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 46 毫秒
1.
K H Johnson  D M Gray 《Biopolymers》1991,31(4):385-395
We analyzed the CD and uv absorption spectra of 5S RNA from Escherichia coli using the method developed in the preceding paper. The analysis of spectra of 5S RNA at 20 degrees C in 0.1M NaClO4, 2.5 mM Na+ (phosphate), pH 7.0, and 0.5 mM MgSO4 gave 7 +/- 3.6 A.U base pairs, 25 +/- 3.6 G.C base pairs, and 7.5 +/- 3.6 G.U base pairs. Estimates of nearest neighbor base pairs were more consistent with the Pieler-Erdmann and the Gewirth-Moore secondary structure models than with the Fox-Woese or the Burns-Luoma-Marshall models. We also examined the structure of 5S RNA as a function of temperature. The melting profile exhibited two transitions--one at about 35 degrees C and one above 50 degrees C. Our spectral data showed that helices I and II were stable during the first transition, and agreed with other data that helix III was the most likely helix to have melted. The results from this in-depth study of 5S RNA indicate that our method of analysis should be useful for studying the secondary structures of other small, unmodified RNAs.  相似文献   

2.
Secondary structure contents of tetanus neurotoxin have been estimated at neutral and acidic pH using circular dichroism (CD) and Fourier transform infrared (FT-IR) spectroscopy. An analysis of the far-ultraviolet CD spectra of the neurotoxin dissolved in 50 mM citrate-phosphate buffer (pH 7.0) revealed 20.0 +/- 2.1% alpha-helix, 50.5 +/- 2.1% beta-pleated sheets, no beta-turns, and 29.5% random coils, which is at considerable variance with results from an earlier detailed study of tetanus neurotoxin's secondary structures (J.P. Robinson, L.A. Holladay, J.H. Hash and D. Puett, J. Biol. Chem. 257 (1982) 407). However, the alpha-helix content estimated in this study is consistent with the earlier studies of Robinson et al. (J.P. Robinson, L.A. Holladay, J.B. Picklesimer and D. Puett, Mol. Cell. Biochem. 5 (1974) 147; J.P. Robinson, J.B. Picklesimer and D. Puett, J. Biol. Chem. 250 (1975) 7435) and with the study by Lazarovici et al. (P. Lazarovici, P. Yanai and E. Yavin, J. Biol. Chem. 262 (1986) 2645), although other secondary structural features do not agree with those of the previous studies. Secondary structure estimation from Fourier transform infrared spectra in both amide I and amide III frequency regions revealed 22-23% alpha-helix, 49-51% beta-pleated sheets and 27-28% random coils, indicating a good correlation with the secondary structure content estimated from CD analysis. Lowering of the pH of the neurotoxin to 5.5 or 4.0 did not result in any noticeable change in the overall secondary structures. However, there were significant pH-induced variations observed in the individual curve-fitted FT-IR bands in the amide III frequency region. For example, the 1302 cm-1 band (relative area, 4.2%) observed at pH 7.0 was shifted to 1297 cm-1 (relative area, 2.2%) at pH 5.5, and the relative area of the band at 1316-1317 cm-1 (alpha-helix) increased by approx. 40%. This study suggests that contrary to earlier reports, tetanus neurotoxin is a beta-pleated sheet dominated structure, and although lower pH does not change the overall contents of the secondary structures, significant conformational alterations are observed.  相似文献   

3.
Raman spectroscopy was employed to examine the secondary structure of the cAMP receptor protein (CRP). Spectra were obtained over the range 400-1900 cm-1 from solutions of CRP and from CRP-cAMP cocrystals. The spectra of CRP dissolved in 30 mM sodium phosphate and 0.15 M NaCl buffered at either pH 6 or pH 8 or dissolved in 0.15-0.2 M NaCl at protein concentrations of 5, 15, and 30 mg/mL were examined. Estimates of the secondary structure distribution were made by analyzing the amide I region of the spectra (1630-1700 cm-1). CRP secondary structure distributions were essentially the same in either pH and at all protein concentrations examined. The amide I analyses indicated a structural distribution of 44% alpha-helix, 28% beta-strand, 18% turn, and 10% undefined for CRP in solution. Raman spectra of CRP-cAMP cocrystals differed from the spectra of CRP in solution. Some differences were assigned to interfering background bands, whereas other spectral differences were attributed to changes in CRP structure. Differences in the amide III region and in the intensity at 935 cm-1 were consistent with alterations in secondary structure. Analysis of the amide I region of the CRP-cAMP cocrystal spectrum indicated a secondary structure distribution of 37% alpha-helix, 33% beta-strand, 17% turn, and 12% undefined. This result is in agreement with a published secondary structure distribution derived from X-ray analysis of CRP-cAMP cocrystals (37% alpha-helix and 36% beta-strand).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Infrared spectra of 2.5 mM solutions of beta-lactoglobulin B were recorded as a function of pH (from pH 2 to pH 13) and as a function of temperature (from -100 degrees C to +90 degrees C). An analysis of the pH- and temperature-induced changes in the secondary structure was performed based on changes in the conformation-sensitive amide I bands of beta-lactoglobulin. Whereas the total amount of beta-structure remains constant (56-59%) between pH 2 and pH 10, the proportions of the various beta-components do change. In particular, the dimerization of the monomeric protein, induced by raising the pH from 2 to 3 , leads to an increase in the intensity of the 1636 cm-1 band (associated with antiparallel beta-sheet), at the expense of the 1626 cm-1 band (associated with exposed beta-strands). Both the thermal and alkaline denaturation of beta-lactoglobulin occur in two distinct stages. Although the spectra (i.e., the structures) after complete thermal or alkaline denaturation are clearly different, the spectrum of the protein after the first stage of thermal denaturation (at about 60 degrees C) is the same as that after the first stage of alkaline denaturation (at pH 11), suggesting a common denaturation intermediate, which probably represents a crossover point in a complex potential hypersurface.  相似文献   

5.
Despite highly conserved active-site structures, members of the plant peroxidase superfamily exhibit a wide range of pH optima. Horseradish peroxidase isozyme C (HRPC) is an ideal peroxidase to investigate the structural determinants of pH stability and activity in superfamily members. Conflicting reports exist on the low-pH stability of HRPC and consequently the pKa of the catalytic distal histidine, which is neutral in active peroxidases. Towards resolving such discrepancies, acid-induced changes in HRPC from two popular commercial suppliers were systematically analyzed. Specifically, FTIR v(CO) and Soret-CD spectra of HRPC-CO and Soret absorption of ferric HRPC were recorded to probe time-dependent heme-pocket changes at pH 3.0 in phosphate, citrate and formate buffers, while the FTIR amide I' and far-UV CD spectra were examined to probe changes in secondary structure. Both HRPC-CO samples exhibited identical pH 7.0 v(CO) bands at 1934 and 1905 cm-1. In the pH 3.0 spectrum of sample A, the 1934 cm-1 band was dominant while a broad 1969 cm-1 band appeared in sample B. The intensity of this band, which is assigned to solvent-exposed heme, was greater in citrate than phosphate buffer, but in formate the 1934 cm-1 band remained dominant. Other spectral changes mirrored the v(CO) trends. No time- or buffer-anion-dependent conformation changes were detected in 1 mM CaCl2, revealing that buffer-anion-dependent leaching of stabilizing Ca2+ from HRPC occurs at pH 3.0. Since the N-glycans present in HRPC are of the flexible protein-surface-shielding type, the variation in low-pH conformational stability of the HRPC samples could be attributed to heterogeneous glycosylation, which was detected by SDS-PAGE. It is further proposed that glycosylation patterns may affect the low-pH stability of class II and III plant peroxidases.  相似文献   

6.
Amyloid-beta (Abeta) is the principal protein constituent of 'senile plaques' and is a suspected mediator in Alzheimer's disease (AD). Senile plaques also contain acetylcholinesterase (AChE; EC 3.1.1.7), which may have a role in promoting Alphabeta-toxicity. We have found that Alphabeta can affect AChE expression in a neuron-like line, the N1E.115 neuroblastoma cell. When 1 micro mAlphabeta 1-42 or 25-35 was added for 24 h to differentiating N1E.115 in culture, AChE activity increased 30-40% in adherent cells, and 100% or more in nonadherent cells. The changes in both tetrameric (G4) and monomeric (G1) AChE forms were comparable. Turnover studies indicated that the elevation of AChE activity reflected slowed AChE degradation rather than accelerated synthesis. With a similar time course, Alphabeta also increased the quantity of muscarinic receptors on the plasma membrane. Immunocytochemistry for a lysosomal membrane protein (LAMP-1) indicated no change in abundance or localization of lysosomes in treated cells. But decreased labeling by pH-sensitive fluorescent dye pointed to an impairment of lysosomal acidification. We consider that the alteration of AChE expression after Alphabeta-exposure could reflect lysosomal dysfunction, and might itself enhance Alphabeta-toxicity.  相似文献   

7.
The effect of temperature, pH, free [Mg(2+)], and ionic strength on the apparent equilibrium constant of arginine kinase (EC 2.7.3.3) was determined. At equilibrium, the apparent K' was defined as [see text] where each reactant represents the sum of all the ionic and metal complex species. The K' at pH 7.0, 1.0 mM free [Mg(2+)], and 0. 25 M ionic strength was 29.91 +/- 0.59, 33.44 +/- 0.46, 35.44 +/- 0. 71, 39.64 +/- 0.74, and 45.19 +/- 0.65 (n = 8) at 40, 33, 25, 15, and 5 degrees C, respectively. The standard apparent enthalpy (DeltaH degrees') is -8.19 kJ mol(-1), and the corresponding standard apparent entropy of the reaction (DeltaS degrees') is + 2. 2 J K(-1)mol(-1) in the direction of ATP formation at pH 7.0, free [Mg(2+)] =1.0 mM, ionic strength (I) =0.25 M at 25 degrees C. We further show that the magnitude of transformed Gibbs energy (DeltaG degrees ') of -8.89 kJ mol(-1) is mostly comprised of the enthalpy of the reaction, with 7.4% coming from the entropy TDeltaS degrees' term (+0.66 kJ mol(-1)). Our results are discussed in relation to the thermodynamic properties of its evolutionary successor, creatine kinase.  相似文献   

8.
R O Ryan  L D Hicks  C M Kay 《FEBS letters》1990,267(2):305-310
Hydrodynamic studies conducted in the analytical ultracentrifuge provided evidence for two populations of lipid transfer particle (LTP) when centrifuged in a buffer solution containing 10 mM Tris, pH 8.0/100 mM KCl. The apparent sedimentation coefficients of the two species was 23.3 S and 15.3 S. Upon changing the buffer pH to 7.0 or 5.7, two species of LTP were still present but the ratio of their relative abundance was altered. When the KCl concentration in the buffer was lowered to 50 mM the sample sedimented as a single species with an apparent S20,w of 22.9 S. In higher ionic strength buffers (10 mM succinate, pH 5.7/500 mM KCl) LTP sedimented with an apparent S20,w of 14.8 S. Further experiments revealed that these two forms are interconvertable as a function of buffer ionic strength. Given previous estimates of the molecular size of LTP we concluded that the slower sedimenting peak observed at high ionic strength represents monomeric LTP while the faster sedimenting material observed at low ionic strength is likely to be an aggregated state of LTP. This interpretation is supported by molecular weight determinations made by sedimentation equilibrium experiments conducted in 10 mM succinate, pH 5.7/500 mM KCl which yielded a particle Mr = 887,000. Circular dichroism spectra of monomeric LTP sample revealed 6% alpha-helix, 49% beta-sheet, 7% beta-turn and 35% random coil while aggregated LTP contained 13% alpha-helix, 66% beta-sheet and 21% random coil. The transfer activity of the two LTP forms was assayed and found to be the same indicating that either the state of LTP aggregation did not affect transfer activity or that upon exposure to a large excess of lipoprotein substrate disaggregation, without loss of activity, occurs.  相似文献   

9.
The primary structure of the human CstF-64 polyadenylation factor contains 12 nearly identical repeats of a consensus motif of five amino acid residues with the sequence MEAR(A/G). No known function has yet been ascribed to this motif; however, according to secondary structure prediction algorithms, it should form a helical structure in solution. To validate this theoretical prediction, we synthesized a 31 amino acid residue peptide (MEARA(6)) containing six repeats of the MEARA sequence and characterized its structure and stability by circular dichroism (CD) spectroscopy and differential scanning calorimetry (DSC). No effects of concentration on the CD or DSC properties of MEARA(6) were observed, indicating that the peptide is monomeric in solution at concentrations up to 2 mM. The far UV-CD spectra of MEARA(6) indicates that at a low temperature (1 degrees C) the MEARA(6) peptide has a relatively high helical content (76% at pH 2.0 and 65% at pH 7.0). The effects of pH and ionic strength on the CD spectrum of MEARA(6) suggest that a number of electrostatic interactions (e.g., i, i + 3 Arg/Glu ion pair, charge-dipole interactions) contribute to the stability of the helical structure in this peptide. DSC profiles show that the melting of MEARA(6) helix is accompanied by positive change in the enthalpy. To determine thermodynamic parameters of helix-coil transition from DSC profiles for this peptide, we developed a new, semiempirical procedure based on the calculated function for the heat capacity of the coiled state for a broad temperature range. The application of this approach to the partial molar heat capacity function for MEARA(6) provides the enthalpy change for helix formation calculated per amino acid residue as 3.5 kJ/mol.  相似文献   

10.
The effect of binding a high mobility group protein (HMG 17) on the stability and conformation of acetylated and control HeLa high molecular weight core chromatin (stripped of H1 and non-histone chromosomal proteins) was studied by circular dichroism and thermal-denaturation measurements. Previously it had been shown that conformational differences exist between native whole chromatin derived from butyrate-treated (acetylated) and control HeLa cells and that these conformational differences disappear by removing H1 and non-histone chromosomal proteins (Reczek, P.R., Weissman, D., Huvos, P.E. and Fasman, G.D. (1982) Biochemistry 21, 993–1002). The circular dichroism spectra and the thermal denaturation profiles of control and acetylated core chromatin were found to be similar. The circular dichroism properties of HMG 17 reconstituted highly acetylated and control core chromatin indicated the same alteration of chromatin structure at low ionic strength (1 mM sodium phosphate/0.25 mM EDTA, pH 7.0). The magnitudes of the decrease in ellipticity were proportional to the amount of HMG 17 bound and were found to be the same for both the acetylated and control core chromatin. Thermal denaturation profiles confirmed this change in structure induced by HMG 17 on control and highly acetylated core chromatin. The thermal denaturation profiles, which were resolved into three component transitions, exhibited a shifting of hyperchromicity from the lower melting transitions to the higher melting transitions, with a concomitant rise in Tm, on HMG 17 binding to both control and acetylated chromatin. The natures of the interactions of HMG 17 at higher ionic strength (50 mM NaCl/0.25 mM EDTA/1 mM sodium phosphate, pH 7.0) with acetylated and control core chromatin were slightly different, as measured by circular dichroism; however, a decrease in ellipticity was observed for both samples upon binding of HMG 17. These observations suggest that acetylation coupled with HMG 17 binding to core chromatin does not loosen chromatin structure. HMG 17 binding to control and acetylated core chromatin produces an overall stabilization and compaction of chromatin structure.  相似文献   

11.
为明确蛛丝蛋白NT结构域的生物学功能,首次研究了MiSp NT结构域在不同pH条件下的二聚化动力学及二级结构特性. 基于蛛丝蛋白MiSp全长编码基因,克隆并分别在BL21(DE3)和Rosetta-gami 2(DE3)中重组表达MiSp NT结构域: BL21(DE3)表达水平较高,约35 mg/L LB培养基,表达产物需经短暂超声和2 mol/L尿素促溶;Rosetta-gami 2(DE3)表达产物可溶性较好,但表达水平仅为BL21 (DE3) 一半左右. 融合蛋白经凝血酶介导的2次Ni-NTA纯化,纯度达到95%以上. Trp 荧光光谱分析表明,MiSpNT在pH为7.0左右时开始二聚化,pH5.7时二聚体为主要构象,pHT约为6.6. MiSp NT在pH 7.5,6.5和5.5条件下的CD图谱相似,均主要为α-helix,表明NT二聚化与二级结构水平的变化没有直接联系. 本研究为进一步探索蛛丝蛋白NT结构域的结构和功能以及成丝机理提供线索.  相似文献   

12.
Human airway lysozyme, purified from pathological bronchial secretions, is characterized by a specific activity 3-fold higher than that of hen egg-white lysozyme. The amino acid composition of human airway lysozyme is identical to that of other human lysozymes. The laser Raman spectra of human airway lysozyme and hen egg-white lysozyme in phosphate buffer solution (pH 7.2) are recorded in the range 300-1900 cm-1 at 488 nm. Drastic intensity differences are observed between the spectra analyzed in the ranges characteristic of the peptide backbone (e.g., beta-sheet; C alpha-C, C alpha-N), and of the aromatic side-chain vibrations (tyrosine, tryptophan). The deconvolution of the Raman amide I band gives secondary structures of 38% and 39% alpha-helix, 25% and 20% beta-sheet, and 37% and 41% undefined structure for the human and hen lysozymes, respectively.  相似文献   

13.
Molecular properties of the enzymic phytohemagglutinin of mung bean   总被引:3,自引:0,他引:3  
Mung bean seeds possess a tetrameric galactose-binding protein that displays two types of activities: (a) a hemagglutinin activity, and (b) an alpha-galactosidase activity. This protein can be reversibly converted by pH changes from a tetrameric form, which possesses both enzymic and hemagglutinin activities, to a monomeric form which possesses enzymic activity only. This observation suggests that the enzymic phytohemagglutinin is an aggregated form of a monomeric alpha-galactosidase. The tetrameric alpha-galactosidase has a pH optimum of about pH 7.0, while the monomeric form displays a pH optimum of 5.6. Circular dichroism difference spectra and inhibition studies suggest that aggregation induces conformational changes in the subunits sufficient to alter their enzymatic properties. The possibility of in vivo changes in subunit equilibria, when combined with the accompanying alterations in activity, provides a new concept worthy of consideration with respect to the physiological role of phytohemagglutinins.  相似文献   

14.
Two 3-hydroxybenzoate-inducible gentisate 1,2-dioxygenases were purified to homogeneity from Pseudomonas alcaligenes NCIB 9867 (P25X) and Pseudomonas putida NCIB 9869 (P35X), respectively. The estimated molecular mass of the purified P25X gentisate 1, 2-dioxygenase was 154 kDa, with a subunit mass of 39 kDa. Its structure is deduced to be a tetramer. The pI of this enzyme was established to be 4.8 to 5.0. The subunit mass of P35X gentisate 1, 2-dioxygenase was 41 kDa, and this enzyme was deduced to exist as a dimer, with a native molecular mass of about 82 kDa. The pI of P35X gentisate 1,2-dioxygenase was around 4.6 to 4.8. Both of the gentisate 1,2-dioxygenases exhibited typical saturation kinetics and had apparent Kms of 92 and 143 microM for gentisate, respectively. Broad substrate specificities were exhibited towards alkyl and halogenated gentisate analogs. Both enzymes had similar kinetic turnover characteristics for gentisate, with kcat/Km values of 44.08 x 10(4) s-1 M-1 for the P25X enzyme and 39.34 x 10(4) s-1 M-1 for the P35X enzyme. Higher kcat/Km values were expressed by both enzymes against the substituted gentisates. Significant differences were observed between the N-terminal sequences of the first 23 amino acid residues of the P25X and P35X gentisate 1,2-dioxygenases. The P25X gentisate 1,2-dioxygenase was stable between pH 5.0 and 7.5, with the optimal pH around 8.0. The P35X enzyme showed a pH stability range between 7.0 and 9.0, and the optimum pH was also 8.0. The optimal temperature for both P25X and P35X gentisate 1, 2-dioxygenases was around 50 degrees C, but the P35X enzyme was more heat stable than that from P25X. Both enzymes were strongly stimulated by 0.1 mM Fe2+ but were completely inhibited by the presence of 5 mM Cu2+. Partial inhibition of both enzymes was also observed with 5 mM Mn2+, Zn2+, and EDTA.  相似文献   

15.
The adsorption of a recombinant cutinase from Fusarium solani pisi onto the surface of 100 nm diameter poly(methyl methacrylate) (PMMA) latex particles was evaluated. Adsorption of cutinase is a fast process since more than 70% of protein molecules are adsorbed onto PMMA at time zero of experiment, irrespective of the tested conditions. A Langmuir-type model fitted both protein and enzyme activity isotherms at 25 degrees C. Gamma(max) increased from 1.1 to 1.7 mg m(-2) and U(max) increased from 365 to 982 U m(-2) as the pH was raised from 4.5 to 9.2, respectively. A decrease (up to 50%) in specific activity retention was observed at acidic pH values (pH 4.5 and 5.2) while almost no inactivation (eta(act) congruent with 87-94%) was detected upon adsorption at pH 7.0 and 9.2. Concomitantly, far-UV circular dichroism (CD) spectra evidenced a reduction in the alpha-helical content of adsorbed protein at acidic pH values while at neutral and alkaline pH the secondary structure of adsorbed cutinase was similar to that of native protein. Fluorescence anisotropy decays showed the release of some constraints to the local motion of the Trp69 upon protein adsorption at pH 8.0, probably due to the disruption of the tryptophan-alanine hydrogen bond when the tryptophan interacts with the PMMA surface. Structural data associated with activity measurements at pH 7.0 and 9.2 showed that cutinase adsorbs onto PMMA particles in an end-on orientation with active site exposed to solvent and full integrity of cutinase secondary structure. Hydrophobic interactions are likely the major contribution to the adsorption mechanism at neutral and alkaline pH values, and a higher amount of protein is adsorbed to PMMA particles with increasing temperature at pH 9.2. The maximum adsorption increased from 88 to 140 mg cutinase per g PMMA with temperature raising from 25 to 50 degrees C, at pH 9.2.  相似文献   

16.
Summary Alanine is formed under anaerobic conditions from glyceraldehyde and ammonium ion in aqueous solutions of sodium phosphate (pH 7.0) or imidazole-imidazolium chloride (pH 7.0) at ambient temperature. In 500 mM imidazole (pH 7.0), alanine synthesis from 10 mM glyceraldehyde and 15 mM ammonium ion is roughly 6 times more rapid in the presence of 10 mM 3-mercaptopropionate (0.62% yield at 60 days) than in its absence (0.10% yield at 60 days). Likewise, the formation of alanine in 500 mM sodium phosphate (pH 7.0) from 5 mM glyceraldehyde and 10 mM ammonium ion is more rapid in the presence of 10 mM N-acetylcysteine than in its absence. In this reaction with N-acetylcysteine, the ratio of the yield of alanine to the yield of lactate is fairly constant. The yield of alanine is about 4.5% that of lactate. Alanine synthesis in the presence of thiol probably proceeds via alanyl thioester, which is produced by rearrangement of the imine of the hemithioacetal of pyruvaldehyde, a product of glyceraldehyde dehydration. The significance of this reaction for molecular evolution is discussed.  相似文献   

17.
The influence of ATP and KCl on the quaternary structure and the enzymatic activity of D-glyceraldehyde-3-phosphate dehydrogenase from yeast(Y-GAPDH) has been studied by ultracentrifugation, gel chromatography and standard optical tests. In 0.1 M imidazole buffer pH 7.0, at low temperature (0°C) both complete deactivation and dissociation to dimers occur in the presence of 2 mM ATP and 0.1 M 2-mercaptoethanol. In 0.067 M phosphate buffer pH 7.0, containing 2 mM ATP and 1 mM dithiothreitol, only slight deactivation paralleled by minor changes of the native quaternary structure is observed. In this same buffer, increasing temperature leads to stabilization of both the tetrameric state and the catalytic activity of the enzyme. Deactivation and dissociation in the presence of 0.15 M KCl (in 0.2 M glycine buffer 9.1 ≥ pH ≥ 8.0) is a function of pH rather than electrolyte concentration; at neutral pH the enzyme is stabilized in its native state. Contrary to earlier assumptions in the literature, ATP and KCl under the above experimental conditions do not appear to play an important role in the in vivo regulation of Y-GAPDH.  相似文献   

18.
The mouse neuroblastoma cell line NB2A produces cellular and secreted acetylcholinesterase (AChE). After incubation of the cells for 4 days the ratio between AChE secreted into the medium and AChE in the cells was 1:1. The cell-associated enzyme could be subdivided into soluble AChE (25%) and detergent-soluble AChE (75%). Both extracts contained predominantly monomeric AChE (4.6S) and minor amounts of tetrameric AChE (10.6S), whereas the secreted AChE in the culture supernatant contained only the tetrameric form. All forms were partially purified by affinity chromatography. It could be demonstrated that the secretory and the intracellular soluble tetramers were hydrophilic, whereas the detergent-soluble tetramer was an amphiphilic protein. On the other hand the soluble and the detergent-soluble monomeric forms were amphiphilic and their activity depended on the presence of detergent. By digestion with proteinase K amphiphilic monomeric and tetrameric AChE could be converted to a hydrophilic form that no longer required detergent for catalytic activity. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of [3H]diisopropylfluorophosphate-labelled AChE gave one band at 64 kilodaltons (kD) under reducing conditions and two additional bands at 120 kD and 140 kD under nonreducing conditions.  相似文献   

19.
The secondary structure content of the recombinant human mu-opioid receptor (HuMOR) solubilized in trifluoroethanol (TFE) and in detergent micelles was investigated by circular dichroism. In both conditions, this G protein-coupled receptor adopts a characteristic alpha-helical structure, with minima at 208 and 222 nm as observed in the circular dichroism spectra. After deconvolution of spectra, the alpha-helix contents were estimated to be in the range of 50% in TFE and in sodium dodecyl sulfate at pH 6. These values are in accordance with the predicted secondary structure content determined for the mu-opioid receptor. A pH-dependent effect was observed on the secondary structure of the receptor solubilized in detergents, which demonstrates the essential role of ionic and hydrophobic interactions on the secondary structure. Circular dichroism spectra of EGFP-HuMOR, a fusion protein between the enhanced green fluorescent protein (EGFP) and the mu-opioid receptor, and EGFP solubilized in TFE were also analyzed as part of this study.  相似文献   

20.
Rat brain acetylcholinesterase (AChE, EC 3.1.1.7) consists of about 80% amphiphilic detergent-soluble (DS-) AChE and 20% hydrophilic salt-soluble (SS-) AChE. DS-AChE contains about 65% tetrameric, 20% dimeric and 10% monomeric, SS-AChE about 40% tetrameric and 60% monomeric forms. N-terminal sequencing of DS- and SS-AChE gave identical N-termini corresponding to the published cDNA sequence of the mature enzyme. The band pattern on SDS-gels is similar to that of AChE from human and bovine brain. SDS-PAGE of hydrophobically labeled DS-AChE revealed the presence of a disulfide bonded hydrophobic membrane anchor of about 20 kDa. Monoclonal antibodies (mAbs) recognizing the anchor-containing subunits of mammalian brain DS-AChE, crossreacted with rat brain DS-AChE but not with SS-AChE. DS- and SS-AChE also reacted with antibodies raised against a peptide comprising the last 10 amino acids of the sequence of bovine brain AChE. Our results led us to conclude that both DS- and SS-AChE from rat brain contain T-type catalytic subunits, and DS-AChE in addition a P-type hydrophobic anchor similar to other mammalian brain DS-AChE.  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