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 31 毫秒
1.
G R Hedwig 《Biopolymers》1992,32(5):537-540
The partial molar volumes of some amino acid side-chains were determined in two recent studies [Makhatadze, G.I., et al. (1990) Biopolymers 30, 1001, and Reading, J.F. & Hedwig, G.R. (1990) J. Chem. Soc. Faraday Trans. 86, 3117] using partial molar volume data, V02, in aqueous solution at 25 degrees C for some peptides of sequence Gly-X-Gly, where X is an amino acid. These side-chain partial molar volumes are critically compared with those obtained using V02 data for amino acids. It is concluded that side-chain partial molar volumes calculated using V02 data for the tripeptides are better estimates of side-chain partial molar volumes in proteins than are those determined using V02 data for amino acids.  相似文献   

2.
The partial molar volumes of tripeptides of sequence glycyl-X-glycine, where X is one of the amino acids alanine, leucine, threonine, glutamine, phenylalanine, histidine, cysteine, proline, glutamic acid, and arginine, have been determined in aqueous solution over the temperature range 10-90 degrees C using differential scanning densitometry . These data, together with those reported previously, have been used to derive the partial molar volumes of the side-chains of all 20 amino acids. The side-chain volumes are critically compared with literature values derived using partial molar volumes for alternative model compounds. The new amino acid side-chain volumes, along with that for the backbone glycyl group, were used to calculate the partial specific volumes of several proteins in aqueous solution. The results obtained are compared with those observed experimentally. The new side-chain volumes have also been used to re-determine residue volume changes upon protein folding.  相似文献   

3.
Several amino acid side-chain hydropathy scales have been devised on the basis of solubility and water/organic solvent partitioning data obtained with free amino acids or side-chain analogs. In nearly all cases, these scales are based upon the structure-additivity assumption; it has been assumed that the transfer free energies of the amino acid side-chains are the same in these model compounds as they are in a polypeptide. This assumption is probably wrong. In the present study, deviations from additivity for amino acid side-chains are demonstrated by comparing a theoretically derived scale, which N-acetylamino acid amides. The results show that the flanking peptide bonds dramatically reduce the hydrophilicity of the polar side-chains, with deviations up to several kilocalories (1 kcal = 4.184 kJ) for the charged side-chains at pH 7.0. Further calculation shows that these deviations are due to reductions of 40 to 85% in the unfavorable transfer free energy of the polar functional groups. In addition, proximity of the neighboring amide bonds in the parent molecule (N-acetylglycine amide) decreases the hydrophilicity of the -CONH-backbone unit by 36%. This decrease is expected to be twice as large for -CONH- units in the interior of a polypeptide backbone. The significance of these observations is: (1) valid hydropathy scales can be obtained only with model peptides; (2) deviations from additivity are expected in all solvent systems, including non-polar solvents that are thought to mimic the interior of a membrane; (3) the spontaneous insertion of polypeptides into membranes is likely to occur much more readily than has been previously thought. In order to estimate the free energy of transferring the side-chains and the polypeptide backbone from water to the interior of a lipid bilayer, the results of this study are used to construct a hydropathy scale based upon the partitioning of solutes between water and non-polar solvents. The validity of hydropathy scales that are based on criteria other than solubility and water/organic solvent partitioning data is also discussed.  相似文献   

4.
Olfactory discrimination of amino acids was investigated in brown bullhead catfish (Ameiurus nebulosus). Based on the magnitude of the observed food search activity of catfish conditioned to single amino acids, the tested compounds were classified as being detected by the catfish as equal to, similar to, or different from the conditioned stimulus. L-Proline (L-Pro)-conditioned brown bullhead catfish discriminated all amino acids from L-Pro, but catfish conditioned to L-valine (L-Val) and L-isoleucine (L-Ile) did not discriminate L-Val from L-Ile nor L-Ile from L-Val; however, all other amino acids tested were always discriminated from these two compounds. Catfish conditioned to L-alanine (L-Ala) discriminated basic, acidic and several neutral amino acids with long side-chains (LCNs) from L-Ala; however, they did not always discriminate L-Ala from all neutral amino acids with short side-chains (SCNs). The L-norleucine (L-nLeu)-conditioned fish responded to L-norvaline (L-nVal), L-methionine (L-Met) and L-Ala similarly to L-nLeu, indicating that these amino acids are detected as similar or identical to L-nLeu. L-nLeu was, however, discriminated from L-Ala in L-Ala-conditioned catfish. Interestingly, L-leucine (L-Leu) was discriminated from the conditioned stimuli, L-Ala, L-Ile and L-Val, indicating independent receptors for L-Leu. Although conditioned catfish discriminated other amino acids from L-arginine hydrochloride (L-Arg), in some tests they were unable to discriminate L-Arg from L-lysine hydrochloride (L-Lys). These results imply the existence of independent olfactory receptive pathways for: (i) L-Pro; (ii) basic amino acids (L-Arg and L-Lys); (iii) L-Leu; (iv) other neutral amino acids with branched side-chains (L-Ile and L-Val); (v) neutral amino acids with long linear side-chains (L-nLeu, L-nVal and L-Met); (vi) neutral amino acids with short side-chains; and (vii) amino acids with sulfhydryl groups (L-Cys and L-homoCys).  相似文献   

5.
Tiunina EIu  Badelin VG 《Biofizika》2005,50(6):965-973
The enthalpies of solvation of 17 amino acids were evaluated by using the sublimation enthalpies of amino acids and the standard enthalpies of their solution in water. An equation was derived, which relates the volume-specific enthalpy of sublimation (deltaH(subl)/V(w)) to the sum of the common bond lengths in molecules (sigman(i)l(i)) of substances examined. The results obtained are interpreted in terms of the effect of hydrophobic and hydrophilic side chain on the interactions between the zwitterions of amino acids and water molecules.  相似文献   

6.
The partial molar heat capacities of the tripeptides of the sequence glycyl-X-glycine, where X is one of the amino acids leucine, threonine, glutamine, phenylalanine, histidine, cysteine, proline, glutamic acid or arginine, and of the two tetrapeptides tetraglycine and glycyltryptophanylglycylglycine in aqueous solution over the temperature range 10-100 degrees C have been determined using high sensitivity scanning microcalorimetry. These results were used to derive the partial molar heat capacities of the various amino acid side-chains. This report completes our programme to derive reliable side-chain heat capacities for all 20 amino acids of proteins over a wide temperature range using the tripeptides Gly-X-Gly as realistic model compounds. Included in the study is a summary of the partial molar heat capacities of all 20 amino acid side-chains. These results, along with the heat capacity of the peptide backbone group, were used to calculate the partial molar heat capacities of some oligopeptides and of the random coil form of some unfolded proteins in water. The calculated heat capacities of the proteins obtained using this new set of heat capacities for the constituent groups are consistent with the heat capacities of the denatured state determined experimentally.  相似文献   

7.
Viscosities for solutions of glycine, DL-alpha-alanine, DL-alpha-amino-n-butyric acid, DL-valine, DL-leucine and L-serine in 5 mol kg(-1) aqueous urea have been determined at 278.15, 288.15, 298.15 and 308.15 K. The viscosity B-coefficients for the amino acids in the aqueous urea solution have been calculated at different temperatures. The effect of temperature on the B-coefficients is discussed on the basis of the Feakins equation. The contribution of solute to the activation parameters (delta mu0*2, deltaH0*2, deltaS0*2) for viscous flow of the solution have been calculated, together with the Gibbs energy, enthalpy and entropy of transfer for the amino acids from the ground-state solvent to the hypothetical viscous transition state solvent. The contributions of the charged end group (NH3+, COO-) and CH2 groups of the amino acids to B-coefficient and delta mu0*2 have been also estimated using the linear correlations between B-coefficient or delta mu0*2 and the number of carbon atoms in the alkyl chains of the amino acids. All the activation parameters are discussed in terms of the solute-solvent interactions in the ground and transition states.  相似文献   

8.
The method of reactive extraction with di-(2-ethylhexyl)phosphoric acid (D2EHPA) for the separation of a range of amino acids is studied. The results obtained on the individual reactive extraction indicated the possibility of the amino acids selective separation as a function of the pH value of aqueous solution and the acidic or basic character of each amino acid. Thus, using multistage extraction, the total separation of the following amino acids groups has been performed: neutral amino acids (l-glycine, l-alanine, l-tryptophan) at pH 5–5.5 (nine extraction stages), basic amino acids (l-lysine, l-arginine) and l-cysteine at pH 4–4.5 (ten extraction stages), l-histidine at pH 3–3.5 (five extraction stages), and acidic amino acids (l-aspartic acid, l-glutamic acid) at pH 2–2.5 (three extraction stages).  相似文献   

9.
A continuum electrostatics approach for molecular dynamics (MD) simulations of macromolecules is presented and analyzed for its performance on a peptide and a globular protein. The approach incorporates the screened Coulomb potential (SCP) continuum model of electrostatics, which was reported earlier. The model was validated in a broad set of tests some of which were based on Monte Carlo simulations that included single amino acids, peptides, and proteins. The implementation for large-scale MD simulations presented in this article is based on a pairwise potential that makes the electrostatic model suitable for fast analytical calculation of forces. To assess the suitability of the approach, a preliminary validation is conducted, which consists of (i) a 3-ns MD simulation of the immunoglobulin-binding domain of streptococcal protein G, a 56-residue globular protein and (ii) a 3-ns simulation of Dynorphin, a biological peptide of 17 amino acids. In both cases, the results are compared with those obtained from MD simulations using explicit water (EW) molecules in an all-atom representation. The initial structure of Dynorphin was assumed to be an alpha-helix between residues 1 and 9 as suggested from NMR measurements in micelles. The results obtained in the MD simulations show that the helical structure collapses early in the simulation, a behavior observed in the EW simulation and consistent with spectroscopic data that suggest that the peptide may adopt mainly an extended conformation in water. The dynamics of protein G calculated with the SCP implicit solvent model (SCP-ISM) reveals a stable structure that conserves all the elements of secondary structure throughout the entire simulation time. The average structures calculated from the trajectories with the implicit and explicit solvent models had a cRMSD of 1.1 A, whereas each average structure had a cRMSD of about 0.8A with respect to the X-ray structure. The main conformational differences of the average structures with respect to the crystal structure occur in the loop involving residues 8-14. Despite the overall similarity of the simulated dynamics with EW and SCP models, fluctuations of side-chains are larger when the implicit solvent is used, especially in solvent exposed side-chains. The MD simulation of Dynorphin was extended to 40 ns to study its behavior in an aqueous environment. This long simulation showed that the peptide has a tendency to form an alpha-helical structure in water, but the stabilization free energy is too weak, resulting in frequent interconversions between random and helical conformations during the simulation time. The results reported here suggest that the SCP implicit solvent model is adequate to describe electrostatic effects in MD simulation of both peptides and proteins using the same set of parameters. It is suggested that the present approach could form the basis for the development of a reliable and general continuum approach for use in molecular biology, and directions are outlined for attaining this long-term goal.  相似文献   

10.
The three-dimensional X-ray structure of the leucine-binding protein (36,900 Mr and 346 residues), an active transport component of Escherichia coli, has been determined by the method of molecular replacement, using the refined structure of the Leu/Ile/Val-binding protein (344 residues) as the model structure. The two amino acid-binding proteins have 80% sequence identity and, although both crystallize in the same space group, they have very different unit cell dimensions. The rotation function yielded one significant peak, which subsequently led to a single self-consistent translation function solution. The model was first refined by the constrained least-squares method, with each of the two domains of the molecule treated separately to allow for any small change in the relative orientation of the two domains. The model was then modified in order to reflect the 72 changes in amino acid side-chains and two insertions in going from the Leu/Ile/Val-binding protein sequence to that of the L-leucine-binding protein. Final structure refinement, using the restrained least-squares technique, resulted in an R-factor of 0.20 for 13,797 reflections to a resolution of 2.4 A. The model is comprised of 2600 protein atoms and 91 solvent molecules. The L-leucine-binding protein structure is, as expected, very similar to the Leu/Ile/Val-binding protein structure; both are in the unliganded conformation with the cleft between the two domains wide open and easily accessible. The superimposing of the structures yields a root-mean-square difference of 0.68 A in the alpha-carbon atoms of the 317 equivalent residues. The five regions of the leucine-binding protein structure that differ by more than 1.6 A from the Leu/Ile/Val-binding protein structure are far from the major portion of the ligand-binding site, which is located in one domain of the bilobate protein. Between the structures, there are three differences in the amino acid side-chains that form the major portion of the substrate-binding sites. These substitutions, by themselves, fail to clearly explain the differences in the specificities for branched aliphatic amino acids.  相似文献   

11.
The amplitude of protein backbone NH group motions on a time-scale faster than molecular tumbling may be determined by analysis of (15)N NMR relaxation data according to the Lipari-Szabo model free formalism. An internet-accessible database has been compiled containing 1855 order parameters from 20 independent NMR relaxation studies on proteins whose three-dimensional structures are known. A series of statistical analyses has been performed to identify relationships between the structural features and backbone dynamics of these proteins. Comparison of average order parameters for different amino acid types indicates that amino acids with small side-chains tend to have greater backbone flexibility than those with large side-chains. In addition, the motions of a given NH group are also related to the sizes of the neighboring amino acids in the primary sequence. The secondary structural environment appears to influence backbone dynamics relatively weakly, with only subtle differences between the order parameter distributions of loop structures and regular hydrogen bonded secondary structure elements. However, NH groups near helix termini are more mobile on average than those in the central regions of helices. Tertiary structure influences are also relatively weak but in the expected direction, with more exposed residues being more flexible on average than residues that are relatively inaccessible to solvent.  相似文献   

12.
Digestion of intact Sindbis virions with α-chymotrypsin produced a single membrane-associated peptide derived from each of the two virion glycoproteins (referred to as RE1 and RE2, or roots derived from E1 and E2, respectively). Amino acid composition data and NH2-terminal sequence analysis established their location at the extreme COOH-terminal end of each glycoprotein. RE1 and RE2 are rich in hydrophobia amino acids and insoluble in aqueous solutions in the absence of detergents, and show differential solubility in organic solvent systems designed for the extraction of lipids. Essentially all of the covalently attached palmitic acid associated with E1 and E2 was found to be clustered in their hydrophobic, membrane-associated roots. Beginning six to seven residues from their NH2 termini, RE1 and RE2 contain uninterrupted sequences of hydrophobic amino acids similar in terms of amino acid composition and length to the transmembrane anchors found in other bitopic integral membrane proteins. By comparing the sequence and composition data obtained here with the sequences of E1 and E2 deduced from complementary DNA sequence analysis (Rice & Strauss, 1981) we can make several observations. First, following their uncharged, putative intramembrane segments (33 and 26 amino acids, respectively), E1 and E2 contain clusters of predominantly basic amino acids. By structural analogy to known transmembrane proteins, E1 probably spans the bilayer but contains only a few residues exposed on the inner face of the virion envelope. In contrast, E2 and PE2 (the precursor to E2), which have been shown to span the bilayer completely, contain an additional 33 COOH-terminal residues, which could be either exposed on the cytoplasmic face of the lipid bilayer or which could loop back into the membrane. This region at the extreme COOH-terminal end of E2, which is protected by the virion envelope from digestion by a-chymotrypsin, contains a second uncharged domain (23 amino acids in length) whose orientation is unknown, but which may be involved in the highly specific interaction of the transmembrane glycoproteins in the plasma membrane with the cytoplasmic nucleocapsid during budding.  相似文献   

13.
A new hexaaza macrocyclic ligand (L) bearing two 2-hydroxypropyl pendants, 6,19-bis(2-hydroxypropyl)-3,6,9,16,19,22-hexaaza-tricyclo-[22.2.2.2(11,14)]triaconta-11,13,24,26,27,29-hexaene has been synthesized and characterized. The macrocyclic ligand was isolated as a colorless crystal, monoclinic, P2(1)/n, with a=10.757(2), b=14.214(3), c=13.746(3) A, beta=101.40(3) degrees, V=2060.3(7) A3, Z=2, R1=0.0695, and wR2=0.1538 [I>2sigma(I)]. Potentiometric studies of the macrocyclic ligand and three types of amino acids, glycine (equal numbers of carboxylate and amino groups), aspartic acid (more carboxylate groups than amino group), and lysine (more amino groups than carboxylate group) have been performed. The stability constants for the new macrocycle and binary complexes of the amino acid with the macrocyclic ligand are reported. Binary complexes are formed in aqueous solution as a result of hydrogen bonding interaction and electrostatic attraction between the host and the guest. The binding Schemes for the recognition of amino acids are suggested. From the results, it seems that this new macrocyclic ligand is able to bind three different amino acids with selectivity in aqueous solution, and the strength of binding is of the order lysine < glycine < aspartic acid.  相似文献   

14.
A new human/rat CRH analogue has been synthesized using the Fmoc/tBu solid-phase synthetic protocol. The sequence of the new peptide differs from the original in two positions, 12 and 15, at which the native amino acids l-phenylalanine 12 and l-leucine 15 have been replaced by the nonprotein amino acids d-phenylalanine and alpha-aminoisobutyric acid (Aib), respectively. The high resolution three-dimensional solution structure of [d-Phe12, Aib15]CRH has been determined by 688 distance constraints (656 meaningful NOE and 32 H-bonds distance limits) and 21 angle constraints. A family of 40 energy-minimized conformers was obtained with average rmsd of 0.39 +/- 0.16 A and 0.99 +/- 0.13 A for backbone and heavy atoms, respectively, and distance penalty functions of 0.42 +/- 0.03 A2. The NMR data acquired in a solvent system of water/trifluoroethanol (34%/66%, v/v) revealed that this 41-polypeptide adopts an almost linear helical structure in solution with helical content which reaches an 84% of the residues. Structural analysis confirmed the existence of two helical peptide fragments. The first was comprised of residues Ile6-Arg16 and the second of residues Glu20-Ile40, forming an angle of 34.2 degrees. The structural differences with respect to the native peptide have been identified in the region d-Phe12-Glu20 where double substitution at positions 12 and 15 seems to perturb the elements of the native 35-residue helix. These structural rearrangements promote non-native intramolecular interactions in the region of the molecule between either the hydrophobic side-chains of d-Phe12, Aib15 and Leu18, or the charged groups of the residue pairs Arg16-Glu20 and His13-Glu17 being responsible for changes in hormonal functionality. This CRH analogue currently exhibits lack of any activity.  相似文献   

15.
The reactions of hydrated electrons (eaq-) with 55 dipeptides and 25 acetyl and formyl amino acids have been studied by e.s.r. and spin-trapping techniques. Gamma-radiolysis of deaerated aqueous solutions was used to generate eaq-, and sodium formate or t-BuOH was added to scavenge the OH radicals. t-Nitrosobutane was employed as the spin-trapping reagent. The radical,--CO---NH--, which is the initial product of the reactions of eaq- with dipeptides, was observed only for val-gly, val-ala, val-leu and ile-ala. For most of the dipeptides this radical converts to the primary deamination radical, CHR'-CONH-CHR-COO-, where R and R' are the side-chains of the common amino acids. In many cases a radical of the type CHR-COO-, formed by secondary deamination, was also observed. Only secondary deamination reactions were observed for dipeptides containing beta-alanine as the amino terminal residue and for acetyl and formyl amino acids. The secondary deamination reactions of eaq- with dipeptides, acetyl and formyl amino acids in aqueous solutions have not been observed previously. This type of reaction is of interest since it brings about main-chain scission in polypeptides and proteins.  相似文献   

16.
Water activity, pH and density of aqueous amino acids solutions   总被引:2,自引:0,他引:2  
The water activity, pH and density of some aqueous amino acid solutions were determined at 25 degrees C in three different types of solvents. Previous published experimental data on water activity and solubility of amino acids in aqueous solutions were used together with data from this work to test the applicability of a group contribution model. The activity coefficients were estimated by the UNIFAC-Larsen model combined with the Debye-Hückel equation, taking also into account the partial dissociation phenomena of species in solution. Interaction energies between the charged species Na(+) and Cl(-) and the specific groups of amino acids (COOH and NH(2)) were adjusted using experimental solubility data.  相似文献   

17.
Partition ratios of 8 free l-amino acids (Gln, Glu, His, Lys, Met, Ser, Thr, and Tyr) were measured in 10 different polymer/polymer aqueous two-phase systems containing 0.15?M NaCl in 0.01?M phosphate buffer, pH 7.4. The solute-specific coefficients representing the solute dipole/dipole, hydrogen-bonding and electrostatic interactions with the aqueous environment of the amino acids were determined by multiple linear regression analysis using a modified linear solvation energy relationship. The solute-specific coefficients determined in this study together with the solute-specific coefficients reported previously for amino acids with non-polar side-chains where used in a Quantitative Structure/Property Relationship analysis. It is shown that linear combinations of these solute-specific coefficients are correlated well with various physicochemical, structural, and biological properties of amino acids.  相似文献   

18.
Amino acids in peptides and proteins display distinct preferences for alpha-helical, beta-strand, and other conformational states. Various physicochemical reasons for these preferences have been suggested: conformational entropy, steric factors, hydrophobic effect, and backbone electrostatics; however, the issue remains controversial. It has been proposed recently that the side-chain-dependent solvent screening of the local and non-local backbone electrostatic interactions primarily determines the preferences not only for the alpha-helical but also for all other main-chain conformational states. Side-chains modulate the electrostatic screening of backbone interactions by excluding the solvent from the vicinity of main-chain polar atoms. The deficiency of this electrostatic screening model of amino acid preferences is that the relationships between the main-chain electrostatics and the amino acid preferences have been demonstrated for a limited set of six non-polar amino acid types in proteins only. Here, these relationships are determined for all amino acid types in tripeptides, dekapeptides, and proteins. The solvation free energies of polar backbone atoms are approximated by the electrostatic contributions calculated by the finite difference Poisson-Boltzmann and the Langevin dipoles methods. The results show that the average solvation free energy of main-chain polar atoms depends strongly on backbone conformation, shape of side-chains, and exposure to solvent. The equilibrium between the low-energy beta-strand conformation of an amino acid (anti-parallel alignment of backbone dipole moments) and the high-energy alpha conformation (parallel alignment of backbone dipole moments) is strongly influenced by the solvation of backbone polar atoms. The free energy cost of reaching the alpha conformation is by approximately 1.5 kcal/mol smaller for residues with short side-chains than it is for the large beta-branched amino acid residues. This free energy difference is comparable to those obtained experimentally by mutation studies and is thus large enough to account for the distinct preferences of amino acid residues. The screening coefficients gamma(local)(r) and gamma(non-local)(r) correlate with the solvation effects for 19 amino acid types with the coefficients between 0.698 to 0.851, depending on the type of calculation and on the set of point atomic charges used. The screening coefficients gamma(local)(r) increase with the level of burial of amino acids in proteins, converging to 1.0 for the completely buried amino acid residues. The backbone solvation free energies of amino acid residues involved in strong hydrogen bonding (for example: in the middle of an alpha-helix) are small. The hydrogen bonded backbone is thus more hydrophobic than the peptide groups in random coil. The alpha-helix forming preference of alanine is attributed to the relatively small free energy cost of reaching the high-energy alpha-helix conformation. These results confirm that the side-chain-dependent solvent screening of the backbone electrostatic interactions is the dominant factor in determining amino acid conformational preferences.  相似文献   

19.
An investigation has been undertaken to determine whether ionizing radiation might engender racemization (radioracemization) of optically active amino acids, along with their well-known radiolysis. We have exposed a number of solid and dissolved optically active amino acids to the ionizing radiation from a 3000-Ci 60Co γ-ray source for periods of time which would engender substantial, but not total radiolysis. γ-Ray doses which caused 55–68% radiolysis of solid amino acids typically engendered 2–5% racemization. Aqueous solutions of the sodium salts of amino acids which underwent 53–66% radiolysis typically showed 5–11% racemization. The corresponding hydrochloride salts in aqueous solution, however, underwent little or no racemization. In aqueous solution both percentage degradation and percentage racemization were approximately proportional to γ-ray dosage within the range employed (1–36 × 106 rads). Mechanisms for the radioracemization of amino acids in the solid state and as dissolved sodium salts are proposed, and the absence of racemization for dissolved hydrochloride salts is rationalized. Implications of these observations with regard to the origin of optical activity by the Vester-Ulbricht β-decay mechanism are discussed, as are their implications regarding the use of diagenetic racemization rates of ancient amino acid samples as criteria for geochronological and geothermometric calculations.  相似文献   

20.
Mass spectrometry (MS) was used to characterise the binding of the 58 kDa protein OppA to 11 peptides with diverse properties. Peptides with two, three and five amino acid residues were added to OppA, and the mass spectra showed that the highest-affinity complexes are formed between OppA and tripeptide ligands. Lower-affinity complexes were observed for OppA and dipeptide ligands, and no complex formation was detected with pentapeptides or a tripeptide in which the N-terminal amino group was acetylated. Tripeptides containing a single d amino acid residue were found not to bind to native OppA. Evidence from the peak width and the, charge in the spectra of the complexes suggests that the bound peptides are encapsulated by the protein in a solvent-filled cavity in the gas phase of the mass spectrometer. Analysis of the proportions of peptide-bound and free proteins under low-energy MS conditions shows a good correlation with solution-phase K(d) measurements where available. Increasing the internal energy of the gas-phase complex led to dissociation of the complex. The ease of dissociation is interpreted in terms of the intrinsic stability of the complex in the absence of the stabilising effects of bulk solvent. The results from this study demonstrate insensitivity to the hydrophobic and ionic properties, of the side-chains of the peptides, in contrast to the investigation of other protein ligand systems by MS. Moreover, these findings are in accord with the physiological role of this protein in allowing into the cell di- and tripeptides containing naturally occurring amino acids, regardless of their sequence, while barring access to potentially harmful peptide mimics.  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