitates the fast entry of the ligand to the coordination site. A histidyl residue is a possible candidate for the ionizable group of the enzyme.  相似文献   

16.
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   

17.
To determine whether a low pH intracellular "sorting" step is required to route peptides into secretory granules, the effects of pH altering drugs on the biosynthesis and secretion of peptides by AtT-20 mouse corticotrope tumor cells and rat intermediate pituitary cells were examined. Doses of each drug maintaining normal protein synthesis and cell morphology, while obliterating the intracellular pH gradients detected by acridine orange fluorescence, were experimentally determined. Regions of the cell rich in secretory granules were localized by immunocytochemistry and were found to coincide with organelles with a low internal pH. Biosynthetic labeling experiments were coupled with immunoprecipitation and sodium dodecyl sulfate polyacrylamide gel analyses to examine the biosynthesis and secretion of corticotropin (ACTH(1-39], alpha-melanotropin, ACTH(18-39), beta-endorphin, gamma-melanotropin, alpha-amidated joining peptide, and the NH2-terminal region of pro-ACTH/endorphin. Chloroquine (20-40 microM) and a mixture of NH4Cl and methylamine (2-5 mM each) dissipated pH gradients but had no effect on the synthetic rate of pro-ACTH/endorphin, the extent and rate of precursor processing to smaller peptides, the rate of basal secretion of the various peptides, or the extent to which secretion of each of the peptides could be stimulated by secretagogues. Monensin (0.1-1 microM) had no discernible effect on intracellular pH gradients yet totally blocked proteolytic processing of pro-ACTH/endorphin. Thus, a monensin-blockable step occurs in peptide processing, presumably in the trans Golgi region; however, a low pH chloroquine-sensitive sorting step is not required for processing or for routing peptides to a stable storage form which can be released in response to secretagogues.  相似文献   

18.
The conformation of the acyclic biscystine peptide S,S'-bis(Boc-Cys-Ala-OMe) has been studied in the solid state by x-ray diffraction, and in solution by 1H- and 13C-nmr, ir, and CD methods. The peptide molecule has a twofold rotation symmetry and adopts an intramolecular antiparallel beta-sheet structure in the solid state. The two antiparallel extended strands are stabilized by two hydrogen bonds between the Boc CO and Ala NH groups [N...O 2.964 (3) A, O...HN 2.11 (3) A, and NH...O angle 162 (3) degrees]. The disulfide bridge has a right-handed conformation with the torsion angle C beta SSC beta = 95.8 (2) degrees. In solution the presence of a twofold rotation symmetry in the molecule is evident from the 1H- and 13C-nmr spectra. 1H-nmr studies, using solvent and temperature dependencies of NH chemical shifts, paramagnetic radical induced line broadening, and rate of deuterium-hydrogen exchange effects on NH resonances, suggest that Ala NH is solvent shielded and intramolecularly hydrogen bonded in CDCl3 and in (CD3)2SO. Nuclear Overhauser effects observed between Cys C alpha H and Ala NH protons and ir studies provide evidence of the occurrence of antiparallel beta-sheet structure in these solvents. The CD spectra of the peptide in organic solvents are characteristic of those observed for cystine peptides that have been shown to adopt antiparallel beta-sheet structures.  相似文献   

19.
The substrate specificities of cyclic GMP-dependent and cyclic AMP-dependent protein kinases have been compared by kinetic analysis using synthetic peptides as substrates. Both enzymes catalyzed the transfer of phosphate from ATP to calf thymus histone H2B, as well as to two synthetic peptides, Arg-Lys-Arg-Ser32-Arg-Lys-Glu and Arg-Lys-Glu-Ser36-Tyr-Ser-Val, corresponding to the amino acid sequences around serine 32 and serine 36 in histone H2B. Serine 38 in the latter peptide was not phosphorylated by either enzyme. Cyclic GMP-dependent kinase and cyclic AMP-dependent kinase catalyzed the incorporation of 1.1 and 2.0 mol of phosphate/mol of histone H2B, respectively. The phosphorylation of histone H2B, respectively. The phosphorylation of histone H2B by cyclic GMP-dependent kinase showed two distinct optima as the magnesium concentration was increased. However, the phosphorylation of either synthetic peptide by this enzyme was depressed at high magnesium concentrations. As the pH of reaction mixtures was elevated from pH 6 to pH 9, the rate of phosphorylation of Arg-Lys-Arg-Ser32-Arg-Lys-Glu by cyclic GMP-dependent kinase continually increased. Acetylation of the NH2 terminus of the peptide did not qualitatively affect this pH profile, but did increase the Vmax value of the enzyme 3-fold. The apparent Km and Vmax values for the phosphorylation of Arg-Lys-Arg-Ser32-Arg-Lys-Glu by cyclic GMP-dependent kinase were 21 microM and 4.4 mumol/min/mg, respectively. The synthetic peptide Arg-Lys-Glu-Ser36-Tyr-Ser-Val was a relatively poor substrate for cyclic GMP-dependent kinase, exhibiting a Km value of 732 microM, although the Vmax was 12 micromol/min/mg. With histone H2B as substrate for the cyclic GMP-dependent kinase, two different Km values were apparent. The Km values for cyclic AMP-dependent kinase for either synthetic peptide were approximately 100 microM, but the Vmax for Arg-Lys-Arg-Ser32-Arg-Lys-Glu was 1.1 mumol/min/mg, while the Vmax for Arg-Lys-Glu-Ser36-Tyr-Ser-Val was 16.5 mumol/min/mg. These data suggest that although the two cyclic nucleotide-dependent protein kinases have similar substrate specificities, the determinants dictated by the primary sequence around the two phosphorylation sites in histone H2B are different for the two enzymes.  相似文献   

20.
R Tuma  G J Thomas  Jr 《Biophysical journal》1996,71(6):3454-3466
The theory, design, and application of a dialysis flow cell for Raman spectroscopy are described. The flow cell permits rapid collection of Raman spectra concurrent with the efflux of small solute molecules or ions into a solution of macromolecules and is well suited to acquisition of data during hydrogen-isotope exchange reactions of biological molecules. Kinetic parameters of the device are described by a diffusion model, which accounts satisfactorily for the observed rates of efflux of deuterium oxide (K2H = 0.30 min-1), calcium ions (KCa = 0.10 min-1) and EGTA (KEGTA = 0.07 min-1). Application to the kinetics of glutamate protonation in a peptide copolymer [poly(Glu, Lys, Tyr)] shows that pH-titration rates as high as 3.3 pH units/min can be monitored. It is also shown that one can extract first-order hydrogen-isotope exchange rate constants from measured second-order exchanges by taking into account the rate of entry of 2H2O effluent into the bulk H2O solution. Deuterium exchanges of the single-stranded polyribonucleotides poly(rA) and poly(rU) and of the double-stranded RNA genome from bacteriophage phi 6 have been investigated. The measured nucleotide base exchange rates are comparable with those determined previously by other methods. The results indicate that base exchanges as fast as approximately 2 min-1 can be determined reliably with the present design. Application of the Raman flow cell to hydrogen-isotope exchange of the basic pancreatic trypsin inhibitor confirms consistency with results obtained previously on this protein by tritiation and NMR techniques.  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