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.
DL -Arginine DL -glutamate monohydrate and DL -arginine DL -aspartate, the first DL -DL amino acid–amino acid complexes to be prepared and x-ray analyzed, crystallize in the space group P1 with a = 5.139(2), b = 10.620(1), c = 14.473(2) Å, α = 101.34(1)°, β = 94.08(2)°, γ = 91.38(2)° and a = 5.402(3), b = 9.933(3), c = 13.881(2) Å, α = 99.24(2)°, β = 99.73(3)°, γ = 97.28(3)°, respectively. The structures were solved using counter data and refined to R values of 0.050 and 0.077 for 1827 and 1739 observed reflections, respectively. The basic element of aggregation in both structures is an infinite chain made up of pairs of molecules. Each pair, consisting of a L - and a D -isomer, is stabilized by two centrosymmetrically or nearly centrosymmetrically related hydrogen bonds involving the α-amino and the α-carboxylate groups. Adjacent pairs in the chain are then connected by specific guanidyl–carboxylate interactions. The infinite chains are interconnected through hydrogen bonds to form molecular sheets. The sheets are then stacked along the shortest cell translation. The interactions between sheets involve two head-to-tail sequences in the glutamate complex and one such sequence in the aspartate complex. However, unlike in the corresponding LL and DL complexes, head-to-tail sequences are not the central feature of molecular aggregation in the DL -DL complexes. Indeed, fundamental differences exist among the aggregation patterns in the LL , the LD , and the DL -DL complexes.  相似文献   

2.
The crystal structures of the complexes of oxalic acid with glycyl-L-histidine and L-histidyl-L-alanine were determined. The three crystallographically independent peptide molecules in the complexes have closed conformations. The terminal carboxyl group of the dipeptide and the oxalate ion in the Gly-His complex exhibit unusual ionization states and are connected by a symmetric O- - -O hydrogen bond. The peptide aggregation in the complex is almost identical to that in the corresponding semisuccinate complex and is similar to one of the predicted aggregation patterns for Ala-Ala, demonstrating that dipeptide aggregation is controlled primarily by main-chain interactions and is substantially unaffected by disturbing influences such as those arising from polar side chains, ions and water molecules. The peptide molecules in the highly pseudosymmetric crystals of the His-Ala complex, however, exhibit a hitherto unobserved aggregation pattern. Thus, in spite of the repeated occurrence of a few patterns, the possibility of the existence of new patterns needs to be taken into account.  相似文献   

3.
L-Lysine D-glutamate crystallizes in the monoclinic space group P2(1) with a = 4.902, b = 30.719, c = 9.679 A, beta = 90 degrees and Z = 4. The crystals of L-lysine D-aspartate monohydrate belong to the orthorhombic space group P2(1)2(1)2(1) with a = 5.458, b = 7.152, c = 36.022 A and Z = 4. The structures were solved by the direct methods and refined to R values of 0.125 and 0.040 respectively for 1412 and 1503 observed reflections. The glutamate complex is highly pseudosymmetric. The lysine molecules in it assume a conformation with the side chain staggered between the alpha-amino and the alpha-carboxylate groups. The interactions of the side chain amino groups of lysine in the two complexes are such that they form infinite sequences containing alternating amino and carboxylate groups. The molecular aggregation in the glutamate complex is very similar to that observed in L-arginine D-aspartate and L-arginine D-glutamate trihydrate, with the formation of double layers consisting of both types of molecules. In contrast to the situation in the other three LD complexes, the unlike molecules in L-lysine D-aspartate monohydrate aggregate into alternating layers as in the case of most LL complexes. The arrangement of molecules in the lysine layer is nearly the same as in L-lysine L-aspartate, with head-to-tail sequences as the central feature. The arrangement of aspartate ions in the layers containing them is, however, somewhat unusual. Thus the comparison between the LL and the LD complexes analyzed so far indicates that the reversal of chirality of one of the components in a complex leads to profound changes in molecular aggregation, but these changes could be of more than one type.  相似文献   

4.
A new form of L-histidine L-aspartate monohydrate crystallizes in space group P22 witha = 5.131(1),b = 6.881(1),c= 18.277(2) Å,β= 97.26(1)° and Z = 2. The structure has been solved by the direct methods and refined to anR value of 0.044 for 1377 observed reflections. Both the amino acid molecules in the complex assume the energetically least favourable allowed conformation with the side chains staggered between the α-amino and α-scarboxylate groups. This results in characteristic distortions in some bond angles. The unlike molecules aggregate into alternating double layers with water molecules sandwiched between the two layers in the aspartate double layer. The molecules in each layer are arranged in a head-to-tail fashion. The aggregation pattern in the complex is fundamentally similar to that in other binary complexes involving commonly occurring L amino acids, although the molecules aggregate into single layers in them. The distribution of crystallographic (and local) symmetry elements in the old form of the complex is very different from that in the new form. So is the conformation of half the histidine molecules. Yet, the basic features of molecular aggregation, particularly the nature and the orientation of head-to-tail sequences, remain the same in both the forms. This supports the thesis that the characteristic aggregation patterns observed in crystal structures represent an intrinsic property of amino acid aggregation.  相似文献   

5.
L-Lysine acetate crystallises in the monoclinic space group P21 with a = 5.411 (1), b = 7.562(1), c = 12.635(2) A and beta = 91.7(1) degrees. The crystal structure was solved by direct methods and refined to an R value of 0.049 using the full matrix least squares method. The conformation and the aggregation of lysine molecules in the structure are similar to those found in the crystal structure of L-lysine L-aspartate. A conspicuous similarity between the crystal structures of L-arginine acetate and L-lysine acetate is that in both cases the strongly basic side chain, although having the largest pK value, interacts with the weakly acidic acetate group leaving the alpha-amino and the alpha-carboxylate groups to take part in head-to-tail sequences. These structures thus indicate that electrostatic effects are strongly modulated by other factors so as to give rise to head-to-tail sequences which have earlier been shown to be an almost universal feature of amino acid aggregation in the solid state.  相似文献   

6.
The new form of L-arginine D-glutamate is monoclinic, P21, witha = 9.941(1),b = 4.668(2),c = 17.307(1) Å,β = 95.27(1)°, and Z = 2. In terms of composition, the new form differs from the old form in that the former is a monohydrate whereas the latter is a trihydrate. The structure has been solved by the direct methods and refined to R = 0.085 for 1012 observed reflections. The conformation of the arginine molecule is the same in both the forms whereas that of the glutamate ion is different. The change in the conformation of the glutamate ion is such that it facilitates extensive pseudosymmetry in the crystals. The molecules arrange themselves in double-layers stabilised by head-to-tail sequences involving main chains, in both the forms. However, considerable differences exist between the two forms in the interface, consisting of side chains and water molecules, between double-layers. A comparative study of the relationship between the crystal structures of L and DL amino acids on the one hand and that between the structures of LL and LD amino acid-amino acid complexes on the other, provides interesting insights into amino acid aggregation and the effect of chirality on it. The crystal structures of most hydrophobic amino acids are made up of double-layers and those of most hydrophilic amino acids contain single layers, irrespective of the chiralities of the amino acids involved. In most cases, the molecules tend to appropriately rearrange themselves to preserve the broad features of aggregation patterns when the chirality of half the molecules is reversed as in the structures of DL amino acids. The basic elements of aggregation in the LL and the LD complexes, are similar to those found in the crystals of L and DL amino acids. However, the differences between the LL and the LD complexes in the distribution of these elements are more pronounced than those between the distributions in the structures of L and DL amino acids.  相似文献   

7.
Crystals of DL-arginine hemisuccinate dihydrate (I)(monoclinic; P2(1)/c; a = 5.292, b = 16.296. c = 15.203 A; beta = 92.89 degrees; Z = 4) and L-arginine hemisuccinate hemisuccinic acid monohydrate (II) (triclinic; P1; a = 5.099; b = 10.222, c = 14.626 A; alpha = 77.31, beta = 89.46, gamma = 78.42 degrees; Z = 2) were grown under identical conditions from aqueous solutions of the components in molar proportions. The structures were solved by direct methods and refined to R = 0.068 for 2585 observed reflections in the case of (I) and R = 0.036 for 2154 observed reflections in the case of (II). Two of the three crystallographically independent arginine molecules in the complexes have conformations different from those observed so far in the crystal structures containing arginine. The succinic acid molecules and the succinate ions in the structures are centrosymmetric and planar. The crystal structure of (II) is highly pseudosymmetric. Arginine-succinate interactions in both the complexes involve specific guanidyl-carboxylate interactions. The basic elements of aggregation in both the structures are ribbons made up of alternating arginine dimers and succinate ions. However, the ribbons pack in different ways in the two structures. (II) presents an interesting case in which two ionisation states of the same molecule coexist in a crystal. The two complexes provide a good example of the effect of change in chirality on stoichiometry, conformation, aggregation, and ionisation state in the solid state.  相似文献   

8.
Diffusion of acetonitrile into an aqueous solution of DL -histidine and succinic acid in 1:3 molar proportions results in the crystals of DL -histidine hemisuccinate dihydrate [triclinic, P1 , a = 7.654(1), b = 8.723(1), c = 9.260(1) Å, α = 77.23(1), β = 72.37(1) and γ = 82.32 (1)°]. The replacement of DL -histidine by L -histidine in the crystallization experiment under identical conditions leads to crystals of L -histidine semisuccinate trihydrate [orthorhombic, P212121, a = 7.030 (1), b = 8.773 (1), and c = 24.332 (3) Å]. The structures were solved using counter data and refined to R values of 0.056 and 0.054 for 2356 and 1778 observed reflections, respectively. Histidine molecules in both the complexes exist in open conformation I. Succinate and semisuccinate ions in them are planar, and exactly or nearly centrosymmetric. In the DL -histidine complex, the amino acid molecules form double ribbons and the succinate ions occupy voids left behind when the double ribbons aggregate, as in inclusion compounds. In the L -histidine complex, the amino acid molecules form columns; so do the semisuccinate ions and water molecules. The two columns interdigitate to form the complex crystal. There are similarities between the molecular aggregation in the complexes and that in the crystals of L - and DL -histidine. However, the presence of succinic acid has the effect of disrupting, partially or totally, head-to-tail sequences involving amino acid molecules. © 1993 John Wiley & Sons, Inc.  相似文献   

9.
The site on influenza virus N9 neuraminidase recognized by NC41 monoclonal antibody comprises 19 amino acid residues that are in direct contact with 17 residues on the antibody. Single sequence changes in some of the neuraminidase residues in the site markedly reduce antibody binding. However, two mutants have been found within the site, Ile368 to Arg and Asn329 to Asp selected by antibodies other than NC41, and these mutants bind NC41 antibody with only slightly reduced affinity. The three-dimensional structures of the two mutant N9-NC41 antibody complexes as derived from the wild-type complex are presented. Both structures show that some amino acid substitutions can be accommodated within an antigen-antibody interface by local structural rearrangements around the mutation site. In the Ile368 to Arg mutant complex, the side-chain of Arg368 is shifted by 2.9 A from its position in the uncomplexed mutant and a shift of 1.3 A in the position of the light chain residue HisL55 with respect to the wild-type complex is also observed. In the other mutant, the side-chain of Asp329 appears rotated by 150 degrees around C alpha-C beta with respect to the uncomplexed mutant, so that the carboxylate group is moved to the periphery of the antigen-antibody interface. The results provide a basis for understanding some of the potential structural effects of somatic hypermutation on antigen-antibody binding in those cases where the mutation in the antibody occurs at antigen-contacting residues, and demonstrate again the importance of structural context in evaluating the effect of amino acid substitutions on protein structure and function.  相似文献   

10.
H Kim  W N Lipscomb 《Biochemistry》1990,29(23):5546-5555
O-[[(1R)-[[N-(Phenylmethoxycarbonyl)-L-alanyl]amino]ethyl] hydroxyphosphinyl]-L-3-phenyllacetate [ZAAP(O)F], an analogue of (benzyloxycarbonyl)-Ala-Ala-Phe or (benzyloxycarbonyl)-Ala-Ala-phenyllactate, binds to carboxypeptidase A with great affinity (Ki = 3 pM). Similar phosphonates have been shown to be transition-state analogues of the CPA-catalyzed hydrolysis [Hanson, J. E., Kaplan, A. P., & Bartlett, P. A. (1989) Biochemistry 28, 6294-6305]. In the present study, the structure of the complex of this phosphonate with carboxypeptidase A has been determined by X-ray crystallography to a resolution of 2.0 A. The complex crystallizes in the space group P2(1)2(1)2(1) with cell dimensions a = 61.9 A, b = 67.2 A, and c = 76.2 A. The structure of the complex was solved by molecular replacement. Refinement of the structure against 20,776 unique reflections between 10.0 and 2.0 A yields a crystallographic residual of 0.193, including 140 water molecules. The two phosphinyl oxygens of the inhibitor bind to the active-site zinc at 2.2 A on the electrophilic (Arg-127) side and 3.1 A on the nucleophilic (Glu-270) side. Various features of the binding mode of this phosphonate inhibitor are consistent with the hypothesis that carboxypeptidase A catalyzed hydrolysis proceeds through a general-base mechanism in which the carbonyl carbon of the substrate is attacked by Zn-hydroxyl (or Zn-water). An unexpected feature of the bound inhibitor, the cis carbamoyl ester bond at the benzyloxycarbonyl linkage to alanine, allows the benzyloxycarbonyl phenyl ring of the inhibitor to interact favorably with Tyr-198. This complex structure is compared with previous structures of carboxypeptidase A, including the complexes with the potato inhibitor, a hydrated keto methylene substrate analogue, and a phosphonamidate inhibitor. Comparisons are also made with the complexes of thermolysin with some phosphonamidate inhibitors.  相似文献   

11.
We have obtained deuterium (2H) nuclear magnetic resonance (NMR) spectra and spin-lattice relaxation times (T1) of L-[epsilon-2H3]methionine, L-[epsilon-2H3]methionine in a D,L lattice, and [S-methyl-2H3]methionine in the crystalline solid state, as a function of temperature, in addition to obtaining 2H T1 and line-width results as a function of temperature on [epsilon-2H3]methionine-labeled sperm whale (Physeter catodon) myoglobins by using the method of magnetic ordering [Rothgeb, T. M., & Oldfield, E. (1981) J. Biol. Chem. 256, 1432-1446]. The results indicate that in the L-amino acid, methyl rotation having an activation energy (delta E) of 8.3 +/- 1 kJ dominates T1 at low temperatures (less than or equal to 10 degrees C), while at higher temperatures an additional large-amplitude side-chain motion occurs which causes changes in the 2H NMR line shape and T1. This motion is inhibited in the D,L lattice, indicating that lattice effects may have a strong effect on the mobility of anhydrous amino acids in the solid state. Further substitution at S delta to form the sulfonium salt [S-methyl-2H3]-methionine causes a large increase in delta E, to 15.9 +/- 2 kJ, a value comparable to the 14-16 kJ found in valine and leucine, which contain the structurally similar isopropyl moiety. These results suggest that the very low barriers to methyl rotation in the methionine side chain are due to long C-S bond lengths and the presence of only two substituents on sulfur, while the anomalous high-temperature behavior is due to a lattice-packing effect. 2H T1 results with methionine-labeled myoglobin are complex, reflecting the presence of fast large-amplitude side-chain motions, in addition to rapid methyl rotation. Our data indicate that Met-55 and Met-131 are motionally inequivalent in crystalline cyanoferrimyoglobin, in contrast to solution NMR results. We have also recorded 13C cross-polarization "magic-angle" sample-spinning NMR spectra of [epsilon-13C]methionine-labeled crystalline cyanoferrimyoglobin (at 37.7 MHz, corresponding to a magnetic field strength of 3.52 T) and of the same protein in aqueous solution. Cross-polarization transfer rates and proton rotating-frame relaxation time results again indicate that Met-55 and Met-131 are motionally inequivalent in the solid state, and the TCH data indicate that Met-55 is more solidlike. However, we find that 13C chemical shifts in solution and those in the crystalline solid state are in very close agreement, suggesting that the average solution and crystal conformations are the same, in the area of Met-55 and Met-131.  相似文献   

12.
A rhymed presentation of the amino acids, consisting of one line per amino acid, gives the structures of the amino acids, their nomenclature, abbreviations, pKs, basic chemistry, dietary requirements and the characteristics of the peptide bond.  相似文献   

13.
The heptacosapeptide amide corresponding to the entire amino acid sequence of chicken secretin was synthesized by assembling four peptide fragments followed by deprotection with thioanisole-mediated trifluoromethanesulfonic acid in TFA. The deprotected peptide was purified by gel-filtration on Sephadex G-25, followed by ion-exchange chromatographies on CM-cellulose and DEAE-Sepharose. Synthetic peptide stimulated the flow of pancreatic alkaline juice and decreased systemic blood pressure in rats.  相似文献   

14.
N V Kumar  G Govil 《Biopolymers》1984,23(10):1979-1993
Coulombic interactions between the side chains of charged amino acids (Arg+, Lys+, and His+) and negatively charged phosphate groups of nucleic acid fragments have been studied theoretically. Diribose monophosphate and dideoxyribose monophosphate are chosen as model systems for single-stranded RNA and DNA, respectively. The interaction energies have been calculated by second-order perturbation theory using simplified formulas for individual terms. The interaction energy in this formalism is a sum of electrostatic, polarization, dispersion, and repulsive energies. Our results show that about 90% of the total interaction energy is contributed by the electrostatic term alone. Contribution from the repulsive term exceeds that from the dispersion term. Calculated interaction energies suggest that Lys+ and His+ form more stable complexes with RNA than with single-stranded DNA. On the other hand, Arg+ has a higher affinity for DNA than for RNA. The affinity of nucleic acids for the three amino acids is in the order Lys+ > His+ > Arg+. Further, the basic amino acid residues form more stable complexes with A-DNA than with B-DNA. The role of the Coulombic interactions in the specific recognition of nucleic acids by proteins is discussed.  相似文献   

15.
16.
N V Kumar  G Govil 《Biopolymers》1984,23(10):2009-2024
Stacking of aromatic amino acids tryptophan (Trp), tyrosine (Tyr), phenylalanine (Phe), and histidine (His) with bases and base pairs of nucleic acids has been studied. Stacking energies of the amino acid–base (or base pair) complexes have been calculated by second-order perturbation theory. Our results show that, in general, the predominant contribution to the total stacking energy comes from the dispersion terms. In these cases, repulsion energy is greater than the sum of electrostatic and polarization energies. In contrast to this, interaction of histidine with the bases and base pairs is largely Coulombic in nature. The complexes of guanine with aromatic amino acids are more stable than the corresponding complexes of adenine. Among pyrimidines, cytosine forms the most stable complexes with the aromatic amino acids. The G · C base pair has the highest affinity with aromatic amino acids among various sets of base pairs. Optimized geometries of the stacked complexes show that the aromatic moieties overlap only partially. The heteroatom of one residue generally overlaps with the other aromatic moiety. There is a considerable degree of configurational freedom in the stacked geometries. The role of stacking in specific recognition of base sequences by proteins is discussed.  相似文献   

17.
18.
The gene (coxII) encoding subunit II of Rhodobacter sphaeroides cytochrome c oxidase (cytochrome aa3) has been isolated by screening a genomic DNA library in phage lambda with a probe derived from coxII of Paracoccus denitrificans. A 2-kb fragment containing coxII DNA was subcloned into the phage M13mp18 and the sequence determined. The 2-kb insert contains the entire coding region for coxII gene, including the ATG start codon and a TGA stop codon. The deduced amino acid (aa) sequence of subunit II of R. sphaeroides shows regions of substantial homology to the corresponding subunit of the bovine mitochondrial oxidase (63% overall) and P. denitrificans oxidase (68% overall). The postulated redox-active copper ion (CuA) binding site involving two Cys and two His residues (as well as an alternative Met residue) is conserved among these species, along with four invariant acidic aa residues (two Asp and two Glu) that may be involved in interactions with cytochrome c, and a region of aromatic residues (Tyr-Gln-Trp-Tyr-Trp-Gly-Tyr-Glu-Tyr) which is postulated to play a role in electron transfer. Hydropathy profile analysis suggests that while the bovine COXII secondary structure contains two transmembrane helices, the R. sphaeroides subunit II has a third such helix that may function as part of a signal sequence, as suggested for P. denitrificans.  相似文献   

19.
20.
The interaction of Cu(II) with di- and tripeptides each containing phenylalanine, tryptophan or histidine in the amino acid chain has been investigated by means of electron spin resonance (ESR) and optical absorption spectroscopy. Cu(II) complexes of dipeptides and tripeptides exhibit different magnetic and optical parameters. Dipeptide complexes have larger gparallel-values and smaller A parallel values than tripeptide complexes. When compared to dipeptide complexes, the d-d band of the central metal ion is blue shifted for tripeptide complexes. There are no significant difference in the behavior of Cu(II) peptide complexes containing phenylalanine or tryptophan. Complexes of histidine containing peptides, however, show modified spectra caused by the participation of the imidazole nitrogen in the coordination to Cu(II). The imidazole nitrogen seems to coordinate in-plane with other coordinating atoms or in an axial position depending on the kind of peptide.  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