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.
BackgroundFragment-based ligand design is used for the development of novel ligands that target macromolecules, most notably proteins. Central to its success is the identification of fragment binding sites that are spatially adjacent such that fragments occupying those sites may be linked to create drug-like ligands. Current experimental and computational approaches that address this problem typically identify only a limited number of sites as well as use a limited number of fragment types.MethodsThe site-identification by ligand competitive saturation (SILCS) approach is extended to the identification of fragment bindings sites, with the method termed SILCS-Hotspots. The approach involves precomputation of the SILCS FragMaps following which the identification of Hotspots, performed by identifying of all possible fragment binding sites on the full 3D structure of the protein followed by spatial clustering.ResultsThe SILCS-Hotspots approach identifies a large number of sites on the target protein, including many sites not accessible in experimental structures due to low binding affinities and binding sites on the protein interior. The identified sites are shown to recapitulate the location of known drug-like molecules in both allosteric and orthosteric binding sites on seven proteins including the androgen receptor, the CDK2 and Erk5 kinases, PTP1B phosphatase and three GPCRs; the β2-adrenergic, GPR40 fatty-acid binding and M2-muscarinic receptors. Analysis indicates the importance of considering all possible fragment binding sites, and not just those accessible to experimental methods, when identifying novel binding sites and performing ligand design versus just considering the most favorable sites. The approach is shown to identify a larger number of known binding sites of drug-like molecules versus the commonly used FTMap and Fpocket methods.General significanceThe present results indicate the potential utility of the SILCS-Hotspots approach for fragment-based rational design of ligands, including allosteric modulators.  相似文献   

2.
Protein docking methods are powerful computational tools to study protein-protein interactions (PPI). While a significant number of docking algorithms have been developed, they are usually based on rigid protein models or with limited considerations of protein flexibility and the desolvation effect is rarely considered in docking energy functions, which may lower the accuracy of the predictions. To address these issues, we introduce a PPI energy function based on the site-identification by ligand competitive saturation (SILCS) framework and utilize the fast Fourier transform (FFT) correlation approach. The free energy content of the SILCS FragMaps represent an alternative to traditional energy grids and they can be efficiently utilized to guide FFT-based protein docking. Application of the approach to eight diverse test cases, including seven from Protein Docking Benchmark 5.0, showed the PPI prediction using SILCS approach (SILCS-PPI) to be competitive with several commonly used protein docking methods indicating that the method has the ability to both qualitatively and quantitatively inform the prediction of PPI. Results show the utility of the SILCS-PPI docking approach for determination of probability distributions of PPI interactions over the surface of both partner proteins, allowing for identification of alternate binding poses. Such binding poses are confirmed by experimental crystal contacts in our test cases. While more computationally demanding than available PPI docking technologies, we anticipate that the SILCS-PPI docking approach will offer an alternative methodology for improved evaluation of PPIs that could be used in a variety of fields from systems biology to excipient design for biologics-based drugs.  相似文献   

3.
Structure of prothrombin fragment 1 refined at 2.8 A resolution   总被引:4,自引:0,他引:4  
The structure of prothrombin fragment 1, solved at 2.8 A resolution (1 A = 0.1 nm) by a combination of multiple and single isomorphous replacement methods utilizing solvent flattening, has been refined by restrained least-squares methods (R = 0.24), solvent not included, using fairly stringent restraints on the molecular geometry and individual thermal parameters. The inner kringle loop possesses significantly lower B-values than the outer loops even though the former also constitutes a surface of the folded kringle structure. This surface forms the Lys sub-site of the fibrin binding site of other kringles. The hydrogen bonding network and ion pair interactions of fragment 1 appear to maintain a compact folded structure among the various loops of the kringle structure. On the other hand, since there is only one hydrogen bond between the kringle and its preceding 30 residues, considerable flexibility is suggested for the Gla-domain consistent with its disorder in crystals. A chitobiose has been located at the Asn77 glycosylation site, but only a single N-acetyl-glucosamine is ordered at Asn101. The lysine binding site region of other kringles is not properly developed in fragment 1, accounting for its lack of Lys/fibrin affinity. Most of the conserved sequence among 11 different kringles is associated with either: (1) protecting the inner loop disulfides Cys87-127, Cys115-139 upon which the folding is based; or (2) a requirement of the lysine binding site. The remainder of the conservation is generally associated with the ten reverse turns of the folding; of these 40 residues, or about half the sequence, 14 are conserved among eight different turns. The intermolecular packing consists of infinite helical columns of fragment 1 molecules related by a crystallographic 4(3) screw axis, which are held together by van der Waals' interactions of aromatic clusters from different molecules related by a crystallographic 2-fold rotation axis.  相似文献   

4.
Three sets of in silico experiments have been conducted to elucidate the binding mechanics of two drugs, (+)-methamphetamine (METH) and amphetamine (AMP) to the single-chain variable fragment (scFv) recently engineered from anti-METH monoclonal antibody mAb6H4 (IgG, κlight chain, Kd = 11 nM). The first set of in silico experiments are long time equilibration runs of scFv:drug complexes and of drug-free scFv both in the solution. They demonstrate how the solution structures of scFv deviate from its crystallographic form with or without drug molecules bound to it. They lead to the prediction that the Arrhenius activation barrier is nearly zero for transitions from the dissociated state to the bound state. The second set of in silico experiments are nonequilibrium dynamics of pulling the drug molecules out of the binding pocket of scFv and the equilibration runs for drugs to fall back into the binding pocket. They demonstrate that extra water molecules (in addition to the two crystallographic waters) exist inside the binding pocket, underneath the drug molecules. These extra waters must have been evaporated from the binding pockets during the crystallization process of the in vitro experiments of structural determination. The third set of in silico experiments are nonequilibrium steered molecular dynamics simulations to determine the absolute binding free energies of METH and AMP to scFv. The center of mass of a drug molecule (METH or AMP) is steered (pulled) towards (forward) and away from (reverse) the binding site, sampling forward and reverse pulling paths. Mechanic work is measured along the pulling paths. The work measurements are averaged through the Brownian dynamics fluctuation dissipation theorem to produce the free-energy profiles of the scFv:drug complexes as a function of the drug-scFv separation. These experiments lead to the theoretical prediction of absolute binding energies of METH and AMP that are in agreement with the in vitro experimental results.  相似文献   

5.
Despite the high density within a typical protein fold, the ensemble of sterically permissible side-chain repackings is vast. Here, we examine the extent of this variability that survives energetic biases due to van der Waals interactions, hydrogen bonding, salt bridges, and solvation. Monte Carlo simulations of an atomistic model exhibit thermal fluctuations among a diverse set of side-chain arrangements, even with the peptide backbone fixed in its crystallographic conformation. We have quantified the torsional entropy of this native-state ensemble, relative to that of a noninteracting reference system, for 12 small proteins. The reduction in entropy per rotatable bond due to each kind of interaction is remarkably consistent across this set of molecules. To assess the biophysical importance of these fluctuations, we have estimated side-chain entropy contributions to the binding affinity of several peptide ligands with calmodulin. Calculations for our fixed-backbone model correlate very well with experimentally determined binding entropies over a range spanning more than 80 kJ/(mol·308 K).  相似文献   

6.
Molecular dynamics (MD) simulations were used to characterize the non-cooperative denaturation of the molten globule A-state of human alpha-lactalbumin by urea. A solvent of explicit urea and water molecules was used, corresponding to a urea concentration of approximately 6M. Three simulations were performed at temperatures of 293K, 360K and 400K, with lengths of 2 ns, 8 ns and 8 ns respectively. The results of the simulations were compared with experimental data from NMR studies of human alpha-lactalbumin and related peptides. During the simulations, hydrogen bonds were formed from the protein to both urea and water molecules as intra-protein hydrogen bonds were lost. Urea was shown to compete efficiently with water as both a hydrogen bond donor and acceptor. Radial distribution functions of water and urea around hydrophobic side chain atoms showed a significant increase in urea molecules in the solvation shell as the side chains became exposed during denaturation. A considerable portion of the native-like secondary structure persisted throughout the simulations. However, in the simulations at 360K and 400K, there were substantial changes in the packing of aromatic and other hydrophobic side chains in the protein, and many native contacts were lost. The results suggest that during the non-cooperative denaturation of the molten globule, secondary structure elements are stabilized by non-specific, non-native interactions.  相似文献   

7.
The three-dimensional structure of the 131-residue rat intestinal fatty acid-binding protein, without bound ligand (apoI-FABP), has been refined with x-ray diffraction data to a nominal resolution of 1.19 A. The final model has a conventional crystallographic R-factor of 16.9% for 34,290 unique reflections [a root mean square (r.m.s.) deviation for bond length of 0.012 A and a r.m.s. deviation of 2.368 degrees for bond angles]. Ninety-two residues are present as components of the protein's 10 anti-parallel beta-strands while 14 residues are part of its two short alpha-helices. The beta-strands and alpha-helices are organized into two nearly orthogonal beta-sheets. Particular attention has been placed in defining solvent structure and the structures of discretely disordered groups in this protein. Two hundred thirty-seven solvent molecules have been identified; 24 are located within apoI-FABP. The refined model includes alternate conformers for 228 protein atoms (109 main-chain, 119 side-chain) and 63 solvent molecules. We have found several aromatic side-chains with multiple conformations located near, or in, the protein's ligand binding site. This observation, along with the fact that these side-chains have a temperature factor that is relatively higher than that of other aromatic residues, suggests that they may be involved in the process of noncovalent binding of fatty acid. The absence of a true hydrophobic core in I-FABP suggests that its structural integrity may be maintained primarily by a hydrogen bonding network involving protein and solvent atoms.  相似文献   

8.
The interaction of benzamide with the isolated components of calf thymus poly(ADP-ribose) polymerase and with liver nuclei has been investigated. A benzamide-agarose affinity gel matrix was prepared by coupling o-aminobenzoic acid with Affi-Gel 10, followed by amidation. The benzamide-agarose matrix bound the DNA that is coenzymic with poly(ADP-ribose) polymerase; the matrix, however, did not bind the purified poly(ADP-ribose) polymerase protein. A highly radioactive derivative of benzamide, the 125I-labelled adduct of o-aminobenzamide and the Bolton-Hunter reagent, was prepared and its binding to liver nuclear DNA, calf thymus DNA and specific coenzymic DNA of poly(ADP-ribose) polymerase was compared. The binding of labelled benzamide to coenzymic DNA was several-fold higher than its binding to unfractionated calf thymus DNA. A DNA-related enzyme inhibitory site of benzamide was demonstrated in a reconstructed poly(ADP-ribose) polymerase system, made up from purified enzyme protein and varying concentrations of a synthetic octadeoxynucleotide that serves as coenzyme. As a model for benzamide binding to DNA, a crystalline complex of 9-ethyladenine and benzamide was prepared and its X-ray crystallographic structure was determined; this indicated a specific hydrogen bond between an amide hydrogen atom and N-3 of adenine. The benzamide also formed a hydrogen bond to another benzamide molecule. The aromatic ring of benzamide does not intercalate between ethyladenine molecules, but lies nearly perpendicular to the planes of stacking ethyladenine molecules in a manner reminiscent of the binding of ethidium bromide to polynucleotides. Thus we have identified DNA as a site of binding of benzamide; this binding is critically dependent on the nature of the DNA and is high for coenzymic DNA that is isolated with the purified enzyme as a tightly associated species. A possible model for such binding has been suggested from the structural analysis of a benzamide-ethyladenine complex.  相似文献   

9.
The periplasmic molecular chaperone protein SurA facilitates correct folding and maturation of outer membrane proteins in Gram-negative bacteria. It preferentially binds peptides that have a high fraction of aromatic amino acids. Phage display selections, isothermal titration calorimetry and crystallographic structure determination have been used to elucidate the basis of the binding specificity. The peptide recognition is imparted by the first peptidyl-prolyl isomerase (PPIase) domain of SurA. Crystal structures of complexes between peptides of sequence WEYIPNV and NFTLKFWDIFRK with the first PPIase domain of the Escherichia coli SurA protein at 1.3 A resolution, and of a complex between the dodecapeptide and a SurA fragment lacking the second PPIase domain at 3.4 A resolution, have been solved. SurA binds as a monomer to the heptapeptide in an extended conformation. It binds as a dimer to the dodecapeptide in an alpha-helical conformation, predicated on a substantial structural rearrangement of the SurA protein. In both cases, side-chains of aromatic residues of the peptides contribute a large fraction of the binding interactions. SurA therefore asserts a recognition preference for aromatic amino acids in a variety of sequence configurations by adopting alternative tertiary and quaternary structures to bind peptides in different conformations.  相似文献   

10.
Cyclomaltoheptaose (cycloheptaamylose) has been crystallized with 1-adamantanemethanol as the guest molecule. The complex crystallized in space group C222(1), with unit-cell dimensions a = 19.162 (13), b = 23.965 (17), and c = 32.597 (27) A. The structure was solved by rotation-translation search-methods. The cyclomaltoheptaose exists as a dimer in the crystal by means of extensive hydrogen-bonding across the secondary hydroxyl ends of two cyclomaltoheptaose molecules. The two halves of the dimer are related by a crystallographic two-fold axis. The primary hydroxyl ends of two adjacent cyclomaltoheptaose molecules are also related by a crystallographic two-fold axis, but do not directly hydrogen bond to one another. Instead, they are held in place by a strong hydrogen bond from the hydroxyl group of the 1-adamantanemethanol to a primary hydroxyl group on an adjacent cyclomaltoheptaose molecule. Other stabilizing hydrogen bonds are formed via three water molecules which are situated at the primary hydroxyl interface, and others that form parallel columns stabilizing the crystal structure. A unique feature of this complex is the presence of trapped water in the cavity at the secondary hydroxyl interface. This water is distributed over 3 disordered sites. Its presence blocks one possible site for the 1-adamantanemethanol, which, instead, binds near the primary hydroxyl end, with its hydroxyl group and part of the adamantane moiety protruding from the cyclomaltoheptaose.  相似文献   

11.
A comprehensive analysis of interfacial water molecules in the structures of 109 unique protein-DNA complexes is presented together with a new view on their role in protein-DNA recognition. Location of interfacial water molecules as reported in the crystal structures and as emerging from a series of molecular dynamics studies on protein-DNA complexes with explicit solvent and counterions, was analyzed based on their acceptor, donor hydrogen bond relationships with the atoms and residues of the macromolecules, electrostatic field calculations and packing density considerations. Water molecules for the purpose of this study have been categorized into four classes: viz. (I) those that contact both the protein and the DNA simultaneously and thus mediate recognition directly; (II) those that contact either the protein or the DNA exclusively via hydrogen bonds solvating each solute separately; (III) those that contact the hydrophobic groups in either the protein or the DNA; and, lastly (IV) those that contact another water molecule. Of the 17,963 crystallographic water molecules under examination, about 6% belong to class I and 76% belong to class II. About three-fourths of class I and class II water molecules are exclusively associated with hydrogen bond acceptor atoms of both protein and DNA. Noting that DNA is polyanionic, it is significant that a majority of the crystallographically observed water molecules as well as those from molecular dynamics simulations should be involved in facilitating binding by screening unfavorable electrostatics. Less than 2% of the reported water molecules occur between hydrogen bond donor atoms of protein and acceptor atoms of DNA. These represent cases where protein atoms cannot reach out to DNA to make favorable hydrogen bond interactions due to packing/structural restrictions and interfacial water molecules provide an extension to side-chains to accomplish hydrogen bonding.  相似文献   

12.
Short contacts of water molecules with the pi-faces of aromatic residues were studied in a set of 75 very high resolution (<1.1 A) protein X-ray crystal structures. For 18 water molecules found at distances to aromatic midpoints <3.5 A, it was attempted to assign the hydrogen bond configuration (without experimental knowledge of the H-atom positions) by inspection of the surrounding. For approximately one-quarter of the cases, evidence for an O-H...pi hydrogen bond was found, another one-quarter does not form such a hydrogen bond, and for the remaining half, no conclusive assignment could be made. The results confirm the relatively frequent occurrence of aromatic hydrogen bonding in biomolecular hydration, but also underline difficulties in hydrogen bond assignment without reliable knowledge of the H-atom positions.  相似文献   

13.
The accurate prediction of protein druggability (propensity to bind high-affinity drug-like small molecules) would greatly benefit the fields of chemical genomics and drug discovery. We have developed a novel approach to quantitatively assess protein druggability by computationally screening a fragment-like compound library. In analogy to NMR-based fragment screening, we dock ∼11000 fragments against a given binding site and compute a computational hit rate based on the fraction of molecules that exceed an empirically chosen score cutoff. We perform a large-scale evaluation of the approach on four datasets, totaling 152 binding sites. We demonstrate that computed hit rates correlate with hit rates measured experimentally in a previously published NMR-based screening method. Secondly, we show that the in silico fragment screening method can be used to distinguish known druggable and non-druggable targets, including both enzymes and protein-protein interaction sites. Finally, we explore the sensitivity of the results to different receptor conformations, including flexible protein-protein interaction sites. Besides its original aim to assess druggability of different protein targets, this method could be used to identifying druggable conformations of flexible binding site for lead discovery, and suggesting strategies for growing or joining initial fragment hits to obtain more potent inhibitors.  相似文献   

14.
The crystal structure of a periplasmic l-aspartate/l-glutamate binding protein (DEBP) from Shigella flexneri complexed with an l-glutamate molecule has been determined and refined to an atomic resolution of 1.0 Å. There are two DEBP molecules in the asymmetric unit. The refined model contains 4462 non-hydrogen protein atoms, 730 water molecules, 2 bound glutamate molecules, and 2 Tris molecules from the buffer used in crystallization. The final Rcryst and Rfree factors are 13.61% and 16.89%, respectively. The structure has root-mean-square deviations of 0.016 Å from standard bond lengths and 2.35° from standard bond angles.The DEBP molecule is composed of two similarly folded domains separated by the ligand binding region. Both domains contain a central five-stranded β-sheet that is surrounded by several α-helices. The two domains are linked by two antiparallel β-strands. The overall shape of DEBP is that of an ellipsoid approximately 55 Å × 45 Å × 40 Å in size.The binding of ligand to DEBP is achieved mostly through hydrogen bonds between the glutamate and side-chain and main-chain groups of DEBP. Side chains of residues Arg24, Ser72, Arg75, Ser90, and His164 anchor the deprotonated γ-carboxylate group of the glutamate with six hydrogen bonds. Side chains of Arg75 and Arg90 form salt bridges with the deprotonated α-carboxylate group, while the main-chain amide groups of Thr92 and Thr140 form hydrogen bonds with the same group. The positively charged α-amino group of the l-glutamate forms salt bridge interaction with the side-chain carboxylate group of Asp182 and hydrogen bond interaction with main-chain carbonyl oxygen of Ser90. In addition to these hydrogen bond and electrostatic interactions, other interactions may also play important roles. For example, the two methylene groups from the glutamate form van der Waals interactions with hydrophobic side chains of DEBP.Comparisons with several other periplasmic amino acid binding proteins indicate that DEBP residues involved in the binding of α-amino and α-carboxylate groups of the ligand and the pattern of hydrogen bond formation between these groups are very well conserved, but the binding pocket around the ligand side chain is not, leading to the specificity of DEBP. We have identified structural features of DEBP that determine its ability of binding glutamate and aspartate, two molecules with different sizes, but discriminating against very similar glutamine and asparagine molecules.  相似文献   

15.
Given thousands of proteins constituting a eukaryotic pathogen, the principal objective for a high-throughput in silico vaccine discovery pipeline is to select those proteins worthy of laboratory validation. Accurate prediction of T-cell epitopes on protein antigens is one crucial piece of evidence that would aid in this selection. Prediction of peptides recognised by T-cell receptors have to date proved to be of insufficient accuracy. The in silico approach is consequently reliant on an indirect method, which involves the prediction of peptides binding to major histocompatibility complex (MHC) molecules. There is no guarantee nevertheless that predicted peptide-MHC complexes will be presented by antigen-presenting cells and/or recognised by cognate T-cell receptors. The aim of this study was to determine if predicted peptide-MHC binding scores could provide contributing evidence to establish a protein’s potential as a vaccine. Using T-Cell MHC class I binding prediction tools provided by the Immune Epitope Database and Analysis Resource, peptide binding affinity to 76 common MHC I alleles were predicted for 160 Toxoplasma gondii proteins: 75 taken from published studies represented proteins known or expected to induce T-cell immune responses and 85 considered less likely vaccine candidates. The results show there is no universal set of rules that can be applied directly to binding scores to distinguish a vaccine from a non-vaccine candidate. We present, however, two proposed strategies exploiting binding scores that provide supporting evidence that a protein is likely to induce a T-cell immune response–one using random forest (a machine learning algorithm) with a 72% sensitivity and 82.4% specificity and the other, using amino acid conservation scores with a 74.6% sensitivity and 70.5% specificity when applied to the 160 benchmark proteins. More importantly, the binding score strategies are valuable evidence contributors to the overall in silico vaccine discovery pool of evidence.  相似文献   

16.
With the aid of receptor-oriented pharmacophore-based in silico screening, we established three pharmacophore maps explaining the binding model of hPNMT and a known inhibitor, SK&F 29661 (Martin et al., 2001). The compound library was searched using these maps. Nineteen selected candidate inhibitors of hPNMT were screened using STD-NMR and fluorescence experiments. An enzymatic activity assay based on HPLC was additionally performed. Consequently, three potential hPNMT inhibitors were identified, specifically, 4-oxo-1,4-dihydroquinoline-3,7-dicarboxylic acid, 4-(benzo[d][1,3]dioxol-5-ylamino)-4-oxobutanoic acid, and 1,4-diaminonaphthalene-2,6-disulfonic acid. These novel inhibitors were retrieved using Map II comprising one hydrogen bond acceptor, one hydrogen bond donor, one lipophilic feature, and shape constraints, including a hydrogen bond between Lys57 of hPNMT and a hydrogen bond donor of the inhibitor, and stacked hydrophobic interactions between the side-chain of Phe182 and an aromatic region of the inhibitor. Water-mediated interactions between Asn267 and Asn39 of hPNMT and the amide or amine group of three potent inhibitors were additional important features for hPNMT activity. The binding model presented here may be applied to identify inhibitors with higher potency. Moreover, our novel compounds are valuable candidates for further lead optimization of PNMT inhibitors.  相似文献   

17.
Two new tricyclic β-aminoacrylate derivatives (2e and 3e) have been found to be inhibitors of Plasmodium falciparum dihydroorotate dehydrogenase (PfDHODH) with Ki 0.037 and 0.15 μM respectively. 1H and 13C NMR spectroscopic data show that these compounds undergo ready cis-trans isomerisation at room temperature in polar solvents. In silico docking studies indicate that for both molecules there is neither conformation nor double bond configuration which bind preferentially to PfDHODH. This flexibility is favourable for inhibitors of this channel that require extensive positioning to reach their binding site.  相似文献   

18.
19.
Virtual screening and QSAR analysis were carried out to investigate the binding features of (2R, 3R, 4S)-2-aminomethylpyrrolidine 3,4-diol and the functionalized pyrrolidine derivatives to the α-mannosidase I and II enzymes. The QSAR models (possessed considerable R2, Q2 values, etc.) suggested that the presence of polar property on the vdW surface (vsurf_W, vsurf_Wp, etc.) of the molecules is important along with the presence of aromatic rings (opr_violation) in the molecules (which also provide hydrophobicity to the molecules). The docking study performed on α-mannosidase I and II enzymes pointed that the main interactions occur by hydrogen bonds, hydrophobic π–π stacking contacts and salt bridges with the cation calcium (for α-mannosidase I) and close interaction with zinc ion (α-mannosidase II), respectively. The bond flexibility orientates the aromatic ring in the molecules toward the hydrophobic cavity for π–π stacking contacts with the aromatic amino acids (Phe528, Phe329 and Phe659 for α-mannosidase I and Trp95, Tyr269, Phe312, Tyr102 for α-mannosidase II). The pharmacophore analysis also supports the results derived from the docking and QSAR studies. Our results suggest that the best compound to inhibit both classes of α-mannosidase is the compound 30, which may be used to design similar and better inhibitors to next generation drugs.  相似文献   

20.
Reaction of a high-temperature solid-phase catalytic isotope exchange in peptides and proteins under the action of the catalytically activated spillover hydrogen was studied. The reaction of human recombinant insulin with deuterium and tritium at 120–140°C resulted in an incorporation of 2–6 isotope hydrogen atoms per one insulin molecule. The distribution of the isotopic label by amino acid residues of the tritium-labeled insulin was determined by the oxidation of the protein S-S-bonds by performic acid, separation of polypeptide chains, their subsequent acidic hydrolysis, amino acid analysis, and liquid scintillation counts of tritium in the amino acids. The isotopic label was shown to be incorporated in all the amino acid residues of the protein, but the higher inclusion was observed for the FVNQHLCGSHLVE peptide fragment (B1–13) of the insulin B-chain, and the His5 and His10 residues of this fragment contained approximately 45% of the whole isotopic label of the protein. Reduction of the S-S-bonds by 2-mercaptoethanol, enzymatic hydrolysis by glutamyl endopeptidase from Bacillus intermedius, and HPLC fractionation of the obtained peptides were also used for the analysis of the distribution of the isotopic label in the peptide fragments of the labeled insulin. Peptide fragments which were formed after the hydrolysis of the Glu-Xaa bond of the B-chain were identified by mass spectrometry. The mass spectrometric analysis of the isotopomeric composition of the deuterium-labeled insulin demonstrated that all the protein molecules participated equally in the reaction of the solid-phase hydrogen isotope exchange. The tritium-labeled insulin preserved the complete physiological activity.  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