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.
An alternate model for surface noncovalent and surface covalent binding of aflatoxin B1 to N(7) of guanine in DNA is proposed. This model considers the out-of-plane motions of C(8) of aflatoxin B1 in those interactions. The covalent intercalated fit of aflatoxin B1 into DNA arises from steric adjustments made by DNA at the covalent intercalation site as well as local strain in the bond angles about N(7) of guanine and C(8) of aflatoxin B1. The bond angle about N(7) deviates modestly from the sp2 value toward the sp3 value. This study suggests that the surface covalent aflatoxin B1-DNA complex serves only a minor role in aflatoxin's precarcinogenic interaction with DNA and is a likely correctable error.  相似文献   

2.
The modes of reaction of the tumorigenic bay region diol epoxide anti-BADE [+/-)-trans-3,4-diol-anti-1,2-epoxy-1,2,3,4-tetrahydrobenz[a]anthr acene) and the less potent tumor initiating diastereomer syn-BADE [+/-)-trans-3,4-diol-syn-1,2-epoxy-1,2,3,4-tetrahydrobenz[a]anthra cene) with native, double-stranded DNA were compared. The bay-region diol epoxide derived from 3-methylcholanthrene (3-MCDE, racemic trans-9,10-diol-anti-7,8-epoxy-7,8,9,10-tetrahydromethylcholanthrene+ ++) was included in this study in order to assess the effects of the methyl and methylene substituents on the reactivity with DNA. Utilizing linear dichroism and other spectroscopic methods, it is shown that all three diol epoxides forn non-covalent complexes with DNA. The diastereomers anti-BADE and syn-BADE form intercalative physical complexes, but the association constant K of the syn-diastereomer is about 6-7 times smaller than for anti-BADE; this effect is ascribed to the bulky quasi-diaxial conformation of the diol epoxide ring in the syn diastereomer. The value of K (4000 M-1) is similar for anti-BADE and 3-MCDE, although the latter is not intercalated in the classical sense since the short axis of the molecule is tilted closer to the axis of the DNA double helix. The conformations of the covalent DNA adducts are interpreted in terms of a quasi-intercalative conformation (site I), and a conformation in which the long axes of the polycyclic molecules are tilted closer to the axis of the helix (site II). Both tumorigens, anti-BADE and 3-MCDE, undergo a marked re-orientation from a non-covalent site I to a covalent site II conformation upon binding chemically with the DNA bases, although a small fraction of the covalent anti-BADE adducts remains quasi-intercalated; in contrast, the alkyl substituents in 3-MCDE not only prevent the formation of intercalative physical complexes, but also the formation of site I covalent adducts. In the case of the less tumorigenic syn-BADE, both the non-covalent complexes and the covalent adducts are of the site I-type. The bay-region diol epoxide of benz[a]anthracene and of 3-methylcholanthrene display a similar pattern of reactivities and covalent adduct conformations as the bay region diol epoxide derivatives of benz[a]pyrene, suggesting that adduct conformation might be an important factor in determining the levels of mutagenic and tumorigenic activities of this class of compounds.  相似文献   

3.
Efficient ophthalmic therapy requires the development of strategies that can provide sufficiently high drug levels in the ocular structures for a prolonged time. This work focuses on the suitability of poly-(cyclo)dextrins as carriers able to solubilize the carbonic anhydrase inhibitor (CAI) ethoxzolamide (ETOX), which is so far used for oral treatment of glaucoma. Topical ocular treatment should notably enhance the efficiency/safety profile of the drug. Natural α-, β- and γ-cyclodextrins and a maltodextrin were separately polymerized using citric acid as cross-linker agent under mild conditions. The resultant hydrophilic polymers exhibited larger capability to solubilize ETOX than the pristine (cyclo)dextrins. Moreover, they provided sustained drug diffusion in artificial lachrymal fluid. Interestingly the poly-(cyclo)dextrins solutions facilitate the loading of remarkably high doses of ETOX in poly(2-hydroxyethyl methacrylate)-based contact lenses. Exploiting ionic interactions between functional groups in the contact lenses and remnant free carboxylic acids in the citric acid linkers of poly-(cyclo)dextrins led to the retention of the drug-loaded poly-(cyclo)dextrins and, in turn, to sustained release for several weeks.  相似文献   

4.
In this study, we show that posttranslational folding of Vesicular Stomatitis virus G protein subunits can involve noncovalent, multimeric complexes as transient intermediates. The complexes are heterogeneous in size (4-21S20,W), contain several G glycopolypeptides, and are associated with BiP/GRP78. The newly synthesized, partially intrachain disulfide-bonded G proteins enter these complexes immediately after chain termination, and are released 1-4 min later as fully oxidized, trimerization-competent monomers. These monomers are properly folded, judging by their binding of conformation-specific mAbs. When the G protein is translated in the presence of DTT, it remains reduced, largely unfolded and aggregated in the ER, but it can fold successfully when the DTT is removed. In this case, contrary to normal folding, the aggregates become transiently disulfide cross-linked. We also demonstrated that the fidelity of the folding process is dependent on metabolic energy. Finally, we established that the G protein of the folding mutant of the Vesicular Stomatitis virus, ts045, is blocked at a relatively late step in the folding pathway and remains associated with oligomeric, BiP/GRP78-containing folding complexes.  相似文献   

5.
Resistin, an adipocyte secreted cysteine rich hormone has been implicated as molecular link between obesity and type 2 diabetes in a murine model. Although, at the protein level mouse and human resistin show remarkable similarities with respect to conserved cysteine residues, the physiological role of human resistin is not yet clear. In the present study we describe the purification and refolding of human recombinant resistin using two different refolding processes. Gel filtration analysis of protein refolded by both the methods revealed that human recombinant resistin, like mouse resistin, has a tendency to form dimers. Interestingly, dimerization of resistin appears to be mediated by both covalent (disulfide bond mediated) and non-covalent interactions as seen on reducing and non-reducing SDS-PAGE. Circular dichroism spectral analysis revealed that human resistin peptide backbone is a mixture of alpha-helical and beta-sheet conformation with significant amounts of unordered structure, similar to the mouse resistin. It is likely that the first cysteine (Cyst22) of human resistin, which is equivalent to mouse Cyst26, may be involved in stabilizing the dimers through covalent interaction.  相似文献   

6.
Duplex DNA incubated with adriamycin, dithiothreitol (DTT), and Fe3+ under aerobic, aqueous conditions yields double-stranded (DS) DNA bands by denaturing polyacrylamide gel electrophoresis (DPAGE) analysis, characteristic of DNAs which are interstrand cross-linked. Another laboratory has provided evidence that formaldehyde produced under these conditions promotes the covalent linkage of adriamycin to one strand of DNA and suggested that this complex results in the anomalous DPAGE behavior. We provide herein strong support for this interpretation. We show: (a) that mixtures of DNA and adriamycin incubated with DTT/Fe3+, H2O2, or formaldehyde all show DS DNA bands on DPAGE, (b) that the DS DNA bands and the formaldehyde-mediated lesion (detected by an indirect, GC-MS analysis) form with similar time courses, and in similar amounts, and (c) that the DNA in the DS DNA bands contains approximately one such lesion per DNA, whereas the single-stranded DNA is devoid of it. These results further support the interpretation that adriamycin does not create interstrand cross-links in DNA, and that the DS DNA observed in DPAGE experiments derives from the formaldehyde-mediated monoadduct.  相似文献   

7.
Covalent binding of reactive metabolites of drugs to proteins has been a predominant hypothesis for the mechanism of toxicity caused by numerous drugs. The development of efficient and sensitive analytical methods for the separation, identification, quantification of drug-protein adducts have important clinical and toxicological implications. In the last few decades, continuous progress in analytical methodology has been achieved with substantial increase in the number of new, more specific and more sensitive methods for drug-protein adducts. The methods used for drug-protein adduct studies include those for separation and for subsequent detection and identification. Various chromatographic (e.g., affinity chromatography, ion-exchange chromatography, and high-performance liquid chromatography) and electrophoretic techniques [e.g., sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE), two-dimensional SDS-PAGE, and capillary electrophoresis], used alone or in combination, offer an opportunity to purify proteins adducted by reactive drug metabolites. Conventionally, mass spectrometric (MS), nuclear magnetic resonance, and immunological and radioisotope methods are used to detect and identify protein targets for reactive drug metabolites. However, these methods are labor-intensive, and have provided very limited sequence information on the target proteins adducted, and thus the identities of the protein targets are usually unknown. Moreover, the antibody-based methods are limited by the availability, quality, and specificity of antibodies to protein adducts, which greatly hindered the identification of specific protein targets of drugs and their clinical applications. Recently, the use of powerful MS technologies (e.g., matrix-assisted laser desorption/ionization time-of-flight) together with analytical proteomics have enabled one to separate, identify unknown protein adducts, and establish the sequence context of specific adducts by offering the opportunity to search for adducts in proteomes containing a large number of proteins with protein adducts and unmodified proteins. The present review highlights the separation and detection technologies for drug-protein adducts, with an emphasis on methodology, advantages and limitations to these techniques. Furthermore, a brief discussion of the application of these techniques to individual drugs and their target proteins will be outlined.  相似文献   

8.
Starch dextrins of different molecular sizes (DPn 311, 142 and 39) were prepared by hydrolyzing a high amylose maize starch in acidic alcohol solutions. The dextrins were dissolved in an aqueous dimethyl sulfoxide solution (90% DMSO), and then the solution was allowed to migrate down into n-butanol separated by a membrane filter. The complex was gradually formed between the dextrin and butanol, and precipitated in the butanol layer. The dextrin–butanol complex yielded V6-I type crystals with broad reflections (d-spacings 1.123, 0.657 and 0.429 nm) under X-ray diffractograms. Platelets of average length less than 100 nm, interspersed in amorphous matrices, were observed in complexes of DPn 311 and 142, but that of DPn 39 showed different morphology, and the formation of complexes was limited. By hydrolyzing the complex of DPn 311 with α-amylase, amorphous matrices were selectively removed, and crystallites of 23–72 nm showing a V6-I X-ray diffraction pattern were obtained. However, crystallites in complexes of DPn 142 and 39 were eroded by amylolysis, forming large aggregates.  相似文献   

9.
The covalent attachment of polyethylene glycol of 5000 daltons to non-essential groups on trypsin produces an adduct that no longer precipitates with anti-trypsin antibody. In comparison with trypsin, polyethylene glycol-trypsin preparations show equal or greater activity against N-alpha-benzoyl-L-arginine ethyl ester, about one-fourth activity against angiotensin II, and little activity against bovine liver catalase. The polyethylene glycol-trypsin adduct dissolves soft blood clots at one-fourth the rate of trypsin. Soybean trypsin inhibitor produces two-thirds inhibition of the adduct under conditions that cause complete inhibition of trypsin.  相似文献   

10.
The binding of the antitumor drug CC-1065 has been studied with nuclear magnetic resonance (NMR) spectroscopy. This study involves two parts, the elucidation of the covalent binding site of the drug to DNA and a detailed investigation of the noncovalent interactions of CC-1065 with a DNA fragment through analysis of 2D NOE (NOESY) experiments. A CC-1065-DNA adduct was prepared, and an adenine adduct was released upon heating. NMR (1H and 13C) analysis of the adduct shows that the drug binds to N3 of adenine by reaction of its cyclopropyl group. The reaction pathway and product formed were determined by analysis of the 13C DEPT spectra. An octamer duplex, d(CGATTAGC.GCTAATCG), was synthesized and used in the interaction study of CC-1065 and the oligomer. The duplex and the drug-octamer complex were both analyzed by 2D spectroscopy (COSY, NOESY). The relative intensity of the NOEs observed between the drug (CC-1065) and the octamer duplex shows conclusively that the drug is located in the minor groove, covalently attached to N3 of adenine 6 and positioned from the 3'----5' end in relation to strand A [d(CGATTA6GC)]. A mechanism for drug binding and stabilization can be inferred from the NOE data and model-building studies.  相似文献   

11.
Single-walled carbon nanotubes (SWNTs) covalently functionalised with polyethylene glycol (PEG) or noncovalently coated with PEGylated lipids were simulated in water and in lipid bilayers at different PEG sizes and grafting densities using coarse-grained force fields. Starting with the random position of three SWNT–PEG complexes in water, larger PEGs at higher grafting densities more significantly inhibit the aggregation of SWNTs because of larger radii of gyration and hydrodynamic radii of the SWNT–PEG complex, which influence the thickness and the wrapping extent of PEG layer. In particular, PEG-functionalised SWNTs, where PEGs are evenly grafted along the SWNT, disperse, while PEG-coated SWNTs aggregate because SWNTs are less covered by randomly adsorbed PEGylated lipids. Simulations of SWNT–PEGs in lipid bilayers show that PEG (Mw = 550 and 2000)-functionalised SWNTs bind to the bilayer surface but do not insert into the bilayer, while PEG-coated SWNTs insert into the bilayer because PEGylated lipids detach from SWNTs and mix with bilayer lipids. These findings support recent experiments at the same PEG size and density, which suggested that PEG-coated SWNTs may form bundles and thus cannot be easily excreted through the renal route, while PEG-functionalised SWNTs may remain individual and thus show more renal excretion.  相似文献   

12.
D A Torchia  C M Deber 《Biopolymers》1972,11(3):653-659
NMR spectra of cyclo(tri-L -prolyl), c-(P)3, show large shifts of the Hα resonance on adding C6D6 to a solution of c-(P)3 in CD2Cl2. CPK models and observed coupling constants indicate a rigid c-(P)3 conformation, independent of solvent composition, suggesting that these shifts result from formation of stereospecific C6D6–c-(P)3 collision complexes in which the c-(P)3 Hα lie near the face of the aromatic ring. The temperature dependence of the Hα shifts and the solvent dependent shifts observed on adding toluene-d8 or nitrobenzene-d5 to the c-(P)3 solution suggest that preferred C6D6–c-(P)3 orientations result from attractive interactions between the electron-rich aromatic ring and the electropositive Hα's and/or δ+ nitrogen atoms in the peptide backbone. Reports of such interactions in increasingly diverse peptide model systems suggest that they may play a role in stabilizing protein structures.  相似文献   

13.
Fourier transform ion cyclotron resonance mass spectrometry (FTICR-MS) was applied for the study of noncovalent chaperone SecB-ligand complexes produced in solution and examined in the gas phase with the aid of electrospray ionization (ESI). Since chaperone proteins are believed to recognize and bind only with ligands with nonnative tertiary structure, this work required careful unfolding of the ligand and subsequent reaction with the intact chaperone (the noncovalent tetrameric protein, SecB). A high denaturant concentration was employed to produce nonnative structures of the OppA, and microdialysis of the resulting solutions containing the chaperone-ligand complexes was carried out to rapidly remove the denaturant prior to analysis. Multistage mass spectrometry was essential to the successful study of these complexes since the initial mass spectra indicated extensive adduction that precluded mass measurements, even after microdialysis. However, low energy collisional activation of the ions in the FTICR trap proved useful for adduct removal, and careful control of excitation level preserved the intact complexes of interest, revealing a 1:1 SecB:OppA stoichiometry. To our knowledge, these results present the first direct observation of chaperone-ligand noncovalent complexes and the highest molecular weight heterogeneous noncovalent complex observed to date by mass spectrometry. Furthermore, these results highlight the capabilities of FTICR for the study of such complex systems, and the development of a greater understanding of chaperone interactions in protein export.  相似文献   

14.
Methods of enzyme molecules covalent immobilization on the surface of liposomes are suggested. The methods permit more protein molecules to be bound than traditional methods of noncovalent immobilization by means of adsorption or incorporation. The liposome membranes preserve their integrity during immobilization. At the same tame, the enzyme bound with the liposome surface via the “spacer” groups completely preserves its ability to interact with a specific macromolecular compound.  相似文献   

15.
The kinetics of reduction of free flavin semiquinones of the individual components of 1:1 covalent and electrostatic complexes of yeast ferric and ferryl cytochrome c peroxidase and ferric horse cytochrome c have been studied. Covalent cross-linking between the peroxidase and cytochrome c at low ionic strength results in a complex that has kinetic properties both similar to and different from those of the electrostatic complex. Whereas the cytochrome c heme exposure to exogenous reductants is similar in both complexes, the apparent electrostatic environment near the cytochrome c heme edge is markedly different. In the electrostatic complex, a net positive charge is present, whereas in the covalent complex, an essentially neutral electrostatic charge is found. Intracomplex electron transfer within the two complexes is also different. For the covalent complex, electron transfer from ferrous cytochrome c to the ferryl peroxidase has a rate constant of 1560 s-1, which is invariant with respect to changes in the ionic strength. The rate constant for intracomplex electron transfer within the electrostatic complex is highly ionic strength dependent. At mu = 8 mM a value of 750 s-1 has been obtained [Hazzard, J. T., Poulos, T. L., & Tollin, G. (1987) Biochemistry 26, 2836-2848], whereas at mu = 30 mM the value is 3300 s-1. This ionic strength dependency for the electrostatic complex has been interpreted in terms of the rearrangement of the two proteins comprising the complex to a more favorable orientation for electron transfer. In the case of the covalent complex, such reorientation is apparently impeded.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
A kinetic comparison of the hydrolase and transferase activities of two bacterial phospholipase D (PLD) enzymes with little sequence homology provides insights into mechanistic differences and also the more general role of Ca(2+) in modulating PLD reactions. Although the two PLDs exhibit similar substrate specificity (phosphatidylcholine preferred), sensitivity to substrate aggregation or Ca(2+), and pH optima are quite distinct. Streptomyces sp. PMF PLD, a member of the PLD superfamily, generates both hydrolase and transferase products in parallel, consistent with a mechanism that proceeds through a covalent phosphatidylhistidyl intermediate where the rate-limiting step is formation of the covalent intermediate. For Streptomyces chromofuscus PLD, the two reactions exhibit different pH profiles, a result consistent with a mechanism likely to involve direct attack of water or an alcohol on the phosphorus. Ca(2+), not required for monomer or micelle hydrolysis, can activate both PLDs for hydrolysis of PC unilamellar vesicles. In the case of Streptomyces sp. PMF PLD, Ca(2+) relieves product inhibition by interactions with the phosphatidic acid (PA). A similar rate enhancement could occur with other HxKx(4)D-motif PLDs as well. For S. chromofuscus PLD, Ca(2+) is absolutely critical for binding of the enzyme to PC vesicles and for PA activation. That the Ca(2+)-PA activation involves a discreet site on the protein is suggested by the observation that the identity of the C-terminal residue in S. chromofuscus PLD can modulate the extent of product activation.  相似文献   

17.
Efimov I  Cronin CN  McIntire WS 《Biochemistry》2001,40(7):2155-2166
Each flavoprotein subunit (alpha or PchF) of the alpha(2)beta(2) flavocytochrome p-cresol methylhydroxylase (PCMH) from Pseudomonas putida contains FAD covalently attached to Tyr384. PCMH oxidizes p-cresol to 4-hydroxybenzyl alcohol, which is oxidized subsequently by PCMH to 4-hydroxybenzaldehyde. The Y384F mutant form of PchF (apo-PchF[Y384F]) displayed stoichiometric noncovalent FAD binding. PchF[Y384F]FAD associated with the cytochrome subunit (beta or PchC) (producing PCMH[Y384F]), although not as avidly as with wild-type PchF containing covalently bound FAD (PchF(C)). Dramatic increases in the two-electron E(m,7) (NHE) values for FAD were observed when it bound noncovalently to either apo-PchF or apo-PchF[Y384F], and the two-electron E(m,7) value for FAD was increased further by about 75 mV upon covalent binding to PchF, i.e., PchF(C). The E(m,7) values increased by approximately 20 and 45 mV, respectively, when PchF(C) and PchF[Y384F]FAD associated with PchC. The two-electron E(m,7) for covalently bound FAD in PCMH is 84 mV, the highest measured for a flavoprotein. The values for the one-electron redox potentials (E(m,7), NHE) for FAD were measured also for various forms of PchF. Under anaerobiosis, the reduction of PchF[Y384F]FAD by substrates was similar to that observed previously for PchF containing noncovalently bound FAD. Stopped-flow kinetic studies indicated a rapid substrate reduction of the FAD and heme in PCMH[Y384F] which produced PchF[Y384F]FAD(rad) x PchC, the mutant enzyme containing the flavin radical and reduced heme. These experiments also revealed a slow reduction of unassociated PchC(ox) by PchF[Y384F]FAD(rad) x PchC. Steady-state kinetic studies of the reaction of PCMH[Y384F] with p-cresol indicated that the K(m) for this substrate was unchanged relative to that of PCMH, but that the k(cat) was diminished by an order of magnitude. The data indicate that the covalent attachment of FAD to PchF assists catalysis by raising the E(m,7) of the flavin. Contributions to this effect likely result from conformational changes.  相似文献   

18.
Nagao T  Tanino T  Iwaki M 《Chirality》2003,15(5):423-428
To determine the effect of arthritis on the disposition of flurbiprofen (FP) and its acyl glucuronide (FPG) as well as formation of covalent adducts with plasma protein, a pharmacokinetic study was carried out in adjuvant-induced arthritic (AA) rats. In control animals the pharmacokinetics of FP were stereoselective following intravenous bolus injection of rac-FP: (-)-(R)-FP showed higher plasma clearance (CL(tot)) and shorter mean residence time (MRT) compared to (+)-(S)-FP. The CL(tot) and clearance for the glucuronide formation (CL(glu)) of both enantiomers in AA rats were extremely increased compared to those in control rats. Increased total clearance in AA rats was due, at least in part, to a remarkable increase in the plasma unbound fraction of FP, consistent with a decrease in the plasma albumin level. The yield of covalent binding of FP to plasma protein in AA rats was less than that in controls, being consistent with the decrease in the plasma acyl glucuronide level.  相似文献   

19.
K Kvalnes-Krick  M S Jorns 《Biochemistry》1986,25(20):6061-6069
Sarcosine oxidase was purified to homogeneity from Corynebacterium sp. P-1, a soil organism isolated by a serial enrichment technique. The enzyme contains 1 mol of noncovalently bound flavin [flavin adenine dinucleotide (FAD)] plus 1 mol of covalently bound flavin [8 alpha-(N3-histidyl)-FAD] per mole of enzyme (Mr 168,000). The two flavins appear to have different roles in catalysis. The enzyme has an unusual subunit composition, containing four dissimilar subunits (Mr 100,000, 42,000, 20,000, and 6000). The same subunits are detected in Western blot analysis of cell extracts prepared in the presence of trichloroacetic acid, indicating that the subunits are a genuine property of the enzyme as it exists in vivo. The presence of both covalent and noncovalent flavin in a single enzyme is extremely unusual and has previously been observed only with a sarcosine oxidase from a soil Corynebacterium isolated in Japan. The enzymes exhibit many similarities but are distinguishable in electrophoretic studies. Immunologically, the enzymes are cross-reactive but not identical. The results indicate that the synthesis of a sarcosine oxidase containing both covalent and noncovalent flavin is not a particularly unusual event in corynebacteria.  相似文献   

20.
To enable studies of conformational changes within multimolecular complexes, we present a simultaneous, four-color single molecule fluorescence methodology implemented with total internal reflection illumination and camera-based, wide-field detection. We further demonstrate labeling histidine-tagged proteins noncovalently with Tris-nitrilotriacetic acid (Tris-NTA)-conjugated dyes to achieve single molecule detection. We combine these methods to colocalize the mismatch repair protein MutSα on DNA while monitoring MutSα-induced DNA bending using F?rster resonance energy transfer (FRET) and to monitor assembly of membrane-tethered SNARE protein complexes.  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