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 21 毫秒
1.
The conformational study on Ac‐pSer‐Pro‐NHMe and Ac‐pThr‐Pro‐NHMe peptides has been carried out using hybrid density functional methods with the implicit solvation reaction field theory at the B3LYP/ 6‐311++G(d,p)//B3LYP/6‐31+G(d) level of theory in the gas phase and in solution (chloroform and water). For both pSer‐Pro and pThr‐Pro peptides in the gas phase and in chloroform, the most preferred conformation has the α‐helical structure for the pSer/pThr residue, the down‐puckered polyproline I structure for the Pro residue, and the cis prolyl peptide bond between the two residues, in which two hydrogen bonds between the phosphate oxygens with the backbone N? H groups seem to play a role. However, the trans conformations that have a single hydrogen bond of the phosphate oxygen with either of two backbone N? H groups become most preferred for both peptides in water. This is because the hydration free energy of the anionic oxygen of the phosphate group is expected to dramatically decrease for the cis conformation upon formation of the hydrogen bond with the backbone N? H groups. These calculated results are consistent with the observations by NMR and IR experiments, suggesting the existence of hydrogen bonds between the charged phosphoryl group and the backbone amide protons in solution. The calculated cis populations of 14.7 and 14.2% and rotational barriers of 19.87 and 20.57 kcal/mol to the cis‐to‐trans isomerization for pSer‐Pro and pThr‐Pro peptides in water, respectively, are consistent with the observed values for pSer‐Pro and pThr‐Pro containing peptides from NMR experiments. However, the hydrogen bond between the prolyl nitrogen and the following amide N? H group, which was suggested to be capable of catalyzing the prolyl isomerization, does not play a role in stabilizing the preferred transition state for the pSer/pThr‐Pro peptides in water. Instead, the amide hydrogen of the NHMe group is involved in a bifurcated hydrogen bond with the anionic oxygen and phosphoester oxygen of the phosphate group. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 330–339, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

2.
Structures of (Pro‐Pro‐Gly)4‐Xaa‐Yaa‐Gly‐(Pro‐Pro‐Gly)4 (ppg9‐XYG) where (Xaa, Yaa) = (Pro, Hyp), (Hyp, Pro) or (Hyp, Hyp) were analyzed at high resolution using synchrotron radiation. Molecular and crystal structures of these peptides are very similar to those of the (Pro‐Pro‐Gly)9 peptide. The results obtained in this study, together with those obtained from related compounds, indicated the puckering propensity of the Hyp in the X position: (1) Hyp(X) residues involved in the Hyp(X):Pro(Y) stacking pairs prefer the down‐puckering conformation, as in ppg9‐OPG, and ppg9‐OOG; (2) Hyp(X) residues involved in the Hyp(X):Hyp(Y) stacking pairs prefer the up‐puckering conformation if there is no specific reason to adopt the down‐puckering conformation. Water molecules in these peptide crystals are classified into two groups, the 1st and 2nd hydration waters. Water molecules in the 1st hydration group have direct hydrogen bonds with peptide oxygen atoms, whereas those in the 2nd hydration group do not. Compared with globular proteins, the number of water molecules in the 2nd hydration shell of the ppg9‐XYG peptides is very large, likely due to the unique rod‐like molecular structure of collagen model peptides. In the collagen helix, the amino acid residues in the X and Y positions must protrude outside of the triple helix, which forces even the hydrophobic side chains, such as Pro, to be exposed to the surrounding water molecules. Therefore, most of the waters in the 2nd hydration shell are covering hydrophobic Pro side chains by forming clathrate structures. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 361–372, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

3.
Conformational preferences and prolyl cis?trans isomerizations of the (2S,4S)‐4‐methylproline (4S‐MePro) and (2S,4R)‐4‐methylproline (4R‐MePro) residues are explored at the M06‐2X/cc‐pVTZ//M06‐2X/6‐31+G(d) level of theory in the gas phase and in water, where solvation free energies were calculated using the implicit SMD model. In the gas phase, the down‐puckered γ‐turn structure with the trans prolyl peptide bond is most preferred for both Ac‐4S‐MePro‐NHMe and Ac‐4R‐MePro‐NHMe, in which the C7 hydrogen bond between two terminal groups seems to play a role, as found for Ac‐Pro‐NHMe. Because of the C7 hydrogen bonds weakened by the favorable direct interactions between the backbone C?O and H? N groups and water molecules, the 4S‐MePro residue has a strong preference of the up‐puckered polyproline II (PPII) structure over the down‐puckered PPII structure in water, whereas the latter somewhat prevails over the former for the 4R‐MePro residue. However, these two structures are nearly equally populated for Ac‐Pro‐NHMe. The calculated populations for the backbone structures of Ac‐4S‐MePro‐NHMe and Ac‐4R‐MePro‐NHMe in water are reasonably consistent with CD and NMR experiments. In particular, our calculated results on the puckering preference of the 4S‐MePro and 4R‐MePro residues with the PPII structures are in accord with the observed results for the stability of the (X‐Y‐Gly)7 triple helix with X = 4R‐MePro or Pro and Y = 4S‐MePro or Pro. The calculated rotational barriers indicate that the cis?trans isomerization may in common proceed through the anticlockwise rotation for Ac‐4S‐MePro‐NHMe, Ac‐4R‐MePro‐NHMe, and Ac‐Pro‐NHMe in water. The lowest rotational barriers become higher by 0.24?1.43 kcal/mol for Ac‐4S‐MePro‐NHMe and Ac‐4R‐MePro‐NHMe than those for Ac‐Pro‐NHMe in water. © 2010 Wiley Periodicals, Inc. Biopolymers 95: 51–61, 2011.  相似文献   

4.
The collagen triple helix has a larger accessible surface area per molecular mass than globular proteins, and therefore potentially more water interaction sites. The effect of deuterium oxide on the stability of collagen model peptides and Type I collagen molecules was analyzed by circular dichroism and differential scanning calorimetry. The transition temperatures (Tm) of the protonated peptide (Pro‐Pro‐Gly)10 were 25.4 and 28.7°C in H2O and D2O, respectively. The increase of the Tm of (Pro‐Pro‐Gly)10 measured calorimetrically at 1.0°C min?1 in a low pH solution from the protonated to the deuterated solvent was 5.1°C. The increases of the Tm for (Gly‐Pro‐4(R)Hyp)9 and pepsin‐extracted Type I collagen were measured as 4.2 and 2.2°C, respectively. These results indicated that the increase in the Tm in the presence of D2O is comparable to that of globular proteins, and much less than reported previously for collagen model peptides [Gough and Bhatnagar, J Biomol Struct Dyn 1999, 17, 481–491]. These experimental results suggest that the interaction of water molecules with collagen is similar to the interaction of water with globular proteins, when the ratio of collagen to water is very small and collagen is monomerically dispersed in the solvent. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 93–101, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

5.
We report here the results on N-acetyl-L-proline-N'-methylamide (Ac-Pro-NHMe) calculated at the HF/6-31+G(d) level with the conductor-like polarizable continuum model (CPCM) of self-consistent reaction field methods to investigate the changes of backbone and prolyl ring along the cis-trans isomerization of the prolyl peptide bond. From the potential energy surface, the barrier to ring flip from the down-puckered conformation to the up-puckered one is estimated to be 2.5 and 3.2 kcal/mol for trans and cis conformers of Ac-Pro-NHMe, respectively. In particular, the ring flip seems to be inaccessible in the intermediate regions between trans and cis conformations, because of higher barriers (approximately 13-19 kcal/mol) to rotation of the prolyl peptide bond. The torsion angles for backbone and prolyl ring vary largely around the transition states at omega' approximately 120 degrees and -70 degrees for the prolyl peptide bond. Three kinds of puckering amplitudes show the same trend of puckering along the cis-trans isomerization although their absolute values are different. In particular, trans and cis conformations have the almost same degree of puckering. The cis populations and barriers to rotation of the prolyl peptide bond for Ac-Pro-NHMe are increased with the increase of solvent polarity, which is mainly ascribed to the decreases of relative free energies for cis conformations and the increase of relative free energies for transition states.  相似文献   

6.
Conformational preferences of the (2S,4R)-4-chloroproline (Clp) and (2S,4S)-4-chloroproline (clp) residues are explored at the M06-2X/cc-pVTZ//M06-2X/6-31+G(d) level of theory in the gas phase and in water, where solvation free energies were calculated using the implicit solvation model, and by an X-ray diffraction study in the solid state. In the gas phase, the down-puckered γ-turn structure with the trans prolyl peptide bond is most preferred for both Ac-Clp-NHMe and Ac-clp-NHMe, in which the C(7) hydrogen bond between two terminal groups seems to play a role, as found for Ac-Pro-NHMe. In water, the Clp residue has a strong preference for the up-puckered PP(II) structure, whereas the up-puckered PP(II) structure prevails a little over the down-puckered PP(II) structure for the clp residue, similar to the Pro residue. Hence, our calculated results on the puckering preference of the Clp and clp residues in water are in accord with the observed results deduced from the relative stabilities of the triple helices of the collagen model peptides. The X-ray structure of Ac-clp-NHMe was found to be the most preferred in water but that of Ac-Clp-NHMe was located as a local minimum with ΔG = 2.0 kcal/mol. In particular, the X-ray structure of Ac-Clp-NHMe was quite different from that of Ac-Clp-OMe but similar to that of Ac-Pro-NHMe. The lowest rotational barriers to the prolyl cis-trans isomerization for Ac-Clp-NHMe become nearly the same as those for Ac-Pro-NHMe in water, whereas the barriers are lower by ~2 kcal/mol for Ac-clp-NHMe. It was found that the cis-trans isomerization may proceed through the clockwise or anticlockwise rotations for Ac-Clp-NHMe and the anticlockwise rotation for Ac-clp-NHMe and Ac-Pro-NHMe in water.  相似文献   

7.
The conformational preference and electronic properties of three L ‐tryptophyl‐containing dipeptides, i.e., glycyl‐L ‐tryptophane (H‐Gly‐Trp‐OH), L ‐alanyl‐L ‐tryptophane (H‐Ala‐Trp‐OH), and L ‐methionyl‐L ‐tryptophane (L ‐Met‐Trp‐OH) in solution depending on the pH of the media are studied both theoretically and experimentally. The effect of the protonation of the COO? and deprotonation of the NH as well as the alkaline hydrolysis of the amide fragment in a strong basic media on the electronic spectra are discussed. Ab initio and density functional theory (DFT) methods as well as the time‐dependent DFT (TD‐DFT) method as a function of the basis set are performed with a view to obtain the geometry and electronic properties of all of the species as well as the intermediate, obtained in the alkaline hydrolysis mechanism. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 727–734, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

8.
Mei Zhang 《Biopolymers》2010,93(2):121-131
Recently, we isolated and purified a neutral polysaccharide (PGN) from edible fungus Pleurotus geestanus. Its structure was characterized by a range of physical–chemical methods, including high performance anion exchange chromatography, uronic acid, and protein analyses, size exclusion chromatography with ultraviolet, refractive index and light scattering detectors, and nuclear magnetic resonance. Our results revealed that PGN is a novel β‐(1→3)‐D ‐glucan with glucose attached to every other sugar residues at Position 6 in the backbone. It has a degree of branching of 1/2. Such structure is different from typical β‐(1→3)‐D ‐glucans schizophyllan and lentinan in which DB is 1/3 and 2/5, respectively. Rheological study showed a very interesting melting behavior of PGN in water solution: heating PGN in water leads to two transitions, in the range of 8–12.5°C and 25–60°C, respectively. The melting behavior and conformational changes were characterized by rheometry, micro‐differential scan calorimetry, atomic force microscopy, static and dynamic light scattering at different temperatures. The first heating‐induced transition corresponds to the disintegration of polymer bundles into small helical clusters, resembling the heating‐induced dissociation of SPG in water at 7°C; the second one might correspond to the dissociation of helical strands to individual chains. The ability of PGN to undergo a conformation/viscosity transition in water upon heating is very valuable to immobilize cells or enzymes or therapeutic DNA/RNA, which makes PGN a potentially useful biomaterial. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 121–131, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

9.
Collagen, which is used as a biomaterial, is the most abundant protein in mammals. We have previously reported that a dendrimer modified with collagen model peptides, (Gly‐Pro‐Pro)5, formed a collagen‐like triple‐helical structure, showing thermal reversibility. In this study, various collagen‐mimic dendrimers of different generations and at different binding ratios were synthesized, to investigate the relationship between the peptide clustering effect and the higher order structure formation. The formation of the higher order structure was influenced by the binding ratios of the peptide to the dendrimer, but was not influenced by the dendrimer generation. A spacer, placed between the dendrimer terminal group and the peptide, negatively contributed to the formation of the higher order structure. The collagen model peptides were also attached to poly(allylamine) (PAA) and poly‐L ‐lysine (poly(Lys)) to compare them with the collagen‐mimic dendrimers. The PAA‐based collagen‐mimic compound, bearing more collagen model peptides than the dendrimer, exhibited a thermally stable higher order structure. In contrast, this was not observed for the collagen‐mimic polymers based on poly(Lys). Therefore, dendrimers and vinyl polymers act as a scaffold for collagen model peptides and subsequently induce higher order structures. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 640–648, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

10.
Calorimetric studies were performed on exon 6 in powdered form and in solution [water and 2,2,2‐trifluoroethanol (TFE), a structure‐inducing solvent or cosolvent]. Dynamic dielectric spectroscopy (DDS) analyses were realized in water and 20% TFE. The major role of solvent–peptide organization is evidenced with these techniques. Calorimetric measurements reveal the structural water organization around the polypeptide as well as the presence of hydrophobic interactions in TFE solution. Dielectric measurements showed for exon 6/water a decrease of relaxations times of bulk solvent implying a faster dynamics with a slight increase of the activation entropy, suggesting that exon 6 probably creates disorder within the solvent. For TFE/water mixtures, an influence of exon 6 on its environment was seen with a relaxation associated with the exon 6/solvent interactions reinforced by storage of 72 h. Finally, exon 6/solvent interactions were clearly observed with additionof TFE. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 943–952, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

11.
Is linker DNA bent in the 30‐nm chromatin fiber at physiological conditions? We show here that electrostatic interactions between linker DNA and histone tails including salt condensation and release may bend linker DNA, thus affecting the higher order organization of chromatin. © 2005 Wiley Periodicals, Inc. Biopolymers 81: 20–28, 2006 This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

12.
Thiol proteinase inhibitors are crucial to proper functioning of all living tissues consequent to their cathepsin regulatory and myriad important biologic properties. Equilibrium denaturation of dimeric goat pancreas thiol proteinase inhibitor (PTPI), a cystatin superfamily variant has been studied by monitoring changes in the protein's spectroscopic and functional characteristics. Denaturation of PTPI in guanidine hydrochloride and urea resulted in altered intrinsic fluorescence emission spectrum, diminished negative circular dichroism, and loss of its papain inhibitory potential. Native like spectroscopic properties and inhibitory activity are only partially restored when denaturant is diluted from guanidine hydrochloride unfolded samples demonstrating that process is partially reversible. Coincidence of transition curves and dependence of transition midpoint (3.2M) on protein concentration in guanidine hydrochloride‐induced denaturation are consistent with a two‐state model involving a native like dimer and denatured monomer. On the contrary, urea‐induced unfolding of PTPI is a multiphasic process with indiscernible intermediates. The studies demonstrate that functional conformation and stability are governed by both ionic and hydrophobic interactions. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 708–717, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

13.
As a member of intrinsically unstructured protein family, β‐casein (β‐CN) contains relatively high amount of prolyl residues, adopts noncompact and flexible structure and exhibits chaperone‐like activity in vitro. Like many chaperones, native β‐CN does not contain cysteinyl residues and exhibits strong tendencies for self‐association. The chaperone‐like activities of three recombinant β‐CNs wild type (WT) β‐CN, C4 β‐CN (with cysteinyl residue in position 4) and C208 β‐CN (with cysteinyl residue in position 208), expressed and purified from E. coli, which, consequently, lack the phosphorylated residues, were examined and compared with that of native β‐CN using insulin and alcohol dehydrogenase as target/substrate proteins. The dimers (β‐CND) of C4‐β‐CN and C208 β‐CN were also studied and their chaperone‐like activities were compared with those of their monomeric forms. Lacking phosphorylation, WT β‐CN, C208 β‐CN, C4 β‐CN and C4 β‐CND exhibited significantly lower chaperone‐like activities than native β‐CN. Dimerization of C208 β‐CN with two distal hydrophilic domains considerably improved its chaperone‐like activity in comparison with its monomeric form. The obtained results demonstrate the significant role played by the polar contributions of phosphorylated residues and N‐terminal hydrophilic domain as important functional elements in enhancing the chaperone‐like activity of native β‐CN. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 623–632, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

14.
Young Kee Kang  In Kee Yoo 《Biopolymers》2014,101(11):1077-1087
Conformational preferences of 9‐ and 14‐helix foldamers have been studied for γ‐dipeptides of 2‐aminocyclohexylacetic acid (γAc6a) residues such as Ac‐(γAc6a)2‐NHMe ( 1 ), Ac‐(Cα‐Et‐γAc6a)2‐NHMe ( 2 ), Ac‐(γAc6a)2‐NHBn ( 3 ), and Ac‐(Cα‐Et‐γAc6a)2‐NHBn ( 4 ) at the M06‐2X/cc‐pVTZ//M06‐2X/6‐31 + G(d) level of theory to explore the influence of substituents on their conformational preferences. In the gas phase, the 9‐helix foldamer H9 and 14‐helix foldamer H14‐z are found to be most preferred for dipeptides 2 and 4 , respectively, as for dipeptides 1 and 3 , which indicates no remarkable influence of the Cα‐ethyl substitution on conformational preferences. The benzyl substitution at the C‐terminal end lead H14‐z to be the most preferred conformer for dipeptides 3 and 4 , whereas it is H9 for dipeptides 1 and 2 , which can be ascribed to the favored C? H···π interactions between the cyclohexyl group of the first residue and the C‐terminal benzyl group. There are only marginal changes in backbone structures and the distances and angles of H‐bonds for all local minima by Cα‐ethyl and/or benzyl substitutions. Although vibrational frequencies and intensities of the dipeptide 4 calculated at both M06‐2X/6‐31 + G(d) and M05‐2X/6‐31 + G(d) levels of theory are consistent with observed results in the gas phase, H14‐z is predicted to be most preferred by ΔG only at the former level of theory. Hydration did not bring the significant changes in backbone structures of helix foldamers for both dipeptide 1 and 4 . It is expected that the different substitutions at the C‐terminal end lead to the different helix foldamers, which may increase the resistance of helical structures to proteolysis and provide the more surface to the helical structures suitable for molecular recognition. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1077–1087, 2014.  相似文献   

15.
Electronic and vibrational circular dichroism are often used to determine the secondary structure of proteins, because each secondary structure has a unique spectrum. Little is known about the vibrational circular dichroic spectroscopic features of the β‐hairpin. In this study, the VCD spectral features of a decapeptide, YYDPETGTWY (CLN025), which forms a stable β‐hairpin that is stabilized by intramolecular weakly polar interactions and hydrogen bonds were determined. Molecular dynamics simulations and ECD spectropolarimetry were used to confirm that CLN025 adopts a β‐hairpin in water, TFE, MeOH, and DMSO and to examine differences in the secondary structure, hydrogen bonds, and weakly polar interactions. CLN025 was synthesized by microwave‐assisted solid phase peptide synthesis with Nα‐Fmoc protected amino acids. The VCD spectra displayed a (?,+,?) pattern with bands at 1640 to 1656 cm?1, 1667 to 1687 cm?1, and 1679 to 1686 cm?1 formed by the overlap of a lower frequency negative couplet and a higher frequency positive couplet. A maximum IR absorbance was observed at 1647 to 1663 cm?1 with component bands at 1630 cm?1, 1646 cm?1, 1658 cm?1, and 1675 to 1680 cm?1 that are indicative of the β‐sheet, random meander, either random meander or loop and turn, respectively. These results are similar to the results of others, who examined the VCD spectra of β‐hairpins formed by DPro‐Xxx turns and indicated that observed pattern is typical of β‐hairpins. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 442–450, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

16.
Recently, ubiquitin was suggested as a promising anti‐inflammatory protein therapeutic. We found that a peptide fragment corresponding to the ubiquitin50–59 sequence (LEDGRTLSDY) possessed the immunosuppressive activity comparable with that of ubiquitin. CD and NMR spectroscopies were used to determine the conformational preferences of LEDGRTLSDY in solution. The peptide mixture, obtained by pepsin digestion of ubiquitin, was even more potent than the intact protein. Although the peptide exhibited a well‐defined conformation in methanol, its structure was distinct from the corresponding 50–59 fragment in the native ubiquitin molecule. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 423–431, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

17.
Synthetic peptides corresponding to the sixth transmembrane segment (TMS6) of secondary‐active transporter MntH (Proton‐dependent Manganese Transporter) from Escherichia coli and its two mutations in the functionally important conserved histidine residue were used as a model for structure–function study of MntH. The secondary structure of the peptides was estimated in different environments using circular dichroism spectroscopy. These peptides interacted with and adopted helical conformations in lipid membranes. Electrophysiological experiments demonstrated that TMS6 was able to form multi‐state ion channels in model biological membranes. Electrophysiological properties of these weakly cation‐selective ion channels were strongly dependent on the surrounding pH. Manganese ion, as a physiological substrate of MntH, enhanced the conductivity of TMS6 channels, influenced the transition between closed and open states, and affected the peptide conformations. Moreover, functional properties of peptides carrying two different mutations of His211 were analogous to in vivo functional characteristics of Nramp/MntH proteins mutated at homologous residues. Hence, a single functionally important TMS can retain some of the functional properties of the full‐length protein. These findings could contribute to understanding the structure–function relationship at the molecular level. However it remains unclear to what extent the peptide‐specific channel activity represents a functional aspect of the full‐length membrane carrier protein. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 718–726, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

18.
19.
Aminoglycoside phosphotransferase(3′)‐IIIa (APH) is the enzyme with broadest substrate range among the phosphotransferases that cause resistance to aminoglycoside antibiotics. In this study, the thermodynamic characterization of interactions of APH with its ligands are done by determining dissociation constants of enzyme–substrate complexes using electron paramagnetic resonance and fluorescence spectroscopy. Metal binding studies showed that three divalent cations bind to the apo‐enzyme with low affinity. In the presence of AMPPCP, binding of the divalent cations occurs with 7‐to‐37‐fold higher affinity to three additional sites dependent on the presence and absence of different aminoglycosides. Surprisingly, when both ligands, AMPPCP and aminoglycoside, are present, the number of high affinity metal binding sites is reduced to two with a 2‐fold increase in binding affinity. The presence of divalent cations, with or without aminoglycoside present, shows only a small effect (<3‐fold) on binding affinity of the nucleotide to the enzyme. The presence of metal–nucleotide, but not nucleotide alone, increases the binding affinity of aminoglycosides to APH. Replacement of magnesium (II) with manganese (II) lowered the catalytic rates significantly while affecting the substrate selectivity of the enzyme such that the aminoglycosides with 2′‐NH2 become better substrates (higher Vmax) than those with 2′‐OH. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 801–809, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

20.
Our previous studies of the potential utility of the CαDα stretch frequency, ν(CD), as a tool for determining conformation in peptide systems (Mirkin and Krimm, J Phys Chem A 2004, 108, 10923–10924; 2007, 111, 5300–5303) dealt with the spectroscopic characteristics of isolated alanine peptides with αR, β, and polyproline II structures. We have now extended these ab initio calculations to include various explicit‐water environments interacting with such conformers. We find that the structure‐discriminating feature of this technique is in fact enhanced as a result of the conformation‐specific interactions of the bonding waters, in part due to our finding (Mirkin and Krimm, J Phys Chem B 2008, 112, 15268) that Cα? Dα…O(water) hydrogen bonds can be present in addition to those expected between water and the CO and NH of the peptide groups. In fact, ν(CD) is hardly affected by the latter bonding but can be shifted by up to 70 cm?1 by the former hydrogen bonds. We also discuss the factors that will have to be considered in developing the molecular dynamics (MD) treatment needed to satisfactorily take account of the influence of outer water layers on the structure of the first‐layer water molecules that hydrogen bond to the peptide backbone. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 791–800, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  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