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 sequence of human myoglobin (Mb) is similar to that of other species except for a unique cysteine at position 110 (Cys(110)). Adding hydrogen peroxide (H(2)O(2)) to human Mb affords Trp(14)-peroxyl, Tyr(103)-phenoxyl, and Cys(110)-thiyl radicals and coupling of Cys(110)-thiyl radicals yields a homodimer through intermolecular disulfide bond formation (Witting, P. K., Douglas, D. J., and Mauk, A. G. (2000) J. Biol. Chem. 275, 20391-20398). Treating a solution of wild type Mb and H(2)O(2) with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) at DMPO:protein /= 100 mol/mol only DMPO-Tyr(103) radicals were present. The DMPO-dependent decrease in DMPO-Cys(110) was matched by a near 1:1 stoichiometric increase in DMPO-Tyr(103). In contrast, reaction of the Y103F human Mb with H(2)O(2) gave no DMPO-Cys(110) at DMPO:protein /= 100 mol/mol (i.e. conditions that consistently gave DMPO-Tyr(103) in the case of wild type Mb). No detectable homodimer was formed by incubation of the Y103F variant with H(2)O(2). However, the homodimer was detected in a mixture of both the Y103F and C110A variants of human Mb upon treatment with H(2)O(2) (C110A:Y103F:H(2)O(2) 2:1:5 mol/mol/mol); the yield of this homodimer increased with increasing ratios of C110A:Y103F. Together, these data suggest that addition of H(2)O(2) to human Mb can produce Cys(110)-thiyl radicals through an intermolecular electron transfer reaction from Cys(110) to a Tyr(103)-phenoxyl radical.  相似文献   

2.
Witting PK  Mauk AG  Lay PA 《Biochemistry》2002,41(38):11495-11503
Myoglobin (Mb) catalyzes a range of oxidation reactions in the presence of hydrogen peroxide (H(2)O(2)) through a peroxidase-like cycle. C110A and Y103F variants of human Mb have been constructed to assess the effects of removing electron-rich oxidizable amino acids from the protein on the peroxidase activity of Mb: a point mutation at W14 failed to yield a viable protein. Point mutations at C110 and Y103 did not result in significant changes to structural elements of the heme pocket, as judged by low-temperature electron paramagnetic spectroscopy (EPR) studies on the ground-state ferric proteins. However, compared to the native protein, the yield of globin radical (globin*) was significantly decreased for the Y103F but not the C110A variant Mb upon reaction of the respective proteins with H(2)O(2). In contrast with our expectation that inhibiting pathways of intramolecular electron transfer may lead to enhanced Mb peroxidase activity, mutation of Y103 marginally decreased the rate constant for reaction of Mb with H(2)O(2) (1.4-fold) as judged by stopped-flow kinetic analyses. Consistent with this decrease in rate constant, steady-state analyses of Y103F Mb-derived thioanisole sulfoxidation indicated decreased V(max) and increased K(m) relative to the wild-type control. Additionally, thioanisole sulfoxidation proceeded with lower stereoselectivity, suggesting that Y103 plays a significant role in substrate binding and orientation in the heme pocket of Mb. Together, these results show that electron transfer within the globin portion of the protein is an important modulator of its stability and catalytic activity. Furthermore, the hydrogen-bonding network involving the residues that line the heme pocket of Mb is crucial to both efficient peroxidase activity and stereospecificity.  相似文献   

3.
The addition of ascorbate to ischemic rat hearts prevents the myocardial damage associated with reoxygenation. H2O2 oxidizes myoglobin (Mb+2) to higher oxidation states (Mb+4 and Mb+5) which are rapidly reduced by ascorbate. It is proposed that the operation of a myoglobin redox cycle, in which H2O2 causes the two-electron oxidation of myoglobin, is a critical determinant of reperfusion injury. Conversely, the reduction of myoglobin, in one-electron steps, may represent an essential protective mechanism against such injury in the heart.  相似文献   

4.
The crucial step in enzymatic transamination is the tautomerization of aldimine/ketimine intermediates, formed between the pyridoxyl coenzyme and the amino/keto acid substrate, which is catalyzed primarily by the active site residue Lys-258 (Malcolm, B. A., and Kirsch, J. F. (1985) Biochem. Biophys. Res. Commun. 132, 915-921; W. L. Finlayson and J. F. Kirsch, in preparation). Tyr-70 is localized in close proximity to Lys-258 and, in addition, forms a hydrogen bond with the coenzyme phosphate. Tyr-70 has been postulated to have an important role in the tautomerization (Kirsch, J. F., Eichele, G., Ford, G. C., Vincent, M. G., Jansonius, J. N., Gehring, H., and Christen, P. (1984) J. Mol. Biol. 174, 497-525). This hypothesis has now been tested by the construction and analysis of a mutant Escherichia coli aspartate aminotransferase in which Tyr-70 has been changed to Phe (Y70F). Y70F retains at least 15% of the maximal activity of the wild type enzyme (WT) (kcat = 170 +/- 15 s-1 for WT versus greater than or equal to 26 +/- 3 s-1 for Y70F and shows increased Michaelis constants for both substrates (KmAsp = 2.5 +/- 0.4 mM; Km alpha Kg = 0.59 +/- 0.08 mM for WT versus KmAsp = 3.9 +/- 0.3 mM; Km alpha Kg = 2.70 +/- 0.02 mM for Y70F (where alpha Kg is alpha-ketoglutarate) ). The spectrophotometrically determined pK a values of the internal aldimines formed between pyridoxal 5'-phosphate (PLP) and Lys-258 are identical for WT and Y70F. In assays where excess L-aspartate and excess PLP are incubated with either WT or Y70F, the mutant enzyme converts the free PLP to free pyridoxamine 5'-phosphate 80-fold faster than WT (k = (3.75 +/- 0.23) X 10(-2)s-1 for Y70F versus (4.90 +/- 0.02) X 10(-4)s-1 for WT). Y70F also converts free pyridoxamine 5'-phosphate to free PLP faster than WT. Thus, Y70F dissociates coenzyme more readily than does WT. It therefore appears that the role of Tyr-70 is mainly in preventing the dissociation of the coenzyme from the enzyme. Tyr-70 does not function in an essential chemical step.  相似文献   

5.
Claudin-5 has recently attracted increasing attention by its potential as a novel treatment target in the early stage of heart failure. However, whether Claudin-5 produces beneficial effects on myocardial ischemia and reperfusion (IR) injury has not been elucidated yet. In this study, we identified reduced levels of Claudin-5 in the hearts of mice subjected to acute myocardial IR injury and murine HL-1 cardiomyocytes subjected to hypoxia and reoxygenation (HR). We then constructed cardiac-specific Cldn5-overexpressing mice using an adeno-associated virus (AAV9) vector and demonstrated that Cldn5 overexpression ameliorated cardiac dysfunction and myocardial damage in mice subjected to myocardial IR injury. Moreover, Cldn5 overexpression attenuated myocardial oxidative stress (DHE and protein levels of Nrf2, HO-1, and NQO1), inflammatory response (levels of MPO, F4/80, Ly6C, and circulating inflammatory cells), mitochondrial dysfunction (protein levels of PGC-1α, NRF1, and TFAM), endoplasmic reticulum stress (protein levels of GRP78, ATF6, and CHOP and p-PERK), energy metabolism disorder (p-AMPK and ACC), and apoptosis (TUNEL assay and protein levels of Bax and Bcl2) in mice subjected to myocardial IR. Next, we generated Cldn5 knockdown cells by lentiviral shRNA and observed that Cldn5 knockdown inhibited cell viability and affected the expression or activation of these IR-related signalings in HL-1 cardiomyocytes subjected to HR. Mechanistically, SIRT1 was proved to be involved in regulating the expression of Claudin-5 by co-immunoprecipitation analysis and Sirt1 knockdown experiments. Our data demonstrated that targeting Claudin-5 may represent a promising approach for preventing and treating acute myocardial IR injury.  相似文献   

6.
Activation of the type 1 histamine (H1) or the type 2 protease-activated (PAR-2) G protein-coupled receptors interrupts E-cadherin adhesion and decreases the transepithelial resistance (TER) of epithelium. Several reports suggest that cadherin adhesive function depends on the association of cadherin with beta-catenin and that this association is regulated by phosphorylation of tyrosines in beta-catenin. We tested the hypothesis that loss of cadherin adhesion and compromise of TER on activation of the H1 or PAR-2 receptor is due to phosphorylation of tyrosines in beta-catenin. L cells were stably transfected to express E-cadherin (L-E-cad cells) and H1 (L-H1-E-cad cells). L cells and Madin-Darby canine kidney (MDCK) cells constitutively express PAR-2. Stably transfected L-E-cad, L-H1-E-cad, and MDCK cells were also stably transfected with FLAG-tagged wild-type (WT) or mutant beta-catenin, converting tyrosine 142, 489, or 654 to the nonphosphorylatable mimetic, phenylalanine (WT, Y142F, Y489F, or Y654F). Activation of H1 or PAR-2 interrupted adhesion to an immobilized E-cadherin-Fc fusion protein of L-H1-E-cad, L-E-cad, and MDCK cells expressing WT or Y142F beta-catenin but did not interrupt adhesion of L-H1-E-cad, L-E-cad, and MDCK cells expressing the Y489F or Y654F mutant beta-catenins. PAR-2 activation decreased the TER of monolayers of MDCK cells expressing WT or Y142F beta-catenin 40-45%. However, PAR-2 activation did not decrease the TER of monolayers of MDCK cells expressing Y489F or Y654F beta-catenin. The protein tyrosine phosphatase PTP1B binds to the cadherin cytoplasmic domain and dephosphorylates beta-catenin. Inhibition of PTP1B interrupted adhesion to E-cadherin-Fc of MDCK cells expressing WT beta-catenin but did not affect the adhesion of MDCK cells expressing Y489F or Y654F beta-catenin. Similarly, inhibition of PTP1B compromised the TER of MDCK cells expressing WT beta-catenin but did not affect the TER of MDCK cells expressing Y489F or Y654F beta-catenin. We conclude that phosphorylation of tyrosines 489 and 654 in beta-catenin is a necessary step in the process by which G protein-coupled H1 and PAR-2 receptors interrupt E-cadherin adhesion. We also conclude that activation of PAR-2 has no effect on the TER without first interrupting E-cadherin adhesion.  相似文献   

7.
Oxidative stress plays a pivotal role in ischemic injury, and p66(ShcA)ko mice exhibit both lower oxidative stress and decreased tissue damage following hind limb ischemia. Thus, it was investigated whether tissue regeneration following acute hind limb ischemia was altered in p66(ShcA)ko mice. Upon femoral artery dissection, muscle regeneration started earlier and was completed faster than in wild-type (WT) control. Moreover, faster regeneration was associated with decreased oxidative stress. Unlike ischemia, cardiotoxin injury induced similar skeletal muscle damage in both genotypes. However, p66(ShcA)ko mice regenerated faster, in agreement with the regenerative advantage upon ischemia. Since no difference between p66(ShcA)wt and knock-out (ko) mice was found in blood perfusion recovery after ischemia, satellite cells (SCs), a resident population of myogenic progenitors, were examined. Similar SCs numbers were present in WT and ko mice. However, in vitro cultured p66(ShcA)ko SCs displayed lower oxidative stress levels and higher proliferation rate and differentiated faster than WT. Furthermore, when exposed to sublethal H(2)O(2) doses, p66(ShcA)ko SCs were resistant to H(2)O(2)-induced inhibition of differentiation. Finally, myogenic conversion induced by MyoD overexpression was more efficient in p66(ShcA)ko fibroblasts compared with WT. The present work demonstrates that oxidative stress and p66(ShcA) play a crucial role in the regenerative pathways activated by acute ischemia.  相似文献   

8.
Classical molecular dynamics simulations were utilized to investigate the structural and dynamical properties of water in the active site of ketosteroid isomerase (KSI) to provide insight into the role of these water molecules in the enzyme-catalyzed reaction. This reaction is thought to proceed via a dienolate intermediate that is stabilized by hydrogen bonding with residues Tyr16 and Asp103. A comparative study was performed for the wild-type (WT) KSI and the Y16F, Y16S, and Y16F/Y32F/Y57F (FFF) mutants. These systems were studied with three different bound ligands: equilenin, which is an intermediate analog, and the intermediate states of two steroid substrates. Several distinct water occupation sites were identified in the active site of KSI for the WT and mutant systems. Three additional sites were identified in the Y16S mutant that were not occupied in WT KSI or the other mutants studied. The number of water molecules directly hydrogen bonded to the ligand oxygen was approximately two in the Y16S mutant and one in the Y16F and FFF mutants, with intermittent hydrogen bonding of one water molecule in WT KSI. The molecular dynamics trajectories of the Y16F and FFF mutants reproduced the small conformational changes of residue 16 observed in the crystal structures of these two mutants. Quantum mechanical/molecular mechanical calculations of (1)H NMR chemical shifts of the protons in the active site hydrogen-bonding network suggest that the presence of water in the active site does not prevent the formation of short hydrogen bonds with far-downfield chemical shifts. The molecular dynamics simulations indicate that the active site water molecules exchange much more frequently for WT KSI and the FFF mutant than for the Y16F and Y16S mutants. This difference is most likely due to the hydrogen-bonding interaction between Tyr57 and an active site water molecule that is persistent in the Y16F and Y16S mutants but absent in the FFF mutant and significantly less probable in WT KSI.  相似文献   

9.
Recent studies suggest that superoxide dismutase (SOD1) may represent a major target of oxidative damage in neurodegenerative diseases. To test the possibility that oxidized species of wild-type (WT) SOD1 might be involved in pathogenic processes, we analyzed the properties of the WT human SOD1 protein after its oxidation in vivo or in vitro by hydrogen peroxide (H2O2) treatment. Using transfected Neuro2a cells expressing WT or amyotrophic lateral sclerosis-linked SOD1 species, we show that exposure to H2O2 modifies the properties of WT SOD1. Western blot analysis of immunoprecipitates from cell lysates revealed that, like mutant SOD1, oxidized WT SOD1 can be conjugated with poly-ubiquitin and can interact with Hsp70. Chromogranin B, a neurosecretory protein that interacts with mutant SOD1 but not with WT SOD1, was co-immunoprecipitated with oxidized WT SOD1 from lysates of Neuro2a cells treated with H2O2. Treatment of microglial cells (line BV2) with either oxidized WT SOD1 or mutant SOD1 recombinant proteins induced tumor necrosis factor-alpha and inducible nitric oxide synthase. Furthermore, exposure of cultured motor neurons to oxidized WT SOD1 caused dose-dependent cell death like mutant SOD1 proteins. These results suggest that WT SOD1 may acquire binding and toxic properties of mutant forms of SOD1 through oxidative damage.  相似文献   

10.
Reaction of sperm whale metmyoglobin (SwMb) with H2O2 produces a ferryl (MbFeIV=O) species and a protein radical and leads to the formation of oligomeric products. The ferryl species is maximally formed with one equivalent of H2O2, and the maximum yields of the dimer (28%) and trimer (17%) with 1 or 2 eq. Co-incubation of the SwMb Y151F mutant with native apoSwMb and H2O2 produced dimeric products, which requires radical transfer from the nondimerizing Y151F mutant to apoSwMb. Autoreduction of ferryl SwMb to the ferric state is biphasic with t = 3.4 and 25.9 min. An intramolecular autoreduction process is implicated at low protein concentrations, but oligomerization decreases the lifetime of the ferryl species at high protein concentrations. A fraction of the protein remained monomeric. This dimerization-resistant protein was in the ferryl state, but after autoreduction it underwent normal dimerization with H2O2. Proteolytic digestion established the presence of both dityrosine and isodityrosine cross-links in the oligomeric proteins, with the isodityrosine links primarily forged by Tyr151-Tyr151 coupling. The tyrosine content decreased by 47% in the dimer and 14% in the recovered monomer, but the yields of isodityrosine and dityrosine in the dimer were only 15.2 and 6.8% of the original tyrosine content. Approximately 23% of the lost tyrosines therefore have an alternative but unknown fate. The results clearly demonstrate the concurrence of intra- and intermolecular electron transfer processes involving Mb protein radicals. Intermolecular electron transfers that generate protein radicals on bystander proteins are likely to propagate the cellular damage initiated by the reaction of metalloproteins with H2O2.  相似文献   

11.
The catalase reaction has been studied in detail by using myoglobin (Mb) mutants. Compound I of Mb mutants (Mb-I), a ferryl species (Fe(IV)=O) paired with a porphyrin radical cation, is readily prepared by the reaction with a nearly stoichiometric amount of m-chloroperbenzoic acid. Upon the addition of H2O2 to an Mb-I solution, Mb-I is reduced back to the ferric state without forming any intermediates. This indicates that Mb-I is capable of performing two-electron oxidation of H2O2 (catalatic reaction). Gas chromatography-mass spectroscopy analysis of the evolved O2 from a 50:50 mixture of H2(18)O2/H2(16)O2 solution containing H64D or F43H/H64L Mb showed the formation of 18O2 (m/e = 36) and 16O2 (m/e = 32) but not 16O18O (m/e = 34). This implies that O2 is formed by two-electron oxidation of H2O2 without breaking the O-O bond. Deuterium isotope effects on the catalatic reactions of Mb mutants and catalase suggest that the catalatic reactions of Micrococcus lysodeikticus catalase and F43H/H64L Mb proceed via an ionic mechanism with a small isotope effect of less than 4.0, since the distal histidine residue is located at a proper position to act as a general acid-base catalyst for the ionic reaction. In contrast, other Mb mutants such as H64X (X is Ala, Ser, and Asp) and L29H/H64L Mb oxidize H2O2 via a radical mechanism in which a hydrogen atom is abstracted by Mb-I with a large isotope effect in a range of 10-29, due to a lack of the general acid-base catalyst.  相似文献   

12.
The macrophage scavenger receptor class A (SR-A) participates in the innate immune and inflammatory responses. This study examined the role of macrophage SR-A in myocardial ischemia/reperfusion (I/R) injury and hypoxia/reoxygenation (H/R)-induced cell damage. SR-A?/? and WT mice were subjected to ischemia (45 min) followed by reperfusion for up to 7 days. SR-A?/? mice showed smaller myocardial infarct size and better cardiac function than did WT I/R mice. SR-A deficiency attenuated I/R-induced myocardial apoptosis by preventing p53-mediated Bak-1 apoptotic signaling. The levels of microRNA-125b in SR-A?/? heart were significantly greater than in WT myocardium. SR-A is predominantly expressed on macrophages. To investigate the role of SR-A macrophages in H/R-induced injury, we isolated peritoneal macrophages from SR-A deficient (SR-A?/?) and wild type (WT) mice. Macrophages were subjected to hypoxia followed by reoxygenation. H/R markedly increased NF-κB binding activity as well as KC and MCP-1 production in WT macrophages but not in SR-A?/? macrophages. H/R induced caspase-3/7 and -8 activities and cell death in WT macrophages, but not in SR-A?/? macrophages. The levels of miR-125b in SR-A?/? macrophages were significantly higher than in WT macrophages. Transfection of WT macrophages with miR-125b mimics attenuated H/R-induced caspase-3/7 and -8 activities and H/R-decreased viability, and prevented H/R-increased p-53, Bak-1 and Bax expression. The data suggest that SR-A deficiency attenuates myocardial I/R injury by targeting p53-mediated apoptotic signaling. SR-A?/? macrophages contain high levels of miR-125b which may play a role in the protective effect of SR-A deficiency on myocardial I/R injury and H/R-induced cell damage.  相似文献   

13.
Hydroxyl radicals (*OH) are involved in the pathogenesis of ischemia-reperfusion injury and are observed in clinical situations, including acute heart failure, stroke, and myocardial infarction. Acute transient exposure to *OH causes an intracellular Ca(2+) overload and leads to impaired contractility. We investigated whether upregulation of sarcoplasmic reticulum Ca(2+)-ATPase function (SERCA) can attenuate *OH-induced dysfunction. Small, contracting right ventricular papillary muscles from wild-type (WT) SERCA1a-overexpressing (transgenic, TG) and SERCA2a heterogeneous knockout (HET) mice were directly exposed to *OH. This brief 2-min exposure led to a transient elevation of diastolic force (F(dia)) and depression of developed force (F(dev)). In WT mice, F(dia) increased to 485 +/- 49% and F(dev) decreased to 11 +/- 3%. In sharp contrast, in TG mice F(dia) increased only to 241 +/- 17%, whereas F(dev) decreased only to 51 +/- 5% (P < 0.05 vs. WT). In HET mice, F(dia) rose more than WT (to 597 +/- 20%, P < 0.05), whereas F(dev) was reduced in a similar amount. After approximately 45 min after *OH exposure, a new steady state was reached: F(dev) returned to 37 +/- 6% and 32 +/- 6%, whereas F(dia) came back to 238 +/- 28% and 292 +/- 17% in WT and HET mice, respectively. In contrast, the sustained dysfunction was significantly less in TG mice: F(dia) and F(dev) returned to 144 +/- 20% and 67 +/- 6%, respectively. Before exposure to *OH, there is decrease in phospholamban (PLB) phosphorylation at Ser16 (pPLBSer16) and PLB phosphorylation at Thr17 (pPLBThr17) in TG mice and an increase in pPLBSer16 and pPLBThr17 in HET mice versus WT. After exposure to *OH there is decrease in pPLBSer16 in WT, TG, and HET mice but no significant change in the level of pPLBThr17 in any group. The results indicate that SERCA overexpression can reduce the *OH-induced contractile dysfunction in murine myocardium, whereas a reduced SR Ca(2+)-ATPase activity aggravates this injury. Loss of pPLB levels at Ser16 likely amplifies the differences observed in injury response.  相似文献   

14.
Lee LV  Poyner RR  Vu MV  Cleland WW 《Biochemistry》2000,39(16):4821-4830
H97N, H95N, and Y229F mutants of L-ribulose-5-phosphate 4-epimerase had 10, 1, and 0.1%, respectively, of the activity of the wild-type (WT) enzyme when activated by Zn(2+), the physiological activator. Co(2+) and Mn(2+) replaced Zn(2+) in Y229F and WT enzymes, although less effectively with the His mutants, while Mg(2+) was a poorly bound, weak activator. None of the other eight tyrosines mutated to phenylalanine caused a major loss of activity. The near-UV CD spectra of all enzymes were nearly identical in the absence of metal ions and substrate, and addition of substrate without metal ion showed no effect. When both substrate and Zn(2+) were present, however, the positive band at 266 nm increased while the negative one at 290 nm decreased in ellipticity. The changes for the WT and Y229F enzymes were greater than for the two His mutants. With Co(2+) as the metal ion, the CD and absorption spectra in the visible region were different, showing little ellipticity in the absence of substrate and a weak absorption band at 508 nm. With substrate present, however, an intense absorption band at 555 nm (epsilon = 150-175) with a negative molar ellipticity approaching 2000 deg cm(2) dmol(-1) appears with WT and Y229F enzymes. With the His mutants, the changes induced by substrate were smaller, with negative ellipticity only half as great. The WT, Y229F, H95N, and H97N enzymes all catalyze a slow aldol condensation of dihydroxyacetone and glycolaldehyde phosphate with an initial k(cat) of 1.6 x 10(-3) s(-1). The initial rate slowed most rapidly with WT and H97N enzymes, which have the highest affinity for the ketopentose phosphates formed in the condensation. The EPR spectrum of enzyme with Mn(2+) exhibited a drastic decrease upon substrate addition, and by using H(2)(17)O, it was determined that there were three waters in the coordination sphere of Mn(2+) in the absence of substrate. These data suggest that (1) the substrate coordinates to the enzyme-bound metal ion, (2) His95 and His97 are likely metal ion ligands, and (3) Tyr229 is not a metal ion ligand, but may play another role in catalysis, possibly as an acid-base catalyst.  相似文献   

15.
Replacement of phenylalanine 87 with alanine or glycine (mutant F87A or F87G) greatly increased the H2O2-supported substrate hydroxylation activity of cytochrome P450BM-3, whose original H2O2-supported activity is hardly detectable. On the other hand, replacement of phenylalanine 87 with valine (mutant F87V) did not. In the oxidation of p-nitrophenoxydodecanoic acid (12-pNCA), the turnover numbers of the mutant F87A in the presence of NADPH and O2, or H2O2 were 493 and 162 nmol/min/nmol, respectively. The H2O2-supported F87A hydroxylation activity was further confirmed with free fatty acids as substrates. Moreover, the stability of F87A in H2O2 solutions also largely increased. The order of the stability of the wild type (WT), F87A, and their substrate (12-pNCA)-binding complexes in H2O2 solutions listed from high to low was F87A, WT, F87A substrate-binding complex, and WT substrate-binding complex. We propose that the free space size in the vicinity of the heme iron significantly influences P450BM-3 H2O2-supported activity and H2O2 inactivation.  相似文献   

16.
The cellular prion protein (PrPC), a protein most noted for its link to prion diseases, has been found to play a protective role in ischemic brain injury. To investigate the role of PrPC in the kidney, an organ highly prone to ischemia/reperfusion (IR) injury, we examined wild-type (WT) and PrPC knockout (KO) mice that were subjected to 30-min of renal ischemia followed by 1, 2, or 3 days of reperfusion. Renal dysfunction and structural damage was more severe in KO than in WT mice. While PrP was undetectable in KO kidneys, Western blotting revealed an increase in PrP in IR-injured WT kidneys compared to sham-treated kidneys. Compared to WT, KO kidneys exhibited increases in oxidative stress markers heme oxygenase-1, nitrotyrosine, and Nε-(carboxymethyl)lysine, and decreases in mitochondrial complexes I and III. Notably, phosphorylated extracellular signal-regulated kinase (pERK) staining was predominantly observed in tubular cells from KO mice following 2 days of reperfusion, a time at which significant differences in renal dysfunction, histological changes, oxidative stress, and mitochondrial complexes between WT and KO mice were observed. Our study provides the first evidence that PrPC may play a protective role in renal IR injury, likely through its effects on mitochondria and ERK signaling pathways.  相似文献   

17.
Cardiomyocyte loss is the main cause of myocardial dysfunction following an ischemia-reperfusion (IR) injury. Mitochondrial dysfunction and altered mitochondrial network dynamics play central roles in cardiomyocyte death. Proteasome inhibition is cardioprotective in the setting of IR; however, the mechanisms underlying this protection are not well-understood. Several proteins that regulate mitochondrial dynamics and energy metabolism, including Mitofusin-2 (Mfn2), are degraded by the proteasome. The aim of this study was to evaluate whether proteasome inhibition can protect cardiomyocytes from IR damage by maintaining Mfn2 levels and preserving mitochondrial network integrity. Using ex vivo Langendorff-perfused rat hearts and in vitro neonatal rat ventricular myocytes, we showed that the proteasome inhibitor MG132 reduced IR-induced cardiomyocyte death. Moreover, MG132 preserved mitochondrial mass, prevented mitochondrial network fragmentation, and abolished IR-induced reductions in Mfn2 levels in heart tissue and cultured cardiomyocytes. Interestingly, Mfn2 overexpression also prevented cardiomyocyte death. This effect was apparently specific to Mfn2, as overexpression of Miro1, another protein implicated in mitochondrial dynamics, did not confer the same protection. Our results suggest that proteasome inhibition protects cardiomyocytes from IR damage. This effect could be partly mediated by preservation of Mfn2 and therefore mitochondrial integrity.  相似文献   

18.
H2O2诱导的线粒体损伤神经元内硫氧还蛋白mRNA水平的变化   总被引:3,自引:0,他引:3  
线粒体缺陷和氧化应激参与了神经退行性疾病的发病机制。叠氮钠(NaN3)是线粒体细胞色素C氧化酶(COX)的特异性抑制剂,能诱导线粒体缺陷。本实验通过细胞活性检测(MTT法),形态学观察,分析H2O2对原代培养的正常神经元及NaN3诱导的线粒体缺陷神经元的损伤作用的差异。并通过RT-PCR半定量法检测H2O2损伤后两类神经元内硫氧还蛋白(Thioredoxin,Trx)mRNA水平的变化,以阐明细胞内这一重要氧化还原调节蛋白在神经元损伤时的作用机制。实验表明,在正常神经元内,H2O2的损伤对Trx表达量的改变似乎不明显;而线粒体缺陷神经元内Trx的表达量下降,且对于H2O2的损伤具有浓度、时间依赖性。提示:在线粒体功能缺陷神经元中,Trx似乎发挥更重要的作用。  相似文献   

19.
Ferric myoglobin undergoes a two-electron oxidation in its reaction with H(2)O(2). One oxidation equivalent is used to oxidize Fe(III) to the Fe(IV) ferryl species, while the second is associated with a protein radical but is rapidly dissipated. The ferryl species is then slowly reduced back to the ferric state by unknown mechanisms. To clarify this process, the formation and stability of the ferryl forms of the Tyr --> Phe and Trp --> Phe mutants of recombinant sperm whale myoglobin (SwMb) were investigated. Kinetic studies showed that all the mutants react normally with H(2)O(2) to give the ferryl species. However, the rapid phase of ferryl autoreduction typical of wild-type SwMb was absent in the triple Tyr --> Phe mutant and considerably reduced in the Y103F and Y151F mutants, strongly implicating these two residues as intramolecular electron donors. Replacement of Tyr146, Trp7, or Trp14 did not significantly alter the autoreduction, indicating that these residues do not contribute to ferryl reduction despite the fact that Tyr146 is closer to the iron than Tyr151 or Tyr103. Furthermore, analysis of the fast phase of autoreduction in the dimer versus recovered monomer of the Tyr --> Phe mutant K102Q/Y103F/Y146F indicates that the Tyr151-Tyr151 cross-link is a particularly effective electron donor. The presence of an additional, slow phase of reduction in the triple Tyr --> Phe mutant indicates that alternative but normally minor electron-transfer pathways exist in SwMb. These results demonstrate that internal electron transfer is governed as much by the tyrosine pK(a) and oxidation potential as by its distance from the electron accepting iron atom.  相似文献   

20.
Myoglobin (Mb) is used as a model system for other heme proteins and the reactions they catalyze. The latest novel function to be proposed for myoglobin is a P450 type hydroxylation activity of aromatic carbons (Watanabe, Y., and Ueno, T. (2003) Bull. Chem. Soc. Jpn. 76, 1309-1322). Because Mb does not contain a specific substrate binding site for aromatic compounds near the heme, an engineered tryptophan in the heme pocket was used to model P450 hydroxylation of aromatic compounds. The monooxygenation product was not previously isolated because of rapid subsequent oxidation steps (Hara, I., Ueno, T., Ozaki, S., Itoh, S., Lee, K., Ueyama, N., and Watanabe, Y. (2001) J. Biol. Chem. 276, 36067-36070). In this work, a Mb variant (F43W/H64D/V68I) is used to characterize the monooxygenated intermediate. A modified (+16 Da) species forms upon the addition of 1 eq of H2O2. This product was digested with chymotrypsin, and the modified peptide fragments were isolated and characterized as 6-hydroxytryptophan using matrix-assisted laser desorption ionization time-of-flight tandem mass spectroscopy and 1H NMR. This engineered Mb variant represents the first enzyme to preferentially hydroxylate the indole side chain of Trp at the C6 position. Finally, heme extraction was used to demonstrate that both the formation of the 6-hydroxytryptophan intermediate (+16 Da) and subsequent oxidation to form the +30 Da final product are catalyzed by the heme cofactor, most probably via the compound I intermediate. These results provide insight into the mechanism of hydroxylation of aromatic carbons by heme proteins, demonstrating that non-thiolate-ligated heme enzymes can perform this function. This establishes Mb compound I as a model for P450 type aromatic hydroxylation chemistry.  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