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 equilibrium unfolding of dimeric yeast glutathione reductase (GR) by guanidine hydrochloride (GdnHCl) was investigated. Unfolding was monitored by a variety of techniques, including intrinsic fluorescence emission, anisotropy and iodide quenching measurements, far-ultraviolet circular dichroism and thiol reactivity measurements. At 1 M GdnHCl, one thiol group of GR became accessible to modification with 5,5′-dithiobis-(2-nitrobenzoic) acid (DTNB), whereas no changes could be detected in the spectroscopic properties (fluorescence, circular dichroism) of the protein. Between 2 and 3 M GdnHCl, two partially folded intermediate states possessing flexible tertiary structures (revealed by fluorescence data) but compact secondary structures (as indicated by circular dichroism measurements) were identified. The quaternary structure of GR in the presence of GdnHCl was also investigated by size-exclusion liquid chromatography. These results indicated the presence of an expanded predissociated dimer at 2.5 M GdnHCl and partially folded monomers at 3 M GdnHCl. Taken together, these results suggest the existence of two molten-globule-like intermediate species (one dimeric and one monomeric) in the unfolding of GR. The results are discussed in terms of the mechanism of GR folding and dimerization.  相似文献   

2.
3.
4.
Rhodopsin (Rho) is a G protein-coupled receptor that initiates phototransduction in rod photoreceptors. High expression levels of Rho in the disc membranes of rod outer segments and the propensity of Rho to form higher oligomeric structures are evident from atomic force microscopy, transmission electron microscopy, and chemical cross-linking experiments. To explore the structural and functional properties of Rho in n-dodecyl-beta-maltoside, frequently used to purify heterologously expressed Rho and its mutants, we used gel filtration techniques, blue native gel electrophoresis, and functional assays. Here, we show that in micelles containing n-dodecyl-beta-maltoside at concentrations greater than 3 mM, Rho is present as a single monomer per detergent micelle. In contrast, in 12 mM 3-[(3-cholamidopropyl)dimethyl-ammonio]-1-propanesulfonate (CHAPS), micelles contain mostly dimeric Rho. The cognate G protein transducin (Gt) appears to have a preference for binding to the Rho dimer, and the complexes fall apart in the presence of guanosine 5'-3-O-(thio)triphosphate. Cross-linked Rho dimers release the chromophore at a slower rate than monomers and are much more resistant to heat denaturation. Both Rho(*) monomers and dimers are capable of activating Gt, and both of them are phosphorylated by Rho kinase. Rho expressed in HEK293 cells is also readily cross-linked by a bifunctional reagent. These studies provide an explanation of how detergent influences the oligomer-dimermonomer equilibrium of Rho and describe the functional characterization of Rho monomers and dimers in detergent.  相似文献   

5.
Full-consensus designed ankyrin repeat proteins were designed with one to six identical repeats flanked by capping repeats. These proteins express well in Escherichia coli as soluble monomers. Compared to our previously described designed ankyrin repeat protein library, randomized positions have now been fixed according to sequence statistics and structural considerations. Their stability increases with length and is even higher than that of library members, and those with more than three internal repeats are resistant to denaturation by boiling or guanidine hydrochloride. Full denaturation requires their heating in 5 M guanidine hydrochloride. The folding and unfolding kinetics of the proteins with up to three internal repeats were analyzed, as the other proteins could not be denatured. Folding is monophasic, with a rate that is nearly identical for all proteins (∼ 400-800 s− 1), indicating that essentially the same transition state must be crossed, possibly the folding of a single repeat. In contrast, the unfolding rate decreases by a factor of about 104 with increasing repeat number, directly reflecting thermodynamic stability in these extraordinarily slow denaturation rates. The number of unfolding phases also increases with repeat number. We analyzed the folding thermodynamics and kinetics both by classical two-state and three-state cooperative models and by an Ising-like model, where repeats are considered as two-state folding units that can be stabilized by interacting with their folded nearest neighbors. This Ising model globally describes both equilibrium and kinetic data very well and allows for a detailed explanation of the ankyrin repeat protein folding mechanism.  相似文献   

6.
Dissociation of alpha beta and alpha gamma dimers of human hemoglobins (Hb) A and F into monomers was studied by alpha chain exchange (Shaeffer, J. R., McDonald, M. J., Turci, S. M., Dinda, D. M., and Bunn, H. F. (1984) J. Biol. Chem. 259, 14544-14547). Unlabeled carbonmonoxy-Hb A was incubated with trace amounts of preparatively purified, native, 3H-alpha subunits in 10 mM sodium phosphate, pH 7.0, at 25 degrees C. At appropriate times, free alpha monomers were separated from Hb A tetramers by anion exchange high performance liquid chromatography. Transfer of radioactivity from the alpha chain pool into Hb A was measured, yielding a first order dimer dissociation rate constant, k2 = (3.2 +/- 0.3) X 10(-3) h-1. The Arrhenius plot of k2 was linear between 7 and 37 degrees C, yielding an enthalpy of activation of 23 kcal/alpha beta dimer. As the chloride concentration was raised from 0 to 0.2 M, the dissociation rate increased 3-fold; with higher salt concentrations, however, the rate gradually returned to baseline. This rate was not altered by raising the pH from 6.5 to 7.2, but as pH was further raised to 8.4, kappa 2 increased about 3-fold. Hb F, which has an increased stability at alkaline pH, dissociated into alpha and gamma monomers 3 times more slowly than Hb A. Moreover, the dimer-monomer dissociation of Hb F was characterized by a significantly reduced pH dependence. These results demonstrate that both alpha beta and alpha gamma dimers of Hb A and Hb F dissociate reversibly into monomers under physiologic conditions. The differential pH dependence for dimer dissociation between Hb A and Hb F suggests that specific amino acid replacement at the alpha 1 gamma 1 interface confers increased resistance to alkaline denaturation.  相似文献   

7.
Nine dihydroartemisinin acetal dimers (614) with diversely functionalized linker units were synthesized and tested for in vitro antiprotozoal, anticancer and antimicrobial activity. Compounds 6, 7 and 11 [IC50: 3.0–6.7 nM (D6) and 4.2–5.9 nM (W2)] were appreciably more active than artemisinin (1) [IC50: 32.9 nM (D6) and 42.5 nM (W2)] against the chloroquine-sensitive (D6) and chloroquine-resistant (W2) strains of the malaria parasite, Plasmodium falciparum. Compounds 10, 13 and 14 displayed enhanced anticancer activity in a number of cell lines compared to the control drug, doxorubicin. The antifungal activity of 7 and 12 against Cryptococcus neoformans (IC50: 0.16 and 0.55 μM, respectively) was also higher compared to the control drug, amphotericin B. The antileishmanial and antibacterial activities were marginal. A number of dihydroartemisinin acetal monomers (1517) and a trimer (18) were isolated as byproducts from the dimer synthesis and were also tested for biological activity.  相似文献   

8.
A combination of enzyme kinetics and 51V NMR spectroscopy was used to identify the species of vanadate that inhibits acid phosphatases. Monomeric vanadate was shown to inhibit wheat germ and potato acid phosphatases. At pH 5.5, the vanadate dimer inhibits the human prostatic acid phosphatase whereas at pH 7.0 it is the vanadate monomer that inhibits this enzyme. The pH-dependent shift in the affinity of the prostatic phosphatase for vanadate is presumably due to deprotonation of an amino acid side chain in or near the binding site resulting in a conformational change in the protein. pH may be a subtle effector of the insulin-like vanadate activity in biological systems and may explain some of the differences in selectivity observed with the protein phosphatases.  相似文献   

9.
A series of 2′-fluorinated adenosine compounds, dAfl, dAflp, pdAfl, dAfl-A, A-dAfl, and dAfl-dAfl, have been investigated by nmr spectroscopies. The 1H-, 19F-, and 31P-nmr data provide structural information from different parts of these moleucles. The pKa of the phosphate group of these two 2′-fluoro-2′-deoxyadenosine monophosphates was found to be the same as that of hte parent adenosine monophosphate. As for the pentose conformation, the 3E population is greatly increased as a result of the fluorine substitution at the C2′ position. However, the populations of conformers of gg (C4′-C5′) and g′g′ (C5′-O5′) and the average angle ?′(C3′-O3′) of the 2′-fluoro compounds remain unchanged as compared to the natural riboadenosine monomer and dimer (A-A). Thefefore, the backbone conformation of the 2′-fluoro-2′-deoxy-adenosine, its monophosphates and dimers, resembles that of RNA. The extent of base-base overlapping in these 2′-fluoro-2′-deoxy-adenosine-containing dimers is also found to be similar to or even greater than A-A. Thus, the conformations of these compounds can be considered as those in the RNA family. These fluorocompounds also serve as models for a careful study on the 19F-nmr in nucleic acid. The 19F chemical-shift values are sensitive to the environment of the fluorine atom such as ionic structure of the neighboring group(s) (phosphate of base), solvation, and ring-ruccent anisotropic effect from the base(s). Qualitatively, the change of the 19F chemical-shift values (up to 2 ppm) is much larger than that of 1H-nmr (up to 0.5 ppm) in the dimers. Using dAfl·poly(U), poly(dAfl)·poly(dAfl), and poly(dAfl)·poly(U) helix–coil transition as model systems, the linewidth of 19F in dAfl- residues reflects effectively the mobility of the unit in the nucleic acid complex as calibrated by uv data and by 1H-nmr. Therefore, application of 19F-nmr spectroscopy on fluorine-substituted nucleic acid can also be used to detect nucleic acid-nucleic acid interaction in complicated systems.  相似文献   

10.
Satumba WJ  Mossing MC 《Biochemistry》2002,41(48):14216-14224
Cro binds to operator sites in lambda DNA as a dimer. Dimerization of this small repressor protein is weak, however, and proline residues in the dimer interface suggest that folding and assembly of active repressors may be complex. Cro and selected variants have been studied by circular dichroism and fluorescence. Fluorescent probes include a unique tryptophan residue in the dimer interface and extrinsic resonance energy transfer probes that monitor dimerization. Both folding and unfolding are characterized by two distinct kinetic phases. Fast processes that are complete within the 5-10 ms dead time of stopped flow experiments account for the majority of the change in the CD signal and abrupt changes in both tryptophan fluorescence and energy transfer. The slow phases show all the hallmarks of proline isomerization. The rates of the slow phases are between 0.005 and 0.02 s(-1), are relatively independent of protein and denaturant concentration, display activation energies of 20 kcal/mol, and are accelerated by the peptidyl-prolyl isomerase SlyD. Although CD measurements indicate that more than 70% of the secondary structure is regained in the refolding burst phase, intermolecular fluorescence resonance energy transfer experiments indicate that less than 25% of these subunits are assembled into dimers. Full folding and dimerization requires isomerization of the non-native prolyl isomers over hundreds of seconds.  相似文献   

11.
The nitrosylation of two water-soluble iron-porphyrins, the anionic Fe(III)-meso-tetrakis(p-sulfonatophenyl)porphyrin (FeTPPS(4)) and the cationic Fe(III)-meso-tetrakis(4-N-methylpyridiniumyl)porphyrin (FeTMPyP), by the nitric oxide donor S-nitroso-N-acetylpenicillamine (SNAP) was studied using optical absorption spectroscopy. The influence of ionic and non-ionic micelles on rates of nitric oxide transfer was investigated. Initially, the effect of the micelles on the pH-dependent equilibrium between monomeric and micro-oxo-dimeric species of the iron-porphyrins was examined. It is not affected in micelle-porphyrin systems with electric charges identical in sign. The non-ionic micelles of polidocanol induce a small negative pK shift. In contrast, the dimerization equilibrium of porphyrins in oppositely charged micellar phases is displaced to lower pH by approximately 2 units. Nitric oxide binding to monomers and micro-oxo-dimers was examined at pH 5.0 and 8.0, respectively. Contrary to nitrosylation by NO gas, SNAP induces reductive nitrosylation. There was no observed NO-Fe(III)porphyrin intermediate. Nitrosylation rates were obtained and compared in aqueous buffer and in micellar systems. Monomers nitrosylate much faster than micro-oxo-dimers. Oppositely charged micelles prevent nitrosylation of the iron-porphyrins or considerably enhance nitrosylation times. Nitrosylation rates are comparable to transnitrosylation rates between several S-nitrosothiols and thiol-containing proteins, suggesting biological relevance for the process.  相似文献   

12.
1. In a recent study, we distinguished two classes of amphiphilic AChE3 dimers in Torpedo tissues: class I corresponds to glycolipid-anchored dimers and class II molecules are characterized by their lack of sensitivity to PI-PLC and PI-PLD, relatively small shift in sedimentation with detergent, and absence of aggregation without detergent. 2. In the present report, we analyze the amphiphlic or nonamphiphilic properties of globular AChE forms in T28 murine neural cells, rabbit muscle, and chicken muscle. The molecular forms were identified by sucrose gradient sedimentation in the presence and absence of detergent and analyzed by nondenaturing charge-shift electrophoresis. Some amphiphilic forms showed an abnormal electrophoretic migration in the absence of detergent, because of the retention of detergent micelles. 3. We show that the amphiphilic monomers (G1a) from these tissues, as well as the amphiphilic dimers (G2a) from chicken muscle, resemble the class II dimers of Torpedo AChE. We cannot exclude that these molecules possess a glycolipidic anchor but suggest that their hydrophobic domain may be of a different nature. We discuss their relationship with other cholinesterase molecular forms.  相似文献   

13.
We analyze the effect of different environmental conditions, sequence lengths and starting configurations on the folding and unfolding pathways of small peptides exhibiting beta turns. We use chignolin and a sequence of peptide G as examples. A variety of different analysis tools allows us to characterize the changes in the folding pathways. It is observed that different harmonic modes dominate not only for different conditions but also for different starting points. The modes remain essentially very similar but their relative importance varies. A detailed analysis from diverse viewpoints including the influence of the particular amino acid sequence, conformational aspects as well as the associated motions yields a global picture that is consistent with experimental evidence and theoretical studies published elsewhere. Patterns of modes that remain stable over a range of temperatures might serve as an additional diagnostic to identify conformations that have reliably adopted a native fold. This could aid in reconstructing the folding process of a complete protein by identifying conformationally determined regions.  相似文献   

14.
Okumura H 《Proteins》2012,80(10):2397-2416
A multibaric‐multithermal molecular dynamics (MD) simulation of a 10‐residue protein, chignolin, was performed. All‐atom model with the Amber parm99SB force field was used for the protein and the TIP3P model was used for the explicit water molecules. This MD simulation covered wide ranges of temperature between 260 and 560 K and pressure between 0.1 and 600 MPa and sampled many conformations without getting trapped in local‐minimum free‐energy states. Folding events to the native β‐hairpin structure occurred five times and unfolding events were observed four times. As the temperature and/or pressure increases, fraction of folded chignolin decreases. The partial molar enthalpy change ΔH and partial molar volume change ΔV of unfolding were calculated as ΔH = 24.1 ± 4.9 kJ/mol and ΔV = ?5.6 ± 1.5 cm3/mol, respectively. These values agree well with recent experimental results. Illustrating typical local‐minimum free‐energy conformations, folding and unfolding pathways were revealed. When chignolin unfolds from the β‐hairpin structure, only the C terminus or both C and N termini open first. It may undergo an α‐helix or 310‐helix structure and finally unfolds to the extended structure. Difference of the mechanism between temperature denaturation and pressure denaturation is also discussed. Temperature denaturation is caused by making the protein transferred to a higher entropy state and making it move around more with larger space. The reason for pressure denaturation is that water molecules approach the hydrophobic residues, which are not well hydrated at the folded state, and some hydrophobic contacts are broken. Proteins 2012;. © 2012 Wiley Periodicals, Inc.  相似文献   

15.
The influence of treatments for extracting non-receptor peripheral proteins on the oligomeric states of the acetylcholine receptor has been studied in receptor-rich membranes from Torpedo marmorata. Conventional alkaline treatment of non-alkylated membranes resulted in the extraction of peripheral proteins (30% of total membrane proteins). Concomitantly, partial conversion of the dimer into the monomer was observed in the absence of exogenous reduction. Alkaline extraction at high ionic strength resulted in a marked decrease in protein solubilization, and no conversion of the dimer to the monomer occurred. Alkaline treatment extracted only one half of the peripheral proteins (15% of total protein) from membranes previously alkylated with N-ethylmaleimide or iodoacetamide, or oxidized by sodium periodate. Conversion of dimer to monomer was totally prevented by these treatments. Similar results were obtained by treatment of the membranes with lithium 3,5-diiodosalicylate. The above effects of alkaline extraction on the acetylcholine receptor can be interpreted in the context of two mutually non-exclusive mechanisms: (a) some of the peripheral proteins may directly participate in the thiol-dependent receptor aggregational states. Their extraction destroys this dynamic control. (b) Extraction of peripheral proteins destabilizes the receptor and makes it more susceptible to inter or intramolecular sulfhydryl-disulfide exchange, leading to the endogenous reduction of a proportion of the dimers.  相似文献   

16.
ATP-binding cassette (ABC) proteins have two nucleotide-binding domains (NBDs) that work as dimers to bind and hydrolyze ATP, but the molecular mechanism of nucleotide hydrolysis is controversial. In particular, it is still unresolved whether hydrolysis leads to dissociation of the ATP-induced dimers or opening of the dimers, with the NBDs remaining in contact during the hydrolysis cycle. We studied a prototypical ABC NBD, the Methanococcus jannaschii MJ0796, using spectroscopic techniques. We show that fluorescence from a tryptophan positioned at the dimer interface and luminescence resonance energy transfer between probes reacted with single-cysteine mutants can be used to follow NBD association/dissociation in real time. The intermonomer distances calculated from luminescence resonance energy transfer data indicate that the NBDs separate completely following ATP hydrolysis, instead of opening. The results support ABC protein NBD association/dissociation, as opposed to constant-contact models.  相似文献   

17.
Redox reactions were carried out in aerobiosis and anaerobiosis between NAD(P) dimers or NAD(P)H and pyrroloquinoline quinone (PQQ) in different buffers. The buffer system and pH significantly affected the oxidation rates of nucleotides and the ESR signal intensity of the PQQ(*) radical formed in anaerobiosis by comproportion between the quinone and quinol forms. The relative reactivity of the four nucleotides toward PQQ was affected by pH and buffer nature. PQQ, which behaves as an electron shuttle from nucleotides to oxygen, was first converted to PQQH(2) and then rapidly reoxidized by oxygen, with formation of hydrogen peroxide. Both NAD(P) dimers and NAD(P)H consumed 1 mol of oxygen per mole of reacted molecule of pyridine nucleotide, yielding 1 or 2 mol of NAD(P)(+) from NAD(P)H or from NAD(P) dimers, respectively. Chelating agents such as EDTA and phytate strongly decreased the reaction rate and the PQQ(*) radical signal intensity. Kinetics carried out in the presence of metal ions showed instead an increased reaction rate in the order Ca(2+) > Mg(2+) > Na(+) > K(+). Spectrofluorimetric measurements of PQQ with increasing concentrations of Ca(2+) showed a fluorescence quenching and shift of the maximum emission toward lower wavelengths, while other metal ions showed minor effects, if any. Therefore, it is demonstrated that Ca(2+) binds to PQQ, probably forming a complex which is more reactive with both one-electron (NAD(P) dimers) or two-electron donors (NAD(P)H) in nonenzymic reactions. It is important to recall that Ca(2+) was already found to play active role in PQQ-containing enzymes.  相似文献   

18.
Most (90 to 95%) human plasma fibronectin (PFn) molecules exist as 450-kDa disulfide-rich dimers comprised of two major types of subunits (A, 220 kDa; B, 215 kDa) that are joined near the COOH terminus by two disulfide bonds. Smaller PFn species (Zone II; 190-235 kDa) consist mainly of monomers and/or a monomeric subunit joined covalently to a smaller peptide remnant presumably derived by proteolysis of a parent 450-kDa molecule. A relatively simple and selective method for preparing functionally active, partially reduced monomeric fibronectin subunits (PR-PFn) by limited and selective reduction of dimeric plasma fibronectin (PFn) has been developed. PR-PFn was prepared by incubating PFn in phosphate-buffered saline, pH 7.4, for 2 h at room temperature in the presence of 17 mM dithiothreitol (DTT). Following S-carboxymethylation or S-carboxyamidomethylation, the material was passed through a gelatin-Sepharose column and nonbinding material was discarded; gelatin-bound material was eluted using a 0 to 2 M KSCN gradient. Residual dimeric species (10-20%) could be separated from monomers in high yield by gel-sieving chromatography on a Sepharose 6B-Cl in the presence of a chaotropic salt, 0.3 M KSCN. Most new SH groups (74-81%) in that fraction of PR-PFn binding to gelatin were localized in proteolytic fragments containing the COOH terminus, thus suggesting that selective cleavage of the interchain disulfide bridges had taken place. The binding affinity of PR-PFn to gelatin- and fibrin-Sepharose was lower than that of dimeric PFn, but the same as that of Zone II PFn and other monomeric gelatin-binding proteolytic derivatives. PR-PFn also bound to heparin-Sepharose and promoted cell attachment and spreading. We conclude that PR-PFn monomers possess the same functional activities as those of the parent chains.  相似文献   

19.
T Kiefhaber  R Quaas  U Hahn  F X Schmid 《Biochemistry》1990,29(12):3061-3070
The slow refolding of ribonuclease T1 was investigated by different probes. Structural intermediates with secondary structure are formed early during refolding, as indicated by the rapid regain of a native-like circular dichroism spectrum in the amide region. This extensive structure formation is much faster than the slow steps of refolding, which are limited in rate by the reisomerization of incorrect proline isomers. The transient folding intermediates were also detected by unfolding assays, which make use of the reduced stability of folding intermediates relative to that of the native protein. The results of this and the preceding paper [Kiefhaber et al. (1990) Biochemistry (preceding paper in this issue)] were used to propose kinetic models for the unfolding and refolding of ribonuclease T1. The unfolding mechanism is based on the assumption that, after the structural unfolding step, the slow isomerizations of two X-Pro peptide bonds occur independently of each other in the denatured protein. At equilibrium a small amount of fast-folding species coexists with three slow-folding species: two with one incorrect proline isomer each and another, dominant species with both these prolines in the incorrect isomeric state. In the mechanism for refolding we assume that all slow-folding molecules can rapidly regain most of the secondary and part of the tertiary structure early in folding. Reisomerizations of incorrect proline peptide bonds constitute the slow, rate-limiting steps of refolding. A peculiar feature of the kinetic model for refolding is that the major unfolded species with two incorrect proline isomers can enter two alternative folding pathways, depending on which of the two reisomerizes first. The relative rates of reisomerization of the respective proline peptide bonds at the stage of the rapidly formed intermediate determine the choice of pathway. It is changed in the presence of prolyl isomerase, because this enzyme catalyzes these two isomerizations with different efficiency and consequently leads to a shift from the very slow to the intermediate refolding pathway.  相似文献   

20.
Single-protein force experiments have relied on a molecular fingerprint based on tethering multiple single-protein domains in a polyprotein chain. However, correlations between these domains remain an issue in interpreting force spectroscopy data, particularly during protein folding. Here we first show that force-clamp spectroscopy is a sensitive technique that provides a molecular fingerprint based on the unfolding step size of four single-monomer proteins. We then measure the force-dependent unfolding rate kinetics of ubiquitin and I27 monomers and find a good agreement with the data obtained for the respective polyproteins over a wide range of forces, in support of the Markovian hypothesis. Moreover, with a large statistical ensemble at a single force, we show that ubiquitin monomers also exhibit a broad distribution of unfolding times as a signature of disorder in the folded protein landscape. Furthermore, we readily capture the folding trajectories of monomers that exhibit the same stages in folding observed for polyproteins, thus eliminating the possibility of entropic masking by other unfolded modules in the chain or domain-domain interactions. On average, the time to reach the I27 folded length increases with increasing quenching force at a rate similar to that of the polyproteins. Force-clamp spectroscopy at the single-monomer level reproduces the kinetics of unfolding and refolding measured using polyproteins, which proves that there is no mechanical effect of tethering proteins to one another in the case of ubiquitin and I27.  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