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.
Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.  相似文献   

2.
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   

3.
4.
We have identified, purified, and characterized three subcomplexes of PA700, the 19 S regulatory complex of the 26 S proteasome. These subcomplexes (denoted PS-1, PS-2, and PS-3) collectively account for all subunits present in purified PA700 but contain no overlapping components or significant levels of non-PA700 proteins. Each subcomplex contained two of the six AAA subunits (Rpt1–6) that form the binding interface of PA700 with the 20 S proteasome, the protease component of the 26 S proteasome. Unlike intact PA700, no individual PA700 subcomplex displayed ATPase activity or proteasome activating activity. However, both activities were manifested by ATP-dependent in vitro reconstitution of PA700 from the subcomplexes. We exploited functional reconstitution to define and distinguish roles of different PA700 subunits in PA700 function by selective alteration of subunits within individual subcomplexes prior to reconstitution. Carboxypeptidase treatment of either PS-2 or PS-3, subcomplexes containing specific Rpt subunits previously shown to have important roles in 26 S proteasome assembly and activation, inhibited these processes but did not affect PA700 reconstitution or ATPase activity. Thus, the intact C termini of both subunits are required for 26 S proteasome assembly and activation but not for PA700 reconstitution. Surprisingly, carboxypeptidase treatment of PS-1 also inhibited 26 S proteasome assembly and activation upon reconstitution with untreated PS-2 and PS-3. These results suggest a previously unidentified role for other PA700 subunits in 26 S proteasome assembly and activation. Our results reveal relative structural and functional relationships among the AAA subunits of PA700 and new insights about mechanisms of 26 S proteasome assembly and activation.The 26 S proteasome is a 2,500,000-Da protease complex that degrades polyubiquitylated proteins by an ATP-dependent mechanism (1, 2). The biochemical processes required for this function are divided between two subcomplexes that compose the holoenzyme (3, 4). The first, called 20 S proteasome or core particle, is a 700,000-Da complex that catalyzes peptide bond hydrolysis (5). The second, called PA700 or 19 S regulatory particle, is a 700,000-Da complex that mediates multiple aspects of proteasome function related to initial binding and subsequent delivery of substrates to the catalytic sites of the 20 S proteasome (6). The 20 S proteasome is composed of 28 subunits representing the products of 14 genes arranged in four axially stacked heteroheptameric rings (7, 8). Each of the two center β rings contains three different protease subunits that utilize N-terminal threonine residues as catalytic nucleophiles (5, 8, 9). These residues line an interior lumen formed by the stacked rings and thus are sequestered from interaction with substrates by a shell of 20 S proteasome subunits.PA700 is composed of 20 different subunits. Six of these subunits, termed Rpt1–6, are AAA2 (ATPases Associated with various cellular Activities) family members that confer ATPase activity to the complex and mediate energy-dependent proteolysis by the 26 S proteasome (2, 10). 26 S proteasome assembly from PA700 and 20 S proteasome requires ATP binding to Rpt subunits (1115). Binding of PA700 to the 20 S proteasome occurs at an axial interface between a heterohexameric ring of the PA700 Rpt subunits and the heteroheptameric outer ring of α-type 20 S proteasome subunits (16). Substrates enter the proteasome through a pore in the center of the α subunit ring that is reversibly gated by conformationally variable N-terminal residues of certain α subunits in response to PA700 binding (12, 1719). Although the degradation of polyubiquitylated proteins requires additional ATP hydrolysis-dependent actions by PA700, the assembled 26 S proteasome displays greatly increased rates of energy-independent degradation of short peptides by virtue of their increased access to catalytic sites via diffusion through the open pore (15, 18, 20).Recently, specific interactions between Rpt and α subunits that determine PA700-20 S proteasome binding and gate opening have been defined. These findings established nonequivalent roles among the six different Rpt subunits for these processes (12, 19). For example, carboxypeptidase A treatment of PA700 selectively cleaves the C termini of two Rpt subunits (Rpt2 and Rpt5) and renders PA700 incompetent for proteasome binding and activation (19). Remarkably, short peptides corresponding to the C terminus of either Rpt2 or Rpt5, but none of the other Rpt subunits, were sufficient to bind to the 20 S proteasome and activate peptide substrate hydrolysis by inducing gate opening (12, 15, 18). The C-terminal peptides of Rpt2 and Rpt5 appear to bind to different and distinct sites on the proteasome and produce additive effects on rates of peptide substrate hydrolysis, suggesting that pore size or another feature of gating can be variably modulated (19). These various results, however, do not specify whether the action of one or the other or both C-terminal peptides is essential for function of intact PA700.In addition to its role in activation, PA700 plays other essential roles in 26 S proteasome function related to substrate selection and processing. For example, PA700 captures polyubiquitylated proteins via multiple subunits that bind polyubiquitin chains (2123). Moreover, to ensure translocation of the bound ubiquitylated protein through the narrow opened substrate access pore for proteolysis, PA700 destabilizes the tertiary structure of the protein via chaperone-like activity and removes polyubiquitin chains via deubiquitylating activities of several different subunits (2430). These various functions appear to be highly coordinated and may be mechanistically linked to one another and to the hydrolysis of ATP by Rpt subunits during substrate processing.Despite support for this general model of PA700 action, there is a lack of detailed knowledge about how PA700 subunits are structurally organized and functionally linked. Previously, we identified and characterized a subcomplex of PA700 called “modulator” that contained two ATPase subunits, Rpt4 and Rpt5, and one non-ATPase subunit, p27 (31). Although this protein was identified by an assay that measured increased PA700-dependent proteasome activation, the mechanistic basis of this effect was not clear. Moreover, the modulator lacked detectable ATPase activity and proteasome activating activity. The latter feature is surprising in retrospect because of the newly identified capacity of Rpt5 to activate the proteasome directly (12, 19). This disparity suggests that specific interactions among multiple PA700 subunits determine the manifestation and regulation of various activities.This study extends our recent findings regarding relative roles of Rpt subunits in the regulation of proteasome function. It also provides new insights and significance to older work that identified and characterized the modulator as a subcomplex of PA700. Our findings unite two different lines of investigation to offer new information about the structure, function, and regulation of 26 S proteasome. They also offer insights about alternative models for assembly of PA700 and 26 S proteasome in intact cells.  相似文献   

5.
The acid-sensing ion channels (ASICs) open in response to extracellular acidic pH, and individual subunits display differential sensitivity to protons and calcium. ASIC1a acts as a high affinity proton sensor, whereas ASIC2a requires substantially greater proton concentrations to activate. Using chimeras composed of ASIC1a and ASIC2a, we determined that two regions of the extracellular domain (residues 87–197 and 323–431) specify the high affinity proton response of ASIC1a. These two regions appear to undergo intersubunit interactions within the multimeric channel to specify proton sensitivity. Single amino acid mutations revealed that amino acids around Asp357 play a prominent role in determining the pH dose response of ASIC1a. Within the same region, mutation F352L abolished PcTx1 modulation of ASIC1a. Surprisingly, we determined that another area of the extracellular domain was required for calcium-dependent regulation of ASIC1a activation, and this region functioned independently of high affinity proton sensing. These results indicate that specific regions play overlapping roles in pH-dependent gating and PcTx1-dependent modulation of ASIC1a activity, whereas a distinct region determines the calcium dependence of ASIC1a activation.The acid-sensing ion channels (ASICs)3 are proton-gated ion channels expressed in neurons throughout the central and peripheral nervous system (13). ASICs are activated by extracellular acidosis, and protons act as ligands triggering channel opening (4). Disruption of the accn2 gene (which encodes ASIC1) dramatically reduces proton-gated currents in central neurons and alters a variety of behaviors, including fear, learning, and memory (5, 6). ASIC1 also contributes to neuronal damage and death during the prolonged acidosis accompanying cerebral ischemia (7). Specifically, mice with disruptions in the accn2 gene display 60% smaller lesion size compared with normal mice in models of stroke (8). Application of PcTx1, a venom peptide that prevents ASIC1a activation, is similarly neuroprotective, even when administered hours after injury (8, 9). Thus, ASIC1a represents a novel pharmacological target for the prevention of neuronal death following stroke.Mammals have four genes encoding ASICs (accn1 to -4) that encode at least six different ASIC subunits (13, 10). Like all members of the DEG/ENaC family, individual ASIC subunits have two transmembrane regions separated by a large cysteine-rich extracellular region. Three ASIC subunits associate to form homomeric or heteromeric channels with distinct biophysical characteristics (1114). Specifically, ASIC1a homomeric channels activate at pH values much closer to neutral pH compared with ASIC2a homomeric channels. The high affinity proton sensitivity of ASIC1a plays a prominent role in acidosis-induced neuronal death, and modulators that alter the pH dose response of ASIC1a affect neuronal sensitivity to prolonged acidosis (8, 9, 15). For example, the neuroprotective venom peptide PcTx1 increases the proton sensitivity of the ASIC1a channel, allowing the channel to desensitize at neutral pH and become unresponsive to subsequent acidic shifts in pH (16, 17). The large extracellular region of ASIC1a is thought to be the site of proton/modulator interaction and governs the characteristics of channel gating (10, 11, 18). However, the exact molecular mechanisms defining ASIC1a activation and the protein domains that are responsible for the apparent proton sensitivity of ASIC1a remain unclear. Here, we used chimeras containing specific regions from both ASIC1a and ASIC2a to identify the specific protein regions that confer high affinity proton sensing, PcTx1 sensitivity, and calcium modulation to ASIC1a.  相似文献   

6.
Phosphorylase kinase (PhK), a 1.3 MDa enzyme complex that regulates glycogenolysis, is composed of four copies each of four distinct subunits (α, β, γ, and δ). The catalytic protein kinase subunit within this complex is γ, and its activity is regulated by the three remaining subunits, which are targeted by allosteric activators from neuronal, metabolic, and hormonal signaling pathways. The regulation of activity of the PhK complex from skeletal muscle has been studied extensively; however, considerably less is known about the interactions among its subunits, particularly within the non-activated versus activated forms of the complex. Here, nanoelectrospray mass spectrometry and partial denaturation were used to disrupt PhK, and subunit dissociation patterns of non-activated and phospho-activated (autophosphorylation) conformers were compared. In so doing, we have established a network of subunit contacts that complements and extends prior evidence of subunit interactions obtained from chemical crosslinking, and these subunit interactions have been modeled for both conformers within the context of a known three-dimensional structure of PhK solved by cryoelectron microscopy. Our analyses show that the network of contacts among subunits differs significantly between the nonactivated and phospho-activated conformers of PhK, with the latter revealing new interprotomeric contact patterns for the β subunit, the predominant subunit responsible for PhK''s activation by phosphorylation. Partial disruption of the phosphorylated conformer yields several novel subcomplexes containing multiple β subunits, arguing for their self-association within the activated complex. Evidence for the theoretical αβγδ protomeric subcomplex, which has been sought but not previously observed, was also derived from the phospho-activated complex. In addition to changes in subunit interaction patterns upon phospho-activation, mass spectrometry revealed a large change in the overall stability of the complex, with the phospho-activated conformer being more labile, in concordance with previous hypotheses on the mechanism of allosteric activation of PhK through perturbation of its inhibitory quaternary structure.In the cascade activation of glycogenolysis in skeletal muscle, phosphorylase kinase (PhK),1 upon becoming activated through phosphorylation, subsequently phosphorylates glycogen phosphorylase in a Ca2+-dependent reaction. This phosphorylation of glycogen phosphorylase activates its phosphorolysis of glycogen, leading to energy production (1). The 1.3 MDa (αβγδ)4 PhK complex was the first protein kinase to be characterized and is among the largest and most complex enzymes known (2). As such, the intact complex has proved to be refractory to high resolution x-ray crystallographic or NMR techniques; however, low resolution structures of the nonactivated and Ca2+-saturated conformers of PhK have been deduced through modeling (3) and solved by means of three-dimensional electron microscopic (EM) reconstruction (47), and they show that the complex is a bilobal structure with interconnecting bridges. Approximate locations of small regions of each subunit in the complex are known (810) and show that the subunits pack head-to-head as apparent αβγδ protomers that form two octameric (αβγδ)2 lobes associating in D2 symmetry (11), although direct evidence that the αβγδ protomers are discrete, functional subcomplexes has been lacking until now.Approximately 90% of the mass of the PhK complex is involved in its regulation. Its kinase activity is carried out by the catalytic core of the γ subunit (44.7 kDa), with the kcat being enhanced up to 100-fold by multiple metabolic, hormonal, and neural stimuli that are integrated through allosteric sites on PhK''s three regulatory subunits, α, β, and δ (12). The small δ subunit (16.7 kDa), which is tightly bound integral calmodulin (13), binds to at least the C-terminal regulatory domain of the γ subunit (γCRD) (14, 15), thereby mediating activation of the catalytic subunit by the obligate activator Ca2+ (16). The α and β subunits, as deduced from DNA sequencing, are polypeptides of 1237 and 1092 amino acids, respectively, with calculated masses prior to post-translational modifications of 138.4 and 125.2 kDa (17, 18). Both subunits can be phosphorylated by numerous protein kinases, including cAMP-dependent protein kinase and PhK itself (2). The α and β subunits are also homologous (38% identity and 61% similarity); however, each subunit has unique phosphorylatable regions that contain nearly all the phosphorylation sites found in these subunits (17, 18).The regulation of PhK activity by both Ca2+ (1923) and phosphorylation has been studied extensively (reviewed in Ref. 24); however, only the structural effects induced by Ca2+ are well characterized (25), primarily through comparison of the non-activated and Ca2+-activated conformers using three-dimensional EM reconstructions (4), small angle x-ray scattering modeling (3), and biophysical (2628) and chemical crosslinking methods (2932). In contrast to the Ca2+-activated versus non-activated conformers, there are no reported structures of phosphorylated PhK to compare against the non-activated form. A very small amount of structural information for phospho-activated PhK derived from chemical crosslinking raises the possibility of phosphorylation-dependent communication between the β and γ subunits: Arg-18 in the N-terminal phosphorylatable region of β was found to be relatively near the γCRD (33). Several lines of evidence suggest that transduction of the activating phosphorylation signal in PhK occurs concomitantly with conformational changes in β (33) that are detected via various methods (10, 34), including chemical crosslinking (35). For example, crosslinking of only the phosphorylated conformer by the short-span crosslinker 1,5-difluoro-2,4-dinitrobenzene results in the formation of β homodimers (35). Correspondingly, more recent two-hybrid screens of the full length β subunit against itself yielded positive binding interactions only for point mutants in which the N-terminal phosphorylatable serine residues were mutated to phosphomimetic glutamates (33). It should be noted, however, that both chemical crosslinking and two-hybrid screening have potential drawbacks in the study of subunit interactions within a multisubunit complex. In the case of the latter, it is difficult when observing homodimeric two-hybrid interactions to determine whether they correspond to naturally occurring interactions between two like subunits within a complex or between two interacting regions within a single subunit of that complex. Studying subunit interactions in a complex through chemical crosslinking comes with its own inherent limitations. For example, an initial mono-derivatization can potentially cause a conformational change in one subunit that might affect the subsequent crosslinking reaction. This is particularly the case if the crosslinker contains a functionality, such as an aromatic group, that can unexpectedly direct it to a specific locus on the protein complex (36, 37). In addition, the spacer arms on many crosslinkers are sufficiently long to confound interpretation as to whether two subunits within a complex are actually in contact. Similarly, it should be proved that any observed crosslinked conjugate is formed from subunits within a complex, as opposed to between complexes (38, 39), a control that is often not run. Thus, it is prudent to analyze subunit interactions within a complex using a variety of approaches.To corroborate, complement, and expand the previous two-hybrid screening and chemical crosslinking studies of PhK''s subunit interactions and to investigate changes in the pattern of subunit interactions induced by phosphorylation, we carried out comparative MS analyses of both intact and partially denatured forms of nonactivated and phospho-activated PhK using mass spectrometers modified specifically to enhance the transmission of large noncovalently bound protein complexes (4042). The array of subunit interactions detected for the nonactivated PhK complex largely replicated those reported in the crosslinking literature for this conformer, both corroborating those earlier studies and validating the use of these MS approaches to study subunit interactions within the PhK complex. Additionally, several novel subcomplexes of PhK were revealed, most notably an αβγδ protomer, which corroborates the observed packing of this subcomplex in the D2 symmetrical (αβγδ)4 native complex (9, 11). Moreover, we show herein that the array of subunit interactions detected for phospho-activated PhK differs significantly from that observed for the nonactivated conformer, with only the former showing extensive self-interactions between and among the regulatory β subunits. As is discussed, this suggests that activation through phosphorylation is associated with increased interprotomeric interactions in the bridged core of the PhK complex (33, 35).  相似文献   

7.
8.
9.
Helicobacter pylori CagA plays a key role in gastric carcinogenesis. Upon delivery into gastric epithelial cells, CagA binds and deregulates SHP-2 phosphatase, a bona fide oncoprotein, thereby causing sustained ERK activation and impaired focal adhesions. CagA also binds and inhibits PAR1b/MARK2, one of the four members of the PAR1 family of kinases, to elicit epithelial polarity defect. In nonpolarized gastric epithelial cells, CagA induces the hummingbird phenotype, an extremely elongated cell shape characterized by a rear retraction defect. This morphological change is dependent on CagA-deregulated SHP-2 and is thus thought to reflect the oncogenic potential of CagA. In this study, we investigated the role of the PAR1 family of kinases in the hummingbird phenotype. We found that CagA binds not only PAR1b but also other PAR1 isoforms, with order of strength as follows: PAR1b > PAR1d ≥ PAR1a > PAR1c. Binding of CagA with PAR1 isoforms inhibits the kinase activity. This abolishes the ability of PAR1 to destabilize microtubules and thereby promotes disassembly of focal adhesions, which contributes to the hummingbird phenotype. Consistently, PAR1 knockdown potentiates induction of the hummingbird phenotype by CagA. The morphogenetic activity of CagA was also found to be augmented through inhibition of non-muscle myosin II. Because myosin II is functionally associated with PAR1, perturbation of PAR1-regulated myosin II by CagA may underlie the defect of rear retraction in the hummingbird phenotype. Our findings reveal that CagA systemically inhibits PAR1 family kinases and indicate that malfunctioning of microtubules and myosin II by CagA-mediated PAR1 inhibition cooperates with deregulated SHP-2 in the morphogenetic activity of CagA.Infection with Helicobacter pylori strains bearing cagA (cytotoxin-associated gene A)-positive strains is the strongest risk factor for the development of gastric carcinoma, the second leading cause of cancer-related death worldwide (13). The cagA gene is located within a 40-kb DNA fragment, termed the cag pathogenicity island, which is specifically present in the genome of cagA-positive H. pylori strains (46). In addition to cagA, there are ∼30 genes in the cag pathogenicity island, many of which encode a bacterial type IV secretion system that delivers the cagA-encoded CagA protein into gastric epithelial cells (710). Upon delivery into gastric epithelial cells, CagA is localized to the plasma membrane, where it undergoes tyrosine phosphorylation at the C-terminal Glu-Pro-Ile-Tyr-Ala motifs by Src family kinases or c-Abl kinase (1114). The C-terminal Glu-Pro-Ile-Tyr-Ala-containing region of CagA is noted for the structural diversity among distinct H. pylori isolates. Oncogenic potential of CagA has recently been confirmed by a study showing that systemic expression of CagA in mice induces gastrointestinal and hematological malignancies (15).When expressed in gastric epithelial cells, CagA induces morphological transformation termed the hummingbird phenotype, which is characterized by the development of one or two long and thin protrusions resembling the beak of the hummingbird. It has been thought that the hummingbird phenotype is related to the oncogenic action of CagA (7, 1619). Pathophysiological relevance for the hummingbird phenotype in gastric carcinogenesis has recently been provided by the observation that infection with H. pylori carrying CagA with greater ability to induce the hummingbird phenotype is more closely associated with gastric carcinoma (2023). Elevated motility of hummingbird cells (cells showing the hummingbird phenotype) may also contribute to invasion and metastasis of gastric carcinoma.In host cells, CagA interacts with the SHP-2 phosphatase, C-terminal Src kinase, and Crk adaptor in a tyrosine phosphorylation-dependent manner (16, 24, 25) and also associates with Grb2 adaptor and c-Met in a phosphorylation-independent manner (26, 27). Among these CagA targets, much attention has been focused on SHP-2 because the phosphatase has been recognized as a bona fide oncoprotein, gain-of-function mutations of which are found in various human malignancies (17, 18, 28). Stable interaction of CagA with SHP-2 requires CagA dimerization, which is mediated by a 16-amino acid CagA-multimerization (CM)2 sequence present in the C-terminal region of CagA (29). Upon complex formation, CagA aberrantly activates SHP-2 and thereby elicits sustained ERK MAP kinase activation that promotes mitogenesis (30). Also, CagA-activated SHP-2 dephosphorylates and inhibits focal adhesion kinase (FAK), causing impaired focal adhesions. It has been shown previously that both aberrant ERK activation and FAK inhibition by CagA-deregulated SHP-2 are involved in induction of the hummingbird phenotype (31).Partitioning-defective 1 (PAR1)/microtubule affinity-regulating kinase (MARK) is an evolutionally conserved serine/threonine kinase originally isolated in C. elegans (3234). Mammalian cells possess four structurally related PAR1 isoforms, PAR1a/MARK3, PAR1b/MARK2, PAR1c/MARK1, and PAR1d/MARK4 (3537). Among these, PAR1a, PAR1b, and PAR1c are expressed in a variety of cells, whereas PAR1d is predominantly expressed in neural cells (35, 37). These PAR1 isoforms phosphorylate microtubule-associated proteins (MAPs) and thereby destabilize microtubules (35, 38), allowing asymmetric distribution of molecules that are involved in the establishment and maintenance of cell polarity.In polarized epithelial cells, CagA disrupts the tight junctions and causes loss of apical-basolateral polarity (39, 40). This CagA activity involves the interaction of CagA with PAR1b/MARK2 (19, 41). CagA directly binds to the kinase domain of PAR1b in a tyrosine phosphorylation-independent manner and inhibits the kinase activity. Notably, CagA binds to PAR1b via the CM sequence (19). Because PAR1b is present as a dimer in cells (42), CagA may passively homodimerize upon complex formation with the PAR1 dimer via the CM sequence, and this PAR1-directed CagA dimer would form a stable complex with SHP-2 through its two SH2 domains.Because of the critical role of CagA in gastric carcinogenesis (7, 1619), it is important to elucidate the molecular basis underlying the morphogenetic activity of CagA. In this study, we investigated the role of PAR1 isoforms in induction of the hummingbird phenotype by CagA, and we obtained evidence that CagA-mediated inhibition of PAR1 kinases contributes to the development of the morphological change by perturbing microtubules and non-muscle myosin II.  相似文献   

10.
High grade gliomas such as glioblastoma multiforme express multiple members of the epithelial sodium channel (ENaC)/Degenerin family, characteristically displaying a basally active amiloride-sensitive cation current not seen in normal human astrocytes or lower grade gliomas. Using quantitative real time PCR, we have shown higher expression of ASIC1, αENaC, and γENaC in D54-MG human glioblastoma multiforme cells compared with primary human astrocytes. We hypothesize that this glioma current is mediated by a hybrid channel composed of a mixture of ENaC and acid-sensing ion channel (ASIC) subunits. To test this hypothesis we made dominant negative cDNAs for ASIC1, αENaC, γENaC, and δENaC. D54-MG cells transfected with the dominant negative constructs for ASIC1, αENaC, or γENaC showed reduced protein expression and a significant reduction in the amiloride-sensitive whole cell current as compared with untransfected D54-MG cells. Knocking down αENaC or γENaC also abolished the high PK+/PNa+ of D54-MG cells. Knocking down δENaC in D54-MG cells reduced δENaC protein expression but had no effect on either the whole cell current or K+ permeability. Using co-immunoprecipitation we show interactions between ASIC1, αENaC, and γENaC, consistent with these subunits interacting with each other to form an ion channel in glioma cells. We also found a significant inhibition of D54-MG cell migration after ASIC1, αENaC, or γENaC knockdown, consistent with the hypothesis that ENaC/Degenerin subunits play an important role in glioma cell biology.Gliomas are the most common primary tumors of the central nervous system. These tumors arise either from astrocytes or their progenitor cells (1). Gliomas are divided into four grades based on the degree of malignancy. Glioblastoma multiforme (GBM),2 Grade IV, is the most frequently occurring, most invasive, and has the worst prognostic outcome with a median survival of approximately one year from diagnosis (2).We have previously reported the presence of an amiloride-sensitive current in glioblastoma cells that is not seen in normal astrocytes or low grade gliomas (3). Amiloride is a potassium sparing diuretic that inhibits sodium channels composed of subunits from the epithelial sodium channel (ENaC)/Degenerin (Deg) family. Amiloride-sensitive Na+ channels are essential for the regulation of Na+ transport into cells and tissues throughout the body. These channels are found in all body tissues; from epithelia, endothelia, osteoblasts, keratinocytes, taste cells, lymphocytes, and brain (4). Apart from the ENaCs, the ENaC/Deg family also includes acid-sensing ion channels (ASICs) which have been found predominantly in neurons (46). Primary malfunctions of ENaC/Deg family members underlie or are involved in the pathophysiology of several human diseases such as salt-sensitive hypertension (7, 8), pseudohypoaldosteronism type I (7), cystic fibrosis (9), chronic airway diseases (10, 11), and flu (12).The ENaC/Deg family subunits share the same structural topology. They all have short intracellular N and C termini, two transmembrane spanning domains, and a large extracellular cysteine-rich loop (4, 5). There are five ENaC subunits termed α, β, γ, δ, and ϵ. Functional ion channels arise from a multimeric assembly of these subunits. The prototypical ENaC channel of the collecting duct principal cell is thought to be αβγENaC (13, 14). The α-ENaC subunit appears to be the core conducting element, whereas the β- and γ-ENaC subunits are associated with trafficking and insertion of the channel in the cell membrane (13, 15, 16). ASICs are homologous to ENaCs and are most prevalently expressed in the brain and nervous system (1719), although they are also found in the retina (2022), testes (23), pituitary gland (24), lung epithelia (22), and bone and cartilage (25). Four ASIC genes have been identified so far, ASIC1–4. Of these, ASIC1–3 has multiple splice variants (19, 22). The crystal structure of chicken ASIC1 has revealed it to be a homotrimer (26). ASICs differ from their ENaC counterparts in that they are transiently activated by extracellular acid (19) and are much less sensitive to inhibition by amiloride (27, 28). Also ASIC1 is inhibited with high affinity by psalmotoxin 1 (PcTX-1), a 40-amino acid peptide found in the venom of the West Indies tarantula, Psalmopoeus Cambridgei (29). ASICs, because they are activated by acidic pH, have been suggested to play a role in chemical pain associated with increased tissue acidification as occurs in ischemia (30, 31). They have also been implicated in touch sensation (32), taste (33), fear-conditioning (6), and learning and memory (34).Our laboratory has proposed that ENaC/Deg channels underlie the basally activated cation current measured in high grade glioma cells (3). We hypothesize that the channels forming this current pathway are composed of a mixture of ASIC and ENaC subunits. RNA profiling of a large number of GBM-derived cell lines and freshly resected tumors have revealed the presence of a myriad of ASIC/ENaC components (3). The basally active current seen in GBM cells can be significantly reduced by amiloride or benzamil (a higher affinity amiloride analog), both of which are inhibitors of the ENaC/Deg family of ion channels (3). PcTX1, a selective ASIC1 blocker, also effectively abolishes the basally active GBM current (35). We have previously shown that ENaC and ASIC subunits can form cross-clade interactions in a heterologous expression system (36). This study aims to probe the composition of the novel ENaC/Deg heteromer in a glioma cell line, D54-MG. Our study postulates that a change in GBM cell electrophysiological properties after subunit knockdown would be indicative of that subunit being a part of the GBM channel. We have sequentially knocked down different ENaC/Deg subunits from the D54-MG glioma cells and measured amiloride-sensitive whole cell current using patch clamp. We found that knocking down various ENaC/Deg subunits significantly reduced the whole cell patch clamp current in glioma cells and changed the resting Na+/K+ permeability of the these cells. After subunit knockdown, glioma cells showed a reduced cell migration as compared with control cells, consistent with our hypothesis that ENaC/Deg subunits play an important role in glioma cell pathophysiology.  相似文献   

11.
12.
13.
The molecular chaperone Hsp27 exists as a distribution of large oligomers that are disassembled by phosphorylation at Ser-15, -78, and -82. It is controversial whether the unphosphorylated Hsp27 or the widely used triple Ser-to-Asp phospho-mimic mutant is the more active molecular chaperone in vitro. This question was investigated here by correlating chaperone activity, as measured by the aggregation of reduced insulin or α-lactalbumin, with Hsp27 self-association as monitored by analytical ultracentrifugation. Furthermore, because the phospho-mimic is generally assumed to reproduce the phosphorylated molecule, the size and chaperone activity of phosphorylated Hsp27 were compared with that of the phospho-mimic. Hsp27 was triply phosphorylated by MAPKAP-2 kinase, and phosphorylation was tracked by urea-PAGE. An increasing degree of suppression of insulin or α-lactalbumin aggregation correlated with a decreasing Hsp27 self-association, which was the least for phosphorylated Hsp27 followed by the mimic followed by the unphosphorylated protein. It was also found that Hsp27 added to pre-aggregated insulin did not reverse aggregation but did inhibit these aggregates from assembling into even larger aggregates. This chaperone activity appears to be independent of Hsp27 phosphorylation. In conclusion, the most active chaperone of insulin and α-lactalbumin was the Hsp27 (elongated) dimer, the smallest Hsp27 subunit observed under physiological conditions. Next, the Hsp27 phospho-mimic is only a partial mimic of phosphorylated Hsp27, both in self-association and in chaperone function. Finally, the efficient inhibition of insulin aggregation by Hsp27 dimer led to the proposal of two models for this chaperone activity.Oligomeric heat shock protein 27 (Hsp27)2 is a ubiquitous mammalian protein with a variety of functions in health and disease (18). These functions include ATP-independent chaperone activity in response to environmental stress, e.g. heat shock and oxidative stress, control of apoptosis, and regulation of actin cytoskeleton dynamics. Hsp27 is a member of the α-crystallin small heat shock protein family of which αB-crystallin is the archetype. These proteins are characterized by an α-crystallin domain of 80–90 residues consisting of roughly eight β-strands that form an intermolecular β-sheet interaction interface within a dimer, the basic building subunit of the oligomer (2, 4, 911).Hsp27 is in equilibrium between high molecular weight oligomers and much lower molecular weight multimers. It has been reported that unphosphorylated Hsp27 includes predominantly a distribution of high molecular species ranging in size from 12-mer to 35-mer (1219). Phosphorylation of Hsp27 at serines 15, 78, and 82 by the p38-activated MAPKAP-2 kinase (2022) or the use of the triple Ser-to-Asp phospho-mimic results in a major shift in the equilibrium toward much smaller multimers (23) and in an alteration of its function (1, 3, 6, 7, 24, 25). The size distribution of the smaller species has been reported to be between monomer and tetramer (1216, 18, 19).Small heat shock proteins, including Hsp27, behave as ATP-independent molecular chaperones during cellular heat shock. They bind partially unfolded proteins and prevent their aggregation until the proteins can be refolded by larger ATP-dependent chaperones or are digested (7, 8, 26). This function includes the up-regulation and/or phosphorylation of Hsp27.It is not entirely clear what the role of Hsp27 size and phosphorylation state plays in its heat shock function because there are conflicting results in the literature. Some in vitro studies concluded that the unphosphorylated oligomeric Hsp27 (or the murine isoform Hsp25) protects proteins against aggregation better than does the phosphorylation mimic (13, 19, 27), whereas others found no difference (16, 28, 29), and still other studies found that the mimic protects better than does the unphosphorylated wild type (27, 30, 31). In-cell studies found that phosphorylation of Hsp27 was essential for thermo-protection of actin filaments (32), and the Hsp27 phosphorylation mimic decreased inclusion body formation better than did unphosphorylated Hsp27 (33). This study was undertaken to investigate the molecular chaperone function of Hsp27 by correlating chaperone activity with Hsp27 size and by comparing fully phosphorylated Hsp27 with its phospho-mimic.  相似文献   

14.
In Archaea, an hexameric ATPase complex termed PAN promotes proteins unfolding and translocation into the 20 S proteasome. PAN is highly homologous to the six ATPases of the eukaryotic 19 S proteasome regulatory complex. Thus, insight into the mechanism of PAN function may reveal a general mode of action mutual to the eukaryotic 19 S proteasome regulatory complex. In this study we generated a three-dimensional model of PAN from tomographic reconstruction of negatively stained particles. Surprisingly, this reconstruction indicated that the hexameric complex assumes a two-ring structure enclosing a large cavity. Assessment of distinct three-dimensional functional states of PAN in the presence of adenosine 5′-O-(thiotriphosphate) and ADP and in the absence of nucleotides outlined a possible mechanism linking nucleotide binding and hydrolysis to substrate recognition, unfolding, and translocation. A novel feature of the ATPase complex revealed in this study is a gate controlling the “exit port” of the regulatory complex and, presumably, translocation into the 20 S proteasome. Based on our structural and biochemical findings, we propose a possible model in which substrate binding and unfolding are linked to structural transitions driven by nucleotide binding and hydrolysis, whereas translocation into the proteasome only depends upon the presence of an unfolded substrate and binding but not hydrolysis of nucleotide.In eukaryotic cells most protein breakdown in the cytosol and nucleus is catalyzed by the 26 S proteasome. This ∼2.5-MDa (1) complex degrades ubiquitin-conjugated and certain non-ubiquitinated proteins in an ATP-dependent manner (2, 3). The 26 S complex is composed of one or two 19 S regulatory particles situated at the ends of the cylindrical 20 S proteasome. Within the 26 S complex, proteins are hydrolyzed in the 20 S proteasome. Tagged substrates, however, first bind to the 19 S regulatory particle, which catalyzes their unfolding and translocation into the 20 S subcomplex (4, 5). The 19 S regulatory particle consists of at least 17 different subunits (1, 6). Nine of these subunits form a “lid,” whereas the other eight subunits, including six ATPases, comprise the base of the 19 S particle. Electron microscopy (710) as well as cross-linking experiments (11, 12) have demonstrated that the six homologous ATPases are associated with the α rings of the 20 S particle.Unlike eukaryotes, Archaea and certain eubacteria contain homologous 20 S particles but lack ubiquitin. Their proteasomes degrade proteins in association with a hexameric ATPase ring complex termed PAN (13). PAN appears to be the evolutionary precursor of the 19 S base, predating the coupling of ubiquitination and proteolysis in eukaryotes (14). In addition, PAN recognizes the bacterial targeting sequence ssrA (in analogy to the polyubiquitin conjugates in eukaryotes) and efficiently unfolds and translocates globular substrates, like green fluorescent protein, when tagged with ssrA (15). In both PAN and the 19 S proteasome regulatory complexes, ATP is essential for substrate unfolding and translocation and for opening of the gated channel in the α ring through which substrates enter the 20 S particle (1517). Because this portal is quite narrow (1820), only extended polypeptides can enter the 20 S proteasome. Consequently, a globular substrate must be unfolded by the associated ATPase complex to be translocated and digested within the 20 S particle.PAN and the six ATPases found at the base of the 19 S particle are members of the AAA+ superfamily of multimeric ATPases which also includes the ATP-dependent proteases Lon and FtsH and the regulatory components of the bacterial ATP-dependent proteases ClpAP, ClpXP, and HslUV (8, 21). For mechanistic studies of the roles of ATP, the simpler archaeal PAN-20 S system offers many technical advantages over the much more complex 26 S proteasome. For example, prior studies of PAN (17, 22) demonstrated that unfolding of globular substrates (e.g. green fluorescent protein-ssrA) requires ATP hydrolysis. The same was also shown for the Escherichia coli ATP-dependent proteases ClpXP (23) and ClpAP (24). We have also shown that unfolding by PAN can take place on the surface of the ATPase ring in the absence of translocation (15). Thus, unfolding seems to proceed independently from protein translocation into the 20 S proteolytic particle. It is noteworthy that other studies suggest that proteins are unfolded by energy-dependent translocation through the ATPase ring (25, 26). These studies have suggested that the translocation of an unfolded polypeptide from the ATPase into the 20 S core is an active process that is coupled to ATP hydrolysis. A key to underline a detailed molecular mechanism for substrate binding, unfolding, and translocation by the proteasome regulatory ATPase complex is improved understanding of its architecture and the nucleotide-dependent structural transitions that afford these functions.To date we and others have failed to generate micrographs suitable for three-dimensional reconstruction of PAN using single-particle EM analysis. Likewise, structural information regarding the three-dimensional architecture and subunit organization within the 19 S particle is very limited. In fact, high resolution three-dimensional information on the 19 S complex is not yet available. Most knowledge available is based on cross-linking experiments (11, 12) as well as EM structural analysis (710), which provided a three-dimensional model outline of the general architecture of the 26 S complex. Unlike the 19 S complex, the structure of the 20 S subcomplex was determined by x-ray crystallography (18, 19). In contrast to the highly homogenous structure of the 20 S complex, the structural heterogeneity and flexibility of the 19 S subcomplex is presumably reflected in multiple conformations, which in turn also contribute to the difficulty in generating a high resolution three-dimensional structural model of the 26 S proteasome. Accordingly, the initial goal of this study was to generate a three-dimensional model of PAN that will allow us to determine its general architecture and to correlate unique conformational transitions within this ATPase with the nucleotide state of the complex (i.e. in the presence of ATPγS, ADP, or in the absence of nucleotides).Smith et al. (27) suggested a general architecture for the PAN-20 S complex based on two-dimensional averaging of a Thermoplasma acidophilum (TA)3 20 S proteasome and Methanococcus jannaschii (MJ) PAN hybrid complex in the presence of ATPγS. Based on side-view projections of that complex, these authors proposed that PAN assumes an overall structure similar to E. coli HslU (2830).We realized that although PAN appears heterogeneous in electron micrographs, it does not occupy all possible orientations when adsorbed to carbon-coated electron microscopy (EM) grids, a prerequisite for single particle analysis. This problem was overcome by applying electron tomography in conjunction with a three-dimensional averaging procedure that accounts for the missing wedge in the Fourier space of electron tomograms (31, 32). The three-dimensional model generated revealed an unexpected architecture leading to a possible molecular mechanism describing the function of PAN and presumably the 19 S ATPases.  相似文献   

15.
16.
17.
Understanding the activation mechanism of Cys loop ion channel receptors is key to understanding their physiological and pharmacological properties under normal and pathological conditions. The ligand-binding domains of these receptors comprise inner and outer β-sheets and structural studies indicate that channel opening is accompanied by conformational rearrangements in both β-sheets. In an attempt to resolve ligand-dependent movements in the ligand-binding domain, we employed voltage-clamp fluorometry on α1 glycine receptors to compare changes mediated by the agonist, glycine, and by the antagonist, strychnine. Voltage-clamp fluorometry involves labeling introduced cysteines with environmentally sensitive fluorophores and inferring structural rearrangements from ligand-induced fluorescence changes. In the inner β-sheet, we labeled residues in loop 2 and in binding domain loops D and E. At each position, strychnine and glycine induced distinct maximal fluorescence responses. The pre-M1 domain responded similarly; at each of four labeled positions glycine produced a strong fluorescence signal, whereas strychnine did not. This suggests that glycine induces conformational changes in the inner β-sheet and pre-M1 domain that may be important for activation, desensitization, or both. In contrast, most labeled residues in loops C and F yielded fluorescence changes identical in magnitude for glycine and strychnine. A notable exception was H201C in loop C. This labeled residue responded differently to glycine and strychnine, thus underlining the importance of loop C in ligand discrimination. These results provide an important step toward mapping the domains crucial for ligand discrimination in the ligand-binding domain of glycine receptors and possibly other Cys loop receptors.Glycine receptor (GlyR)3 chloride channels are pentameric Cys loop receptors that mediate fast synaptic transmission in the nervous system (1, 2). This family also includes nicotinic acetylcholine receptors (nAChRs), γ-aminobutyric acid type A and type C receptors, and serotonin type 3 receptors. Individual subunits comprise a large ligand-binding domain (LBD) and a transmembrane domain consisting of four α-helices (M1–M4). The LBD consists of a 10-strand β-sandwich made of an inner β-sheet with six strands and an outer β-sheet with four strands (3). The ligand-binding site is situated at the interface of adjacent subunits and is formed by loops A–C from one subunit and loops D–F from the neighboring subunit (3).The activation mechanism of Cys loop receptors is currently the subject of intense investigation because it is key to understanding receptor function under normal and pathological conditions (4, 5). Based on structural analysis of Torpedo nAChRs, Unwin and colleagues (6, 7) originally proposed that agonist binding induced the inner β-sheet to rotate, whereas the outer β-sheet tilted slightly upwards with loop C clasping around the agonist. These movements were thought to be transmitted to the transmembrane domain via a differential movement of loop 2 (β1-β2) and loop 7 (β6-β7) (both part of the inner β-sheet) and the pre-M1 domain (which is linked via a β-strand to the loop C in the outer sheet). The idea of large loop C movements accompanying agonist binding is supported by structural and functional data (3, 813). However, a direct link between loop C movements and channel gating has proved more difficult to establish. Although computational modeling studies have suggested that this loop may be a major component of the channel opening mechanism (1418), experimental support for this model is not definitive. Similarly, loop F is also thought to move upon ligand binding, although there is as yet no consensus as to whether these changes represent local or global conformational changes (11, 1921). Recently, a comparison of crystal structures of bacterial Cys loop receptors in the closed and open states revealed that although both the inner and outer β-sheets exhibit different conformations in closed and open states, the pre-M1 domain remains virtually stationary (22, 23). It is therefore relevant to question whether loop C, loop F, and pre-M1 movements are essential for Cys loop receptor activation.Strychnine is a classical competitive antagonist of GlyRs (24, 25), and to date there is no evidence that it can produce LBD structural changes. In this study we use voltage-clamp fluorometry (VCF) to compare glycine- and strychnine-induced conformational changes in the GlyR loops 2, C, D, E, and F and the pre-M1 domain in an attempt to determine whether they signal ligand-binding events, local conformational changes, or conformational changes associated with receptor activation.In a typical VCF experiment, a domain of interest is labeled with an environmentally sensitive fluorophore, and current and fluorescence are monitored simultaneously during ligand application. VCF is ideally suited for identifying ligand-specific conformational changes because it can report on electrophysiologically silent conformational changes (26), such as those induced by antagonists. Indeed, VCF has recently provided valuable insights into the conformational rearrangements of various Cys loop receptors (19, 21, 2733).  相似文献   

18.
19.
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   

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