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 62 毫秒
1.
Abstract: Identification of residues in the skeletal muscle nicotinic acetylcholine receptor (AChR) that bind snake venom a-neurotoxin antagonists of acetylcholine [e.g., α-bungarotoxin (α-BTx)] provides structural information about the neurotransmitter binding region of the receptor. Using synthetic peptides of the human AChR α-subunit region 177–208, we previously localized a pharmacologically specific binding site for α-BTx in segment 185–199. To define in more detail the residues that influence the binding of α-BTx to this region, we prepared 16 peptide analogues of the α-subunit segment 185–200, with the amino acid Lalanine sequentially replacing each native amino acid. Circular dichroism spectroscopy did not reveal changes in the secondary structure of the peptides except for the analogue in which Pro194 was substituted with alanine. This implies that any change in α-BTx binding could be attributed to replacement of the native residue's side chain by alanine's methyl group, rather than to a change in the structure of the peptide. The influence of each substitution with alanine was determined by comparing the analogue to the parental sequence α 185–200 in solution-phase competition with native human AChR for binding of 125I-labeled α-BTx. The binding of α-BTx by analogue peptides with alanine substituted for Tyr190, Cys192, or Cys193 was greatly diminished. Binding of α-BTx to peptides containing alanine replacements at Val188, Thr189, Pro194, Asp195, or Tyr198 was also reduced significantly (p < 0.003). An unanticipated finding was that substitution of alanine for Ser191 significantly increased α-BTx binding (p < 0.003). The data imply that these nine amino acids influence the binding of the antagonist, α-BTx, to the nicotinic acetylcholine receptor of human skeletal muscle, and confirm previous reports for certain contact residues for α-BTX that were found in region α181-200 of the Torpedo AChR.  相似文献   

2.
The N-terminal domain of HIV-1 glycoprotein 41?000 (FP; residues 1-23; AVGIGALFLGFLGAAGSTMGARSCONH2) participates in fusion processes underlying virus-cell infection. Here, we use physical techniques to study the secondary conformation of synthetic FP in aqueous, structure-promoting, lipid and biomembrane environments. Circular dichroism and conventional, 12C-Fourier transform infrared (FTIR) spectroscopy indicated the following α-helical levels for FP in 1-palmitoyl-2-oleoylphosphatidylglycerol (POPG) liposomes∼hexafluoroisopropanol (HFIP)>trifluoroethanol (TFE)>phosphate-buffered saline (PBS). 12C-FTIR spectra also showed disordered FP structures in these environments, along with substantial β-structures for FP in TFE or PBS. In further experiments designed to map secondary conformations to specific residues, isotope-enhanced FTIR spectroscopy was performed using a suite of FP peptides labeled with 13C-carbonyl at multiple sites. Combining these 13C-enhanced FTIR results with molecular simulations indicated the following model for FP in HFIP: α-helix (residues 3-16) and random and β-structures (residues 1-2 and residues 17-23). Additional 13C-FTIR analysis indicated a similar conformation for FP in POPG at low peptide loading, except that the α-helix extends over residues 1-16. At low peptide loading in either human erythrocyte ghosts or lipid extracts from ghosts, 13C-FTIR spectroscopy showed α-helical conformations for the central core of FP (residues 5-15); on the other hand, at high peptide loading in ghosts or lipid extracts, the central core of FP assumed an antiparallel β-structure. FP at low loading in ghosts probably inserts deeply as an α-helix into the hydrophobic membrane bilayer, while at higher loading FP primarily associates with ghosts as an aqueous-accessible, β-sheet. In future studies, 13C-FTIR spectroscopy may yield residue-specific conformations for other membrane-bound proteins or peptides, which have been difficult to analyze with more standard methodologies.  相似文献   

3.
One proposed toxic mechanism of Bacillus thuringiensis Cry δ-endotoxins involves pore formation in target membranes by the α4-α5 transmembrane hairpin constituting their pore-forming domain. Here, nine selected charged and uncharged polar residues in the pore-lining α4 of the Cry4Aa mosquito-active toxin were substituted with Ala. All mutant toxins, i.e., D169A, R171A, Q173A, H178A, Y179A, H180A, Q182A, N183A and E187A, were over-expressed in Escherichia coli as 130-kDa protoxin inclusions at levels comparable to the wild-type toxin. Bioassays against Aedes aegypti larvae revealed that only H178A and H180A mutants displayed a drastic reduction in biotoxicity, albeit almost complete insolubility observed for H178A, but not for H180A inclusions. Further mutagenic analysis showed that replacements of His180 with charged (Arg, Lys, Asp, Glu), small uncharged polar (Ser, Cys) or small non-polar (Gly, Val) residues severely impaired the biotoxicity, unlike substitutions with relatively large uncharged (Asn, Gln, Leu) or aromatic (Phe, Tyr, Trp) residues. Similar to the trypsin-activated wild-type toxin, both bio-active and -inactive H180 mutants were still capable of releasing entrapped calcein from lipid vesicles and producing cation-selective channels with ~130-pS maximum conductance. Analysis of the Cry4Aa structure revealed the existence of a hydrophobic cavity near the critical His180 side-chain. Analysis of simulated structures revealed that His180-to-smaller residue conversions create a gap disrupting such cavity's hydrophobicity and hence structural arrangements of the α4-α5 hairpin. Altogether, our data disclose a critical involvement in Cry4Aa-biotoxicity of His180 exclusively present in the lumen-facing α4 for providing proper environment for the α4-α5 hairpin prior to membrane-inserted pore formation.  相似文献   

4.
The 10,000-nucleotide RNA genome of the Prague strain, subgroup B (PR-B) of Rous sarcoma virus, was found to contain 11.6 ± 0.5 residues of m6Ap by quantitative analysis of 32P-labeled virion RNA after complete RNAase digestion. Approximately ten of the m6Ap residues are located, without obvious clustering, in that region of the genome between 500 and 4000 nucleotides from the 3′ poly(A) end. The src gene, which is required for transformation, and part of the env gene, which codes for the major viral envelope glycoprotein, have previously been mapped in this region of the viral genome. A transformation-defective deletion mutant of PR-B Rous sarcoma virus, which lacks the src gene, has 7.0 ± 0.2 m6Ap residues per RNA subunit. This supports our mapping of a portion of the m6A residues in src and suggests that this methylation is specific to certain regions of the genome. The possible significance of this result for Rous sarcoma virus RNA processing and translation is discussed.  相似文献   

5.
A water-soluble glucan, AR-Glucan, from the roots of Angelica acutiloba was obtained homogeneous as determined by ultracentrifugal analysis, electrophoresis, and gel filtration. AR-Glucan was composed Of d-glucose, and its MW was estimated to be 13 500. Methylation analysis indicated that AR-Glucan contained 4-O- and 4,6-di-O-substituted glucosyl residues. 1H and 13C NMR data accorded with the results of methylation analysis, and the glycosidic linkages in AR-Glucan were shown to have the α-configuration. The results of β-amylase, α-amylase, and pullulanase treatments of AR-Glucan showed that it contained (1 → 4) linked α-d-glucosyl side chains of long chain length such as amylopectin. Thus, AR-Glucan is a (1 → 4) linked α-d-glucan to which are attached glucosyl side chains at O-6 of the glucosyl residues of the main chain.  相似文献   

6.
Raman spectra in the region 1000–150 cm?1 were measured for copoly(D ,L -alanines) with the D -residue contents, 3, 7, 10, and 20%, and compared with the spectrum of the α-helical poly-L -alanine. The 532- and 378-cm?1 peaks were assigned to the L -residues with a right-handed α-helix-like local conformation or to the D -residues with a left-handed α-helix-like local conformation. From the intensity of the latter peak the contents of these local conformations were estimated as a function of the D -residue contents for the copolymers. The 264-cm?1 peak, which has been assigned to the breathing vibration of the α-helical poly-L -alanine, shows a marked decrease in its intensity upon the introduction of the D residues. This result suggests that the overall deformation vibration of the α-helix arises from rather long sequences of the L - and D -alanine residues with the α-helical conformation and that the intensity of this vibration depends on the content of these sequences in the copolymers.  相似文献   

7.
The l-α-Asp residues in peptides or proteins are prone to undergo nonenzymatic reactions to form l-β-Asp, d-α-Asp, and d-β-Asp residues via a succinimide five-membered ring intermediate. From these three types of isomerized aspartic acid residues, particularly d-β-Asp has been widely detected in aging tissue. In this study, we computationally investigated the cyclization of α- and β-Asp residues to form succinimide with dihydrogen phosphate ion as a catalyst (H2PO4). We performed the study using B3LYP/6-31 + G(d,p) density functional theory calculations. The comparison of the activation barriers of both residues is discussed. All the calculations were performed using model compounds in which an α/β-Asp-Gly sequence is capped with acetyl and methylamino groups on the N- and C-termini, respectively. Moreover, H2PO4 catalyzes all the steps of the succinimide formation (cyclization-dehydration) acting as a proton-relay mediator. The calculated activation energy barriers for succinimide formation of α- and β-Asp residues are 26.9 and 26.0 kcal mol 1, respectively. Although it was experimentally confirmed that β-Asp has higher stability than α-Asp, there was no clear difference between the activation barriers. Therefore, the higher stability of β-Asp residue than α-Asp residue may be caused by an entropic effect associated with the succinimide formation.  相似文献   

8.
The C-peptide of ribonuclease A (residues 1 to 13) is obtained by cyanogen bromide cleavage at Met13, which converts methionine to a mixture of homoserine lactone (giving C-peptide lactone) and homoserine carboxylate (giving C-peptide carboxylate). The helix-forming properties of C-peptide lactone have been reported. The helix is formed intramolecularly in aqueous solution, is stabilized at low temperatures (0 to 20 °C) and also by a pH-dependent interaction between sidechains. The C-peptide lactone helix is about 1000-fold more stable than expected from “host-guest” data for helix formation in synthetic polypeptides.Here we report the failure of C-peptide carboxylate to form an α-helix in comparable conditions. Formation of a salt-bridge between the α-COO? group and the imidazolium ring of His12+ appears to be responsible for the suppression of helix formation. The presence of the Hse13-COO? … His12+ salt-bridge in C-peptide carboxylate is shown by 1H nuclear magnetic resonance titration of the amide proton resonances of His12 and Hse13, and is expected from model peptide studies. The most probable reason why C-peptide carboxylate does not form an α-helix is that the Hse13-COO? … His12+ salt-bridge competes successfully with a helix stabilizing salt-bridge (Glu9? … His12+).S-peptide (residues 1 to 20 of ribonuclease A) does form an α-helix with properties similar to those of the C-peptide (lactone) helix, which shows that the lactone ring of C-peptide lactone is not needed for helix formation.These results support the hypothesis that a Glu9? … His12+ salt-bridge stabilizes the C-peptide (lactone) helix, and they show that specific interactions between side-chains can be important in preventing as well as in promoting α-helix formation.  相似文献   

9.
Recent publications defined requirements for inter-subunit contacts in a benzodiazepine-sensitive GABAA receptor (GABAARα1β3γ2). There is strong evidence that the heteropentameric receptor contains two α1, two β3, and one γ2 subunit. However, the available data do not distinguish two possibilities: When viewed clockwise from an extracellular viewpoint the subunits could be arranged in either γ2β3α1β3α1 or γ2α1β3α1β3 configurations. Here we use molecular modeling to thread the relevant GABAAR subunit sequences onto a template of homopentameric subunits in the crystal structure of the acetylcholine binding protein (AChBP). The GABAA sequences are known to have 15-18% identity with the acetylcholine binding protein and nearly all residues that are conserved within the nAChR family are present in AChBP. The correctly aligned GABAA sequences were threaded onto the AChBP template in the γ2β3α1β3α1 or γ2α1β3α1β3  arrangements. Only the γ2α1β3α1β3 arrangement satisfied three known criteria: (1) α1 His102 binds at the γ2 subunit interface in proximity to γ2 residues Thr142, Phe77, and Met130; (2) α1 residues 80-100 bind near γ2 residues 91-104; and (3) α1 residues 58-67 bind near the β3 subunit interface. In addition to predicting the most likely inter-subunit arrangement, the model predicts which residues form the GABA and benzodiazepine binding sites.  相似文献   

10.
α-Synuclein is the major component of the intracellular Lewy body inclusions present in Parkinson disease (PD) neurons. PD involves the loss of dopaminergic neurons in the substantia nigra and the subsequent depletion of dopamine (DA) in the striatum. DA can inhibit α-synuclein fibrillization in vitro and promote α-synuclein aggregation into soluble oligomers. We have studied the mechanism by which DA mediates α-synuclein aggregation into soluble oligomers. Reacting α-synuclein with DA increased the mass of α-synuclein by 64 Da. NMR showed that all four methionine residues were oxidized by DA, consistent with the addition of 64 Da. Substituting all four methionines to alanine significantly reduced the formation of DA-mediated soluble oligomers. The 125YEMPS129 motif in α-synuclein can modulate DA inhibition of α-synuclein fibrillization. However, α-synuclein ending before the 125YEMPS129 motif (residues 1–124) could still form soluble oligomers. The addition of exogenous synthetic YEMPS peptide inhibited the formation of soluble oligomers and resulted in the YEMPS peptide being oxidized. Therefore, the 125YEMPS129 acts as an antioxidant rather than interacting directly with DA. Our study defines methionine oxidation as the dominant mechanism by which DA generates soluble α-synuclein oligomers and highlights the potential role for oxidative stress in modulating α-synuclein aggregation.  相似文献   

11.
The human interferon (huIFN) σ-4 α25–62/α164–166 is a genetically engineered hybrid that consists of residues 5–62 of huIFN α2 and residues 64–166 of huIFN α1. This variant contains four cysteine residues at positions 29, 86, 99 and 139, but does not contain the cysteine at position 1 that is characteristic of naturally occurring huIFN α subtypes. This novel recombinant hybrid was purified fromEscherichia coli to greater than 95% homogeneity. The purification was based on ethanol extraction of a trichloroacetic acid precipitate and Matrex Gel Blue A chromatography followed by either a selective precipitation or DEAE-Sepharose chromatography. The purified protein that was treated with 2-mercaptoethanol exhibited two closely migrating bands on sodium dodecyl sulfate-polyacrylamide gel electrophoresis with apparent molecular weight values of 17 800 and 17 100, both of which exhibited antiviral activity. Electrophoresis performed without prior reduction with 2-mercaptoethanol indicated only a minor extent of intermolecular disulfide bonding. The purified protein exhibited a high specific antiviral activity of 7·107 units/mg when assayed on human fibroblast cells and, in distinction to the parental huIFN α2, it also demonstrated antiviral activity on murine L929 cells. The level of antiproliferative activity of huIFN δ-4 α25–62/α164–166 on various cell lines of different histological origin appeared to be more comparable to that of huIFN α1 than huIFN α2. The data suggest that huIFN δ-4 α25–62/α164–166 hybrid may be a useful tool for understanding huIFN structure-function relations.  相似文献   

12.
A. Darke  E. G. Finer 《Biopolymers》1975,14(3):441-455
1H, 2H, 13C, and 81Br nmr measurements of mixtures of poly-L -lysine hydrobromide with water have been carried out over a range of temperatures and water contents. When n (number of molecules of water per residue) ~13 at room temperature, a transition occurs from a gel to a liquid phase. The liquid phase contains polymer molecules that are flexible, but contain more intramolecular structure than the same molecules in trifluoracetic acid solution. The gel phase contains junction zones of hexagonally packed α-helices, linked by flexible regions of polypeptide chain. The α-helical residues impart to their associated water molecules a slight anisotropy of motion, which is dectable by 2H nmr. These residues bind up to about seven molecules of water each; the other six required to complete the gel–liquid transition space out the polymer molecules, allowing increased segmental motion of the residues in the flexible regions. This increased motion reduces the energy of the flexible regions and thus increases the proportion of residues in them (increasing the temperature has the same effect); the transition occurs when insufficient residues remain in the α-helical junction zones.  相似文献   

13.
The effects of Cu2+ binding and the utilization of different force fields when modeling the structural characteristics of α-syn12 peptide were investigated. To this end, we performed extensive temperature replica exchange molecular dynamics (T-REMD) simulations on Cu2+-bound and unbound α-syn12 peptide using the GROMOS 43A1, OPLS-AA, and AMBER03 force fields. Each replica was run for 300 ns. The structural characteristics of α-syn12 peptide were studied based on backbone dihedral angle distributions, free-energy surfaces obtained with different reaction coordinates, favored conformations, the formation of different Turn structures, and the solvent exposure of the hydrophobic residues. The findings show that AMBER03 prefers to sample helical structures for the unbound α-syn12 peptide and does not sample any β-hairpin structure for the Cu2+-bound α-syn12 peptide. In contrast, the central structure of the major conformational clusters for the Cu2+-bound and unbound α-syn12 peptide according to simulations performed using the GROMOS 43A1 and OPLS-AA force fields is a β-hairpin with Turn9-6. Cu2+ can also promote the formation of the β-hairpin and increase the solvent exposure of hydrophobic residues, which promotes the aggregation of α-syn12 peptide. This study can help us to understand the mechanisms through which Cu2+ participates in the fibrillation of α-syn12 peptide at the atomic level, which in turn represents a step towards elucidating the nosogenesis of Parkinson’s disease.
Figure
The representative structures of Cu2+-bound and unbound α-syn12 peptide using three different force fields  相似文献   

14.
Synthetic peptides have been used to investigate the site specificity of highly purified virus induced protein kinase, a recently discovered protein kinase isolated from cells infected with α-herpesviruses. The enzyme from cells infected with pseudorabies virus can catalyse the phosphorylation of both seryl and threonyl residues in peptides that contain several arginyl residues on the amino-terminal side of the target residue. At least two arginyl residues are required, and the best substrates examined contain four to six such residues. Virus induced protein kinase differs in site specificity from protein kinase C in being unable to phosphorylate peptides in which multiple arginyl residues are on the carboxyl-terminal side of the target residue, or to phosphorylate peptides in which the arginyl residues are replaced by ornithyl residues. Virus induced protein kinase from cells infected with herpes simples virus type I had similar substrate preferences to virus induced protein kinase from cells infected with pseudorabies virus. Although virus induced protein kinase and the cyclic AMP-dependent protein kinase have several peptide substrates in common, their relative preferences for these (as indicated by Km values) were found to be very different.  相似文献   

15.
Summary

The two characterized crustacean pigment-dispersing hormones (α-PDH; β-PDH) are octadecapeptides which differ in primary structure at six positions. Assays for melanophore pigment-dispersing activity showed β-PDH to be 21-fold more potent than α-PDH. In an effort to explain the difference in potencies between the two PDHs, we synthesized and purified six analogs of α-PDH (Leu4?, Leu11?, Lys13?, Asn16?, Asp17?, and Glu3, Leu4? α-PDH) in which the amino acid residues of α-PDH were substituted with those of β-PDH. Four analogs (Leu11?, Lys13?, Asn16?, and Asp17? α-PDH) possessed melanophore-dispersing activity equivalent to α-PDH. Leu4? α-PDH and Glu3, Leu4? α-PDH were 2.4? and 4-fold more potent than α-PDH, respectively. Glu3-α-PDH was 3.3-fold more potent than α-PDH (Jorenby et al., 1987). These results suggest that the 21-fold increase in activity of β-PDH over α-PDH is due to an interactive effect of two or more substitutions rather than from the product of the effects brought about by individual substitutions.  相似文献   

16.
Arrestins rapidly bind phosphorylated activated forms of their cognate G protein-coupled receptors, thereby preventing G protein coupling and often switching signaling to other pathways. Amphipathic α-helix I (residues 100-111) has been implicated in receptor binding, but the mechanism of its action has not been determined yet. Here we show that several mutations in the helix itself and in adjacent hydrophobic residues in the body of the N-domain reduce arrestin1 binding to light-activated phosphorylated rhodopsin (P-Rh?). On the background of phosphorylation-independent mutants that bind with high affinity to both P-Rh? and light-activated unphosphorylated rhodopsin, these mutations reduce the stability of the arrestin complex with P-Rh?, but not with light-activated unphosphorylated rhodopsin. Using site-directed spin labeling, we found that the local structure around α-helix I changes upon binding to rhodopsin. However, the intramolecular distances between α-helix I and adjacent β-strand I (or the rest of the N-domain), measured using double electron-electron resonance, do not change, ruling out relocation of the helix due to receptor binding. Collectively, these data demonstrate that α-helix I plays an indirect role in receptor binding, likely keeping β-strand I, which carries several phosphate-binding residues, in a position favorable for its interaction with receptor-attached phosphates.  相似文献   

17.
A Na+-specific and Na+-stimulated active α-aminoisobutyric acid transport system was reconstituted from plasma membranes isolated from mouse fibroblast BALB/c 3T3 cells transformed by simian virus 40. The plasma membranes were treated with dimethylmaleic anhydride and then extracted with 2% cholate. The cholate-solubilized supernatant proteins were combined with exogenous phospholipids and eluted through a Sephadex G-50 column. This yielded reconstituted vesicles which in the presence of Na+ could actively transport α-aminoisobutyric acid as shown by the transient accumulation above the equilibrium level (overshoot). The overshoot was not obtained with other monovalent cations such as K+, Li+, and choline+. The electrochemical effect of the lipophilic anion, SCN?, led to greater α-aminoisobutyric acid uptake as compared to that observed with Cl? or SO42?. The Na+-stimulated transport of a-aminoisobutyric acid was a saturable process with an apparent Km of 2 mm. Studies of the inhibition of α-aminoisobutyric acid transport by other amino acids showed that methylaminoisobutyric acid [specifically transported by A system (alanine preferring)]had a pronounced inhibitory effect on a-aminoisobutyric acid uptake in contrast to the slight inhibitory effect produced by phenylalanine [primarily transported by L system (leucine preferring)]. The results show that the reconstituted vesicles, prepared from partially purified membrane proteins and exogenous phospholipids, regained the same important transport properties of native membrane vesicles, i.e., Na+-specific and Na+-stimulated concentrative α-aminoisobutyric acid uptake.  相似文献   

18.
A set of 18 synthetic uniform overlapping peptides spanning the entire extracellular part (residues 1–210) of the α-subunit of human acetylcholine receptor were studied for their binding activity of125I-labeled α-bungarotoxin and cobratoxin. A major toxin-binding region was found to reside within peptide α122–138. In addition, low-binding activities were obtained with peptides α34–49 and α194–210. It is concluded that the region within residues α122–138 constitutes a universal major toxin-binding region for acetylcholine receptor of various species.  相似文献   

19.
Nε-acetylation occurs on select lysine residues in α-crystallin of the human lens and alters its chaperone function. In this study, we investigated the effect of Nε-acetylation on advanced glycation end product (AGE) formation and consequences of the combined Nε-acetylation and AGE formation on the function of α-crystallin. Immunoprecipitation experiments revealed that Nε-acetylation of lysine residues and AGE formation co-occurs in both αA- and αB-crystallin of the human lens. Prior acetylation of αA- and αB-crystallin with acetic anhydride (Ac2O) before glycation with methylglyoxal (MGO) resulted in significant inhibition of the synthesis of two AGEs, hydroimidazolone (HI) and argpyrimidine. Similarly, synthesis of ascorbate-derived AGEs, pentosidine and Nε-carboxymethyl lysine (CML), was inhibited in both proteins by prior acetylation. In all cases, inhibition of AGE synthesis was positively related to the degree of acetylation. While prior acetylation further increased the chaperone activity of MGO-glycated αA-crystallin, it inhibited the loss of chaperone activity by ascorbate-glycation in both proteins. BioPORTER-mediated transfer of αA- and αB-crystallin into CHO cells resulted in significant protection against hyperthermia-induced apoptosis. This effect was enhanced in acetylated and MGO-modified αA- and αB-crystallin. Caspase-3 activity was reduced in α-crystallin transferred cells. Glycation of acetylated proteins with either MGO or ascorbate produced no significant change in the anti-apoptotic function. Collectively, these data demonstrate that lysine acetylation and AGE formation can occur concurrently in α-crystallin of human lens, and that lysine acetylation improves anti-apoptotic function of α-crystallin and prevents ascorbate-mediated loss of chaperone function.  相似文献   

20.
The N-terminal α-amino groups of β1-bungarotoxin (β1-Bgt) fromBungarus multicinctus venom were modified with trinitrobenzene sulfonic acid and the modified derivative was separated by high performance liquid chromatography. The trinitrophenylated (TNP) derivative contained two TNP groups at the α-amino groups of A chain and B chain and showed a marked decrease in enzymatic activity. Methionine residues at positions 6 and 8 of the A chain were oxidized with chloramine T or cleaved with cyanogen bromide to remove the N-terminal octapeptide. Oxidation of methionine residues and removal of the N-terminal octapeptide caused a precipitous decrease in enzymatic activity, whereas antigenicity remained unchanged. The presence of dihexanoyllecithin influenced the interaction between β1-Bgt and 8-antilinonaphthalene sulfonate (ANS) and revealed that β1-Bgt consists of two types of ANS-binding sites, one at the substrate binding site of the A chain and the other might be at the B chain. The modified derivatives still retained their affinity for Ca2+ and ANS, indicating that the N-terminal region is not involved in Ca2+ and substrate binding. A fluorescence study revealed that the α-amino group of the A chain was in the vicinity of substrate binding site and that the TNP α-amino groups were in proximity to Trp-19 of the A chain. In addition, the study showed that the N-terminal region is important for stabilizing the architectural environment of Trp-19. The results, together with the proposal that Trp-19 of the A chain is involved in substrate binding, suggest that the N-terminal region of the A chain plays a crucial role in maintaining a functional active site for β1-Bgt.  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