首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The crystal state conformations of three peptides containing the α,α-dialkylated residues. α,α-di-n-propylglycine (Dpg) and α,α-di-n-butylglycine (Dbg), have been established by x-ray diffraction. Boc-Ala-Dpg-Alu-OMe (I) and Boc-Ala-Dbg-Ala-OMe (III) adopt distorted type II β-turn conformations with Ala (1) and Dpg/Dbg (2) as the corner residues. In both peptides the conformational angles at the Dxg residue (I: ? = 66.2°, ψ = 19.3°; III: ? = 66.5°. ψ = 21.1°) deviate appreciably from ideal values for the i + 2 residue in a type II β-turn. In both peptides the observed (N…O) distances between the Boc CO and Ala (3) NH groups are far too long (1: 3.44 Å: III: 3.63 Å) for an intramolecular 4 → 1 hydrogen bond. Boc-Ala-Dpg-Ata-NHMe (II) crystallizes with two independent molecules in the asymmetric unit. Both molecules HA and HB adopt consecutive β-turn (type III-III in HA and type III-I in IIB) or incipient 310-helical structures, stabilized by two intramolecular 4 → 1 hydrogen bonds. In all four molecules the bond angle N-Cα-C′ (τ) at the Dxg residues are ≥ 110°. The observation of conformational angles in the helical region of ?,ψ space at these residues is consistent with theoretical predictions. © 1995 John Wiley & Sons, Inc.  相似文献   

2.
Malonic and diaminomethane residues, equivalent to the two possible retro modifications of a glycine unit, with an inverted peptide group, present particular conformations that differ from those found in glycine and, in general, in α-amino acids. In both cases the φi and ψi torsional angles have restricted values as deduced from inspection of the Cambridge Structural Data Bank and from compounds studied by us. Thus, both ψi angles tend to be equal to 115° (or −115°) in the malonyl residues, whereas the φi angles tend to be equal to 88° (or −88°) in the diaminomethane residues. These results are in agreement with previous experimental data on polymers, but in the case of malonyl residues they differ from theoretical calculations on isolated molecules. The experimental data for both residues can be represented in a way similar to the usual Ramachandran plot, which will be useful in analyzing the incorporation of these residues into proteins. When side chains are present in either type of residue, they are similar to conventional α-amino acids, although the orientation of the peptide groups is different. In such cases they acquire conformations similar to those found in peptide residues in the α-helix and β-sheet conformations, although other conformations are also possible. © 1998 John Wiley & Sons, Inc. Biopoly 45: 149–155, 1998  相似文献   

3.
α,β‐Dehydroamino acid esters occur in nature. To investigate their conformational properties, a systematic theoretical analysis was performed on the model molecules Ac‐ΔXaa‐OMe [ΔXaa = ΔAla, (E)‐ΔAbu, (Z)‐ΔAbu, ΔVal] at the B3LYP/6‐311+ + G(d,p) level in the gas phase as well as in chloroform and water solutions with the self‐consistent reaction field‐polarisable continuum model method. The Fourier transform IR spectra in CCl4 and CHCl3 have been analysed as well as the analogous solid state conformations drawn from The Cambridge Structural Database. The ΔAla residue has a considerable tendency to adopt planar conformations C5 (?, ψ ≈ ? 180°, 180°) and β2 (?, ψ ≈ ? 180°, 0°), regardless of the environment. The ΔVal residue prefers the conformation β2 (?, ψ ≈ ? 120°, 0°) in a low polar environment, but the conformations α (?, ψ ≈ ? 55°, 35°) and β (?, ψ ≈ ? 55°, 145°) when the polarity increases. The ΔAbu residues reveal intermediate properties, but their conformational dispositions depend on configuration of the side chain of residue: (E)‐ΔAbu is similar to ΔAla, whereas (Z)‐ΔAbu to ΔVal. Results indicate that the low‐energy conformation β2 is the characteristic feature of dehydroamino acid esters. The studied molecules constitute conformational patterns for dehydroamino acid esters with various side chain substituents in either or both Z and E positions. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
To obtain general rules of peptide design using α,β-dehydro-residues, a sequence with two consecutive ΔPhe-residues, Boc-L -Val-ΔPhe–ΔPhe- L -Ala-OCH3, was synthesized by azlactone method in solution phase. The peptide was crystallized from its solution in an acetone/water mixture (70:30) in space group P61 with a=b=14.912(3) Å, c= 25.548(5) Å, V=4912.0(6) Å3. The structure was determined by direct methods and refined by a full matrix least-squares procedure to an R value of 0.079 for 2891 observed [I?3σ(I)] reflections. The backbone torsion angles ?1=?54(1)°, ψ1= 129(1)°, ω1=?177(1)°, ?2 =57(1)°, ψ2=15(1)°, ω2 =?170(1)°, ?3=80(1)°, ψ3 =7(2)°, ω3=?177(1)°, ?4 =?108(1)° and ψT4=?34 (1)° suggest that the peptide adopts a folded conformation with two overlapping β-turns of types II and III′. These turns are stabilized by two intramolecular hydrogen bonds between the CO of the Boc group and the NH of ΔPhe3 and the CO of Val1 and the NH of Ala4. The torsion angles of ΔPhe2 and ΔPhe3 side chains are similar and indicate that the two ΔPhe residues are essentially planar. The folded molecules form head-to- tail intermolecular hydrogen bonds giving rise to continuous helical columns which run parallel to the c-axis. This structure established the formation of two β-turns of types II and III′ respectively for sequences containing two consecutive ΔPhe residues at (i+2) and (i+3) positions with a branched β-carbon residue at one end of the tetrapeptide.  相似文献   

5.
α,β-Dehydro amino acid residues are known to constrain the peptide backbone to the β-bend conformation. A pentapeptide containing only one α,β dehydrophenylalanine (ΔPhe) residue has been synthesized and crystallized, and its solid state conformation has been determined. The pentapeptide Boc-Leu-Phe-Ala-ΔPhe-Leu-OMe (C39H55N5O8, Mw = 721.9) was crystallized from aqueous methanol. Monoclinic space group was P21, a = 10.290(2)°, b = 17.149(2)°, c = 12.179(2) Å, β = 96.64(1)° with two molecules in the unit cell. The x-ray (Mo Kα, λ = 0.7107A) intensity data were collected using a CAD4 diffractometer. The crystal structure was determined by direct methods and refined using least-squares technique. R = 4.4% and Rw = 5.4% for 4403 reflections having |F0| ≥ 3σ(|F0|). All the peptide links are trans and the pentapeptide molecule assumes 310-helical conformation. The mean ?,ψ values, averaged over the first four residues, are ?64.4°, ?22.4° respectively. There are three 4 → 1 intramolecular hydrogen bonds, characteristic of 310,-helix. In the crystal, the peptide helices interact through two head-to-tail. N? H? O intermolecular hydrogen bonds. The peptide molecules related by 21, screw symmetry form a skewed assembly of helices. © 1995 John Wiley & Sons, Inc.  相似文献   

6.
We report here the synthesis and molecular structure in the solid state of fully protected tripeptides containing Cα,α-diphenylglycine (Dph), namely Z-Aib-Dph-Gly-OMe (Aib: Cα,α-dimethylgrycine) and Bz-Dph-Dph-Gly-OMe. The molecular conformation around the Dph residue, containing two bulky substituents, is fully extended, while the Aib residue, containing two smaller groups on the Cα atom, adopts the typical 310/α-helical conformation. Gly residues, without substituents on the Cα atom, show different conformational preferences. Each residue seems to behave, from a conformational point of view, independently from the presence of the other residues, and thus mixed local conformations (folded and extended) are present in the crystals. The nonconventional peptide synthesis, using the Ugi reaction, is also reported. © 1994 John Wiley & Sons, Inc.  相似文献   

7.
Over the last several years we have used spin labeling as a means for exploring the structure of helical peptides. Two nitroxide labels are engineered into a peptide sequence and distances are ranked with electron spin resonance (ESR). We have found that there is a significant amount of 310–helix in 16–residue model peptides containing only L –amino acids. This review covers several facets of the methodology including spin labeling strategy, interpretation of ESR spectra and the influence of molecular dynamics on the spectral line shapes. Also covered are recent findings of a length–dependent 3l0-helix → α-helix transition and the role of Arg+ in the stabilization of specific helix structures. © 1994 John Wiley & Sons, Inc.  相似文献   

8.
A single-crystal x-ray diffraction analysis of Boc-L -Ala-D -aIle-L -Ile-OMe has been carried out. The analysis has shown (a) that the tripeptide molecules have in part an α-extended conformation, the torsion angles of the L -Ala and D -aIle residues being φ1 = ?75.1° and ψ1 = ?25.8° and φ2 = 67.3° and ψ2 = 44.1°, respectively, and (b) that the molecules are organized in rippled planes where they occur in relative antiparallel orientation linked together side by side by H bonds. This molecular organization of the tripeptide corresponds closely to that of an antiparallel α-pleated sheet, and likely constitutes the first example of a structure of this kind for which a characterization at the atomic level has been achieved. A molecular dynamics study has shown that the molecular conformation of the tripeptide in the crystalline state is determined primarily by intermolecular interactions. © 1994 John Wiley & Sons, Inc.  相似文献   

9.
A complete series of terminally blocked, monodispersed homo-oligopeptides (to the pentamer level) from the sterically demanding, medium-ring alicyclic Cα,α-disubstituted glycine 1-aminocyclooctane-1-carb oxylic acid (Ac8c), and two Ala/Ac8c tripeptides, were synthesized by solution methods and fully characterized. The preferred conformation of all the oligopeptides was determined in deuterochloroform solution by IR absorption and 1H-NMR. The molecular structures of the amino acid derivative Z-Ac8c-OH, the dipeptide pBrBz- (Ac8c)2-OH and the tripeptide pBrBz-(Ac8c)3-OtBu were assessed in the crystal state by X-ray diffraction. Conformational energy computations were performed on the monopeptide Ac-Ac8c-NHMe. Taken together, the results obtained strongly support the view that the Ac8c residue is an effective β-turn and helix former. A comparison is also made with the conformational preferences of α-aminoisobutyric acid, the prototype of Cα, α-disubstituted glycines, and of the other members of the family of 1-aminocycloalkane-1-carboxylic acids (Acnc, with n=3, 5–7) investigated so far. The implications for the use of the Ac8c residue in peptide conformational design are considered.  相似文献   

10.
The solid state conformations of cyclo[Gly–Proψ[CH2S]Gly–D –Phe–Pro] and cyclo[Gly–Proψ[CH2–(S)–SO]Gly–D –Phe–Pro] have been characterized by X-ray diffraction analysis. Crystals of the sulfide trihydrate are orthorhombic, P212121, with a = 10.156(3) Å, b = 11.704(3) Å, c = 21.913(4) Å, and Z = 4. Crystals of the sulfoxide are monoclinic, P21, with a = 10.662(1) Å, b = 8.552(3) Å, c = 12.947(2) Å, β = 94.28(2), and Z = 2. Unlike their all-amide parent, which adopts an all-trans backbone conformation and a type II β-turn encompassing Gly-Pro-Gly-D -Phe, both of these peptides contain a cis Gly1-Pro2 bond and form a novel turn structure, i.e., a type II′ β-turn consisting of Gly–D –Phe–Pro–Gly. The turn structure in each of these peptides is stabilized by an intramolecular H bond between the carbonyl oxygen of Gly1 and the amide proton of D -Phe4. In the cyclic sulfoxide, the sulfinyl group is not involved in H bonding despite its strong potential as a hydrogen-bond acceptor. The crystal structure made it possible to establish the absolute configuration of the sulfinyl group in this peptide. The two crystal structures also helped identify a type II′ β-turn in the DMSO-d6 solution conformers of these peptides. © 1993 John Wiley & Sons, Inc.  相似文献   

11.
A single chiral cyclic α,α‐disubstituted amino acid, (3S,4S)‐1‐amino‐(3,4‐dimethoxy)cyclopentanecarboxylic acid [(S,S)‐Ac5cdOM], was placed at the N‐terminal or C‐terminal positions of achiral α‐aminoisobutyric acid (Aib) peptide segments. The IR and 1H NMR spectra indicated that the dominant conformations of two peptides Cbz‐[(S,S)‐Ac5cdOM]‐(Aib)4‐OEt ( 1) and Cbz‐(Aib)4‐[(S,S)‐Ac5cdOM]‐OMe (2) in solution were helical structures. X‐ray crystallographic analysis of 1 and 2 revealed that a left‐handed (M) 310‐helical structure was present in 1 and that a right‐handed (P) 310‐helical structure was present in 2 in their crystalline states. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
The side chain conformations of α-helical poly(L -glutamic acid) esters $ \rlap{--}[NHCH(CH_2 CH_2 COOR)CO\rlap{--}]_x $, carrying a homologous series of ester residues such as R = ? (CH2)n? with n = 1–3, have been studied in the lyotropic liquid crystalline state (chloroform 20 v/v%) by the deuterium nmr method. In order to study the surface chirality of the molecule, the phenyl groups situated at the terminal of the side chain have been deuterated. From the observed deuterium quadrupolar splittings, the average inclination θp of the para-axis of the phenyl group with respect to the α-helical backbone was elucidated. A distinct odd–even oscillation in the quantity such as 〈 cos2 θp〉 was observed with the number of methylene units n. A rotational isomeric state analysis has indicated that the observed orientational correlation arises from the interdependence of the neighboring bond rotation along the side chain. Preference of the “extended” conformations is also enhanced by the mutual conformational exclusion of neighboring side chains.  相似文献   

13.
Conformations of the α-l -Rhap(1-2)-β-d -Glc1-OMe and β-d -Galp(1-3)-β-d -Glc1-OMe disaccharides and the branched title trisaccharide were examined in DMSO-d6 solution by 1H-nmr. The distance mapping procedure was based on rotating frame nuclear Overhauser effect (NOE) constraints involving C- and O-linked protons, and hydrogen-bond constraints manifested by the splitting of the OH nmr signals for partially deuteriated samples. An “isotopomer-selected NOE” method for the unequivocal identification of mutually hydrogen-bonded hydroxyl groups was suggested. The length of hydrogen bonds thus detected is considered the only one motionally nonaveraged nmr-derived constraint. Molecular mechanics and molecular dynamics methods were used to model the conformational properties of the studied oligosaccharides. Complex conformational search, relying on a regular Φ,Ψ-grid based scanning of the conformational space of the selected glycosidic linkage, combined with simultaneous modeling of different allowed orientations of the pendant groups and the third, neighboring sugar residue, has been carried out. Energy minimizations were performed for each member of the Φ,Ψ grid generated set of conformations. Conformational clustering has been done to group the minimized conformations into families with similar values of glycosidic torsion angles. Several stable syn and anti conformations were found for the 1→2 and 1→3 bonds in the studied disaccharides. Vicinal glycosylation affected strongly the occupancy of conformational states in both branches of the title trisaccharide. The preferred conformational family of the trisaccharide (with average Φ,Ψ values of 38°, 17° for the 1→2 and 48°, 1° for the 1→3 bond, respectively) was shown by nmr to be stabilized by intramolecular hydrogen bonding between the nonbonded Rha and Gal residues. © 1998 John Wiley & Sons, Inc. Biopoly 46: 417–432, 1998  相似文献   

14.
Some theoretical studies have predicted that the conformational freedom of the α-aminoisobutyric acid (H-Aib-OH) residue is restricted to the α-helical region of the Ramachandran map. In order to obtain conformational experimental data, two model peptide derivatives, MeCO-Aib-NHMe 1 and ButCO-LPro-Aib-NHMe 2 , have been investigated. The Aib dipeptide 1 crystallizes in the monoclinic system (a = 12.71 Å, b = 10.19 Å, c = 7.29 Å, β = 110.02°, Cc space group) and its crystal structure was elucidated by x-ray diffraction analysis. The azimuthal angles depicting the molecular conformation (? = ?55.5°, ψ = ?39.3°) fall in the α-helical region of the Ramachandran map and molecules are hydrogen-bonded in a three-dimensional network. In CCl4 solution, ir spectroscopy provides evidence for the occurrence of the so-called 5 and C7 conformers stabilized by the intramolecular ii and i + 2 → i hydrogen bonds, respectively. The tripeptide 2 was studied in various solvents [CCl4, CD2Cl2, CDCl3, (CD3)2SO, and D2O] by ir and pmr spectroscopies. It was shown to accommodate predominantly the βII folded state stabilized by the i + 3 → i hydrogen bond. All these experimental findings indicate that the Aib residue displays the same conformational behavior as the other natural chiral amino acid residues.  相似文献   

15.
A series of terminally blocked peptides (to the pentamer level) from l ‐Ala and the cyclic Cα,α‐disubstituted Gly residue Afc and one Gly/Afc dipeptide have been synthesized by solution method and fully characterized. The molecular structure of the amino acid derivative Boc‐Afc‐OMe and the dipeptide Boc‐Afc‐Gly‐OMe were determined in the crystal state by X‐ray diffraction. In addition, the preferred conformation of all of the model peptides was assessed in deuterochloroform solution by FT‐IR absorption and 1H‐NMR. The experimental data favour the conclusion that the Afc residue tends to adopt either the fully‐extended (C5) or a folded/helical structure. In particular, the former conformation is highly populated in solution and is also that found in the crystal state in the two compounds investigated. A comparison with the structural propensities of the strictly related Cα,α‐disubstituted Gly residues Ac5c and Dϕg is made and the implications for the use of the Afc residue in conformationally constrained analogues of bioactive peptides are briefly examined. A spectroscopic (UV absorption, fluorescence, CD) characterization of this novel aromatic Cα,α‐disubstituted Gly residue is also reported. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
The preferred conformations of N-acetyl-N′-methyl amides of some dialkylglycines have been determined by empirical conformational-energy calculations; minimum-energy conformations were located by minimizing the energy with respect to all the dihedral angles of the molecules. The conformational space of these compounds is sterically restricted, and low-energy conformations are found only in the regions of fully extended and helical structures. Increasing the bulkiness of the substituents on the Cα, the fully extended conformation becomes gradually more stable than the helical structure preferred in the cases of dimethylglycine. This trend is, however, strongly dependent on the bond angles between the substituents on the Cα atom: In particular, helical structures are favored by standard values (111°) of the N-Cα-C′ angle, while fully extended conformations are favored by smaller values of the same angle, as experimentally observed, for instance, in the case of α,α-di-n-propylglycine.  相似文献   

17.
The conformational properties of α,α-dialkylated amino acid residues possessing acyclic (diethylglycine, Deg: di-n-propylglycine, Dpg; di-n-butylglycine, Dbg) and cyclic (1-amino-cycloalkane-1-carboxylic acid, Acnc) side chains have been compared in solution. The five peptides studied by nmr and CD spectroscopy are Boc-Ala-Xxx-Ala-OMe, where Xxx = Deg(I). Dpg (II), Dbg (III), Ac6c (IV), and Ac7c (V). Delineation of solvent-shielded NH groups have been achieved by solvent and temperature dependence of NH chemical shifts in CDCl3 and (CD3)2SO and by paramagnetic radical induced line broadening in pepiide III. In the Dxg peptides the order of solvent exposure of NH groups is Ala(1) > Ala(3) > Dxg(2), whereas in the Acnc peptides the order of solvent exposure of NH groups is Ala(1) > Acnc(2) > Ala(3). The nmr results suggest that Acnc peptides adopt folded β-turn conformations with Ala(1) and Acnc(2) occupying i + 1 and i + 2 positions. In contrast, the Dxg peptides favor extended C5 conformations. The conformational differences in the two series are clearly borne out in CD studies. The solution conformations of peptides I-III are distinctly different from the β-turn structure observed in crystals. Low temperature nmr spectra recorded immediately after dissolution of crystals of peptide II provide evidence for a structural transition. Introduction of an additional hydrogen-bonding function in Boc-Ala-Dpg-Ala-NHMe (VI) results in a stabilization of a consecutive β-turn or incipient 310-helix in solution. © 1995 John Wiley & Sons, Inc.  相似文献   

18.
Today there are several different experimental scales for the intrinsic α-helix as well as β-strand, propensities of the 20 amino acids obtained from the thermodynamic analysis of various model systems. These scales do not compare well with those extracted from statistical analysis of three-dimensional structure databases. Possible explanations for this could be the limited size of the databases used, the definitions of intrinsic propensities, or the theoretical approach. Here we report a statistical determination of α-helix and β-strand propensities derived from the analysis of a database of 279 three-dimensional structures. Contrary to what has been generally done, we have considered a particular residue as in α-helix or β-strand conformation by looking only at its dihedral angles (?–ψ matrices). Neither the identity nor the conformation of the surrounding residues in the amino acid sequence has been taken into consideration. Pseudoenergy empirical scales have been calculated from the statistical propensities. These scales agree very well with the experimental ones in relative and absolute terms. Moreover, its correlation with the average of the experimental scales for α-helix or β-strand is as good as the correlations of the individual experimental scales with the average. These results show that by using a large enough database and a proper definition for the secondary structure propensities, it is possible to obtain a scale as good as any of experimental origin. Interestingly the ?–ψ analysis of the Ramachandran plot suggests that the amino acids could have different β-strand propensities in different subregions of the β-strand area. © 1994 Wiley-Liss, Inc.  相似文献   

19.
Elucidating protein function from its structure is central to the understanding of cellular mechanisms. This involves deciphering the dependence of local structural motifs on sequence. These structural motifs may be stabilized by direct or water‐mediated hydrogen bonding among the constituent residues. π‐Turns, defined by interactions between (i) and (i + 5) positions, are large enough to contain a central space that can embed a water molecule (or a protein moiety) to form a stable structure. This work is an analysis of such embedded π‐turns using a nonredundant dataset of protein structures. A total of 2965 embedded π‐turns have been identified, as also 281 embedded Schellman motif, a type of π‐turn which occurs at the C‐termini of α‐helices. Embedded π‐turns and Schellman motifs have been classified on the basis of the protein atoms of the terminal turn residues that are linked by the embedded moiety, conformation, residue composition, and compared with the turns that have terminal residues connected by direct hydrogen bonds. Geometrically, the turns have been fitted to a circle and the position of the linker relative to its center analyzed. The hydroxyl group of Ser and Thr, located at (i + 3) position, is the most prominent linker for the side‐chain mediated π‐turns. Consideration of residue conservation among homologous sequences indicates the terminal and the linker positions to be the most conserved. The embedded π‐turn as a binding site (for the linker) is discussed in the context of “nest,” a concave depression that is formed in protein structures with adjacent residues having enantiomeric main‐chain conformations. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 441–453, 2014.  相似文献   

20.
Proline-induced constraints in alpha-helices   总被引:9,自引:0,他引:9  
L Piela  G Némethy  H A Scheraga 《Biopolymers》1987,26(9):1587-1600
The disrupting effect of a prolyl residue on an α-helix has been analyzed by means of conformational energy computations. In the preferred, nearly α-helical conformations of Ac-Ala4-Pro-NHMe and of Ac-Ala7-Pro-Ala7-NHMe, only the residue preceding Pro is not α-helical, while all other residues can occur in the α-helical A conformation; i.e., it is sufficient to introduce a conformational change of only one residue in order to accommodate proline in a distorted α-helix. Other low-energy conformations exist in which the conformational state of three residues preceding proline is altered considerably; on the other hand, another conformation in which these three residues retain the near-α-helical A-conformational state (with up to 26° changes of their dihedral angles ? and ψ, and a 48° change in one ω from those of the ideal α-helix) has a considerably higher energy. These conclusions are not altered by the substitution of other residues in the place of the Ala preceding Pro. The conformations of the peptide chain next to prolyl residues in or near an α-helix have been analyzed in 58 proteins of known structure, based on published atomic coordinates. Of 331 α-helices, 61 have a Pro at or next to their N-terminus, 21 have a Pro next to their C-terminus, and 30 contain a Pro inside the helix. Of the latter, 16 correspond to a break in the helix, 9 are located inside distorted first turns of the helix, and 5 are parts of irregular helices. Thus, the reported occurrence of prolyl residues next to or inside observed α-helices in proteins is consistent with the computed steric and energetic requirements of prolyl peptides.  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