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 687 毫秒
1.
A block copolymer [γ-Et-DL -Glu]m [L -Trp]n was prepared using N-carboxy anhydrides (NCA) of L -tryptohan and γ-ethyl DL -glutamate. The block copolymer, dissolved in trifluoroethanol (TFE)–dichloroacetic acid (DCA) mixtures, exhibited a sharp change in the specific rotation at 546 mμ when the solvent composition reached 70–75% DCA content. Optical rotatory dispersion (ORD) and circular dichroism (CD) measurement were carried out in TFE solution in the spectral range 180–350 mμ. Indole side-chain chromophores were found to be optically active in the polymer. On the other hand, these groups exhibit very small optical activity in the model compound C6H3? CH2? O? CO? (L -Trp)2? O? CH3. Indole groups therefore appear to be in a dissymmetric environment only in the polymer. From these data it was concluded that poly-L -Trp is in some type of helical conformation in TFE. Strong overlapping of CD bands from side-chain chromophores and peptides chromophores in the wavelength range 185–240 mμ does not allow definite conclusions to be drawn about the type of helical conformation which exists in poly-L -Trp in TFE solution.  相似文献   

2.
Min Zhang  Tuck C. Wong 《Biopolymers》1993,33(12):1901-1908
High-resolution proton spectra at 500 MHz of two tachykinin peptides, substance P methyl ester (SPOMe) and [Nle10]-neurokinin A (4–10), have been obtained in dimethylsulfoxide (DMSO), and for SPOMe, also in 2, 2, 2-trifluoroethanol (TFE)/water mixtures. Complete chemical shift assignments for these peptides were made based on two-dimensional (2D) nmr techniques, correlated spectroscopy and total COSY. J coupling measurement and nuclear Overhauser effect spectroscopy (NOESY) were then used to determine the conformation of these peptides in the various solvents. Based on the J coupling, NOE correlations, and temperature coefficients of the NH resonances, it is concluded that these two peptides exist in DMSO at room temperature as a mixture of conformers that are primarily extended. For SPOMe in TFE/water with high TFE content, however, helical structures are found to be present, and they become quite clear at temperatures between 270 and 280 K. The variation of the 13C chemical shifts of the Cα (the secondary shift) with TFE contents corroborates this conclusion. The NOE and Cα shifts show that the main helical region for SPOMe lies between 4P and 9G. The C-terminus segment L? M? NH2 is found to be quite flexible, which appears to be quite common for neurokinin-1 selective peptides. © 1993 John Wiley & Sons, Inc.  相似文献   

3.
CD spectroscopic studies on protected peptides containing lysine and serine, or phosphoserine, and on serine-containing fragments of the neurofilament protein midsized subunit, both in the unphosphorylated and phosphorylated form, are reported. The introduction of the phosphoryl group was not found to have a significant spectral effect in aqueous solution. In trifluoroethanol (TFE), spectral shifts toward unordered (type U) spectra or the appearance of distorted spectra likely reflect the adoption of aperiodic polypeptide conformations due to salt bridge(s) between negatively charged phosphoserine and positive lysine side-chain groups. A turn-stabilizing effect of phosphorylation was also observed. CD-monitored titration experiments in TFE revealed a high conformational sensitivity of phosphopeptides toward Ca2+ ions. The appearance of the unordered spectra or spectral shifts were the sign of a bulk disordering effect of Ca2+ ions. Spectra with specific spectroscopic features reflect the formation of Ca2+complexes and the adoption of ordered unique backbone conformations. When ordered structures were obtained on addition of Ca2+ ions, the observed CD curves showed a resemblance to the spectrum of β-pleated sheets. This may originate from chain extension and the formation of β-pleated sheet segments fixed by Ca2+ bridges between PO3H groups of adjacent peptide chains. The data clearly show that the effect of the Ca2+ ions is highly specific: the sequence, chain length, presence and distribution of charged side-chain groups, degree and site of phosphorylation, and environmental factors appear to be determining in the process of chain extension or β-sheet formation. © 1993 John Wiley & Sons, Inc.  相似文献   

4.
The CD spectra of the peptides Boc-X-(Aib-X)n-OMe (n = 1, 2, 3) and Boc-(Aib-X)5-OMe, where X = L -Ala or L -Val have been examined in several solvents. The X = Ala and Val peptides behave similarly in all solvents, suggesting that the Aib residues dominate the folding preferences of these peptides. The decapeptides adopt helical conformations in methanol and trifluoroethanol, with characteristic negative CD bands at 222 and 205 nm. In the heptapeptides, similar spectra with reduced intensities are observed. Comparison with nmr studies suggest that estimates of helical content in oligopeptides by CD methods may lead to erroneous conclusions. The pentapeptides yield solvent-dependent spectra indicative of conformational perturbations. Peptide association in dioxane results in an unusual spectrum with a single negative band at 210 nm for the decapeptides. Disaggregation is induced by the addition of methanol or water to dioxane solutions. Aggregation of the heptapeptides is less pronounced in dioxane, suggesting that a critical helix length may be necessary to promote association stabilized by helix dipole–dipole interactions.  相似文献   

5.
CD and nmr studies have been carried out on aqueous trifluoroethanol (TFE) solutions of bradykinin (BK) and a bradykinin antagonist. The CD results exhibit a striking effect of TFE on the spectra of BK, with sequence Arg-Pro-Pro-Gly-Phe-Ser-Pro-Phe-Arg, and the BK antagonist, with sequence D -Arg-Arg-Pro-Hyp-Gly-Thi-D -Ser-D -Cpg-Cpg-Arg [where Hyp is 4-hydroxy-L -proline; Thi refers to β-(2-thienyl)-L -alanine and Cpg refers to α-cyclopentylglycine]. The effect of increasing concentration of TFE in water on the difference ellipticity at 222 nm was examined and showed that BK may be a mixture of at least two different conformers, one of which largely forms when the TFE concentration is increased beyond 80%. The linear extrapolation of 100% of the difference ellipticity of BK at low TFE concentrations yields a value in agreement with that shown by the BK antagonist, indicating that the conformation of BK at the lower TFE concentrations is similar to that of the BK antagonist. The conformational analysis was carried out using both one-dimensional and two-dimensional 1H-nmr techniques. The total correlation spectroscopy (TOCSY) spectrum of BK in a 60/40% (v/v) TFE/H2O solution at 10°C and a nuclear Overhauser effect spectroscopy (NOESY) spectrum that shows only sequential Hα(i) – NH(i + 1) or the Hα(i) – Hδδ′(i + 1) NOEs indicate that the majority of the molecules adopt an all-trans extended conformation. The TOCSY for BK in the 95/5% (v/v) TFE/H2O solution shows that there are two major conformations in the solution with about equal population. The NOESY experiment shows two new important cross peaks for one conformation, namely Pro2(α)-Pro3 (α) and the Pro2(α)-Gly4(NH), indicating a cis Pro2-Pro3 bond and a type VI β-turn between residues Arg1 and Gly4 involving cis proline at position 3, respectively. The low temperature coefficient of Gly4 for this conformation suggests the presence of an intramolecular hydrogen bond, therefore a type VIa β-turn is present. The other conformation is all trans and extended. The BK antafonist shows difference CD spectra in TFE solutions referred to H2O that are superficially indicative of a β-bend. However, nmr speaks against this possibility, as only one set of peaks were observed in the TOCSY and NOESY experiments, indicating an all-trans extended confirmation over the range of TFE concentrations. The BK-antagonist CD data suggest that solvent perturbation of the CD of an extended confirmation perturbation of the optical activity of the thienyl moiety of the peptide since the CD spectrum of N-acetyl-β-thienyl-L -alanine N-methylamide is strongly perturbed by TFE. The present results again demonstrate the complementary relationship between CD and nmr. © 1994 John Wiley & Sons, Inc.  相似文献   

6.
The structures of the first and the second transmembrane segment of the bovine mitochondrial oxoglutarate carrier (OGC) were studied by circular dichroism (CD) and nuclear magnetic resonance (NMR) spectroscopies. Peptides 21–46 and 78–108 of its primary sequence were synthesized and structurally characterized in membrane-mimetic environments. CD data showed that at high concentrations of TFE (>50%) and SDS (>2%) both peptides assume α-helical structures, whereas in more hydrophilic environments only peptide 78–108 has a helical structure. 1H-NMR spectra of the two peptides in TFE/water and SDS were fully assigned, and the secondary structures of the peptides were obtained from nuclear Overhauser effects, 3JαH-NH coupling constants and αH chemical shifts. The three-dimensional solution structures of the peptides in TFE/water were generated by distance geometry calculations. A well-defined α-helix was found in the region K24-V39 of peptide 21–46 and in the region A86–F106 of peptide 78–108. We cannot exclude that in intact OGC the extension of these helices is longer. The helix of peptide 21–46 is essentially hydrophobic, whereas that of peptide 78–108 is predominantly hydrophilic.  相似文献   

7.
Protein folding problem remains a formidable challenge as main chain, side chain and solvent interactions remain entangled and have been difficult to resolve. Alanine‐based short peptides are promising models to dissect protein folding initiation and propagation structurally as well as energetically. The effect of N‐terminal diproline and charged side chains is assessed on the stabilization of helical conformation in alanine‐based short peptides using circular dichroism (CD) with water and methanol as solvent. A1 (Ac–Pro–Pro–Ala–Lys–Ala–Lys–Ala–Lys–Ala–NH2) is designed to assess the effect of N‐terminal homochiral diproline and lysine side chains to induce helical conformation. A2 (Ac–Pro–Pro–Glu–Glu–Ala–Ala–Lys–Lys–Ala–NH2) and A3 (Ac–d Pro–Pro–Glu–Glu–Ala–Ala–Lys–Lys–Ala–NH2) with N‐terminal homochiral and heterochiral diproline, respectively, are designed to assess the effect of Glu...Lys (i , i  + 4) salt bridge interactions on the stabilization of helical conformation. The CD spectra of A1 , A2 and A3 in water manifest different amplitudes of the observed polyproline II (PPII) signals, which indicate different conformational distributions of the polypeptide structure. The strong effect of solvent substitution from water to methanol is observed for the peptides, and CD spectra in methanol evidence A2 and A3 as helical folds. Temperature‐dependent CD spectra of A1 and A2 in water depict an isodichroic point reflecting coexistence of two conformations, PPII and β‐strand conformation, which is consistent with the previous studies. The results illuminate the effect of N‐terminal diproline and charged side chains in dictating the preferences for extended‐β, semi‐extended PPII and helical conformation in alanine‐based short peptides. The results of the present study will enhance our understanding on stabilization of helical conformation in short peptides and hence aid in the design of novel peptides with helical structures. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
A series of 14 residue amphipathic α-helical peptides, in which the sidechains of glutamic acid and lysine have been covalently joined, was synthesized in order to determine the effect of spacing, position and orientation of these lactam bridges. It was found that although an (i, i+3) spacing would position the lactam bridge on the same face of the helix, these lactams with 18-member rings were actually helix-destabilizing regardless of position or location. On the other hand, (i, i+4) lactams with 21-member rings were helix-stabilizing but this was dependent on orientation. Glutamic acid-lysine lactams increased the helical content of the peptide when compared with their linear homologue in benign conditions (50 mM KH2PO4, 100 mM KCl, pH 7). Two Glu-Lys (i, i+4) lactams located at the N- and C-termini gave rise to a peptide with greater than 99% helical content in benign conditions. Peptides with Lys-Glu oriented lactams were random structures in benign conditions but in the presence of 50% TFE could be induced into a helical conformation. The stability of the single-stranded α-helices, as measured by thermal denaturations in 25% TFE indicated that Glu-Lys oriented lactam bridges stabilized the helical conformation relative to the linear unbridged peptide. One Glu-Lys lactam in the middle of the peptide was more effective at stabilizing helical structure than two Glu-Lys lactams positioned one at each end of the molecule. The lactams with the Lys-Glu orientation were destabilizing relative to the unbridged peptide. This study demonstrates that correct orientation and position of a lactam bridge is critical in order to design peptides with high helical content in aqueous media.  相似文献   

9.
Bombolitins are five naturally occurring heptadecapeptides acting at the membrane level and able to increase the activity of phospholipase A2. As for other peptides with similar function, the biological activity of bombolitins seems to be mainly due to their ability to form amphipathic helical structures. We synthesized and tested the retro sequence of bom-bolitin I (retro-bombolitin I). This peptide showed an activity similar to that of the natural sequence and was able to adopt a helical structure in the presence of an amphipathic environment consisting of SDS micelles. The secondary structure of this peptide was fully characterized by CD and nmr spectroscopy. © 1994 John Wiley & Sons, Inc.  相似文献   

10.
In order to better understand the driving forces that determine the alignment of amphipathic helical polypeptides with respect to the surface of phospholipid bilayers, lysine-containing peptide sequences were designed, prepared by solid-phase chemical synthesis, and reconstituted into membranes. CD spectroscopy indicates that all peptides exhibit a high degree of helicity in the presence of SDS micelles or POPC small unilamellar vesicles. Proton-decoupled (31)P-NMR solid-state NMR spectroscopy demonstrates that in the presence of peptides liquid crystalline phosphatidylcholine membranes orient well along glass surfaces. The orientational distribution and dynamics of peptides labeled with (15)N at selected sites were investigated by proton-decoupled (15)N solid-state NMR spectroscopy. Polypeptides with a single lysine residue adopt a transmembrane orientation, thereby locating this polar amino acid within the core region of the bilayer. In contrast, peptides with > or = 3 lysines reside along the surface of the membrane. With 2 lysines in the center of an otherwise hydrophobic amino acid sequence the peptides assume a broad orientational distribution. The energy of lysine discharge, hydrophobic, polar, and all other interactions are estimated to quantitatively describe the polypeptide topologies observed. Furthermore, a molecular modeling algorithm based on the hydrophobicities of atoms in a continuous hydrophilic-hydrophobic-hydrophilic potential describes the experimentally observed peptide topologies well.  相似文献   

11.
Antibacterial properties of the secretion from the female reproductive accessory glands of medfly Ceratitis capitata are mostly ascribed to the presence of two peptides, ceratotoxin A and B, which exhibit a strong activity against gram-positive and gram-negative bacterial strains, and show sequence and function homology with cecropins, melittin, and magainins. CD experiments performed in different solvents indicate the presence of a significant content of helical structures in organic solvent. Two-dimensional nmr results for ceratotoxin A in methanol show a helical behavior for the 8–25 region of the peptide. A Ramachandran classification of each residue for the structures obtained from distance geometry calculations lead to the definition of four structural families in which the central segment 10–19 is always helical and differences refer to residues 8–9 and 19–23. A sequence analysis of the two ceratotoxins and a systematic search on the protein data bank revealed the occurrence of a KX-hydrophobic-hydrophobic-P motif that seems to be important for helix stabilization. © 1996 John Wiley & Sons, Inc.  相似文献   

12.
The solution structure of a peptide fragment corresponding to the 38–59 region of porcine phospholipase A2 has been investigated using CD, nmr chemical shifts, and nuclear over-hauser effects (NOEs). This isolated fragment of phospholipase forms an α-helix spanning residues 38–55, very similar to the one found in the native protein, except for residues 56–58, which were helical in the crystal but found random in solution. Addition of triflouro-ethanol (TFE) merely increased helix population but it did not redefine helix limits. To investigate how the folding information, in particular that concerning eventual helix start and stop signals, was coded in this particular amino acid sequence, the helices formed by synthetic peptides reproducing sections of this phospholipase 38–59 fragment, namely 40–59, 42–59, 38–50, and 45–57, were characterized using NOEs and helix populations quantitatively evaluated on different peptide chain segments using nmr chemical shifts in two solvents (H2O and 30% TFE/H2O). A set of nmr spectra was also recorded and assigned under denaturing conditions (6Murea) to obtain reliable values for the chemical shifts of each peptide in the random state. Based on chemical shift data, it was concluded that the helix formed by the phospholipase 38–59 fragment was not abruptly, but progressively, destabilized all along its length by successive elimination of residues at the N end, while the removal of residues at the C end affected helix stability more locally and to a lesser extent. These results are consistent with the idea that there are not single residues responsible for helix initiation or helix stability, and they also evidence an asymmetry for contributions to helix stability by residues located at the two chain ends. The restriction of molecular mobility caused by linking with a disulphide bridge at Cys 51 two identical 38–59 peptide chains did not increase helix stability. The helix formed by the covalently formed homodimer was very similar in length and population to that formed by the monomer. © 1994 John Wiley & Sons, Inc.  相似文献   

13.
The structures of the fifth and sixth transmembrane segments of the bovine mitochondrial oxoglutarate carrier (OGC) and of the hydrophilic loop that connects them were studied by CD and NMR spectroscopies. Peptides F215-R246, W279-K305 and P257-L278 were synthesized and structurally characterized. CD data showed that at high concentrations of TFE and SDS all peptides assume α-helical structures. 1H-NMR spectra of the three peptides in TFE/water were fully assigned and the secondary structures of the peptides were obtained from nuclear Overhauser effects, 3JαH-NH coupling constants and αH chemical shifts. The three-dimensional solution structures of the peptides were generated by distance geometry calculations. A well-defined α–helix was found in the region L220-V243 of peptide F215-R246 (TMS-V), in the region P284-M303 of peptide W279-K305 (TMS-VI) and in the region N261-F275 of peptide P257-L278 (hydrophilic loop). The helix L220-V243 exhibited a sharp kink at P239, while a little bend around P291 was observed in the helical region P284-M303. Fluorescence studies performed on peptide W279-K305, alone and together with other transmembrane segments of OGC, showed that the W279 fluorescence was quenched upon addition of peptide F215-R246, but not of peptides K21-K46, R78-R108 and P117-A149 suggesting a specific interaction between TMS-V and TMS-VI of OGC.  相似文献   

14.
R W Storrs  D Truckses  D E Wemmer 《Biopolymers》1992,32(12):1695-1702
Helix propagation of the S-peptide sequence (residues 1-19 of ribonuclease A) in 2,2,2-trifluoroethanol (TFE) solutions has been investigated with CD and nmr Overhauser effect spectroscopies. In this study, the S-peptide helix is covalently initiated at the N-terminus through disulfide bonds to a helix scaffold derived from the N-terminal sequence of the bee venom peptide apamin. The entire S-peptide sequence of this hybrid sequence peptide becomes helical at high proportions of TFE. Residues 14-19 of the S-peptide are not helical in the free peptide in TFE, nor are they helical in ribonuclease A. The "helix stop" signal encoded by the S-peptide sequence near residue 13 does not persist at high TFE with this hybrid sequence peptide. The helix-stabilizing effects of TFE are due at least in part to facilitated propagation of an extant helix. This stabilizing effect appears to be a general solvation effect and not due to specific interaction of the helical peptide with TFE. Specifically these data support the idea that TFE destabilizes the coil state by less effective hydrogen bonding of the peptide amide to the solvent.  相似文献   

15.
The N‐terminal 1–34 segments of both parathyroid hormone (PTH) and parathyroid hormone‐related protein (PTHrP) bind and activate the same membrane receptor in spite of major differences between the two hormones in their amino acid sequence. Recently, it was shown that in (1–34)PTH/PTHrP segmental hybrid peptides, the N‐terminal 1–14 segment of PTHrP is incompatible with the C‐terminal 15–34 region of PTH leading to substantial reduction in potency. The sites of incompatibility were identified as positions 5 in PTH and 19 in PTHrP. In the present paper we describe the synthesis, biological evaluation, and conformational characterization of two point‐mutated PTH/PTHrP 1–34 hybrids in which the arginine residues at positions 19 and 21 of the native sequence of PTHrP have been replaced by valine (hybrid V21) and glutamic acid (hybrid E19), respectively, taken from the PTH sequence. Hybrid V21 exhibits both high receptor affinity and biological potency, while hybrid E19 binds weakly and is poorly active. The conformational properties of the two hybrids were studied in aqueous solution containing dodecylphosphocholine (DPC) micelles and in water/2,2,2‐trifluoroethanol (TFE) mixtures. Upon addition of TFE or DPC micelles to the aqueous solution, both hybrids undergo a coil‐helix transition. The maximum helix content in 1 : 1 water/TFE, obtained by CD data for both hybrids, is ∼ 80%. In the presence of DPC micelles, the maximum helix content is ∼ 40%. The conformational properties of the two hybrids in the micellar system were further investigated by combined 2D‐nmr, distance geometry (DG), and molecular dynamics (MD) calculations. The common structural motif, consisting of two helical segments located at N‐ and C‐termini, was observed in both hybrids. However, the biologically potent hybrid V21 exhibits two flexible sites, centered at residues 12 and 19 and connecting helical segments, while the flexibility sites in the weakly active hybrid E19 are located at position 11 and in the sequence 20–26. Our findings support the hypothesis that the presence and location of flexibility points between helical segments are essential for enabling the active analogs to fold into the bioactive conformation upon interaction with the receptor. © 1999 John Wiley & Sons, Inc. Biopoly 50: 525–535, 1999  相似文献   

16.
1H-nmr spectra for a series of Boc-L -(Met)n-OMe (n = 2–9) homo-oligopeptides have been observed in the helix-supporting solvent trifluoroethanol (TFE) at millimolar concentrations. Interfering solvent peaks were eliminated using two decoupling frequencies to selectively remove the methylene and hydroxyl protons of the solvent. Comparisons with specifically α-deuterated homo-oligopeptides gave complete assignments of the NH region of the Boc-Metn-OMe oligomers up to the heptapeptide. Analysis of chemical shifts, coupling constants, and temperature dependence of chemical shifts suggests that up to the hexapeptide, similar structures exist in deuterochloroform and TFE. In contrast, nmr parameters at the heptapeptide for several internal residues differ in these solvents. These results suggest that a C7 to α-helix transition may occur in TFE as the chain length of the methionine oligopeptides increases.  相似文献   

17.
We synthesized porcine neuropeptide Y (pNPY) N-terminal fragments by solid-phase synthesis techniques and analyzed them for solution Conformational properties by CD and 1H-nmr spectroscopy. The analogues pNPY1–9 and pNPY1–14 displayed CD spectra indicative of random structures and showed no evidence for induced α-helical structures in trifluoroethanol (TFE) up to 50%. However, the CD spectra of pNPY1-9 suggested a Conformational shift in tetrahydrofuran. Although in aqueous solution the CD spectra of pNPY1–21 indicated random structures with induction of only a small percentage of α-helix in aqueous TFE, pNPY1-25 displayed 13% a-helical structure in aqueous solution that increased to 40 and 41% by the addition of TFE and methanol, respectively. The nmr spectra of pNPY1-9 and the proline region of pNPY1–25 indicated extended structures with all-trans conformers at Pro5 and Pro8 for pNPY1–9 and at Pro5, Pro8, and Pro13 for pNPY1–25; in each case the Tyrl-Pro2 amide bond was in both cis and trans conformations. However, observed nuclear Overhauser effect correlations and UN exchange experiments indicated an α-helical segment in pNPY1–25 initiated by Pro 13 and extending from residues 14 to 25. Thus, the N-terminal polyproline region of NPY has no propensity to fold into a regular secondary structure, although Pro 13 is a helix initiator, a result consistent with the proposed role of this amino acid in the NPY structural model. © 1995 John Wiley & Sons, Inc.  相似文献   

18.
Abstract

NMR and CD spectroscopy have been used to examine the conformation of the peptide, β(12–28), (VHHQKLVFFAEDVGSNK) in aqueous and 60% TFE/40% H20 solution at pH 2.4. In 60% TFE solution, the peptide is helical as confirmed by the CD spectrum and by the pattern of the NOE cross peaks detected in the NOESY spectrum of the peptide. In aqueous solution, the peptide adopts a more extended and flexible conformation. Broadening of resonances at low temperature, temperature-dependent changes in the chemical shifts of several of the CHα resonances and the observation of a number of NOE contacts between the hydrophobic side-chain protons of the peptide are indicative of aggregation in aqueous solution. The behavior of β(12–28) in 60% TFE and in aqueous solution are consistent with the overall conformation and aggregation behavior reported for the larger peptide fragment, β(1–28) and the parent β-amyloid peptide.  相似文献   

19.
RNA and DNA helicases manipulate or translocate along single strands of nucleic acids by grasping them using a conserved structural motif. We have examined the available crystal structures of helicases of the two principal superfamilies, SF1 and SF2, and observed that the most conserved interactions with the nucleic acid occur between the phosphosugar backbone of a trinucleotide and the three strand‐helix loops within a (β‐strand/α‐helix)3 structural module. At the first and third loops is a conserved hydrogen‐bonded feature called a thr‐motif, often seen at α‐helical N‐termini, with the threonine as the N‐cap residue. These loops can be aligned with few insertions or deletions, and their main chain atoms are structurally congruent amongst the family members and between the two modules found as tandem pairs in all SF1 and SF2 proteins. The other highly conserved interactions with nucleic acid involve mainchain NH groups, often at the helical N‐termini, interacting with phosphate groups. We comment on how the sequence motifs that are commonly used to identify helicases map to locations on the module and discuss the implications of the conserved orientation of nucleic acid on the surface of the module for directional stepping along DNA or RNA. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

20.
The receptor for the α-factor mating pheromone of the yeast Saccharomyces cerevisiae consists of 431 amino acid residues and is a member of a family of membrane proteins predicted to have seven transmembrane helices. Fragments of the receptor corresponding to two of the transmembrane helices [residues 246–269 (M6) and 273–302 (M7)], two of the interhelical loops [residues 107–125 (E2) and 191–206 (E3)], and to a portion of the carboxyl terminus [residues 350–372 (CT)] were synthesized using solid-phase methodologies and purified to near homogeneity. CD was used to characterize the secondary structure of these peptides in trifluoroethanol (TFE), in TFE/water mixtures, in sodium dodecyl sulfate (SDS), and in the presence of dimyristoyl phosphatidylcholine (DMPC) liposomes. In TFE, M6 and M7 exhibited CD spectra consistent with highly helical peptides, whereas CT was partially helical. In contrast, E2 and E3 were either disordered or aggregated in this solvent. M6 did not partition well into DMPC vesicles whereas M7 remained helical. Both M6 and M7 assumed helical conformations in 25 mM SDS. The loop neptides and the carboxyl terminus peptide were either in a β-structure or disordered in the presence of lipid. These findings represent the first biophysical evidence for conformations assumed by specific segments of the STE2 receptor protein. © 1994 John Wiley & Sons, Inc.  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