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.
The major unfolded form of ribonuclease A is known to show well-populated structural intermediates transiently during folding at 0°–10°C. We describe here how the exchange reaction between D2O and peptide NH protons can be used to trap folding intermediates. The protons protected from exchange during folding can be characterized by 1H-nmr after folding is complete. The feasibility of using 1H-nmr to resolve a set of protected peptide protons is demonstrated by using a specially prepared sample of ribonuclease S in D2O in which only the peptide protons of residues 7–14 are in the 1H-form. All eight of these protected peptide protons are H-bonded. Resonance assignments made on isolated peptides containing these residues have been used to identify the protected protons. Other sets of protected protons trapped in the 1H-form can also be isolated by differential exchange, using either ribonuclease A or S. Earlier model compound studies have indicated that H-bonded folding intermediates should be unstable in water unless stabilized by additional interactions. Nevertheless, peptides derived from ribonuclease A that contain residues 3–13 do show partial helix formation in water at low temperatures. We discuss the possibility that specific interactions between side chains can stabilize short α-helixes by nucleating the helix, and that specific interactions may also define the helix boundaries at early stages in folding.  相似文献   

2.
The synthesis of the cation-binding cyclic octapeptide, cyclo(Glu-Sar-Gly-(N-decyl)Gly)2 is reported. This peptide, containing two ionizable Glu carboxyl side chain protons per molecule, can form neutral cation complexes with divalent ions via protonmetal exchange. Solubilized in chloroform solution, the peptide has been found to extract calcium from an aqueous phase (pH 8.5, 100 mM Tris) generally on a 11 molar basis. By contrast, under comparable conditions but with other metal chlorides, the peptide does not extract magnesium, sodium, or potassium. That the extraction proceeds via proton-metal exchange was demonstrated by the absence of (radioactive) chloride ion from the organic phase. Parallel sets of experiments performed with the naturally-occurring ionophore A23187 reaffirmed that the latter substance extracts calcium and magnesium with nearly equal propensity.  相似文献   

3.
The effect of Ca2+ ion on structural fluctuation of a milk Ca2+-binding protein, α-lactalbumin, under native conditions was investigated by comparing hydrogen-exchange reactions of tryptophan residues in the apo-form without Ca2+ and in the holo-form at 1 mM CaCl2 at pH 7.0 in the presence of 0.1M Na+. The reactions were followed by measuring time-dependent absorption changes at 298–300 nm due to the 2H-1H exchange of the tryptophan imino protons and were found to be biphasic under all the conditions examined. Two of the four tryptophan protons are insensitive to Ca2+ concentration and show a relatively fast exchange rate. The other two protons are much more extensively protected (a protection degree of 103–105) and are markedly affected by the presence of Ca2+. Examinations of the temperature dependence and pH dependence of the individual exchange rates have been utilized for elucidating the exchange mechanism. The fast protons show a low activation energy reaction with so-called EX2 kinetics. The exchange reaction of the slow protons is accompanied by a high activation energy, and the exchange mechanism of the protons depended on the presence or absence of stabilizing Ca2+ ions—the EX1 kinetics for the apo-protein and the EX2 kinetics for the holo-protein at 1 mM Ca2+. The exchange reaction in the thermally unfolded state was also found to be biphasic, but the fast phase, which has an exchange rate in the fully exposed state, becomes predominant with decreasing temperature. By taking this fact and using a structural unfolding model of hydrogen exchange, the present results are fully consistent with thermodynamic parameters of the thermal transition and kinetic parameters of refolding reactions induced by concentration jumps of guanidine hydrochloride obtained in previous studies. It is demonstrated that the reaction of the slow protons in the native state is mediated by a transient global unfolding equivalent to the “thermal” unfolding under a native condition and that switching of the exchange mechanism from the EX1 to EX2 kinetics results from acceleration of the refolding rate with an increase in Ca2+ concentration. The transient global unfolding takes place even under a strongly native condition, e.g., at a temperature 20° below the beginning of the thermal transition.  相似文献   

4.
The exchange broadening of backbone amide proton resonances of a 23-mer fusion peptide of the transmembrane subunit of HIV-1 envelope glycoprotein gp41, gp41-FP, was investigated at pH 5 and 7 at room temperature in perdeuterated sodium dodecyl sulfate (SDS) micellar solution. Comparison of resonance peaks for these pHs revealed an insignificant change in exchange rate between pH 5 and 7 for amide protons of residues 4 through 14, while the exchange rate increase at neutral pH was more prominent for amide protons of the remaining residues, with peaks from some protons becoming undetectable. The relative insensitivity to pH of the exchange for the amide protons of residues 4 through 14 is attributable to the drastic reduction in [OH–] in the micellar interior, leading to a decreased exchange rate. The A15-G16 segment represents a transition between these two regimes. The data are thus consistent with the notion that the peptide inserts into the hydrophobic core of a membrane-like structure and the A15-G16 dipeptide is located at the micellar-aqueous boundary.  相似文献   

5.
Two glutamic acid-rich fusion peptide analogs of influenza hemagglutinin were synthesized to study the organization of the charged peptides in the membranous media. Fluorescence and gel electrophoresis experiments suggested a loose association between the monomers in the vesicles. A model was built which showed that a positional difference of 3, 7 and 4, 8 results in the exposure of Glu3 and Glu7 side chains to the apolar lipidic core. Supportive results include: first, pKa values of two pH units higher than reference value in aqueous medium for Glu3 and Glu7 CγH, whereas the deviation of pKa from the reference value for Glu4 and Glu8 CγH is substantially smaller; second, Hill coefficients of titration shift of these protons indicate anti-cooperativity for Glu3 and Glu7 side chain protons but less so for Glu4 and Glu8, implying a strong electrostatic interaction between Glu3 and Glu7 possibly resulting from their localization in an apolar environment; third, positive and larger titration shift for NH of Glu3 is observed compared to that of Glu4, suggesting stronger hydrogen bond between the NH and the carboxylic group of Glu3 than that of Glu4, consistent with higher degree of exposure to hydrophobic medium for the side chain of Glu3.  相似文献   

6.
We report proton magnetic resonance studies of a series of lysine oligopeptides in H2O solution. At pH 5 the protonated ε-amino groups are seen as broad resonances; the peptide NH proton resonances are split by spin–spin coupling with the Cα-H proton, and appear at positions which depend on position in the chain and on chain length. Assignments were made by the europium shift method, and we observed the expected effect of catalysis by the terminal —NH3+ of exchange of the adjacent peptide NH. Coupling constants and the temperature coefficient of chemical shift values were consistent with a non-hydrogen-bonded structure for the oligolysines. The rate and mechanism of NH hydrogen exchange were investigated by line-broadening measurements of the peptide protons as a function of pH. Exchange was found to be OH? catalyzed, with large differences in the rate depending on position in the chain. Preliminary studies of the complex between double-helical d(pA)3pGpC(pT)3 and tetra(L -lysine) were performed using 1H- and 31P-nmr techniques. Pmr spectra of the complex at pH values ranging from 3.98 to 8.15 showed very complicated patterns. Downfield shifts and reduction in exchange rates were observed for several tetra(L -lysine) protons. 31P-nmr spectra of the complex reveal an upfield shift of 1 ppm for 3′-5′ phosphate diester resonances on complexation. 31P T1 relaxation times change little on complex formation at low temperature but are altered at higher temperature.  相似文献   

7.
The hydrogen exchange kinetics of the S-peptide in ribonuclease S can be measured by first tritiating the S-peptide in the absence of S-protein and then allowing it to recombine rapidly with S-protein. Afterwards the exchange reactions of this specific segment of ribonuclease S can be studied. The exchange kinetics of bound S-peptide are complex, indicating that different protons exchange at markedly different rates. The terminal exchange reaction, involving at least five highly protected protons, has been studied as a function of pH.At low concentrations of ribonuclease S the exchange kinetics become concentration-dependent, owing to the dissociation of the S-peptide. Although the fraction of free S-peptide is always very small, its rate of exchange is several orders of magnitude faster than that of bound S-peptide, and the concentration dependence of the exchange kinetics is readily measurable. It provides a highly sensitive method for determining small dissociation constants (KD). Values of KD ranging from 10?6m at pH 2.7, 0 °C, to 2 × 10?10m at pH 7.0, 0 °C, are reported here. Our value for KD at pH 7.0, 0 °C, confirms the data and extrapolation to 0 °C of Hearn et al. (1971).At high concentrations of ribonuclease S the terminal exchange reaction is independent of concentration. It probably results from a local unfolding reaction of the bound S-peptide. Above pH 4 the strong pH dependence of KD closely resembles that of the apparent equilibrium constant for this local unfolding reaction. The latter may be one step in the dissociation process and we present such a model for ribonuclease S dissociation.Measurement of concentration-dependent exchange kinetics should provide a useful method of determining small dissociation constants in other systems: for example, in studies of protein-nucleic acid interactions.  相似文献   

8.
Lanthanide DOTA-tetraglycinate (LnDOTA-(gly)4 ) complexes contain four magnetically equivalent amide protons that exchange with protons of bulk water. The rate of this base catalyzed exchange process has been measured using chemical exchange saturation transfer (CEST) NMR techniques as a function of solution pH for various paramagnetic LnDOTA-(gly)4 complexes to evaluate the effects of lanthanide ion size on this process. Complexes with Tb(III), Dy(III), Tm(III) and Yb(III) were chosen because these ions induce large hyperfine shifts in all ligand protons, including the exchanging amide protons. The magnitude of the amide proton CEST exchange signal differed for the four paramagnetic complexes in order, Yb>Tm>Tb>Dy. Although the Dy(III) complex showed the largest hyperfine shift as expected, the combination of favorable chemical shift and amide proton CEST linewidth in the Tm(III) complex was deemed most favorable for future in vivo applications where tissue magnetization effects can interfere. TmDOTA-(gly)4 at various concentrations was encapsulated in the core interior of liposomes to yield lipoCEST particles for molecular imaging. The resulting nanoparticles showed less than 1% leakage of the agent from the interior over a range of temperatures and pH. The pH versus amide proton CEST curves differed for the free versus encapsulated agents over the acidic pH regions, consistent with a lower proton permeability across the liposomal bilayer for the encapsulated agent. Nevertheless, the resulting lipoCEST nanoparticles amplify the CEST sensitivity by a factor of ∼104 compared to the free, un-encapsulated agent. Such pH sensitive nano-probes could prove useful for pH mapping of liposomes targeted to tumors.  相似文献   

9.
In the present paper we investigate the influence of sample pH on the conformational and dynamical properties of the pseudotripeptide H-Tyr-TicΨ[CH2NH]Phe-OH(TIP[Ψ]:Tic: l, 2, 3, 4,-tetrahydroisoquinoline-3-carboxylic acid) using various one- and two-dimensional nmt techniques in conjunction with molecular modeling. Studies were conducted at three different pH levels-corresponding to the zwitterionic peptide containing a formal positive charge(pH 3. 1).the deprotonated molecule(pH 9. 1), and a situation at neutral pH(pH 7. 2) involving both protonated and deprotonated states of the reduced peptide bond. Analysis of the one-dimensional1H-nmr spectra reveals that in solution TIP[Ψ]is in slow dynamic exchange between conformations containing cis and trans configurations of the Tyr-Tic bond. An nmr pH dependence study of the cis:trans ratio indicated that the exchange process was governed by the protonation state of the reduced bond amine. From the nmr data, reduced peptide bond pKavalues of 6. 5 and 7. 5 were determined for the cis and trans conformers, respectively. It was concluded that conformations containing a trans Tyr-Tic bond are stabilized at law pH by an intramolecular hydrogen bond between the Tyr carbonyl and the reduced peptide bond protonated amine. This observation was corroborated by molecular mechanics investigations that revealed low energy trans structures compatible with nmr structural data, and furthermore, were consistently characterized by the existence of a strong N+ H?O? C interaction closing a seven-membered cycle. The dynamics of cis-trans isomerization about the Tyr-Tic peptide bond were probed by nmr exchange experiments. The selective presaturation of exchanging resonances carried out at several temperatures between 50 and 70°C allowed the determination of isomerization rate constants as well as thermodynamic activation parameters. ΔG values were in close agreement with the cis → trans energy barrier found in X-Pro peptide fragments (~83 kJ/mol).A large entropic barrier determined for the trans → cis conversion of TIP[Ψ](5. 7 JK?1 mol?1 at pH 3. 1; 6. 5 JK?1 mol?1 at pH 9. 1) is discussed in terms of decreased solvent molecular ordering around the conformers possessing a trans Tyr-Tic bond. Evidence that the neutral form of the reduced peptide bond gains rigidity upon protonation was obtained from relaxation measurements in the rotating frame. TJp measurements of several protons in the vicinity of the reduced peptide bond were made as a function of spin-lock field. Quantitative analysis of the relaxation data indicated that chemical shift fluctuations in the 10?4-10?5s range were more pronounced in the case of deprotonated TIP[Ψ]. Results of molecular dynamics simulations in addition to 3 J αβ coupling constant measurements support the experimentally observed greater flexibility in the C-terminal region of TIP[Ψ]. © 1995 John Wiley & Sons, Inc.  相似文献   

10.
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   

11.
The exchange rate of the indole nitrogen proton with solvent water protons was measured as a function of pH and temperature for tryptophan, N-methyl tryptophan, glycyl tryptophan, tryptophanamide, tryptophylglycine, and tryptophylglycyl glycine. The nmr observation was by long-pulse Fourier transform methods, and kinetics were inferred from saturation recovery, H2O transfer of saturation, and linewidth. There are observable differences between the rates of these compounds, but all are describable within a factor of two by specific acid and base-catalyzed rates kH = 100 and kOH = 108 1/mol-sec at 27°C. It is concluded that this behaviour is representative of this proton on the indole side chain in a random-chain peptide exposed to water.  相似文献   

12.
We show for the first time that the secondary structure of the Alzheimer beta-peptide is in a temperature-dependent equilibrium between an extended left-handed 3(1) helix and a flexible random coil conformation. Circular dichroism spectra, recorded at 0.03 mM peptide concentration, show that the equilibrium is shifted towards increasing left-handed 3(1) helix structure towards lower temperatures. High resolution nuclear magnetic resonance (NMR) spectroscopy has been used to study the Alzheimer peptide fragment Abeta(12-28) in aqueous solution at 0 degrees C and higher temperatures. NMR translation diffusion measurements show that the observed peptide is in monomeric form. The chemical shift dispersion of the amide protons increases towards lower temperatures, in agreement with the increased population of a well-ordered secondary structure. The solvent exchange rates of the amide protons at 0 degrees C and pH 4.5 vary within at least two orders of magnitude. The lowest exchange rates (0.03-0.04 min(-1)) imply that the corresponding amide protons may be involved in hydrogen bonding with neighboring side chains.  相似文献   

13.
The object of this study was to analyze the dynamics of Al and protons in the rhizosphere of maize cultivated in a simple acid substrate, so as to allow the use of a dynamic model of the functioning of a rhizosphere consisting of an organic phase (an agarose gel) and a mineral phase (an amorphous aluminium hydroxide). Two cultivars of maize (Zea mays L.), one Al-sensitive and the other Al-tolerant, were cultivated on this substrate in the presence of different proportions of NH 4 + and NO 3 - , which served to acidify the rhizosphere to a greater or lesser extent. The state of the agarose gel and of the cell walls of the roots were monitored using an ion exchange model which had previously been calibrated for each substrate. The experiment showed that Al and protons reduce root growth and the Ca and Mg content in the root, while relative growth varies little between pH 4.0 and pH 4.5. The model showed that competition between Al and protons for the binding sites of the cell walls might account for these results. The sensitivity of the model to the rate of Al(OH)3 dissolution and to the cation exchange capacity of the culture substrate was tested by numerical simulation. When roots release protons and dissolve Al(OH)3 in the rhizosphere, there is little possibility of Al desorption by protons on the cell walls at pHs compatible with good root growth of maize, plant specie sensitive to Al and H. Furthermore, the phytotoxicity of the different forms of Al hydroxides should be considered only in taking into account the dynamics of the whole system, in particular the solubilisation of Al in the rhizosphere. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

14.
Intermediate amyloidogenic states along the amyloid β peptide (Aβ) aggregation pathway have been shown to be linked to neurotoxicity. To shed more light on the different structures that may arise during Aβ aggregation, we here investigate surfactant-induced Aβ aggregation. This process leads to co-aggregates featuring a β-structure motif that is characteristic for mature amyloid-like structures. Surfactants induce secondary structure in Aβ in a concentration-dependent manner, from predominantly random coil at low surfactant concentration, via β-structure to the fully formed α-helical state at high surfactant concentration. The β-rich state is the most aggregation-prone as monitored by thioflavin T fluorescence. Small angle x-ray scattering reveals initial globular structures of surfactant-Aβ co-aggregated oligomers and formation of elongated fibrils during a slow aggregation process. Alongside this slow (minutes to hours time scale) fibrillation process, much faster dynamic exchange (kex ∼1100 s−1) takes place between free and co-aggregate-bound peptide. The two hydrophobic segments of the peptide are directly involved in the chemical exchange and interact with the hydrophobic part of the co-aggregates. Our findings suggest a model for surfactant-induced aggregation where free peptide and surfactant initially co-aggregate to dynamic globular oligomers and eventually form elongated fibrils. When interacting with β-structure promoting substances, such as surfactants, Aβ is kinetically driven toward an aggregation-prone state.  相似文献   

15.
The extent to which proteins aggregate into distinct structures ranging from prefibrillar oligomers to amyloid fibrils is key to the pathogenesis of many age-related degenerative diseases. We describe here for the Alzheimer's disease-related amyloid β peptide (Aβ) an investigation of the sequence-based determinants of the balance between the formation of prefibrillar aggregates and amyloid fibrils. We show that by introducing single-point mutations, it is possible to convert the normally harmless Aβ40 peptide into a pathogenic species by increasing its relative propensity to form prefibrillar but not fibrillar aggregates, and, conversely, to abolish the pathogenicity of the highly neurotoxic E22G Aβ42 peptide by reducing its relative propensity to form prefibrillar species rather than mature fibrillar ones. This observation can be rationalized by the demonstration that whereas regions of the sequence of high aggregation propensity dominate the overall tendency to aggregate, regions with low intrinsic aggregation propensities exert significant control over the balance of the prefibrillar and fibrillar species formed, and therefore play a major role in determining the neurotoxicity of the Aβ peptide.  相似文献   

16.

Background

In the membrane-bound enzyme cytochrome c oxidase, electron transfer from cytochrome c to O2 is linked to proton uptake from solution to form H2O, resulting in a charge separation across the membrane. In addition, the reaction drives pumping of protons across the membrane.

Methods

In this study we have measured voltage changes as a function of pH during reaction of the four-electron reduced cytochrome c oxidase from Rhodobacter sphaeroides with O2. These electrogenic events were measured across membranes containing purified enzyme reconstituted into lipid vesicles.

Results

The results show that the pH dependence of voltage changes (primarily associated with proton transfer) during O2 reduction does not match that of the previously studied absorbance changes (primarily associated with electron transfer). Furthermore, the voltage changes decrease with increasing pH.

Conclusions

The data indicate that cytochrome c oxidase does not pump protons at high pH (10.5) (or protons are taken from the “wrong” side of the membrane) and that at this pH the net proton-uptake stoichiometry is ∼ 1/2 of that at pH 8. Furthermore, the results provide a basis for interpretation of results from studies of mutant forms of the enzyme.

General significance

These results provide new insights into the function of cytochrome c oxidase.  相似文献   

17.
In order to establish whether p.m.r. spectroscopy is useful for identifying Amadori- and Heyns-rearrangement products, the p.m.r. spectra at 220 MHz of 16 rearrangement products derived from d-glucose or d-fructose and amino acids have been investigated. At pH 3, the protons of the NCH2 group of N-substituted 1-amino-1-deoxy-d-fructose (Amadori-rearrangement products) resonate at δ 3.25–3.60 in D2O and are shifted upfield by 0.3–0.6 p.p.m. at pH 9. These protons exchange with deuterium. Also, in D2O there is an equilibrium of the acyclic, furanose, and pyranose structures, the last being favoured. At pH ? 7, the equilibrium is completely shifted to the β-pyranose form, which adopts exclusively the 2C5 conformation. At pH 3, the equilibrium favours the β-furanose form. At pH 3, H-1e and H-1a of N-substituted 2-amino-2-deoxy-d-glucoses (Heyns-rearrangement products) resonate at δ 5.55 and 5.04, respectively. At pH 9, the signal for H-2 is shifted upfield by 0.2–0.7 p.p.m. In D2O solution, these compounds exist as an equilibrium of α- and β-pyranose forms in the 4C1 conformation. The α anomer is stabilised by the amino acid group at position 2. At pH 3, the αβ-ratio is 2–4:1, and, at pH 9, 1.0–1.1:1.  相似文献   

18.
A general method to explore the dynamic nature of amyloid fibrils is described, combining hydrogen/deuterium exchange and nuclear magnetic resonance spectroscopy to determine the exchange rates of individual amide protons within an amyloid fibril. Our method was applied to fibrils formed by the amyloid-β(1-40) peptide, the major protein component of amyloid plaques in Alzheimer’s disease. The fastest exchange rates were detected among the first 14 residues of the peptide, a stretch known to be poorly structured within the fibril. Considerably slower exchange rates were observed in the remainder of the peptide within the β-strand-turn-β-strand motif that constitutes the fibrillar core.  相似文献   

19.
The preceding article shows that there are eight highly protected amide protons in the S-peptide moiety of RNAase S at pH 5, 0 degrees C. The residues with protected NH protons are 7 to 13, whose amide protons are H-bonded in the 3 to 13 alpha-helix, and Asp 14, whose NH proton is H-bonded to the CO group of Val47. We describe here the exchange behavior of these eight protected protons as a function of pH. Exchange rates of the individual NH protons are measured by 1H nuclear magnetic resonance in D2O. A procedure is used for specifically labeling with 1H only these eight NH protons. The resonance assignments of the eight protons are made chiefly by partial exchange, through correlating the resonance intensities in spectra taken when the peptide is bound and when it is dissociated from S-protein in 3.5 M-urea-d4, in D2O, pH 2.3, -4 degrees C. The two remaining assignments are made and some other assignments are checked by measurements of the nuclear Overhauser effect between adjacent NH protons of the alpha-helix. There is a transition in exchange behavior between pH 3, where the helix is weakly protected against exchange, and pH 5 where the helix is much more stable. At pH 3.1, 20 degrees C, exchange rates are uniform within the helix within a factor of two, after correction for different intrinsic exchange rates. The degree of protection within the helix is only 10 to 20-fold at this pH. At pH 5.1, 20 degrees C, the helix is more stable by two orders of magnitude and exchange occurs preferentially from the N-terminal end. At both pH values the NH proton of Asp 14, which is just outside the helix, is less protected by an order of magnitude than the adjacent NH protons inside the helix. Opening of the helix can be observed below pH 3.7 by changes in chemical shifts of the NH protons in the helix. At pH 2.4 the changes are 25% of those expected for complete opening. Helix opening is a fast reaction on the n.m.r. time scale (tau much less than 1 ms) unlike the generalized unfolding of RNAase S which is a slow reaction. Dissociation of S-peptide from S-protein in native RNAase S at pH 3.0 also is a slow reaction. Opening of the helix below pH 3.7 is a two-state reaction, as judged by comparing chemical shifts with exchange rates. The exchange rates at pH 3.1 are predicted correctly from the changes in chemical shift by assuming that helix opening is a two-state reaction. At pH values above 3.7, the nature of the helix opening reaction changes. These results indicate that at least one partially unfolded state of RNAase S is populated in the low pH unfolding transition.  相似文献   

20.
pH titration by 1H-NMR spectroscopy of the peptide hormone bradykinin was carried out in 2H2O. Assignment of all α-proton signals and of most of the other resonances permitted the extraction of vicinal coupling constants 3Jαβ,β′ from which side chain conformation of all residues could be followed and analyzed as a function of pH. It is shown that the ionization of the terminal COOH group affects simultaneously the Arg-9 and Phe-8 chemical shifts and side chain orientation, and the non-equivalence of the Gly-4 methylene protons. Cooperative effects along the peptide backbone or a folded structure of the C-terminal part of bradykinin could explain this effect.  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