首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The structural flexibility of creatine kinase has been investigated with the covalent hydrophobic probe 2-[4′-(2″-iodoacetamido) phenyl] aminonaphthalene-6-sulfonic acid (IAANS) which reacts at vastly different rates with the two subunits to give a protein conjugate with fluorescence characteristic of reaction with a site in a hydrophobic cleft. Binding of purine nucleotides greatly enhances the probe fluorescence while pyrimidine nucleotides quench the fluorescence. Small anions bind to nucleotide-free creatine kinase near the location of the transferable phosphoryl group and quench both the IAANS fluorescence of modified creatine kinase and the tryptophan fluorescence of native creatine kinase. Chloride and nitrate non-competitively inhibit MgADP binding both with and without creatine. Fluorescence energy transfer demonstrates that the active sites of creatine kinase are well separated and become further apart after the nucleotide-induced conformational change.  相似文献   

2.
Loop L5 of kinesin is located near the ATPase site, in common with kinesins of various animal species. The rice plant-specific kinesin K16 also has a corresponding loop that is slightly shorter than that of mouse brain kinesin. The present study was designed to monitor conformational changes in loop L5 during ATP hydrolysis. For this purpose, we introduced one reactive cysteine into the L5 of rice kinesin and modified it with fluorescent probes. The cysteine in L5 was labeled with a fluorescent probe 2-(4'(iodoacetamide) anilino-naphthalene-6-sulfonic acid sodium salt) [IAANS]. IAANS was incorporated into L5 at an almost equimolar ratio in the absence of nucleotides. In contrast, the incorporated amount was reduced to 0.62 and 0.32 mol IAANS/mol motor domain in the presence of ATP and ADP, respectively. Upon nucleotide addition, the fluorescent intensity of IAANS incorporated into L5 was significantly reduced to 63% and 51% for ATP and ADP, respectively. These results suggest that L5 of rice kinesin significantly changes its conformation during ATP hydrolysis.  相似文献   

3.
The enzyme rhodanese (EC 2.8.1.1) could be reversibly refolded from urea in the presence of lauryl maltoside, beta-mercaptoethanol, and sodium thiosulfate. The unfolding/folding transition monitored using intrinsic fluorescence was resolved into two two-state transitions with midpoints at 3.6 and 5.0 M urea. The analysis assumed an intermediate with an emission maximum at 345 nm. Monitoring anisotropy of intrinsic fluorescence also gave an asymmetric transition. Activity followed one two-state transition centered at 3.6 M urea with no major change of secondary structure. Without thiosulfate or mercaptoethanol, there was one two-state transition at 5.0 M urea giving a species, in dilute urea, with a fluorescence maximum at 345 nm. This intermediate slowly relaxed toward 335 nm (t1/2 = 85 min) if only thiosulfate was absent but without regaining activity. Subsequent addition of thiosulfate led to a first-order recovery of activity (t1/2 = 75 min). Thus, a possible folding intermediate can be trapped which displays increased access of water and solutes to its fluorescent tryptophans. This intermediate conformer, which is flexible, has considerable secondary structure, is inactive, has exposed hydrophobic surfaces, and requires specific reducing conditions to regain full activity. Refolding probably involves an initial, rapid, hydrophobic collapse with acquisition of secondary structure to form the intermediate, followed by slower adjustment to the native global conformation. Final reactivation requires reduction localized at the active site.  相似文献   

4.
The enzyme rhodanese was investigated for the conformational transition associated with its urea unfolding. When rhodanese was treated with 0 or 3 M urea, the activity was not significantly affected. 4.25 M urea treatment led to a time-dependent loss of activity in 60 min. Rhodanese was completely inactivated within 2 min in 6 M urea. The 1,1'-bi(4-anilino)naphthalene-5,5'-disulfonic acid fluorescence intensity was not significantly increased during 0, 3, and 6 M urea equilibrations, and the fluorescence was dramatically increased with 4.25 M urea, indicating that hydrophobic surfaces are exposed. After 0 and 3 M urea equilibration, rhodanese was not significantly proteolyzed with trypsin. Treatment with 4.25 M urea led to simultaneous formation of major 12-, 15.9-, 17-, and 21.2-kDa fragments, followed by progressive emergence of smaller peptides. The N termini of the 17- and 21.2-kDa bands were those of intact rhodanese. The N terminus of the 15.9-kDa band starts at the end of the interdomain tether. The 12-kDa band begins with either residue 183 or residue 187. The size and sequence information suggest that the 17- and 15.9-kDa bands correspond to the two domains. The 21.2- and 12-kDa bands appear to be generated through one-site tryptic cleavage. It is concluded that urea disrupts interaction between the two domains, increasing the accessibility of the interdomain tether that can be digested by trypsin. The released domains have increased proteolytic susceptibility and produce smaller peptides, which may represent subdomains of rhodanese.  相似文献   

5.
Aspergillus niger produces multiple forms of polygalacturonases with molecular masses ranging from 30 to 60 kDa. The high molecular weight polygalacturonase (61 ± 2 kDa) from A. niger possesses a pH optimum of 4.3 and a pI of 3.9. The enzyme exhibited high sensitivity, both in terms of activity and structure, in the pH range of 4.3–7.0. The enzyme was irreversibly inactivated at pH 7.0. The enzyme is predominantly rich in parallel β structure. There is unfolding of the enzyme molecule between 4.3 and 7.0 resulting in irreversible loss of secondary and tertiary structure with the exposure of hydrophobic surfaces. ANS binding measurements, intrinsic fluorescence and acrylamide quenching measurements have confirmed the unfolding and exposure of hydrophobic surfaces. The midpoint of pH transition for both activity and secondary structure is 6.2 ± 0.1. The pH-induced changes of polygalacturonase confirm the role of histidine residues in structure and activity of the enzyme. The irreversible nature of inactivation is due to the unfolding induced exposure of hydrophobic surfaces leading to association/aggregation of the molecule. Size exclusion chromatography measurements have established the association of enzyme at higher pH. Urea induced unfolding measurements at pH 4.3 and 7.0 have confirmed the loss in stability as we approach neutral pH.  相似文献   

6.
The conformations of sulfur-free and sulfur-containing rhodanese were followed with and without the detergent lauryl maltoside after guanidinium chloride (GdmCl) addition to 5 M to study the apparent irreversibility of denaturation. Without lauryl maltoside, sulfur-containing rhodanese denatured in a transition giving, at approximately 2.3 M GdmCl, 50% of the total denaturation induced change observed by activity, CD, or intrinsic fluorescence. Sulfur-free rhodanese gave more complex behavior by intrinsic fluorescence and CD. CD showed loss of secondary structure in a broad, complex, and apparently biphasic transition extending from 0.5 to 3 M GdmCl. The interpretation of the transition was complicated by time-dependent aggregation due to noncovalent interactions. Results with the apolar fluorescence probe 2-anilinonaphthalene-8-sulfonic acid, implicated apolar exposure in aggregation. Sulfhydryl reactivity indicated that low GdmCl concentrations induced intermediates affecting the active site conformation. Lauryl maltoside prevented aggregation with no effect on activity or any conformational parameter of native enzyme. Transitions induced by GdmCl were still observed and consistent with several phases. Even in lauryl maltoside, an increase in apolar exposure was detected by 2-anilinonaphthalene-8-sulfonic acid, and by protein adsorption to octyl-Sepharose well below the major unfolding transitions. These results are interpreted with a model in which apolar interdomain interactions are disrupted, thereby increasing active site accessibility, before the intradomain interactions.  相似文献   

7.
The effects of high hydrostatic pressure (HHP) and urea on conformational transitions of human alpha-thrombin structure were studied by fluorescence spectroscopy and by measuring the catalytic activity of the enzyme. Treatment of thrombin with urea produced a progressive red shift in the center of mass of the intrinsic fluorescence emission spectrum, with a maximum displacement of 650 cm(-1). HHP (270 MPa) shifted the centre of mass by only 370 cm(-1). HHP combined with a subdenaturing urea concentration (1.5 m) displaced the centre of mass by approximately 750 cm(-1). The binding of the fluorescent probe bis(8-anilinonaphthalene-1-sulfonate) to thrombin was increased by 1.8-, 4.0-, and 2.7-fold after treatment with high urea concentration, HHP or HHP combined with urea, respectively, thus suggesting that all treatments convert the enzyme to partially folded intermediates with exposed hydrophobic regions. On the other hand, treatment of thrombin with urea (but not HHP) combined with dithiothreitol progressively displaced the fluorescent probe, thus suggesting that this condition converts the enzyme to a completely unfolded state. Urea and HHP also led to different conformations when changes in the thrombin catalytic site environment were assessed using the fluorescence emission of fluorescein-d-Phe-Pro-Arg-cloromethylketone-alpha-thrombin: addition of urea up to 2 m gradually decreased the fluorescence emission of the probe to 65% of the initial intensity, whereas HHP caused a progressive increase in fluorescence. Hydrolysis of the synthetic substrate S-2238 was enhanced (35%) in 2 m urea and gradually abolished at higher concentrations, while HHP (270 MPa) inhibited the enzyme's catalytic activity by 45% and abolished it when 1.5 m urea was also present. Altogether, analysis of urea and HHP effects on thrombin structure and activity indicates the formation of dissimilar intermediate states during denaturation by these agents.  相似文献   

8.
Reversible folding of rhodanese. Presence of intermediate(s) at equilibrium   总被引:4,自引:0,他引:4  
For the first time completely reversible unfolding was achieved for guanidinium chloride-denatured rhodanese using a systematically defined protocol. These conditions included beta-mercaptoethanol, lauryl maltoside, and sodium thiosulfate. All components were required to get more than the previous best reactivation with lauryl maltoside of 17% (Tandon, S., and Horowitz, P. (1986) J. Biol. Chem. 261, 15615-15681). Non-coincidental transition curves were obtained by monitoring different parameters including: (i) variation in the activity, (ii) shifts of the fluorescence wavelength maximum, and (iii) variation in ellipticity at 220 nm. The transition followed by the fluorescence wavelength maximum was asymmetric and resolvable into two separate transitions. A thermodynamic analysis was used to define the energetics of the two processes. Studies with the fluorescent "apolar" probe 1,8ANS are consistent with the appearance of organized hydrophobic surfaces following the first transition. Near UV CD measurements indicated that the first transition is associated with a loss of dyssymmetry around at least some of the tryptophans. Thus, the unfolding of rhodanese is complex, and there are detectable intermediate(s) during the process. These results suggest that reversible unfolding occurs in two discrete stages: 1) loss of tertiary interactions and activity, with retention of secondary structure, and 2) loss of secondary structure. The available x-ray structure suggests that the first transition can be associated with changes in the domain interactions, which may modulate the effectiveness of helix dipoles in lowering the pKa of the active site sulfhydryl.  相似文献   

9.
Molecular dynamics simulations of calcium-free calmodulin in solution   总被引:3,自引:0,他引:3  
A 4-ns molecular dynamics simulation of calcium-free calmodulin in solution has been performed, using Ewald summation to treat electrostatic interactions. Our simulation results were mostly consistent with solution experimental studies, including NMR, fluorescence and x-ray scattering. The secondary structures within the N- and C-terminal domains were conserved in the simulation, with trajectory structures similar to the NMR-derived model structure 1CFD. However, the relative orientations of the domains, for which there are no NMR restraints, differed in details between the simulation and the 1CFD model. The most interesting information provided by the simulations is that the dynamics of calcium-free calmodulin in solution is dominated by slow rigid body reorientations of the domains. The interdomain distance fluctuated between 29 and 39 A, and interdomain orientation angle, defined as the pseudo-dihedral formed by the four calcium binding sites, varied between -2 degrees and 108 degrees. Similarly, the domain linker region also exhibited significant fluctuations, with its length varying in the 34-45 A range and its bend angle in the 10-100 degrees range. The simulations are in accord with fluorescence results suggesting that calcium-free calmodulin is more compact and more flexible than the calcium activated form. Surprisingly, quite similar solvent accessibilities of the hydrophobic patches were seen in the calcium-free trajectory described in this work and previously generated calcium-loaded calmodulin simulations. Thus, our simulations suggest a reexamination of the standard model of the structural change of calmodulin upon calcium binding, involving exposure of the hydrophobic patches to solvent.  相似文献   

10.
Abstract

A 4-ns molecular dynamics simulation of calcium-free calmodulin in solution has been performed, using Ewald summation to treat electrostatic interactions. Our simulation results were mostly consistent with solution experimental studies, including NMR, fluorescence and x-ray scattering. The secondary structures within the N- and C-terminal domains were conserved in the simulation, with trajectory structures similar to the NMR-derived model structure 1CFD. However, the relative orientations of the domains, for which there are no NMR restraints, differed in details between the simulation and the 1CFD model. The most interesting information provided by the simulations is that the dynamics of calcium-free calmod- ulin in solution is dominated by slow rigid body reorientations of the domains. The interdomain distance fluctuated between 29 and 39 Å, and interdomain orientation angle, defined as the pseudo-dihedral formed by the four calcium binding sites, varied between ?2° and 108°. Similarly, the domain linker region also exhibited significant fluctuations, with its length varying in the 34–45 Å range and its bend angle in the 10–100° range. The simulations are in accord with fluorescence results suggesting that calcium-free calmodulin is more compact and more flexible than the calcium activated form. Surprisingly, quite similar solvent accessibilities of the hydrophobic patches were seen in the calcium-free trajectory described in this work and previously generated calcium-loaded calmodulin simulations. Thus, our simulations suggest a reexamination of the standard model of the structural change of calmodulin upon calcium binding, involving exposure of the hydrophobic patches to solvent.  相似文献   

11.
Fluorescence investigations of calmodulin hydrophobic sites   总被引:3,自引:0,他引:3  
Calmodulin activation of target enzymes depends on the interaction between calmodulin hydrophobic regions and some enzyme areas. The Ca2+ induced exposure of calmodulin hydrophobic sites was studied by means of 2-p-toluidinylnaphthalene-6-sulfonate, a fluorescent probe. Scatchard and Job plots showed that the calmodulin-Ca42+ complex bound two molecules of this hydrophobic probe, with KD congruent to 1.4 X 10(-4) M. These sites are not totally exposed until calmodulin has bound four Ca2+ per molecule, so the conformational change is not over before the four specific Ca2+ - binding sites are saturated with Ca2+.  相似文献   

12.
The effect of urea on glucose dehydrogenase from Bacillus megaterium has been studied by following changes in enzymatic activity, conformation and state of aggregation. It was found that the denaturation process involves several transitions. At very low urea concentrations (below 0.5 M), where the enzyme is fully active and tetrameric, there is a conformational change as monitored by an increase in intensity of the tryptophan fluorescence and a maximum exposure of organized hydrophobic surfaces as reported by the fluorescence of 4,4'-dianilino-1,1'-binaphthyl-5.5'-disulfonic acid. At slightly higher urea concentrations (0.75-2 M), a major conformational transition occurs, as monitored by circular dichroism and fluorescence measurements, in which the enzyme activity is completely lost and is concomitant with the formation of interacting intermediates that lead to a highly aggregated state. Increasing urea concentrations cause a complete dissociation to lead first a partially and eventually the complete unfolded monomer. These phenomena are fully reversible by dilution of denaturant. It is concluded that after urea denaturation, the folding/assembly pathway of glucose dehydrogenase occurs with the formation of intermediate species in which transient higher aggregates appear to be involved.  相似文献   

13.
The structural and functional properties of arginine kinase (AK) in alkaline conditions in the absence or presence of salt have been investigated. The conformational changes of AK during alkaline unfolding and salt-induced folding at alkaline pH were monitored using intrinsic fluorescence emission, binding of the fluorescence probe 1-anilino-8-naphthalenesulfonate and circular dichroism. The results for the alkaline unfolded enzyme showed that much lower pH (11.0) was required to cause the complete loss of AK activity than was required to cause an obvious conformational change of the enzyme. Compared with the completely unfolded state in 5 M urea, the high pH denatured enzyme had some residual secondary and tertiary structure even at pH 13.0. Increasing the ionic strength by adding salt at pH 12.75 resulted in the formation of a relatively compact tertiary structure and a little new secondary structure with hydrophobic surface enhancement. These results indicate that the partially folded state formed under alkaline conditions may have similarities to the molten globule state which is compact, but it has a poorly defined tertiary structure and a native-like secondary structure.  相似文献   

14.
The multidomain structure of soybean LOX1 was examined over the pH range 1-12. Lipoxygenase-1 activity was reversible over broad pH range of 4-10 due to the reversibility of conformational states of the molecule. Below pH 4.0, due to collapse in hydrophobic interactions, the enzyme unfolded to an irreversible conformation with the properties of molten globule state with a mid point of transition at pH 2.4. This intermediate state lost iron irreversibly. In alkaline pH at 11.5, LOX1 underwent partial unfolding with the exposure of cysteine residues with subsequent oxidation of a pair of cysteine residues in the C-terminal domain and this intermediate showed some properties of molten globule state and retained 35% of activity. Beyond pH 12.0, the enzyme was completely inactivated irreversibly due to irreversible conformational changes. The pH-dependent urea-induced unfolding of LOX1 suggested that LOX1 was more stable at pH 7.0 and least stable at pH 9.0. Furthermore, the urea-induced unfolding of LOX1 indicated that the unfolding was biphasic due to pH-dependent domain interactions and involved sequential unfolding of domains. The loss of enzyme activity at pH 4. 0 and 7.0 occurred much earlier to unfolding of the C-domain at all pHs studied. The combination of urea-induced unfolding measurements and limited proteolysis experiments suggested that at pH 4.0, the domains in LOX1 were less interactive and existed as tightly folded units. Furthermore, these results confirmed the contribution of ionic interactions in the interdomain contacts.  相似文献   

15.
Urea-induced unfolding of lipoxygenase-1 (LOX1) at pH 7.0 was followed by enzyme activity, spectroscopic measurements, and limited proteolysis experiments. Complete unfolding of LOX1 in 9 M urea in the presence of thiol reducing or thiol modifying reagents was observed. The aggregation and oxidative reactions prevented the reversible unfolding of the molecule. The loss of enzyme activity was much earlier than the structural loss of the molecule during the course of unfolding, with the midpoint concentrations being 4.5 and 7.0 M for activity and spectroscopic measurements, respectively. The equilibrium unfolding transition could be adequately fitted to a three-state, two-step model (N left arrow over right arrow I left arrow over right arrow U) and the intermediate fraction was maximally populated at 6.3 M urea. The free energy change (DeltaG(H(2)O)) for the unfolding of native (N) to intermediate (I) was 14.2 +/- 0.28 kcal/mol and for the intermediate to the unfolded state (U) was 11.9 +/- 0.12 kcal/mol. The ANS binding measurements as a function of urea concentration indicated that the maximum binding of ANS was in 6.3 M urea due to the exposure of hydrophobic groups; this intermediate showed significant amount of tertiary structure and retained nearly 60% of secondary structure. The limited proteolysis measurements showed that the initiation of unfolding was from the C-terminal domain. Thus, the stable intermediate observed could be the C-terminal domain unfolded with exposed hydrophobic domain-domain interface. Limited proteolysis experiments during refolding process suggested that the intermediate refolded prior to completely unfolded LOX1. These results confirmed the role of cysteine residues and domain-domain interactions in the reversible unfolding of LOX1. This is the first report of the reversible unfolding of a very large monomeric, multi-domain protein, which also has a prosthetic group.  相似文献   

16.
The antagonist carazolol has been used as a fluorescent probe for the binding site of the beta-adrenergic receptor (beta AR). The fluorescence properties of carazolol are dominated by the emission of the carbazole group, with the fine structure of the spectrum, but not the quantum yield, sensitive to the environment of the probe. The fluorescence emission spectrum of the bound probe is consistent with an extremely hydrophobic environment in the binding site of the receptor. Binding of carazolol to the purified beta AR increases the polarization of the fluorophore. Exposure to collisional quenchers has demonstrated the bound carazolol to be completely inaccessible to the solvent. Furthermore, the fluorescence of bound carazolol is not quenched by exposure to sodium nitrite, a F?rster energy acceptor which has an R0 value of 11.7 A with carazolol. Thus, physical analysis of the binding site of the beta AR by carazolol fluorescence indicates that the antagonist binds to the beta AR in a rigid hydrophobic environment which is buried deep within the core of the protein.  相似文献   

17.
To investigate the mechanisms by which hydrostatic pressure inhibits (Na,K)-ATPase, we measured enzyme activity, as a function of pressure and temperature, of purified (Na,K)-ATPase from dog kidney and eel electroplax, and we monitored protein conformation, possible subunit interactions, and the fluidity of the membrane with fluorescent probes. The (Na,K)-ATPase and p-nitrophenylphosphatase activities were inhibited reversibly by pressures below 1.5 kilobars (eel enzyme) and 2.5 kilobars (dog kidney enzyme). Above these pressures, the enzymes were inactivated irreversibly. The plots of 1n(activity) versus pressure were curvilinear; this indicates that the reversible inhibition by pressure involves two or more rate-limiting steps. The calculated activation volumes varied with temperature and pressure and were larger for the (Na,K)-ATPase activity compared to the p-nitrophenylphosphatase activity. The fluorescence polarization of three hydrophobic probes decreased with increasing temperature (10-36 degrees C) and increased with increasing pressure (10(-3)-1.5 kilobars), reversibly, without any evidence of a lipid phase transition. Plots of enzyme activity versus fluorescence polarization of the lipid probes showed an inverse relationship; this indicates that enzyme activity was directly related to the fluidity of the membrane as measured by the lipid probes. Pressure had no effect on the fluorescence polarization of two cardiac glycoside probes nor on the efficiency of resonance energy transfer between either donor and acceptor cardiac glycosides specifically bound to the ouabain sites of different alpha-subunits, or tryptophan and the bound cardiac glycoside probe. These results suggest that dissociation of dimeric alpha-subunits is not related to the inhibition by pressure, and that the cardiac glycoside-complexed enzyme conformation is stabilized by pressure. It is concluded that increased pressure decreases the membrane fluidity which hinders conformational transitions associated with rate-limiting steps of the (Na,K)-ATPase reaction. It is proposed that ion-bound or -occluded forms of (Na,K)-ATPase are stabilized by pressure because they occupy a smaller volume.  相似文献   

18.
The presence of hydrophobic sites in fetuin, ovine submaxillary mucin and two homogeneous canine tracheal mucins was established by fluorescence probe techniques. The interaction between the above-mentioned glycoproteins and two hydrophobic fluorescent compounds, sodium mansate and mansylphenylalanine, was accompanied by an enhancement in fluorescence and a shift of the fluorescence maxima to shorter wavelengths. The introduction of a phenylalanine residue to the mansyl group enhanced the binding affinity of the probe for the hydrophobic sites of these glycoproteins as evidenced by lower values for the dissociation constants. The high molecular weight (581 600) tracheal mucin, which had the highest carbohydrate content (80%) of all the glycoproteins investigated, exhibited the highest fluorescence enhancement and the largest number of binding sites for these fluorescent probes.  相似文献   

19.
The coenzyme-linked fluorescence of aromatic-L-amino-acid decarboxylase decays non-exponentially. The decay of both native and NaBH4 reduced samples can only be fitted by two exponentials each roughly accounting for about half of the total fluorescence. Denaturation of the reduced protein with 8 M urea makes the fluorescence decay mono-exponential, like that observed for the reference compound pyridoxamine-5-phosphate. An extra pyridoxyl moiety can be bound to the enzyme after incubation with excess pyridoxal phosphate and reduction with NaBH4. This sample is almost twice as fluorescent and shows also two lifetimes. After denaturation only one fluorescence lifetime is observed. The presence of two non-equivalent pyridoxal sites in the native enzyme can be postulated. The heterogeneous decay behaviour of the pyridoxyl moiety in the enzyme together with the variability of lifetime shown, makes this fluorophore an even more interesting fluorescent probe for proteins.  相似文献   

20.
Chen X  Wolfgang DE  Sampson NS 《Biochemistry》2000,39(44):13383-13389
To elucidate the cholesterol oxidase-membrane bilayer interaction, a cysteine was introduced into the active site lid at position-81 using the Brevibacterium enzyme. To eliminate the possibility of labeling native cysteine, the single cysteine in the wild-type enzyme was mutated to a serine without any change in activity. The loop-cysteine mutant was then labeled with acrylodan, an environment-sensitive fluorescence probe. The fluorescence increased and blue-shifted upon binding to lipid vesicles, consistent with a change into a more hydrophobic, i.e., lipid, environment. This acrylodan-labeled cholesterol oxidase was used to explore the pH, ionic strength, and headgroup dependence of binding. Between pH 6 and 10, there was no significant change in binding affinity. Incorporation of anionic lipids (phosphatidylserine) into the vesicles did not increase the binding affinity nor did altering the ionic strength. These experiments suggested that the interactions are primarily driven by hydrophobic effects not ionic effects. Using vesicles doped with either 5-doxyl phosphatidylcholine, 10-doxyl phosphatidylcholine, or phosphatidyl-tempocholine, quenching of acrylodan fluorescence was observed upon binding. Using the parallax method of London [Chattopadhyay, A., and London, E. (1987) Biochemistry 26, 39-45], the acrylodan ring is calculated to be 8.1 +/- 2.5 A from the center of the lipid bilayer. Modeling the acrylodan-cysteine residue as an extended chain suggests that the backbone of the loop does not penetrate into the lipid bilayer but interacts with the headgroups, i.e., the choline. These results demonstrate that cholesterol oxidase interacts directly with the lipid bilayer and sits on the surface of the membrane.  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