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 31 毫秒
1.
The one-electron oxidation of methionine (Met) plays an important role in the redox reactions of Met in peptides and proteins under conditions of oxidative stress, e.g., during the metal-catalyzed oxidation of beta-amyloid peptide (beta A). However, little information is available with regard to mechanisms and product formation during the metal-catalyzed oxidation of Met. Here, we demonstrate that two-electron oxidation of Met in Fenton reactions, carried out aerobically by [Fe(II)(EDTA)](2-) and H(2)O(2) (EDTA = ethylenediaminetetra acetate) is the consequence of two consecutive one-electron transfer reactions carried out by either free or complexed hydroxyl radicals, followed by the reaction of an intermediary sulfur-nitrogen bonded radical cation (sulfuranyl radical) with O(2). The model peptide Met-Met represents an ideal substrate for these investigations as its one-electron oxidation, followed by reaction with molecular oxygen, produces unique intermediates, azasulfonium diastereomers, which can be chemically isolated before hydrolysis to sulfoxide occurs.  相似文献   

2.
Methionine regulates copper/hydrogen peroxide oxidation products of Abeta.   总被引:5,自引:0,他引:5  
Metal-catalysed oxidation (MCO) may play a causative role in the pathogenesis of Alzheimer's disease (AD). Amyloid beta peptide (Abeta), the major biomarker of AD, in the presence of copper ions reduces Cu(2+) to Cu(+) and catalyses the formation of H(2)O(2) that subsequently induces radicals through Fenton chemistry. Abeta is also subject to attack by free radicals, where the presence of Cu(2+) in conjunction with H(2)O(2) catalyses oxygenation, primarily at the methionine sulfur atom. This work investigates MCO of Abeta, to gain further insight into the role of oxidative stress in AD. By combining a fluorescence assay with gel electrophoresis to monitor MCO reactions of Abeta (1-28) in the presence and absence of methionine it was determined that methionine can both protect some residues against MCO and promote the oxidation of Tyr(10) specifically. Electrospray ionization mass spectrometric analysis of methionine MCO products indicated the formation of methionine sulfoxide, methionine sulfone and related hydroxylated products. Similar products could be formed from the oxidation of Met(35) of Abeta and may relate to changes in properties of the peptide following MCO.  相似文献   

3.
Oxidative degradation of collagen and the model peptides by Cu(II)/H2O2 has been studied. The depolymerization of collagen was predominantly observed by use of gel filtration chromatography. Polyproline was used as a model for collagen, and the oxidative modification was examined by amino acid analysis. Glutamic acid and gamma-aminobutyric acid were identified in the hydrolysates of oxidized polyproline. The formation of glutamic acid was reduced by treatment with NaBH4. The model peptide, (Pro-Pro-Gly)10, was also degraded by Cu(II)/H2O2, and a new N-terminal glycine was generated in proportion to the reaction time. Hydroxyl radical scavengers show only partial inhibition of the degradation of (Pro-Pro-Gly)10. In order to estimate the fragmentation mechanism, we used N-tert-butoxycarbonyl (Boc)-L-prolylglycine as a model for collagen and (Pro-Pro-Gly)10. The degradation products were isolated and characterized. Then N-tert-Boc-2-pyrrolidone, which provides gamma-aminobutyric acid by acid hydrolysis, was identified. The formation of a 2-pyrrolidone compound from oxidized Boc-L-prolylglycine is direct evidence for the scission of the peptide bond. The time-dependent formation of N-tert-Boc-2-pyrrolidone and liberation of glycine from N-tert-Boc-L-prolylglycine exposed to Cu(II)/H2O2 was observed. These results suggest that the cleavage of the peptide bond (Pro-Gly) was caused by oxidation of the proline residue, which led to the formation of the 2-pyrrolidone compound. We confirmed that proline oxidation leads to the fragmentation of proteins, accompanied by the formation of a 2-pyrrolidone structure.  相似文献   

4.
The classical tachykinins, substance P, neurokinin A and neurokinin B are predominantly found in the nervous system where they act as neurotransmitters and neuromodulators. Significantly reduced levels of these peptides were observed in neurodegenerative diseases and it may be suggested that this reduction may also result from the copper(II)-catalyzed oxidation. The studies of the interaction of copper(II) with neurokinin A and the copper(II)-catalyzed oxidation were performed. Copper(II) complexes of the neurokinin A (His-Lys-Thr-Asp-Ser-Phe-Val-Gly-Leu-Met-NH2) and acetyl-neurokinin A (Ac-His-Lys-Thr-Asp-Ser-Phe-Val-Gly-Leu-Met-NH2) were studied by potentiometric, UV-Vis (UV-visible), CD (circular dichroism) and EPR spectroscopic methods to determine the stoichiometry, stability constants and coordination modes in the complexes formed. The histidine residue in first position of the peptide chain of neurokinin A coordinates strongly to Cu(II) ion with histamine-like {NH2, NIm} coordination mode. With increasing of pH, the formation of a dimeric complex Cu2H2L2 was found but this dimeric species does not prevent the deprotonation and coordination of the amide nitrogens. In the Ac-neurokinin A case copper(II) coordination starts from the imidazole nitrogen of the His; afterwards three deprotonated amide nitrogens are progressively involved in copper coordination. To elucidate the products of the copper(II)-catalyzed oxidation of the neurokinin A and Ac-neurokinin A, liquid chromatography-mass spectrometry (LC-MS) method and Cu(II)/hydrogen peroxide as a model oxidizing system were employed.Oxidation target for both studied peptides is the histidine residue coordinated to the metal ions. Both peptides contain Met and His residues and are very susceptible on the copper(II)-catalyzed oxidation.  相似文献   

5.
We have studied the metal-catalyzed oxidation (MCO) of brain-derived neurotrophic factor (BDNF) with regard to target sites and potential conformational changes of the protein. The exposure of BDNF to three different levels of ascorbate/Cu(II)/O2 [20 microM Cu(II), 2 mM ascorbate (level 1); 20 microM Cu(II), 4 mM ascorbate (level 2); 40 microM Cu(II), 4 mM ascorbate (level 3)], chosen based on the extent of chemical modification of Met and His, respectively, resulted in the exclusive oxidation of a buried Met residue, Met92, at level 1 but in the predominant oxidation of His at level 3. His modification had a significant impact on the structure of BDNF, as quantified by CD and ANSA fluorescence measurements, while Met oxidation had not, also assessed through complementary oxidation of BDNF through hydrogen peroxide. Our ultimate objective was the correlation of the surface exposure of an oxidized His residue in a protein with potential effects on the conformational integrity of the oxidized protein. In a series of three proteins, human growth hormone (hGH), human relaxin (hR1x), and BDNF, we have now observed that His oxidation is paralleled by significant conformational changes when the target His residue is more surface exposed (hR1x, BDNF) while conformational consequences of His modification are less significant when the target His residues are more buried in the interior of the protein (hGH).  相似文献   

6.
Abstract The majority of extant life forms thrive in an O(2)-rich environment, which unavoidably induces the production of reactive oxygen species (ROS) during cellular activities. ROS readily oxidize methionine (Met) residues in proteins/peptides to form methionine sulphoxide [Met(O)] that can lead to impaired protein function. Two methionine sulphoxide reductases, MsrA and MsrB, catalyse the reduction of the S and R epimers, respectively, of Met(O) in proteins to Met. The Msr system has two known functions in protecting cells against oxidative damage. The first is to repair proteins that have lost activity due to Met oxidation and the second is to function as part of a scavenger system to remove ROS through the reversible oxidation/reduction of Met residues in proteins. Bacterial, plant and animal cells lacking MsrA are known to be more sensitive to oxidative stress. The Msr system is considered an important cellular defence mechanism to protect against oxidative stress and may be involved in ageing/senescence. MsrA is present in all known eukaryotes and eubacteria and a majority of archaea, reflecting its essential role in cellular life. MsrB is found in all eukaryotes and the majority of eubacteria and archaea but is absent in some eubacteria and archaea, which may imply a less important role of MsrB compared to MsrA. MsrA and MsrB share no sequence or structure homology, and therefore probably emerged as a result of independent evolutionary events. The fact that some archaea lack msr genes raises the question of how these archaea cope with oxidative damage to proteins and consequently of the significance of msr evolution in oxic eukaryotes dealing with oxidative stress. Our best hypothesis is that the presence of ROS-destroying enzymes such as peroxiredoxins and a lower dissolved O(2) concentration in those msr-lacking organisms grown at high temperatures might account for the successful survival of these organisms under oxidative stress.  相似文献   

7.
The N-terminal, extracellular regions of eukaryotic high affinity copper transport (Ctr) proteins vary in composition of the Cu(i) binding amino acids: methionine, histidine, and cysteine. To examine why certain amino acids are exploited over others in Ctrs from different organisms, the relative Cu(i) binding affinity and the dependence of binding on pH were examined for 3 peptides of the sequence MG(2)XG(2)MK, where X is either Met, His, or Cys. Cu(i) affinity was examined using an ascorbic acid oxidation assay, an electrospray ionization mass spectrometry technique, and spectrophotometric titration with a competitive Cu(i) chelator. The relative affinities of the peptides with Cu(i) reveal a trend whereby Cys > His > Met at pH 7.4 and Cys > Met > His at pH 4.5. Ligand geometry and metric parameters were determined with X-ray absorption spectroscopy. Susceptibility of the peptides to oxidation by hydrogen peroxide and copper-catalyzed oxidative conditions was evaluated by mass spectrometry. These results support hypotheses as to why certain Cu(i) binding amino acids are preferred over others in proteins expressed at different pH and exposed to oxidative environments. The results also have implications for interpreting site-directed mutagenesis studies aimed at identifying copper binding amino acids in copper trafficking proteins.  相似文献   

8.
Aggregation of Aβ peptides into amyloid plaques is considered to trigger the Alzheimer’s disease (AD), however the mechanism behind the AD onset has remained elusive. It is assumed that the insoluble Aβ aggregates enhance oxidative stress (OS) by generating free radicals with the assistance of bound copper ions. The aim of our study was to establish the role of Met35 residue in the oxidation and peptide aggregation processes. Met35 can be readily oxidized by H2O2. The fibrillization of Aβ with Met35 oxidized to sulfoxide was three times slower compared to that of the regular peptide. The fibrils of regular and oxidized peptides looked similar under transmission electron microscopy. The relatively small inhibitory effect of methionine oxidation on the fibrillization suggests that the possible variation in the Met oxidation state should not affect the in vivo plaque formation. The peptide oxidation pattern was more complex when copper ions were present: addition of one oxygen atom was still the fastest process, however, it was accompanied by multiple unspecific modifications of peptide residues. Addition of copper ions to the Aβ with oxidized Met35 in the presence of H2O2, resulted a similar pattern of nonspecific modifications, suggesting that the one-electron oxidation processes in the peptide molecule do not depend on the oxidation state of Met35 residue. Thus, it can be concluded that Met35 residue is not a part of the radical generating mechanism of Aβ–Cu(II) complex.  相似文献   

9.
Fluorescent cell analytical techniques require the incorporation of a fluorophore into the target molecule without causing a significant change in the native conformation. Many short peptides have a limited number of reactive groups that can be labeled without affecting the biological activity. In this work we present several methods for labeling beta-amyloid peptides (betaA[25-35], betaA[1-40]) and their derivatives (LPFFD, RIIGL and RVVIA) with different chromophores exclusively at the N-terminus. In the case of liquid-phase labeling, fluorescein isothiocyanate was used. The side-chain amino function of Lys, if present in the sequence, was protected with an Fmoc group, whereby the hydrophobic character of the peptide was further increased. The labeling reaction was carried out in an appropriate deaggregating solvent, DMSO. For solid-phase labeling, 5(6)-carboxyfluorescein and 7-amino-4-methyl-3-coumarinylacetic acid were applied. Several cleavage cocktails were tested for removal of the labeled amyloid peptides from the resin in order to completely suppress the oxidation of Met.  相似文献   

10.
Pseudomonas aeruginosa azurin binds copper so tightly that it remains bound even upon polypeptide unfolding. Copper can be substituted with zinc without change in protein structure, and also in this complex the metal remains bound upon protein unfolding. Previous work has shown that native-state copper ligands Cys112 and His117 are two of at least three metal ligands in the unfolded state. In this study we use isothermal titration calorimetry and spectroscopic methods to test if the native-state ligand Met121 remains a metal ligand upon unfolding. From studies on a point-mutated version of azurin (Met121Ala) and a set of model peptides spanning the copper-binding C-terminal part (including Cys112, His117 and Met121), we conclude that Met121 is a metal ligand in unfolded copper-azurin but not in the case of unfolded zinc-azurin. Combination of unfolding and metal-titration data allow for determination of copper (Cu(II) and Cu(I)) and zinc affinities for folded and unfolded azurin polypeptides, respectively.  相似文献   

11.
We report here that in vitro exposure of monomeric actin to hydrogen peroxide leads to a conversion of 6 of the 16 methionine residues to methionine sulfoxide residues. Although the initial effect of H2O2 on actin is the oxidation of Cys374, we have found that Met44, Met47, Met176, Met190, Met269, and Met355 are the other sites of the oxidative modification. Met44 and Met47 are the methionyl sites first oxidized. The methionine residues that are oxidized are not simply related to their accessibility to the external medium and are found in all four subdomains of actin. The conformations of subdomain 1, a region critical for the functional binding of different actin-binding proteins, and subdomain 2, which plays important roles in the polymerization process and stabilization of the actin filament, are changed upon oxidation. The conformational changes are deduced from the increased exposure of hydrophobic residues, which correlates with methionine sulfoxide formation, from the perturbations in tryptophan fluorescence, and from the decreased susceptibility to limited proteolysis of oxidized actin.  相似文献   

12.
DuBois JL  Klinman JP 《Biochemistry》2005,44(34):11381-11388
The copper amine oxidases (CAOs) catalyze the O(2)-dependent, two-electron oxidation of amines to aldehydes at an active site that contains Cu(II) and topaquinone (TPQ) cofactor. TPQ arises from the autocatalytic, post-translational oxidation of a tyrosine side chain in the active site. Monooxygenation within the ring of tyrosine at a single Cu(II) site is unique in biology and occurs as an early step in the formation of TPQ. The mechanism of this reaction has been further examined in the CAO from Hansenula polymorpha (HPAO). When a Clark electrode fitted to a custom-made, gastight apparatus over a range of initial concentrations of O(2) was used, rates of O(2) consumption at levels greater than air are seen to be reduced relative to earlier results, yielding K(D)(apparent) = 216 microM for O(2). This is consistent with a mechanism in which O(2) binds reversibly to the active site, triggering a conformational change that promotes ligation of tyrosinate to Cu(II). The activated Cu(II)-tyrosinate species has been proposed to react with O(2) in a rate-limiting step, although it was also possible that breakdown of a putative peroxy-intermediate controlled TPQ formation. To test the latter hypothesis, Cu(II)-free HPAO was prepared with 3,5-ring-[(2)H(2)]-tyrosine incorporated throughout the primary sequence. The absence of an isotope effect on the rate of TPQ formation eliminates cleavage of this C-H bond in a proposed Cu(II)-aryl-peroxide intermediate as a rate limiting step. The role of methionine 634, previously found to moderate O(2) binding during the catalytic cycle, is shown here to serve a similar function in TPQ formation. As with catalysis, the rate of TPQ formation correlates with the volume of the hydrophobic side chain at position 634, implicating similar binding sites for O(2) during catalysis and cofactor biogenesis.  相似文献   

13.
The biotransformation of butylated hydroxyanisole (BHA), a possible carcinogenic food antioxidant, includes o-demethylation to 2-tert-butyl(1,4)hydroquinone (TBHQ) which can subsequently be oxidized to 2-tert-butyl(1,4)paraquinone (TBQ). In this study, we have examined the capacity of Cu, a nuclei- and DNA-associated transition metal, to mediate the oxidation of TBHQ. In phosphate buffered saline (PBS), autooxidation of TBHQ to TBQ was not detectable, while Cu(II) at micromolar concentrations strongly catalyzed the oxidation of TBHQ to TBQ. Oxidation of TBHQ by Cu(II) was accompanied by the utilization of O(2) and the concomitant generation of H(2)O(2). Using electron spin resonance spectroscopy, it was observed that Cu(II) mediated the one electron oxidation of TBHQ to a semiquinone anion radical. The formation of a semiquinone anion radical, the utilization of O(2) and the generation of H(2)O(2) and TBQ could be completely blocked by bathocuproinedisulfonic acid (BCS) and reduced glutathione (GSH), two Cu(I)-chelators. 4-Pyridyl-1-oxide-N-tert-butylnitrone (POBN)-spin trapping experiments showed that the reaction of TBHQ with Cu(II) resulted in the generation of POBN-CH(3) and POBN-CH(OH)CH(3) adducts in the presence of dimethyl sulfoxide (DMSO) and ethanol, respectively, suggesting the formation of hydroxyl radical or a similar reactive intermediate. The formation of POBN-CH(3) adduct from the TBHQ/Cu(II)+DMSO could be completely inhibited by catalase, GSH or BCS, indicating that the hydroxyl radical or its equivalent is generated from the interaction of H(2)O(2) with Cu(I). Incubation of supercoiled phiX-174 plasmid DNA with the TBHQ/Cu(II) resulted in extensive DNA strand breaks, which could be prevented by catalase or BCS. Incubation of rat hepatocytes with TBHQ in PBS led to increased formation of 8-hydroxy-2'-deoxyguanosine (8-OHdG) in nuclear DNA. The TBHQ-induced formation of 8-OHdG was markedly reduced in the presence of cell permeable Cu(I)-specific chelator, bathocuproine or neocuproine, suggesting that a Cu(II)/Cu(I) redox mechanism may also be involved in the induction of oxidative DNA damage by TBHQ in hepatocytes. Taken together, the above results conclusively demonstrate that the activation of TBHQ by Cu(II) results in the formation of TBQ, semiquinone anion radical and reactive oxygen species (ROS), and that the ROS formed may participate in oxidative DNA damage in both isolated DNA and intact cells. These reactions may contribute to the carcinogenicity as well as other biochemical activities observed with BHA in animals. To our knowledge this study provides the first evidence that endogenous cellular Cu may be capable of bioactivating TBHQ, leading to oxidative DNA damage in cultured cells.  相似文献   

14.
Amyloid aggregation of α-synuclein (AS) is one of the hallmarks of Parkinson’s disease. The interaction of copper ions with the N-terminal region of AS promotes its amyloid aggregation and metal-catalyzed oxidation has been proposed as a plausible mechanism. The AS(1–6) fragment represents the minimal sequence that models copper coordination to this intrinsically disordered protein. In this study, we evaluated the role of methionine residues Met1 and Met5 in Cu(II) coordination to the AS(1–6) fragment, and in the redox activity of the Cu–AS(1–6) complex. Spectroscopic and electronic structure calculations show that Met1 may play a role as an axial ligand in the Cu(II)–AS(1–6) complex, while Met5 does not participate in metal coordination. Cyclic voltammetry and reactivity studies demonstrate that Met residues play an important role in the reduction and reoxidation processes of this complex. However, Met1 plays a more important role than Met5, as substitution of Met1 by Ile decreases the reduction potential of the Cu–AS(1–6) complex by ~80 mV, causing a significant decrease in its rate of reduction. Reoxidation of the complex by oxygen results in oxidation of the Met residues to sulfoxide, being Met1 more susceptible to copper-catalyzed oxidation than Met5. The sulfoxide species can suffer elimination of methanesulfenic acid, rendering a peptide with no thioether moiety, which would impair the ability of AS to bind Cu(I) ions. Overall, our study underscores the important roles that Met1 plays in copper coordination and the reactivity of the Cu–AS complex.  相似文献   

15.
Radiolabelled peptides with high specificity and affinity towards receptors that are overexpressed by tumour cells are used in nuclear medicine for the diagnosis (imaging) and therapy of cancer. In some cases, the sequences of peptides under investigations contain methionine (Met), an amino acid prone to oxidation during radiolabelling procedures. The formation of oxidative side products can affect the purity of the final radiopharmaceutical product and/or impair its specificity and affinity towards the corresponding receptor. The replacement of Met with oxidation resistant amino acid analogues, for example, norleucine (Nle), can provide a solution. While this approach has been applied successfully to different radiolabelled peptides, a Met → Nle switch only preserves the length of the amino acid side chain important for hydrophobic interactions but not its hydrogen‐bonding properties. We report here the use of methoxinine (Mox), a non‐canonical amino acid that resembles more closely the electronic properties of Met in comparison to Nle. Specifically, we replaced Met15 by Mox15 and Nle15 in the binding sequence of a radiometal‐labelled human gastrin derivative [d ‐Glu10]HG(10‐17), named MG11 (d ‐Glu‐Ala‐Tyr‐Gly‐Trp‐Met‐Asp‐Phe‐NH2). A comparison of the physicochemical properties of 177Lu‐DOTA[ X 15]MG11 ( X = Met, Nle, Mox) in vitro (cell internalization/externalization properties, receptor affinity (IC50), blood plasma stability and logD) showed that Mox indeed represents a suitable, oxidation‐stable amino acid substitute of Met in radiolabelled peptide conjugates. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
The amyloid beta peptide is toxic to neurons, and it is believed that this toxicity plays a central role in the progression of Alzheimer's disease. The mechanism of this toxicity is contentious. Here we report that an Abeta peptide with the sulfur atom of Met-35 oxidized to a sulfoxide (Met(O)Abeta) is toxic to neuronal cells, and this toxicity is attenuated by the metal chelator clioquinol and completely rescued by catalase implicating the same toxicity mechanism as reduced Abeta. However, unlike the unoxidized peptide, Met(O)Abeta is unable to penetrate lipid membranes to form ion channel-like structures, and beta-sheet formation is inhibited, phenomena that are central to some theories for Abeta toxicity. Our results show that, like the unoxidized peptide, Met(O)Abeta will coordinate Cu2+ and reduce the oxidation state of the metal and still produce H2O2. We hypothesize that Met(O)Abeta production contributes to the elevation of soluble Abeta seen in the brain in Alzheimer's disease.  相似文献   

17.
The Cu,Zn SOD catalyzes the bicarbonate-dependent oxidation of a wide range of substrates by H2O2. A mechanism in accord with this activity has been described. It involves the generation of a strong oxidant (Cu(I)O, Cu(II)OH, or Cu(III)) by reaction of the active site Cu with H2O2, followed by oxidation of bicarbonate to CO3-* that in turn diffuses from the active site to oxidize the various substrates in free solution. Recently, an alternative mechanism, entailing firmly bound HCO3- and CO3-*, has been proposed [J. Biol. Chem. 278 (2003) 21032-21039]. We present data supporting the diffusible CO3-* and discuss the properties of this system that can be accommodated in this way and that preclude bound intermediates.  相似文献   

18.
Recent studies have shown that the "calcium-sensor" protein calmodulin (CaM) suffers an age-dependent oxidation of methionine (Met) to methionine sulfoxide (MetSO) in vivo. However, MetSO did not accumulate on the Met residues that show the highest solvent-exposure. Hence, the pattern of Met oxidation in vivo may give hints as to which reactive oxygen species and oxidation mechanisms participate in the oxidation of this important protein. Here, we have exposed CaM under a series of different reaction conditions (pH, [Ca(2+)], [KCl]) to various biologically relevant reactive oxygen species and oxidizing systems (peroxides, HOCl, peroxynitrite, singlet oxygen, metal-catalyzed oxidation, and peroxidase-catalyzed oxidation) to investigate whether one of these systems would lead to an oxidation pattern of CaM similar to that observed in vivo. However, generally, these oxidizing conditions led to a preferred or exclusive oxidation of the C-terminal Met residues, in contrast to the oxidation pattern of CaM observed in vivo. Hence, none of the employed oxidizing conditions was able to mimic the age-dependent oxidation of CaM in vivo, indicating that other, yet unidentified oxidation mechanisms may be important in vivo. Some oxidizing species showed a quite-remarkable diastereoselectivity for the formation of either L-Met-D-SO or L-Met-L-SO. Diastereoselectivity was dependent on the nature of the oxidizing species but was less a function of the location of the target Met residue in the protein. In contrast, diastereoselective reduction of L-Met-D-SO by protein methionine sulfoxide reductase (pMSR) was efficient regardless of the position of the L-Met-D-SO residue in the protein and the presence or absence of calcium. With only the L-Met-D-SO diastereomer being a substrate for pMSR, any preferred formation of L-Met-L-SO in vivo may cause the accumulation of MetSO unless the oxidized protein is substrate for (accelerated) protein turnover.  相似文献   

19.
In an attempt to investigate the role of histone H2B in Cu(II) induced toxicity and carcinogenesis, we synthesized the terminally blocked peptides H2B32-62 (SRKESYSVYVYKVLKQVH48PDTGISSKAMGIM) and Η2Β94-125 (IQTAVRLLLPGELAKH110AVSEGTKAVTKYTSS), mimicking the N-terminal histone-fold domain and C-terminal tail of histone H2B, respectively and studied their interaction with Cu(II) ions by means of potentiometric titrations and spectroscopic techniques (UV-visible, CD and EPR). Both peptides, H2B32-62 and H2B94-125, interacted efficiently with Cu(II) ions, forming several species from pH 4 to 11, with His48 and His110 serving as anchors for metal binding. In H2B32-62, the effective Cu(II) binding is emphasized by the formation of a soluble Cu(II)-H2B32-62 complex, unlike the unbound peptide that precipitated over pH 7.9. At physiological pH, both peptides form tetragonal 3N species with a {NIm, 2N} coordination mode. At this pH, H2B32-62 presented the formation of coordination isomers, differentiated by the presence in one of them, of an axial coordination of the carboxylate group of Asp50. Copper binding with both H2B32-62 and H2B94-125 may induce a conformational change in the peptides' original structure. At physiological conditions, this effect may interfere with nucleosome's structure and dynamics, including the ubiquitination of Lys120 which is linked to gene silencing.  相似文献   

20.
The purpose of our study was to determine the effects of specific truncations on the structural properties of human betaA3-crystallin. The following eight deletion mutants of betaA3-crystallin were generated: (i) N-terminal extension (NTE) 21 amino acids (betaA3[21] mutant), (ii) NTE 22 amino acids (betaA3[22] mutant), (iii) NTE (betaA3[N] mutant), (iv) NTE plus motif I (betaA3[N+I] mutant), (v) NTE plus motifs I and II (betaA3[N+I+II] mutant), (vi) NTE plus motifs I and II and connecting peptide (betaA3[N+I+II+CP] mutant), (vii) motifs III and IV (betaA3[III+IV] mutant), and (viii) motif IV (betaA3 [IV] mutant). The DNA sequencing and MALDI-TOF mass spectrometric methods confirmed desired specific deletions, and the purified mutant proteins exhibited a single band during SDS-PAGE analysis. When ANS bound, all the mutant proteins exhibited fluorescence quenching and a red shift, suggesting that the truncations caused changes in the exposed hydrophobic patches. The CD spectra showed that deletion of either NTE or the N-terminal domain (motifs I and II) had a relatively weaker effect on the structural stability than deletion of the C-terminal domain (motifs III and IV). Intrinsic Trp fluorescence spectral studies suggested changes in the microenvironment of the mutant proteins following truncations. HPLC multiangle light scattering analyses showed that truncation led to higher-order aggregation compared to that in the wild-type protein. Equilibrium unfolding and refolding of WT betaA3 with urea were best fit to a three-state model with transition midpoints at 2.2 and 3.1 M urea. However, the two transition midpoints of betaA3[21] and betaA3[22] and betaA3[N] mutants were similar to those of the wild type, suggesting that these truncations had a minimal effect on structural stabilization. Further, the mutant proteins containing the N-terminal domain (i.e., betaA3[III+IV] and betaA3[IV] mutants) exhibited higher transition midpoints compared to the transition midpoints of the mutant protein with the C-terminal domain (i.e., betaA3[N+I+II+CP] mutant). The results suggested that the N-terminal domain is relatively more stable than the C-terminal domain in betaA3-crystallin.  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