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.
2.
The C-terminal domain of subunit ε of the bacterial FoF1 ATP synthase is reported to be an intrinsic inhibitor of ATP synthesis/hydrolysis activity in vitro, preventing wasteful hydrolysis of ATP under low-energy conditions. Mutants defective in this regulatory domain exhibited no significant difference in growth rate, molar growth yield, membrane potential, or intracellular ATP concentration under a wide range of growth conditions and stressors compared to wild-type cells, suggesting this inhibitory domain is dispensable for growth and survival of Escherichia coli.FoF1 ATP synthases are ubiquitous enzymes that synthesize ATP using a transmembrane electrochemical potential of protons or proton motive force (PMF) generated by the respiratory chain across the cytoplasmic membrane of bacteria, the thylakoid membrane of chloroplasts, or the mitochondrial inner membrane (4, 5, 37). The enzyme consists of two parts: membrane-embedded Fo subcomplex (a complex of subunits a, b, and c in bacteria) and hydrophilic F1 subcomplex (composed of subunits α, β, γ, δ, and ε). The enzyme is also known as a molecular motor, which is composed of the stator subcomplex (α, β, δ, a, and b) and the rotor subcomplex (γ, ε, and c), and its rotation is coupled to ATP synthesis and proton flow across the membrane (20, 31, 52). The reaction of the enzyme is reversible; ATP is hydrolyzed into ADP and inorganic phosphate, the rotor subcomplex rotates in reverse, and protons are extruded to the periplasmic side, resulting in the generation of PMF. Although some bacteria utilize the reverse reaction under particular conditions, the primary function of FoF1 ATP synthase is generation of ATP from the PMF. Therefore, the direction of the activity of FoF1 ATP synthase is regulated to avoid wasteful ATP hydrolysis.Subunit ε in bacterial FoF1 has been known to be an intrinsic inhibitor of F1 and FoF1 complex (18, 21, 23) and is proposed to have a regulatory function (10, 11, 42). Although the inhibitory effects of subunit ε vary among species, in general, ε inhibits ATP hydrolysis activity while repressing ATP synthesis activity to a lesser degree (14, 27). This regulatory function of the ε subunit is mediated almost exclusively by the C-terminal region of ε, which is comprised of two antiparallel α-helices (18, 49, 50). Biochemical and crystallographic studies have revealed that the C-terminal helices can adopt two different conformations (34, 46, 47, 48). In the retracted conformation, the α-helices form a hairpin-like structure and sit on the N-terminal β-sandwich domain of the ε subunit. When the ε subunit exhibits an inhibitory effect, it adopts a more extended conformation in which the C-terminal α-helices extend along the γ subunit, which composes the central stalk. It has also been shown that basic, positively charged residues on the second α-helix of the ε subunit interact with negatively charged residues in the DELSEED segment of subunit β to exert the inhibitory effect (12).Escherichia coli mutants deleted in the entire ε subunit exhibit a reduced growth rate and growth yield, and this effect is proposed to be a result of a deficiency in assembly of the Fo and F1 complexes (21). The N-terminal β-sandwich domain of the ε subunit is responsible for the assembly of Fo and F1 and is therefore important for efficient coupling between proton translocation through Fo and ATP synthesis/hydrolysis in F1 (15, 39). Deletion of the ε subunit leads to dissociation of the FoF1 complex and wasteful ATP hydrolysis by free (cytoplasmic) F1 and dissipation of PMF through free Fo (21, 22, 51).While the importance of the entire ε subunit in the whole-cell physiology of E. coli is fairly well established, the role of the regulatory C-terminal region of ε has received little attention and warrants investigation to determine if the regulatory functions (e.g., inhibition of ATP hydrolysis) observed in vitro are manifested in the physiology of E. coli under various growth conditions. To address this question, we constructed isogenic E. coli mutants that were deleted in the C-terminal region of ε subunit (εDC) and used these strains to compare physiological properties of wild-type versus εDC cells under a wide range of environmental conditions and stressors.  相似文献   

3.
4.
The present study tests the hypothesis that the structure of extracellular domain Loop 2 can markedly affect ethanol sensitivity in glycine receptors (GlyRs) and γ-aminobutyric acid type A receptors (GABAARs). To test this, we mutated Loop 2 in the α1 subunit of GlyRs and in the γ subunit of α1β2γ2GABAARs and measured the sensitivity of wild type and mutant receptors expressed in Xenopus oocytes to agonist, ethanol, and other agents using two-electrode voltage clamp. Replacing Loop 2 of α1GlyR subunits with Loop 2 from the δGABAAR (δL2), but not the γGABAAR subunit, reduced ethanol threshold and increased the degree of ethanol potentiation without altering general receptor function. Similarly, replacing Loop 2 of the γ subunit of GABAARs with δL2 shifted the ethanol threshold from 50 mm in WT to 1 mm in the GABAA γ-δL2 mutant. These findings indicate that the structure of Loop 2 can profoundly affect ethanol sensitivity in GlyRs and GABAARs. The δL2 mutations did not affect GlyR or GABAAR sensitivity, respectively, to Zn2+ or diazepam, which suggests that these δL2-induced changes in ethanol sensitivity do not extend to all allosteric modulators and may be specific for ethanol or ethanol-like agents. To explore molecular mechanisms underlying these results, we threaded the WT and δL2 GlyR sequences onto the x-ray structure of the bacterial Gloeobacter violaceus pentameric ligand-gated ion channel homologue (GLIC). In addition to being the first GlyR model threaded on GLIC, the juxtaposition of the two structures led to a possible mechanistic explanation for the effects of ethanol on GlyR-based on changes in Loop 2 structure.Alcohol abuse and dependence are significant problems in our society, with ∼14 million people in the United States being affected (1, 2). Alcohol causes over 100,000 deaths in the United States, and alcohol-related issues are estimated to cost nearly 200 billion dollars annually (2). To address this, considerable attention has focused on the development of medications to prevent and treat alcohol-related problems (35). The development of such medications would be aided by a clear understanding of the molecular structures on which ethanol acts and how these structures influence receptor sensitivity to ethanol.Ligand-gated ion channels (LGICs)2 have received substantial attention as putative sites of ethanol action that cause its behavioral effects (612). Research in this area has focused on investigating the effects of ethanol on two large superfamilies of LGICs: 1) the Cys-loop superfamily of LGICs (13, 14), whose members include nicotinic acetylcholine, 5-hydroxytryptamine3, γ-aminobutyric acid type A (GABAA), γ-aminobutyric acid type C, and glycine receptors (GlyRs) (10, 11, 1520) and 2) the glutamate superfamily, including N-methyl d-aspartate, α-amino-3-hydroxyisoxazolepropionic acid, and kainate receptors (21, 22). Recent studies have also begun investigating ethanol action in the ATP-gated P2X superfamily of LGICs (2325).A series of studies that employed chimeric and mutagenic strategies combined with sulfhydryl-specific labeling identified key regions within Cys-loop receptors that appear to be initial targets for ethanol action that also can determine the sensitivity of the receptors to ethanol (712, 18, 19, 2630). This work provides several lines of evidence that position 267 and possibly other sites in the transmembrane (TM) domain of GlyRs and homologous sites in GABAARs are targets for ethanol action and that mutations at these sites can influence ethanol sensitivity (8, 9, 26, 31).Growing evidence from GlyRs indicates that ethanol also acts on the extracellular domain. The initial findings came from studies demonstrating that α1GlyRs are more sensitive to ethanol than are α2GlyRs despite the high (∼78%) sequence homology between α1GlyRs and α2GlyRs (32). Further work found that an alanine to serine exchange at position 52 (A52S) in Loop 2 can eliminate the difference in ethanol sensitivity between α1GlyRs and α2GlyRs (18, 20, 33). These studies also demonstrated that mutations at position 52 in α1GlyRS and the homologous position 59 in α2GlyRs controlled the sensitivity of these receptors to a novel mechanistic ethanol antagonist (20). Collectively, these studies suggest that there are multiple sites of ethanol action in α1GlyRs, with one site located in the TM domain (e.g. position 267) and another in the extracellular domain (e.g. position 52).Subsequent studies revealed that the polarity of the residue at position 52 plays a key role in determining the sensitivity of GlyRs to ethanol (20). The findings with polarity in the extracellular domain contrast with the findings at position 267 in the TM domain, where molecular volume, but not polarity, significantly affected ethanol sensitivity (9). Taken together, these findings indicate that the physical-chemical parameters of residues at positions in the extracellular and TM domains that modulate ethanol effects and/or initiate ethanol action in GlyRs are not uniform. Thus, knowledge regarding the physical-chemical properties that control agonist and ethanol sensitivity is key for understanding the relationship between the structure and the actions of ethanol in LGICs (19, 31, 3440).GlyRs and GABAARs, which differ significantly in their sensitivities to ethanol, offer a potential method for identifying the structures that control ethanol sensitivity. For example, α1GlyRs do not reliably respond to ethanol concentrations less than 10 mm (32, 33, 41). Similarly, γ subunit-containing GABAARs (e.g. α1β2γ2), the most predominantly expressed GABAARs in the central nervous system, are insensitive to ethanol concentrations less than 50 mm (42, 43). In contrast, δ subunit-containing GABAARs (e.g. α4β3δ) have been shown to be sensitive to ethanol concentrations as low as 1–3 mm (4451). Sequence alignment of α1GlyR, γGABAAR, and δGABAAR revealed differences between the Loop 2 regions of these receptor subunits. Since prior studies found that mutations of Loop 2 residues can affect ethanol sensitivity (19, 20, 39), the non-conserved residues in Loop 2 of GlyR and GABAAR subunits could provide the physical-chemical and structural bases underlying the differences in ethanol sensitivity between these receptors.The present study tested the hypothesis that the structure of Loop 2 can markedly affect the ethanol sensitivity of GlyRs and GABAARs. To accomplish this, we performed multiple mutations that replaced the Loop 2 region of the α1 subunit in α1GlyRs and the Loop 2 region of the γ subunit of α1β2γ2 GABAARs with corresponding non-conserved residues from the δ subunit of GABAAR and tested the sensitivity of these receptors to ethanol. As predicted, replacing Loop 2 of WT α1GlyRs with the homologous residues from the δGABAAR subunit (δL2), but not the γGABAAR subunit (γL2), markedly increased the sensitivity of the receptor to ethanol. Similarly, replacing the non-conserved residues of the γ subunit of α1β2γ2 GABAARs with δL2 also markedly increased ethanol sensitivity of GABAARs. These findings support the hypothesis and suggest that Loop 2 may play a role in controlling ethanol sensitivity across the Cys-loop superfamily of receptors. The findings also provide the basis for suggesting structure-function relationships in a new molecular model of the GlyR based on the bacterial Gloeobacter violaceus pentameric LGIC homologue (GLIC).  相似文献   

5.
6.
Codon optimization was used to synthesize the blh gene from the uncultured marine bacterium 66A03 for expression in Escherichia coli. The expressed enzyme cleaved β-carotene at its central double bond (15,15′) to yield two molecules of all-trans-retinal. The molecular mass of the native purified enzyme was ∼64 kDa as a dimer of 32-kDa subunits. The Km, kcat, and kcat/Km values for β-carotene as substrate were 37 μm, 3.6 min−1, and 97 mm−1 min−1, respectively. The enzyme exhibited the highest activity for β-carotene, followed by β-cryptoxanthin, β-apo-4′-carotenal, α-carotene, and γ-carotene in decreasing order, but not for β-apo-8′-carotenal, β-apo-12′-carotenal, lutein, zeaxanthin, or lycopene, suggesting that the presence of one unsubstituted β-ionone ring in a substrate with a molecular weight greater than C35 seems to be essential for enzyme activity. The oxygen atom of retinal originated not from water but from molecular oxygen, suggesting that the enzyme was a β-carotene 15,15′-dioxygenase. Although the Blh protein and β-carotene 15,15′-monooxygenases catalyzed the same biochemical reaction, the Blh protein was unrelated to the mammalian β-carotene 15,15′-monooxygenases as assessed by their different properties, including DNA and amino acid sequences, molecular weight, form of association, reaction mechanism, kinetic properties, and substrate specificity. This is the first report of in vitro characterization of a bacterial β-carotene-cleaving enzyme.Vitamin A (retinol) is a fat-soluble vitamin and important for human health. In vivo, the cleavage of β-carotene to retinal is an important step of vitamin A synthesis. The cleavage can proceed via two different biochemical pathways (1, 2). The major pathway is a central cleavage catalyzed by mammalian β-carotene 15,15′-monooxygenases (EC 1.14.99.36). β-Carotene is cleaved by the enzyme symmetrically into two molecules of all-trans-retinal, and retinal is then converted to vitamin A in vivo (35). The second pathway is an eccentric cleavage that occurs at double bonds other than the central 15,15′-double bond of β-carotene to produce β-apo-carotenals with different chain lengths, which are catalyzed by carotenoid oxygenases from mammals, plants, and cyanobacteria (6). These β-apo-carotenals are degraded to one molecule of retinal, which is subsequently converted to vitamin A in vivo (2).β-Carotene 15,15′-monooxygenase was first isolated as a cytosolic enzyme by identifying the product of β-carotene cleavage as retinal (7). The characterization of the enzyme and the reaction pathway from β-carotene to retinal were also investigated (4, 8). The enzyme activity has been found in mammalian intestinal mucosa, jejunum enterocytes, liver, lung, kidney, and brain (5, 9, 10). Molecular cloning, expression, and characterization of β-carotene 15,15′-monooxygenase have been reported from various species, including chickens (11), fruit flies (12), humans (13), mice (14), and zebra fishes (15).Other proteins thought to convert β-carotene to retinal include bacterioopsin-related protein (Brp) and bacteriorhodopsin-related protein-like homolog protein (Blh) (16). Brp protein is expressed from the bop gene cluster, which encodes the structural protein bacterioopsin, consisting of at least three genes as follows: bop (bacterioopsin), brp (bacteriorhodopsin-related protein), and bat (bacterioopsin activator) (17). brp genes were reported in Haloarcula marismortui (18), Halobacterium sp. NRC-1 (19), Halobacterium halobium (17), Haloquadratum walsbyi, and Salinibacter ruber (20). Blh protein is expressed from the proteorhodopsin gene cluster, which contains proteorhodopsin, crtE (geranylgeranyl-diphosphate synthase), crtI (phytoene dehydrogenase), crtB (phytoene synthase), crtY (lycopene cyclase), idi (isopentenyl diphosphate isomerase), and blh gene (21). Sources of blh genes were previously reported in Halobacterium sp. NRC-1 (19), Haloarcula marismortui (18), Halobacterium salinarum (22), uncultured marine bacterium 66A03 (16), and uncultured marine bacterium HF10 49E08 (21). β-Carotene biosynthetic genes crtE, crtB, crtI, crtY, ispA, and idi encode the enzymes necessary for the synthesis of β-carotene from isopentenyl diphosphate, and the Idi, IspA, CrtE, CrtB, CrtI, and CrtY proteins have been characterized in vitro (2328). Blh protein has been proposed to catalyze or regulate the conversion of β-carotene to retinal (29, 30), but there is no direct proof of the enzymatic activity.In this study, we used codon optimization to synthesize the blh gene from the uncultured marine bacterium 66A03 for expression in Escherichia coli, and we performed a detailed biochemical and enzymological characterization of the expressed Blh protein. In addition, the properties of the enzyme were compared with those of mammalian β-carotene 15,15′-monooxygenases.  相似文献   

7.
The temperature-dependent rotation of F1-ATPase γ subunit was observed in Vmax conditions at low viscous drag using a 60-nm gold bead (Nakanishi-Matsui, M., Kashiwagi, S., Hosokawa, H., Cipriano, D. J., Dunn, S. D., Wada, Y., and Futai, M. (2006) J. Biol. Chem. 281, 4126–4131). The Arrhenius slopes of the speed of the individual 120° steps and reciprocal of the pause length between rotation steps were very similar, indicating a flat energy pathway followed by the rotationally coupled catalytic cycle. In contrast, the Arrhenius slope of the reciprocal pause length of the γM23K mutant F1 was significantly increased, whereas that of the rotation rate was similar to wild type. The effects of the rotor γM23K substitution and the counteracting effects of βE381D mutation in the interacting stator subunits demonstrate that the rotor-stator interactions play critical roles in the utilization of stored elastic energy. The γM23K enzyme must overcome an abrupt activation energy barrier, forcing it onto a less favored pathway that results in uncoupling catalysis from rotation.F-ATPase (FoF1), consisting of the catalytic sector F13β3γδϵ) and the transmembrane proton transport sector Fo (ab2c10), synthesizes or hydrolyzes ATP coupled with proton transport (for reviews, see Ref. 16). As Abrahams et al. (7) discovered in the first high resolution x-ray structure, a critical feature of the F1-ATPase is the inherent asymmetry of the three β subunits in different conformations, βTP, βDP, and βE, referring to the nucleotide bound in each catalytic site, ATP, ADP, or empty, respectively. A rotational mechanism has been firmly established mostly based on direct observation in single molecule experiments of the behavior of the rotor complex ϵγc10, relative to the stator complex α3β3δab2 (reviewed in Ref. 1). ATP hydrolysis-dependent rotation of the γ and ϵ subunits in purified bacterial F1 (8, 9), the ϵγc10 complex in detergent solubilized FoF1 (1013), and the ϵγc10 complexin FoF1 in lipid bilayers (14) were shown experimentally by single molecule observations using fluorescent actin filament as a probe. Relative rotation of the single copy Fo a subunit was also shown in F0F1, which was immobilized through the ring of ∼10 c subunits, suggesting that the rotor and stator are interchangeable mechanical units (14). ATP synthesis by F-ATPase is believed to follow the reverse mechanism of ATP hydrolysis because mechanically induced rotation of the γ subunit in immobilized F1 in the presence of ADP and Pi results in net ATP synthesis (15, 16). There remain many questions about the mechanism of coupling between catalysis and transport via mechanical rotation. In particular, the mechanism of coupling H+ transport to rotation of the subunit c10 ring is still not well understood (4).In contrast, there is considerably more information on the mechanism of coupling catalysis to γ and ϵ subunit rotation. Observations of γ subunit rotation in the catalytic F1 sector are consistent with Boyer''s binding change model (17); thus coupling between the chemistry and rotation can be assessed by studies of the soluble F1, and these findings relate to the mechanism of the entire ATP synthase complex. The γ subunit rotates relative to the α3β3 hexamer in distinct 120° steps. A 120° rotation step consisting of pause and rotation substeps appears to correspond to the hydrolysis of one ATP, assuming that three ATP molecules are hydrolyzed per 360° revolution (18). Additional pauses observed at low ATP concentrations are attributed to the “ATP waiting” dwell (19). Yasuda et al. (19) and Shimabukuro et al. (20) further resolved that each 120° step occurred in two substeps: an 80° substep whose onset was dependent upon the Mg·ATP concentration, and a 40° substep, which was not affected by substrate concentration (19). The pause before the 80° substep, the ATP waiting dwell became shorter with increasing [Mg·ATP]. In contrast, the pause duration before the 40° rotation step was modulated by the slow hydrolysis rate of ATPγS2 or by the catalytic site mutant βE190D (in the Bacillus PS3 F1), which was found to significantly increase the length of the catalytic dwell (20). These data together indicate that the dwell before the 40° step is the “catalytic dwell” (20) and defines the order of the substeps during the 120° rotation step observed in high Mg·ATP concentrations (21).In this paper, we address the question of when the rate-limiting step of steady state catalysis occurs, with respect to the rotational behavior. Pre-steady state analysis of the burst kinetics of ATP hydrolysis at nearly Vmax conditions demonstrated that the rate-limiting transition state occurs after the reversible hydrolysis/synthesis step and before release of phosphate (Pi) (22, 23). The rate-limiting step is likely associated with a rotation step because a γ-β cross-linked enzyme is still able to undergo the initial ATP hydrolysis, but the rotation-impeded enzyme is unable to release Pi (23). Significantly, the kinetics of steady state hydrolysis can only be assessed when the Mg·ATP concentration is high enough to fill all three catalytic sites. The only model consistent with these data is one that involves all three catalytic sites. During each 120° catalytic cycle, one site binds ATP, a different site carries out reversible hydrolysis/synthesis, and the third site releases product Pi and ADP (22, 23).Steady state analyses, which take advantage of a particular γ subunit mutation γM23K (24), are consistent with this model. Replacement of the conserved γMet-23 with lysine causes an uncoupling between catalysis and γ subunit rotation caused by altered interactions between γ and β subunits (25). Importantly, Al-Shawi and Nakamoto (26) and Al-Shawi et al. (25, 27) found that the γM23K mutation strongly affected the rate-limiting transition state of steady state ATP hydrolysis and ATP synthesis. The slope of the Arrhenius plots and thus the energy of activation were significantly increased in the mutant enzyme. Several second site suppressor mutations, mostly in the γ subunit (28, 29) but also in the β subunits (30, 31), were genetically identified because they restored coupled ATP synthesis. Significantly, all were in the γ-β interface. Thermodynamic analyses found that the second site suppressors generally compensated for the primary γM23K mutations by reducing the increased activation energy (25, 27, 31). Although most of the second site mutations were found distant from the γM23K site, the x-ray crystal structures (7) suggested that γM23K may directly interact with conserved βGlu-381. As expected, replacement of βGlu-381 with aspartate also suppressed the uncoupling effects of γM23K (31).To identify the rate-limiting transition state step in the rotational behavior, we analyzed the temperature dependence of the γM23K mutant in Vmax conditions observed in single molecule experiments. Interestingly, direct observation of this mutant using the micron-length actin filaments did not detect differences in the rotation behavior at room temperature (9). In contrast, we find in the data presented here that there is dramatic effect of the mutation on the temperature dependence of the length of the catalytic dwell or pause between the 120° rotation steps. This is likely because of two factors: first, we used a bead small enough not to invoke a drag on the rotation (32), and second, the temperature dependence of the rate of the rotation steps is critical for the analyses of the mechanism.  相似文献   

8.
F1-ATPase is a rotary molecular motor in which the γ-subunit rotates against the α3β3 cylinder. The unitary γ-rotation is a 120° step comprising 80 and 40° substeps, each of these initiated by ATP binding and ADP release and by ATP hydrolysis and inorganic phosphate release, respectively. In our previous study on γ-rotation at low temperatures, a highly temperature-sensitive (TS) reaction step of F1-ATPase from thermophilic Bacillus PS3 was found below 9 °C as an intervening pause before the 80° substep at the same angle for ATP binding and ADP release. However, it remains unclear as to which reaction step the TS reaction corresponds. In this study, we found that the mutant F1(βE190D) from thermophilic Bacillus PS3 showed a clear pause of the TS reaction below 18 °C. In an attempt to identify the catalytic state of the TS reaction, the rotation of the hybrid F1, carrying a single copy of βE190D, was observed at 18 °C. The hybrid F1 showed a pause of the TS reaction at the same angle as for the ATP binding of the incorporated βE190D, although kinetic analysis revealed that the TS reaction is not the ATP binding step. These findings suggest that the TS reaction is a structural rearrangement of β before or after ATP binding.F1-ATPase (F1)2 is an ATP-driven rotary motor protein. The subunit composition of the bacterial F1-ATPase is α3β3γδϵ, and the minimum complex of F1-ATPase as a rotary motor is α3β3γ subcomplex. This motor protein forms the FoF1-ATP synthase complex by binding to another rotary motor, namely, Fo, which is driven by the proton flux resulting from the proton motive force across the membranes (14). Under physiological conditions, where the proton motive force is sufficiently large, Fo forcibly rotates F1-ATPase in the reverse direction of F1-ATPase, leading the reverse reaction of ATP hydrolysis, i.e. ATP synthesis from ADP and inorganic phosphate (Pi). When the proton motive force diminishes or F1 is isolated from Fo, F1-ATPase hydrolyzes ATP to rotate the γ-subunit against the α3β3 stator ring in the counterclockwise direction as viewed from the Fo side (5). The catalytic sites are located at the interface of the α- and β-subunits, predominantly on the β-subunit (6). Each β-subunit carries out a single turnover of ATP hydrolysis during the γ-rotation of 360° following the common catalytic reaction pathway, whereas they are 120° different in the catalytic phase. In this manner, the three β-subunits undergo different reaction steps of ATP hydrolysis upon each rotational step. The rotary motion of the γ-subunit has been demonstrated by biochemical (7) and spectroscopic methods (8) and directly proved in single-molecule observation studies (5).Since the establishment of the single-molecule rotation assay, the chemomechanical coupling scheme of F1 has been studied extensively by resolving the rotation into discrete steps. The stepping rotation was first observed under an ATP-limiting condition where F1 makes discrete 120° steps upon ATP binding (9). Then, high speed imaging of the rotation with a small probe of low friction was performed, which revealed that the 120° step comprises 80 and 40° substeps, each initiated by ATP binding, and two unknown consecutive reactions, respectively (10). This finding necessitated the identification of the two reactions that trigger the 40° substep. Hence, the rotation assay was performed using a mutant, namely F1(βE190D), and a slowly hydrolyzed ATP analog, namely ATPγS (11). Glutamate 190 of the β-subunit of F1, derived from thermophilic Bacillus PS3 and the corresponding glutamates from other F1-ATPases (Glu-181 of F1 from Escherichia coli and Glu-188 of F1 from bovine mitochondria), has been identified as one of the most critical catalytic residues for ATP hydrolysis (6, 1215). When this glutamate was substituted with aspartic acid, which has a shorter side chain than that of glutamate, the ATP cleavage step of F1 was drastically slowed. In the rotation assay, this mutant showed a distinct long pause before the 40° substep. ATPγS also caused a long pause before the 40° substep. These observations established that the 40° substep is initiated by hydrolysis. Accordingly, the pause angles before the 80 and 40° substeps are referred to as to the binding angle and the catalytic angle, respectively. Then, the rotation assay was performed in the presence of a high amount of Pi in the solution. It was shown that Pi rebinding caused the long pause at the catalytic angle, suggesting that Pi is released before the 40° substep (16).However, the reaction scheme of F1 cannot be established by simply assigning each reaction step to either the binding angle or the catalytic angle, because each reaction step must be assigned to one of the three binding or catalytic angles when considering the 360° cyclic reaction scheme of each β-subunit. Direct information about the timing of ADP release was obtained by simultaneous imaging of fluorescently labeled nucleotides and γ rotation, which showed that each β retains ADP until the γ rotates 240° after binding of the nucleotide as ATP and releases ADP between 240 and 320° (16, 17). Another powerful approach is the use of a hybrid F1 carrying a mutant β that causes a characteristic pause during the rotation. In a previous study, the hybrid F1 carrying a single copy of β(E190D), α3β2β(E190D)γ, showed a distinct pause caused by the slow hydrolysis of β(E190D) at +200° from the ATP binding angle of the mutant β (18). From this observation, it was confirmed that each β executes the chemical cleavage of the bound ATP at +200° from the angle where the ATP binds to β. The asymmetric feature of the pause of the hybrid F1 was also utilized in other experiments as a marker in the rotational trajectory to correlate the rotational angle and the conformational state of β (19) or to determine the state of F1 in the crystal structures as the pausing state at catalytic angle (20).Recently, we have found a new reaction intermediate of F1 rotation as a clear intervening pause before the 80° substep in the rotation assay below 9 °C (21). Furuike et al. (22) also observed the TS reaction in a high speed imaging experiment. The rate constant of this reaction was remarkably sensitive to temperature, giving a Q10 factor around 19. When ADP was added to solution, the pause before the 80° substep was prolonged, whereas the solution Pi caused a longer pause before the 40° substep (21). Although this result can be explained by assuming that the temperature-sensitive (TS) reaction is ADP release, it was not decisive for the identification of the TS reaction.In this study, we found that the mutant F1(βE190D) also exhibits the distinct pause of the TS reaction but at a higher temperature than for the wild-type F1, i.e. at 18 °C. This feature was advantageous in identifying the angle position of the TS reaction in the catalytic cycle for each β-subunit coupled with the 360° rotation. Taking advantage of the feature of the hybrid F1, we analyzed the rotational behavior of the hybrid F1 at 18 °C in order to assign the angle position of the TS reaction in the catalytic cycle of the 360° rotation, and we have shown that the TS reaction is not directly involved in the ADP release but in some conformational rearrangement before or after ATP binding step.  相似文献   

9.
10.
11.
12.
A central question in Wnt signaling is the regulation of β-catenin phosphorylation and degradation. Multiple kinases, including CKIα and GSK3, are involved in β-catenin phosphorylation. Protein phosphatases such as PP2A and PP1 have been implicated in the regulation of β-catenin. However, which phosphatase dephosphorylates β-catenin in vivo and how the specificity of β-catenin dephosphorylation is regulated are not clear. In this study, we show that PP2A regulates β-catenin phosphorylation and degradation in vivo. We demonstrate that PP2A is required for Wnt/β-catenin signaling in Drosophila. Moreover, we have identified PR55α as the regulatory subunit of PP2A that controls β-catenin phosphorylation and degradation. PR55α, but not the catalytic subunit, PP2Ac, directly interacts with β-catenin. RNA interference knockdown of PR55α elevates β-catenin phosphorylation and decreases Wnt signaling, whereas overexpressing PR55α enhances Wnt signaling. Taken together, our results suggest that PR55α specifically regulates PP2A-mediated β-catenin dephosphorylation and plays an essential role in Wnt signaling.Wnt/β-catenin signaling plays essential roles in development and tumorigenesis (13). Our previous work found that β-catenin is sequentially phosphorylated by CKIα4 and GSK3 (4), which creates a binding site for β-Trcp (5), leading to degradation via the ubiquitination/proteasome machinery (3). Mutations in β-catenin or APC genes that prevent β-catenin phosphorylation or ubiquitination/degradation lead ultimately to cancer (1, 2).In addition to the involvement of kinases, protein phosphatases, such as PP1, PP2A, and PP2C, are also implicated in Wnt/β-catenin regulation. PP2C and PP1 may regulate dephosphorylation of Axin and play positive roles in Wnt signaling (6, 7). PP2A is a multisubunit enzyme (810); it has been reported to play either positive or negative roles in Wnt signaling likely by targeting different components (1121). Toward the goal of understanding the mechanism of β-catenin phosphorylation, we carried out siRNA screening targeting several major phosphatases, in which we found that PP2A dephosphorylates β-catenin. This is consistent with a recent study where PP2A is shown to dephosphorylate β-catenin in a cell-free system (18).PP2A consists of a catalytic subunit (PP2Ac), a structure subunit (PR65/A), and variable regulatory B subunits (PR/B, PR/B′, PR/B″, or PR/B‴). The substrate specificity of PP2A is thought to be determined by its B subunit (9). By siRNA screening, we further identified that PR55α, a regulatory subunit of PP2A, specifically regulates β-catenin phosphorylation and degradation. Mechanistically, we found that PR55α directly interacts with β-catenin and regulates PP2A-mediated β-catenin dephosphorylation in Wnt signaling.  相似文献   

13.
14.
The scorpion α-toxin Lqh2 (from Leiurus quinquestriatus hebraeus) is active at various mammalian voltage-gated sodium channels (Navs) and is inactive at insect Navs. To resolve the molecular basis of this preference we used the following strategy: 1) Lqh2 was expressed in recombinant form and key residues important for activity at the rat brain channel rNav1.2a were identified by mutagenesis. These residues form a bipartite functional surface made of a conserved “core domain” (residues of the loops connecting the secondary structure elements of the molecule core), and a variable “NC domain” (five-residue turn and the C-tail) as was reported for other scorpion α-toxins. 2) The functional role of the two domains was validated by their stepwise construction on the similar scaffold of the anti-insect toxin LqhαIT. Analysis of the activity of the intermediate constructs highlighted the critical role of Phe15 of the core domain in toxin potency at rNav1.2a, and has suggested that the shape of the NC-domain is important for toxin efficacy. 3) Based on these findings and by comparison with other scorpion α-toxins we were able to eliminate the activity of Lqh2 at rNav1.4 (skeletal muscle), hNav1.5 (cardiac), and rNav1.6 channels, with no hindrance of its activity at Nav1.1–1.3. These results suggest that by employing a similar approach the design of further target-selective sodium channel modifiers is imminent.The pivotal role of voltage-gated sodium channels (Navs)4 in excitability mark them as major targets for a large variety of toxins that bind at distinct receptor sites and modify their gating (1). These channels are large membrane proteins made of a pore-forming α-subunit of ∼260 kDa and auxiliary β-subunits of ∼30 kDa. The α-subunit is composed of four homologous domains (D1–D4), each consisting of six α-helical transmembrane segments (S1–S6) connected by intracellular and extracellular loops. A key feature in Navs function is their ability to rapidly activate and inactivate, leading to transient increase in Na+ conductance through the cell membrane. This mechanism is attributed to the ability of the positively charged S4 voltage sensors to move across the membrane in response to changes in membrane potential (1, 2).In mammals, at least nine genes encode a variety of Nav subtypes (1, 3), whose expression varies greatly in different tissues (Nav1.1–1.3 mainly in the central nervous system; Nav1.6 in both central and peripheral neurons; Nav1.7 in the peripheral nervous system; Nav1.8 and Nav1.9 in sensory neurons; Nav1.4 and Nav1.5 in skeletal and cardiac muscles, respectively). Nav subtypes are distributed heterogeneously in the human brain and their expression is regulated under developmental and pathological conditions (1, 35). In addition, many disorders in humans result from abnormal function due to mutations in various Nav genes (68). Thus, ligands that show specificity for Nav subtypes may be used for their identification at various tissues and as leads for design of specific drugs. This requires that the bioactive surfaces of these ligands be resolved along with molecular details that determine their specificity.Among the wide range of Nav modifiers, those derived from scorpion venoms play an important role in studying channel activation (β-toxins) and inactivation (α-toxins) (911). The channel site of interaction with scorpion α-toxins, named neurotoxin receptor site-3 (12), is shared also by structurally unrelated toxins from sea anemone and spider venoms (13, 14), which raises questions as to its architecture and boundaries. Based on the findings that site-3 toxins eliminate a gating charge component associated with the movement of D4/S4 (15, 16), and that this segment plays a critical role in coupling channel inactivation to activation (17), scorpion α-toxins were postulated to inhibit channel inactivation by hindering the outward movement of this segment during depolarization (9).Scorpion α-toxins constitute a class of structurally and functionally related 61–67-residue long polypeptides reticulated by four conserved disulfide bridges. Despite a common βαββ core (10, 18, 19) these toxins are highly diverse in sequence and preference for insect and mammalian Navs. Indeed, the α-toxin class is divided to pharmacological groups according to their toxicity in insects and mice brain and ability to compete on binding at insect and mammalian Navs (10) (supplemental Fig. S1): (i) classical anti-mammalian toxins, such as Aah2 (from Androctonus australis hector) and Lqh2 (from Leiurus quinquestriatus hebraeus), which bind with high affinity to Navs at rat brain synaptosomes and are practically non-toxic to insects; (ii) α-toxins, such as LqhαIT, which strongly affect insect Navs and are weak in mammalian brain; and (iii) α-like toxins, such as Lqh3 and BmKM1 (from Buthus martensii Karsch), which are active in both mammalian brain and insects.Efforts to identify α-toxin residues involved in the interaction with the Nav receptor site-3 revealed a generally common bioactive surface divided to two topologically distinct domains: a conserved “core domain” formed by residues of the loops connecting the secondary structure elements of the molecule core, and a variable “NC domain” formed by the five-residue turn (residues 8–12) and the C-tail (2023). These analyses raised the hypothesis that a protruding conformation of the NC domain correlates with high activity at insect Navs, whereas a flat conformation of this domain appears in α-toxins active at the brain channel rNav1.2a (21). The correlation of this structural difference with toxin preference for Nav subtypes was corroborated by constructing the bioactive surface of LqhαIT on the scaffold of the anti-mammalian α-toxin Aah2 ending up with a chimera (Aah2LqhαIT(face)) active on insects, whose NC domain is in the protruding conformation (21). Despite this result, the molecular requirements that enable high affinity binding of classical α-toxins to mammalian Navs have not been clarified, and only initial data about the channel region that constitutes receptor site-3 is available (Refs. 2426; also see Ref. 10 for review).Lqh2 is a 64-residue long toxin from L. quinquestriatus hebraeus (Israeli yellow scorpion) (27) that is almost identical in sequence (96% identity) to the most active anti-mammalian toxin, Aah2, whose structure and action are documented (18, 28, 29). By functional expression and mutagenesis we uncovered residues on the Lqh2 exterior that are putatively involved in bioactivity. By construction of these residues on the scaffold of the anti-insect toxin LqhαIT we confirmed their bioactive role and differentiated those that determine toxin potency from those contributing to toxin efficacy. Comparison to other α-toxins was then instrumental for the design of an Lqh2 mutant that exhibits high specificity for the neuronal channels hNav1.1, rNav1.2a, and rNav1.3.  相似文献   

15.
Transgenic expression of human amyloid β (Aβ) peptide in body wall muscle cells of Caenorhabditis elegans has been used to better understand aspects of Alzheimer disease (AD). In human aging and AD, Aβ undergoes post-translational changes including covalent modifications, truncations, and oligomerization. Amino truncated Aβ is increasingly recognized as potentially contributing to AD pathogenesis. Here we describe surface-enhanced laser desorption ionization-time of flight mass spectrometry mass spectrometry of Aβ peptide in established transgenic C. elegans lines. Surprisingly, the Aβ being expressed is not full-length 1–42 (amino acids) as expected but rather a 3–42 truncation product. In vitro analysis demonstrates that Aβ3–42 self-aggregates like Aβ1–42, but more rapidly, and forms fibrillar structures. Similarly, Aβ3–42 is also the more potent initiator of Aβ1–40 aggregation. Seeded aggregation via Aβ3–42 is further enhanced via co-incubation with the transition metal Cu(II). Although unexpected, the C. elegans model of Aβ expression can now be co-opted to study the proteotoxic effects and processing of Aβ3–42.Numerous studies support a role for aggregating Aβ3 in mediating the toxicity that underlies AD (1, 2). However, several key questions remain central to understanding how AD and Aβ pathology are related. What is the connection between Aβ aggregation and toxicity? Is there a specific toxic Aβ conformation or species? How and why does aging impact on Aβ precipitation? Significant effort to address these questions has been invested in the use of vertebrate and simple invertebrate model organisms to simulate neurodegenerative diseases through transgenic expression of human Aβ (3). From these models, several novel insights into the proteotoxicity of Aβ have been gained (47).Human Aβ (e.g. in brain, cerebrospinal fluid, or plasma) is not found as a single species but rather as diverse mixtures of various modified, truncated, and cross-linked forms (810). Specific truncations, covalent modifications, and cross-linked oligomers of Aβ have potentially important roles in determining Aβ-associated neurotoxicity. For example, N-terminal truncations of Aβ have increased abundance in AD, rapidly aggregate, and are neurotoxic (9, 11). Furthermore, the N-terminal glutamic acid residue of Aβ3–42 can be cyclized to pyroglutamate (Aβ3(pE)-42) (12), which may be particularly important in AD pathogenesis (13, 14). Aβ3(pE)-42 is a significant fraction of total Aβ in AD brain (15), accounting for more than 50% of Aβ accumulated in plaques (16). Aβ3(pE)-42 seeds Aβ aggregation (17), confers proteolytic resistance, and is neurotoxic (13). Recently, glutaminyl cyclase (QC) has been proposed to catalyze, in vivo, pyroglutamate formation of Aβ3(pE)-40/42 (14, 18). Aβ1–42 itself cannot be cyclized by QC to Aβ3(pE)-42 (19), unlike Aβ that commences with an N-terminal glutamic acid-residue (e.g.3–42 and Aβ11–42) (20). QC has broad expression in mammalian brain (21, 22), and its inhibition attenuates accumulation of Aβ3(pE)-42 into plaques and improves cognition in a transgenic mouse model of AD that overexpresses human amyloid precursor protein (14). N-terminal truncations at position 3 have been reported in senile plaques (23, 24); however, the process that generates Aβ3–42 is unknown. Currently there are no reported animal models of Aβ3–42 expression.Advances in surface-enhanced laser desorption ionization-time of flight mass spectrometry (SELDI-TOF MS) analysis now facilitate accurate identification of particular Aβ species. Using this technology, we examined well characterized C. elegans transgenic models of AD that develop amyloid aggregates (25, 26) to see whether the human Aβ they express is post-translationally modified.  相似文献   

16.
AMP-activated protein kinase (AMPK) is a heterotrimeric complex playing a crucial role in maintaining cellular energy homeostasis. Recently, homodimerization of mammalian AMPK and yeast ortholog SNF1 was shown by us and others. In SNF1, it involved specific hydrophobic residues in the kinase domain αG-helix. Mutation of the corresponding AMPK α-subunit residues (Val-219 and Phe-223) to glutamate reduced the tendency of the kinase to form higher order homo-oligomers, as was determined by the following three independent techniques in vitro: (i) small angle x-ray scattering, (ii) surface plasmon resonance spectroscopy, and (iii) two-dimensional blue native/SDS-PAGE. Recombinant protein as well as AMPK in cell lysates of primary cells revealed distinct complexes of various sizes. In particular, the assembly of very high molecular mass complexes was dependent on both the αG-helix-mediated hydrophobic interactions and kinase activation. In vitro and when overexpressed in double knock-out (α1−/−, α2−/−) mouse embryonic fibroblast cells, activation of mutant AMPK was impaired, indicating a critical role of the αG-helix residues for AMPK activation via its upstream kinases. Also inactivation by protein phosphatase 2Cα was affected in mutant AMPK. Importantly, activation of mutant AMPK by LKB1 was restored by exchanging the corresponding and conserved hydrophobic αG-helix residues of LKB1 (Ile-260 and Phe-264) to positively charged amino acids. These mutations functionally rescued LKB1-dependent activation of mutant AMPK in vitro and in cell culture. Our data suggest a physiological role for the hydrophobic αG-helix residues in homo-oligomerization of heterotrimers and cellular interactions, in particular with upstream kinases, indicating an additional level of AMPK regulation.The maintenance of energy homeostasis is a basic requirement of all living organisms. The AMP-activated protein kinase (AMPK)2 is crucially involved in this essential process by playing a central role in sensing and regulating energy metabolism on the cellular and whole body level (16). AMPK is also participating in several signaling pathways associated with cancer and metabolic diseases, like type 2 diabetes mellitus, obesity, and other metabolic disorders (79).Mammalian AMPK belongs to a highly conserved family of serine/threonine protein kinases with homologs found in all eukaryotic organisms examined (1, 3, 10). Its heterotrimeric structure includes a catalytic α-subunit and regulatory β- and γ-subunits. These subunits exist in different isoforms (α1, α2, β1, β2, γ1, γ2, and γ3) and splice variants (for γ2 and γ3) and can thus assemble to a broad variety of heterotrimeric isoform combinations. The α- and β-subunits possess multiple autophosphorylation sites, which have been implicated in regulation of subcellular localization and kinase activation (1115). The most critical step of AMPK activation, however, is phosphorylation of Thr-172 within the activation segment of the α-subunit kinase domain. At least two AMPK upstream kinases (AMPKKs) have been identified so far, namely the tumor suppressor kinase LKB1 in complex with MO25 and STRAD (16) and Ca2+/calmodulin-dependent protein kinase kinase-2 (CamKK2) (17). Furthermore, the transforming growth factor-β-activated kinase 1 was also shown to activate AMPK using a variety of in vitro approaches (18), but the physiological relevance of these findings remains unclear. Besides direct phosphorylation of Thr-172, AMPK activity is stimulated by the allosteric activator AMP, which can bind to two Bateman domains formed by two pairs of CBS domains within the γ-subunit (1922). Hereby bound AMP not only allosterically stimulates AMPK but also protects Thr-172 from dephosphorylation by protein phosphatase 2Cα (PP2Cα) and thus hinders inactivation of the kinase (19, 22, 23). Consequently, on the cellular level, AMPK is activated upon metabolic stress increasing the AMP/ATP ratio. Furthermore, AMPK activation can also be induced by several chemical compounds, like nucleoside 5-aminoimidazole-4-carboxamide-1-β-d-ribofuranoside (24) and the anti-diabetic drug Metformin (2528). In addition, the small molecule compound A-769662 was recently developed as a direct allosteric activator of AMPK (29, 30).Previous work in our groups proposed a model of AMPK regulation by AMP, which incorporates the major functional features and the latest structural information (31). The latter mainly included truncated core complexes of AMPK from different species (3235). Further valuable structural information is provided by the x-ray structures of the isolated catalytic domains, in particular of the human AMPK α2-subunit (Protein Data Bank code 2H6D) and its yeast ortholog SNF1 (36, 37). The kinase domain of SNF1 is capable of forming homodimers in the protein crystal, as well as in vitro in solution, in a unique way, which has not been observed previously in any other kinase (36). The dimer interface is predominantly formed by hydrophobic interactions of the loop-αG region, also known as subdomain X situated on the large kinase lobe (36, 38, 39), and it mainly involves Ile-257 and Phe-261. Because the T-loop activation segment was buried within the dimer interface, it was suggested that the dimeric state of the SNF1 catalytic domain represents the inactive form of the kinase. Intriguingly, it was shown in our groups by small angle x-ray scattering that AMPK self-organizes in a concentration-dependent manner to form homo-oligomers in solution (31). However, the interface responsible for oligomerization of the AMPK heterotrimer has remained elusive.Here we further investigate the distinct oligomeric states of the AMPK heterotrimer and suggest a possible regulatory function for this process. Most importantly, we provide conclusive evidence for participation of αG-helix residues in the recognition of AMPK by its upstream kinases LKB1 and CamKK2.  相似文献   

17.
Transient exposure of β-cells to oxidative stress interrupts the transduction of signals normally coupling glucose metabolism to insulin secretion. We investigated putative persistence of effects induced by one transient oxidative stress (200 μm H2O2, 10 min) on insulin secreting cells following recovery periods of days and weeks. Three days after oxidative stress INS-1E cells and rat islets exhibited persistent dysfunction. In particular, the secretory response to 15 mm glucose was reduced by 40% in INS-1E cells stressed 3 days before compared with naïve cells. Compared with non-stressed INS-1E cells, we observed reduced oxygen consumption (−43%) and impaired glucose-induced ATP generation (−46%). These parameters correlated with increased mitochondrial reactive oxygen species formation (+60%) accompanied with down-regulation of subunits of the respiratory chain and decreased expression of genes responsible for mitochondrial biogenesis (TFAM, −24%; PGC-1α, −67%). Three weeks after single oxidative stress, both mitochondrial respiration and secretory responses were recovered. Moreover, such recovered INS-1E cells exhibited partial resistance to a second transient oxidative stress and up-regulation of UCP2 (+78%) compared with naïve cells. In conclusion, one acute oxidative stress induces β-cell dysfunction lasting over days, explained by persistent damages in mitochondrial components.Pancreatic β-cells are poised to sense blood glucose to regulate insulin exocytosis and thereby glucose homeostasis. The conversion from metabolic signals to secretory responses is mediated through mitochondrial metabolism (1). Failure of the insulin secreting β-cells, a common characteristic of both type 1 and type 2 diabetes, derives from various origins, among them mitochondrial impairment secondary to oxidative stress is a proposed mechanism (2).Oxidative stress is characterized by a persistent imbalance between excessive production of reactive oxygen species (ROS)3 and limited antioxidant defenses. Examples of ROS include superoxide (O2−̇), hydroxyl radical (OH), and hydrogen peroxide (H2O2). Superoxide can be converted to less reactive H2O2 by superoxide dismutase (SOD) and then to oxygen and water by catalase (CAT), glutathione peroxidase (GPx), and peroxiredoxin, which constitute antioxidant defenses. Increased oxidative stress and free radical damages have been proposed to participate in the diabetic state (3). In type 1 diabetes, ROS are implicated in β-cell dysfunction caused by autoimmune reactions and inflammatory cytokines (4). In the context of type 2 diabetes, excessive ROS could promote deficient insulin synthesis (5, 6) and apoptotic pathways in β-cells (5, 7). Of note, ROS fluctuations may also contribute to physiological control of cell functions (8), including the control of insulin secretion (9). It should also be stressed that metabolism of physiological nutrient increases ROS without causing deleterious effects on cell function. However, uncontrolled increase of oxidants, or reduction of their detoxification, leads to free radical-mediated chain reactions ultimately triggering pathogenic events. Pancreatic β-cells are relatively weak in expressing free radical-quenching enzymes SOD, CAT, and GPx (10, 11), rendering those cells particularly susceptible to oxidative attacks (12). Mitochondria are not only the main source of cellular oxidants, they are also the primary target of ROS (13, 14).Mitochondria are essential for pancreatic β-cell function, and damages to these organelles are well known to markedly alter glucose-stimulated insulin secretion (15). The mitochondrial genome constitutes one of the targets, encoding for 13 polypeptides essential for the integrity of electron transport chain (16). Damages to mitochondrial DNA (mtDNA) induce mutations that in turn may favor ROS generation, although the contribution of mtDNA mutations to ROS generation remains unclear. We previously reported that patient-derived mitochondrial A3243G mutation, causing mitochondrial inherited diabetes, is responsible for defective mitochondrial metabolism associated with elevated ROS levels and reduced antioxidant enzyme expression (17). On the other hand, mtDNA mutator mice exhibit accelerated aging without changes in superoxide levels in embryonic fibroblasts (18), showing that ROS generation can be dissociated from mtDNA mutations.In humans, mitochondrial defects typically appear with aging (19), accompanied by sustained ROS generation and progressive oxidant-induced damages (20). In support of this “mitochondrial theory of aging” (21), accumulating evidence shows that in older individuals mitochondria are altered, both morphologically and functionally (22). These age-related mitochondrial changes are foreseen to play a role in the late onset diabetes. In a rat model of intrauterine growth retardation, a vicious cycle between accumulation of mtDNA mutations and elevation of ROS production has been associated to β-cell abnormalities and the onset of type 2 diabetes in adulthood (23). Similarly, mitochondrion-derived ROS impair β-cell function in the Zucker diabetic fatty rat (24). Altogether, these observations point to ROS action as a triggering event inducing mitochondrial dysfunction and ultimately resulting in the loss of the secretory response in β-cells (14).In vitro, oxidative stress applied to β-cells rapidly interrupts the transduction of signals normally coupling glucose metabolism to insulin secretion (12, 25). Specifically, we reported that INS-1E β-cells and rat islets subjected to a 10-min H2O2 exposure exhibit impaired secretory response associated with mitochondrial dysfunction appearing already during the first minutes of oxidative stress (12). In the context of the mitochondrial theory of aging (21, 26), it is important to know whether transient exposure to H2O2 could possibly induce persistent modifications of mitochondrial function. Cells surviving an oxidative stress might carry defects leading to progressive loss of β-cell function. In the present study, we asked the simple but unanswered question if a short transient oxidative stress could induce durable alterations of the mitochondria and thereby chronically impair β-cell function. INS-1E β-cells and rat islets were transiently exposed to H2O2 for 10 min and analyzed after days and weeks of standard tissue culture.  相似文献   

18.
19.
20.
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