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 234 毫秒
1.
Ion mobility measurements have been performed for protonated polyalanine peptides (A10 + H+, A15 + H+, A20 + H+, A25 + H+, and A15NH2 + H+) as a function of temperature using a new high-temperature drift tube. Peaks due to helices and globules were found at room temperature for all peptides, except for A10 + H+ (where only the globule is present). As the temperature is increased, the helix and globule peaks broaden and merge to give a single narrow peak. This indicates that the two conformations interconvert rapidly at elevated temperatures. The positions of the merged peaks show that A15 + H+ and A15NH2 + H+ spend most of their time as globules when heated, while A20 + H+ and A25 + H+ spend most of their time as helices. The helix/globule transitions are almost certainly accompanied by intramolecular proton transfer, and so, these results suggest that the proton becomes mobile (able to migrate freely along the backbone) at around 450 K. The peptides dissociate as the temperature is increased further to give predominantly the bn(+), b(n-1)(+), b(n-2)(+), ... series of fragment ions. There is a correlation between the ease of fragmentation and the time spent in the helical conformation for the An + H+ peptides. Helix formation promotes dissociation because it pools the proton at the C-terminus where it is required for dissociation to give the observed products. In addition to the helix and globule, an antiparallel helical dimer is observed for the larger peptides. The dimer can be collisionally dissociated by injection into the drift tube at elevated kinetic energies.  相似文献   

2.
Helix unfolding in unsolvated peptides   总被引:1,自引:0,他引:1  
The conformations of unsolvated Ac-K(AGG)(5)+H(+) and Ac-(AGG)(5)K+H(+) peptides (Ac = acetyl, A = alanine, G = glycine, and K = lysine) have been examined by ion mobility measurements over a wide temperature range (150-410 K). The Ac-K(AGG)(5)+H(+) peptide remains a globule (a compact, roughly spherical structure) over the entire temperature range, while both an alpha-helix and a globule are found for Ac-(AGG)(5)K+H(+) at low temperature. As the temperature is raised the alpha-helix unfolds. Rate constants for loss of the helix (on a millisecond time scale) have been determined as a function of temperature and yield an Arrhenius activation energy and preexponential factor of 38.2 +/- 1.0 kJ mol(-1) and 6.5 +/- 3.7 x 10(9) s(-1), respectively. The alpha-helix apparently does not unfold directly into the globule, but first converts into a long-lived intermediate which survives to a significantly higher temperature before converting. According to molecular dynamics simulations, there is a partially untwisted helical conformation that has both a low energy and a well-defined geometry. This special structure lies between the helix and globule and may be the long-lived intermediate.  相似文献   

3.
The energetics, dynamics, and infrared spectroscopy of the shared proton in different chemical environments is investigated using molecular dynamics simulations. A three-dimensional potential energy surface (PES) suitable for describing proton transfer between an acceptor and a donor oxygen atom is combined with an all-atom force field to carry out reactive molecular dynamics simulations. The construction of the fully dimensional PES is inspired from the established mixed quantum mechanics/molecular mechanics treatment of larger systems. The "morphing potential" method is used to transform the generic PES for proton transfer along an O...H+...O motif into a three-dimensional PES for proton transfer in protonated diglyme. Using molecular dynamics simulations at finite temperature, the gas phase infrared spectra are calculated for both species from the Fourier transform of the dipole moment autocorrelation function. For protonated diglyme the modes involving the H+ motion are strongly mixed with other degrees of freedom. At low temperature, the O...H+...O asymmetric stretching vibration is found at 870 cm-1, whereas for H5O2+ this band is at 724 cm-1. As expected, the vibrational bands of protonated diglyme show no temperature dependence whereas for H5O2+ at T = 100 K the proton transfer mode is found at 830 cm-1, in good agreement with 861 cm-1 from very recent molecular dynamics simulations.  相似文献   

4.
Ion mobility measurements and molecular dynamics simulations were performed for unsolvated A4G7A4 + H+ and Ac-A4G7A4 + H+ (Ac = acetyl, A = alanine, G = glycine) peptides. As expected, A4G7A4 + H+ adopts a globular conformation (a compact, random-looking, three-dimensional structure) over the entire temperature range examined (100-410 K). Ac-A4G7A4 + H+ on the other hand is designed to have a flat energy landscape with a marginally stable helical state. This peptide shows at least four different conformations at low temperatures (<230 K). The two conformations with the largest cross sections are attributed to - and partial -helices, while the one with the smallest cross section is globular. The other main conformation may be partially helical. Ac-A4G7A4 + H+ becomes predominantly globular at intermediate temperatures and then becomes more helical as the temperature is raised further. This unexpected behavior may be due to the helix having a higher vibrational entropy than the globular state, as predicted by some recent calculations (Ma, B.; Tsai, C.-J.; Nussinov, R. Biophys. J. 2000, 79, 2739-2753).  相似文献   

5.
Light-induced proton pumping in bacteriorhodospin is carried out through five proton transfer steps. We propose that the proton transfer to Asp85 from the Schiff base in the L-to-M transition is accompanied by the relocation of a water cluster on the cytoplasmic side of the Schiff base from a site close to the Schiff base in L to the Phe219-Thr46 region in M. The water cluster present in L, formed at 170 K, is more rigid than that at room temperature. This may be responsible for blocking the conversion of L to M at 170 K. In the photocycle at room temperature, this water cluster returns to the site close to the Schiff base in N, with a rigid structure similar to that of L at 170 K. The increase in the proton affinity of Asp85, which is a prerequisite for the one-way proton transfer in the M-to-N transition, is suggested to be facilitated by a structural change which disrupts interactions between Asp212 and the Schiff base, and between Asp212 and Arg82. We propose that this liberation of Asp212 is accompanied by a rearrangement of the structure of water molecules between Asp85 and Asp212, stabilizing the protonated Asp85 in M.  相似文献   

6.
To elucidate the mechanism of the exceptional behavior of lysine for the ionization (protonation) yields in matrix-assisted laser desorption/ionization (MALDI) observed by Nishikaze and Takayama [Rapid Commun. Mass Spectrom. 2006, 20, 376], the temperature dependences of proton affinity (PA) and gas phase basicity for 20 amino acids are theoretically analyzed with correlated ab initio molecular orbital method under ideal gas condition. We have found that two different conformations, the linear structure with elongation of the side chain and the folded one having intramolecular hydrogen bonding, play important roles for the exceptional behavior of lysine. At low temperatures of around 298 K, the most stable conformation of the protonated lysine is the folded structure due to the formation of intramolecular hydrogen bonding. Meanwhile, at high temperatures, the Gibbs free energy of linear structure of protonated lysine becomes lower than that of the folded one because of the increment of vibrational entropic contribution. To explicitly take account of the contribution of the free energies, we have proposed the effective PA values thermally averaged using the ratio of Boltzmann distributions for two conformations. Since the effective PA value for lysine drastically decreases as the temperature increases above 1000 K, the linear correlation is clearly obtained between our effective PA values at high temperature and the ion yields in MALDI.  相似文献   

7.
Ion mobility measurements and molecular dynamic simulations have been performed for a series of peptides designed to have helix-turn-helix motifs. For peptides with two helical sections linked by a short loop region: AcA(14)KG(3)A(14)K+2H(+), AcA(14)KG(5)A(14)K+2H(+), AcA(14)KG(7)A(14)K+2H(+), and AcA(14)KSar(3)A(14)K+2H(+) (Ac = acetyl, A = alanine, G = glycine, Sar = sarcosine and K = lysine); a coiled-coil geometry with two anti-parallel helices is the lowest energy conformation. The helices uncouple and the coiled-coil unfolds as the temperature is raised. Equilibrium constants determined as a function of temperature yield enthalpy and entropy changes for the unfolding of the coiled-coil. The enthalpy and entropy changes depend on the length and nature of the loop region. For a peptide with three helical sections: protonated AcA(14)KG(3)A(14)KG(3)A(14)K; a coiled-coil bundle with three helices side-by-side is substantially less stable than a geometry with two helices in an antiparallel coiled-coil and the third helix collinear with one of the other two.  相似文献   

8.
Ion mobility measurements have been used to examine the conformations present for unsolvated Ac-(AG)(7)A+H(+) and (AG)(7)A+H(+) peptides (Ac = acetyl, A = alanine, and G = glycine) over a broad temperature range (100-410 K). The results are compared to those recently reported for Ac-A(4)G(7)A(4)+H(+) and A(4)G(7)A(4)+H(+), which have the same compositions but different sequences. Ac-(AG)(7)A+H(+) shows less conformational diversity than Ac-A(4)G(7)A(4)+H(+); it is much less helical than Ac-A(4)G(7)A(4)+H(+) at the upper end of the temperature range studied, and at low temperatures, one of the two Ac-A(4)G(7)A(4)+H(+) features assigned to helical conformations is missing for Ac-(AG)(7)A+H(+). Molecular dynamics simulations suggest that the different conformational preferences are not due to differences in the stabilities of the helical states, but differences in the nonhelical states: it appears that Ac-(AG)(7)A+H(+) is more flexible and able to adopt lower energy globular conformations (compact random looking three-dimensional structures) than Ac-A(4)G(7)A(4)+H(+). The helix to globule transition that occurs for Ac-(AG)(7)A+H(+) at around 250-350 K is not a direct (two-state) process, but a creeping transition that takes place through at least one and probably several intermediates.  相似文献   

9.
We used high-resolution proton-detected multidimensional NMR to study the solvent-exposed parts of a seven-helical integral membrane proton pump, proteorhodopsin (PR). PR samples were prepared by growing the apoprotein on fully deuterated medium and reintroducing protons to solvent-accessible sites through exchange with protonated buffer. This preparation leads to NMR spectra with proton resolution down to ca. 0.2 ppm at fast spinning (28 kHz) in a protein back-exchanged at a level of 40%. Novel three-dimensional proton-detected chemical shift correlation spectroscopy allowed for the identification and resonance assignment of the solvent-exposed parts of the protein. Most of the observed residues are located at the membrane interface, but there are notable exceptions, particularly in helix G, where most of the residues are susceptible to H/D exchange. This helix contains Schiff base-forming Lys231, and many conserved polar residues in the extracellular half, such as Asn220, Tyr223, Asn224, Asp227, and Asn230. We proposed earlier that high mobility of the F-G loop may transiently expose a hydrophilic cavity in the extracellular half of the protein, similar to the one found in xanthorhodopsin. Solvent accessibility of helix G is in line with this hypothesis, implying that such a cavity may be a part of the proton-conducting pathway lined by this helix.  相似文献   

10.
The structure and spectral signatures of the protonated homodimer of pyridine in its complex with a poorly coordinating anion have been studied in solution in CDF(3)/CDClF(2) down to 120 K and in a single crystal. In both phases, the hydrogen bond is asymmetric. In the solution, the proton is involved in a fast reversible transfer that determines the multiplicity of NMR signals and the sign of the primary H/D isotope effect of --0.95 ppm. The proton resonates at 21.73 ppm that is above any value reported in the past and is indicative of a very short hydrogen bond. By combining X-ray diffraction analysis with model computations, the position of the proton in the crystal has been defined as d(N-H) = 1.123 ? and d(H···N) = 1.532 ?. The same distances have been estimated using a (15)N NMR correlation. The frequency of the protonic out-of-plane bending mode is 822 cm(-1) in agreement with Novak's correlation.  相似文献   

11.
The elimination of water from the carboxyl group of protonated diglycine has been investigated by density functional theory calculations. The resulting structure is identical to the b(2) ion formed in the mass spectrometric fragmentation of protonated peptides (therefore named "b2" in this study). The most stable geometry of the fragment ion ("b2") is an O-protonated diketopiperazine. However, its formation is kinetically disfavored as it requires a free energy of 58.2 kcal/mol. The experimentally observed N-protonated oxazolone is 3.0 kcal/mol less stable. The lowest energy pathway for the formation of the "b2" ion requires a free energy of 37.5 kcal/mol and involves the proton transfer from the amide oxygen of protonated diglycine to the hydroxyl oxygen. Fragmentation initiated by proton transfer from the terminal nitrogen has also a comparable free energy of activation (39.4 kcal/mol). Proton transfer initiating the fragmentation, from the highly basic terminal nitrogen or amide oxygen to the less basic hydroxyl oxygen is feasible at energies reached in usual mass spectrometric experiments. Amide N-protonated diglycine structures are precursors of mainly y(1) ions rather than "b2" ions. In the lowest energy fragmentation channels, proton transfer to the hydroxylic oxygen, bond breaking and formation of an oxazolone ring occur concertedly but asynchronously. Proton transfer to hydroxyl oxygen and cleavage of the corresponding C-O bond take place at the early stages of the fragmentation step, while ring closure to form an oxazolone geometry occurs at the later stages of the transition. The experimentally observed low kinetic energy release is expected to be due to the existence of a strongly hydrogen bonded protonated oxazolone-water complex in the exit channel. Whereas the threshold energy for "b2" ion formation (37.1 kcal/mol) is lower than for the y(1) ion (38.4 kcal/mol), the former requires a tight transition state with an activation entropy, DeltaS++ = -1.2 cal/mol.K and the latter has a loose transition state with DeltaS++ = +8.8 cal/mol.K. This leads to y(1) being the major fragment ion over a wide energy range.  相似文献   

12.
Ion-water-cluster properties are investigated both through the multistate empirical valence bond potential and a polarizable model. Equilibrium properties of the ion-water clusters H+(H2O)100, Na+(H2O)100, Na+(H2O)20, and Cl-(H2O)17 in the temperature region 100-450 K are explored using a hybrid parallel basin-hopping and tempering algorithm. The effect of the solid-liquid phase transition in both caloric curves and structural distribution functions is investigated. It is found that sodium and chloride ions largely reside on the surface of water clusters below the cluster melting temperature but are solvated into the interior of the cluster above the melting temperature, while the solvated proton was found to have significant propensity to reside on or near the surface in both the liquid- and solid-state clusters.  相似文献   

13.
The conformation change picture of human islet amyloid polypeptide (hIAPP) is outlined using molecular dynamics simulation, and the structural influences of L16Q, S20G, and L16Q‐S20G mutations on the conformation of hIAPP are analyzed. Particularly, the conformational changes of the amyloidogenic‐related regions of residues 15–17 and 20–29 are emphasized. Our studies find that, for WT hIAPP, residues 15–17 always maintain a stable α‐helix structure, residues 20–25 structurally fluctuate between turn and 5‐helix, and residues 26–29 mainly adopt coil and bend structures. The hydrogen bonds between the polar groups of hIAPP, long‐rang van der Waals forces between the residues, and hydrophobic interactions between the residues of hIAPP are important driving forces to maintain the secondary structure of hIAPP. The replacement of leucine16 by glutamine stabilizes the helix structure of residues 15–17 and 20–23 of hIAPP monomer, and the structure of residues 24–29 fluctuates between helix and turn. The relatively stable helix structures of residues 15–17 and 20–29 are supposed to be beneficial for L16Q hIAPP to resist the aggregation as observed in the experiment. The substitution of serine20 by glycine drives residues 15–17 and 20–29 of hIAPP to transform from helix structure to β‐strands or coil structures with higher extension and flexibility, which may promote the aggregation of hIAPP as the experiments reported. These results are significant to understand the aggregation mechanism of hIAPP monomer into the dimer, trimer, oligomers and fibrils associated with the type 2 diabetes at the atomic level.  相似文献   

14.
(1)H and (13)C NMR data for N-substituted morpholines 1-20 were measured using 1D (DEPT, 1D NOE difference) and 2D NMR spectroscopic methods including (1)H-(1)H COSY, long-range (1)H-(1)H COSY, NOESY, gHMBC and gHMQC experiments. At room temperature the (1)H NMR spectra of protonated compounds 2 and 9 show the chair conformation for the morpholine ring. Spin-spin coupling constants were deduced from the resolution-enhanced proton spectra.  相似文献   

15.
<正>In this work,the formation sites,helical parameters and hydrogen bond positions of Konjac glucomannan molecular helices were investigated using molecular dynamic simulation method.To our interest,the KGM chain is mainly composed by local left and right helix structures. The formation sites of KGM chain might locate at the chain-segments containing acetyl groups,and the left helix is the favorable conformation of KGM.Temperature-dependent molecule conformation study indicates that the right helix is dominant when the temperature is lower than 343 K,above which,however,the left helix is dominating(right helix disappears).In addition,intramolecular hydrogen bonds in the left helix can be found at the-OH groups on C(2),C(4)and C(6)of mannose residues;comparably,the intramolecular hydrogen bonds in the right helix can be mainly observed at the-OH groups on C(4)and C(6)of the mannose residues and C(3)of the glucose residues.In conclusion,molecular dynamic simulation is an efficient method for the microscopic conformation study of glucomannan molecular helices.  相似文献   

16.
A multistate molecular mechanics method is introduced to model the possible competition between various protonation sites in gas-phase biomolecules with excess protons. The method relies on the Amber force field for each site and is calibrated against density-functional theory benchmark calculations at the 6-31+G(d,p) level. In its adiabatic version, where it has similarities with constant-pH algorithms, the model predicts that the small protonated Ala(4)-Lys peptide, unprotected at the N-terminus, changes protonation site above 400 K. In the larger [Ala(9)-Lys+H ](+) peptide, the proton remains at the lysine amine group in a favored charge/electric dipole conformation. In the three-state Ala(4)-Lys-Ala(4)-Lys peptide, the excess proton is found to be partially delocalized on the amine group of the first lysine and on the N-terminus. The statistical properties of the protonated peptides are found to significantly depend on the localized character of the proton. Finally, the model is extended by considering possible couplings between the protonation sites, in an empirical valence-bond version. Strong couplings can stabilize the peptides into unexpected proton-bound conformations over broad ranges of temperature.  相似文献   

17.
Steady-state emission and time-correlated single-photon counting (TCSPC) are used to measure the temperature dependence of the proton-transfer rate of wt-GFP in H2O and D2O. As the temperature decreases, the proton-transfer rate from the protonated form slows down. At about 80 K, the rate is about 10-fold slower than the rate at room temperature. At lower temperatures of 70 K down to 13 K (the lowest temperature studied), the rate of proton transfer is almost temperature independent. We explain the temperature dependence of the proton-transfer rate by an intermolecular vibration assisted tunneling mechanism. We attribute the specific intermolecular vibration to the oscillation of two oxygen atoms: the chromophore's phenol ring and the nearby water molecule. The kinetic isotope effect is about 5 and is almost temperature independent.  相似文献   

18.
Electron transfer from the electron donor of p-phenylenediamine (PPD) to the electron acceptor of (H+)3[PMo12O40]3- forms a one-electron-reduced Keggin cluster of [PMo12O40]4-, bearing a S = 1/2 spin, while proton transfer from the proton donor of (H+)3[PMo12O40]3- to the proton acceptor of PPD yielded mono- and diprotonated cations of 4-aminoanilinium (HPPD+) and p-phenylenediammonium (H2PPD2+). By introduction of crown ether receptors during the crystallization process, supramolecular cations of (HPPD+)(crown ethers) and/or (H2PPD2+)(crown ethers) were successfully introduced into three new alpha-[PMo12O40]4- salts of (H2PPD2+)2([12]crown-4)4[PMo12O40]4- (1), (HPPD+)4([15]crown-5)4[PMo12O40]4- (2), and (HPPD+)2(H2PPD2+)([18]crown-6)4[PMo12O40]4- (3) as the countercation. The protonated states of PPD and molecular-assembly structures of the supramolecular cations depended on the size of the crown ethers. In salt 3, a novel mixed-protonated state of HPPD+ and H2PPD2+ was confirmed to be complexed in the cation structure. According to the changes in the cation structures, the anion arrangements were modulated from those of the two-dimensional layer for salt 1 to the isolated cluster for salts 2 and 3. The temperature-dependent magnetic susceptibilities of salts 1-3 were consistent with the isolated spin arrangements of [PMo12O40]4-. The electronic spectra of salts 1-3 indicated the intervalence optical transition from pentavalent Mo(V) to hexavalent Mo(VI) ions within the [PMo12O40]4- cluster. Temperature-dependent electron spin resonance spectra of salt 2 revealed the delocalization-localization transition of the S = 1/2 spin at 60 K. The spin on the [PMo12O40]4- cluster was localized on a specific Mo(V) site below 60 K, which was thermally activated with an activation energy of 0.015 eV.  相似文献   

19.
The melting of alkyl chains in the saturated fatty acid zinc soaps of different chain lengths, Zn(C(n)H(2n+1)COO)(2); n = 11, 13, 15, and 17, have been investigated by powder X-ray diffraction, differential scanning calorimetry, and vibrational spectroscopy. These compounds have a layer structure with the alkyl chains arranged as tilted bilayers and with all methylene chains adopting a planar, all-trans conformation at room temperature. The saturated fatty acid zinc soaps exhibit a single reversible melting transition with the associated enthalpy change varying linearly with alkyl chain length, but surprisingly, the melting temperature remaining constant. Melting is associated with changes in the conformation of the alkyl chains and in the nature of coordination of the fatty acid to zinc. By monitoring features in the infrared spectra that are characteristic of the global conformation of the alkyl chains, a quantitative relation between conformational disorder and melting is established. It is found that, irrespective of the alkyl chain length, melting occurs when 30% of the chains in the soap are disordered. These results highlight the universal nature of the melting of saturated fatty acid zinc soaps and provide a simple explanation for the observed phenomena.  相似文献   

20.
The protonation of [Ni(SC(6)H(4)R-4)(triphos)](+) (triphos = PhP[CH(2)CH(2)PPh(2)](2); R = NO(2), Cl, H, Me, or MeO) by [lutH](+) (lut = 2,6-dimethylpyridine) to form [Ni(S(H)C(6)H(4)R-4)(triphos)](2+) is an equilibrium reaction in MeCN. Kinetic studies, using stopped-flow spectrophotometry, reveal that the reactions occur by a two-step mechanism. Initially, [lutH](+) rapidly binds to the complex (K(2)(R)) in an interaction which probably involves hydrogen-bonding of the acid to the sulfur. Subsequent intramolecular proton transfer from [lutH](+) to sulfur (k(3)(R)) is slow because of both electronic and steric factors. The X-ray crystal structures of [Ni(SC(6)H(4)R-4)(triphos)](+) (R = NO(2), H, Me, or MeO) show that all are best described as square-planar complexes, with the phenyl substituents of the triphos ligand presenting an appreciable barrier to the approach of the sterically demanding [lutH](+) to the sulfur. The kinetic characteristics of the intramolecular proton transfer from [lutH](+) to sulfur have been investigated. The rate of intramolecular proton transfer exhibits a nonlinear dependence on Hammett sigma(+), with both electron-releasing and electron-withdrawing 4-R-substituents on the coordinated thiolate facilitating the rate of proton transfer (NO(2) > Cl > H > Me < MeO). The rate constants for intramolecular proton transfer correlate well with the calculated electron density of the sulfur. The temperature dependence of the rate of the intramolecular proton transfer reactions shows that deltaH() is small but increases as the 4-R-substituent becomes more electron-withdrawing [deltaH = 4.1 (MeO), 6.9 (Me), 11.4 kcal mol(-)(1) (NO(2))], while DeltaS() becomes progressively less negative [deltaS = -50.1 (MeO), -41.2 (Me), -16.4 (NO(2)) cal K(-)(1) mol(-)(1)]. Studies with [lutD](+) show that the rate of intramolecular proton transfer varies with the 4-R-substituent [(k(3)(NO)2)(H)/(k(3)(NO)2)(D) = 0.39; (k(3)(Cl))(H)/(k(3)(Cl))(D) = 0.88; (k(3)(Me))(H)/(k(3)(Me))(D) = 1.3; (k(3)(MeO))(H)/(k(3)(MeO))(D) = 1.2].  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