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 46 毫秒
1.
2.
Sorghum (Sorghum bicolor L. Moench) has two isozymes of the cyanogenic β-glucosidase dhurrinase: dhurrinase-1 (Dhr1) and dhurrinase-2 (Dhr2). A nearly full-length cDNA encoding dhurrinase was isolated from 4-d-old etiolated seedlings and sequenced. The cDNA has a 1695-nucleotide-long open reading frame, which codes for a 565-amino acid-long precursor and a 514-amino acid-long mature protein, respectively. Deduced amino acid sequence of the sorghum Dhr showed 70% identity with two maize (Zea mays) β-glucosidase isozymes. Southern-blot data suggested that β-glu-cosidase is encoded by a small multigene family in sorghum. Northern-blot data indicated that the mRNA corresponding to the cloned Dhr cDNA is present at high levels in the node and upper half of the mesocotyl in etiolated seedlings but at low levels in the root—only in the zone of elongation and the tip region. Light-grown seedling parts had lower levels of Dhr mRNA than those of etiolated seedlings. Immunoblot analysis performed using maize-anti-β-glucosidase sera detected two distinct dhurrinases (57 and 62 kD) in sorghum. The distribution of Dhr activity in different plant parts supports the mRNA and immunoreactive protein data, suggesting that the cloned cDNA corresponds to the Dhr1 (57 kD) isozyme and that the dhr1 gene shows organ-specific expression.  相似文献   

3.
4.
Studies were conducted to identify a 64-kD thylakoid membrane protein of unknown function. The protein was extracted from chloroplast thylakoids under low ionic strength conditions and purified to homogeneity by preparative sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Four peptides generated from the proteolytic cleavage of the wheat 64-kD protein were sequenced and found to be identical to internal sequences of the chloroplast-coupling factor (CF1) α-subunit. Antibodies for the 64-kD protein also recognized the α-subunit of CF1. Both the 64-kD protein and the 61-kD CF1 α-subunit were present in the monocots barley (Hordeum vulgare), maize (Zea mays), oat (Avena sativa), and wheat (Triticum aestivum); but the dicots pea (Pisum sativum), soybean (Glycine max Merr.), and spinach (Spinacia oleracea) contained only a single polypeptide corresponding to the CF1 α-subunit. The 64-kD protein accumulated in response to high irradiance (1000 μmol photons m−2 s−1) and declined in response to low irradiance (80 μmol photons m−2 s−1) treatments. Thus, the 64-kD protein was identified as an irradiance-dependent isoform of the CF1 α-subunit found only in monocots. Analysis of purified CF1 complexes showed that the 64-kD protein represented up to 15% of the total CF1 α-subunit.  相似文献   

5.
6.
Aggregatibacter actinomycetemcomitans is a gram-negative opportunistic oral pathogen. It is frequently associated with subgingival biofilms of both chronic and aggressive periodontitis, and the diseased sites of the periodontium exhibit increased levels of the proinflammatory mediator interleukin (IL)-1β. Some bacterial species can alter their physiological properties as a result of sensing IL-1β. We have recently shown that this cytokine localizes to the cytoplasm of A. actinomycetemcomitans in co-cultures with organotypic gingival mucosa. However, current knowledge about the mechanism underlying bacterial IL-1β sensing is still limited. In this study, we characterized the interaction of A. actinomycetemcomitans total membrane protein with IL-1β through electrophoretic mobility shift assays. The interacting protein, which we have designated bacterial interleukin receptor I (BilRI), was identified through mass spectrometry and was found to be Pasteurellaceae specific. Based on the results obtained using protein function prediction tools, this protein localizes to the outer membrane and contains a typical lipoprotein signal sequence. All six tested biofilm cultures of clinical A. actinomycetemcomitans strains expressed the protein according to phage display-derived antibody detection. Moreover, proteinase K treatment of whole A. actinomycetemcomitans cells eliminated BilRI forms that were outer membrane specific, as determined through immunoblotting. The protein was overexpressed in Escherichia coli in both the outer membrane-associated form and a soluble cytoplasmic form. When assessed using flow cytometry, the BilRI-overexpressing E. coli cells were observed to bind 2.5 times more biotinylated-IL-1β than the control cells, as detected with avidin-FITC. Overexpression of BilRI did not cause binding of a biotinylated negative control protein. In a microplate assay, soluble BilRI bound to IL-1β, but this binding was not specific, as a control protein for IL-1β also interacted with BilRI. Our findings suggest that A. actinomycetemcomitans expresses an IL-1β-binding surface-exposed lipoprotein that may be part of the bacterial IL-1β-sensing system.  相似文献   

7.
8.
Classic and novel protein kinase C (PKC) isozymes contain two zinc finger motifs, designated “C1a” and “C1b” domains, which constitute the recognition modules for the second messenger diacylglycerol (DAG) or the phorbol esters. However, the individual contributions of these tandem C1 domains to PKC function and, reciprocally, the influence of protein context on their function remain uncertain. In the present study, we prepared PKCδ constructs in which the individual C1a and C1b domains were deleted, swapped, or substituted for one another to explore these issues. As isolated fragments, both the δC1a and δC1b domains potently bound phorbol esters, but the binding of [3H]phorbol 12,13-dibutyrate ([3H]PDBu) by the δC1a domain depended much more on the presence of phosphatidylserine than did that of the δC1b domain. In intact PKCδ, the δC1b domain played the dominant role in [3H]PDBu binding, membrane translocation, and down-regulation. A contribution from the δC1a domain was nonetheless evident, as shown by retention of [3H]PDBu binding at reduced affinity, by increased [3H]PDBu affinity upon expression of a second δC1a domain substituting for the δC1b domain, and by loss of persistent plasma membrane translocation for PKCδ expressing only the δC1b domain, but its contribution was less than predicted from the activity of the isolated domain. Switching the position of the δC1b domain to the normal position of the δC1a domain (or vice versa) had no apparent effect on the response to phorbol esters, suggesting that the specific position of the C1 domain within PKCδ was not the primary determinant of its activity.One of the essential steps for protein kinase C (PKC)2 activation is its translocation from the cytosol to the membranes. For conventional (α, βI, βII, and γ) and novel (δ, ε, η, and θ) PKCs, this translocation is driven by interaction with the lipophilic second messenger sn-1,2-diacylglycerol (DAG), generated from phosphatidylinositol 4,5-bisphosphate upon the activation of receptor-coupled phospholipase C or indirectly from phosphatidylcholine via phospholipase D (1). A pair of zinc finger structures in the regulatory domain of the PKCs, the “C1” domains, are responsible for the recognition of the DAG signal. The DAG-C1 domain-membrane interaction is coupled to a conformational change in PKC, both causing the release of the pseudosubstrate domain from the catalytic site to activate the enzyme and triggering the translocation to the membrane (2). By regulating access to substrates, PKC translocation complements the intrinsic enzymatic specificity of PKC to determine its substrate profile.The C1 domain is a highly conserved cysteine-rich motif (∼50 amino acids), which was first identified in PKC as the interaction site for DAG or phorbol esters (3). It possesses a globular structure with a hydrophilic binding cleft at one end surrounded by hydrophobic residues. Binding of DAG or phorbol esters to the C1 domain caps the hydrophilic cleft and forms a continuous hydrophobic surface favoring the interaction or penetration of the C1 domain into the membrane (4). In addition to the novel and classic PKCs, six other families of proteins have also been identified, some of whose members possess DAG/phorbol ester-responsive C1 domains. These are the protein kinase D (5), the chimaerin (6), the munc-13 (7), the RasGRP (guanyl nucleotide exchange factors for Ras and Rap1) (8), the DAG kinase (9), and the recently characterized MRCK (myotonic dystrophy kinase-related Cdc42-binding kinase) families (10). Of these C1 domain-containing proteins, the PKCs have been studied most extensively and are important therapeutic targets (11). Among the drug candidates in clinical trials that target PKC, a number such as bryostatin 1 and PEP005 are directed at the C1 domains of PKC rather than at its catalytic site.Both the classic and novel PKCs contain in their N-terminal regulatory region tandem C1 domains, C1a and C1b, which bind DAG/phorbol ester (12). Multiple studies have sought to define the respective roles of these two C1 domains in PKC regulation, but the issue remains unclear. Initial in vitro binding measurements with conventional PKCs suggested that 1 mol of phorbol ester bound per mole of PKC (13-15). On the other hand, Stubbs et al., using a fluorescent phorbol ester analog, reported that PKCα bound two ligands per PKC (16). Further, site-directed mutagenesis of the C1a and C1b domains of intact PKCα indicated that the C1a and C1b domains played equivalent roles for membrane translocation in response to phorbol 12-myristate 13-acetate (PMA) and (-)octylindolactam V (17). Likewise, deletion studies indicated that the C1a and C1b domains of PKCγ bound PDBu equally with high potency (3, 18). Using a functional assay with PKCα expression in yeast, Shieh et al. (19) deleted individual C1 domains and reported that C1a and C1b were both functional and equivalent upon stimulation by PMA, with either deletion causing a similar reduction in potency of response, whereas for mezerein the response depended essentially on the C1a domain, with much weaker response if only the C1b domain was present. Using isolated C1 domains, Irie et al. (20) suggested that the C1a domain of PKCα but not those of PKCβ or PKCγ bound [3H]PDBu preferentially; different ligands showed a generally similar pattern but with different extents of selectivity. Using synthesized dimeric bisphorbols, Newton''s group reported (21) that, although both C1 domains of PKCβII are oriented for potential membrane interaction, only one C1 domain bound ligand in a physiological context.In the case of novel PKCs, many studies have been performed on PKCδ to study the equivalency of the twin C1 domains. The P11G point mutation of the C1a domain, which caused a 300-fold loss of binding potency in the isolated domain (22), had little effect on the phorbol ester-dependent translocation of PKCδ in NIH3T3 cells, whereas the same mutation of the C1b caused a 20-fold shift in phorbol ester potency for inducing translocation, suggesting a major role of the C1b domain for phorbol ester binding (23). A secondary role for the C1a domain was suggested, however, because mutation in the C1a domain as well as the C1b domain caused a further 7-fold shift in potency. Using the same mutations in the C1a and C1b domains, Bögi et al. (24) found that the binding selectivity for the C1a and C1b domains of PKCδ appeared to be ligand-dependent. Whereas PMA and the indole alkaloids indolactam and octylindolactam were selectively dependent on the C1b domain, selectivity was not observed for mezerein, the 12-deoxyphorbol 13-monoesters prostratin and 12-deoxyphorbol 13-phenylacetate, and the macrocyclic lactone bryostatin 1 (24). In in vitro studies using isolated C1a and C1b domains of PKCδ, Cho''s group (25) described that the two C1 domains had opposite affinities for DAG and phorbol ester; i.e. the C1a domain showed high affinity for DAG and the C1b domain showed high affinity for phorbol ester. No such difference in selectivity was observed by Irie et al. (20).PKC has emerged as a promising therapeutic target both for cancer and for other conditions, such as diabetic retinopathy or macular degeneration (26-30). Kinase inhibitors represent one promising approach for targeting PKC, and enzastaurin, an inhibitor with moderate selectivity for PKCβ relative to other PKC isoforms (but still with activity on some other non-PKC kinases) is currently in multiple clinical trials. An alternative strategy for drug development has been to target the regulatory C1 domains of PKC. Strong proof of principle for this approach is provided by multiple natural products, e.g. bryostatin 1 and PEP005, which are likewise in clinical trials and which are directed at the C1 domains. A potential advantage of this approach is the lesser number of homologous targets, <30 DAG-sensitive C1 domains compared with over 500 kinases, as well as further opportunities for specificity provided by the diversity of lipid environments, which form a half-site for ligand binding to the C1 domain. Because different PKC isoforms may induce antagonistic activities, inhibition of one isoform may be functionally equivalent to activation of an antagonistic isoform (31).Along with the benzolactams (20, 32), the DAG lactones have provided a powerful synthetic platform for manipulating ligand: C1 domain interactions (31). For example, the DAG lactone derivative 130C037 displayed marked selectivity among the recombinant C1a and C1b domains of PKCα and PKCδ as well as substantial selectivity for RasGRP relative to PKCα (33). Likewise, we have shown that a modified DAG lactone (dioxolanones) can afford an additional point of contact in ligand binding to the C1b domain of PKCδ (34). Such studies provide clear examples that ligand-C1 domain interactions can be manipulated to yield novel patterns of recognition. Further selectivity might be gained with bivalent compounds, exploiting the spacing and individual characteristics of the C1a and C1b domains (35). A better understanding of the differential roles of the two C1 domains in PKC regulation is critical for the rational development of such compounds. In this study, by molecularly manipulating the C1a or C1b domains in intact PKCδ, we find that both the C1a and C1b domains play important roles in PKCδ regulation. The C1b domain is predominant for ligand binding and for membrane translocation of the whole PKCδ molecule. The C1a domain of intact PKCδ plays only a secondary role in ligand binding but stabilizes the PKCδ molecule at the plasma membrane for downstream signaling. In addition, we show that the effect of the individual C1 domains of PKCδ does not critically depend on their position within the regulatory domain.  相似文献   

9.
10.
Calliostoma tupinamba isa new species from Southeastern Brazil, ranging from southern Rio de Janeiro to northern São Paulo, and found only on coastal islands, on rocks and sessile invertebrates at 3 to 5 meters of depth. Shell and soft part morphology is described here in detail. Calliostoma tupinamba is mainly characterized by a depressed trochoid shell; eight slightly convex whorls; a sharply suprasutural carina starting on the third whorl and forming a peripheral rounded keel; and a whitish, funnel-shaped and deep umbilicus, measuring about 5%–10% of maximum shell width. Calliostoma tupinamba resembles Calliostoma bullisi Clench & Turner, 1960 in shape, but differs from it in being taller and wider, having a smaller umbilicus and lacking a strong and large innermost spiral cord at its base. Finally, an identification key of Brazilian Calliostoma species is presented.  相似文献   

11.
12.
The acid-sensing ion channel 1a (ASIC1a) is widely expressed in central and peripheral neurons where it generates transient cation currents when extracellular pH falls. ASIC1a confers pH-dependent modulation on postsynaptic dendritic spines and has critical effects in neurological diseases associated with a reduced pH. However, knowledge of the proteins that interact with ASIC1a and influence its function is limited. Here, we show that α-actinin, which links membrane proteins to the actin cytoskeleton, associates with ASIC1a in brain and in cultured cells. The interaction depended on an α-actinin-binding site in the ASIC1a C terminus that was specific for ASIC1a versus other ASICs and for α-actinin-1 and -4. Co-expressing α-actinin-4 altered ASIC1a current density, pH sensitivity, desensitization rate, and recovery from desensitization. Moreover, reducing α-actinin expression altered acid-activated currents in hippocampal neurons. These findings suggest that α-actinins may link ASIC1a to a macromolecular complex in the postsynaptic membrane where it regulates ASIC1a activity.Acid-sensing ion channels (ASICs)2 are H+-gated members of the DEG/ENaC family (13). Members of this family contain cytosolic N and C termini, two transmembrane domains, and a large cysteine-rich extracellular domain. ASIC subunits combine as homo- or heterotrimers to form cation channels that are widely expressed in the central and peripheral nervous systems (14). In mammals, four genes encode ASICs, and two subunits, ASIC1 and ASIC2, have two splice forms, a and b. Central nervous system neurons express ASIC1a, ASIC2a, and ASIC2b (57). Homomeric ASIC1a channels are activated when extracellular pH drops below 7.2, and half-maximal activation occurs at pH 6.5–6.8 (810). These channels desensitize in the continued presence of a low extracellular pH, and they can conduct Ca2+ (9, 1113). ASIC1a is required for acid-evoked currents in central nervous system neurons; disrupting the gene encoding ASIC1a eliminates H+-gated currents unless extracellular pH is reduced below pH 5.0 (5, 7).Previous studies found ASIC1a enriched in synaptosomal membrane fractions and present in dendritic spines, the site of excitatory synapses (5, 14, 15). Consistent with this localization, ASIC1a null mice manifested deficits in hippocampal long term potentiation, learning, and memory, which suggested that ASIC1a is required for normal synaptic plasticity (5, 16). ASICs might be activated during neurotransmission when synaptic vesicles empty their acidic contents into the synaptic cleft or when neuronal activity lowers extracellular pH (1719). Ion channels, including those at the synapse often interact with multiple proteins in a macromolecular complex that incorporates regulators of their function (20, 21). For ASIC1a, only a few interacting proteins have been identified. Earlier work indicated that ASIC1a interacts with another postsynaptic scaffolding protein, PICK1 (15, 22, 23). ASIC1a also has been reported to interact with annexin II light chain p11 through its cytosolic N terminus to increase cell surface expression (24) and with Ca2+/calmodulin-dependent protein kinase II to phosphorylate the channel (25). However, whether ASIC1a interacts with additional proteins and with the cytoskeleton remain unknown. Moreover, it is not known whether such interactions alter ASIC1a function.In analyzing the ASIC1a amino acid sequence, we identified cytosolic residues that might bind α-actinins. α-Actinins cluster membrane proteins and signaling molecules into macromolecular complexes and link membrane proteins to the actincytoskeleton (for review, Ref. 26). Four genes encode α-actinin-1, -2, -3, and -4 isoforms. α-Actinins contain an N-terminal head domain that binds F-actin, a C-terminal region containing two EF-hand motifs, and a central rod domain containing four spectrin-like motifs (2628). The C-terminal portion of the rod segment appears to be crucial for binding to membrane proteins. The α-actinins assemble into antiparallel homodimers through interactions in their rod domain. α-Actinins-1, -2, and -4 are enriched in dendritic spines, concentrating at the postsynaptic membrane (2935). In the postsynaptic membrane of excitatory synapses, α-actinin connects the NMDA receptor to the actin cytoskeleton, and this interaction is key for Ca2+-dependent inhibition of NMDA receptors (3638). α-Actinins can also regulate the membrane trafficking and function of several cation channels, including L-type Ca2+ channels, K+ channels, and TRP channels (3941).To better understand the function of ASIC1a channels in macromolecular complexes, we asked if ASIC1a associates with α-actinins. We were interested in the α-actinins because they and ASIC1a, both, are present in dendritic spines, ASIC1a contains a potential α-actinin binding sequence, and the related epithelial Na+ channel (ENaC) interacts with the cytoskeleton (42, 43). Therefore, we hypothesized that α-actinin interacts structurally and functionally with ASIC1a.  相似文献   

13.
The Escherichia coli envelope stress response is controlled by the alternative sigma factor, σE, and is induced when unfolded outer membrane proteins accumulate in the periplasm. The response is initiated by sequential cleavage of the membrane-spanning antisigma factor, RseA. RseB is an important negative regulator of envelope stress response that exerts its negative effects onσE activity through its binding to RseA. In this study, we analyze the interaction between RseA and RseB. We found that tight binding of RseB to RseA required intact RseB. Using programs that performed global and local sequence alignment of RseB and RseA, we found regions of high similarity and performed alanine substitution mutagenesis to test the hypothesis that these regions were functionally important. This protocol is based on the hypothesis that functionally dependent regions of two proteins co-evolve and therefore are likely to be sequentially conserved. This procedure allowed us to identify both an N-terminal and C-terminal region in RseB important for binding to RseA. We extensively analyzed the C-terminal region, which aligns with a region of RseA coincident with the major RseB binding determinant in RseA. Both allele-specific suppression analysis and cysteine-mediated disulfide bond formation indicated that this C-terminal region of similarity of RseA and RseB identifies a contact site between the two proteins. We suggest a similar protocol can be successfully applied to pairs of non-homologous but functionally linked proteins to find specific regions of the protein sequences that are important for establishing functional linkage.The Escherichia coli σE-mediated envelope stress response is the major pathway to ensure homeostasis in the envelope compartment of the cell (1-3). σE regulon members encode periplasmic chaperones and proteases, the machinery for inserting β-barrel proteins into the outer membrane and components controlling the synthesis and assembly of LPS (4-6). This pathway is highly conserved among γ-proteobacteria (6).The σE response is initiated when periplasmic protein folding and assembly is compromised (7-9). During steady state growth, σE is inhibited by its antisigma factor, RseA, a membrane-spanning protein whose cytoplasmic domain binds to σE with picomolar affinity (10-13). Accumulation of unassembled porin monomers serves as a signal to activate the DegS protease to cleave RseA in its periplasmic domain (14, 15). This initiates a proteolytic cascade in which RseP cleaves periplasmically truncated RseA near or within the cytoplasmic membrane to release the RseAcytoplasmicE complex, and cytoplasmic ATP-dependent proteases complete the degradation of RseA thereby releasing active σE (16-19).RseB, a second negative regulator of the envelope stress response (11, 20, 21), binds to the periplasmic domain of RseA with nanomolar affinity. RseB is an important regulator of the response (2, 22, 23). It prevents RseP from degrading intact RseA, thereby ensuring that proteolysis is initiated only when the DegS protease is activated by a stress signal (21). Additionally, RseB prevents activated DegS from cleaving RseA, suggesting that interaction of RseB with RseA must be altered before the signal transduction cascade is activated (23).The goal of the present studies was to explore how RseB binds to RseA. The interaction partner of RseB is the unstructured periplasmic domain of RseA (RseA-peri). Within RseA-peri, amino acids ∼169-186 constitute a major binding determinant to RseB (23, 24). This peptide alone binds RseB with 6 μm affinity, and deleting this region abrogates binding to RseB (23). Additional regions of RseA-peri also contribute to RseB binding, as intact RseA-peri binds with 20 nm affinity to RseB (23). Much less is known about the regions of RseB required for interaction with RseA. RseB is homodimeric two-domain protein, whose large N-terminal domain shares structural homology with LolA, a protein that transports lipoproteins to outer membrane (24, 25). The smaller C-terminal domain is connected to the N-terminal domain by a linker, and the two domains share a large interface, which may facilitate interdomain signaling. Glutaraldehyde cross-linking studies indicate that the C-terminal domain interacts with RseA, but the regions of interaction were not identified (25).In the present report, we study the interaction of RseB and RseA. We establish that both domains of RseB interact with RseA-peri. Using a global sequence alignment, we discovered several regions in RseA and RseB that had high sequence similarity, despite the low overall sequence similarity between these two proteins, a finding that was independently confirmed by a local sequence similarity algorithm. This suggested that these regions were functionally dependent, and we performed a set of mutagenesis experiments designed to test this idea. Our studies of the binding properties of these mutants revealed that regions in both the N terminus and C terminus of RseB modulate interaction with RseA. Moreover, genetic suppression analysis and cysteine-mediated disulfide bond formation suggest that the region of RseA/B with highest similarity (RseA residues 165-191 (major binding determinant in RseA) and RseB residues 233-258) are interacting partners.  相似文献   

14.
Previous studies of photosynthetic acclimation to elevated CO2 have focused on the most recently expanded, sunlit leaves in the canopy. We examined acclimation in a vertical profile of leaves through a canopy of wheat (Triticum aestivum L.). The crop was grown at an elevated CO2 partial pressure of 55 Pa within a replicated field experiment using free-air CO2 enrichment. Gas exchange was used to estimate in vivo carboxylation capacity and the maximum rate of ribulose-1,5-bisphosphate-limited photosynthesis. Net photosynthetic CO2 uptake was measured for leaves in situ within the canopy. Leaf contents of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco), light-harvesting-complex (LHC) proteins, and total N were determined. Elevated CO2 did not affect carboxylation capacity in the most recently expanded leaves but led to a decrease in lower, shaded leaves during grain development. Despite this acclimation, in situ photosynthetic CO2 uptake remained higher under elevated CO2. Acclimation at elevated CO2 was accompanied by decreases in both Rubisco and total leaf N contents and an increase in LHC content. Elevated CO2 led to a larger increase in LHC/Rubisco in lower canopy leaves than in the uppermost leaf. Acclimation of leaf photosynthesis to elevated CO2 therefore depended on both vertical position within the canopy and the developmental stage.  相似文献   

15.
Recent studies have revealed that in G protein-coupled receptor signalings switching between G protein- and β-arrestin (βArr)-dependent pathways occurs. In the case of opioid receptors, the signal is switched from the initial inhibition of adenylyl cyclase (AC) to an increase in AC activity (AC activation) during prolonged agonist treatment. The mechanism of such AC activation has been suggested to involve the switching of G proteins activated by the receptor, phosphorylation of signaling molecules, or receptor-dependent recruitment of cellular proteins. Using protein kinase inhibitors, dominant negative mutant studies and mouse embryonic fibroblast cells isolated from Src kinase knock-out mice, we demonstrated that μ-opioid receptor (OPRM1)-mediated AC activation requires direct association and activation of Src kinase by lipid raft-located OPRM1. Such Src activation was independent of βArr as indicated by the ability of OPRM1 to activate Src and AC after prolonged agonist treatment in mouse embryonic fibroblast cells lacking both βArr-1 and -2. Instead the switching of OPRM1 signals was dependent on the heterotrimeric G protein, specifically Gi2 α-subunit. Among the Src kinase substrates, OPRM1 was phosphorylated at Tyr336 within NPXXY motif by Src during AC activation. Mutation of this Tyr residue, together with mutation of Tyr166 within the DRY motif to Phe, resulted in the complete blunting of AC activation. Thus, the recruitment and activation of Src kinase by OPRM1 during chronic agonist treatment, which eventually results in the receptor tyrosine phosphorylation, is the key for switching the opioid receptor signals from its initial AC inhibition to subsequent AC activation.Classical G protein-coupled receptor (GPCR)2 signaling involves the activation of specific heterotrimeric G proteins and the subsequent dissociation of α- and βγ-subunits. These G protein subunits serve as the activators and/or inhibitors of several effector systems, including adenylyl cyclases, phospholipases, and ion channels (1). However, recent studies have shown that GPCR signaling deviates from such a classical linear model. For example, in kidney and colonic epithelial cells, protease-activated receptor 1 can transduce its signals through either Gαi/o or Gαq subunits via inhibition of small GTPase RhoA or activation of RhoD. Thus, RhoA and RhoD act as molecular switches between the negative and positive signaling activity of protease-activated receptor 1 (2). Another example is the ability of β2-adrenergic receptor to switch from Gs-dependent pathways to non-classical signaling pathways by coupling to pertussis toxin-sensitive Gi proteins in a cAMP-dependent protein kinase/protein kinase C phosphorylation-dependent manner. In this case, the phosphorylation-induced switch in G protein coupling provides the receptor access to alternative signaling pathways. For β2-adrenergic receptors, this leads to a Gi-dependent activation of MAP kinase (3, 4). Furthermore the involvement of protein scaffolds, such as β-arrestins in the MAP kinase cascade, could also alter the GPCR signaling (58). Hence the formation of “signaling units” or “receptosomes” would influence the GPCR signaling process and destination.For opioid receptors, which are members of the rhodopsin GPCR subfamily receptors, signal switching is also observed. Normally opioid receptors inhibit AC activity, activate the MAP kinases and Kir3 K+ channels, inhibit the voltage-dependent Ca2+ channels, and regulate other effectors such as phospholipase C (9). However, during prolonged agonist treatment, not only is there a blunting of these cellular responses but also a compensatory increase in intracellular cAMP level, which is particularly significant upon the removal of the agonist or the addition of an antagonist such as naloxone (1012). This compensatory adenylyl cyclase activation phenomenon has been postulated to be responsible for the development of drug tolerance and dependence (13). The observed change from receptor-mediated AC inhibition to receptor-mediated AC activation reflects possible receptor signal switching. Although the exact mechanism for such signal changes has yet to be elucidated, activation of specific protein kinases and subsequent phosphorylation of AC isoforms (14, 15) and other signaling molecules (16) have been suggested to be the key for observed AC activation. Among all the protein kinases studied, involvement of protein kinase C, MAP kinase, and Raf-1 has been implicated in the activation of AC (1719). Alternative mechanisms, such as agonist-induced receptor internalization and the increase in the constitutive activities of the receptor, also have been suggested to play a role in increased AC activity after prolonged opioid agonist treatment (20). Earlier studies also implicated the switching of the opioid receptor from Gi/Go to Gs coupling during chronic agonist treatment (21). Regardless of the mechanism, the exact molecular events that lead to the switching of opioid receptor from an inhibitory response to a stimulatory response remain elusive.Src kinases, which are members of the nonreceptor tyrosine kinase family, have been implicated in GPCR function because several Src family members such as cSrc, Fyn, and Yes have been reported to be activated by several GPCRs, including β2- (22) and β3 (23)-adrenergic, M2- (24) and M3 (25)-muscarinic, and bradykinin receptors (26). The GPCRs that are capable of activating Src predominantly couple to Gi/o family G proteins (27). Src kinases appear to associate with, and be activated by, GPCRs themselves either through direct interaction with intracellular receptor domains or by binding to GPCR-associated proteins, such as G protein subunits or β-arrestins (27). Src kinase has been reported to be activated by κ- (28) and δ (29)-opioid receptors and regulate the c-Jun kinase and MAP kinase activities. Src kinase within the nucleus accumbens has been implicated in the rewarding effect and hyperlocomotion induced by morphine in mice (30). However, it is not clear whether the Src kinase is activated and involved in the signal transduction in AC activation after chronic opioid agonist administration.Previously we reported that the lipid raft location of the receptor and the Gαi2 proteins are two prerequisites for the observed increase in AC activity during prolonged agonist treatment (31, 32). Because various protein kinases including Src kinases and G proteins have been shown to be enriched in lipid rafts (33), the roles of these cellular proteins in the eventual switching of opioid receptor signals from inhibition to stimulation of AC activity were examined in the current studies. We were able to demonstrate that the association with and subsequent activation of Src kinase by the μ-opioid receptor (OPRM1), which leads to eventual tyrosine phosphorylation of OPRM1, are the cellular events required for the switching of opioid receptor signaling upon chronic agonist treatment.  相似文献   

16.
A new role is reported for CP12, a highly unfolded and flexible protein, mainly known for its redox function with A4 glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Both reduced and oxidized CP12 can prevent the in vitro thermal inactivation and aggregation of GAPDH from Chlamydomonas reinhardtii. This mechanism is thus not redox-dependent. The protection is specific to CP12, because other proteins, such as bovine serum albumin, thioredoxin, and a general chaperone, Hsp33, do not fully prevent denaturation of GAPDH. Furthermore, CP12 acts as a specific chaperone, since it does not protect other proteins, such as catalase, alcohol dehydrogenase, or lysozyme. The interaction between CP12 and GAPDH is necessary to prevent the aggregation and inactivation, since the mutant C66S that does not form any complex with GAPDH cannot accomplish this protection. Unlike the C66S mutant, the C23S mutant that lacks the N-terminal bridge is partially able to protect and to slow down the inactivation and aggregation. Tryptic digestion coupled to mass spectrometry confirmed that the S-loop of GAPDH is the interaction site with CP12. Thus, CP12 not only has a redox function but also behaves as a specific “chaperone-like protein” for GAPDH, although a stable and not transitory interaction is observed. This new function of CP12 may explain why it is also present in complexes involving A2B2 GAPDHs that possess a regulatory C-terminal extension (GapB subunit) and therefore do not require CP12 to be redox-regulated.CP12 is a small 8.2-kDa protein present in the chloroplasts of most photosynthetic organisms, including cyanobacteria (1, 2), higher plants (3), the diatom Asterionella formosa (4, 5), and green (1) and red algae (6). It allows the formation of a supramolecular complex between phosphoribulokinase (EC 2.7.1.19) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH),3 two key enzymes of the Calvin cycle pathway, and was recently shown to interact with fructose bisphosphate aldolase, another enzyme of the Calvin cycle pathway (7). The phosphoribulokinase·GAPDH·CP12 complex has been extensively studied in Chlamydomonas reinhardtii (8, 9) and in Arabidopsis thaliana (10, 11). In the green alga C. reinhardtii, the interaction between CP12 and GAPDH is strong (8). GAPDH may exist as a homotetramer composed of four GapA subunits (A4) in higher plants, cyanobacteria, and green and red algae (6, 12), but in higher plants, it can also exist as a heterotetramer (A2B2), composed of two subunits, GapA and GapB (13, 14). GapB, up to now, has exclusively been found in Streptophyta, but recently two prasinophycean green algae, Ostreococcus tauri and Ostreococcus lucimarinus, were also shown to possess a GapB gene, whereas CP12 is missing (15). The GapB subunit is similar to the GapA subunit but has a C-terminal extension containing two redox-regulated cysteine residues (16). Thus, although the A4 GAPDHs lack these regulatory cysteine residues (13, 14, 1720), they are also redox-regulated through its interaction with CP12, since the C terminus of this small protein resembles the C-terminal extension of the GapB subunit. The regulatory cysteine residues for GapA are thus supplied by CP12, as is well documented in the literature (1, 8, 11, 16).CP12 belongs to the family of intrinsically unstructured proteins (IUPs) (2126). The amino acid composition of these proteins causes them to have no or few secondary structures. Their total or partial lack of structure and their high flexibility allow them to be molecular adaptors (27, 28). They are often able to bind to several partners and are involved in most cellular functions (29, 30). Recently, some IUPs have been described in photosynthetic organisms (31, 32).There are many functional categories of IUPs (22, 33). They can be, for instance, involved in permanent binding and have (i) a scavenger role, neutralizing or storing small ligands; (ii) an assembler role by forming complexes; and (iii) an effector role by modulating the activity of a partner molecule (33). These functions are not exclusive; thus, CP12 can form a stable complex with GAPDH, regulating its redox properties (8, 34, 35), and can also bind a metal ion (36, 37). IUPs can also bind transiently to partners, and some of them have been found to possess a chaperone activity (31, 38). This chaperone function was first shown for α-synuclein (39) and for α-casein (40), which are fully disordered. The amino acid composition of IUPs is less hydrophobic than those of soluble proteins; hence, they lack hydrophobic cores and do not become insoluble when heated. Since CP12 belongs to this family, we tested if it was resistant to heat treatment and finally, since it is tightly bound to GAPDH, if it could prevent aggregation of its partner, GAPDH, an enzyme well known for its tendency to aggregate (4144) and consequently a substrate commonly used in chaperone studies (45, 46).Unlike chaperones, which form transient, dynamic complexes with their protein substrates through hydrophobic interactions (47, 48), CP12 forms a stable complex with GAPDH. The interaction involves the C-terminal part of the protein and the presence of negatively charged residues on CP12 (35). However, only a site-directed mutagenesis has been performed to characterize the interaction site on GAPDH. Although the mutation could have an indirect effect, the residue Arg-197 was shown to be a good candidate for the interaction site (49).In this report, we accordingly used proteolysis experiments coupled with mass spectrometry to detect which regions of GAPDH are protected by its association with CP12. To conclude, the aim of this report was to characterize a chaperone function of CP12 that had never been described before and to map the interaction site on GAPDH using an approach that does not involve site-directed mutagenesis.  相似文献   

17.
A new genus of Isotomidae, Bellisotoma gen. n., is described. The new genus is a member of the Proisotoma genus complex and is characterized by a combination of having a bidentate mucro with wide dorsal lamellae that join clearly before the end of mucronal axis without forming a tooth and one strong ventral rib with basal notch that articulates with dens; having abundant chaetotaxy on both faces of dens; and abundant tergal sensilla. Bellisotoma gen. n. shows a furcula adapted to a neustonic mode of life, and may be a Isotopenola-like derivative adapted to neustonic habitats. Subisotoma joycei Soto-Adames & Giordano, 2011 and Ballistura ewingi James, 1933 are transferred to the new genus.  相似文献   

18.
The cyst wall of Entamoeba invadens (Ei), a model for the human pathogen Entamoeba histolytica, is composed of fibrils of chitin and three chitin-binding lectins called Jacob, Jessie3, and chitinase. Here we show chitin, which was detected with wheat germ agglutinin, is made in secretory vesicles prior to its deposition on the surface of encysting Ei. Jacob lectins, which have tandemly arrayed chitin-binding domains (CBDs), and chitinase, which has an N-terminal CBD, were each made early during encystation. These results are consistent with their hypothesized roles in cross-linking chitin fibrils (Jacob lectins) and remodeling the cyst wall (chitinase). Jessie3 lectins likely form the mortar or daub of the cyst wall, because 1) Jessie lectins were made late during encystation; 2) the addition to Jessie lectins to the cyst wall correlated with a marked decrease in the permeability of cysts to nucleic acid stains (DAPI) and actin-binding heptapeptide (phalloidin); and 3) recombinant Jessie lectins, expressed as a maltose-binding proteins in the periplasm of Escherichia coli, caused transformed bacteria to agglutinate in suspension and form a hard pellet that did not dissociate after centrifugation. Jessie3 appeared as linear forms and rosettes by negative staining of secreted recombinant proteins. These findings provide evidence for a “wattle and daub” model of the Entamoeba cyst wall, where the wattle or sticks (chitin fibrils likely cross-linked by Jacob lectins) is constructed prior to the addition of the mortar or daub (Jessie3 lectins).  相似文献   

19.
A new species of Profundulus, Profundulus kreiseri (Cyprinodontiformes: Profundulidae), is described from the Chamelecón and Ulúa Rivers in the northwestern Honduran highlands. Based on a phylogenetic analysis using cytochrome b and the presence of synapomorphic characters (dark humeral spot, a scaled preorbital region and between 32-34 vertebrae), this new species is placed in the subgenus Profundulus, which also includes Profundulus (Profundulus) oaxacae, Profundulus (Profundulus) punctatus and Profundulus (Profundulus) guatemalensis. Profundulus kreiseri can be distinguished from other members of the subgenus Profundulus by having less than half of its caudal fin densely scaled. Profundulus kreiseri can further be differentiated from Profundulus (Profundulus) oaxacae and Profundulus (Profundulus) punctatus by the absence of rows of dark spots on its flanks. The new species can further be differentiated from Profundulus (Profundulus) guatemalensis by the presence of fewer caudal- and pectoral-fin rays. The new species is distinguished from congeners of the profundulid subgenus Tlaloc (viz., Profundulus (Tlaloc) hildebrandi, Profundulus (Tlaloc) labialis, Profundulus (Tlaloc) candalarius and Profundulus (Tlaloc) portillorum) by having a scaled preorbital region and a dark humeral spot. Profundulus kreiseri and Profundulus portillorum are the only two species of Profundulus that are endemic to the region south of the Motagua River drainage in southern Guatemala and northwestern Honduras.  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